Anda di halaman 1dari 263

Advances in Industrial Control

For further volumes:


www.springer.com/series/1412
Other titles published in this series:

Digital Controller Implementation Modelling and Control of Mini-Flying


and Fragility Machines
Robert S.H. Istepanian and James F. Pedro Castillo, Rogelio Lozano and
Whidborne (Eds.) Alejandro Dzul
Optimisation of Industrial Processes Ship Motion Control
at Supervisory Level Tristan Perez
Doris Sáez, Aldo Cipriano and Andrzej W.
Ordys Hard Disk Drive Servo Systems (2nd Ed.)
Ben M. Chen, Tong H. Lee, Kemao Peng
Robust Control of Diesel Ship Propulsion and Venkatakrishnan Venkataramanan
Nikolaos Xiros
Measurement, Control, and
Hydraulic Servo-systems Communication Using IEEE 1588
Mohieddine Mali and Andreas Kroll John C. Eidson
Model-based Fault Diagnosis in Dynamic Piezoelectric Transducers for Vibration
Systems Using Identification Techniques Control and Damping
Silvio Simani, Cesare Fantuzzi and Ron J. S.O. Reza Moheimani and Andrew J.
Patton Fleming
Strategies for Feedback Linearisation Manufacturing Systems Control Design
Freddy Garces, Victor M. Becerra, Stjepan Bogdan, Frank L. Lewis, Zdenko
Chandrasekhar Kambhampati and Kevin Kovačić and José Mireles Jr.
Warwick
Windup in Control
Robust Autonomous Guidance Peter Hippe
Alberto Isidori, Lorenzo Marconi
and Andrea Serrani Nonlinear H2 /H∞ Constrained Feedback
Control
Dynamic Modelling of Gas Turbines
Murad Abu-Khalaf, Jie Huang
Gennady G. Kulikov and Haydn A.
and Frank L. Lewis
Thompson (Eds.)
Practical Grey-box Process Identification
Control of Fuel Cell Power Systems
Torsten Bohlin
Jay T. Pukrushpan, Anna G. Stefanopoulou
and Huei Peng Control of Traffic Systems in Buildings
Sandor Markon, Hajime Kita, Hiroshi Kise
Fuzzy Logic, Identification and Predictive
and Thomas Bartz-Beielstein
Control
Jairo Espinosa, Joos Vandewalle Wind Turbine Control Systems
and Vincent Wertz Fernando D. Bianchi, Hernán De Battista
Optimal Real-time Control of Sewer and Ricardo J. Mantz
Networks Advanced Fuzzy Logic Technologies
Magdalene Marinaki and Markos in Industrial Applications
Papageorgiou Ying Bai, Hanqi Zhuang and Dali Wang
Process Modelling for Control (Eds.)
Benoît Codrons Practical PID Control
Computational Intelligence in Time Series Antonio Visioli
Forecasting
Ajoy K. Palit and Dobrivoje Popovic (continued after Index)
Alfonso Baños r Antonio Barreiro

Reset Control
Systems
Alfonso Baños Antonio Barreiro
Fac. Informática Depto. Ingeniería de Sistemas y Automática
Depto. Informática y Sistemas ETSII
Grupo de Informática Industrial Universidad de Vigo
Universidad de Murcia Campus de Lagoas-Marcosende
Murcia 30071 Vigo 36310
Spain Spain
abanos@um.es abarreiro@uvigo.es

ISSN 1430-9491 Advances in Industrial Control


ISBN 978-1-4471-2216-6 e-ISBN 978-1-4471-2250-0
DOI 10.1007/978-1-4471-2250-0
Springer London Dordrecht Heidelberg New York

British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library

Library of Congress Control Number: 2011940675

© Springer-Verlag London Limited 2012


Apart from any fair dealing for the purposes of research or private study, or criticism or review, as per-
mitted under the Copyright, Designs and Patents Act 1988, this publication may only be reproduced,
stored or transmitted, in any form or by any means, with the prior permission in writing of the publish-
ers, or in the case of reprographic reproduction in accordance with the terms of licenses issued by the
Copyright Licensing Agency. Enquiries concerning reproduction outside those terms should be sent to
the publishers.
The use of 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 laws and regulations and therefore free
for general use.
The publisher makes no representation, express or implied, with regard to the accuracy of the information
contained in this book and cannot accept any legal responsibility or liability for any errors or omissions
that may be made.

Cover design: VTeX UAB, Lithuania

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


To Angeles and Alfonso (A. Baños)
To Rosa, María, and Marta (A. Barreiro)
Series Editors’ Foreword

The series Advances in Industrial Control aims to report and encourage technology
transfer in control engineering. The rapid development of control technology has
an impact on all areas of the control discipline. New theory, new controllers, actu-
ators, sensors, new industrial processes, computer methods, new applications, new
philosophies all lead to new challenges. Much of this development work resides in
industrial reports, feasibility study papers and the reports of advanced collaborative
projects. The series offers an opportunity for researchers to present an extended ex-
position of such new work in all aspects of industrial control for wider and rapid
dissemination.
The name of Isaac Horowitz is usually associated with quantitative feedback the-
ory (QFT), but his first-order reset element (FORE) is probably not so well known.
This Horowitz FORE compensator was a generalisation of the seminal Clegg Inte-
grator (1958) in which integral control was coupled with a reset mechanism to in-
culcate additional performance and design benefits. These two compensators are the
original inspiration and source for the results and applications reported in this Ad-
vances in Industrial Control monograph, Reset Control Systems, by Alfonso Baños
and Antonio Barreiro. In fact, the authors’ whole reset control research project
seems to have been much motivated by original meetings in Spain with Professor
Horowitz.
The monograph is a comprehensive presentation of reset control systems, cov-
ering the fundamental theoretical foundations, design rules and applications experi-
ence. The monograph has four chapters on theoretical fundamentals. Opening these
is an invaluable introductory chapter (Chap. 1) that guides the reader through the
historical development of reset control. This particular chapter also sets out the the-
oretical context of reset systems and identifies their relationship with other related
theoretical fields, for example impulsive and hybrid systems. Chapters. 2 to 4 con-
tain the core theoretical results of the monograph and set out, possibly for the first
time, a complete systems analysis framework for reset control.
Design for reset control is treated in Chap. 5, where the focus is on identifying
the type of systems for which reset control offers good design and performance
benefits. Alongside this; sets of useful compensator tuning rules are presented for

vii
viii Series Editors’ Foreword

different compensator and system types. The chapter also investigates the properties
of the authors’ own novel PI + CI reset compensator.
In control engineering, the assessment of usefulness, effectiveness, design ease
and performance gain is through real applications studies. In this respect, the case
studies of Chap. 6 provide a trio of demanding “benchmark” applications. The au-
thors expose all the steps of trying to accomplish successful reset control system
designs. The applications are from process control (a pilot heat exchanger plant),
control by tele-operation (a remote controlled process application that uses the Inter-
net for communication) and utility process operations (a solar collector field control
problem). This chapter allows both the practitioner and the theoretician to see how
reset control works in practice, and the insights provided will undoubtedly inspire
new developments for the reset control field.
This invaluable and comprehensive study of reset control systems by Professors
Baños and Barreiro joins several other recent Advances in Industrial Control mono-
graphs that give more fundamental presentations of new potential applicable topics.
One such monograph is Process Control by Jie Bao and Peter L. Lee (ISBN 978-1-
84628-892-0, 2007) that was received enthusiastically; another is Fractional-order
Systems and Controls by Concepción A. Monje, YangQuan Chen, Blas M. Vinagre,
Dingyü Xue and Vicente Feliu (ISBN 978-1-84996-334-3, 2010), and a third is,
Internet-based Control Systems by Shuang-Hua Yang (ISBN 978-1-84996-358-9,
2011). The industrial control engineer, the control academic or control postgraduate
researcher may find these three “fundamentals with applications” texts along with,
Reset Control Systems essential volumes for inclusion in a complete control library.
Industrial Control Centre, Glasgow, Scotland, UK M.J. Grimble
M.A. Johnson
Foreword

As part of the small community of control researchers, often referred to as the QFT-
ers, I first met Alfonso Baños in 1997 at an international QFT conference in Glas-
gow, Scotland. We kept in touch throughout the years via emails, and in 2008 he
arranged for his PhD student, Joaquín Carrasco, to spend a few months in my re-
search lab. Joaquín’s thesis focused on reset control, and he wanted to learn about
our activities in that area. Since then, I continued to follow their work and have
come to fully appreciate their contributions to the area of reset control.
My foray into reset control research was somewhat accidental. While working
on challenging industrial robust control problems, we ‘stumbled’ upon a series of
papers from the 1970s written by the late professor I. Horowitz, the founder of QFT,
and his students. I recall that right away we noticed something unique about these
papers. Horowitz presented nontraditional solutions to overcome the limitations
of linear, time-invariant (LTI) control. Horowitz was already familiar with Clegg-
integrator and similar concepts that were introduced a few decades earlier on purely
empirical grounds. Horowitz’s results on Clegg-integrator and First-Order Reset El-
ement (FORE) offered the practicing control engineer a systematic and transparent
procedure for using such nonlinear elements in feedback control design. In our own
work, we initially attempted to use these ideas in experiments, and motivated by
successful results, we decided to investigate if one can explain such results theoreti-
cally. Orhan Beker, a PhD student of CV Hollot and I, was able to prove for the first
time that reset control can overcome classic LTI performance limitations in a feed-
back system comprising a plant with an integrator. I believe that these encouraging
results may have spurned other groups, including Baños’s and Barreiro’s, to probe
deeper into this area which falls under the more general subjects of hybrid control
systems and impulsive control.
This book presents a compilation of numerous results derived by the authors and
colleagues. The material is written in self-contained manner and offers the interested
reader an exposition of current state-of-the-art results. The authors present stability
theory in progression starting from reset-times independent results and reset-times
dependent results, to the more general passivity-based results. A whole chapter is
devoted to time-delay systems where LTI limitations are especially harsh and stabil-

ix
x Foreword

ity conditions take the form of linear matrix inequalities. Numerical examples are
sprinkled throughout the presentation so to bring life into the theory.
In addition to the book’s theoretical focus, the authors appear to have made de-
liberate effort to remain true to the original motivation behind Horowitz’s original
work to provide the control engineer with a practical technique for design of reset
control systems in applications. They achieve this goal using describing function
analysis as a design tool, extend the notion of reset to a partial reset compensator
where only some of the compensator states are reset, and introduce a new reset com-
pensator that consists of a PI compensator whose integrator is reset only at a fraction
of Clegg integrator reset times. An entire chapter presents experimental applications
of reset control systems in a temperature control of a heat exchanger, a bilateral tele-
operation, and a temperature control of a solar collector field.
In closing, I would like to thank the authors for asking me to write this forward.
This well written book strikes a fine balance between sophisticated theory and en-
gineering design and therefore should be accessible to most readers. In writing this
book, the authors have succeeded in continuing Horowitz’s philosophy for feedback
control research.
Amherst, USA Yossi Chait
Preface

Although the origin of reset control systems goes back to 1958 with the founding
work of Clegg, the subject is still in its infancy. The seminal works of Horowitz
and coworkers in the 1970s were the first attempts to build a synthesis theory for
basic reset compensators such as the Clegg integrator or the FORE (first-order reset
element), but the lack of a control theory, approaching basic questions such as well-
posedness and stability, for several decades has put aside of the mainstream further
developments in reset control, both in theory and applications. It is not until the
late 1990s that reset control starts to develop with the works of Chait, Hollot and
coworkers, giving a significative impetus to the field. In the meantime, the area of
hybrid control systems and impulsive control also give numerous results, having
clear connections with reset control. For the last decade, a number of international
groups have been working actively on reset control, and fortunately, reset control
has started to be seen as an attractive control design technique with a significant
potential for practical control applications.
The first direct contact of the authors with reset control was in September 2000,
when Isaac Horowitz was visiting our group in Spain. We were actively working in
nonlinear QFT at that time, that uses linear compensation, and in several fruitful dis-
cussions he gave us clear arguments for the many benefits of using nonlinear/reset
compensation to attack the “nonlinear tiger” and also to overcome fundamental lim-
itations of LTI systems. We started to seriously work on reset control in 2006, and
since then our goal has been to develop a formal theory to cover basic theoretical
aspects, and also to define simple compensator structures and tuning rules for them,
with the focus on some practical problems, including applications in process control
and teleoperation, and in general systems with time-delays.
The first chapter of the book is an introduction to reset control systems, pursuing
two objectives. The first objective is to give a quick and simple description of what
a reset control is, and to provide basic explanations on why and when it is conve-
nient to use this strategy. This objective is covered by the first two sections and is
summarized in this key idea: a reset control is a simple nonlinear control technique
very effective for linear plants subject to fundamental design limitations. The sec-
ond objective of the chapter is to give a brief survey on the literature on analysis

xi
xii Preface

and design of reset control systems. The historical perspective begins with the early
ideas on reset control, including the popular Clegg integrator and the first-order re-
set element (FORE) introduced by Horowitz and coworkers. In addition, the first
series of rigorous results on analysis and design of reset controllers using a state-
space description are given, including full reset and partial reset compensators. In
addition, relationships between reset control and the wider field of impulsive and
hybrid control systems will be analyzed from different points of view.
In Chap. 2, a definition of a reset control system, or a reset system in general, is
given. In general, as it is common in impulsive systems, reset systems may exhibit
different types of solutions, in particular having complex patterns such as beating,
deadlock, and Zenoness. In control practice, this type of behavior is considered
pathological and thus several conditions will be given for reset control systems to
be well-posed. On the other hand, important properties of reset systems may be
derived by analyzing the reset instants that correspond to a given initial condition.
These patterns will be also analyzed, and their relationship with the observability
and reachability of the base linear system will be shown.
Chapter 3 is devoted to the stability problem of reset control systems with finite-
dimensional base systems. The stability problem is addressed from different, com-
plementary points of view: (i) internal or Lyapunov stability, (ii) external or input–
output stability with passivity analysis, and (iii) stability by the describing function
method. Internal stability techniques are subdivided into techniques giving rise to
stability conditions that do not depend directly on the reset instants (reset-times in-
dependent), or alternatively, are reset-times dependent. The first case is obtained di-
rectly using continuous time Lyapunov functions (that gives rise to the so-called Hβ
condition), while the second case (reset-times dependent) requires a discretization
at the after-reset values and a subsequent discrete-time Lyapunov analysis. Then,
the input–output L2 stability is studied, and a number of results are presented in
connection with passivity and dissipativity properties of reset feedback loops. Fi-
nally, the standard describing function tool is used for approximately predicting the
appearance or absence of oscillations.
Stability of time-delay systems under reset control is approached in Chap. 4.
Since reset control is able to overcome fundamental limitations, and time-delay
is one source of such limitations, then it is of great interest to study the problem
of delayed reset systems. The stability is addressed by choosing an appropriate
Lyapunov–Krasovskii functional, and by imposing that the functional should de-
crease in the continuous and reset modes. The resulting conditions take the form of
linear matrix inequalities, and, depending on the chosen functional, these LMIs can
be delay-dependent or delay-independent. In both cases, those LMIs, derived from
time-domain stability conditions, are translated into equivalent frequency-domain
conditions by means of adequate tools, like the Kalman–Yakubovich–Popov lemma,
or passivity techniques. From the latter frequency-domain conditions, useful inter-
pretations are exhibited regarding the achieved robustness, in terms of scaled small-
gain or positive realness of certain subsystems. Finally, several examples illustrate
the application of the stability conditions, showing the potentials of reset control
when applied to time-delay systems.
Preface xiii

In Chap. 5, reset compensation has been used to overcome limitations of LTI


compensation. In this chapter, a new reset compensator, referred to as PI + CI, is
introduced. It basically consists of adding a Clegg integrator to a PI compensator,
with the goal of improving the closed loop response by using the nonlinear charac-
teristic of this element. It turns out that by resetting a percentage of the integral term
in a PI compensator, a significant improvement can be obtained over a well-tuned
PI compensator in some relevant practical cases, such as systems with dominant
lag and integrating systems. The main goal is the development of PI + CI tuning
rules for basic dynamic systems in a wide range of applications, including first- and
higher-order plus dead time systems. In addition, a number of design improvements
such us the use of a fixed or variable reset band, the integration with QFT, and the
use of a variable reset percentage are discussed.
Finally, in Chap. 6, several practical applications of reset control systems will be
developed, all based on the PI + CI compensator: a temperature control system of
a heat exchanger, a bilateral teleoperation control system, and finally, a temperature
control of a solar collector field. The first two applications have been tested by
means of experiments in plants, while the third has been tested by using a (well-
proven) simulator of the field.
This book is a compendium of the several works developed by our groups in the
last five years. These works have been performed in collaboration with Joaquín Car-
rasco, Angel Vidal, Alejandro Fernández, Juan Ignacio Mulero, Sebastián Dormido,
José Carlos Moreno, Manuel Berenguel, and Arjan van der Schaft. In fact, they are
responsible for many parts of this book. The authors also acknowledge the sup-
port of ‘Ministerio de Ciencia e Innovación’ (Spanish government) under the joint
projects DPI2004-07670, DPI2007-66455, and DPI2010-20466.
Thus, the book has been planned as a means to systematize and make available
to a wider audience a number of publications that are disseminated over several
journals and conference proceedings. It is intended for control researchers interested
in a solid introduction to reset control, including the several approaches available
in the literature, and also for control engineers interested in application of simple
and efficient control techniques that may overcome fundamental limitations of the
universally used PI/PID compensators.
Murcia, Spain Alfonso Baños
Vigo, Spain Antonio Barreiro
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 What Is Reset Control? . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Why Reset Control? . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Overcoming Fundamental Limitations in the Time Domain 9
1.2.2 Overcoming Limitations in the Frequency Domain . . . . . 11
1.3 Early Ideas on Reset Control . . . . . . . . . . . . . . . . . . . . 20
1.3.1 Reset Control with the Clegg Integrator . . . . . . . . . . . 20
1.3.2 Reset Control with the First-Order-Reset-Element (FORE) 23
1.4 First General Approaches to Analysis of Reset Control Systems . . 25
1.4.1 General Setup and Asymptotic Stability . . . . . . . . . . . 25
1.5 Reset Systems as Impulsive Systems . . . . . . . . . . . . . . . . 27
1.5.1 Theory of Impulsive Differential Equations . . . . . . . . . 29
1.5.2 Impulsive Control Theory . . . . . . . . . . . . . . . . . . 31
1.5.3 Impulsive Dynamical Systems . . . . . . . . . . . . . . . . 34
1.6 Reset Systems as Hybrid Systems . . . . . . . . . . . . . . . . . . 37
1.7 Other Recent Results on Reset Control . . . . . . . . . . . . . . . 43
1.7.1 Design Based on L2 and H2 Performance . . . . . . . . . 44
1.7.2 Design Based on L2 Gain and Nonzero References . . . . 47
1.7.3 Design Based on Fixed Reset Instants tk . . . . . . . . . . 48
1.8 Preview of the Chapters . . . . . . . . . . . . . . . . . . . . . . . 50
2 Definition of Reset Control System and Basic Results . . . . . . . . . 57
2.1 Preliminaries and Problem Setup . . . . . . . . . . . . . . . . . . 57
2.1.1 Reset Control System Solutions . . . . . . . . . . . . . . . 59
2.1.2 Characterization of Reset Intervals . . . . . . . . . . . . . 60
2.2 Zenoness, Beating, and Deadlock . . . . . . . . . . . . . . . . . . 63
2.2.1 Well-posedness: Beating and Deadlock . . . . . . . . . . . 64
2.2.2 Zeno Solutions . . . . . . . . . . . . . . . . . . . . . . . . 67
2.3 Reset Instants and the After-Reset Surface Dimension . . . . . . . 73
2.3.1 dim(MR ) = 1 . . . . . . . . . . . . . . . . . . . . . . . . 74
2.3.2 dim(MR ) = 2 . . . . . . . . . . . . . . . . . . . . . . . . 76

xv
xvi Contents

2.3.3 dim(MR ) ≥ 3 . . . . . . . . . . . . . . . . . . . . . . . . 80
2.4 Reset Control Systems with Exogenous Inputs . . . . . . . . . . . 82
2.4.1 A Well-posed Reset Control System with Exogenous Input 85
2.4.2 A Reset Control System with Zeno Solutions . . . . . . . . 86
3 Stability of Reset Control Systems . . . . . . . . . . . . . . . . . . . 93
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.2 Lyapunov Stability . . . . . . . . . . . . . . . . . . . . . . . . . . 94
3.2.1 Reset-Times Independent Conditions . . . . . . . . . . . . 94
3.2.2 Reset Times-Dependent Stability Criteria . . . . . . . . . . 99
3.2.3 Stabilization of Reset Control Systems . . . . . . . . . . . 105
3.3 Reset Control Systems with Inputs/Passivity Analysis . . . . . . . 113
3.3.1 Full Reset Compensators . . . . . . . . . . . . . . . . . . 117
3.3.2 Partial Reset Compensators . . . . . . . . . . . . . . . . . 120
3.3.3 L2 -stability of the Reset Control System . . . . . . . . . . 123
3.3.4 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
3.4 Describing Function Analysis . . . . . . . . . . . . . . . . . . . . 127
3.4.1 FORE and Clegg Integrator . . . . . . . . . . . . . . . . . 128
3.4.2 Reset Compensators with Reset Band . . . . . . . . . . . . 129
3.4.3 Limit Cycle Analysis . . . . . . . . . . . . . . . . . . . . 136
3.4.4 Justification of the Describing Function . . . . . . . . . . . 139
4 Stability of Time-Delay Reset Control Systems . . . . . . . . . . . . 147
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.2 Problem Motivation and Statement . . . . . . . . . . . . . . . . . 149
4.3 Delay-Independent Conditions in the Time Domain . . . . . . . . 153
4.4 Delay-Independent Conditions in the Frequency Domain . . . . . . 155
4.4.1 The Hβ -condition . . . . . . . . . . . . . . . . . . . . . . 156
4.4.2 The Generalized Hβ -condition . . . . . . . . . . . . . . . 158
4.4.3 Interpretation of the Stability Conditions . . . . . . . . . . 159
4.5 Example: Delay-Independent Stability . . . . . . . . . . . . . . . 162
4.6 Delay-Dependent Conditions in the Time Domain . . . . . . . . . 164
4.7 Delay-Dependent Conditions in the Frequency Domain . . . . . . 168
4.8 Example: Delay-Dependent Stability . . . . . . . . . . . . . . . . 175
5 Design of Reset Control Systems . . . . . . . . . . . . . . . . . . . . 181
5.1 The PI + CI Compensator . . . . . . . . . . . . . . . . . . . . . . 181
5.1.1 PI + CI Tuning for First Order Plants . . . . . . . . . . . . 184
5.1.2 First Order Plus Deadtime (FOPDT) Systems . . . . . . . . 188
5.1.3 High Order Systems . . . . . . . . . . . . . . . . . . . . . 195
5.1.4 Integrating Systems . . . . . . . . . . . . . . . . . . . . . 198
5.1.5 Summary of Tuning Rules . . . . . . . . . . . . . . . . . . 202
5.2 Design Improvements . . . . . . . . . . . . . . . . . . . . . . . . 202
5.2.1 Fixed Reset Band . . . . . . . . . . . . . . . . . . . . . . 203
5.2.2 Variable Reset Band/Advanced Reset . . . . . . . . . . . . 205
5.2.3 Variable Reset Percentage . . . . . . . . . . . . . . . . . . 206
5.2.4 Robust Control Design Based on QFT . . . . . . . . . . . 208
Contents xvii

6 Application Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211


6.1 Temperature Control in a Heat Exchanger . . . . . . . . . . . . . . 211
6.1.1 Process Model . . . . . . . . . . . . . . . . . . . . . . . . 212
6.1.2 PI + CI Design . . . . . . . . . . . . . . . . . . . . . . . . 216
6.2 Teleoperation of a Gantry Crane . . . . . . . . . . . . . . . . . . . 224
6.2.1 Passive Teleoperation and System Overview . . . . . . . . 225
6.2.2 Reset Procedure . . . . . . . . . . . . . . . . . . . . . . . 228
6.2.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
6.3 Control of Solar Collector Fields . . . . . . . . . . . . . . . . . . 237
6.3.1 The Solar Collector Field . . . . . . . . . . . . . . . . . . 238
6.3.2 PI + CI Design . . . . . . . . . . . . . . . . . . . . . . . . 239
6.3.3 PI + CI Compensator with Variable Reset Percentage . . . 242
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Chapter 1
Introduction

1.1 What Is Reset Control?


A reset controller, as the name suggests, is no more than a standard controller en-
dowed with a reset mechanism, that is, a strategy that resets to zero the controller
state (or part of it) when some condition holds. The reset condition (the event that
triggers the reset action) is typically the zero crossing of the controller input, but
other choices are possible as well.
To explain it with a simple example, consider the control system in Fig. 1.1. The
plant is a standard linear time invariant (LTI) system with a transfer function P (s).
For the reset controller, represented customarily with a broken box, its underlying
continuous mode of operation is given as well by a LTI transfer function C(s). But
the broken box indicates that it is a reset controller, so a reset mechanism has to be
specified.
For the sake of concreteness, suppose that the base linear system C(s) is given
by a first order transfer function
K
C(s) = , (1.1)
s +a
and thus the base linear dynamics are described by u̇ = −au + Ke. Suppose that
the reset condition is, as usual, the zero crossing of the input (e = 0) and that the
reset action is the zero reset of the controller state. Thus, the equations that define
this reset controller (a first-order-reset-element or FORE [31]) are:

u̇(t) = −au(t) + Ke(t) when e(t) = 0,
(1.2)
u(t + ) = 0 when e(t) = 0.
The first equation describes the continuous dynamics or the flow mode. The sec-
ond equation defines the discrete or impulsive dynamics, also called jump mode
because at the time instants tk (k = 1, 2, . . .) when the error crosses zero (e(tk ) = 0)
the controller state jumps from u(tk− ) to u(tk+ ) = 0.
Now let us see what happens in this reset control system, and compare its be-
havior with the purely linear base system. Let us take the same values as in [18],
K = a = 1 for the controller and P (s) = (s + 1)/(s(s + 0.2)) for the plant. The

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 1


DOI 10.1007/978-1-4471-2250-0_1, © Springer-Verlag London Limited 2012
2 1 Introduction

Fig. 1.1 Standard reset


control system formed by a
linear plant P (s) and a reset
controller with base linear
system C(s)

responses y(t) to a unit step reference r(t) applied at t = 1 are plotted in Fig. 1.2.
This figure shows that the base linear system is very underdamped, with a large over-
shoot, greater than 70%. The reset system response (dotted line in Fig. 1.2) is much
more damped, the main overshoot decreases to 40%, and the remaining oscillations
disappear very quickly.
Figure 1.3 plots the corresponding control signals, showing that the convergence
to zero is much faster for the reset system than for the linear system. This figure
illustrates more clearly the discontinuities or jumps of the reset system from u(tk− )
to u(tk+ ) = 0, where tk are the reset times.
The reset times are the instants tk when e(tk ) = 0 or, looking at Fig. 1.2, when
y(tk ) = r(tk ) = 1. This figure shows the improvement achieved with reset control:
an acceptable, bounded overshoot, but with the same speed of response (and small
rise time) than the linear system, much more underdamped. In this way, reset control

Fig. 1.2 Step response y(t) of the reset system (solid) and base linear system (dotted)
1.1 What Is Reset Control? 3

Fig. 1.3 Control action u(t) of the reset system (solid) and base linear system (dotted)

appears as a solution for overcoming the typical linear design trade-offs between
bandwidth and robustness, or other competing objectives. We will discuss in the
next section the subject of fundamental limitations and reset control.
An intuitive explanation of the benefits of reset control can be understood looking
at Fig. 1.2. The initial part of the step response of both the base linear system and
the reset system is very fast, which is a good and desirable property. But after the
first time the output reaches the reference, that is y(t1 ) = r(t1 ) = 1, and the behavior
of the linear and reset systems is completely different. The linear system has a lot
of “inertia”, the linear states are far from the steady state values y = 1, ẏ = 0, u = 0
and consequently a large overshoot appears.
On the other hand, the reset controller performs a strong resetting action from
u(t1− ) = 0 to u(t1+ ) = 0. This has the effect of “discharging” or “emptying” the con-
troller state so that the step response produces a much smaller overshoot. A similar
favorable effect appears at the second reset time t2 , and so on. The reset system
reaches steady state after the fourth reset time t4 (with a very small error) but the
linear system is still very oscillatory at t = t4 and its settling time is much worse.
Although the previous example shows more advantages in favor of reset control,
several skeptical doubts could arise: (i) It might be that there exists another linear
controller, different from C(s) = 1/(s + 1), that performs equally well or better than
4 1 Introduction

Fig. 1.4 Basic reset


integrator with reset condition
input c(t) and after-reset
value input a(t)

the reset controller, and (ii) It might be that, starting from a well-tuned linear control
system, the inclusion of a reset provokes instability.
Actually, these two grounds for concern are justified. Regarding (ii), one can
find examples of stable linear control systems that become unstable if some reset
is applied. This tells us that reset should be designed with great care, if we want to
improve system performance and robustness. Regarding (i), it is true that there exist
powerful design methods for linear controllers that easily find solutions C(s) that
satisfy the given design objectives. In control practice, reset control may perform
better, but it also may perform worse than a well designed linear compensator, and
thus in general reset should be used carefully.
The key point is that there exist some classes of linear plants P (s) that are very
difficult to control with linear controllers and are subject to linearly unsolvable
trade-offs between competing design objectives. These problematic classes of sys-
tems include plants containing integrators, right half-plane poles or zeros, or time
delays. The corresponding unsolvable trade-offs are called linear fundamental lim-
itations. These fundamental limitations are well known in the literature (see, e.g.,
[2], [27], [37]) and impose hard restrictions on the system performance. For ex-
ample, if a system has an open-loop positive zero s = c > 0, then for any linear
controller the closed-loop step response exhibits an undershoot Mu , and (see [27])
Mu ≥ (1 − δ)/(exp(cts ) − 1) where ts is the settling time into the band [1 − δ, 1 + δ].
So we cannot make ts and Mu (good speed and robustness) simultaneously small,
and things get worse as c → 0. These limitations are derived assuming linear plants
and controllers, thus in principle it might be possible that a nonlinear controller (like
a reset controller) overcomes the limitation and outperforms any linear solution.
This is the main point that justifies the use of reset control. In the following section,
the relations between fundamental limitations and reset control will be presented in
more detail.
Before that, let us look at some other possible reset controllers in order to gain
insight into certain dynamical aspects of reset systems. This presentation will be
intuitive and informal; the formal treatment of well-posedness, analysis and design
will be addressed later in this book. The basic block for simulation or block diagram
realization of reset systems is presented in Fig. 1.4. In the same way as standard
integrators are basic dynamic blocks for realization of finite-dimensional systems,
the reset systems require two extra inputs apart from the usual input e(t), the signal
that is being integrated. As Fig. 1.4 shows, there is a second input c(t), the reset
condition, that has the effect that the zero crossings of a signal c(t), that is, c(tk ) = 0,
define the reset instants t = tk (k = 1, 2, . . .) and trigger the reset action. The third
input a(t) defines the after-reset value to be applied at the output: u(tk+ ) = a(tk ).
1.1 What Is Reset Control? 5

Fig. 1.5 Simulation of a reset integrator: e(t) is the standard input, c(t) gives the reset condition,
and a(t) the after-reset value for the output u(t)

This is a basic block that can be built, for example, in MATLAB-SIMULINK, or


can be programmed easily in any other environment. Formally, the equations of this
basic reset integrator are:

⎨ u̇(t) = e(t) when c(t) = 0,
+
u(t ) = a(t) when c(t) = 0, (1.3)

u(0) = u0 .
If the input signals are sufficiently regular (forget for the moment possible dif-
ficulties arising from discontinuities or pathologies in them) then the simulation of
this reset integrator is very easy. To illustrate the role of each signal, Fig. 1.5 plots
a simulation example: the output u(t) integrates the main input (a slow sinusoidal),
but resets the output value at the zero crossings of c(t), to the after-reset values given
at those moments by a(t). The initial condition is u(0) = 0.
Normally, the after-reset value a(t) is 0. Only in special cases is this third input
required, for example, in partial reset systems, that is, when u(tk+ ) = p · u(tk− ) for
some number p, called the percentage of reset. The realization of this strategy re-
quires the use of the third input, and some easy arrangement for circumventing the
algebraic loop a = pu. When the after-reset input is omitted, the default value a = 0
is assumed and the reset becomes zero reset. Furthermore, if we identify the two in-
6 1 Introduction

Fig. 1.6 Clegg integrator

puts c(t) = e(t), that is, the reset instants appear at the zero crossings of the same
signal that is integrated, then we reach the well-known reset integrator or Clegg
integrator [22]. Figure 1.6 shows the block representing the Clegg integrator.
The Clegg integrator is simply given by

u̇(t) = e(t) when e(t) = 0,
(1.4)
u(t + ) = 0 when e(t) = 0

with u(0) = u0 . Figure 1.7 shows the response of the integrator to a sinusoidal input,
and with u0 = 0.
Using Clegg integrators, every finite-dimensional reset system can be built in the
form of a block diagram realization based on this primitive block. For example, the
FORE system in (1.2) can be represented by the block diagram in Fig. 1.8.

Fig. 1.7 Simulation of a Clegg integrator for a sinusoidal input


1.1 What Is Reset Control? 7

Fig. 1.8 Block diagram of


the FORE
(first-order-reset-element)
with base linear system
K/(s + a)

Clegg integrators and FOREs are very simple, low-order, particular cases of reset
controllers. A general reset controller with linear base dynamics can be described
by the set of equations:

⎪ ẋ(t) = Ax(t) + Be(t) when e(t) = 0,

⎨ x(t + ) = A x(t)
ρ when e(t) = 0,
(1.5)

⎪ u(t) = Cx(t) + De(t),

x(0) = x0
where e(t), u(t) are the input and output signals, and x(t) ∈ Rn is the state vector.
The base linear system of (1.5) is given by the transfer function C(s) = C(sI −
A)−1 B + D, defined by the first and third equation, describing the flow mode and
holding almost all the time, when e(t) = 0. The impulsive dynamics, or the jump
mode, is triggered by e(t) = 0. If some of the states are affected by reset and some
are not, we can, without loss of generality, reorder states, for example, non-reset
states first and reset states last. In this way, the reset matrix Aρ takes the form
Aρ = diag(1, . . . , 1, 0, . . . , 0) (1.6)
 
 

n1 n2
with n1 + n2 = n. If n2 = 0, we have a linear controller. If 0 < n2 < n, we have
a so-called partial reset controller and only part of the states are reset states. If
n2 = n, the controller is called full reset and x(tk+ ) = 0 ∈ Rn . The FORE in Fig. 1.8
is a particular case of a full reset system with scalar parameters (A, B, C, D, Aρ ) =
(−a, K, 1, 0, 0).
In summary, to conclude this section, a reset controller, or a reset system, is a
dynamical system (typically linear and finite-dimensional) endowed with a reset
mechanism. The reset mechanism has to be carefully designed by answering two
questions:
• When to apply reset?—Reset condition This amounts to defining the reset con-
dition input c(t) in Fig. 1.4 so that c(tk ) = 0 defines the reset instants tk (k =
1, 2, . . .). Usually, it is made equal to the tracking error of the control system
c(t) = e(t), but other modifications are possible (combinations of e(t) and ė(t),
deadbands or thresholds for e(t), etc.).
• How to apply reset?—Reset action This amounts to defining the after-reset value
signal a(t) in Fig. 1.4 so that u(tk+ ) = a(t). Normally, reset is applied as zero
reset, a(t) = 0. Sometimes it is useful to perform only a percentage of reset,
a(t) = pu(t) so that u(tk+ ) = p(tk )u(tk ) with 0 < p(t) < 1.
The designer has to choose judiciously the reset conditions and reset actions
in order to improve the system performance, and also in order to ensure well-
posedness. The well-posedness could be lost through several causes: algebraic
8 1 Introduction

loops, accumulation of reset instants tk → t∞ < ∞, etc., giving rise to different


pathologies (deadlock, livelock, Zenoness). There are simple recipes (called reg-
ularizations) to eliminate these anomalies. These well-posedness aspects will be
discussed in the following chapter. Now that reset systems are introduced, let us see
in the following section the main reasons why these type of systems are useful in
control.

1.2 Why Reset Control?


The main reason for using reset controllers is that, just by including the mechanism
of resetting, they are able to overcome fundamental limitations in linear systems.
This idea was present from the very beginning: Clegg’s integrator [22] was intro-
duced as a solution for improving feedback performance, due to its ability to provide
the same modulus attenuation as a linear integrator (−20 dB/dec) but with a phase
(≈ −38°) much more favorable in terms of phase margins and robustness. This abil-
ity for overcoming linear limitations was also implicit in the subsequent proposals of
reset control: from the first-order-reset-element (FORE) by Horowitz and coworkers
(see [31], [32]) to the many recent generalizations of reset controllers.
In spite of being the key reason for using this class of controllers, the literature
that specifically studies the overcoming of design limitations by reset controllers is
very scarce: only a few works show explicitly certain linear limitation and how the
reset controller outperforms it. On the other hand, the topic of linear fundamental
limitations has been widely treated in the literature, and there is a large number of
references that quantify linear limitations in the time and frequency domains; see,
for example, [2], [37], [27]. It is out of the scope of this book to give a detailed
presentation on the topic of fundamental limitations, the interested reader can find
more information in the aforementioned references.
The purpose of this section is to give a brief glimpse of the problem, by pre-
senting a representative choice of some time domain and frequency domain linear
trade-offs, and then to motivate the use of reset controllers by showing how they can
solve the trade-offs and outperform the linear solutions. A complete study, quanti-
tative and systematic, on how reset control outperforms linear limitations deserves
further research. We provide here only a presentation based on examples, to give the
reader the main idea and to justify the interest on reset control. Before proceeding
with the section, two observations have to be made.
First, notice that, to overcome linear limitations, other hybrid approaches apart
from reset are possible [25]. We only claim here that reset control is an efficient and
simple choice among other hybrid control strategies. Second, as concluded in [17],
the fact that certain reset controllers overcome some linear limitations could intro-
duce as a drawback a new different set of limitations, specific to reset control. This is
true if we consider standard reset controllers (that reset to zero when tracking error
is zero). But if we consider general reset controllers (that can be implemented with
the reset integrator in Fig. 1.4) the freedom in the choice of the reset condition and
reset action enlarges the design possibilities. Actually, the set of all ‘generic’ reset
1.2 Why Reset Control? 9

Fig. 1.9 Standard linear


control system

controllers contains as particular case the subset of linear ones (obtained putting, for
example, c(t) = 1 ∀t, so that reset condition is never met). In this way, reset con-
trollers form a superset, or extension, of linear ones, so the set of achievable design
objectives is obviously larger. A recent rigorous study on performance improvement
by means of reset can be found in [1].

1.2.1 Overcoming Fundamental Limitations in the Time Domain

The linear fundamental limitations of control systems are classified into time do-
main and frequency domain limitations. The systems that are ‘hard to control’ (hav-
ing right half-plane poles or zeros, or time delays) present simultaneously both types
of limitations, but it is customary to address the limitations separately, studying the
restrictions on the achievable time responses or frequency responses. Let us start
with time domain. First, consider the standard (one degree of freedom) feedback
control system in Fig. 1.9, with the sensitivity S(s) and complementary sensitivity
T (s) closed-loop transfer functions given by
1 C(s)P (s)
S(s) = , T (s) = , (1.7)
1 + C(s)P (s) 1 + C(s)P (s)
and satisfying S(s) + T (s) = 1. It follows that −S(s) is the closed-loop transfer
function from noise n(t) to tracking error e(t), and T (s) is the closed-loop transfer
function from reference r(t) to plant output y(t).
Leaving aside the restrictions arising from actual limitations in the real physical
components (for example, precision and range of sensors or actuators), there are
structural limitations affecting the closed-loop in Fig. 1.9, appearing even in ideal
or nominal conditions. A detailed exposition of linear limitations can be found in
[27], [37]. We present here a selection of results taken from [27, Sect. 8.6].
These restrictions arise from interpolation constraints, that is, from transmission
of poles or zeros from the open-loop transfer functions P (s) or C(s) to the closed-
loop ones. These roots will be called uncanceled poles or uncanceled zeros. It can
be seen that certain types of uncanceled poles or zeros transmitted to closed-loop
impose severe restriction on certain closed-loop time responses.
Notice that the attempt to cancel a plant pole or zero (with a controller zero or
pole, respectively), when lying in the right complex half-plane, is impossible with
perfect accuracy, and hence totally unacceptable as it leads to instability in practice.
On the other hand, although such approximate cancellation in the left half-plane
10 1 Introduction

does not produce instability, in [35] it is shown that there is a trade-off between in-
put disturbance rejection (better when not canceling but shifting the slow undesired
poles) and robustness (which is better when canceling and not shifting the poles).
According to [27] (Sect. 8.6), we have the following interpolation constraints in
the sensitivity functions:
(i) The complementary sensitivity T (s) has a zero at all uncanceled zeros of P (s).
(ii) The sensitivity S(s) has a zero at all uncanceled poles of P (s).
These two basic properties can be particularized to open-loop integrators, open-
loop general poles or zeros, or imaginary-axis poles and zeros. For concreteness, let
us see what happens with open-loop general poles and zeros, reproducing part of
Lemma 8.3 in [27]:

Proposition 1.1 Consider the feedback system in Fig. 1.9, with all the closed-loop
poles to the left of −α < 0. Assume that the controller has at least one pole at the
origin. For an uncanceled plant zero z0 or plant pole η0 satisfying Re(z0 ) > −α or
Re(η0 ) > −α, and for a unit reference step, we have the following:

(i) 0 e(t)e−z0 t dt = z10 ;

(ii) 0 e(t)e−η0 t dt = 0;

(iii) 0 y(t)e−z0 t dt = 0.

The following observations are consequences of the previous constraints [27].


First, if −α < z0 < 0, that is, z0 is an uncanceled minimum phase zero, then the
error must change sign because the initial value of the integral (i) is positive (since
e(0) = 1) and the total integral (from 0 to ∞) is negative. In other words, we have
the following linear limitation:
L1 Any uncanceled minimum phase zero z0 < 0, to the right of the closed-loop
poles, implies overshoot in the step response.
A similar discussion in the case z0 > 0 leads to the following limitation:
L2 Any uncanceled non-minimum phase zero gives rise to undershoot in the step
response and, if z0 > 0 is small, to a positive large value of the error integral (i).
Finally, from (ii):
L3 Any uncanceled pole η0 , to the right of the closed-loop poles, must produce
overshoot in the step response.
Using the previous limitations,
∞introducing
t ∞a suitably defined settling time ts , parti-
tioning the integral in (i) as 0 = 0s + ts , and manipulating bounds, more quan-
titative results are reached (see Lemmas 4.2 and 4.3 in [27] or Proposition 3 in
[35]) showing that there are restrictions affecting the overshoot Mp (or undershoot
Mu ) and the settling time ts . In other words, if we make good ts (small ts and fast
response) then the overshoot Mp (or undershoot Mu ) becomes bad, and vice versa.
It can be shown that reset control systems provide a very simple mechanism
for overcoming the previous linear limitations. The results that follow are based
1.2 Why Reset Control? 11

on [17], where a simple example is presented, showing that reset control does not
suffer from the time-domain limitations of linear controllers. So, consider as in [17]
a linear feedback system of Fig. 1.9 with a plant:
1
P (s) =
s
and a controller C(s) that stabilizes the closed loop. It is known that the tracking
error e(t) to a unit step reference r(t) = 1 ∀t ≥ 0 satisfies:

1
e(t) dt = ,
0 Kv
where the velocity gain is given by Kv = lims→0 sC(s)P (s). Now, consider the
definition of rise time tr given by

t
tr = sup τ : y(t) ≤ ∀t ∈ [0, τ ] .
τ τ
Then [17] proves that
2
tr > ⇒ the unit step response y(t) overshoots. (1.8)
Kv
From this result, [17] derives a set of design objectives impossible to meet with
linear controllers:
Design problem: Given the plant P (s) = 1/s, obtain a controller that stabilizes
the closed loop and has
(i) steady-state error to unit ramp no greater than 1,
(ii) rise time of y(t), in response to unit step, greater than 2, and
(iii) no overshoot in the step response of y(t).
From (1.8) it is clear that no stabilizing linear controller C(s) exists that fulfills
all these specifications [17]. At the same time, [17] presents a very simple reset
controller (FORE) that achieves these specifications:

u̇(t) = −u(t) + e(t) when e(t) = 0,
(1.9)
u(t + ) = 0 when e(t) = 0.
Figure 1.10 plots the unit step response of the reset control system (top), showing
how rise-time and overshoot specifications are met (ramp steady state error is met
as well). For comparison, the bottom plot in Fig. 1.10 shows the response of the
base linear control system, with C(s) = 1/(s + 1), that, according to (1.8), must
overshoot and is subject to this linear limitation.

1.2.2 Overcoming Limitations in the Frequency Domain

The previous section has briefly presented some fundamental limitations in the time
domain. They imply constraints in the achievable time domain specifications (re-
strictions on overshoot, rise time and settling time, etc.) that cannot be solved with
linear controllers.
12 1 Introduction

Fig. 1.10 Example in [17]. Unit step responses: (top) reset control and (bottom) linear control

Likewise, this section introduces their frequency domain counterparts, and shows
how reset controllers can overcome these limitations. The presentation is based on
[2] and [3]. The reader can also consult [27] and, for a more detailed treatment, [37].
The frequency domain limitations, for the typical control system in Fig. 1.9, arise
from the structure of the sensitivities S(s) and T (s) (1.7) that impose certain restric-
tions on the modulus or phase of these complex-valued functions. Some restrictions
are consequences of the Bode’s integral formulas, and some others are related to the
Bode’s gain–phase relation.
Let us see first the Bode’s integral formulas and their effect on performance lim-
itations in the frequency domain [3].
Proposition 1.2 Consider the feedback system in Fig. 1.9. Assume that the loop
transfer function L(s) = C(s)P (s) satisfies
lim sL(s) = 0.
s→0

Then the sensitivity function S(s) = 1+L(s)


1
satisfies the integral identity
∞ 
log |S(iω)| dω = π pk , (1.10)
0
where the summation is over all right half-plane poles of L(s).
1.2 Why Reset Control? 13

Equation (1.10) implies, following [3], that the sensitivity function S(iω) has to
distribute is attenuation over different frequencies. Recall that |S(iω)| is the closed-
loop frequency attenuation from noise n(t) to error e(t) (or from disturbance d(t) to
plant input) in Fig. 1.9. So if S(iω) is made smaller for some frequencies, then nec-
essarily is has to become larger at other frequencies, in order that the total integral
in (1.10) remains constant.
This means [3] that if disturbance attenuation is improved in one frequency range,
then it will be worse at another. This property is known as the waterbed effect. The
total integral in (1.10) is greater (worse attenuation ability) as the number of right
half-plane poles pk increases, or as their values pk increase, so that fast unstable
poles are worse than slow ones, as intuitively predicted. In the best case, when L(s)
has no poles in the right half-plane, we have:

log |S(iω)| dω = 0. (1.11)
0

The Bode’s integral limitation can be given a nice geometrical interpretation [3]
as shown in Fig. 1.11 for the case L(s) = 10/((s + 2)(s − 1)). The figure plots
log |S(iω)| versus ω in linear scale. The horizontal line log |S(iω)| = 0 is also plot-
ted as a reference. Since L(s) has a right half-plane pole at p = 1, then the total

integral in (1.10) is equal to π pk = π . Thus, the area of the curve above the
horizontal line log |S(iω)| = 0 (positive contribution to the integral) minus the area

below the horizontal line (negative contribution) must be equal to π pk = π .
This means that if we want to make |S(iω)| < 1 at lower frequencies (which
is typically desired for disturbance attenuation) then the appearance of disturbance
amplification |S(iω)| > 1 is unavoidable at higher frequencies in order to balance
the integral constraint (1.10) (waterbed effect). Even in the best case (1.11) with
no poles of L(s) in the right half-plane, the negative value of the integral at low
frequencies has to be compensated by a positive value at high frequencies.
If we consider the complementary sensitivity function T (s) = L(s)/(1 + L(s)),
there exists as well a similar Bode’s integral formula [3]:
∞ 1
log |T (iω)|
dω = π , (1.12)
0 ω2 zi

where the summation is taken over all the right half-plane zeros of L(s).
Now, an alternative to the Bode’s integral constraints, in order to make evident
the fundamental limitations of linear design in the frequency domain, is to consider
a well-known restriction affecting minimum phase systems: the Bode’s gain–phase
relation. In what follows, an exposition on fundamental limitations derived from
the Bode’s gain–phase relation is presented based on the results in [2] (see also [3]).
Subsequently, some examples will be given, showing that reset control systems are
able to overcome these limitations and outperform linear controllers.
14 1 Introduction

Fig. 1.11 Waterbed effect: plot of log |S(iω)| versus ω for L(s) = 10/((s + 2)(s − 1))

Proposition 1.3 (Bode’s gain–phase relation) Assume that the transfer function
G(s) is minimum phase, that is, it does not have poles and zeros in the right half-
plane. Then the phase is uniquely given by the gain curve by means of the relation

π ∞ d log |G(iω)|
arg G(iω0 ) = f (ω, ω0 ) d log ω, (1.13)
2 0 d log ω
where the weighting kernel is
 
2  ω + ω0 

f (ω, ω0 ) = 2 log  .
π ω − ω0 

Notice that the kernel weights the frequencies close to ω, for if ω → ω0 then
f (ω, ω0 ) → ∞, and if ω → ±∞ then f (ω, ω0 ) → 0. From this interpretation of
the weight, it can be seen that the relation (1.13) gives rise to the approximate Bode’s
gain–phase relation, for minimum phase systems:
π d log |G(iω)|
arg G(iω0 ) ≈ . (1.14)
2 d log ω
If we define
d log |G(iω)|
n(ω) := (1.15)
d log ω
1.2 Why Reset Control? 15

as the slope of the gain curve, in logarithmic scale, then the approximate Bode’s
gain–phase relation, for minimum phase systems, says:
π
arg G(iω0 ) ≈ n(ω). (1.16)
2
In other words, if G(iω) at a given point ω has −20 dB/dec (respectively, −30,
−40, . . .dB/dec), then we can predict an approximate phase of −π/2 rad (respec-
tively, −3π/4, −π, . . . rad), a property that is well-known when sketching Bode
plots by hand.

1.2.2.1 Limitations Derived from the Approximate Gain–Phase Relation

In this subsection, we present a brief exposition, based on [2] and [3], on frequency
domain limitations specifically derived from the approximate relation (1.16). After
that, it will be shown, following [24], that reset control systems are able to overcome
these limitations.
If a plant P (s) has poles or zeros in the right half-plane, or time-delays, it is called
a non-minimum phase system and it is subject to some fundamental limitations.
Following [2], let us factor the loop transfer function L(s) = C(s)P (s) of one such
system as L(s) = Lm (s)La (s), where Lm (s) is the minimum phase part and La (s)
the non-minimum phase part. We normalize this factorization so that |La (iω)| = 1
(all-pass).
If the non-minimum phase factor comes from a pure time delay h, we can identify
La (s) = e−sh . If the plant has one non-minimum phase zero, L(s) = L1 (s)(z − s)
with z > 0, then we can factor L(s) = Lm (s)La (s) with Lm (s) = L1 (s)(z + s) and
the all-pass factor
z−s
La (s) = .
z+s
Likewise, for an open-loop pole in the right half-plane s = p > 0, the all-pass
factor is La (s) = (s + p)/(s − p). If the open-loop has several poles or zeros in
the right half-plane, or time delays, then the all-pass factor La (s) contains several
all-pass factors as the previously presented ones. Following [2], we characterize the
bandwidth by the gain crossover frequency ωc as given by
|L(iωc )| = |Lm (iωc )| = 1
(since La is all-pass). Let us introduce here
d log |L(iω)| d log |Lm (iω)|
n(ω) = =
log ω log ω
and let nc = n(ωc ) be the ‘crossover slope’. Two measures of robust stability are
given by the guaranteed phase margin φm and by the crossover slope nc . The nega-
tive feedback loop formed around L(s) is more robustly stable as φm is larger (60°
better than 45°) and nc is more negative (−20 dB/dec better than −10 dB/dec). Two
relations hold:
16 1 Introduction

arg L(iωc ) ≥ −π + φm , (1.17)


π
arg Lm (iω) ≈ n(ω) . (1.18)
2
The first inequality ensures the guaranteed phase margin φm , and the second one
is the Bode’s relation. If we use arg L = arg Lm + arg La and Bode’s relation (with
an equality = in place of ≈), we reach
π
− arg La (iωc ) ≤ π − φm + arg Lm (iωc ) = π − φm + nc . (1.19)
2
This last relation,
π
− arg La (iωc ) ≤ π − φm + nc =: c(φm , nc ), (1.20)
2
is called the design inequality in [2]. If we particularize La (iωc ), we typically obtain
restrictions on the crossover frequency ωc that limit in some way the achievable
bandwidth of the linear control system.
For example, if the plant has a non-minimum phase zero and the all-pass factor
is La (s) = (z − s)/(z + s) then, from arg La (iωc ) = −2 arctan(ωc /z), we reach
   
π − φm + nc π/2 c(φm , nc )
ωc ≤ z tan = z tan . (1.21)
2 2
Notice that the limitation imposed by the coefficient c(·,·) becomes more restrictive
as the robustness requirements (φm , nc ) are stronger. For example, if (φm , nc ) =
(45°, −10 dB/dec), then c = π/2 ≈ 1.57, but if (φm , nc ) = (60°, −20 dB/dec) then
c = π/6 ≈ 0.52. The restriction (1.21) also becomes stronger as the zero approaches
the origin (slow non-minimum phase zeros are bad) [2].
Similar limitations could be derived from unstable open-loop poles or from time
delays, see [2]. Therefore, the fact that a plant contains right half-plane poles or
zeros, or time delays, makes it very difficult to control these types of systems, at
least if linear controllers are used.

1.2.2.2 Overcoming Frequency Domain Limitations with Reset

How can a reset control contribute to solving this problem? In this subsection, based
on [24] and [42], we present briefly some examples of reset systems that are not
subject to the limitations for linear controllers. Checking the previous steps of the
inequalities (1.19) and (1.20), it follows that the limitation could be alleviated if we
could replace arg Lm (iωc ) by a quantity less negative (larger) than nc π/2.
Unfortunately, no linear system is able to solve this, minimum phase linear sys-
tems Lm (s) are restricted to the approximate Bode relation (1.16). So the only way
to achieve a significative reduction of the limitation is by using some new type of
nonlinear system Lr (the reset controller). It will be useful to define an approximate
frequency response Lr (ω) of Lr (describing function analysis could provide this
approximation of the frequency response, a detailed analysis of reset control system
with the describing function will be given in Chap. 3).
1.2 Why Reset Control? 17

The reset system Lr can be used in place of Lm , or combined in series with


it Lr (ω)Lm (iω). Since the angles and decibel slopes nr , n of the systems Lr (ω),
Lm (iω) behave additively, it follows that the Bode’s relation (1.19) is drastically
relaxed (and the limitation outperformed) if
π
arg Lr (ω) nr (ω) , (1.22)
2
at least at the frequency ω = ωc .
Fortunately, it is well documented that a variety of reset controllers (FORE,
Clegg, etc.) satisfy the previous relation. For example, the Clegg integrator (1.4)
LCI (an integrator with reset) has an approximate describing function
1.62 −i38°
LCI (ω) = e , (1.23)
ω
that is, arg LCI (iω) = −38° and nCI (ω) = −20 dB/dec, which means a valuable
improvement of about 52° in phase lag, compared to the linear integrator, with phase
lag −90°.
Consider a very simple example: Let us compare two negative feedback loops,
the first one containing a standard integrator Lm (s) = 1/s and a pure delay La (s) =
e−sh ; and the second one with a Clegg integrator Lr = LCI and a delay La (s) =
e−sh as well. In both cases, the modulus slope has to be nc = −20 dB/dec.
Let us particularize (1.19), with φm = π/3 = 60°, having in mind that
arg La (iω) = arg e−iωh = −ωh. Thus we reach
π π
ωc h ≤ π − − ≈ 0.52 with linear integrator, (1.24)
3 2
π π
ωc h ≤ π − − 38° ≈ 1.43 with Clegg integrator. (1.25)
3 180°
Thus, for the same phase margin and delay, the reset system outperforms the
linear one. The bandwidth limitation ωc h ≤ 0.52 is improved to ωc h ≤ 1.43 by a
factor 1.43/0.52 = 2.75 (almost three times larger) provided by reset control.
Going further with this idea, one can think if it is possible to improve the Clegg
integrator, providing also −20 dB/dec but with a phase lag still less negative than
−38°. One simple idea for achieving this is replacing the reset condition e(t) = 0 by
a new condition u(t) = 0 where u(t) is obtained passing e(t) through some block
Fc (s) providing phase advance. Thus, consider now the Clegg Integrator with ad-
vanced reset [24] (or with variable reset band [42]):
ẋ(t) = e, u(t) = 0,
+
x = 0, u(t) = 0, (1.26)
u(t) = L −1 [Fc (s)E(s)] ,

where L , L −1 are the direct and inverse Laplace transforms, and Fc (s) is an an-
ticipative block, for example, Fc (s) = Kp + Kd s, which introduces a phase lead or
phase advance on e(t) (if Fc (s) ≡ 1 then u = e, and we recover the standard Clegg
18 1 Introduction

Fig. 1.12 Pure time-delay


plant P (s) stabilized by:
Advanced Clegg integrator
(top), Clegg integrator
(center), and standard
integrator (bottom)

Integrator). The describing function of an Advanced Clegg Integrator with a phase


advance of φ = φ(ω) = arg Fc (iω) is:
  
1 4 cos2 φ 4 cos φ sin φ
LACI (ω) = −i 1− . (1.27)
ω π π
The particular case when φ = 0 stays constant recovers the standard Clegg inte-
grator (1.23). If φ = 0 is held approximately constant over a frequency range then
|LACI | = k/ω, and we obtain −20 dB/dec as the standard integrator. The key fea-
ture is the behavior of the phase, the plot of arg LACI versus φ reveals that the less
negative phase is obtained for φ ≈ +30° giving arg LACI ≈ −25°.
Thus, by designing the phase lead block Fc (s) so that it provides φ ≈ +30° along
the frequency range of interest (frequency where gain margin is defined), we will
obtain LACI ≈ ω1 e−i25° , which is the same attenuation as that of an integrator but
with 90° − 25° = 65° of phase improvement, a very important advantage, regarding
fundamental limitations.

Example 1.1 To conclude this section, let us translate the ideas above into a simple
but quantitative design problem. Let us consider as shown in Fig. 1.12 a one-second
time-delay plant P (s) = e−s and three possible controllers with adjustable gain: an
integrator, a Clegg integrator, and an advanced Clegg integrator. These are not ex-
amples of real reset control systems that work well in practice because among other
reasons the reset action destroys the steady state property for tracking external refer-
ences. Other real examples of useful practical application of reset will be presented
later in the book. So, consider that we are only interested in the stability of the zero
solution, and how restrictive the trade-off between speed (ωc ) and robustness (φm )
is in each of the three cases.
If the design objective is chosen to be a gain crossover frequency ωc of, say,
4π/9 rad/s = 80°/s, then it is easy to see that the corresponding gains for the inte-
grator and Clegg are respectively k = kI = 4π/9, k = kCI = (4π/9)/1.62, and the
resulting phase margins φm are respectively 10° and 62°.
1.2 Why Reset Control? 19

Fig. 1.13 Bode plots of the delay and three integrators example

On the other hand, to take full advantage of the advanced Clegg, one has to
choose an adequate lead filter Fc (s), for example, a PD: Fc (s) = Kp + Kd s. As
Fc (s) defines the signal so that its zero crossings induce resetting, Fc (s) can be nor-
malized as Fc (s) = 1 + sT . The phase advance is φ(ω) = arg(1 + j ωT ). As the
maximum advance is at φ = 30° = π/6, and we would like to exploit this advance
at ωc (for maximizing the phase margin), the best choice is φ(ωc ) = 30° which is
√ √
solved by ωc T = 1/ 3 implying T = 9/(4 3π). Having fixed the lead compen-
sator Fc (s), it is easy to obtain the gain kACI = 1.32 achieving ωc = 4π/9 crossover
frequency.
Figure 1.13 plots the resulting open-loop responses of the solutions with a stan-
dard integrator, a Clegg, and an advanced Clegg. All the three achieve the same
crossover frequency ωc = 4π/9 = 1.4 rad/s, but the corresponding margins are re-
spectively 10°, 62° and 75°, showing the advantages provided by (advanced) reset-
ting for overcoming fundamental trade-offs. The corresponding time plots, with a
zero reference and x(0) = 1, are shown in Fig. 1.14.
20 1 Introduction

Fig. 1.14 Time plots of the delay and three integrators example

1.3 Early Ideas on Reset Control


In this section, we present a brief glimpse of the early ideas on reset control. It is
based on [21] that reviews the contributions to reset control of Prof. I. Horowitz,
and it is based also on [32] and [31] by Horowitz and coworkers.
It is recognized [21] that reset controllers began with the Clegg integrator (Clegg,
1959) [22]; however, it was not until the work of Horowitz that a quantitative de-
sign procedure was developed, first around the Clegg integrator in Krishnan and
Horowitz (1974) [32], and then around the first-order-reset-element (FORE) in
Horowitz and Rosenbaum (1975) [31]. Since the final design procedures in [32]
and [31] are strongly based on Quantitative Feedback Theory (QFT), which is be-
yond the scope of this book, this section is focused on the ideas by Horowitz and
coworkers regarding the Clegg and FORE controllers: its behavior, interpretation,
analysis, and features.

1.3.1 Reset Control with the Clegg Integrator

In [32], Krishnan and Horowitz proposed for the first time a complete and quantita-
tive control design procedure based on the Clegg integrator. The block diagram of
1.3 Early Ideas on Reset Control 21

Fig. 1.15 (Top) The two


degree of freedom structure in
Krishnan and Horowitz [32];
(bottom) structure for G with
linear integrator in parallel
with Clegg integrator (CI)

the considered control system is shown in Fig. 1.15 (top), where we can see a two-
degree-of-freedom structure with prefilter F (s) acting on the reference r(t), output
disturbance d(t), and measurement noise n(t). The plant (possibly including linear
factors from the controller) is given by P (s), and G is the controller that could be
linear or nonlinear.
The justification of the use of reset control is presented in [32] as a simple mecha-
nism to alleviate the limitations of linear design. More precisely, the effect of sensor
noise n(t) at the plant input u(t) in Fig. 1.15 (top) is given, in the linear case, by
U (s) G(s)
B(s) := =− ,
N (s) 1 + L(s)
where L(s) = G(s)P (s). Typically, L(s) is designed such that |L(j ω)| 1 over
a sufficiently large frequency range. This may require |G(j ω)| 1 over a larger
frequency range. As a consequence, large G in the range where L is small (and
1 + L ≈ 1) causes, from the above equation, large B and amplification of noise
tending to saturate the plant.
Hence, it is desirable to attenuate L as fast as possible outside the frequency
range where specifications require large L. However, from Bode gain–phase re-
lation, a fast attenuation of |L| is accompanied by a very negative phase, which
compromises stability.
These linear limitations justify the use of a nonlinear element whose describing
function has a smaller phase than that of a linear element with the same magnitude
characteristic, and such a nonlinear element is the Clegg integrator. More precisely,
Krishnan and Horowitz propose as nonlinear controller G in Fig. 1.15 (top), the
configuration in Fig. 1.15 (bottom), formed by a linear integrator b/s in parallel
with the Clegg integrator (CI), and a linear compensator G(s). It was well known
that the describing function of the CI is (see Sect. 3.4):
1.62e−j 38° 1.62 j 52°
CI (j ω) := = e ,
ω jω
that is, the same magnitude slope as an integrator, but with phase −38°, which
means 52° less than that of the linear integrator. One simple example shows the
benefit of a Clegg integrator [32]: Consider an unforced typical negative feedback
loop with two linear integrators L(s) = (1/s)(1/s). Then, the autonomous response
22 1 Introduction

is oscillating and does not converge to zero. On the other hand, if either of the
integrators is replaced by a Clegg integrator, that is, L = (1/s) ◦ (CI ) or L = (CI ) ◦
(1/s), then the system is stable and all initial conditions tend to zero in finite time.
Notice that in order that the sinusoidal describing function CI (j ω) is represen-
tative of the approximate response of the CI, the input should have a mean value of
zero. For this reason, [32] proposes the parallel configuration in Fig. 1.15 (bottom),
containing a linear integrator b/s.
Another interesting interpretation of the dynamic response of the Clegg integrator
is given by Krishnan and Horowitz in [32], and consists of noticing that it can be
understood as a linear integrator which is affected by an adequate train of corrective
Dirac impulses, as an additional input. More precisely, let 0 ≤ t1 < t2 < · · · < tn be
the zero crossings of e(t), the input to the CI. Then x(t), its output, is
t t  n tk+1
x(t) = e(τ ) dτ = e(τ ) dτ − e(τ ) dτ.
tn 0 k=1 tk
tk+1
Hence, if we define xk := x(tk ) = tk e(τ ) dτ , we reach
 t 
n

x(t) = e(τ ) − xk δ(t − tk ) dτ. (1.28)
0 k=1

Thus the Clegg integrator can be replaced by a linear integrator corrected by a


train of Dirac impulses as additional input, the weight of the impulse xk being the
value of the CI output at tk . This is a very interesting interpretation proposed in
[32]. This idea will be exploited later in this book, in Chap. 4 on stability analysis
of time-delay systems under reset control.
Another important contribution is the analysis of the noise response of the Clegg
integrator and its application to the design procedure. The motivation for this noise
analysis is derived from the formulation of the design problem: it consists of a de-
terministic and a stochastic objective. The deterministic objective is: Given the plant
and the range of parameter uncertainty, find feasible controllers that satisfy specifi-
cations in the form of bounds on the responses to command and disturbance inputs.
The stochastic objective is in an optimal sense: Among all previous feasible con-
trollers, find the one which minimizes the rms value of the noise appearing at the
input of the plant in Fig. 1.15 (top), originating from sensor noise n(t).
Let x(t), y(t) be the input and output of the Clegg integrator, and let v(t) = t − tk
the interval between t and the immediate previous zero crossing tk :
t
y(t) = x(τ ) dτ.
t−v(t)

If x(·) is a stationary random process, the interval v(t) is itself a stationary random
process. The computations of the statistics of v(·) as a function of the statistics of
x(·) have been reported only for a few processes. If x(·) is a zero mean Gaussian
Markov process with autocorrelation
Rxx (τ ) = ae−α|τ | ,
1.3 Early Ideas on Reset Control 23

then the probability density function of v is


2αe−αv
p(v) = √ ,
π 1 − e−2αv
implying the mean value v̄ = log 2/α and variance σv2 = π/(12α 2 ). Thus for highly
uncorrelated processes (α → ∞) the variance tends to zero, and it seems reasonable
to replace v(t) by v̄ in the integral defining the CI output which leads, among other
results, to
Syx (ω) 1 − e−j ωv̄
= , (1.29)
Sxx (ω) jω
the right-hand side being equal to the random describing function of the CI. These
results on noise analysis are extended to the parallel of CI with an integrator in
Fig. 1.15 (bottom), and finally to the noise response of the whole loop. In this way,
Krishnan and Horowitz present a complete design procedure which clearly reveals
that the optimum nonlinear solution with a Clegg integrator (the feasible controller
that minimizes the rms value of noise at the plant input) outperforms the optimum
solution among only linear controllers.

1.3.2 Reset Control with the First-Order-Reset-Element (FORE)

The analysis and design procedures were subsequently extended by Horowitz and
Rosenbaum in [31] to the analysis and design of reset control systems with a first-
order-reset-element (FORE). Since the approach is similar, regarding general justi-
fications and ideas, and these aspects have already been summarized in the previous
subsection, we will describe here only the basic design procedure.
To facilitate this description we will follow the report by Chait and Hollot in [21]
on the Horowitz design procedure, and to avoid technical details related to the QFT
theory, we will illustrate the procedure by means of one of the application examples.
The design of the FORE proceeds in two steps:
• First, design a linear controller C(s) so that the linear closed-loop response satis-
fies disturbance rejection and noise suppression specifications (at the expense of
violating the required stability margins).
• Second, introduce the factors s+b1
(s +b)C(s), leaving unaltered the linear system.
Apply the linear control law given by (s + b)C(s) at the plant input. Convert the
preceding block s+b 1
into a FORE (introducing resetting). Choose the pole s = −b
in the FORE to reduce overshoot and improve stability margins.
The second step is crucial and it is important to understand that performance
improvement does not come with blind reset: a badly tuned reset controller results
in worse performance than that of the underlying LTI system.
To get an idea of what this procedure is able to do, let us briefly describe an
application example on control of a flexible mechanical shaft. The plant consists of
24 1 Introduction

three inertias J3 , J2 , J1 connected with flexible shafts. The servo motor drives the
inertia J3 and a tachometer measures the speed of the inertia J1 , resulting in the
fifth-order linear plant:
46083950
P (s) = .
(s + 1.524)(s 2 + 3.1s + 2820)(s 2 + 3.62s + 9846)
The following specifications were proposed to highlight the trade-offs in linear
design and the improvement achieved with reset control:
1. Bandwidth: The gain crossover frequency, defined by |C(j ωc )P (j ωc )| = 1, must
satisfy ωc > 3π .
2. Disturbance rejection: For low frequencies, ω ≤ π ,
 
 y(j ω) 
 
 d(j ω)  ≤ 0.2,
where y(j ω), d(j ω) denote the Fourier transforms of the plant output and dis-
turbance, respectively, of the feedback loop (Fig. 1.15).
3. Noise suppression: For high frequencies, ω ≥ 10π ,
 
 y(j ω) 
 
 n(j ω)  ≤ 0.3.
4. Asymptotic performance: Zero steady-state tracking error to constant references
and disturbances.
5. Overshoot: Overshoot in output to a reference step should be less than 20% (no
prefilter is permitted).
Using classical loop shaping techniques, it is impossible to meet all of the above
specifications with linear and time-invariant controllers. To illustrate this difficulty,
two seventh order controllers are proposed, C1 (s) and C2 (s). C1 (s) only fails to
satisfy the noise suppression specification and, when trying to arrange this with
C2 (s), this arrangement comes at the expense of the phase margin, and hence the
overshoot specification is violated.
The ultimate reason for this infeasibility is related to the Bode’s gain–phase rela-
tionship, that imposes a restriction on the linear designs. Turning to the reset control
design and using the previously described design procedure, a valid choice for the
first step is the controller C2 (s) since it satisfies all the design specifications except
the overshoot constraint.
Consequently, in the second step, the direct path in the feedback loop is reorga-
nized as
1
[(s + b)C2 (s)]P (s)
s +b
so that the base-linear loop transfer function remains unaltered: L(s) = C2 (s)P (s).
The actual linear control law implemented at the plant input is given by (s +b)C2 (s),
1
and the first order system s+b is endowed with reset and thus transformed into a
FORE. In [21], an analysis of the influence of the FORE base linear pole at s = −b
is performed, and the optimum value b = 14 is obtained. The final solution is proven
to be satisfactory and all the specifications are met.
1.4 First General Approaches to Analysis of Reset Control Systems 25

To conclude this section, let us quote some statements in [21] on the pioneering
work of Horowitz on reset controllers in the 1970s: ‘History shows that Prof. Isaac
Horowitz was often ahead of the curve in his feedback control research . . . Horowitz
motivated their use [reset controllers] by showing that with qualitative design, they
can exhibit better performance trade-offs than those in linear, time-invariant sys-
tems’.

1.4 First General Approaches to Analysis of Reset Control


Systems
Although the works by Horowitz on reset control were not based on describing
function or any other kind of approximate analysis, they considered only feedback
systems with either Clegg or FORE elements, subject to step inputs, and computed
overshoot assuming second-order dominance. Furthermore, the main obstacle for
widespread dissemination of reset control was, at that time, the lack of a general the-
ory, not restricted to Clegg and FORE, over basic questions such as well-posedness
and stability of reset control systems.
This theoretical framework for general reset control was subsequently developed
in a number of works by Y. Chait, C.V. Hollot, and coworkers in the University of
Massachusetts at Amherst. The developed framework is contained in a large number
of journal papers and conference proceedings, and in several Ph.D. theses. These
researchers considered for the first time higher order controllers and partial state
resetting, and achieved a number of original results on stability and performance of
reset systems.
The approach by Chait, Hollot, and coworkers addresses reset control systems,
and by this we mean here feedback control systems where the zero crossing of the
error signal triggers the resetting of some of the controller states. This is different
(although highly related) to the approach in the field of impulsive dynamical systems
where reset instants may arise from other independent mechanisms, not necessarily
related to zero crossing of feedback errors. This approach to reset systems as special
cases of impulsive systems will be treated in the following section.
In this section, we summarize the main results in the framework elaborated by
Chait, Hollot, and coworkers. For brevity, we cannot report here a detailed account
of the results in all the publications; see [18], [21], and references therein for de-
tails. For clarity, we choose [18] as the main reference for the following exposition
that addresses sequentially the topics of setup, stability, steady-state, and transient
performance of reset control systems.

1.4.1 General Setup and Asymptotic Stability

The framework elaborated by Chait and Hollot [21] can be considered as an exten-
sion of the works by Horowitz where for the first time general reset controllers (with
26 1 Introduction

Fig. 1.16 Reset control


system studied by Beker,
Hollot, Chait, and Han in [18]

arbitrary order and partial reset) were addressed. To deal with these general cases,
the authors in [18] and [21] adopt a modeling setup based on state space instead of
transfer functions.
They showed that a reset control system like the one in Fig. 1.16 can be repre-
sented by a closed-loop model based on state space variables in the form:
ẋ(t) = Acl x(t) + Bcl w(t) if x(t) ∈
/ M (t),
 +
x t = AR x(t) if x(t) ∈ M (t), (1.30)
yp (t) = Ccl x(t).
This is an impulsive-differential equation (IDE) where the first line is the ordi-
nary differential equation describing the continuous mode, the second line defines
the reset actions, or instantaneous state jumps, and the third line simply gives the
plant output yp . The closed-loop matrices (Acl , Bcl , Ccl ) are easily obtained from
the state-space matrices of the plant and of the base linear controller. The exoge-
nous signal is w(t) = r(t) − n(t) − d(t). The complete state vector x(t) is formed
by stacking the states of the plant and of the controller.
The number of plant and controller states are np and nr , respectively, and
n = np + nr is the total state dimension. Within the controller, it may happen that
only part of the state is affected by reset. Let nρ ≤ nr be the number of controller
states to be reset. In this way, Chait, Hollot, and coworkers introduced the concept
of a partial reset: when nρ < nr , not all the controller states are reset. Thus, the
framework in [18] generalizes the analysis by Horowitz to controllers with arbitrary
dimension nr and to reset laws that can be either ‘full reset’ (nρ = nr ) or ‘partial re-
set’ (nρ < nr ). By an appropriate reordering of states, the reset states can be stacked
in the last positions, so that the reset matrix AR takes the form:
AR = diag(1, . . . , 1, 0, . . . , 0) (1.31)
with nρ 0s. Another interesting proposal in [18] is the definition of the so-called
reset surface M (t):
 
M (t) = ξ ∈ Rn : e(t) = 0, (I − AR )ξ = 0 (1.32)
where the feedback tracking error is
e(t) = r(t) − n(t) − d(t) − yp (t) = w(t) − Ccl x(t).
The first condition in the definition of M (t) is e(t) = 0, which shows the de-
pendence on exogenous signals via e(t) or w(t) and also shows the time-varying
nature of M (t) when the exogenous signals are not constant. The second condition,
(I − AR )ξ = 0 is imposed for well-posedness, to avoid the so-called re-resetting,
1.5 Reset Systems as Impulsive Systems 27

that is, the fact that immediately after a reset, the state vector could satisfy again
the reset condition, implying formally an infinite sequence of resettings to be per-
formed instantaneously, also referred to as beating. Thus, to avoid beating, the term
(I − AR )ξ = 0 is added in the definition of M (t). As a consequence of this defini-
tion,
 
x(t) ∈ M (t) ⇒ x t + ∈ / M (t).
In addition, the stability of reset control systems is approached, firstly from an
internal or asymptotic point of view, and secondly from an external or input–output
point of view. Regarding internal stability, the unforced or autonomous version of
the reset system (1.30) is considered:
ẋ(t) = Acl x(t) if x(t) ∈
/ M,
 + (1.33)
x t = AR x(t) if x(t) ∈ M ,
where
 
M = ξ ∈ Rn : Ccl ξ = 0, (I − AR )ξ = 0 .
As a tool for addressing internal stability, a general theorem for arbitrary Lya-
punov functions is given, followed by a particularization to quadratic stability.
This condition is called the Hβ condition and is easily testable by using linear
matrix inequalities, or directly by frequency domain plots. When reset affects one
single state (nρ = 1), the problem is reduced to the search for the single variable
β ∈ R.
A property that follows immediately from the Hβ condition is that the class of
quadratically stable reset control systems requires their base-linear systems to be
stable, that is, Acl stable. On the other hand, the authors pointed out that reset can
stabilize an unstable base-linear system, and showed it with an example. Some other
examples show that there exist reset control systems that are stable but not quadrat-
ically stable. In spite of these examples, it is noticed that the class of quadratically
stable systems is rich and contains many common and practical cases of reset con-
trol applications. Furthermore, it can be seen that the Hβ condition can be easily
incorporated into the Horowitz design scheme introduced in the previous subsec-
tion.
The Hβ condition has been also proven to be useful in establishing other in-
teresting properties of reset control systems, for example regarding some type of
input-state stability. There are an important number of results by Chait, Hollot, and
coworkers regarding other aspects of performance in the steady state and transient
responses, that cannot be included here, for brevity. The interested reader can find
this information in [18], [21], and in the references therein.

1.5 Reset Systems as Impulsive Systems


In the two previous Sects. 1.3 and 1.4, we have reviewed, for the approximate period
1974–2004, the field that we call here ‘reset control systems’, and by this we mean
28 1 Introduction

feedback control systems with resetting applied at the zero-crossings of the tracking
error signal. At the same time, the field of ‘reset control systems’, as previously de-
fined, could be classified as a subarea of the much wider field of ‘impulsive systems’.
This term of ‘impulsive systems’ is used in different areas of applied mathematics
and dynamical systems theory, and there is an important body of results and a large
literature on the subject. It is out of the scope of this section to present a detailed
survey of results on impulsive systems.
‘Reset control systems’ can be considered as a particular class of ‘impulsive sys-
tems’ characterized by the fact that the reset condition is triggered by the zero cross-
ing of the tracking error signal, in a typical feedback loop. As the error signal is
a function of the state for unforced systems (and the forcing external signals can
be described by ‘exosystems’) we can say that in ‘reset control systems’ the reset
events are time-independent and only state-dependent. On the other hand, the re-
search in impulsive theory considers impulses applied at instants depending only
on time or both time-dependent and state-dependent. More generally, as reported in
Bainov and Simeonov (1989) [4], impulsive systems may be classified as (i) systems
with impulses at fixed instants, (ii) systems with impulse effect at variable instants,
and (iii) autonomous systems with impulse effects. Hence, the class (i) has time-
dependent impulses, the class (ii) has both time- and state-dependent impulses, and
class (iii) which contains ‘reset control systems’ has only state-dependent impulses.
Traditionally, most of the research effort has been dedicated to cases (i) and (ii), and
thus case (iii), which includes reset control, is less developed.
Summarizing, ‘reset control systems’ form a particular class of ‘impulsive sys-
tems’ with a specific state-dependent mechanism that triggers reset. Furthermore,
the field of reset control considers implicitly as the main motivation the idea that
the reset action is an effective way for overcoming fundamental limitations in linear
feedback. This motivation regarding control limitations is not present in the gen-
eral literature on impulsive control. In spite of the previous differences between the
points of view of ‘reset control’ and ‘impulsive systems’, there are strong and deep
connections between the concrete problems addressed by both fields. Thus, the re-
search on reset control can profit from employing results on impulsive systems as
addressed in the corresponding areas of applied mathematics and mathematical con-
trol theory.
Having this is mind, we include here three subsections based on three relevant
monographs in the field: Theory of impulsive differential equations by Lakshmikan-
tham, Bainov, and Simeonov [33], Impulsive control theory by Yang [43], and Im-
pulsive and hybrid dynamical systems by Haddad, Chellaboina, and Nersesov [30].
The following comments and discussions on these three books are focused only on
aspects regarding specifically the class of ‘reset control systems’, that is, we discuss
how reset control can be treated within the framework presented in the books, and
how general results could be particularized, if possible, to reset control.
1.5 Reset Systems as Impulsive Systems 29

Fig. 1.17 Impulses at fixed


times

Fig. 1.18 Impulses at


variable times

1.5.1 Theory of Impulsive Differential Equations

In the book Theory of Impulsive Differential Equations by Lakshmikantham, Bainov


and Simeonov [33], the impulsive differential systems are classified in the three
following classes:
(i) Systems with impulses at fixed times. This type of system is described by

ẋ = f (t, x) if t = tk ,
(1.34)
Δx = Ik (x) if t = tk ,
where tk , for k = 1, 2, 3, . . ., is a sequence of prescribed reset times and
Δx(tk ) = x(tk+ ) − x(tk ) are the state ‘jumps’ correspondingly applied. Fig-
ure 1.17 illustrates a typical time response of a system in this class.
(ii) Systems with impulses at variable times. These systems are given formally by

ẋ = f (t, x) if t = τk (x),
(1.35)
Δx = Ik (x) if t = τk (x),
where τk (x) is a sequence of state-dependent reset times having the property
that
τk (x) < τk+1 (x), lim τk (x) = ∞
k→∞

for all x. Figure 1.18 shows a possible time response of one such a system. No-
tice that solutions may hit the same surface τk (x) several times and also present
complex behaviors called ‘pulse phenomenon’ and ‘confluence’. In general,
a simplifying assumption is to consider functions τk to be continuous over Rn .
30 1 Introduction

Fig. 1.19 Impulses at a


general reset surface
h(t, x) = 0

Fig. 1.20 Autonomous


systems with impulses at a
state-dependent surface

The class (ii) can be extended to a more general class in the form:

ẋ = f (t, x) if h(t, x) = 0,
(1.36)
Δx = Ik (x) if h(t, x) = 0,
where h(t, x) = 0 might represent a more general, arbitrary time-dependent reset
surface. This situation is illustrated by Fig. 1.19.
Finally, we have
(iii) Autonomous systems with impulses. These systems are described by

ẋ = f (t, x), if x ∈
/ M,
(1.37)
Δx = Ik (x), if x ∈ M,
where M is an arbitrary subset in the domain of definition of x ∈ Rn . Fig-
ure 1.20 plots a possible trajectory, in the one-dimensional case, n = 1 and
x ∈ R. Although the analysis of an autonomous system with impulse effects
may appear to be easier than the other cases, in general it has been proven
harder since in general the continuity assumption over functions τk cannot be
achieved.
In summary, the classification proposed by [33] is based on the resetting mecha-
nism, and considers three possibilities: only time-dependent reset (class (i), t = tk ),
only state-depending reset (class (iii), x ∈ M) and both, time- and state-reset (class
(ii), t = τk (x)).
The reset control systems that we are interested in, as depicted in Fig. 1.16 or
defined by (1.30), fit well into the third class (iii). The unforced system (1.30) with-
out exogenous input (w = 0) is directly in the form (1.37). The same is true in the
case of constant (step) exogenous inputs. Furthermore, for the typical control inputs
(step, ramp, sinusoid, etc.) it is always possible to express them as the output of
an appropriately defined system (the so-called ‘exosystem’). As a consequence, the
1.5 Reset Systems as Impulsive Systems 31

zero crossing of the error that defines the surface M (t) in (1.32) can be expressed
as a condition on the system states and on the states of the exosystem. In this way,
the reset condition can be made time-independent so that reset systems belong to the
class (iii) of autonomous systems with impulses.
Most of the results in [33] apply to the cases (i) and (ii). Regarding stability,
which is a basic property required for a control system, the results in [33] for class (i)
systems cannot be applied to reset control since the reset times in reset control are
not prescribed, but are state-dependent. In addition, most of the literature about
state-dependent impulsive systems has been developed for systems which can be
described by functions τk with some specific properties, for example, it is usual to
assume that they are continuous over their domain. As it will be seen in detail in
Chap. 2, in general this is not the case for reset control systems.
To conclude this section, the book [33] presents a systematic treatment of impul-
sive systems, addressing well-posedness and continuity of solutions, stability cri-
teria, comparison results, vector Lyapunov functions, and many other aspects. The
interested reader can find a detailed treatment of these topics in [33], [4], and in the
references therein.

1.5.2 Impulsive Control Theory

Another very interesting monograph on impulsive systems is Impulsive Control The-


ory (2001) by T. Yang [43]. This book presents an approach in many parts similar
to the one in the previously discussed reference [33], that is, focused on impulsive
systems of classes (i) and (ii). But [43] has a point of view much closer to the field
of control, than that of [33], with a more mathematical emphasis.
In fact, as T. Yang notices in the preface to [43], there are several reasons why the
field of control has not fully had the benefit from the results in the field on impulsive
systems. One reason is that for many years impulsive control has been restricted to
very specific problems, like mechanical systems with impacts or spacecraft control.
Another reason is that the early research activities on impulsive control were re-
ported as Russian literature, not well-known to the English speaking community.
Even after publication of many English monographs on impulsive control, they did
not attract at the beginning the control community because the examples were too
limited and they targeted mainly mathematicians as potential readers. But nowadays,
the ideas on impulsive control are much more disseminated into the control, com-
munity and the number of control applications has increased, including problems
such as chaotic communication systems or nanoelectronic devices.
Impulsive systems are classified into the same classes that [33] uses: (i) impulses
at fixed times, (ii) impulses at variable times, and (iii) autonomous systems with
impulses (which are called discontinuous dynamical systems in [43]). But, at the
same time, an interesting parallel classification is presented in [43] based on the
way the control action enters the system. In the case of impulses at variable times
32 1 Introduction

t = τk (x), the classification is as follows. A type-I impulsive control system is given


by:

⎨ ẋ = f (t, x) if t = τk (x),
(I) Δx = U (k, y) if t = τk (x), (1.38)

y = g(t, x),
where the control input is implemented as sudden jumps in the state, given by the
impulsive control law U (k, y). A type-II impulsive control system is given by:

⎨ ẋ = f (t, x, ũ) if t = τk (x),
(II) Δx = U (k, y) if t = τk (x), (1.39)

y = g(t, x, ũ) if ũ = γ (t, x),
so that there are actually two control laws, the impulsive control law U (k, y) and
the continuous control law ũ = γ (t, y). A type-III impulsive control system is given
by:

⎨ ẋ = f (t, x, ũ) if t = τk (x),
(III) Δx = jk (x) if t = τk (x), (1.40)

y = g(t, x, ũ) if ũ = γ (t, x),
where for this type of system the control ũ is continuous and what is impulsive is
the plant itself. Most of the results are presented for systems with impulses at fixed
t = tk or variable t = τk (x) times with some regularity condition over functions τk ,
but not for autonomous systems with impulses, hence they are not applicable to the
reset control systems as in Fig. 1.16 or defined by (1.30).
On the other hand, it is out the scope of this section to report a detailed account
of the results in [43] but, just to give an idea of one among the many approaches
presented, let us see a simple example. The example is taken from Chap. 6 and
illustrates the concept of practical stability and the comparison system.
Given a system of type-II (1.39) and given the pair (μ, ν) with 0 < μ < ν, the
system is said to be practically stable when
x0  < μ ⇒ x(t, t0 , x0 ) < ν,
where x(t, t0 , x0 ) is the system solution for t ≥ t0 starting at x(t0 ) = x0 . Hence, the
concept of practical stability is weaker than that of asymptotic stability and only im-
poses boundedness on the state evolution. Consider the following two-dimensional
chaotic system [43]:

ẋ = y − f (x),
(1.41)
ẏ = −εx − ζy + η sin ωt,
where
1
f (x) = bx + (a − b)(|x + 1| − |x − 1|),
2
with a = −1.02, b = −0.55, ε = 1, ζ = 1.015, η = 0.15, and ω = 0.75. This is a
chaotic system, with the chaotic attractor shown in Fig. 1.21.
1.5 Reset Systems as Impulsive Systems 33

Fig. 1.21 Chaotic attractor of an uncontroller chaotic system in Yang [43]

The system is controlled in an impulsive fashion, as a type-II system in the form:



ẋ = y − f (x)
, if t = τi ,
ẏ = −εx − ζy + ηsin ωt,
(1.42)
x(τi+ ) = di x(τi )
, if t = τi .
y(τi+ ) = di y(τi ),
Using the Lyapunov function V = |x| + |y|, the stability of the two-dimensional
system can be derived from the stability of a certain one-dimensional comparison
system. The practical stability of the controlled system (1.42) with respect to (μ, ν)
is implied by the practical stability of the following comparison system:

ẇ = φw + θ if t = τi ,
+ (1.43)
w(τi ) = di w(τi ) if t = τi ,
where φ = max{|ε + |a||, |1 − ζ |} and θ = |η|.
Sufficient stability conditions are derived for the impulsive practical stabilization
of the chaotic attractor. A simplification is obtained letting the resetting coefficients
constant di = d and the reset instants to be fixed and equally spaced, i.e., τi+1 − τi =
δ for all i. It can be seen that a valid stabilizing parameter choice (d, δ) is given, for
any μ < ∞ and for some ν, by the conditions
34 1 Introduction

Fig. 1.22 Partial stability of the impulsively controlled chaotic system in [43]

d < e−φδ ,
θ |1 − eφδ |
< v.
φ 1 − deφδ
A valid solution is ν = 1 and (d, δ) = (0.06, 1.01). So we can conclude that no
matter how large the initial condition is, (x0 , y0 ) < μ < ∞, the resetting controls
in (1.42) achieve the practical stability (boundedness) (x(t), y(t)) < ν = 1 around
the origin. The simulation results in Fig. 1.22 show how the uncontrolled attractor
is destroyed and the controlled solution remains within the specified bounds.

1.5.3 Impulsive Dynamical Systems

An important number of results on stability and the control of impulsive systems


are contained in the book Impulsive and Hybrid Dynamical Systems by Haddad,
Chellaboina, and Nersesov [30], and in many other references by the authors and
coworkers (see [29], [30] and the references therein). One of the main distinctive
features in [30], compared to the monographs discussed in the previous subsections,
is the strong emphasis put on energy-based modeling and the control of impulsive
systems. The book contains an in-depth presentation of a dissipativity theory for
1.5 Reset Systems as Impulsive Systems 35

analysis and design of impulsive systems. The book also addresses impulsive sys-
tems from the point of view of nonnegative and compartmental systems, optimal
control, disturbance rejection, robustness against uncertainties, periodic solutions,
and so on.
In spite of this, again the developed results are not directly applicable to general
reset control systems (the approach is based on a number of regularity conditions
that give some restrictions over the functions τk ). Some brief comments about dis-
sipativity theory (modeling, definitions, and basic principles) will be given in the
following. It is out of scope of this subsection to give a detailed account of the
many aspects treated in [30]. However, to give an idea of the powerful technique of
energy-based control, we will include a simple but illustrative example.
An impulsive system is described, in its most general form, by the equations
ẋ(t) = fc (x(t)) if (t, x(t)) ∈
/ S,
(1.44)
Δx(t) = fd (x(t)) if (t, x(t)) ∈ S ,
where the resetting set S can be both time- and state-dependent. In order to have
well-posedness of the resetting times, two assumptions are imposed. The first as-
sumption A1 imposes that if a trajectory reaches the closure of S , at a point not
in S , then the trajectory must be directed away from S . The second assumption
A2 imposes that when a trajectory intersects S , it jumps out of S .
The general system (1.44) contains as a particular case the state-dependent reset
systems, obtained replacing the reset condition (t, x(t)) ∈ S by certain x(t) ∈ Z ,
where Z is the subset of the state-space where reset takes place. In any case, these
are autonomous or unforced systems, and in order to address dissipativity, it is nec-
essary to include input and output signals, giving rise to models in the form:
ẋ(t) = fc (x(t)) + Gc (x(t))uc (t) if (x(t), uc (t)) ∈
/Z,
Δx(t) = fd (x(t)) + Gd (x(t))ud (t) if (x(t), uc (t)) ∈ Z ,
(1.45)
yc (t) = hc (x(t)) + Jc (x(t))uc (t) if (x(t), uc (t)) ∈
/Z,
yd (t) = hd (x(t)) + jd (x(t))ud (t) if (x(t), uc (t)) ∈ Z .
The impulsive system (1.45) is said to be dissipative with respect to continuous
and discrete supply rates sc and sd if
T 
0≤ sc (uc (t), yc (t)) dt + sd (uc (tk ), yc (tk )), (1.46)
t0 k
for all T ≥ t0 and for all input and output, continuous and discrete, signals
uc , ud , yc , yd , in their corresponding signal spaces. Basically, it is an extension of
the usual dissipation inequality in which the power into the system contains the
usual continuous term, typically sc (uc (t), yc (t)) = yc (t) uc (t), plus discrete energy
packets sd (ud (tk ), yd (tk )) applied at the reset instants tk .
A complete framework for dissipativity of impulsive systems is given, in-
cluding generalizations of the available and required storage functions, Kalman–
Yakubovich–Popov conditions, stability of feedback interconnections, port-
36 1 Introduction

Fig. 1.23 Mechanical


equivalence of an example
in [30]

controlled Hamiltonian systems, and so on. For brevity, it is not possible to include
more details but, before concluding this subsection, we will include an insightful ex-
ample, just to give an idea of how powerful and intuitive this approach is. Consider
a second order Lienard plant:
q̈ + f (q(t)) = u(t), (1.47)
with q ∈ Rn and f : Rn → Rnbeing infinitely differentiable, f (q) = 0 if and only if
∂fi ∂fj
q = 0, and ∂qj = ∂qi for all i, j . The equations represent a mechanical system with
n particles of unit mass coupled by elastic forces given by f (·). Defining p = q̇, the
Hamiltonian is obtained as the sum of the kinetic and potential energies
q
1
H (p, q) = p  p + f  (σ ) dσ, (1.48)
2 0
where the integral is independent of the path, from the assumptions on f . The pro-
posed feedback controller is a copy or replica of the plant, but endowed with a reset
mechanism. More precisely,
q̈c + g(qc (t) − q(t)) = 0 if (q(t), p(t), qc (t), pc (t)) ∈
/Z,
Δqc (t) = q(t) − qc (t) if (q(t), p(t), qc (t), pc (t)) ∈ Z ,
(1.49)
Δpc (t) = −pc (t) if (q(t), p(t), qc (t), pc (t)) ∈ Z ,
u(t) = g(qc (t) − q(t)).
To gain insight into the plant–controller feedback system resulting, notice that in
the scalar case, the continuous dynamics, without the impulsive equations, represent
two unit masses with positions q, qc and velocities p, pc connected as in Fig. 1.23,
by means of two possibly nonlinear springs with restoring forces f (q), g(qc − q).
Notice that the nonimpulsive feedback system (without resetting) is a Hamilto-
nian lossless system and the total energy (plant energy + controller energy) remains
constant. Hence along the solutions of the system, the four states q, p = q̇, qc , and
pc = q̇c change in such a way that the total energy is constant, so that sometimes
the energy flows from the plant to the controller, and sometimes from the controller
back to the plant.
In order to achieve stabilization of the physical plant to the equilibrium (q, p) =
(0, 0), is necessary to implement in the virtual controller some kind of dissipation
of the virtual energy stored in it. It can be seen that resetting is an efficient way of
doing so. Two things have to be decided: when and how to reset. Since we want
to destroy the virtual energy in the controller, once the instant tk is decided, the
1.6 Reset Systems as Hybrid Systems 37

maximum dissipation is achieved by fully emptying the controller energy (kinetic


plus potential), in other words, the appropriate resettings are:
   
qc (t) → qc t + = q(t), pc (t) → pc t + = 0. (1.50)
These resettings are given in the form of incremental impulses in (1.49), it re-
mains to decide when to apply reset. Since the energy may flow in both directions
between the plant and the controller, it makes little sense to apply reset at instants
when the continuous dynamics are extracting energy from the plant. In fact, it is
proposed to start the controller with zero energy (qc (0) = q(0), pc (0) = 0, and thus
the energy at the beginning goes out of the plant) and reset the controller at each
q −q
moment when its energy Hc (qc , pc ) = pc pc /2 + 0 c g  (σ ) dσ achieves a max-
imum, that is, when
d  
0 = Hc (qc , pc ) = − g(qc − q) p. (1.51)
dt
This gives rise to the resetting set Z
   
Z = (q, p, qc , pc ) : g(qc − q) p = 0, (q − qc , pc ) = (0, 0) , (1.52)
which contains the maximum condition 0 = dH /dt plus the condition to avoid
re-resetting (beating). To see how these ideas work, let us particularize to a two-
mass system with nonlinear plant spring f (q) = q + q 3 and a linear virtual spring
g(qc − q) = 3(qc − q). The results of the simulations are shown in Figs. 1.24, 1.25,
1.26, 1.27.

1.6 Reset Systems as Hybrid Systems


In the tutorial by Goebel et al. [26], a modeling of hybrid systems is given that, in
the words of the authors, emphasizes the robustness of asymptotic stability to data
perturbation, disturbances, and measurement error. In general, a hybrid system has
the form

ẋ ∈ F (x) if x ∈ C,
+ (1.53)
x ∈ G(x) if x ∈ D,
where the dynamics is given by both a constrained differential inclusion and a con-
strained difference inclusion, with F and G being the set-valued maps, and C and D,
the flow and the jump set, respectively, subsets of Rn . Among other basic assump-
tions, it is assumed that C and D are closed sets in Rn , and in general D may overlap
with the interior of C, thus the jump set D enables rather than forces jumps.
The general model (1.53) is representative of a large class of hybrid systems as
it has been shown in [26]. In particular, the Clegg integrator and the FORE are first
modeled in this framework in [44], where the FORE compensator, with input e and
output xr , is given by

ẋr = λr xr + e if exr ≥ 0,
(1.54)
x+ = 0 if exr ≤ 0.
38 1 Introduction

Fig. 1.24 Example in Haddad, Chellaboina, and Nersesov [30]. Plant position and velocity

In addition, a plant with state xp ∈ Rn−1 is considered and described by



ẋp = Ap xp + Bpu u + Bdu d,
(1.55)
y = Cp x p ,
and finally, a closed-loop reset system is given by the hybrid system

⎨ ẋ = Ax + Bd d if x ∈ C,
x+ = Ar x if x ∈ D, (1.56)

y = Cx,
where d is an input disturbance to the plant, the flow set C is given by C = {x ∈
Rn |xT Mx ≥ 0}, and the jump set by D = {x ∈ Rn |xT Mx ≤ 0} with
 
0 −CpT
M= . (1.57)
−Cp 0
This representation of a reset system gives solutions compatible with previous
definition of reset systems, as long as the initial condition of the FORE (or Clegg
integrator) is zero. The advantage is that, without loss of generality, the dynamics of
the compensator is refined to be in the sector {(xr , e) ∈ R2 |xr e ≥ 0} instead of the
whole plane R2 . The implication of this is a less conservative approach to stability,
as developed in the mentioned works. In addition, temporal regularization has been
1.6 Reset Systems as Hybrid Systems 39

Fig. 1.25 Example in Haddad, Chellaboina, and Nersesov [30]. Controller position and velocity

used in order to the system solutions to be well defined for forward time. As a result,
the reset system is given by

⎨ τ̇ = 1, ẋ = Ax + Bd d if x ∈ C or τ ≤ ρ,
τ + = 0, x+ = Ar x if x ∈ D and τ ≥ ρ, (1.58)

y = Cx.
For this formulation of reset control systems, several results about bounded-input
bounded-state and bounded-input bounded-output stability are developed. In partic-
ular, it is shown that if the following pair of LMIs
⎧ T 
⎪ A P +P A+τF M P B C T
⎨ T
B P −γ I 0 < 0,
C 0 −γ I (1.59)

⎩ T
AR P AR − P − τR M ≤ 0

is feasible for some P = P T > 0, τF , τR ≥ 0, and γ > 0, then there exists some
ρ > 0 small enough such as the reset control system (1.58) is finite gain L2 -stable
from d to y. It has been also shown that even for simple examples the above state-
ment needs some refinement by using piecewise quadratic Lyapunov functions, and
that in general it may be necessary to solve a large number of LMIs to obtain a tight
bound of the gain.
40 1 Introduction

Fig. 1.26 Example in Haddad, Chellaboina, and Nersesov [30]. Control signal

For example, consider an integrator controlled by a FORE (Example 2 in [44]):



⎨ τ̇ = 1, x˙1 = x2 + d, x˙2 = −x1 + βx2 if x1 x2 ≤ 0 or τ ≤ ρ,
τ + = 0, x1 + = x1 , x2 + = 0 if x1 x2 ≥ 0 and τ ≥ ρ, (1.60)

y = x1 ,
 0 −1
where the flow and jump sets are obtained for M = −1 0
. Note that FORE has
1
a base linear system with transfer function s−β , and that the closed-loop base sys-
1
tem is given by the transfer function s 2 −βs+1 , thus the base closed-loop system is
unstable for β ≥ 0. In addition, for β ≥ 2 the dominant poles are real.
It has been shown that for any value of β ∈ R the L2 -gain from d to y is well-
defined and that it tends to zero as β tends to infinity (for positive values of β,
the base closed-loop system is unstable, and thus it does not have a well-defined
L2 -gain). Figures 1.28–1.29 show the time evolution of the closed-loop state of the
reset control system for a rectangular pulse as disturbance, for β = 2. In this case,
the obtained L2 -gain from d to y is 0.58. As a result, the reset control system can
beat the base linear control system, having less L2 -gain from d to y and thus having
a better disturbance rejection.
1.6 Reset Systems as Hybrid Systems 41

Fig. 1.27 Example in Haddad, Chellaboina, and Nersesov [30]. Plant, controller and total energy

It should be pointed out that for higher order systems, and especially partial reset
compensators, the reset condition based on the sign change in the product of the
input and output signals results in solutions of the reset control systems that are dif-
ferent from the reset condition based on the zero crossing of the input (which is the
classical condition used by Clegg and Horowitz), even though the initial condition
of the compensator state is zero. Consider, for example, (Fig. 1.30) a reset compen-
sator given by a parallel connection of an integrator and a Clegg integrator (this is a
compensator structure that has been used, for example, in [32]).
This is a simple partial reset compensator, and may be described by (with reset
condition e(t) = 0)
⎧  1−pr 

⎨ ẋ(t) = pr e(t) if e(t) = 0,
 
⎪ x(t + ) = 10 00 x(t) if e(t) = 0, (1.61)

u = ( 1 1 ) x,

where the base linear compensator is an integrator, and pr is a parameter with value
0 ≤ pr ≤ 1. Note that for pr = 0 the reset compensator corresponds to a linear
42 1 Introduction

Fig. 1.28 Example 2 in Nesic, Zaccarian, and Teel [44]. Time evolution of states x1 and x2 for a
rectangular pulse disturbance

integrator, and that for pr = 1 it is a CI. On the other hand, using the alternative
reset condition, the reset compensator is given by
⎧  1−pr 

⎨ τ̇ (t) = 1, ẋ(t) = pr e(t) if e(t)u(t) ≥ 0 or τ ≤ ρ,
 
⎪ τ (t + ) = 0, x(t + ) = 10 00 x(t) if e(t)u(t) ≤ 0 and τ ≥ ρ, (1.62)

u = ( 1 1 ) x.

Figure 1.31 shows a time response of both implementations for a sinusoidal input
and zero initial conditions. Both responses are identical up to the first reset, then
the compensator (1.62) performs several resets every ρ time units, and thus both
responses diverge up to next common reset. Note that due to the partial reset the
signs of the input and output are different for a finite time interval after the first
reset, and in fact there would be deadlock for ρ = 0 with the reset condition based
on the sign change.
1.7 Other Recent Results on Reset Control 43

Fig. 1.29 Example 2 in Nesic, Zaccarian, and Teel [44]. Time evolution of disturbance d and
closed-loop output y = x1

Fig. 1.30 I + CI
compensator

1.7 Other Recent Results on Reset Control

To conclude the historical survey of reset control, this section presents several recent
results to the analysis and design of reset control systems. Although the number of
recent publications on reset control is not very large, it is not possible, for brevity
reasons, to give a complete account of all of them. For the sake of concreteness,
a review of three works is given in the following: the first two works are somehow
extensions and are based on the reset system description as a hybrid system outlined
in Sect. 1.6, while the third work introduces reset control systems with a simple
resetting law (reset is performed at fixed instants). More specifically, the first work
[1] provides a general framework for reset systems design based on L2 and H2
gain concepts; the second work [34] also approaches the design from the L2 sense
and specializes to constant or varying, a priori known, references; finally, the design
technique in [28] adopts fixed reset instants tk and finds the controller parameters
from the optimization of a quadratic performance index.
44 1 Introduction

Fig. 1.31 Sinusoidal response of I + CI with different resetting law

1.7.1 Design Based on L 2 and H 2 Performance

This subsection presents a summary of some relevant ideas in [1]. This work con-
tributes new design techniques for reset controllers that generalize previous results
in [36, 44], using the same resetting law. The approach is also based on certain LMIs
that provide computable upper bounds of the L2 and H2 norms.
The plant–controller configuration is based on the typical H∞ framework shown
in Fig. 1.32. The block K denotes the SISO controller, with input y(t) ∈ R and
output u(t) ∈ R. The block P denotes the MIMO augmented plant that includes
the system to be controlled and possible input and output weighting filters. The
vector of exogenous signals w(t) ∈ Rnw typically includes references, disturbances,
and sensor noise, or their pre-weighted signals. Finally, z ∈ Rnz is the vector of
controlled outputs that collects signals (tracking error, actuation, plant output, etc.)
that are to be minimized in some sense (L2 or H2 ).
The augmented plant P is modeled as a MIMO linear system (with the common
assumption that there is no direct feedthrough from u to y) and the reset controller
K is described by a linear base system combined with the reset to zero of part of the
controller states. The resetting law is based on the approach reported in Sect. 1.6,
1.7 Other Recent Results on Reset Control 45

Fig. 1.32 Plant–Controller


configuration in Aangenent et
al. [1]

Fig. 1.33 Partitioning of the


(y, u)-space. The gray area is
the reset set. The white area
is the flow set, partitioned in
angular regions C1 , . . . , CN ,
for PWQ Lyapunov analysis

that is, the resets are applied when input and output have opposite signs. Figure 1.33
shows the reset set (in gray) or the set R  in the (y, u) plane where reset is applied,
D : y ≤ 0, u ≥ 0, or y ≥ 0, u ≤ 0,
and the flow set (white area) C  or the set for the continuous mode of operation,
C : y ≤ 0, u ≤ 0, or y ≥ 0, u ≥ 0.
Both sets intersect on the lines u = 0 and y = 0 where reset takes precedence over
flow. Analogously to the developments given in [36, 44] and outlined in Sect. 1.6,
the flow set is divided into a number of angular regions C  = C1 ∪ · · · ∪ CN in order
to apply a piecewise quadratic (PWQ) Lyapunov analysis that will be commented
on later. Combining the plant and controller equations as in Fig. 1.32, the model of
the closed-loop system Σ is obtained. This system Σ is described by a base linear
closed loop ẋ = Ax + Bw, z = Cx + Dw and a resetting law x + = AR x, where x
contains the plant and controller states and (A, B, C, D, AR ) are derived from the
plant and controller models.
The reset set D and flow set C for the closed-loop system Σ can be expressed in
terms of the state x and the input w

D: TR (x, w) ≥ 0, or TR (x, w) ≤ 0,


C: Tf (x, w) ≥ 0, or Tf (x, w) ≤ 0,
for some 2 × (nx + nw ) matrices TR , Tf .
Well-posedness is addressed considering inputs w(t) belonging to the set of real
Bohl functions, that is, signals that can be expressed in the form w(t) = H eF t v
for some (H, F, v) of adequate dimensions. Then a condition based on x(0) and
on a finite number of derivatives w(0), ẇ(0), ẅ(0), . . . is presented that guarantees
local existence of solutions. After these basic results, several stability results are
46 1 Introduction

developed using L2 and H2 analysis. For the L2 analysis, recall that the L2 gain
Σ∞ of the system Σ is given by
z2
Σ∞ = sup
0<w2 <∞ w2

for any input–output pair (w, z) of the system Σ , with initial state x(0) = 0. Thus, it
represents the input–output amplification in the sense of the L2 norms w2 , z2 .
Stabilization of the closed-loop reset system Σ in Fig. 1.32 requires a finite gain
Σ∞ < ∞, and optimal performance requires minimizing Σ∞ < ∞ with re-
spect to the free parameters in the controller K. Unfortunately, the computation of
the gain of a nonlinear system is extremely difficult, in general. However, a tech-
nique for estimating upper bounds γ of Σ∞ is proposed based on dissipativity
with respect to quadratic Lyapunov storage functions. To this end, quadratic Lya-
punov functions V (x) = x  P x with P > 0 are postulated, resulting in two LMIs
(continuous and discrete). These two LMIs are easily computable in the two param-
eters γ , P . From a bisection search on γ , checking feasibility w.r.t. P > 0, a proce-
dure for computing an upper bound of Σ∞ is obtained.
In order to improve accuracy in the bounding Σ∞ < γ and to avoid conserva-
tiveness, more general storage functions V (x) can be used. A detailed treatment for
piecewise quadratic (PWQ) Lyapunov functions is given. In this way, each angular
sector in the flow set in Fig. 1.33 (and the corresponding regions in the (w, x)-space)
are assigned to a Lyapunov function Vi (x) = x  Pi x with i = 1, . . . , N . The treat-
ment is more involved and the reduction of conservatism requires in some cases to
include strictly proper input weighting filters, but finally, a set of LMIs is obtained
that are applied to representative examples suggesting that are quite accurate for the
L2 gain characterization.
Regarding the H2 gain analysis, the first thing that had to be defined is the in-
terpretation of the H2 norm for nonlinear reset systems. One of the possible inter-
pretations of the H2 norm of linear SISO systems, given by stable strictly proper
transfer functions G(s) = C(sI − A)−1 B, is
 ∞ 1/2
G2 = |g(t)| dt
2
,
0

that is, the total energy of g(t), the SISO impulse response of G(s). This impulsive
response can also be written as g(t) = CeAt B, so it can be identified with g(t) =
CeAt x0 , that is, the autonomous response with initial condition x0 = B ∈ Rn .
In the MIMO case, a similar interpretation holds based on initial conditions x0
equal to the columns Bi of B, or equal to sums of columns Bi , Bj , . . . of B. Inspired
by this interpretation, the following definition is proposed: the H2 norm of the reset
system Σ in Fig. 1.32 corresponding to an initial value x0 ∈ Rn is the total output
energy defined as
 ∞ 1/2
Σ2,x0 = z (t)z(t) dt ,
0
1.7 Other Recent Results on Reset Control 47

Fig. 1.34 Reset control


system in [34]

where z is the output trajectory corresponding to zero input (w = 0) and initial


state x0 . In this way, the H2 analysis of reset systems can follow a treatment very
similar to the L2 analysis.
Typically, the H2 problems are posed with a diagram like the one in Fig. 1.32,
were P is the augmented plant and the input w is a vector containing the reference,
the disturbance, and the sensor noise. Inputs are assumed to be known a priori and
this knowledge is captured in the input filters. Possible input filters include:
• unit step W (s) = s+ε
1
,
• unit ramp W (s) = (s+ε)2 ,
1

• sine wave W (s) = (s+ε)ω2 +ω2 ,

with ε > 0 a small offset. A complete and detailed H2 analysis is developed based
on standard Lyapunov and PWQ Lyapunov functions and providing computable
LMIs so that, given the initial state x0 and the bound γ , the feasibility of the LMIs
implies the bound Σ2,x0 ≤ γ . Iterating with various initial conditions and/or con-
troller parameters, and by a bisection search on γ , these LMIs provide accurate,
powerful tools for performance optimization on reset control systems.

1.7.2 Design Based on L 2 Gain and Nonzero References

The approach in [34] is in many parts similar to that in [1], and also uses the frame-
work outlined in Sect. 1.6. Both works share the same resetting law (based on the
sign of input and output), the L2 point of view, the Lyapunov and PWQ Lyapunov
analysis, and the formalism leading finally to LMI analysis conditions.
What makes it different is some specific treatment of references and disturbances,
and a different L2 tracking criterion. This subsection presents briefly some details
of the interesting work in [34], in order to complement the section with additional,
useful tools, for L2 analysis of reset systems.
The reset control system setup is shown in Fig. 1.34. The strictly proper plant
is given by (Ap , Bp , Cp ), and the base linear model of the controller is given by
(Ar , Br , Cr , Dr ).
The flow or continuous mode is given by e · yr ≥ 0, and the jump, or reset, mode
by e · yr ≤ 0. With zero reference (or constant reference after coordinate change)
these sets are translated to the closed-loop state-space x = (x  
p , xr ) ∈ R , and thus
n
 
the flow set F is given by x Mx ≥ 0 and the jump set J is given by x Mx ≤ 0.
48 1 Introduction

For well-posedness, it is assumed that x ∈ J ⇒ x+ ∈ F , and it is shown that


Dr ≥ 0 is a sufficient condition for that. Also, it is assumed that the disturbances are
L2 -bounded:

1
d22 = d  (t)d(t) dt ≤ ,
0 δ
for some finite 1/δ. The target L2 tracking criterion is a bound in the form
e22 + ė22 ≤ η, (1.63)
which is shown to imply asymptotic tracking, that is, limt→∞ e(t) = 0. Since r(t)
is assumed to be known a priori, it is directly included in the model that in this
way only considers the error response e(t) with respect to bounded disturbances
d(t). Let us briefly see the case of constant r = r0 (the case r(t) = r0 e−εt is also
considered). The closed-loop flow and jump modes are respectively given by
ẋ = Ax + Br0 , x+ = AJ x,
and thus the equilibrium state is xe = −A−1 Br0 . Introducing the coordinate change
x̃ = x − xe and considering disturbances d(t), the closed-loop equations take the
form (flow and jump):
x̃˙ = Ax̃ + Bd d, x̃+ = AJ x̃, ẽ = −C x̃.
This is a hybrid system from d to e, and it is analyzed by means of the Lyapunov
function V (x̃) = x̃ P x̃ and the quantity
1  
L = V (x̃) + e e + ė ė − γ d  d.
γ
It is required that L < 0 along the continuous trajectories of the system (state in
the flow set) and non-increasing variation ΔV (x̃) = V (x̃+ ) − V (x̃) ≤ 0 along resets
(state in the jump set). By processing the inequalities for these two conditions, two
computable LMIs are obtained and thus their feasibility is a sufficient condition:
• If d = 0, the reset control system is globally asymptotically stable.
• If d = 0, then a relation like (1.63) holds in the form
e22 + ė22 ≤ γ x̃
0 P x̃0 + γ d2 ,
2 2
(1.64)
which in particular implies asymptotic tracking limt→∞ e(t) = 0.
To conclude this subsection, we refer the reader to [34] for the treatment with
PWQ Lyapunov functions (to reduce conservativeness), and for time-varying refer-
ences in the form r(t) = r0 e−εt , and for illustrative numerical examples and simu-
lations.

1.7.3 Design Based on Fixed Reset Instants tk

In the following, a brief summary of the results given in [28] and [45] will be pro-
vided. The main differences with the previous design approaches to reset control
1.7 Other Recent Results on Reset Control 49

are that the reset instants tk are a priori fixed and known, and that the after reset
values of the controller state xr (tk+ ) are not zero, but calculated online for optimal
performance.
For the sake of concreteness, let us follow the presentation in [45]. The system
considered is the typical reset control system like the one in Fig. 1.34. No distur-
bances are considered (d = 0), and in addition the reference input is assumed to be
constant r(t) = r. The closed-loop reset system is then described by:
ẋ = Ax + Br if t = tk ,
 +
x tk = M k x + N k r if t = tk ,
with pre-specified reset times t1 < t2 < · · ·. Note that in this case the reset control
system analysis is much simpler, indeed it is a particular type of an impulsive system
with impulses at fixed times (see Sect. 1.5.1), and thus a number of results including
well-posedness and stability can be directly applied. For example, internal stabil-
ity, with zero reference r = 0, can be derived by using an auxiliary discrete-time
(referred to as induced difference system):
+     
x tk+1 = Mk+1 eA(tk+1 −tk ) x tk+ =: Lk+1 x tk+ .
If Δtk := tk − tk−1 < ΔT for some fixed ΔT < ∞ and for all k, then the reset
control system is (asymptotically) stable if and only if the induced difference system
is (asymptotically) stable. As a corollary, in the simplest case in which both Δtk = δ
and Mk = M are constant, the system is asymptotically stable if and only if all the
eigenvalues of MeAδ =: L have modulus strictly smaller than one, that is, L is Schur
stable.
The main idea in [45] and [28] is to let free the controller after-reset state xr (tk+ )
(not necessarily equal to zero), and to compute xr (tk+ ) in order to minimize a
quadratic performance index in the form:
tk+1
Jk = e (tk+1 )P0 e(tk+1 ) + ė (tk+1 )Q0 ė(tk+1 ) + e (s)P1 e(s) ds,
tk

where the tracking error is e(t) = r(t) − y(t) = r(t) − Cx(t). The index Jk can
be seen as a moving index for the kth time interval [tk , tk+1 ] containing two ter-
minal contributions (weighted by P0 , Q0 ) and a distributed contribution (the one
weighted by P1 ). The choice of xr (tk+ ) at the beginning of the interval is the one
that minimizes Jk along the interval [tk , tk+1 ], in a moving horizon optimization
that resembles in part the idea in model-based predictive control (MPC).
For the solution, it is assumed that for any reference values r ∈ R there exists a
steady state value xss such that
Axss + Br = 0, Cxss − r = 0,
which, in the case of stable A, amounts to G(0) = 1, that is, unit dc gain of the
closed loop G(s) = C(sI − A)−1 B. The solution for xr (tk+ ) that minimizes Jk has
the form
   −1
xr tk+ = − Γk22 Γk21 (xp − xpss ) + xrss ,
50 1 Introduction

where the subscripts p, r, ss stand for the plant, reset controller, and steady-state,
respectively, and Γk21 , Γk21 are matrices that can be obtained from the problem data
(A, B, C, P0 , Q0 , P1 ).
To conclude this section, some simulation results that correspond to a real exper-
imental application (a PZT microactuator positioning stage) are shown. The dynam-
ics of the microactuator plant can be described by the state space model:

⎨ ẋ1 (t) = x2 (t),
ΣP : ẋ2 (t) = −a1 x1 (t) − a2 x2 (t) + bu(t),

y(t) = x1 (t),
where x1 , x2 are position and velocity of the moving stage and u is the control input,
and the parameters are a1 = 106 , a2 = 1810, and b = 3 × 106 . On the other hand,
the reset controller has a PI as base linear controller, and a periodic reset action:

⎨ ẋr (t) = e(t) if t = tk ,
+
ΣR : xr (tk ) = E1 x1 (t) + E2 x2 (t) + Gr if t = tk ,

u(t) = ki xr (t) + kp e(t).
The PI gains were tuned to kp = 0.08 and ki = 300. The optimization method in
[45] was applied to an index Jk with weights P0 = 2.1, Q0 = 10−6 , and P1 = 0. The
reset time interval was fixed constant to Δtk = tk − tk−1 = 1 ms. Then, the optimal
solution is given by the constant matrices E1 = −2.8 × 10−4 , E2 = −6.8 × 10−7 ,
and G = 0.0014.
Figures 1.35 and 1.36 show the simulation results for the position output and con-
trol signal, respectively, in three cases: the linear control, the standard reset control,
and the reset control with reset at fixed instants and variable reset.
It can be seen that the linear controller (the PI controller used as base linear
system in the reset controllers) is fast, but very underdamped. On the other hand,
the standard reset controller (that resets to zero at the zero crossings of the error) is
completely inadequate, producing sustained oscillations with underdamping around
the steady state values. Finally, the solution presented by [45] (with Δtk = 1 ms and
the weighting matrices given above), called an ‘improved reset controller’, gives
rise to clearly the best response. It is even faster than the linear solution, but with
zero overshoot. To be completely fair, it should be noted that the standard reset
controller is not the best possible design; note that the steady-state oscillation is due
to the fact that there is no integrator in the loop, thus it is not a good idea to reset
the integrator to zero, a potentially better design would be a PI + CI controller (see
Chap. 5). In addition, the proposed compensator uses state feedback, which means
that, in principle, it is more sensitive to plant uncertainty if an observer is used.

1.8 Preview of the Chapters


In Chap. 2, the definition of a reset control system, or reset system in general, is
given. In general, as it is common to impulsive systems, reset systems may exhibit
different types of solutions, in particular having complex patterns such as beating,
1.8 Preview of the Chapters 51

Fig. 1.35 Step responses of the example in Zheng et al. [45]

deadlock, and Zenoness. In control practice, this type of behavior is considered


pathological and thus several conditions will be given for reset control systems to
be well-posed. On the other hand, important properties of reset systems may be
derived by analyzing the reset instants that correspond to a given initial condition.
These patterns will be also analyzed, and their relationship with the observability
and reachability of the base linear system will be shown. This material is based on
the work [15].
Chapter 3 is devoted to the stability problem of reset control systems with finite-
dimensional base systems. The stability problem is addressed from different, com-
plementary points of view: (i) internal or Lyapunov stability, (ii) external or input–
output stability with passivity analysis, and (iii) stability by the describing function
method. Internal stability techniques are subdivided into techniques giving rise to
stability conditions that do not depend directly on the reset instants (reset-times in-
dependent) or, alternatively, are reset-times dependent. The first case is obtained
directly using continuous time Lyapunov functions (giving rise to the so-called Hβ
condition), while the second case (reset-times dependent) requires a discretization
at the after-reset values and a subsequent discrete-time Lyapunov analysis. Then,
the input–output L2 stability is studied, and a number or results are presented in
connection with passivity and dissipativity properties of reset feedback loops. Fi-
nally, the standard describing function tool is used for approximately predicting the
appearance or absence of oscillations, and as a good practical tool to evaluate the
52 1 Introduction

Fig. 1.36 Control inputs for the step responses of the example in Zheng et al. [45]

phase lead obtained by reset compensation. This material is based on several pub-
lished works [7, 9, 11, 13, 14, 19, 20].
Stability of time-delay systems under reset control is approached in Chap. 4.
Since reset control is able to overcome fundamental limitations, and time-delay is
one source of such limitations, it is of great interest to study the problem of delayed
reset systems. The stability is addressed by choosing an appropriate Lyapunov–
Krasovskii functional, and by imposing that the functional should decrease in the
continuous and reset modes. The resulting conditions take the form of linear matrix
inequalities, and, depending on the chosen functional, these LMIs can be delay-
dependent or delay-independent. In both cases, those LMIs, derived from time-
domain stability conditions, are translated into equivalent frequency-domain con-
ditions by means of adequate tools, like the Kalman–Yakubovich–Popov lemma, or
passivity techniques. From the latter frequency-domain conditions, useful interpre-
tations are exhibited regarding the achieved robustness in terms of scaled small-gain
or positive realness of certain subsystems. Finally, several examples illustrate the
application of the stability conditions, showing the potentials of reset control when
applied to time-delay systems. The chapter is based on [5, 10, 12, 16].
In Chap. 5, reset compensation has been used to overcome limitations of LTI
compensation. In this chapter, a new reset compensator, referred to as PI + CI, is
introduced. It basically consists of adding a Clegg integrator to a PI compensator
with the goal of improving the closed-loop response by using the nonlinear charac-
References 53

teristic of this element. It turns out that by resetting a percentage of the integral term
in a PI compensator, a significant improvement can be obtained over a well-tuned
PI compensator in some relevant practical cases, such as systems with dominant lag
and integrating systems. The main goal is the development of PI + CI tuning rules
for basic dynamic systems in a wide range of applications, including first and higher
order plus dead time systems. In addition, a number of design improvements such
us the use of a fixed or variable reset band, the integration with QFT, and the use of a
variable reset percentage are discussed. This material is based on [6, 8, 24, 38–41].
Finally, in Chap. 6, several practical applications of reset control systems will be
developed, all based on the PI + CI compensator: a temperature control system of
a heat exchanger, a bilateral teleoperation control system, and finally, a temperature
control of a solar collector field. The first two applications have been tested by
means of experiments in plants, while the third has been tested by using a (well-
proven) simulator of the field. The chapter is based on [23, 24, 38, 39, 41, 42].

References
1. Aangenent, W.H.T.M., Witvoet, G., Heemels, W.P.M.H., van de Molengraft, M.J.G., Stein-
buch, M.: Performance analysis of reset control systems. Int. J. Robust Nonlinear Control
20(11), 1213–1233 (2009)
2. Åström, K.J.: Limitations on control system performance. Eur. J. Control 6, 2–20 (2000)
3. Åström, K.J., Murray, R.M.: Feedback Systems. An Introduction for Scientists and Engineers.
Princeton University Press, Princeton (2008)
4. Bainov, D.D., Simeonov, P.S.: Systems with Impulse Effect: Stability, Theory and Applica-
tions. Ellis Horwood, Chichester (1989)
5. Baños, A., Barreiro, A.: Delay-independent stability of reset control systems. In: 32nd Annual
Conference of IEEE Industrial Electronics Society, Paris, France (2006)
6. Baños, A., Vidal, A.: Definition and tuning of PI + CI reset controller. In: European Control
Conference, Kos, Grecia (2007)
7. Baños, A., Carrasco, J., Barreiro, A.: Reset times-dependent stability of reset control system.
In: European Control Conference, Kos, Grecia (2007)
8. Baños, A., Vidal, A.: Design of PC + CI reset compensators for second order plants. In: IEEE
International Symposium on Industrial Electronics, Vigo, España (2007)
9. Baños, A., Carrasco, J., Barreiro, A.: Reset times-dependent stability of reset control with
unstable base systems. In: IEEE International Symposium on Industrial Electronics, Vigo,
España (2007)
10. Baños, A., Barreiro, A.: Delay dependent stability of reset control systems. In: American
Control Conference, New York, EE.UU (2007)
11. Baños, A., Dormido, S., Barreiro, A.: Stability analysis of reset control systems with reset
band. In: 3rd IFAC Conference on Analysis and Design of Hybrid Systems, Zaragoza, España
(2009)
12. Baños, A., Barreiro, A.: Delay-independent stability of reset systems. IEEE Trans. Autom.
Control 54(2), 341–346 (2009)
13. Baños, A., Barreiro, A.: Limit cycles analysis in reset systems with reset band. Nonlinear
Anal. Hybrid Syst. (2010). doi:10.1016/j.nahs.2010.07.004
14. Baños, A., Carrasco, J., Barreiro, A.: Reset times-dependent stability of reset control systems.
IEEE Trans. Autom. Control 56(1), pp. 217–223 (2011)
15. Baños, A., Mulero, J.I.: On the well-posedness of reset control systems. Technical Report
TR-DIS-1-2011, University of Murcia (2011)
54 1 Introduction

16. Barreiro, A., Baños, A.: Delay-dependent stability of reset systems. Automatica 46(1), 216–
221 (2010)
17. Beker, O., Hollot, C.V., Chait, Y., Han, H.: Plant with integrator: an example of reset control
overcoming limitations of linear systems. IEEE Trans. Autom. Control 46(11), 1797–1799
(2001)
18. Beker, O., Hollot, C.V., Chait, Y., Han, H.: Fundamental properties of reset control systems.
Automatica 40, 905–915 (2004)
19. Carrasco, J., Baños, A., van der Schaft, A.: A passivity approach to reset control of nonlinear
systems. In: 34th Annual Conference of the IEEE Industrial Electronics Society, Orlando,
EE.UU (2008)
20. Carrasco, J., Baños, A., Barreiro, A.: Stability of reset control systems with inputs. In: 16th
IEEE Mediterranean Conference on Control and Automation, Ajaccio, Francia (2008)
21. Chait, Y., Hollot, C.V.: On Horowitz’s contributions to reset control. Int. J. Robust Nonlinear
Control 12, 335–355 (2002)
22. Clegg, J.C.: A nonlinear integrator for servomechanism. Trans. AIEE, Part II 77, 41–42 (1958)
23. Fernández, A.F., Barreiro, A., Baños, A., Carrasco, J.: Reset control for passive teleoperation
applications in process control. In: 34th Annual Conference of the IEEE Industrial Electronics
Society, Orlando, EE.UU (2008)
24. Fernández, A.F., Barreiro, A., Baños, A., Carrasco, J.: Reset control for passive bilateral tele-
operation. IEEE Trans. Ind. Electron. (2010). doi:10.1109/TIE.2010.2077610
25. Feuer, A., Goodwin, G.C., Salgado, M.: Potential benefits of hybrid control for linear time
invariant plants. In: Proc. Amer. Control Conf., Alburquerque, pp. 2790–2794 (1997)
26. Goebel, R., Sanfelice, R., Teel, A.R.: Hybrid dynamical systems. IEEE Control Syst. Mag.
(2009)
27. Goodwin, G.C., Graebe, S.F., Salgado, M.E.: Control System Design. Prentice-Hall, Upper
Saddle River (2001)
28. Guo, Y., Wang, Y., Zheng, J., Xie, L.: Stability analysis, design and application of reset control
systems. In: Proceedings of the IEEE Int. Conf. on Control and Automation, Guangzhou,
China, May 30–June 1 (2007)
29. Haddad, W.M., Nersesov, S.G., Chellaboina, V.S.: Energy-based control for hybrid port-
controlled Hamiltonian systems. Automatica 39, 1425–1435 (2003)
30. Haddad, W.M., Chellaboina, V.S., Nersesov, S.G.: Impulsive and Hybrid Dynamical Systems.
Stability, Dissipativity and Control. Princeton University Press, Princeton (2006)
31. Horowitz, I.M., Rosenbaum, P.: Nonlinear design for cost of feedback reduction in systems
with large parameter uncertainty. Int. J. Control 24(6), 977–1001 (1975)
32. Krishnan, K.R., Horowitz, I.M.: Synthesis of a nonlinear feedback system with significant
plant-ignorance for prescribed system tolerances. Int. J. Control 19(4), 689–706 (1974)
33. Lakshmikantham, V., Bainov, D.D., Simeonov, P.S.: Theory of Impulsive Differential Equa-
tions. World Scientific, Singapore (1989)
34. Loquen, T., Tarbouriech, S., Prieur, Ch.: Stability of reset control systems with nonzero ref-
erence. In: Proceedings of the 47th IEEE Conf. on Decision and Control, Cancun, Mexico
(2008)
35. Middleton, R.H., Graebe, S.F.: Slow stable open loop poles: to cancel or not to cancel. Auto-
matica 35, 877–886 (1999)
36. Nešić, D., Zaccarian, L., Teel, A.R.: Stability properties of reset systems. Automatica 44(8),
2019–2026 (2008)
37. Seron, M., Braslavsky, J.H., Goodwin, G.C.: Fundamental Limitations in Filtering and Con-
trol. Springer, London (1997)
38. Vidal, A., Baños, A., Moreno, J.C., Berenguel, M.: PI + CI compensation with variable re-
set: application on solar collector fields. In: 34th Annual Conference of the IEEE Industrial
Electronics Society, Orlando, EE.UU (2008)
39. Vidal, A., Baños, A.: QFT-based design of PI + CI reset compensator: applications in pro-
cess control. In: 16th IEEE Mediterranean Conference on Control and Automation, Ajaccio,
Francia (2008)
References 55

40. Vidal, A., Baños, A.: Stability of reset control systems with variable reset: application to
PI + CI compensation. In: European Control Conference, Budapest, Hungary (2009)
41. Vidal, A., Baños, A.: Reset compensation applied on industrial heat exchangers. In: 14th IEEE
International Conference on Emerging Technologies and Factory Automation, Mallorca, Spain
(2009)
42. Vidal, A., Baños, A.: Reset compensation for temperature control: experimental applications
on heat exchangers. Chem. Eng. J. 159(1–3), 170–181 (2010)
43. Yang, T.: Impulsive Control Theory. Lecture Notes in Control and Information Sciences,
vol. 272. Springer, Berlin (2001)
44. Zaccarian, L., Nesic, D., Teel, A.R.: First order reset elements and the Clegg integrator revis-
ited. Proc. Am. Control Conf. 1, 563–568 (2005)
45. Zheng, J., Guo, Y., Fu, M., Wang, Y., Xie, L.: Improved reset control design for a PZT posi-
tioning stage. In: Proceedings of the 16th IEEE Int. Conf. on Control Applications, Singapore
(2007)
Chapter 2
Definition of Reset Control System and Basic
Results

2.1 Preliminaries and Problem Setup

The main focus of the book will be on the use of single-input single-output reset
compensators having as reset condition the classical condition of zero input. There-
fore, the main goal will be to analyze and design reset control systems with linear
and time invariant base systems. This chapter will be devoted to the definition of
a reset control system, developing basic conditions for a system to be well-posed.
In addition, an analysis of the dependence of zero crossing instants with respect
to initial conditions will be given, showing the complex patterns that may result
depending on the dimension of the after-reset surface.
Consider the feedback control system of Fig. 2.1, where the system P , with state
xp , is described by

ẋp (t) = Ap xp (t) + Bp u(t), xp (0) = xp0 ,
(P ) (2.1)
y(t) = Cp xp (t)

and the reset compensator R, with state xr , is modeled in principle by the impulsive
differential equation

ẋr (t) = Ar xr (t) + Br e(t) if e(t) = 0,
(R) (2.2)
xr (t + ) = Aρ xr (t) if e(t) = 0,

where xr (0) = xr0 and v(t) = Cr xr (t). Here np is the dimension of the state xp , and
nr is the dimension of the state xr . In addition, xr (t + ), or simply x+ r , is the value
xr (t + ε) with ε → 0+ . Aρ is a diagonal matrix with diagonal elements (Aρ )ii = 0 if
the compensator state component (xr )i is to be reset, and (Aρ )ii = 1 otherwise, i =
1, . . . , nr . In general, it is assumed that the first nρ̄ compensator state components
are not reset, and the last nρ compensator states are reset or set to zero at the reset
instants t in which the compensator input e(t) is zero. Thus, nr = nρ̄ + nρ and Aρ
is given by Aρ = diag(Inρ̄ , Onρ ). For the particular case corresponding to nρ̄ = 0

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 57


DOI 10.1007/978-1-4471-2250-0_2, © Springer-Verlag London Limited 2012
58 2 Definition of Reset Control System and Basic Results

Fig. 2.1 Reset controller R


applied to an LTI plant

and nr = nρ , that is, if Aρ is the zero matrix, the reset compensator will be referred
to as full reset. Otherwise, it will be referred to as partial reset.
The closed-loop autonomous unforced system is given by e(t) = −y(t), u(t) =
v(t), where by definition the closed-loop state of dimension n = np + nr is
 
xp
x= . (2.3)
xr

The result is the closed-loop system (without exogenous inputs)



ẋ(t) = Ax(t) if x(t) ∈
/ M,
(2.4)
x(t + ) = AR x(t) if x(t) ∈ M
x   A B C
with x(0) = x0 and y(t) = Cx(t), and where x0 = xp0 r0
, A = −BrpCp Ap r r , AR =
diag(Inp , Aρ ) = diag(Inp , (Inρ̄ , Onρ )), and C = (Cp 0).
Therefore, to complete the closed-loop system equations, the set M , which will
be referred to later as the reset surface, needs to be defined. Another set, the after-
reset surface MR , also plays an important role in the definition of closed-loop sys-
tem solutions. Note that reset actions occur when the state x(t) contacts the re-
set surface M at some instant t, that is, x(t) ∈ M , and then the state jumps to
AR x(t) ∈ MR . In general, the set MR will be defined as

MR = R(AR ) ∩ N (C), (2.5)

where R(X) and N (X) stand for the image and the null subspace of the linear
operator given by the matrix X, respectively. Thus, MR is the set of states x that
belong both to the null space of C (and then the output is y = Cx = 0) and to the
image space of AR (they are the after-reset states). In addition, the set M is defined
as
M = N (C) \ MR (2.6)
where the after-reset states are removed from the reset surface. Otherwise, an infinite
number of resets would be produced after a reset action; in general, the reset system
(2.4) can exhibit rather complex behaviors that have been referred to as beating or
pulse phenomena in the literature on impulsive systems [1, 10]. They are related to
the fact that reset instants may not be well defined or may not be distinct; in addi-
tion, even if the reset instants are well defined and are distinct, they may converge
to a finite number and then the reset system exhibits Zenoness. This topic will be
analyzed in detail in Sect. 2.2.2; however, a common solution in practice to avoid
these behaviors is to use time regularization.
2.1 Preliminaries and Problem Setup 59

A time regularization solution can be simply constructed adopting the scheme


proposed in [12]: the system (2.4) is modified including a temporal restriction over
the reset instants by simply avoiding resets if some minimum time between resets
Δm has not passed. Thus, the closed-loop system will be described by

Δ̇(t) = 1, ẋ(t) = Ax(t), (x(t) ∈
/ M ) ∨ (Δ ≤ Δm ),
(2.7)
Δ(t + ) = 0, x(t + ) = AR x(t), (x(t) ∈ M ) ∧ (Δ > Δm )

with Δ(0) = 0, x(0) = x0 , and y(t) = Cx(t), and where the reset action is only
performed if the state x(t) contacts the reset surface M at some reset instant tj , j =
1, 2, . . . , and in addition every reset interval Δk = tk − tk−1 satisfies Δk > Δm ,
k = 1, 2, . . . , where Δm > 0 is some given constant.

2.1.1 Reset Control System Solutions

The reset control system given by (2.4) or (2.7) is a special class of a system with
impulse effects, or an impulsive system. There is a large literature on impulsive sys-
tems [1, 5, 8, 10, 11, 14]. Most of the work has been done in the area of systems
with impulses at fixed instants, or systems with impulses dependent on the state.
The case in which reset and after-reset surfaces are time independent, which is usu-
ally referred to as autonomous impulsive systems, has attracted considerably less
attention in spite of being relevant for engineering applications including control
systems, this being the case of reset control systems. The framework developed in
[8] will be (partly) adopted here. The LTI system described by the first equation in
(2.4) will be referred to as the linear base system, or simply the base system, while
the second equation in (2.4) will be referred to as the resetting law.
Let D ⊂ Rn be an initial conditions set, and Ix0 ⊆ [0, ∞), with x0 ∈ Rn , a dense
subset of [0, ∞) such that Jx0 = [0, ∞) \ Ix0 is a countable set with a finite or infi-
nite number of elements; in fact, this set will be the set of reset times corresponding
to the initial condition x0 . In general, an initial condition x0 may produce a finite
or infinite number of resets, depending on whether the set Jx0 is finite or infinite.
A function x : [0, ∞) × D → Rn is a solution to the reset control system (2.4) if the
following conditions are satisfied:
1. x(·, x0 ) is left-continuous in t, that is, limτ →t − x(τ, x0 ) = x(t, x0 ) for all x0 ∈ D
and t ∈ (0, ∞).
2. x(·, x0 ) is differentiable in t, and dx(t,x
dt
0)
= Ax(t, x0 ), for all t ∈ Ix0 .
+
3. x(t , x0 ) = AR x(t, x0 ), for all t ∈ Jx0 .
In addition, functions τk : D → [0, ∞), k = 1, 2, . . . , are defined such that τk (x0 )
is the kth reset instant of the solution x(·, x0 ). Therefore, functions Δk : Rn →
[0, ∞), k = 1, 2, . . . , are defined such that Δk (x0 ) = τk (x0 ) − τk−1 (x0 ) is the kth
reset interval, with τ0 (x0 ) = 0 for any x0 ∈ Rn . The following result is directly
adapted from [1] with minimal effort.
60 2 Definition of Reset Control System and Basic Results

Proposition 2.1 For any initial condition x0 ∈ D , assume that τ1 (x0 ) < τ2 (x0 ) <
· · · < τk (x0 ) < · · · , and τk (x0 ) → ∞ as k → ∞, then there exists a unique solution
x(·, x0 ) (x(·) in short) to the reset control system (2.4) that can be written in the
form x(t, x0 ) = W (t, x0 )x0 for t > 0, where the transition matrix W (t, x0 ) is given
by
W (t, x0 ) = eA(t−τk (x0 )) AR eA(τk (x0 )−τk−1 (x0 )) AR · · · AR eAτ1 (x0 ) (2.8)
for t ∈ (τk (x0 ), τk+1 (x0 )].

Proof Since reset instants are directly determined by the initial condition, for a
given initial condition x0 ∈ D and its reset instants tk = τk (x0 ), k = 1, 2, . . . , the
system (2.4) is an impulsive system with impulses at fixed instants t1 , t2 , . . . , where
by assumption 0 = t0 < t1 < t2 < · · · . A direct application of Theorem 3.6 and
Corollary 3.2 in [1] gives the result. 

Note that in general if the formulation (2.7) for reset system is used then the
reset intervals are lower-bounded by the time regularization constant Δm , that is,
Δk (x0 ) > Δm , k = 1, 2, . . . for any x0 ∈ R, and thus the assumptions of Propo-
sition 2.1 are clearly satisfied. If time regularization is not performed, deadlock
and beating could be present if τk+1 (x0 ) = τk (x0 ) for some initial condition x0 and
some k; or even a Zeno solution may be obtained if τk (x0 ) < ∞ for k → ∞.
Alternatively, if the initial instant is t0 = 0 and the initial condition is x(t0 ) = x0 ,
then a solution x(t, t0 , x0 ) is defined in a similar way.
In Sect. 2.2, a detailed analysis of beating, deadlock, and Zeno solutions in the
reset system (2.4) will be given. It will be shown that under mild conditions it does
not have these behaviors and that time regularization is not necessary to have well-
defined solutions for forward time.

2.1.2 Characterization of Reset Intervals

An important question that has not been previously approached in the reset control
literature is the analysis of reset instants for a given reset control system. For ex-
ample, given the base system of (2.4), it is not evident if every initial condition will
cross the reset surface M at some finite instant. Usually, since at least a reset action
is wanted for initial conditions in some set D , a common practice is to design the
base closed-loop system to have a pair of dominant complex poles, and thus to force
crossings after some arbitrarily large time.
In the following, it will be shown that this practice is, in fact, theoretically sup-
ported, and in addition an upper bound over resets intervals will be computed. In
general, the problem of computing crossings of the solution of a linear system gov-
erned by ẋ(t) = Ax(t), with an initial condition x0 ∈ R, with a given hyperplane,
is a particular instance of the reachability problem for linear systems. In general,
for arbitrary values of the state matrix A, it has been shown to be an open problem,
referred to as the continuous Skolem–Pisot problem [4, 9]. As discussed above,
2.1 Preliminaries and Problem Setup 61

this is a central problem in reset control where a base system has to be designed for
the reset system to perform crossings with a reset surface.
For the analysis of the crossings with the hyperplane defined by the row vector C,
that is, the instants t > 0 at which CeAt x0 = 0, it is convenient to use the equation
Ce(A+λI )t x0 = 0 for some λ ∈ R, obtaining the same results [4]. This is a simple,
but key simplifying result because all the eigenvalues of A can be assumed to have
non-positive real parts without loss of generality, for some λ properly chosen. In
addition, it has also been shown [4] that if for some initial condition x0 the matrix A
does not have dominant real eigenvalues and (A, C, x0 ) is reduced, the continuous
Skolem–Pisot problem always has a solution for that initial condition.
By definition, an eigenvalue of A is dominant if it is the rightmost placed eigen-
value of A in the complex plane. In addition, the term CeAt x0 can always be split as
CeAt x0 = y1 (t) + y2 (t), where by definition the dominant term y1 (t) is not identi-
cally zero if (C, A, x0 ) is reduced, and in addition y2 (t) tends to zero exponentially
fast as t increases. This means that a reset control system with such a matrix A will
always produce crossings for the initial condition x0 . Note that the set of states x0
for which (A, C, x0 ) is not reduced has a linear subspace structure, and will be re-
ferred to as R̄. Note that if the modes corresponding to dominant eigenvalues are
unobservable then (A, C, x0 ) is not reduced for any x0 ∈ Rn , and thus R̄ = Rn ,
but in the case they are observable, (A, C, x0 ) may not be reduced for some initial
condition x0 ∈ Rn , and in general R̄ is not the empty set.
In general, if the matrix A ∈ Rn×n has eigenvalues λ1 , λ2 , . . . , λs such that ki =
index(λi ), then it can be expressed as
⎛ ⎞
Q1
⎟ 
s

A = (P1 . . . Ps )J ⎝ ... ⎠ = Pi J (λi )Qi , (2.9)
Qs i=0

where J is the Jordan form, and J (λi ) the Jordan segment associated to the eigen-
value λi . In addition, for a function f : R → R such that f (λi ), f  (λi ), . . . ,
f (ki −1) (λi ) exist for each i = 1, . . . , s, the value of f (A) can be determined by
using a generalization of the spectral theorem for non-diagonalizable matrices

 i −1
s k
f (j ) (λi )
f (A) = (A − λi I )j Gi , (2.10)
j!
i=1 j =0

where Gi = Pi Qi is the spectral projector corresponding to the eigenvalue λi .


In the case in which the matrix A is diagonalizable, the expression of f (A) is
much simpler since index(λi ) = 1, for i = 1, . . . , s. In this case,

s
f (A) = f (λi )Gi , (2.11)
i=1

where in addition the spectral projectors are simply given by Gi = vi wTi , with vi
and vj being the right-hand eigenvector and the left-hand eigenvector corresponding
to the eigenvalue λi , respectively.
62 2 Definition of Reset Control System and Basic Results

In the following, it will be shown how a reset control system produces an infinite
number of resets, and that the reset intervals are upper-bounded, if the base system
does not have real dominant eigenvalues.

Proposition 2.2 Consider the reset control system (2.4), where the matrix A does
not have real dominant eigenvalues, and a nonempty closed set of initial conditions
D ⊂ Rn \ R̄, then reset intervals are uniformly upper-bounded, that is, Δk (x0 ) =
τk+1 (x0 ) − τk (x0 ) < ΔM , k = 1, 2, . . . , for any x0 ∈ D and some finite constant
ΔM > 0.

Proof It is assumed, without loss of generality, that A has dominant imaginary


eigenvalues. The case in which A has a pair of dominant eigenvalues will be
considered in the following (with multiplicity not necessarily equal to one), the
more general case of multiple dominant eigenvalues is a bit more involved but
follows a similar reasoning. The eigenvalues of A are ordered in such a way that
its spectrum is σ (A) = {λ1 , λ2 , . . . , λs } = {iβ1 , −iβ1 , α3 + iβ3 , . . . , αs + iβs }, and
αl < 0, l = 3, . . . , s. In addition, let kl , l = 1, . . . , s, be the index of each eigenvalue.
Then, the output y(t) of the reset control system can be expressed (using (2.9) and
(2.10) for computing eAt ) as

 l −1 j λl t
s k
t e
y(t) = C(A − λl I )j Gl x0 , (2.12)
j!
l=1 j =0

where Gl is the spectral projector associated to the eigenvalue λl . It can be split in


two parts as y(t) = y1 (t) + y2 (t), where

 l −1 j iβl t
2 k
t e
y1 (t) = C(A − iβl I )j Gl x0 (2.13)
j!
l=1 j =0

and
 l −1 j αl t iβl t
s k
t e e
y2 (t) = C(A − (αl + iβl )I )j Gl x0 . (2.14)
j!
l=3 j =0

Here y2 (t) tends to zero exponentially fast as t increases since αl < 0, l = n +


1, . . . , s, and thus it is clear that for any ε2 > 0 it is always possible to choose an
instant t2 large enough such that |y2 (t)| < ε2 x0 , for t ≥ t2 .
 −1 (A−iβl I )j Gl
In addition, defining constant matrices Zlj = jkl=0 j! ∈ C and matrix
kl −1
polynomials Pl (t) = ( j =0 Zlj t ), the dominant term of the output y1 (t) can be
j

expressed as
 
y1 (t) = C P1 (t)eiβl t + P2 (t)e−iβl t x0
= C(P1,R (t) cos(β1 t) + P1,I (t) sin(β1 t))x0 , (2.15)
2.2 Zenoness, Beating, and Deadlock 63

where P2 (t) = P̄1 (t), P1,R = 2 Real{P1 }, and P1,I = 2 Imag{P1 }. Now, for instants
tm = m 2πβ1 , with m ∈ N, the value of the dominant term is given simply by y1 (tm ) =
CP1,R (tm )x0 , and since (C, A, x0 ) is reduced for any x0 ∈ D , and D is closed,
there must exist some lower bound ε1 > 0 such that |y1 (tm )| > ε1 x0 , for any m
and x0 ∈ D . This can be shown by contradiction, if it were false then it would exist
a sequence of states {x01 , x02 , . . . } → x0 such as for any t > 0, CP1,R (t)x0n  → 0
as n → ∞. Thus, since D is closed it must contain a state x0 such as (C, A, x0 ) is
not reduced, which is a contradiction. Also note that for an m1 large enough, the sign
of y1 (tm ) is constant for m > m1 (for example, choosing tm1 as the Cauchy’s bound
of the roots of the polynomial CP1,R (t)x0 ). It will be assumed that y1 (tm1 ) > 0,
otherwise tm = m πβ , m ∈ N, may be chosen. Therefore, it is possible to find a large
ε2 x0 
enough instant t2 such as ε2 < ε1 and thus | yy21 (t(tmm )) | < ε1 x0  = ε2
ε1 < 1, and then it is
y2 (tm )
true that y(tm ) = y1 (tm )(1 + provided that tm > max{tm1 , t2 }. A similar
y1 (tm ) ) > 0,
reasoning may be applied to show that y(tn ) < 0, with tn = n βπ1 , for some tn >
max{tn1 , t2 }, where n1 is computed in a similar way. As a result, the output of the
reset system y(t) = CeAt x0 , for some initial condition x0 ∈ Rn , is equal to 0 for t <
tm , with the upper bound tm being independent of the initial condition x0 . Then,
since the reset action is only active for t > Δm , Δ1 (x0 ) = τ1 (x0 ) < Δm + tm =: ΔM .
Obviously, the rest of the reset intervals are also upper-bounded by ΔM , and thus
the proof is complete. 

2.2 Zenoness, Beating, and Deadlock


In control practice, it is required that the reset control system solutions x(t, x0 ) be
well defined in the sense that they exist and are unique for any t > 0, as given in
Proposition 2.1. As it has been previously discussed, this may be complicated by the
fact that these solutions may exhibit complex phenomena such as non-continuability
of solutions or deadlock, beating or livelock, and Zenoness. These concepts will be
used as defined in [8]. Deadlock occurs when the state x(t) cannot evolve in time
because no continuation, continuous or discrete, is possible. To avoid deadlock, a
typical assumption (adapted from [8]) is that
(A1) ∀x(t) ∈ MR , ∃ε > 0 such that ∀δ ∈ (0, ε) x(t + δ) ∈
/ M,
that is, the after-reset states evolve with the continuous base dynamics for some
finite time interval. Beating appears when the system solution encounters the reset
surface after resetting, which in our case is simply avoided by assuming that
(A2) MR ∩ M = ∅,
that is, assuming that the after-reset states are not elements of the reset surface.
Finally, a Zeno solution exists if a system solution has infinitely many reset actions
in a finite time.
By definition, the reset control system (2.4), with sets MR and M , is well-
posed if it satisfies conditions (A1) and (A2) (and thus it does not exhibit beating or
64 2 Definition of Reset Control System and Basic Results

deadlock, but Zeno solutions may exist in principle). Note that in the reset control
system (2.4), the reset action occurs at the instants t at which the output y(t) is zero.
The set MR is defined as in (2.5), that is, as the set of states that belong both to the
null space of C (and then the output is y = Cx = 0) and to the image space of AR
(they are the after reset states). In addition, the set M will be defined as in (2.6),
where the points that are the after-reset states are removed from the reset surface to
satisfy condition (A2), otherwise an infinite number of resets may be produced after
a reset action, and beating would be present.

2.2.1 Well-posedness: Beating and Deadlock

To have well-defined solutions to the reset systems as given in last section, reset
instants have to be well defined and distinct. In general, two phenomena that have
to be avoided are deadlock (non-continuability of solutions), and beating or live-
lock. Another important type of solutions like Zeno solutions will be treated in next
section. In general, a reset control system as defined by (2.4)–(2.6) does not exhibit
beating, once the surfaces MR and M are defined according to (2.5) and (2.6),
respectively. On the other hand, additional assumptions have to be made to avoid
deadlock.

Proposition 2.3 The reset control system (2.4)–(2.6) is well-posed if the after-reset
surface MR is a subset of the observable subspace of the linear base system, that is,

MR ∩ N (Obase ) = {0} (2.16)

where
⎛ ⎞
C
⎜ CA ⎟
⎜ ⎟
Obase = ⎜ .. ⎟.
⎝ . ⎠
CAn−1

Proof Beating is avoided by defining the after-reset surface MR by (2.6). Thus, the
proof is centered around deadlock. In general, given any initial condition x0 ∈ D , the
reset surface M is first contacted at the instant t1 = τ1 (x0 ). Then the reset instant
t1 is simply given by t1 = inf{t > 0|x(t, x0 ) ∈ M }, and thus is a solution of the
equation CeAt1 x0 = 0. In addition, it must be true that eAt1 x0 ∈/ R(AR ) according
to the definitions of the after-reset and reset surfaces as given by (2.5) and (2.6),
respectively. For simplicity, consider first that the closed-loop state matrix A has
distinct eigenvalues, then the matrix exponential may be computed by using the
Caley–Hamilton method, that is,

eAt1 = α0 I + α1 A + · · · + αn−1 An−1 , (2.17)


2.2 Zenoness, Beating, and Deadlock 65

where αi , i = 0, . . . , n − 1, are given by


⎧ λt

⎪ e 1 1 = α0 + α1 λ1 + · · · + αn−1 λn−1
1 ,



⎨ eλ2 t1 = α0 + α1 λ2 + · · · + αn−1 λn−1 ,
2
(2.18)
⎪ ···




⎩ λn t1
e = α0 + α1 λn + · · · + αn−1 λn−1
n .

Using the vector notation λT = (λ1 λ2 . . . λn ), α T = (α0 α1 . . . αn−1 ) and eλt1 =


 n λi ti e , where e stands for the unit vector (0 . . . 0 1 0 . . . 0)T in which the
i=1 e i i
ith component is 1, (2.18) can be compactly written as

eλt1 = V (λ)T α, (2.19)

where V (λ) is a (nonsingular) Vandermonde matrix. Now, by eliminating α from


(2.17) and (2.19) the equation 0 = CeAt1 x0 is transformed into
⎛ ⎞
C
⎜ CA ⎟
⎜ ⎟ T
0 = αT ⎜ .. ⎟ x0 = eλt1 U (λ)Obase x0 , (2.20)
⎝ . ⎠
CAn−1

where U (λ) = V (λ)−1 . Now, if condition (2.16) is satisfied, then the right-hand
side of (2.20) is an analytical function (in fact, a sum of exponentials) that is not
zero for all t ≥ 0. As a result, it has no isolated zeros, and then the reset instant
t1 is lower-bounded. In other words, solutions of (2.20) may not be obtained for
arbitrarily small values of the reset instant t1 , and thus deadlock does not occur if
condition (2.16) is satisfied.
In the case in which the eigenvalues of A may be repeated, a similar ar-
gument may be applied. Note that eAt1 may be written as the infinite series
 Ai t1i ∞ λi t1i
D(A) = ∞ i=0 i! . Thus the polynomial D(λ) = i=0 i! can be factorized as
D(λ) = Q(λ)P (λ) + R(λ), with R(λ) = 0, or deg(R) < deg(P  ) = n. jIn addi-
tion, R has degree no greater than n − 1, and thus R(λ) = n−1 j =0 αj λ . Since
the characteristic polynomial is zero for the eigenvalues of A, D(λk ) = R(λk ) for
 λik t1i n−1
k = 0, 1, . . . , n − 1. And then D(λk ) = ∞
j
i=0 i! = e
λk t1 = R(λ ) =
k j =0 αj λk
for k = 0, 1, . . . , n − 1. This can be compactly expressed as V T (λ)α = eλt1 , and the
expression (2.19) is obtained.
Now, if A has r different eigenvalues with respective % multiplicity order ni , and as
a consequence the characteristic polynomial is p(λ) = ri=1 (λ − λi )ni , then again
there exist unique polynomials Q and R such as D(λ) = Q(λ)P (λ) + R(λ) where
D(λ) = eλt1 and R = 0 or deg(R) < deg(P ). Here R can be expressed as R(λ) =
 n−1 i
i=0 αi λ , where the coefficients are unique. Since p and its derivatives up to order
66 2 Definition of Reset Control System and Basic Results

nr are zero at λi ,

d j D(λi ) d j R(λi )
= ∀i = 1, 2, . . . , r, ∀j = 0, 1, . . . , ni − 1. (2.21)
dλj dλj
This can be expressed by μ = W α, where

 i −1
r n
j
μ= ei eλi t1 ⊗ ej λi , (2.22)
i=1 j =0
⎛ ⎞
 r ni −1 j
∂ V (λ) ⎠
W =⎝ ei ⊗ ej eTi j
. (2.23)
i=1 j =0 ∂λi

By using arguments based on the Lagrange–Hermite interpolation problem, it


can be shown that, in fact, the matrix W is invertible. And then an expression sim-
ilar to (2.20) may be obtained. Using similar arguments as those after (2.20), the
proposition is proved for the general case of repeated eigenvalues. 

Obviously, a reset control system will be well-posed if the base linear system is
observable. But some unobservable base linear systems can also define well-posed
reset control systems. Therefore, note that in the proof of Proposition 2.3 no par-
ticular structure of the matrices A, C, and AR has been used. Thus, the result is in
general valid for any reset system given by (2.4) with arbitrary values of those matri-
ces. In the following, two examples corresponding to an ill-posed (not well-posed)
reset system and a well-posed reset control system are given.

2.2.1.1 Example (Ill-posed Reset System)

This example is used in [12] for analyzing some weak points in the definition of
reset systems given in [3]. Consider a reset system (2.4) with the following system
matrices
⎛ ⎞ ⎛ ⎞
−1 0 0 1 0 0  
A = ⎝ 0 −1 −1 ⎠ , AR = ⎝ 0 1 0 ⎠ , C = 1 0 0 , (2.24)
0 1 −1 0 0 0

where in addition the sets MR and M are defined according to (2.5) and (2.6) as
 
MR = R(AR ) ∩ N (C) = span (0, 1, 0)T ,

and
 &  
M = N (C) \ MR = span (0, 1, 0)T , (0, 0, 1)T span (0, 1, 0)T .

Note that this reset system cannot be realized as a control reset system with the
structure of Fig. 2.1.
2.2 Zenoness, Beating, and Deadlock 67

In [12], it is correctly argued that for any initial condition x0 = (0, a, 0)T ∈ MR ,
the solution is ill-defined since the continuous dynamics makes the system instantly
evolve to the set MR and then the system instantly resets once the reset surface
is reached, making infinitely many resets without leaving the reset surface; in fact,
there is deadlock. Note that this is due to the fact that the after-reset surface MR is
a subset of the unobservable subspace of the linear base system, which is given in
this case by
⎛⎛ ⎞⎞ ⎛⎛ ⎞⎞
C 1 0 0
N ⎝⎝ CA ⎠⎠ = N ⎝⎝ −1 0 0 ⎠⎠ = N (C) ⊃ MR . (2.25)
CA2 1 0 0

2.2.1.2 Example (Well-posed Reset Control System)

This example, adapted from [3], shows how an unobservable linear base system
may define a well-posed reset system, as long as the unobservable subspace does
not contain after-reset states. Consider a reset control system (2.4)–(2.6) with
⎛ ⎞ ⎛ ⎞
0 0 1 1 0 0  
A = ⎝ 1 −0.2 1 ⎠ , AR = ⎝ 0 1 0 ⎠ , C= 010 (2.26)
0 −1 −1 0 0 0

that has an unobservable mode corresponding to a stable pole–zero cancellation in


the linear base system, where the plant has a transfer function P (s) = s(s+0.2)
s+1
, and
the base compensator is C(s) = s+1 1
(corresponding to a first order reset element—
FORE). In addition, the after-reset and reset surfaces are given by MR = R(AR ) ∩
N (C) = span{(1, 0, 0)T } and M = N (C) \ MR = span{(1, 0, 0)T , (0, 0, 1)T } \
span{(1, 0, 0)T }, respectively. In this case, the set MR is not a subset of the linear
base system unobservable subspace given by
⎛⎛ ⎞⎞ ⎛⎛ ⎞⎞
C 0 1 0  
N ⎝⎝ CA ⎠⎠ = N ⎝⎝ 1 −0.2 1 ⎠⎠ = span (1, 0, −1)T .
CA2 −0.2 −0.96 −0.2
(2.27)
As a result, Proposition 2.3 may be used to ensure that the system is well-posed.
Figure 2.2 shows the system solutions corresponding to two initial conditions.

2.2.2 Zeno Solutions

In principle, the reset control system (2.4) may exhibit Zeno solutions even in the
case where it is well-posed (assuming that condition (2.16) is satisfied). Zeno solu-
tions are solutions to (2.4) that have an infinite number of jumps in a compact time
interval. However, as it will be shown in the following, condition (2.16) is sufficient
to avoid the existence of Zeno solutions.
68 2 Definition of Reset Control System and Basic Results

Fig. 2.2 System solution for the well-posed reset system example

Proposition 2.4 The reset system (2.4)–(2.6) does not have Zeno solutions if MR ∩
N (Obase ) = {0}.
Proof The basic idea of the proof consists of showing that the reset system (2.4)–
(2.6), with an initial condition in MR , can only have finite sequences of reset in-
tervals Δk , k = 1, 2, . . . , m − 1 such as Δm−1 < Δm−2 < · · · < Δ1 = ε, for some
ε > 0 arbitrarily small but fixed, and some finite positive integer m. In fact, in the
following it will be shown that at most there will be sequences of length m − 1, with
m being the dimension of the after-reset surface MR .
Without loss of generality, it is considered that the plant state equations (2.1) are
given in an observer canonical form, that is,
⎛ ⎞ ⎛ ⎞
0 0 ··· 0 −a0 b0
⎜1 0 ··· 0 −a1 ⎟ ⎜ b1 ⎟  
⎜ ⎟ ⎜ ⎟
Ap = ⎜ . . . . . ⎟ , Bp = ⎜ . ⎟ , Cp = 0 0 . . . 1 ,
⎝. .
. . . . .. .
. ⎠ ⎝ . ⎠ .
0 0 · · · 1 −anp −1 bnp −1
(2.28)
then C = (0, 0, . . . , 1, 0, . . . , 0) and thus
⎛ ⎞ ⎛ ⎞
C 0 0 ··· 0 0 1 0 ... 0
⎜ CA ⎟ ⎜ 0 0 · · · 0 1 −anp −1 X . . . X ⎟
⎜ ⎟ ⎜ ⎟
⎜ .. ⎟ ⎜ .. .. .. .. .. .. .. . . .. ⎟
Obase = ⎜ ⎜ ⎟ ⎜ . .⎟
. ⎟=⎜. . . . . . . ⎟,
⎜ CAnp −1 ⎟ ⎜ 1 −an −1 · · · X X X X · · · X ⎟
⎝ ⎠ ⎝ p ⎠
.. .. .. .. . . . . .
.. . . . ..
. . . . .. .. ..
(2.29)
where X stands for a non (necessarily) zero term.
2.2 Zenoness, Beating, and Deadlock 69

For simplicity, the case of full reset is approached at first. Thus, an after-reset
state x ∈ MR is given by

x = (x1 , x2 , . . . , xnp −2 , xnp −1 , 0, 0, . . . , 0)T (2.30)

for some values x1 , . . . , xnp −1 ∈ R, with np being the number of plant states. Thus,
m = np − 1 in the case of full reset.
Let us start with the case m = 1, which corresponds to second order plants and
full reset compensators of arbitrary order. In this case, if the reset control system
is well-posed, it is well known that the reset is periodic after reaching the set MR
from any initial condition; thus starting from MR , the reset will be periodic and
Zeno solutions are not possible. In fact, there exists no initial condition in the set
MR that contacts M in an arbitrarily small time since reset intervals are constant.
The case m = 2 is analyzed in the following. Consider an initial condition x0 =
x1 ∈ MR , that is, x1 = (x1 , x2 , 0, 0, . . . , 0)T . If the solution x(t, 0, x1 ) contacts the
reset surface MR at time t1 = ε1 , thus Δ1 = ε1 , for some ε1 > 0 arbitrarily small,
then
ε12
0 = CeAε1 x1 = Cx1 + ε1 CAx1 + CA2 x1 + · · · . (2.31)
2
Since the control system (2.4)–(2.6) is well-posed, the right-hand side of (2.31)
is not identically zero for any x1 ∈ MR . Now using the special structure given in
(2.29), one obtains
ε1  
0 = x2 + x1 + O ε12 , (2.32)
2
where the terms of order ε12 and higher maybe be neglected, in principle. Note that
for (2.32) to be satisfied for an arbitrarily small ε1 > 0 it must be true that x1 = 0
and x2 = 0.
In addition, the following after-reset state x2 is given by x2 = AR x(t1 , 0, x1 ) =
(x1 + O(ε12 ), x2 + ε1 x1 + O(ε12 ), 0, 0, . . . , 0)T . Repeating the argument, the solution
x(t, t1 , x2 ) will again contact M at the instant t2 = t1 + Δ2 . If Δ2 = ε2 ≤ ε1 for
some ε2 > 0, then it is verified that
ε2  
0 = x2 + ε1 x1 + x1 + O ε12 , (2.33)
2

where the properties O(ε22 ) = O(ε12 ) for ε2 ≤ ε1 and O(kε) = O(ε), for a real con-
stant k, have been used. Now, using (2.32) and (2.33), the result is that given some
ε1 > 0 arbitrarily small, ε2 = −ε1 + O(ε12 ) < 0, which is absurd. Thus, by con-
tradiction it is true that ε2 > ε1 , and thus any initial condition in the set MR that
produces a first reset interval ε1 > 0 arbitrarily small, gives a larger second reset
interval ε2 > 0. Thus Zeno solutions do not exist in this case either.
In the rest of the proof, the terms O(ε1m ) are directly neglected for simplicity. For
the case m = 3, consider an initial condition x0 = x1 = (x1 , x2 , x3 , 0, 0, . . . , 0)T ∈
MR . Applying a similar argument, the result is now that if a sequence of decreasing
70 2 Definition of Reset Control System and Basic Results

reset intervals {ε1 , ε2 , ε3 } exists, with ε1 ≥ ε2 ≥ ε3 > 0 and ε1 being arbitrarily


small, then

ε1 ε2
x3 +
x2 + 1 x1 = 0,
2 6
   2 2 
ε2 ε1 ε1 ε2 ε2
x3 + ε1 + x2 + + + x1 = 0, (2.34)
2 2 2 6
   
ε3 (ε1 + ε2 )2 (ε1 + ε2 )ε3 ε32
x3 + ε1 + ε2 + x2 + + + x1 = 0,
2 2 2 6

and now, eliminating x1 , x2 , and x3 , after some computation, ε3 is given as a func-


tion of ε1 and ε2 by the second order equation

(ε1 + ε2 )ε2 + (ε1 + 2ε2 )ε3 + ε32 = 0 (2.35)

having the solutions ε3 = −ε2 < 0 and ε3 = −(ε1 + ε2 ) < 0, which is a contradic-
tion. Thus, no initial condition in MR can produce a sequence of resets intervals
that converge to zero, and again Zeno solutions do not exist for the case m = 3.
For the general case in which the dimension of MR is m, with initial state x0 =
x1 = (x1 , x2 , . . . , xm , 0, 0, . . . , 0)T , a similar reasoning results in the set of equations


m
ε1m−1
xk = 0,
(m + 1 − k)!
k=1


m 
i
ε2m−i ε1i−k
xk = 0,
(m + 1 − i)!(i − k)! (2.36)
i=1 k=1

...

m 
i
ε m−i (ε1 + · · · + εm−1 )i−k
m
xk = 0,
(m + 1 − i)!(i − k)!
i=1 k=1

which results in an algebraic equation of order m in εm , with the solutions εm =


−εm−1 , εm = −(εm−1 + εm−2 ), . . . , εm = −(εm−1 + εm−2 + · · · + ε1 ). And again,
a sequence of reset intervals {ε1 , ε2 , . . . , εm } with ε1 ≥ ε2 ≥ · · · ≥ εm > 0 and ε1
arbitrarily small cannot exist, showing that a Zeno solution is not possible in the
full-reset case.
The case of partial-reset can be conveniently transformed into the full-reset form
by a change of coordinates, by a simple resorting of coordinates so that the bijectiv-
ity is guaranteed. We will consider the system structure decomposition by writing
the states as xT = (xTp , xρ̄ T , xTρ ) where xp ∈ Rnp stands for the states of the plant,
xρ̄ ∈ Rnρ̄ for the non-resetting compensator states, and xρ ∈ Rnρ for the resetting
2.2 Zenoness, Beating, and Deadlock 71

compensator states. Define the linear transformation T from Rn to Rn such that


⎛ ⎞ ⎛ ⎞
xp xρ̄
T x = T ⎝ xρ̄ ⎠ = ⎝ xp ⎠ = z, (2.37)
xρ xρ

that is,
⎛ ⎞
0nρ̄ ×np Inρ̄ ×nρ̄ 0nρ̄ ×nρ

T = Inp ×np 0np ×nρ̄ 0np ×nρ ⎠ . (2.38)
0nρ ×np 0nρ ×nρ̄ Inρ ×nρ
Note that T is a square matrix, all of whose entries are 0 or 1, and in each row
and column of T there is precisely one 1. This means that T is a permutation matrix.
Clearly, such a matrix is unitary, hence orthogonal, so T T = T −1 . The nonsingular
matrix T allows us to rewrite the dynamical system via a similarity transformation
(congruence transformation):

ż(t) = Āz(t) / M˜,


if z(t) ∈
 +
z t = ĀR z(t) if z(t) ∈ M˜, (2.39)
y(t) = CT z(t) = C̄z(t),
T

where Ā = T AT T , ĀR = T AR T T , and C̄ = CT T , and in addition the reset sur-


face is transformed into M˜ = {z ∈Rn : T T z ∈ M }. Note that C̄ = CT T = enp +nρ̄
so that the output is not changed by the transformation, i.e., y(t) = znp +nρ̄ (t) as
expected. Henceforth, (2.39) is in full-reset form. Since observability is invariant
under similarity transformations, it is clear that (2.4)–(2.6) is well-posed if and only
if (2.39) is well-posed. Finally, to complete the proof it is necessary to show that the
observability matrix has the structure given in (2.29) (using state transformations
x 
if needed). This is simply done by considering the substate z1 = xpρ̄ . In general,
z  w 
there always exists a state transformation of z = xρ1 to w = xρ1 such that the
state submatrix corresponding to z1 is in the observability staircase form, and thus
the observability matrix has the structure given in (2.29) once unobservable states
are eliminated. This concludes the proof. 

2.2.2.1 Example: Well-posed Reset Control System with Partial Reset


x1 
Consider a reset control system (2.4)–(2.6) where the plant, with state xp = x2 , is
given by the state space model
   
0 1 1  
Ap = , Bp = , Cp = 0 1 , (2.40)
1 −1 0
72 2 Definition of Reset Control System and Basic Results
x3 
and the reset compensator, with state xr = x4 , is given by
     
0 0 1   1 0
Ar = , Br = , Cr = 1 1 , Aρ = ,
0 0 1 0 0
(2.41)
that is, the reset control system has a partial reset compensator: it is a parallel con-
nection of an integrator and a Clegg integrator, where only the state x4 is set to zero
at the reset instants.
The closed-loop system is defined by the matrices
⎛ ⎞ ⎛ ⎞
0 1 1 1 1 0 0 0
⎜ 1 −1 0 0 ⎟ ⎜0 1 0 0⎟  
A=⎜ ⎝ 0 −1 0 0 ⎠ ,
⎟ AR = ⎜⎝0 0 1 0⎠,
⎟ C= 0100 ,
0 −1 0 0 0 0 0 0
(2.42)
and the closed-loop state x = ( x1 x2 x3 x4 )T .
This reset control system is well-posed, since
⎧⎛ ⎞ ⎛ ⎞⎫ ⎧⎛ ⎞⎫

⎪ 1 0 ⎪⎪ ⎪
⎪ 0 ⎪ ⎪
⎨⎜ ⎟ ⎜ ⎟⎬ ⎨⎜ ⎬
⎜ 0⎟ ⎜0⎟ ⎜ 0 ⎟ ⎟
MR = span ⎝ ⎠ , ⎝ ⎠ , N (Obase ) = span ⎝ , (2.43)

⎪ 0 1 ⎪⎪ ⎪
⎪ −1 ⎠⎪ ⎪
⎩ ⎭ ⎩ ⎭
0 0 1

and then MR ∩ N (Obase ) = {0}. Following the reasoning given in the proof of
Proposition 2.4, the closed-loop state x can be transformed into a state z in which
the observability matrix has the form (2.29). In this case, this is obtained with z =
( x3 x1 x2 x4 )T . Thus, the initial conditions that produce a crossing in an arbitrarily
small time ε > 0 are of the form z1 = ( 1 − 2ε 0 0 )T , or equivalently,
* ε +T
x1 = − 0 1 0 . (2.44)
2
Now, the second after-reset state is given by
* ε +T * ε +T
x2 = AR eAε − 0 1 0 = 010 , (2.45)
2 2
and, according to Proposition 2.4, x2 cannot produce a new crossing in a time less
than or equal to ε. This fact can be verified by computing solutions to the implicit
equation 0 = CeAt ( α 0 1 0 )T for t, given α ∈ R. The solution is shown in Fig. 2.3,
where t = τ1 (( α 0 1 0 )T ) is given.
Note that for t to have an arbitrarily small value, an initial condition x1 in the
after-reset surface must have the form (2.44), that is, α = −ε/2 in Fig. 2.3. Then, as
a result the state after the first reset x2 has the form (2.45), that is, α = +ε/2 (see
also Fig. 2.3). And then the value of the second reset instant can be obtained from
Fig. 2.3. The result is that if the first reset instant is arbitrarily small, then the second
reset instant is arbitrarily close to 3.15.
2.3 Reset Instants and the After-Reset Surface Dimension 73

Fig. 2.3 First reset instant as a function of α

2.3 Reset Instants and the After-Reset Surface Dimension

In general, reset instants tk = τk (x0 ), k = 1, 2, . . . , can take different and complex


patterns for different initial conditions x0 ∈ D , and in fact this is a key property of
reset control systems since it determines the way in which the reset instants evolve
and also some important properties. For example, the fact that τ1 (x0 ) has a dis-
continuity in the example of the last section (Fig. 2.3) is directly related with the
non-existence of Zeno solutions.
A useful property is that functions τk (x0 ) are homogeneous (of degree 0) since
τk (αx0 ) = τk (x0 ), for any α > 0 and k = 1, 2, . . . . If the set of initial conditions is D ,
this means that the computation of reset intervals can be simplified, for example, to
those states that are elements of the unit ball (centered at the origin). On the other
hand, the set of not reduced states R̄, as defined before Proposition 2.2, can be
obtained by using the spectral projectors of the corresponding eigenvalues.
In the rest of this section, it is assumed that the reset control systems are well-
posed and that their base systems have a state matrix A with a simple pair of com-
plex dominant eigenvalues λ1 = α1 + iβ1 and λ2 = α1 − iβ1 (with index 1). As far as
the computation of the set of not reduced states R̄ is concerned, it may be assumed
that α1 = 0 without loss of generality. Thus, for an initial condition x0 the output
dominant term before the first reset instant is given by (using (2.15))
 
y1 (t) = C Re{G1 } cos(β1 t) + Im{G1 } sin(β1 t) x0 , (2.46)
74 2 Definition of Reset Control System and Basic Results

where G1 is the spectral projector given by G1 = v1 wT1 , and v1 and w1 are the right
and the left eigenvectors, respectively. Thus, the set R̄ can be simply computed as
 
C Re{G1 }
R̄ = N , (2.47)
C Im{G1 }

where Re{G1 } and Im{G1 } stands for the real and imaginary parts of G1 , respec-
tively. And thus the set of reduced states is given simply by Rn \ R̄.
The order of the reset control system, and in particular the dimension of the
after-reset surface, is key in the analysis of the reset instants, thus in the following
different cases corresponding to dimensions of the after-reset surface 1, 2, and ≥ 3
are treated separately.

M R) = 1
2.3.1 dim(M

In this case, since the after-reset surface has dimension 1, any initial condition
x0 ∈ MR can be generated by one vector u, that is, x0 = αu for some α ∈ R. Thus,
simply by using the homogeneity property of the functions tk = τk (x0 ), k = 1, 2, . . . ,
it is clear that τk (x0 ) = τk (αu) = τk (u), that is, they are all constant functions over
the set MR , as long as (C, A, u) is reduced (otherwise there are no crossings). In
other words, starting from an initial condition in the after-reset surface, the reset
instants are periodic.
For initial conditions that are not elements of the after-reset surface, the first reset
instant is in general different. Using Proposition 2.2, if the reset system has a pair of
dominant complex eigenvalues then the first reset instant is uniformly bounded, that
is, τ1 (x0 ) < Δ, for some upper bound Δ > 0 and for any initial condition x0 such
that (C, A, x0 ) is reduced. In the case in which (C, A, x0 ) is not reduced, τ1 (x0 ) is
not necessarily bounded, that is, the initial condition may not cross the reset surface.

2.3.1.1 Example
 
Consider the reset control system (2.4)–(2.6) where the plant, with state xp = xx12 ,
is given by the state space model
   
0 0 1  
Ap = , Bp = , Cp = 0 1 , (2.48)
1 −1 0

and the reset compensator, with state xr = x3 , is given by

Ar = −1, Br = 1, Cr = 1, Aρ = 0, (2.49)

that is, the reset compensator is FORE, and the state x3 is set to zero at the reset
instants.
2.3 Reset Instants and the After-Reset Surface Dimension 75

The closed-loop system is given by the matrices


⎛ ⎞ ⎛ ⎞
0 0 1 1 0 0  
A = ⎝1 −1 0 ⎠ , AR = ⎝ 0 1 0 ⎠ , C= 010 , (2.50)
0 −1 −1 0 0 0

and the closed-loop state is x = ( x1 x2 x3 )T .


It can be easily checked that the system is well-posed since the base linear system
is observable. By definition, the after-reset surface is given by
⎧⎛ ⎞⎫
⎨ 1 ⎬
MR = span ⎝ 0 ⎠ . (2.51)
⎩ ⎭
0

Now, the subspace of not reduced states R̄ is computed. The closed-loop state
matrix A has the eigenvalues λ1 = −0.12 + j 0.74, λ2 = −0.12 − j 0.74, and λ3 =
−1.75. Thus, the reset control system has two complex dominant eigenvalues, with
their spectral projectors being
⎛ ⎞
0.41 − j 0.28 0.15 + j 0.34 0.12 − j 0.41
G1 = ⎝ 0.12 − 0.j 41 0.29 + j 0.14 −0.16 − j 0.34 ⎠ , G2 = G1 .
0.16 + j 0.34 −0.27 + j 0.07 0.29 + j 0.14
C Re{G1 } (2.52)
The subspace of not reduced states is R̄ = N C Im{G1 } , which in this example
has dimension 1 and is given by
⎧⎛ ⎞⎫
⎨ −0.41 ⎬
R̄ = span ⎝ 0.55 ⎠ . (2.53)
⎩ ⎭
0.72

Since, according to (2.51) and (2.53), every nonzero after-reset state is reduced,
it results in that the reset instants are periodic as discussed above. However, note that
if the initial condition is not an after-reset state, two types of solutions may occur
(see Figs. 2.4–2.5):

• solutions with no crossings if


⎛ ⎞
−0.41
x0 = α ⎝ 0.55 ⎠ ,
0.72

for some real number α, and


• solutions with a first crossing at a finite time and then an infinite number of peri-
odic crossings.
76 2 Definition of Reset Control System and Basic Results

Fig. 2.4 Closed-loop output for three different initial conditions

Fig. 2.5 Control input for three different initial conditions

M R) = 2
2.3.2 dim(M

The case of dim(MR ) = 2 is slightly more involved since, as it will be described


below, the resets instants are not periodic in general, and a large variety of solutions
may appear regarding the number and structure of crossings associated to a given
initial condition. By simplicity, the particular case of a third order plant and a full
reset compensator is considered. Without loss of generality, consider an observer
2.3 Reset Instants and the After-Reset Surface Dimension 77

canonical form realization of the plant, that is,


⎛ ⎞ ⎛ ⎞
0 0 −a0 b0  
Ap = ⎝ 1 0 −a1 ⎠ , Bp = ⎝ b1 ⎠ , Cp = 0 0 1 . (2.54)
0 1 −a2 b2

Thus an after-reset state x ∈ MR is given by

x = (x1 , x2 , 0, 0, . . . , 0)T (2.55)


 
for some values xx12 ∈ R2 . Using again the fact that tk = τk (x0 ), k = 1, 2, . . . , are
homogeneous, for the systems solutions with an initial condition in the after-reset
surface it is enough to check the reset instants
x1  in some subset of R2 , for example,
ρ cos θ 
the unit circle. Using polar coordinates x2 = ρ sin θ , for ρ ∈ [0, ∞), θ ∈ [0, 2π),
it is true that
⎛⎛ ⎞⎞ ⎛⎛ ⎞⎞
ρ cos θ cos θ
⎜⎜ ρ sin θ ⎟⎟ ⎜⎜ sin θ ⎟⎟
⎜⎜ ⎟⎟ ⎜⎜ ⎟⎟
⎜⎜ 0 ⎟⎟ ⎜⎜ ⎟⎟
τk ⎜⎜ ⎟⎟ = τk ⎜⎜ 0 ⎟⎟ (2.56)
⎜⎜ .. ⎟⎟ ⎜⎜ .. ⎟⎟
⎝⎝ . ⎠⎠ ⎝⎝ . ⎠⎠
0 0
for k = 1, 2, . . . . As a result the reset instants are functions of the single parameter θ .
In general, the system’s solutions may exhibit no crossings, a finite number of
crossings, or infinitely many crossings. If the initial condition is an after-reset state,
that is, x0 ∈ MR , no crossings or a finite number of crossings may occur if some of
the after-reset states are not reduced. Otherwise, an infinite number of crossings is
produced.

2.3.2.1 Example

Consider again the reset control system with a FORE compensator, with parameters
Ar = −1, Br = 1, and Cr = 1, and a plant given by the state space model
⎛ ⎞ ⎛ ⎞
0 0 −0.35 3  
Ap = ⎝ 1 0 −2.40 ⎠ , Bp = ⎝ 1 ⎠ , Cp = 0 0 1 . (2.57)
0 1 −4.35 0

The reset control system is well-posed since the base linear system is observable.
Using the parameter θ as in (2.56), the function τ1 can be computed by solving the
implicit equation
⎛ ⎞
cos θ
⎜ sin θ ⎟
CeAt1 ⎜⎝ 0 ⎠=0
⎟ (2.58)
0
78 2 Definition of Reset Control System and Basic Results

Fig. 2.6 Reset instants corresponding to after-reset states as a function of the parameter θ

for t1 = τ1 (θ ). The result is given in Fig. 2.6, where it can be seen that the mapping
τ1 has a discontinuity at θ = π , and in addition it results in the reset instants being
uniformly bounded for initial conditions in the after-reset surface. This is due to the
fact that all the after-reset states are reduced, as it will be seen below.
The after-reset surface is given by
⎧⎛ ⎞ ⎛ ⎞⎫

⎪ 1 0 ⎪ ⎪
⎨ ⎜ 0 ⎟ ⎜ 1 ⎟⎬
MR = span ⎜ ⎟,⎜ ⎟
⎝ 0 ⎠ ⎝ 0 ⎠⎪ (2.59)

⎪ ⎪
⎩ ⎭
0 0

and the subspace of not reduced states by


⎧⎛ ⎞ ⎛ ⎞⎫
  ⎪
⎪ 0.20 −0.81 ⎪

⎨⎜ ⎟⎬
C Re{G1 } 0.31 ⎟ ⎜
⎟ , ⎜ 0.47 ⎟ ,
R̄ = N = span ⎜⎝ 0.92 ⎠ ⎝ −0.05 ⎠⎪ (2.60)
C Im{G1 } ⎪
⎪ ⎪
⎩ ⎭
0.16 0.35

thus it can be easily checked that MR ∩ R̄ = {∅}, which means that all the after-reset
states are reduced. In Figs. 2.7 and 2.8, several simulations corresponding to differ-
ent initial conditions in the after-reset surface MR are shown, including closed-loop
outputs and control inputs. Note that since the after-reset states are reduced, there are
an infinite number of resets corresponding to each initial condition, and as indicated
above, a bound over the reset intervals may be found.
In the case in which the initial condition is not an after-reset surface state, it may
occur that no crossings are produced. This is, in fact, the case for initial conditions in
the set R̄ given by (2.60). Otherwise, an infinite number of crossings are produced
2.3 Reset Instants and the After-Reset Surface Dimension 79

Fig. 2.7 Closed-loop output for three initial conditions in the after-reset surface

Fig. 2.8 Control input for three initial conditions in the after-reset surface

in this example. Different simulations for not reduced initial conditions are shown
in Figs. 2.9–2.10
80 2 Definition of Reset Control System and Basic Results

Fig. 2.9 Closed-loop output for three initial conditions (not reduced)

Fig. 2.10 Control input for three initial conditions (not reduced)

M R) ≥ 3
2.3.3 dim(M

For higher order reset control systems, the functions τk , k = 1, 2, . . . , depend on


more than one parameter, and in general they can be described by the values of τ1
over the unit ball in Rm , where m is the number of the after-reset states. In general,
the function τ1 can exhibit an infinite number of discontinuities over that domain
2.3 Reset Instants and the After-Reset Surface Dimension 81

even in the case m = 2, resulting in very complex patterns of the resets instants as a
function of the initial condition. A detailed analysis of these patterns is still an open
issue. In the following, an example corresponding to a second order partial reset
compensator and a third order plant is analyzed.

2.3.3.1 Example

In this example, a PI + CI compensator is used. It consists of a parallel connection


of a PI compensator and a Clegg integrator, and it is a partial reset compensator.
Consider the PI + CI compensator with a state space realization given by
   
0 0 1
Ar = , Br = ,
0 0 1
 
  1 0
Cr = 0.2 0.2 , Dr = 0.2, Aρ = (2.61)
0 0

and a third order plant given by


⎛ ⎞ ⎛ ⎞
−2.20 −1.32 −0.72 1  
Ap = ⎝ 2 0 0 ⎠, Bp = ⎝ 0 ⎠ , Cp = 0 0 1.33 .
0 1 0 0
(2.62)
In addition, closed-loop state matrices are
⎛ ⎞
−2.20 −1.32 −0.99 0.20 0.2
⎜ 2 0 0 0 0 ⎟
⎜ ⎟  

A=⎜ 0 1 0 0 0 ⎟ C = 0 0 1.33 0 0 , (2.63)
⎟,
⎝ 0 0 −1.33 0 0 ⎠
0 0 −1.33 0 0

and
⎛ ⎞
1 0 0 0 0
⎜0 1 0 0 0⎟
⎜ ⎟
AR = ⎜
⎜0 0 1 0 0⎟⎟. (2.64)
⎝0 0 0 1 0⎠
0 0 0 0 0
In this case, the base system is unobservable, with the unobservable subspace in
this case being given by
⎧⎛ ⎞⎫

⎪ 0 ⎪ ⎪
⎪⎜
⎪ ⎪
⎨⎜ 0 ⎟ ⎟⎬


N (Obase ) = span ⎜ 0 ⎟ ; ⎟ (2.65)

⎪ ⎝ −1 ⎠⎪ ⎪

⎪ ⎪

⎩ ⎭
1
82 2 Definition of Reset Control System and Basic Results

Fig. 2.11 Closed-loop output for three initial conditions in the after reset surface

however, the reset control system is well-posed since the after-reset surface is
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫

⎪ 1 0 0 ⎪
⎪⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎪
⎪ ⎪
⎨⎜ 0 ⎟ ⎜ 1 ⎟ ⎜ 0 ⎟⎪ ⎬
⎜ ⎟ ⎜ ⎟ ⎜
MR = span ⎜ 0 ⎟ , ⎜ 0 ⎟ , ⎜ 0 ⎟ ⎟ (2.66)
⎪⎝ ⎠ ⎝ ⎠ ⎝ ⎠⎪
⎪ ⎪

⎪ 0 0 1 ⎪

⎩ ⎭
0 0 0

and MR ∩ N (Obase ) = {0}. In addition, it can be checked that the after reset states
are not reduced, in other words, assuming initial conditions in the after-reset surface
there are always an infinite number of crossings.
In Fig. 2.11, closed-loop outputs corresponding to different initial conditions are
obtained. Also in Fig. 2.12 the different control inputs are shown. Note that in gen-
eral, as the dimension of the after-reset surface increases, the structure of the reset
instants becomes not that simple, for example, in the interval [0, 15]s one initial con-
dition does not produce a crossing while the others produce five reset actions. As a
conclusion, as the dimension of the after reset surface increases, the reset instants
occur with more complex patterns.

2.4 Reset Control Systems with Exogenous Inputs

As it is usual in control practice, reset control systems are driven by external or ex-
ogenous inputs such as reference or disturbance signals. In this case, well-posedness
of the reset control system can be analyzed using the arguments given in previous
sections, if the exogenous inputs are generated by an exosystem.
2.4 Reset Control Systems with Exogenous Inputs 83

Fig. 2.12 Control input for three initial conditions in the after-reset surface

Consider the reset control system of Fig. 2.1, where the plant and the reset com-
pensator are given by (2.1) and (2.2), respectively, and with a reference input r and
a disturbance input d generated by exosystems, with the state space models

ẇ1 (t) = A1 w1 (t), w1 (0) = w10 ,
(2.67)
r(t) = C1 w1 (t),

with w1 ∈ Rm1 , and



ẇ2 (t) = A2 w2 (t), w2 (0) = w20 ,
(2.68)
d(t) = C2 w2 (t),

with w2 ∈ Rm2 . These exosystems allow the generation of signals like steps, ramps,
sinusoids, . . . . Now, the feedback connection is given by e = r − y and u = v + d,
and the base linear closed-loop system may be described by
   
0 0
ẋ(t) = Ax(t) + r(t) + d(t) (2.69)
Br Bp
 
where x = xxpr , and xp and xr are the plant and compensator states. In addition,
the reset instants are defined as those instants t at which the closed-loop output
y(t) = Cx(t) is equal to the reference signal r(t) = C1 w1 (t). Define the augmented
state z as z = ( wT1 wT2 xT )T . Then, the reset map can be defined in the augmented
84 2 Definition of Reset Control System and Basic Results

state space by
⎛ ⎞
I 0 0 0
⎜0 I 0 0 ⎟
ĀR = ⎜
⎝0
⎟ (2.70)
0 I 0 ⎠
0 0 0 Aρ
and the after-surface M¯R and the reset surface M¯ as

M¯R = R(ĀR ) ∩ N (C̄),


(2.71)
M¯ = N (C̄) \ M¯R

with C̄ = ( C1 0 −Cp 0 ). Finally, the closed-loop system in the augmented state


space is given by

ż(t) = Āz(t) if z(t) ∈
/ M,
(2.72)
z(t + ) = ĀR z(t) if z(t) ∈ M
with the state space matrix
⎛ ⎞
A1 0 0 0
⎜ 0 A2 0 0 ⎟
Ā = ⎜
⎝ 0
⎟. (2.73)
Bp C2 Ap Bp Cr ⎠
Br C1 0 −Br Cp Ar

In the augmented space state representation, Propositions 2.3 and 2.4 can be
directly applied, giving the next result.

Proposition 2.5 Consider the reset control system of Fig. 2.1, with the plant and
reset compensator given by (2.1) and (2.2), respectively, and with inputs r and d
generated by the exosystems (2.67)–(2.68). If M¯R ∩ N (Ōbase ) = {0}, where Ōbase
is the observability matrix
⎛ ⎞

⎜ C̄ Ā ⎟
⎜ ⎟
Ōbase = ⎜ .. ⎟ (2.74)
⎝ . ⎠
C̄ Ān+m1 +m2 −1

then
1. The reset system is well-posed.
2. The reset system does not have Zeno solutions.
 
Proof The augmented state can be partitioned as z = xz1r , and thus the state matrix
(2.73) is partitioned as
 
Āp B̄p Cr
Ā = , (2.75)
Br C̄p Ar
2.4 Reset Control Systems with Exogenous Inputs 85

where
⎛ ⎞ ⎛ ⎞
A1 0 u0 0  

Āp = 0 A2 u 0 ⎠ , B̄p = ⎝ 0 ⎠ , C̄p = C1 0 −Cp . (2.76)
0 Bp C2 uAp Bp

Part 1 of the proposition is a direct application of Proposition 2.3 to the system (2.72)
defined by the matrices Ā, ĀR , and C̄. On the other hand, since the state matrix Ā
has the structure given in (2.75), it always possible to make a state transformation of
z1 in such a way that corresponding principal minor be in the observer form. Thus,
the reasoning used in the proof of Proposition 2.4 may be used to prove Part 2. 

In general, a simple sufficient condition for the well-posedness of a reset control


system with exogenous inputs it that the augmented closed-loop system (2.72) be
observable, that is, the matrix Ōbase be full rank.

2.4.1 A Well-posed Reset Control System with Exogenous Input

A simple example is shown here to illustrate Proposition 2.5. Consider the reset
control system of Fig. 2.1, consisting of the feedback interconnection of a Clegg
integrator and an integrator. The Clegg integrator has the state equations


⎪ ẋr (t) = r(t), xr (t) − r(t) = 0,

xr (t + ) = 0, xr (t) − r(t) = 0, (2.77)



v(t) = xr (t).

In addition, consider a sinusoidal reference input r(t) = a sin(ωt + φ), for some
given constants a, ω > 0, and φ. It is given by the exosystem
  0 ω  a sin φ 
ẇ1 (t) = −ω 0
w1 (t), w1 (0) = a cos φ ,
(2.78)
r(t) = ( 1 0 )w1 .

Since a disturbance signal is not considered in this example, Proposition 2.5 can
be used by eliminating the row and column blocks corresponding to the disturbance
exosystem in the matrices Ā, ĀR , and C̄. The result is
⎛ ⎞
0 ω 0 0
⎜ −ω 0 0 0⎟  
Ā = ⎜
⎝ 0
⎟, C̄ = 1 0 −1 0 . (2.79)
0 0 1⎠
1 0 −1 0
86 2 Definition of Reset Control System and Basic Results

Fig. 2.13 Reference signal and closed-loop output corresponding to a well-posed reset control
system

Now, the observability matrix of the (augmented) base system is


⎛ ⎞
1 0 −1 0
⎜ 0 ω 0 −1 ⎟
Ōbase = ⎜
⎝ −(1 + ω2 )
⎟ (2.80)
0 1 0 ⎠
0 −ω(1 + ω2 ) 0 1

which is full rank for any ω > 0 (and does not depend on a), and thus the sys-
tem is well-posed for any sinusoidal reference input. Figures 2.13 and 2.14 show
a simulation of the reset control system for ω = 0.5 rad/s, with an initial condition
w1 (0) = (1 − 1)T for the exosystem, and with zero initial condition for the Clegg
integrator and the integrator.

2.4.2 A Reset Control System with Zeno Solutions

In general, Proposition 2.5 gives a simple and checkable condition for well-
posedness of the reset control system of Fig. 2.1 with exogenous inputs, for the
plant and compensator given by (2.1) and (2.2), respectively. For well-posedness,
the result may be also used for any reset system that can be expressed as (2.72) with
arbitrary values of Ā, A¯R , and C̄, since no particular structure of these matrices is
used to prove the result. However, for avoiding Zeno solutions the structure of these
matrices, related with (2.1) and (2.2), is a central part of the result. Thus, a reset
control system with a different structure may have Zeno solutions, this is the case,
for example, corresponding to a non-strictly proper plant in Fig. 2.1.
2.4 Reset Control Systems with Exogenous Inputs 87

Fig. 2.14 Control input corresponding to a well-posed reset control system

In the following, an example developed in [6] is shown here to illustrate the


existence of Zeno solutions in reset control systems with external inputs, where the
plant is simply P (s) = 1. The resulting feedback system is a Clegg integrator with
unity feedback. In addition, a sinusoidal reference input is also considered. The
state space models (2.77) and (2.78) are again used, as a result the augmented state
matrices are in this case (the row and column block corresponding to the plant are
simply removed)
⎛ ⎞ ⎛ ⎞
0 ω 0 1 0 0  
Ā = ⎝ −ω 0 0 ⎠ , ĀR = ⎝ 0 1 0 ⎠ , C̄ = 1 0 −1 , (2.81)
1 0 −1 0 0 0

and the observability matrix of the augmented base system is given by


⎛ ⎞
1 0 −1
Ōbase = ⎝ −1 ω 1 ⎠ (2.82)
1 − ω −ω −1
2

which is full rank for any ω > 0. Although Proposition 2.5 cannot be used in this
case because the plant does not fit the model (2.2), the well-posedness can be assured
since Ōbase is full rank. However, the well-posedness argument cannot be used for
assessing the existence of Zeno solutions. In fact, this system has a Zeno solution
with sequences of reset instants {tk }, k = 1, 2, . . . , converging to t ∗ = nπ for every
integer n ≥ 1.
Figure 2.15 (top) shows the output response y(t) of the closed loop (and output of
the Clegg integrator) to the sinusoidal reference w(t) = sin ωt. It can be shown that
a Zeno behavior appears to the left of every t = kπ , k = 1, 2, 3, . . . . The solution
88 2 Definition of Reset Control System and Basic Results

Fig. 2.15 (Top) Response y(t) to a sinusoidal reference of a feedback loop with a Clegg integrator;
(Bottom) detailed zoom of the three first visible resets close to t = π

starting from the right of every y(kπ) = 0 escapes the Zeno behavior until it again
reaches another Zeno point at t = (k + 1)π .
Figure 2.15 (bottom) plots a detailed zoom of the first three visible resets close to
t = π . These reset times are t1 ≈ 2.28, t2 ≈ 2.92, and t3 ≈ 3.12. In fact, there exists
an infinite sequence of resets {tk } converging to t∞ = π .
It is illustrative to see in Fig. 2.16 the time response of this system in a three-
dimensional plot that shows the time evolution of the state (w11 , w12 , xp ) =
(r, ṙ, y) . The first two coordinates are from the exosystem, and the first one is
the reference r(t) to the closed loop, in this case r(t) = sin ωt and ṙ(t) = cos ωt,
with ω = 1. Thus the trajectory lies entirely within the cylinder r 2 + ṙ 2 = 1.
The three-dimensional plot reveals clearly that, in the state space (r, ṙ, y),
there are actually two Zeno accumulation points, namely Z1 = (0, 1, 0) and Z2 =
(0, −1, 0). The first one corresponds in the time domain (Fig. 2.15) to the zero
crossings of the reference for t = 2kπ and the second one for t = (2k + 1)π , with k
integer.
Figure 2.16 also shows two relevant planes: the null space N (C) of C =
(1, 0, −1), that is, the plane y = r that triggers the reset action, and the plane MR
given by y = 0 where the state is projected immediately after a reset y(tk+ ) = 0.
2.4 Reset Control Systems with Exogenous Inputs 89

Fig. 2.16 Trajectory (r(t), ṙ(t), y(t)) with Zeno behavior

The analysis of Zeno behavior in dynamical systems is an involved topic, and


we will not address it here rigorously. Notice from the previous proposition in this
section that the reset control systems that we are interested in (well-posed) do not
have Zeno behavior. However, in order to give an idea of the asymptotic behavior
close to the Zeno limit, let us complete the example with a simplified study based
on the Poincaré map. We refer the interested reader to the literature, for example,
[8], [7], and [13] for a full exploration of these features.
From Fig. 2.16 it is clear that from any initial condition the state evolves until it
reaches the reset condition at (rk , ṙk , yk ) with yk = rk and then resets to the point
xk = (rk , ṙk , 0). Then, it flows again until a new reset condition holds and a new
reset moves the state to xk+1 = (rk+1 , ṙk+1 , 0). The sequence of after-reset points
{xk } defines a discrete-time map, or iteration xk → xk+1 , called the Poincaré map
of the reset system. Note that, since rk2 + ṙk2 = 1 for all k, the Poincaré sequence
{xk } ∈ P is here defined on a one-dimensional manifold P given by r 2 + ṙ 2 = 1
and y = 0.
Since P is one-dimensional, the Poincaré map can be determined also from the
time domain plots of r(t) and y(t) in Fig. 2.15, for if we determine the sequence of
reset times {tk }, then the Poincaré sequence is {xk } = {(sin tk , cos tk , 0)}. In Fig. 2.15
(bottom), we see the first three reset times t1 ≈ 2.28, t2 ≈ 2.92, and t3 ≈ 3.12. These
tk tend from the left to t∞ = π . Let us suppose that we start at t = tk with y(tk ) = 0.
90 2 Definition of Reset Control System and Basic Results

The evolution of the Clegg integrator state is governed by:

ẏ = r(t) − y(t) = sin t − y(t), y(tk ) = 0.

Since tk tends to π , a key simplifying assumption is that, in the limit, we can ap-
proximate sin t by −(t − π). This simplification is also helpful because the Poincaré
maps give rise to implicit equations not solvable analytically. In this way, close to
t = π , we have:
ẏ = −(t − π) − y(t), y(tk ) = 0,
having the solution

y(t) = (tk − π − 1)e−(t−tk ) − t + π + 1, t ≥ tk .

To determine the next reset time tk+1 , we have to impose the condition y(tk+1 ) =
r(tk+1 ). Again we replace r(tk+1 ) = sin tk+1 by −(tk+1 −π). Thus the Poincaré map
tk → tk+1 is given implicitly by the condition

−(tk+1 − π) = (tk − π − 1)e−(tk+1 −tk ) − tk+1 + π + 1.

Introduce the change of variables tk − π = dk , with dk → 0− . Using the asymp-


totic approximation e−(tk+1 −tk ) = e−(dk+1 −dk ) ≈ 1 − (dk+1 − dk ) gives rise to

−dk+1 = (dk − 1)(1 − dk+1 + dk ) − dk+1 + 1,

from which we can solve explicitly dk+1 = dk2 /(dk − 1). This law for dk → 0−
approaches dk+1 = −dk2 , or equivalently, approaches the limit relation

|tk+1 − π| = |tk − π|2

which proves that there exists an infinite sequence of reset times tk tending to π and
governed, in the limit, by a quadratic recurrence law. This explains why in Fig. 2.15
there are only a few visible resets: to detect the reset at tk = π + dk , we should
implement a simulation step size smaller than |dk |, that, from the quadratic law
|dk+1 | = |dk |2 , decreases very fast with k, implying a strong computational cost.

References
1. Bainov, D.D., Simeonov, P.S.: Systems with Impulse Effect: Stability, Theory and Applica-
tions. Ellis Horwood Limited, Chichester (1989)
2. Baños, A., Mulero, J.I.: On the well-posedness of reset control systems. Technical Report
TR-DIS-1-2011, University of Murcia (2011)
3. Beker, O., Hollot, C.V., Chait, Y., Han, H.: Fundamental properties of reset control systems.
Automatica 40, 905–915 (2004)
4. Bell, P.C., Delvenne, J.-C., Jungers, R.M., Blondel, V.D.: The continuous Skolem–Pisot
problem: on the complexity of reachability for linear ordinary differential equations (2009).
arXiv:0809.2189v2 [maths.DS]
References 91

5. Branicky, M.S., Borkar, V.S., Mitter, S.K.: A unified framework for hybrid control: model and
optimal control theory. IEEE Trans. Autom. Control 43(1), 31–45 (1998)
6. Carrasco, J.: Stability of reset control systems. Ph.D. Thesis (2009) (in Spanish)
7. Grizzle, J.W., Abba, G., Plestan, F.: Asymptotically stable walking for biped robots. Analysis
via systems with impulse effects. IEEE Trans. Autom. Control 46, 51–64 (2001)
8. Haddad, W.M., Chellaboina, V., Nersesov, S.G.: Impulsive and Hybrid Dynamical Systems:
Stability, Dissipativity, and Control. Princeton University Press, Princeton (2006)
9. Hainry, E.: Reachability in linear dynamical systems. In: Computability in Europe 2008. Lec-
tures Notes in Computer Science, vol. 5028. Springer, Berlin (2008)
10. Lakshmikanthan, V., Bainov, D.D., Simeonov, P.S.: Theory of Impulsive Differential Equa-
tions. World Scientific, Singapore (1989)
11. Michel, A., Hu, B.: Towards a stability theory of general hybrid dynamical systems. Automat-
ica 35, 371–384 (1999)
12. Nešić, D., Zaccarian, L., Teel, A.R.: Stability properties of reset systems. In: Proc. of 16th
IFAC World Congress, Prague, Czech Republic (2005)
13. Schumacher, J.M.: Time-scaling symmetry and Zeno solutions. Automatica 45, 1237–1242
(2009)
14. Yang, T.: Impulsive Control Theory. Lecture Notes in Control and Information Sciences,
vol. 272. Springer, Berlin (2001)
Chapter 3
Stability of Reset Control Systems

3.1 Introduction
Stability is a main concern in control systems, and thus stability of reset control
systems is a key issue both from an analysis and a design point of view. In the
classical works of Clegg and Horowitz, the main stability tool was the describing
function that, although giving only approximate results, provides information about
the phase lead characteristic of reset compensators. In the seminal work [18], the sta-
bility problem is explicitly formulated, concluding that the design problem “will be
solved more satisfactorily when stability criteria are developed for feedback loops
containing nonlinear elements such as FORE ( first order reset element)”.
In the last decade, the stability problem has been approached from different per-
spectives, and nowadays a number of results are available for reset control systems,
including systems with time delays. A basic question is about the relationship be-
tween the stability of the reset control system and its base linear system. It is impor-
tant to emphasize that the stability of a reset control system is not always guaranteed
by the stability of the base system; in other words, reset actions can destabilize a sta-
ble base linear system, thus reset should be carefully used in practice. Conversely,
a proper reset control design may stabilize an unstable base linear system, thus sta-
bility of the base linear system is neither sufficient nor necessary for the stability of
the reset control system.
This chapter addresses the stability of reset control systems with finite-dimen-
sional base systems. Section 3.2.1 deals with reset-times independent conditions,
that is, stability conditions that are not explicitly based on the reset instants, includ-
ing the Hβ condition developed in [9]. In Sect. 3.2.2, we study stability conditions
depending on the reset time intervals and based on the induced discrete-time sys-
tem obtained by sampling the reset control system at the after-reset instants. This
approach is further extended in Sect. 3.2.3 to the stabilization problem, imposing
stabilizing upper bounds (resp., lower bounds) on the reset intervals, that is, forc-
ing (resp., avoiding) reset actions. Section 3.3 considers input–output L2 stability
by using passivity-based techniques and providing a number of useful results on
the passivity and dissipativity of different reset systems. Finally, the chapter con-
cludes with Sect. 3.4 where an analysis by the describing function is presented,

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 93


DOI 10.1007/978-1-4471-2250-0_3, © Springer-Verlag London Limited 2012
94 3 Stability of Reset Control Systems

applied to different types of reset compensators, including fixed/variable reset band


and variable reset. The describing function is especially useful for evaluating in the
frequency domain the overcoming of linear modulus-phase limitations, showing the
phase lead property of reset compensators.

3.2 Lyapunov Stability


In this section, the stability of the autonomous reset control system (2.4)–(2.6) will
be analyzed, assuming that it is well-posed. For any equilibrium point, both the
(continuous) linear base dynamics and the (discrete) resetting dynamics need to be
at an equilibrium, that is, for any equilibrium point xe of (2.4), both ẋe = 0 and
xe = AR xe are satisfied. In general, the origin xe = 0 will be the equilibrium point
of the reset control system (2.4). Note that xe = 0 is an element of the after reset
surface MR , and thus 0 ∈ / M.
The following stability definitions are standard, with the notation adopted from
[9] and [27]. The equilibrium point xe = 0 of the reset system, or simply the reset
system, is stable if for any ε > 0 there exists a δ(ε) such that if x0  < δ then
x(t, x0 ) < ε for all t ≥ 0, where x(t, x0 ) is the solution of (2.4) at instant t and
for the initial condition x(0) = x0 . On the other hand, the equilibrium point xe =
0 of the reset system, or again simply the reset system, is asymptotically stable
if it is stable and, in addition, there exists some number η such that if x0  < η
then x(t, x0 ) → 0 as t → ∞. Since the origin xe = 0 is the only equilibrium point,
stability is equivalent to global stability.
The stability of a discrete-time linear system given by x(k + 1) = A(k)x(k),
x(k0 ) = x0 is considered following [27]. It is uniformly stable if there exists a fi-
nite positive constant γ such that for any k ≥ k0 and any initial condition x0 the
corresponding solution satisfies x(k) ≤ γ x0 . It is uniformly asymptotically sta-
ble if, in addition, there exists a constant 0 ≤ λ < 1 such that for any k0 and x0 the
solution satisfies x(k) ≤ γ λk−k0 x0 , for any k ≥ k0 .

3.2.1 Reset-Times Independent Conditions

Recently, the reset control systems stability problem has been successfully ad-
dressed in [9], [10] for general reset compensators, allowing full or partial state
reset. This work is significant since stability can be checked by a simple condi-
tion: the (strictly) positive realness of a transfer functions matrix Hβ , referred to
as the Hβ -condition. The Hβ -condition is based on a quadratic Lyapunov function
that must be decreasing over all the state space along the system trajectories and
non-increasing on the reset jumps.
In [24], the condition has been relaxed by using a different definition of reset and
piecewise quadratic Lyapunov functions. However, this definition of reset system is
not equivalent to the original (see Sect. 1.6) and it will not be developed here.
3.2 Lyapunov Stability 95

3.2.1.1 The Hβ -condition

The Hβ stability condition is applied to the reset control system (2.4), where in
principle the reset surface M is defined by
 
M = x ∈ Rn : Cx = 0, (I − AR )x = 0 , (3.1)

to avoid beating. According to this definition, M is the set of closed-loop states


in the null space N (C) that does not contain fixed points of AR . Note that this
definition is equivalent to (2.6). In addition, it will be assumed that the system is
well-posed, that is, MR ∩ Obase = {0}, to avoid deadlock and also the existence of
Zeno solutions.
The basic approach for addressing stability is a Lyapunov theorem that requires
the Lyapunov function to be decreasing in the continuous and jump modes, but with
no explicit assumption on the reset times [9]. In this way, the derived results are
reset-times independent:

Proposition 3.1 Let V : Rn → R be a continuously differentiable, positive-definite,


radially unbounded function such that
 
∂V
V̇ (x) := Ax < 0, for x = 0, (3.2)
∂x

and
ΔV (x) := V (AR x) − V (x) ≤ 0, for x ∈ M . (3.3)
Then the reset control system (2.4)–(2.6) is asymptotically stable.

This basic result can be specialized to the quadratic Lyapunov case. The reset
control system (2.4)–(2.6) is said to be quadratically stable if it satisfies (3.2)–(3.3)
for some V (x) = x P x with P > 0. Recall that the row C has the structure

C = (Cp , 01×nρ̄ , 01×nρ ), (3.4)

where Cp is 1 × np and np is the number of plant states. The number of controller


states is nρ̄ + nρ , where nρ is the number of controller reset states (note that nρ̄ > 0
in the partial reset case). Thus, the reset matrix takes the form

AR = diag(0np ×np , 0nρ̄ ×nρ̄ , Inρ ×nρ ), (3.5)

and the following stability result is obtained, where the matrices C0 and B0 are
defined by
⎛ ⎞
0np ×nρ
C0 = (βCp , 0nρ ×nρ̄ Pρ ), B0 = ⎝ 0nρ̄ ×nρ ⎠ . (3.6)
Inρ ×nρ
96 3 Stability of Reset Control Systems

Proposition 3.2 The following conditions are equivalent:


(i) The reset system (2.4)–(2.6) is quadratically stable.
(ii) There exists a constant β ∈ Rnρ ×1 such that the restricted Lyapunov equation

P > 0, A P + P A < 0, (3.7)

B0 P = C0 (3.8)
has a solution for P .
(iii) (Hβ -condition) There exist a positive-definite matrix Pρ ∈ Rnρ ×nρ and β ∈
Rnρ ×1 such that
Hβ (s) := C0 (sI − A)−1 B0 (3.9)
is a strictly positive real (SPR) transfer function, where (A, B0 ) is controllable
and (A, C0 ) is observable.

Proof The complete proof can be found in [9], but the main steps are as follows.
First, the equivalence between (3.2) and (3.7) is immediate. Second, (3.3) can be
expressed in the form x (AR P AR − P )x ≤ 0 for all x ∈ M , which is equivalent,
by continuity, to the same inequality for all x ∈ N (C). Processing this inequality
gives rise to (3.8).
Finally, the equivalence between (3.7), (3.8), and (3.9) follows from the well-
known Kalman–Yakubovich–Popov lemma, or positive real lemma (see, for exam-
ple, [19]), which is a standard tool for passing between time-domain (Lyapunov)
and frequency domain conditions. 

Recall that the transfer function matrix Hβ (s), with dimension nρ × nρ , is strictly
positive real when Hβ (s −ε) is positive real for some ε > 0. In our case, since Hβ (s)
is strictly proper, the Hβ -condition amounts to [19]:
• Hβ (s) is Hurwitz, that is, all poles of elements of Hβ (s) have negative real parts;
• Hβ (j ω) + Hβ (−j ω) > 0 for all finite ω ∈ R;
• limω→∞ ω2 (Hβ (j ω) + Hβ (−j ω)) > 0.
The availability of frequency domain stability conditions, like the SPR condition
on Hβ above, is important for many well-known reasons: integration of analysis
and design, treatment of uncertainty, etc. Note that the Hβ -condition gives quadratic
stability by checking if Hβ is SPR for some β and some Pρ . Since the Hβ -condition
is easily checkable especially in the scalar case corresponding to the reset of only
one state (nρ = 1), it has been proved to be an efficient method for stability analysis
of reset control systems. In general, Hβ (s) will be a transfer function matrix with
dimension nρ × nρ , that is, with size equal to the number of reset states.
The Hβ -condition results in that the quadratic Lyapunov function V (x) = xT P x
is decreasing over the time intervals (tk , tk+1 ) and non-increasing at the reset in-
stants tk , k = 1, 2, . . . . Similar stability conditions have also been previously used
in [23] and the seminal work [2], in the context of impulsive systems with impulses
at fixed instants.
3.2 Lyapunov Stability 97

In spite of its usefulness, the Hβ -condition has some limitations. Firstly, since
the quadratic function V (x) has a continuous derivative over Rn , it is clear that (3.2)
is equivalent to xT (AT P + P A)x < 0 for all x ∈ Rn and x = 0, and thus the base
linear system must be always asymptotically stable. Secondly, note that even in the
case in which the base linear system and the reset system are asymptotically stable,
it might happen that the Hβ condition does not hold. This is, for example, the case if
the reset action induces small positive increments ΔV (x) > 0 that are compensated
by large decrements between reset instants, due to V̇ (x) < 0, so that the net effect is
a net decrement of V (x), and thus the reset system is stable.
As a result, the Hβ -condition has two basic limitations: (i) it does not apply to
reset systems with unstable base system, and (ii) it is only a sufficient condition,
and thus there exist stable reset systems for which an SPR Hβ does not exist. In
the following, two examples are developed: an example in which the Hβ -condition
gives stability of a reset system; and a second example of a stable reset system with
a stable base system in which the Hβ -condition does not apply.

Example: Hβ is SPR Consider Example 6 in [10], which uses a reset control sys-
tem with plant P (s) = 1/s and base linear controller Rb (s) = 1/(s + 1), connected
in negative feedback, without exogenous inputs. If the state vector is x = (x1 , x2 )
with x1 the plant state and x2 the (reset) controller state, then it results in a reset
system like that in (2.4)–(2.6) with
   
0 1 1 0  
A= , AR = , C= 10 . (3.10)
−1 −1 0 0

Since the base linear system is observable, the reset control system is well-posed.
In addition, C0 and B0 are given by
 
0  
B0 = , C0 = β 1 , (3.11)
1

and thus it is easily obtained that (A, B0 ) is controllable and (A, C0 ) is observable.
In addition, from (3.9), Hβ is simply given by (for this case nρ = 1 and then Pρ = 1
without loss of generality):
 
  −1 0 s +1−β
Hβ (s) = β 1 (sI − A) = 2 ,
1 s +s +1

and its frequency response is

ω2 ω(β − ω2 )
Hβ (j ω) = + j .
(1 − ω2 )2 + ω2 (1 − ω2 )2 + ω2
Finally, since Hβ (s) is Hurwitz, Re{Hβ (j ω)} > 0 for all finite ω, and further-
more, limω→∞ ω2 Re{Hβ (j ω)} = 1 > 0 for any value of β ∈ R, we can conclude
that Hβ is SPR, and as a consequence, the control system is quadratically stable and
thus asymptotically stable.
98 3 Stability of Reset Control Systems

In this example, the after-reset surface MR and reset surface M are given by

MR = N (C) ∩ R(AR ) = {0},


  ,
1 (3.12)
M = N (C) \ MR = span {0}.
0

Note that since the after-reset surface is the origin 0, the state evolution of this
system is very simple, the equilibrium point is reached after the first reset if a reset
occurs. In addition, since√the matrix A does not have real dominant eigenvalues, its
eigenvalues are − 12 ± j 23 , and (A, C, x0 ) is reduced for any initial condition x0 ,
so the reset surface is always reached in a finite time.
This is a very simple example of a reset control system that is asymptotically
stable, and this stability is perfectly captured by the Hβ -condition. But it might
happen that the Hβ -condition fails in detecting stability, as seen in the following
example.

Example: Hβ is not SPR This example shows a case of a reset control system
with a stable base system in which the Hβ -condition does not apply, and thus the
stability of the reset system cannot be asserted. As it will be shown in Sect. 3.2.2.1
(see the second example), this case corresponds indeed to a stable reset system.
Consider a reset control system with a base compensator Rb (again correspond-
ing to a FORE compensator) and a plant P as given by the transfer functions:
3.2 s + 0.25
Rb (s) = , P (s) = . (3.13)
s −1 s 2 + 2s
A state-space realization for the closed-loop system is given by matrices A, AR ,
and C:
⎛ ⎞ ⎛ ⎞
−2 0 1.6 1 0 0  
A=⎝ 1 0 0 ⎠, AR = ⎝ 0 1 0 ⎠ , C = 1 0.25 0 ,
−2 −0.5 1 0 0 0
(3.14)
where A is Hurwitz (all its eigenvalues have strictly negative real part). In this ex-
ample, the compensator is full reset, and so nρ̄ = 0. In addition, np = 2 and nρ = 1.
Thus β is a real number, and
   
C0 = βCp Pρ = β 0.25β 1 , (3.15)

where Pρ = 1 without loss of generality. In addition,


   
B0 = 0np 0nρ̄ Inρ = 0 0 1 . (3.16)

The transfer function Hβ (s) after a simple computation is

5s 2 + (8β + 10)s + 2β
Hβ (s) = . (3.17)
5s 3 + 5s 2 + 6s + 4
3.2 Lyapunov Stability 99

Fig. 3.1 Minimum of the real part of Hβ (j ω) for ω ∈ [0, ∞) vs β

In Fig. 3.1, it is shown that although the base system is stable (A is Hurwitz) the Hβ -
condition is not satisfied since there does not exist any β ∈ R such as Re{Hβ (j ω)} >
0 for any ω > 0.

3.2.2 Reset Times-Dependent Stability Criteria

The results in the previous section are limited by the requirement that the Lyapunov
function has to be decreasing both in the continuous flow and at the reset changes.
Although in many practical cases this is the natural and expected behavior, there
are stable reset systems such that the Lyapunov function sometimes increases in the
continuous flow or at the reset instants.
On the other hand, the results in the previous section do not depend explicitly on
the reset instants, so that in many cases the important influence of the reset intervals
on the system stability cannot be adequately captured.
For the two mentioned reasons, this section presents an approach to reset systems
stability that explicitly takes into account the reset instants and reset intervals. The
resulting conditions are thus called reset times-dependent stability criteria.
To this end, stability of the reset control system is based on the stability of a time
varying discrete time system representing the dynamics of the reset system between
consecutive after-resets instants. Related work has been recently developed in [16],
but there reset is simplified to be performed at fixed instants, and thus earlier results
developed in [2] can be directly used.
100 3 Stability of Reset Control Systems

In this section, stability conditions will be developed by using an auxiliary time-


variant discrete time system, where the dynamics of the auxiliary system will explic-
itly depend on reset time intervals. This approach is further extended in Sect. 3.2.3
to the stabilization problem, imposing restrictions over reset intervals to be in some
designed set. A simple observation of the reset control system equation (2.4), and
its solution as given in Proposition 2.1, shows that if it is sampled at the after-reset
instants tk+ , k = 0, 1, 2, . . . , the sampled system is given by
+   
x tk+1 , x0 = AR eA(τk+1 (x0 )−τk (x0 )) x tk+ , x0 (3.18)

since x(t, x0 ) = eA(t−tk ) x(tk+ , x0 ), for t ∈ (tk , tk+1 ].


In addition, if the assumptions of Proposition 2.2 are satisfied, more specifically
the base system has no real dominant modes and (A, C, x0 ) is reduced for any ini-
tial condition x0 ∈ Rn \ {0}, there exist a well defined sequence of reset intervals
(Δ1 (x0 ), Δ2 (x0 ), . . . ), in the sense that they are upper-bounded by some constant
ΔM . Strictly speaking, it is only necessary to introduce some upper bound over
reset intervals if the base system is unstable, otherwise the reset system will be un-
stable. Therefore, since in general the problem of determining if an initial condition
x0 crosses the surface M is an open issue, in this work it will be assumed that if
the base system is unstable then it has the property that for any initial condition
x0 ∈ MR , its solution x(t, x0 ) = CeAt x0 is equal to zero at some instant tc ≤ ΔM ,
for some constant ΔM > 0. A reset control system with either a stable base system
or an unstable base system with that property will be referred to as regular. Note
that in particular reset control systems having an unstable base linear systems with
complex dominant modes, with (A, C, x0 ) reduced for any x0 ∈ Rn \{0}, are regular.
In the case of low order reset control systems in which the dimension of the after
reset surface MR is m = 0 or m = 1, the sampled system (3.18) results in an auxil-
iary discrete time system with particularly simple dynamics. The case in which the
dimension of the reset surface is m = 0, that is, MR = {0}, corresponds to a first or-
der plant and a full reset compensator; in this case, the equilibrium point 0 is reached
at the first reset instant if there is a crossing, thus the stability problem reduces to
the reachability of the reset surface and the following result easily follows.

Proposition 3.3 A regular reset control system with MR = {0} is asymptotically


stable. In addition, a reset control system with MR = {0} which is not regular is
not stable.

The case m = 1 has been analyzed in Sect. 2.3.1. In this case, for any initial
condition x0 , reset intervals are periodic with period Δ after the first reset inter-
val Δ1 (x0 ) = τ1 (x0 ), and thus the auxiliary discrete-time system associated to the
sampled system (3.18) is time invariant. It is given by

xd (k + 1) = AR eAΔ xd (k), xd (0) = AR eAτ1 (x0 ) x0 . (3.19)

As a result the stability of the reset control system is given by the stability of the
auxiliary time-invariant discrete time system, taking into account that the reset sur-
3.2 Lyapunov Stability 101

face is reached from any nonzero initial condition in the case in which the base
system is not stable. The following result easily follows.

Proposition 3.4 A regular reset control system with m = 1 is asymptotically stable


if the matrix AR eAΔ is Schur stable, that is, all its eigenvalues are strictly inside the
unit circle. In addition, a reset control system with m = 1 is not stable if it is not
regular, or is regular and AR eAΔ is not Schur stable.

In general, for the case m > 1, the reset intervals sequences may exhibit complex
patterns as has been discussed in Sects. 2.3.2 and 2.3.3. However, the stability prob-
lem of the reset control system can be reduced to solutions with initial conditions
in the after reset surface MR . In addition, if the base system is not stable, the regu-
larity condition is needed to assure that MR is uniformly reachable in a finite time;
otherwise, the reset control system is not stable. On the other hand, the auxiliary
time-varying discrete time system will be given by

xd (k + 1) = AR eAΔk+1 xd (k), xd (0) = x0 (3.20)

where (Δ1 , Δ2 , . . . ) is a sequence of reset intervals corresponding to some initial


condition. In the following, the stability of the reset control system will be analyzed
in terms of the stability of a family of auxiliary discrete-time systems in which all the
possible sequences of reset intervals are included. Since in general for the stability
of the auxiliary systems it will be necessary that reset intervals be lower-bounded,
the original reset control system will be modified to include a dwell-time constant
Δm . In this way, the reset control system to be considered in the following is

Δ̇(t) = 1, ẋ(t) = Ax(t) (x(t) ∈
/ M ) ∨ (Δ ≤ Δm ),
+ +
(3.21)
Δ(t ) = 0, x(t ) = AR x(t) (x(t) ∈ M ) ∧ (Δ > Δm ).

Proposition 3.5 A regular reset control system (3.21) is (asymptotically) stable if


the time-varying discrete time system (3.20) is uniformly (asymptotic) stable for any
sequence (Δ1 , Δ2 , . . . ), where Δk > Δm , k = 1, 2, . . . .

Proof If the discrete time system given by (3.20) is uniformly stable for any se-
quence (Δ1 , Δ2 , . . . ), with Δk > Δm , k = 1, 2, . . . then it is true that xd (k) ≤
γ xd (k0 ), k ≥ k0 , for any k0 and xd (k0 ) = x0 , and some γ > 0. In particular, for
any x0 ∈ Rn and k0 = 0, reset intervals will be given by Δk = Δk (x0 ) = tk − tk−1 =
τk (x0 ) − τk−1 (x0 ), k = 1, 2, . . . , making a sequence (Δ1 , Δ2 , . . . ) with Δk > Δm by
assumption. Then, the solution of the system (2.4) at the after-reset instants satisfies
x(tk+ , x0 ) = xd (k) ≤ γ x0 , k ≥ 0. Now, using standard arguments based on
the Gronwall inequality (see, for example, [32]) it is possible to bound the solution
between two consecutive reset instants. Since the solution x(t, x0 ) at some instant
t ∈ (tk , tk+1 ] is given by
t
 
x(t, x0 ) = x tk+ , x0 + Ax(s, x0 ) ds, (3.22)
tk
102 3 Stability of Reset Control Systems

by taking norms, we get



-  - t
x(t, x0 ) ≤ -x t + , x0 - + α
k x(s, x0 ) ds, (3.23)
tk

where the induced norm A is always bounded by some real number α, that
is, A ≤ α. Finally, using Gronwall inequality, x(t, x0 ) ≤ x(tk+ , x0 )eα(t−tk ) ,
showing that the solution x(t) is always bounded in every interval t ∈ (tk , tk+1 ], re-
sulting in x(t, x0 ) ≤ γ eα(t−tk ) x0 , where γ and α do not depend on the initial
condition x0 ∈ M . In addition, since the reset control system is regular, if the base
linear system is unstable, that is, A is not Hurwitz, the solution x(t, x0 ) satisfies

x(t, x0 ) ≤ γ eαΔk x0  ≤ γ eαΔM x0  (3.24)

for t ≥ 0, including reset instants. Thus, it is obtained that the system (2.4) is stable.
Note that in the case in which A is Hurwitz, reset intervals do not need to be upper-
bounded since the exponential is always bounded, and stability of (2.4) directly
follows.
Asymptotic stability follows by using similar arguments and the fact that
x(tk+ , x0 ) ≤ γ λk x0 , k ≥ 0 for any x0 ∈ Rn , with 0 ≤ λ < 1, by uniformly
asymptotic stability of the discrete time system (3.19). Then, as t → ∞, it is true
that k → ∞ since t ∈ (tk , tk+1 ] and tk+1 − tk = Δk+1 (x0 ) > Δm for any x0 ∈ Rn ,
and in addition since the reset control system is regular,

x(t, x0 ) ≤ γ λk eαΔk x0  ≤ γ λk eα·ΔM x0 , (3.25)

thus x(t, x0 ) → 0 for any x0 ∈ Rn as t → ∞. 

Proposition 3.6 A regular reset control system is asymptotically stable if there exist
a positive definite matrices sequence Pk , k = 1, . . . , such that ηI ≤ Pk ≤ ρI and

eA k AR Pk+1 AR eAΔk − Pk ≤ −εI (3.26)

for some finite positive constants η, ρ, and ε, and for any reset interval sequence
(Δ1 , Δ2 , . . . ), where Δk > Δm , k = 1, 2, . . . .

Proof It is a direct application of Proposition 3.5 and discrete time-varying system’s


stability results as given, for example, in [27, Theorem 23.3]. 

In comparison with the reset times-independent results in the previous section,


notice that there the stability conditions are:
 T T
x (A P + P A)x < 0 if x ∈
/ M \ {0},
(3.27)
xT (ATR P AR − P )x ≤ 0 if x ∈ M ,

which implies that the quadratic Lyapunov function V (x) = xT P x is decreasing


over the time intervals (tk , tk+1 ) and non-increasing at the reset instants tk , k =
3.2 Lyapunov Stability 103

1, 2, . . . . These stability conditions have also been previously published in [23] and
the seminal work [2], in the context of impulsive systems with impulses at fixed
instants.
It is clear that from the first condition of (3.27) the base linear system must always
be asymptotically stable. Note that even in the case in which the base linear system
is asymptotically stable, the second condition of (3.27) may be not compatible in
general with the solutions of the first condition.
In addition, in the previous section it has been shown that (3.27) is equivalent
to the Hβ -condition. Thus, the main contribution of Proposition 3.5 (and Proposi-
tion 3.6) is to circumvent the Hβ -condition by allowing that the continuous time
evolution between two consecutive reset instants, in (tk , tk+1 ), k = 1, 2, . . . , is not
necessarily associated with a Lyapunov function (quadratic or not) in those intervals,
thus results are less conservative: (i) the proposition may be applied to unstable base
linear systems, (ii) it may be applied to stable base linear systems in which condition
(3.27) cannot be applied, as we will see in the two following examples.

3.2.2.1 Examples

Unstable Base System Consider a reset control system as given by (2.4), where
A, AR , and C are given by
⎛ ⎞ ⎛ ⎞
−2 0 1 1 0 0  
A=⎝ 1 0 0⎠, AR = ⎝ 0 1 0⎠, C= 001 . (3.28)
0 −1 1 0 0 0

This corresponds to the feedback system of Fig. 2.1 with P (s) = 1


s(s+2) and a FORE
with base compensator Rb (s) = 1
s−1 .
In this case, the reset control system is regular
since although A is not Hurwitz it has a pair of dominant complex eigenvalues, its
spectrum is σ (A) = {0.57 + 0.37i, 0.57 − 0.37i, −2.15}; and, in addition, (C, A, x0 )
is reduced for x0 = [x01 0 0]T ∈ MR and x01 = 0, since in this case equation (2.7)
is given (after some computation) by
 
y(t) = 0.42e0.57t cos(0.37t) + 0.37 e0.57t sin(0.37t) − 0.42 e−2.15t x01 .

Thus, after the first reset, reset instants are periodic with period Δ = 2.3 (the first
zero of y(t) that is bigger than Δm = 0.1) for any initial condition. Note that AR eAΔ
is Schur stable (its eigenvalues are strictly inside the unit circle), but that the reset
control system is not regular since the after-reset surface is not reached for initial
conditions x0 ∈ span{(−0.898, 0.418, 0.133)T }. Thus, Proposition 3.4 can be used
to state that the reset control system is unstable. If the initial condition set were
restricted to MR [7] then the reset control system would be stable since AR eAΔ is
Schur stable.
104 3 Stability of Reset Control Systems


Fig. 3.2 Max. eigenvalue of eA AR P AR eAΔ − P + εI vs Δ

Stable Base System This example is given to show how the Hβ -condition can be
overtaken for reset control systems with stable base system. Consider a reset control
system as given by (3.21) with a dwell-time constant Δm = 0.1 and matrices
⎛ ⎞ ⎛ ⎞
−2 0 1.6 1 0 0  
A = ⎝ 0.5 0 0 ⎠, AR = ⎝ 0 1 0 ⎠ , C= 001 (3.29)
−2 −1 1 0 0 0

corresponding to Rb (s) = s−1 3.2


and P (s) = s+0.25
s 2 +2s
. It can be easily checked that, al-
though A is Hurwitz (and thus the reset control system is regular), the Hβ -condition
is not satisfied, that is, there does not exist any β ∈ R (β is scalar in this case) such
that Re{Hβ (j ω)} = Re[ β 0 1 ](j ωI − A)−1 [ 0 0 1 ]T > 0 (see the second example
in Sect. 3.2.1).
Now, it has been found that (see the search procedure given in next section) (3.26)
is feasible for constant matrices Pk = P , where P is given by
⎛ ⎞
22.07 7.91 −7.01
P = ⎝ 7.91 41.96 1.22 ⎠ (3.30)
−7.01 1.22 46.86

and for any Δk ≥ 0.03, k = 1, 2, . . . . Figure 3.2 shows the maximum eigenvalue of

eA Δ AR P AR eAΔ − P + εI vs Δ, for ε = 0.001. It is always negative for Δ ≥ 0.03.
As a result, Proposition 3.6 guarantees that the reset control system is asymp-
totically stable. As it will be developed in next section, for the case of stable base
3.2 Lyapunov Stability 105

Fig. 3.3 Base system (dashed line), and reset system (solid line)

systems this method can be used to compute the values of the dwell-time constant
Δm that stabilizes the reset control system. In this example, stability is assured for
any Δm ≥ 0.03. Figure 3.3 shows a time simulation of the closed-loop output for
the base and the reset control system.

3.2.3 Stabilization of Reset Control Systems

In this section, Proposition 3.5 and Proposition 3.6 will be used to adjust the dwell-
time constant Δm in such a way that, by avoiding reset actions at some crossings
with the reset surface, stability of the reset control system is achieved. The case in
which the base system is stable is first approached, note that for the unstable case a
separate treatment is necessary, since it is not enough to avoid reset actions, it is also
necessary to force resets at some instants even if the reset surface is not crossed.

3.2.3.1 Stable Base System

Proposition 3.7 For the reset control system (3.21) with A Hurwitz, there always
exists a value Δm > 0 of the dwell-time constant such as for any Δ > Δm the in-
equality

eA AR P AR eAΔ − P ≤ −εI (3.31)
106 3 Stability of Reset Control Systems

is satisfied for some P > 0 and ε > 0. Therefore, the reset control system (3.21) is
asymptotically stable for that Δm .

Proof Since A is Hurwitz and then the reset control system is regular, stability di-
rectly follows from Proposition 3.6, by taking Pk = P , for k = 1, 2, . . . . Note that
since A is Hurwitz, the matrix eAΔ can be made arbitrarily small in norm, by choos-
ing Δ large enough. Thus, for any ε > 0 there always exists a Δm such for any
Δ > Δm it is true that eAΔ  ≤ ε. Some valid explicit relations Δ = f (ε) could be
derived using some canonical form of A and eAΔ (diagonal, Jordan form, etc.) and
bounding |eλi Δ |, where λi are the (stable) eigenvalues of A. Then, for the sequence
P1 = P2 = · · · = P for some P > 0, a Δ large enough can be chosen (Δ > Δm ) so
that the left side of the inequality (3.31) approaches 0−P = −P , which is obviously
smaller than −εI for small ε > 0. 

A Search Procedure for Δm Proposition 3.7 gives a condition that is easy to


check in practice by using available algorithms for solving LMIs. However, some
care has to be taken because as it is well known the norm eAΔ  is not necessarily
monotonic with respect to Δ. For A Hurwitz, when Δ → ∞ it is true that eAΔ → 0,
but this does not imply monotonicity of eAΔ  as a function of Δ.
Before the search procedure, a lower bound of the solution (ΔLB ) can be found
by the map of the feasibility of the LMI (3.31) vs Δ. This lower bound is defined by
ΔLB ∈ R such that for all Δ ≥ ΔLB the LMI (3.31) is feasible for some P (Δ).
A search procedure for Δm is proposed as follows:
1. Initial value: Let us start with a large initial value ΔL so that the LMI (3.31) is
feasible for the single value Δ = ΔL with respect to some P = PL > 0, and also
for all Δ > ΔL the LMI (3.31) is feasible.
2. Feasible values of P : A set of feasible values of P is obtained simply by sam-
pling the interval (0, ΔL ] and solving the LMI (3.31) for P with a given sam-
pled value of Δ. More precisely, with some step size Δ1 , a set of reset intervals
{Δ1,1 = Δ1 , Δ2,1 = 2Δ1 , Δ3,1 = 3Δ1 , . . . } ⊂ (0, ΔL ] is chosen. Now, for every
Δk,1 a matrix Pk > 0 is computed by solving the LMI (3.31). As a result, a set
of feasible pairs (Δk,1 , Pk ), k = 1, 2, . . . , is obtained. Let P = {P1 , P2 , . . .}.
3. Feasible values of Δ: The next step is to compute the minimum value Δm,k
such that for any Δ > Δm,k the LMI (3.31) is feasible for some Pk ∈ P. In
order to obtain that result, every matrix Pk ∈ P generates a matrix Qk (Δ) =

eA Δ AR Pk AR eAΔ − Pk + εI . Then, the interval (0, ΔL ] is sampled with another
step size Δ2 such that the maximum eigenvalue of Qk (ΔL − iΔ2 ), where i =
0, 1, . . . , is computed and i is increased until the maximum is non-negative for
some value of i = ik + 1; therefore, for all Pk ∈ P, a Δm,k = ΔL − (ik )Δ2 is
found.
Again the LMI has to be checked with some accuracy Δ2 (not necessarily
equal to Δ1 ), for Δ ∈ {Δ1,2 = Δ2 , Δ2,2 = 2Δ2 , Δ3, = 3Δ2 , . . . }. As a result,
a set of pairs (Δk,m , Pk ), k = 1, 2, . . . , is now obtained.
4. Finally, the desired Δm is the minimum value of Δm,k for k = 1, 2, . . . , obtained
for some km , and Pkm is the associated matrix.
3.2 Lyapunov Stability 107

The number of LMIs to be checked is of order Δ2L /(Δ1 Δ2 ). This procedure


guaranties that a minimum value Δm and a matrix P = Pkm exist such that the LMI
(3.31) is feasible for Δ > Δm . On the other hand, note that Proposition 3.7 gives a
simple and practical method for designing stabilizing reset control systems. Once
the bound Δm has been computed, the reset control system should only reset when
reset intervals are larger than Δm .

Stabilization by Removing Reset Actions at Some Crossings As it is well


known, there exist cases in which reset actions can destabilize a stable base closed-
loop system. This is the case of the example to be analyzed in the following,
which is borrowed from [9]. It will be shown how stability of the reset system
can be assured by simply avoiding those reset intervals that are not large enough.
Consider a standard reset control system where the base compensator Rb (corre-
sponding to a FORE compensator) and the plant P are given by Rb (s) = s+1 1
and
(s+1)(s+10) 2
P (s) = s 4 +4s 3 +6s 2 +5s−98 , respectively, corresponding to a state-space realization
with matrices A and AR :
⎛ ⎞
−4 −0.750 −0.312 1.531 4
⎜ 8 0 0 0 0 ⎟
⎜ ⎟
A=⎜ 0 ⎜ 2 0 0 0 ⎟⎟,
⎝ 0 0 4 0 0 ⎠
⎛−0.25 −0.656 ⎞−1.875 −0.391 −1 (3.32)
1 0 0 0 0
⎜0 1 0 0 0⎟
⎜ ⎟
AR = ⎜ ⎜0 0 1 0 0⎟.

⎝0 0 0 1 0⎠
0 0 0 0 0

Note that the reset control system is regular since A is Hurwitz. It has been found
that the LMI (3.31) is feasible for Δ ≥ 27.02 and some P > 0. As a result of the
search procedure a matrix Pm has been found:
⎛ ⎞
0.4680 0.0594 0.0395 0.0296 0.0556
⎜ 0.0594 0.3630 0.0034 0.119 −0.0161 ⎟
⎜ ⎟
Pm = 10 ⎜ 0.0395 0.0034 0.4300 0.0597 0.1110 ⎟
4⎜
⎟ (3.33)
⎝ 0.0296 0.1190 0.0597 0.170 0.2810 ⎠
0.0556 −0.0161 0.1110 0.2810 1.0600

making the LMI (3.31) feasible for this Pm and Δ ≥ Δm = 27.02. Figure 3.4 shows

the maximum eigenvalue of eA Δ AR P AR eAΔ − P , which is always negative for
values of Δ ≥ 27.02.
Since the reset control system produces crossings with the reset surface with
intervals that are smaller than 27.02, it is not guaranteed that the reset control system
is stable in this case. Figure 3.5 shows a simulation for this system, comparing the
closed loop outputs both for the base and the reset control system, for Δm = 0.1.
Using again Proposition 3.7, it can be concluded that the closed-loop reset system is
108 3 Stability of Reset Control Systems


Fig. 3.4 Max. eigenvalue of eA AR P AR eAΔ − P + εI vs Δ

Fig. 3.5 Closed loop output for Δm = 0.1

asymptotically stable if the reset intervals Δk , k = 0, 1, . . . , are always greater than


Δm = 27.1.
Figure 3.6 shows a simulation of the reset control system avoiding resets when
the reset interval is smaller than Δm = 27.1. Note that the system, although stable,
3.2 Lyapunov Stability 109

Fig. 3.6 Reset control system output for Δm = 27.1

Fig. 3.7 Reset control system output for Δm = 30

is very oscillatory since the reset intervals are close to the stability limit. Figure 3.7
shows a simulation for Δm = 30. In this case, the system behavior is much less
oscillatory due to the fact that reset intervals are far away from the stability limit.
110 3 Stability of Reset Control Systems

3.2.3.2 Unstable Base System

In the previous subsection, stability of reset control systems has been considered for
stable base systems, that is, for the case in which the matrix A is Hurwitz. In this
subsection, Proposition 3.5 is used to elaborate a sufficient condition for stability of
reset control systems in the case in which the closed loop matrix A is not Hurwitz,
and thus the base linear system is unstable. It should be noted that when A is not
Hurwitz the Hβ -condition, developed in [9], is not applicable.
In addition, in the case in which A is not Hurwitz and thus the base system is not
stable, solutions of (3.31) are typically given for some bounded set (usually a union
of intervals) I ⊂ R for a given definite positive matrix P . First, the case in which the
reset control system is regular is investigated. In this case, the reset control system
(3.21) is redefined to allow resets only at the instants in which Δ ∈ I :

Δ̇(t) = 1, ẋ(t) = Ax(t) (x(t) ∈
/ M ) ∨ (Δ ∈ / I ),
(3.34)
Δ(t + ) = 0, x(t + ) = AR x(t) (x(t) ∈ M ) ∧ (Δ ∈ I ),

and thus the stability of reset control system easily follows.

Proposition 3.8 For a regular reset control system (3.34), assume that I ⊂ R is a
closed bounded set such that for any scalar Δ ∈ I the inequality (3.31) is satisfied
for some ε > 0, and a positive-definite matrix P . Then, the reset control system is
asymptotically stable.

In the case in which the base system is not regular, that is, when there exist initial
conditions that do not reach the reset surface and thus the reset control system is
directly unstable since the base system is not stable, some reset may be forced in
order to stabilize the system. In this case, let ΔI be the upper bound of I , that is,
ΔI = max{I }, then the reset control system may be modified, resulting in

Δ̇(t) = 1, ẋ(t) = Ax(t) (x(t) ∈/ M ) ∨ (Δ ∈/ I ) ∧ (Δ = ΔI ),
+ +
(3.35)
Δ(t ) = 0, x(t ) = AR x(t) (x(t) ∈ M ) ∧ (Δ ∈ I ) ∨ (Δ = ΔI ).

Proposition 3.9 For a reset control system (3.35), assume that I ⊂ R is a closed
bounded set such that for any scalar Δ ∈ I the inequality (3.31) is satisfied for some
ε > 0, and a positive-definite matrix P . Then the reset control system is asymptoti-
cally stable.

A search procedure similar to that developed in the previous section can be used
to obtain an interval I , although there are some particularities (see [4]). In the fol-
lowing, two examples are developed. The first example consists of a lower order
base system (with m = 1), in particular the plant is a second order model with an in-
tegrator and a real pole. In the second example, a higher order base system (m = 4)
is considered.
3.2 Lyapunov Stability 111

A Low Order Base System Consider a reset control system with the base com-
pensator Rb and the plant P given by
8 1
Rb (s) = , P (s) = . (3.36)
s +1 (s + 2)s
The (unstable) closed-loop base system is given by the transfer function
8
T (s) = . (3.37)
s3 + 3s 2+ 2s + 8
A state-space realization of the base system transfer is obtained as

(Rb ) Ar = −1, Br = 4, Cr = 2, Dr = 0, (3.38)


   
−2 0 1  
(P ) Ap = , Bp = , Cp = 0 1 , Dp = 0, (3.39)
1 0 0
from where the closed-loop state matrices are given by
⎛ ⎞ ⎛ ⎞
−2 0 2 1 0 0  
A=⎝ 1 0 0 ⎠, AR = ⎝ 0 1 0 ⎠ , C= 010 . (3.40)
0 −4 −1 0 0 0

For the positive-definite matrix P given by


⎛ ⎞
29.18 10.56 1.49
P = ⎝ 10.56 176.23 −19.39 ⎠ , (3.41)
1.49 −19.39 93.07

it has been found that the resulting set I is given by

I = [0.92, 1.72] ∪ [3.16, 3.51] (3.42)

as shown in Fig. 3.8.


In this example, there are initial conditions that do not produce crossings with
the surface reset, and thus the reset control system is not regular. As a result, Propo-
sition 3.9 can be used to show that the reset control system (3.35) is stable with
I given by (3.42) and ΔI = 3.51. Figure 3.9 shows a time simulation comparing
closed-loop outputs for the reset control system and its base system.

A High Order Base System Consider the reset control system with base compen-
sator Rb and plant P given by
350 s +3
Rb (s) = , P (s) = (3.43)
s +1 s5 + 13s 4 + 87s 3 + 305s 2 + 350s
that results in a unstable base control system, for example, the (unstable) closed-
loop transfer function from the reference input to the output is given by
112 3 Stability of Reset Control Systems


Fig. 3.8 Maximum eigenvalue of eA AR P AR eAΔ − P + εI vs Δ

Fig. 3.9 Closed-loop output for the linear base control system (dashed line), and the reset control
system (solid line)

350(s + 3)
T (s) = . (3.44)
s6 + 14s 5 + 100s 4 + 392s 3 + 655s 2 + 700s + 1050
3.3 Reset Control Systems with Inputs/Passivity Analysis 113


Fig. 3.10 Maximum eigenvalue of eA AR P AR eAΔ − P + εI vs Δ

In this case, it has been found that for the positive-definite matrix P
⎛ ⎞
16.22 1.16 2.14 2.42 0.79 3.08
⎜ 1.16 14.68 1.46 5.44 1.93 1.74 ⎟
⎜ ⎟
⎜ 2.14 1.46 9.77 2.92 10.78 2.26 ⎟
P =⎜ ⎜ ⎟ (3.45)
2.42 5.44 2.92 9.22 10.45 −0.68 ⎟
⎜ ⎟
⎝ 0.79 1.93 10.78 10.45 28.11 −9.55 ⎠
3.08 1.74 2.26 −0.68 −9.55 27.35

the resulting set I is

I = [1.12, 1.92] ∪ [3.37, 3.96] ∪ [5.72, 6.01] ∪ [8.05, 8.08] (3.46)

as shown in Fig. 3.10.


On the other hand, this reset control system is not regular since the base system
is unstable and in addition there exist initial conditions that do not reach the surface
reset. Using again Proposition 3.9, the reset control system (3.35) is stable with I
given by (3.46) and ΔI = 8.08. Figure 3.11 shows a time simulation comparing
closed-loop outputs for the reset control system and its base system.

3.3 Reset Control Systems with Inputs/Passivity Analysis


Regarding L2 -stability of reset systems with inputs, a number of works have been
developed that give results for particular cases of reset compensators and/or inputs.
114 3 Stability of Reset Control Systems

Fig. 3.11 Closed-loop output for the linear base control system (dashed line), and the reset control
system (solid line)

The work [24] approaches the problem for compensators in which its output has
the same sign as its input, and the zero reference case is considered. In addition,
in [1], L2 -stability conditions for the case of nonzero references are given. The
conservatism given by the Hβ -condition is improved for that kind of system.
On the other hand, dissipative systems theory was developed in [33], where the
passive system concept which was developed in electric circuit and mechanical sys-
tems was extended to abstract systems. That theory is widely used in interconnected
systems due to the fact that the feedback interconnection of two passive systems is
a passive system. In [29], passivity techniques are shown to be a powerful tool for
nonlinear control. Dissipative systems theory has been developed for a hybrid sys-
tem in [17], where dissipative hybrid system’s definitions are developed for classic
concepts such as supply rate. A few works have dealt with passive systems theory
and hybrid systems [35], [11], [25], [8]. In spite of the fact that a single input–output
approach to passive systems theory cannot be sufficient for more general hybrid sys-
tems, this section will show how several passive properties can be obtained for reset
compensators.
The goals of this section are: (a) to obtain stability conditions that are applicable
to feedback interconnections of reset compensators and nonlinear plants and (b) to
find passive reset compensators that can be used in passive control techniques. The
passivity approach will be used for these purposes. Easily checkable passivity con-
ditions on the base linear compensator will be presented. In general, the feedback
system given by Fig. 3.12 will be considered, where w and d are the reference and
disturbance inputs, respectively. R is a single-input single-output (SISO) reset com-
pensator to be defined later on, and P is an SISO plant.
3.3 Reset Control Systems with Inputs/Passivity Analysis 115

Fig. 3.12 Reset controller R


applied to an LTI plant

∞ space L2 consists of all measurable functions


By definition, the linear
f (·) : R+ → R such that 0 |f (t)|2 dt < ∞, with the L2 -norm  ·  : L2 → R+
∞ 1
given by f  = ( 0 |f (t)|2 dt) 2 . Let ΠT be a projector operator such that
ΠT (f )(t) = f (t), for 0 ≤ t ≤ T , and ΠT (f )(t) = 0, for t > T . Then the extended
space L2,e is given by all functions f (·) : R+ → R such that ΠT (f ) belongs to L2 .
The feedback interconnection is given simply by

e(t) = w(t) − y(t), u(t) = v(t) + d(t). (3.47)

The feedback system of Fig. 3.12 is L2 -stable (with respect to inputs w and d)
if for every input signals w ∈ L2 and d ∈ L2 , the outputs u ∈ L2 and y ∈ L2 . In
addition, it is finite-gain stable if there exists a positive constant γ > 0 such that
y2 + u2 ≤ γ (w2 + d2 ).
Here the plant P is represented by the state space model

ẋp = f (xp , u) if u ∈ R,
(P ) (3.48)
y = g(xp , u) if y ∈ R,

where np is the dimension of the state xp , f : Rnp × R → Rnp is locally Lips-


chitz, g : Rnp × R → R is continuous, f (0, 0) = 0, and g(0, 0) = 0. In addition,
following the notation given in [17], the reset compensator dynamic is described by
three elements: (a) a continuous-time dynamical equation, (b) a difference equation,
and (c) a resetting law. Whenever the resetting law is not held, the system flows
continuously; otherwise, when the resetting law is satisfied, the system jumps in a
discrete way. An LTI base compensator will be considered in this work, then the
reset compensator R is given by the impulsive differential equation (IDE):


⎪ ẋr = Ar xr + Br e if e = 0,

(R) xr+ = Aρ xr if e = 0, (3.49)



u = Cr xr + Dr e,

where nr is the dimension of the state xr , Aρ is a diagonal matrix with zeros in


the states to be reset and ones in the rest of compensator states, nρ is defined as
116 3 Stability of Reset Control Systems

the dimension of the reset subspace, and nρ is defined as the dimension of the non-
reset subspace (nρ + nρ = nr ). When Aρ = 0, R will be referred to as a full reset
compensator; otherwise, it will be referred to as a partial reset compensator.
The first equation in (3.49) describes the continuous compensator dynamic at the
non-resetting instants, and the second equation gives the reset operation as a jump
of the compensator state at the reset instants. Note that reset instants are produced
when the compensator input is zero. The base compensator is simply obtained by
elimination of the reset actions in (3.49) and thus has the transfer function Rbl (s) =
Cr (sI − Ar )−1 Br + Dr , where it is assumed that (Ar , Br , Cr , Dr ) is a minimal
realization.
It is assumed that the feedback system is well-posed and thus Zeno solutions are
avoided. Alternatively, time regularization may be imposed (see (2.7)), where the
resetting law is deactivated during a time Δm after every reset action. As a result, it
is assumed that there exists a finite number of resets over every finite time interval,
thus the reset actions will be given by the countable set {t1 , t2 , . . . , tk , . . . } where
tk+1 − tk > Δm for all k = 1, 2, . . . and limt→∞ tk = ∞.
A system H : L2,e → L2,e , with input u and output y = H u is said to be passive
if there exists a constant β ≤ 0 such that
T
u (t)y(t) dt ≥ β, ∀T ≥ 0, ∀u ∈ L2 . (3.50)
0

If there are constants δ ≥ 0 and ε ≥ 0 such that


T T T
 
u (t)y(t) dt ≥ β + δ u (t)u(t) dt + ε y  (t)y(t) dt, (3.51)
0 0 0

for all functions u, and all T ≥ 0, then the system is input strictly passive (ISP)
if δ > 0, output strictly passive (OSP) if ε > 0, and very strictly passive (VSP) if
δ > 0 and ε > 0. In particular, for a linear time-invariant system with a transfer
function given by H (s) = C(sI − A)−1 + D, with A Hurwitz and the pair (A, B)
controllable, it is true that [21]
1. The system is passive if and only if Re[H (j ω)] ≥ 0 for all ω.
2. The system is ISP if and only if there is a δ > 0 such that Re[H (j ω)] ≥ δ > 0
for all ω.
3. The system is OSP if and only if there is an ε such that Re[H (j ω)] ≥ ε|H (j ω)|2
for all ω.
Note that these three properties are verifiable in a Nyquist diagram: if H (j ω) is in
the closed right half-plane then the system is passive; if H (j ω) is in Re[H (j ω)] ≥
δ > 0, the system is ISP; and if H (j ω) is inside the circle with center at s = 2ε
1
and
radius r = 2ε , then the system is OSP.
1

When the system is in state space representation, the passivity of the input–output
maps as before is replaced by the following notions from dissipative systems theory.
3.3 Reset Control Systems with Inputs/Passivity Analysis 117

Consider the system H : L2,e → L2,e given by



ẋ = f (x, u) if u ∈ R,
(H ) (3.52)
y = g(x, u) if y ∈ R,

where f and g have the same properties as in (3.48). H is said to be dissipa-


tive respect to a supply rate w(t) = w(u(t), y(t)) if there exists a storage function
V (x) ≥ 0 such that the following dissipation inequality holds
T
V (x) ≤ V (x(0)) + w(u(t), y(t)) dt; ∀u, ∀x(0), ∀T ≥ 0. (3.53)
0

Following [29], the relationship between dissipative and passive systems is given
by the choice of a particular supply rate:
1. The system H is passive if it is dissipative with respect to supply rate wp = u y
and V (0) = 0.
2. The system H is input strictly passive (ISP) if it is dissipative with respect to
supply rate wi = u y − εu u, for some ε > 0.
3. The system H is output strictly passive (OSP) if it is dissipative with respect to
supply rate wo = u y − δy  y, for some δ > 0.
4. The system H is very strictly passive (VSP) if it is dissipative with respect to
supply rate wv = u y − δu u − εy  y, for some ε > 0 and δ > 0.
Comparing (3.50) and (3.53), since V (x) ≥ 0, the relationship between β and the
initial condition is clear, namely, β = −V (x(0)). An important result is the Passivity
Theorem (see [29]): the feedback interconnection of Fig. 3.12 is finite-gain stable if

εR + δP > 0,
εP + δR > 0,

where εR , δR , εP , and δP are such that they satisfy (3.51) for the systems R and P ,
respectively.

3.3.1 Full Reset Compensators

The main goal of this section will be to analyze the passivity properties of reset
compensators as described by (3.49). Firstly, conditions on reset compensators are
derived in order to show when they are (strictly) passive systems. Full reset com-
pensators and partial reset compensators are studied in different sections. Finally,
the passivity theorem will be used to show stability properties of the feedback sys-
tem given in Fig. 3.12. Notice that plants can be nonlinear in general, and will be
described by (3.48).
In spite of the fact that passivity theory (as an input–output theory) is not ap-
propriate in general for hybrid systems, it can be successfully used for full reset
118 3 Stability of Reset Control Systems

compensators given by (3.49). This is due to the fact that the system loses all of its
memory at every reset time, i.e., the base compensator is restarted with zero initial
condition. Therefore, a space states description is not necessary because the output
compensator just depends on the input since the last reset time. This is the key point
in the next proposition.

Proposition 3.10 A full reset compensator R (given by (3.49) with Aρ = 0) is pas-


sive, ISP, OSP, or VSP if the base compensator is passive, ISP, OSP, or VSP, respec-
tively.

Proof The result is proved for the VSP case, the rest of the cases follow similar
arguments. If the base compensator is VSP, then there exist δ > 0 and ε > 0 such
that (3.51) is satisfied for all T ≥ 0 and for all u ∈ L2 . Since, by assumption, in
every finite time interval there are a finite number of resets instants {0 < t1 , t2 , . . . }
with 0 < t1 < t2 < · · · and limk→∞ tk = ∞, there exists a finite integer k such that
for every T ∈ (tk , tk+1 ] the integration interval [0, T ] can be divided into a finite
number of subintervals
T t1  ti+1
i=k−1
u (t)y(t) dt = u (t)y(t) dt + u (t)y(t) dt
0 0 i=1 ti
T
+ u (t)y(t) dt. (3.54)
tk

In addition, since the base system is time invariant and VSP by assumption, it is true
that
t1 t1 t1
 
u (t)y(t) dt ≥ β + δ u (t)u(t) dt + ε y  (t)y(t) dt, (3.55)
0 0 0
ti+1 ti+1 ti+1
u (t)y(t) dt ≥ δ u (t)u(t) dt + ε y  (t)y(t) dt, (3.56)
ti ti ti
and
T T T
u (t)y(t) dt ≥ δ u (t)u(t) dt + ε y  (t)y(t) dt (3.57)
tk tk tk

for any u ∈ L2 and for i = 1, 2, . . . , k − 1. In (3.56) and (3.57), β = 0 has been


taken due to the system having null initial conditions at ti+ for all i = 1, 2, . . . , k. In
addition, the integral limits ti+ can be taken as ti by integral properties. Then, after
direct substitution of (3.55), (3.56), and (3.57) into (3.54), we at once conclude that
T T T
 
u (t)y(t) dt ≥ β + δ u (t)u(t) dt + ε y  (t)y(t) dt. (3.58)
0 0 0

Thus, the reset compensator is a VSP system with the same constants β, δ, and ε as
its base compensator. 
3.3 Reset Control Systems with Inputs/Passivity Analysis 119

Remark 3.1 Note that the formulation of passivity, ISP, OSP, and VSP given in [29]
(Sects. 2.2 and 3.1) is completely general, and thus does not involve any assumption
regarding the continuity of system trajectories of the reset system R as given by
(3.49). In particular, the state trajectories may be discontinuous at the reset instants,
and passivity definitions and the passivity theorem can be directly applied to reset
systems in order to obtain input–output stability properties.

Remark 3.2 Note that in the proof of the previous proposition, linearity of the base
compensator has been used simply to state that it has null initial conditions at ti+ for
all i = 1, 2, . . . . Thus, it holds for any full-reset compensator, not necessarily linear,
as long as β = 0 or, using the dissipativity framework to be elaborated in the next
subsection, the storage function of the compensator satisfies V (0) = 0.

Some examples of reset compensators are analyzed in the following.

3.3.1.1 The First Order Reset Element (FORE)

A typical reset compensator is the first order reset element (FORE) introduced in
[18, 20]. The transfer function of its base compensator is

k
FORE(s) = (3.59)
s+b
where it will be assumed that k, b > 0. The real part of its frequency response is
given by
kb
Re[FORE(j ω)] = , (3.60)
b2 + ω2
and thus, applying Proposition 3.10, it is directly obtained that
• FORE is passive since Re[FORE(j ω)] > 0 for all ω,
• FORE is not ISP because limω→∞ Re[FORE(j ω)] = 0,
• FORE is OSP, since for ε = bk

b
Re[FORE(j ω)] = |FORE(j ω)|2 .
k

3.3.1.2 A Reset Lag Compensator (Reset-PI)

In order to obtain VSP reset compensators, it is necessary that their base compen-
sators have the same number of poles as zeros. A possible choice is a base lag
compensator
1+Ts
PI(s) = k (3.61)
1+γTs
120 3 Stability of Reset Control Systems

where γ ∈ (1, ∞). Its frequency response real part is given by

1 + γ (T ω)2
Re[PI(j ω)] = k , (3.62)
1 + (γ T ω)2

then it is easy to check that Re[PI(j ω)] > min(1, γk ), and finally,

1 + (T ω)2
| PI(j ω)|2 = k 2 . (3.63)
1 + (γ T ω)2

Applying Proposition 3.10, it is now obtained that


• Reset-PI is passive due to Re[PI(j ω)] > 0 for all w,
• Reset-PI is ISP with δ = min(1, γk ),
• Reset-PI is OSP with ε = γk .

3.3.1.3 A Reset-PID Compensator

As example of higher order reset element, consider as base a PID compensator with
the transfer function
1 + Ti s 1 + Td s
PID(s) = kp γ (3.64)
1 + γ Ti s 1 + αTd s

where 0 ≤ Td < Ti , 1 ≤ β < ∞, and 0 < α ≤ 1. This base compensators is VSP,


then the full reset compensator based on it, referred to as reset-PID, is VSP. Several
other cases are possible for particular combination of parameters, a particular study
of the Nyquist plot is necessary in each case.

3.3.2 Partial Reset Compensators

In the partial reset compensator case, where only some of the compensator states are
reset, dissipativity theory needs to be used (in contrast with the full reset case). Note
that after each reset action, partial reset results in a possibly nonzero initial condition
for the compensator state in comparison (as opposed to full reset, which produces a
zero initial condition). Thus, for every initial condition after reset, a different input–
output system has to be taken into consideration. As a result, the compensator state
has to be included in the formulation and the result of Proposition 3.10 is not appli-
cable. Dissipative systems theory leads to a solution of this problem due to the fact
that the dissipativity property holds for every initial condition.

Proposition 3.11 A partial reset compensator R (given by (3.3) with the reset ma-
trix Aρ = diag(Inρ , 0nρ )) is passive, ISP, OSP, or VSP if its base compensator is
3.3 Reset Control Systems with Inputs/Passivity Analysis 121

dissipative with respect to supply rate wp , wi , wo , or wv , respectively, and a stor-


age function V (x) that satisfies

V (Aρ x) ≤ V (x) (3.65)

for every x ∈ Rnr .

Proof Again the case VSP is considered, the remaining cases follow similar argu-
ments. Since, by assumption, the base compensator is dissipative with respect the
storage function V and the supply rate wv = u y − δu u − εy  y for some δ > 0
and ε > 0, the following inequality is satisfied
T
V (x(T )) ≤ V (x(0)) + wv (u(t), y(t)) dt, ∀u ∈ L2,e , ∀x(0) ∈ Rnr , ∀T ≥ 0.
0
(3.66)
For a given input u(t), by assumption there is a finite number of reset times ti , i =
1, 2, . . . , k, in the interval [0, T ], with T ∈ (tk , tk+1 ]. Thus, it is true that
T
  
V (x(T )) ≤ V x tk+ + wv (t) dt (3.67)
tk

since in the interval [tk+ , T ] the reset compensator behaves like its base compensator
with x(tk+ ) the initial condition. Applying a similar argument, it is true that
ti
  + 
V (x(ti )) ≤ V x ti−1 + wv (t) dt (3.68)
ti−1

for i = 2, . . . , k, and also


ti
V (x(t1 )) ≤ V (x(0)) + wv (t) dt. (3.69)
0

Then, applying condition (3.65) at the reset instants, it is obtained that V (x(ti+ )) =
V (Aρ x(ti )), i = 1, . . . , k. Using this fact and (3.67)–(3.69), the result is

    +  tk T
V x(T ) ≤ V x tk−1 + wv (t) ds + wv (t) dt
tk−1 tk

  +  T
= V x tk−1 + wv (t) dt (3.70)
tk−1

  +  T
≤ V x tk−2 + wv (t) dt ≤ · · · ,
tk−2

and then it is concluded that the reset compensator satisfies


T
0 ≤ V (x(T )) ≤ V (x(0)) + wv (t) dt.
0
122 3 Stability of Reset Control Systems

As a result, it is true that


T T T
u (t)y(t) dt ≥ β + δ u (t)u(t) dt + ε y  (t)y(t) dt, (3.71)
0 0 0

where β = −V (x(0)), and thus the reset compensator is VSP. 

Remark 3.3 A necessary and sufficient condition for the satisfaction of (3.65) can
be given as follows for a convex storage function V . Note that if V is convex then
V satisfies (3.65) for all x ∈ Rnr if and only if ∂V
∂x (Aρ x) ⊥ ker Aρ (see [26]). In the
linear case (quadratic V ) and Aρ being given as the projection on the first vector
component, this is equivalent to the decoupled quadratic function V given by:
 
Q11 0
V (x) = x  x (3.72)
0 Q22

where Q11 and Q22 are positive-definite matrices such that Q11 ∈ Rnρ ×nρ and
Q22 ∈ Rnρ ×nρ . It is clear that V satisfies (3.65) since
 
 0 0
V (Aρ x) − V (x) = −x x ≤ 0, ∀x ∈ Rnr .
0 Q22

Remark 3.4 A well-known condition for an LTI system with state-space represen-
tation (A, B, C, D) to be SPR is that
 
−QA − A Q QB − C 
>0 (3.73)
B Q − C D + D

for some Q > 0. In addition, the system (A, B, C, D) is dissipative with respect
to the storage function V (x) = x  Qx (see [21]). Note that if the LMI (3.73) is
 
satisfied for a block diagonal Q = Q011 Q022 with the structure given by (3.72), then
Proposition 2 can be applied and as a result the partial reset compensator with base
system (A, B, C, D) will be VSP.

Some examples of passive partial reset compensators are given in the following.

3.3.2.1 Partial Reset PIIR Compensator

Consider the high order reset elements with base compensator PII R defined as

1 + γ1 Ti s 1 + γ2 Tr s
PII R (s) = k (3.74)
1 + T i s 1 + Tr s

where γ1 ∈ (0, 1) and γ2 ∈ (0, 1). Now, the transfer function is not enough to define
the reset compensator because, with the same Ti , Tr , γ1 , and γ2 , infinitely many
3.3 Reset Control Systems with Inputs/Passivity Analysis 123

Fig. 3.13 Reset choice in


PII R (s)

reset compensators can be defined. Thus, a state space description has to be given.
In particular, consider the base compensator

1 + 2s 1 + 0.012s
PII R (s) = (3.75)
1 + 3s 1 + s
given by the block diagram of Fig. 3.13, where the left block gives the state x1 = u1
to be reset.
Then, the compensator is defined by its state space matrices given by
   
−0.333 0.247 0.003
Ar = , Br = ,
0 −1 1
 
Cr = 0.444 0.658 , Dr = 0.008,

and
 
1 0
Aρ = . (3.76)
0 0
Note that by simply checking the Nyquist plot of the system (Ar , Br , Cr , Dr ) it
is straightforward to show that it is SPR (see Fig. 3.14). In addition, a decoupled
storage function has been found that satisfies (3.72), which is given by
 
1.20 0
V (x) = x  x. (3.77)
0 8.94

As a result, an application of Proposition 3.11 (see Remark 3.4) guarantees that


the partial reset compensator is VSP.

3.3.3 L 2 -stability of the Reset Control System

In this subsection, L2 -stability will be studied for the feedback system of Fig. 3.12,
by directly using Propositions 3.10 and 3.11, and the passivity theorem. Both the
full reset and the partial reset cases will be investigated.

Proposition 3.12 The reset feedback control system of Fig. 3.12, with R being a
full reset compensator with base compensator Rb , is finite-gain stable if some of the
following conditions are satisfied:
• Rb is ISP and P is ISP.
• Rb is OSP and P is OSP.
124 3 Stability of Reset Control Systems

Fig. 3.14 Nyquist plot of PII R (s)

• Rb is VSP and P is passive.


• Rb is passive and P is VSP.

In contrast to the full reset case, for the partial reset compensator a state space
description is necessary, the input–output map is not enough. In this case, as it was
shown in Proposition 3.11, a condition over the storage function has to be added.

Proposition 3.13 The reset feedback control system of Fig. 3.12, with R being a
partial reset compensator with base linear compensator Rb , is finite-gain L2 stable
if
1. Rb is dissipative with respect to the supply rate wi and a storage function V that
satisfies (3.65), and P is ISP.
2. Rb is dissipative with respect to the supply rate wo and a storage function that
satisfies (3.65), and P is OSP.
3. Rb is dissipative with respect to the supply rate wv and a storage function that
satisfies (3.65) and P is passive.
4. Rb is dissipative with respect to the supply rate wp and a storage function holds
(3.65) and P is VSP.

Note that if a system G is OSP (with parameters δG = 0 and εG > 0), then G
is finite-gain L2 stable with gain ≤ ε1G (see, for example, Theorem 2.2.14 in [29]).
Let us apply this to Propositions 3.12 and 3.13, for example, in the last case where
P is VSP (with parameters εP and δP ), and Rbl is passive (and thus δR = εR = 0).
Then, after simple manipulations we get that the closed-loop system (with inputs
3.3 Reset Control Systems with Inputs/Passivity Analysis 125

Fig. 3.15 Online pH process

w, d, and outputs u, y) is finite-gain L2 stable with gain less than or equal to


min{εP ,δP } = max{ εP , δP }. The other cases of Propositions 3.12 and 3.13 are similar.
1 1 1

As a result, the gain of the reset control system is less than or equal to the gain of the
base control system (while the precise computation of the gain of the closed-loop
gain should be performed by other means).

3.3.4 Example

In this example, Propositions 3.11–3.12 will be used to investigate the input–output


stability of a reset control system in which the plant is nonlinear; in particular, a reset
compensator for a pH control system will be designed based on the above described
passivity properties. Firstly, consider a stream (with flow rate F ) of distilled water
that is mixed with a titrating stream (with flow rate u) of a strong acid in a continuous
stirred tank reactor (CSTR) [34] with the following assumptions (see Fig. 3.15):
(a) the mixture is perfect and instantaneous, (b) the pH sensor in placed very near
to acid injection, and (c) the flow rate of titrating stream is negligible compared
with the flow rate of the process stream, i.e., F + u F . Under these restrictions,
the process can be modeled as a linear plant in cascade with a nonlinear passive
memoryless system (see Fig. 3.16).
The tank will be considered as the part of the pipe where acid is injected. Due
to conservation of charge, the concentration of the charge sum (W ) is taken as the
state of the linear subsystem, i.e., W = [H+ ] − [OH− ]. The dynamics of the CSTR
is given by
V Ẇ = −F (W − Wdw ) + uWa (3.78)
where V is the volume of the tank, F is the flow rate of the process stream, u is the
flow rate of the titrating stream and Wa is the concentration of charge in the titrating
stream. In particular, for distilled water Wdw = 0, but also a stream process of strong
base (Wb < 0) can be studied if −Wb " Wa . Furthermore, the output of the linear
system equals the state, yL = W . The relationship between this concentration and
126 3 Stability of Reset Control Systems

Fig. 3.16 Final model of pH


process under some
considerations

the pH is given by
 √ 
W+ W 2 + 4 × 10−14
pH(W ) = − log . (3.79)
2

Taking an operation point, with a flow rate constant uss , an equilibrium point is
obtained as −F Wss + uss Wa = 0. Therefore, Wss = uFss Wa . Working with incre-
ments with respect to the equilibrium point, ΔW = W − Wss , the new dynamics are
given by
˙ = −FΔW + ΔuWa
V ΔW (3.80)
where Δu = u − uss .
The nonlinear memoryless function is defined by ΔpH = pH (Wss ) − pH (W ).
The system was inverted in order for an increment in the input to result in an in-
crement in the output. Finally, the model was recast as the block interconnection of
Fig. 3.16, where L is an LTI system given by:

ẋ = −ax + bu,
(L) (3.81)
y = x,
Wa
where a = F
V and b = V . The static nonlinearity is given by

φ(x) = pH(Wss ) − pH(x), (3.82)

from which it follows that xφ(x) > 0 for all x = 0. It can be easily shown (see, for
example, [19] or [21]) that the system of Fig. 3.16 with L and φ given by (3.81)–
(3.82) is passive.
A numerical example is developed for the next data: V = 0.1 l, F = 5 × 10−2 l/s,
Wa = 1 mol/l, uss = 0.5 × 10−4 l/s. The linear plant from (3.81) is

20
L(s) = , (3.83)
2s + 1
and the static nonlinearity is given by
  √ 
x+ x 2 + 4 × 10−14
φ(x) = 3 − − log10 . (3.84)
2

A reset-PII R compensator has been also designed for tracking some pH ref-
erences, consisting in deviations about the equilibrium point. By using Proposi-
tion 3.12, finite-gain L2 -stability of the feedback system can be directly guaran-
teed since reset-PII R is VSP (see Sect. 3.3.2.1). A simulation of the control system
3.4 Describing Function Analysis 127

Fig. 3.17 Simulation of pH process with a partial reset compensator

is shown in Fig. 3.17, where in addition an important characteristic of reset com-


pensation is clearly shown: the possibility of obtaining very fast responses without
overshooting, overcoming fundamental limitations of LTI compensation.

3.4 Describing Function Analysis


The analysis of reset control systems with the describing function (DF) will follow
standard methods, and in the following it will be assumed that the reader is familiar
with the DF analysis (see, for example, [32] for a good introduction to the subject).
Although it is well-known that the describing function may fail in some cases, it has
proved to be an efficient method and has been formally justified in many practical
cases (see [14, 22]), including some types of switched systems (see [28]).
As it has been emphasized in this book, the main motivation for the use of reset
compensation is to improve the performance of a previously designed LTI control
system, with the goal to reset some states of a base LTI compensator to improve
control system performance both in terms of speed of response and relative stabil-
ity. In general, these specifications will be impossible to achieve by means of LTI
compensation. It turns out that the describing function technique, in spite of being
approximate, is a good practical tool to evaluate the phase lead obtained by reset
compensation. As it will be seen below, the fact that for reset compensators (with
128 3 Stability of Reset Control Systems

Fig. 3.18 FORE sinusoidal


response

zero error crossing as the resetting law) the describing function is only a function of
the input frequency (it does not depend on the input amplitude) makes the analysis
very simple, and thus is a good tool for a first stability analysis.
In the following, the describing functions of several basic reset compensators,
such as the Clegg integrator and FORE, are obtained.

3.4.1 FORE and Clegg Integrator

The FORE describing function is computed in this section, the Clegg integrator
describing function is then obtained as a particular case. FORE is a simple reset
compensator with a first order base compensator given by FOREbase (s) = s+aK
. As
it is well known (see, for example, [32]), for a given system its DF is calculated
as the ratio between the fundamental component of its sinusoidal response and the
sinusoidal input. The response v of the reset compensator FORE to the sinusoidal
input e(t) = E sin(ωt) is given in the s-domain simply by

K Eω
V (s) = , (3.85)
s + a s 2 + ω2

and the time response v(t) will consist of two terms: one corresponding to the tran-
sient response given by the compensator modes, and the other corresponding to the
steady response given by the sinusoidal input modes. Since the reset instants will
be given by the crossings of the sinusoidal input with zero, and thus they will be
periodic with fundamental period 2π/ω (see Fig. 3.18), it turns out that v(t) will
also be periodic with that period.
To compute the FORE describing function, a simple procedure is to add to the
base frequency response the contribution of the transitory terms vt (t) in v(t) over
ω ]. By symmetry (see Fig. 3.18), it will be computed by
a period, for example [0, 2π
3.4 Describing Function Analysis 129

using the interval [0, πω ], where

KEω −at
vt (t) = e . (3.86)
a 2 + ω2

The result is
π/ω
1
vt (t)e−j ωt ω dt j 2Kω2  π
π 0
= 1 + e−a ω , (3.87)
E
2j
π(a + j ω)(a + ω )
2 2

and finally adding to the frequency response of the base system, the describing func-
tion of FORE results as
 
2ω2 (1 + e−a ω )
π
K
FORE(ω) = 1+j . (3.88)
a + jω π(a 2 + ω2 )

Note the simple form of the describing function that only depends on the fre-
quency. This allows us to make direct comparisons of the FORE describing func-
tion and the frequency response of its base system. The Clegg integrator describing
function can be simply obtained by setting K = 1 and a = 0 in (3.88):
 
1 4
CI(ω) = 1+j , (3.89)
jω π

and since 1 + j π4 = 1.62ej 51.85 , it is clear that Clegg integrator gives a phase lead
of almost 52° over an integrator (it also increases the gain with a factor of 1.62
approximately). Figure 3.19 shows this fundamental property of the Clegg integrator
and FORE, that is, the achievement of a phase lead up to 52◦ with respect their base
linear compensator.

3.4.2 Reset Compensators with Reset Band

In control practice, the compensator implementation is usually done by using a reset


band. In addition, it has been noted that the use of a reset band may improve stability
and performance in systems with time-delays [30] due to the fact that phase lead can
be even improved by using a reset band; see Sect. 5.2 for a more detailed justification
of how reset band can improve reset control system stability and performance. In
this section, the focus will be on the computation of the DF for a number of basic
reset compensators with a reset band. In these cases, the resetting law is given by
the crossing of the error signal when entering some (fixed or variable) band.
130 3 Stability of Reset Control Systems

Fig. 3.19 Nichols plot of FORE and CI describing functions

Reset control systems with a reset band are feedback control systems (see
Fig. 2.1) where the compensator R is given by the impulsive differential equation


⎪ ẋr (t) = Ar xr (t) + Br e(t) if (e(t), ė(t)) ∈
/ Bδ ,

(R) xr (t + ) = Aρ xr (t) if (e(t), ė(t)) ∈ Bδ , (3.90)



v(t) = Cr xr (t) + Dr e(t),

where the reset band surface Bδ is given by Bδ = {(x, y) ∈ R2 |(x = −δ ∧ y >


0) ∨ (x = δ ∧ y < 0)}, with δ some non-negative real number. In this way, the com-
pensator states are reset at the instants at which its input is entering into the reset
band. In general, the reset band surface Bδ will consist of two reset lines Bδ+ and
Bδ− in the plane, as shown in Fig. 3.20.
In the particular case δ = 0, the standard reset compensator is obtained. On the
other hand, if δ is big enough in relation to the error amplitude, then no reset action
is produced, and the reset compensator reduces to its base compensator given by the
transfer function Rb (s) = Cr (sI − Ar )−1 Br + Dr .
The following notation will be used to distinguish between reset compensators
with and without reset band: Rδ will denote a reset compensator with reset band Bδ ,
3.4 Describing Function Analysis 131

Fig. 3.20 Reset band

Fig. 3.21 FORE with reset


band sinusoidal response

whereas R0 (or simply R) is the standard reset compensator without reset band.
Note that Rb stands for the base reset compensator. For simplicity, the describing
function (DF) of a FOREδ compensator is considered first.

3.4.2.1 FOREδ

To obtain the describing function with a reset band, again the sinusoidal response
needs to be calculated. In this case, note that since FOREδ is time-invariant, the
FOREδ response to e(t) = E sin(ωt) over the period [−tδ , 2π ω − tδ ] is the same
(except for a translation) as its response to e(t − tδ ) over the period [0, 2π ω ] (see
Fig. 3.21). Thus, FOREδ DF can be obtained as the contribution of the transitory
terms in the response to e(t − tδ ) = E cos(ωtδ ) sin(ωt) − E sin(ωtδ ) cos(ωt), plus
the base frequency response. But note that the transitory terms consist of two terms:
one term v1t corresponding to the response to E cos(ωtδ ) sin(ωt), and a second term
vt2 with the response to −E sin(ωtδ ) cos(ωt).
δ ) −at
By using (3.86), the term vt1 is directly given by vt1 = KEa 2cos(ωt
+ω2
e , and thus
using again symmetry arguments, its harmonic balance (corresponding to the fun-
132 3 Stability of Reset Control Systems

damental component) is given by


π/ω KEω2 cos(ωtδ )
(1 + e−a ω )
π
1
0 vt1 (t)e−j ωt ω dt π(a+j ω)(a 2 +ω2 )
π
2π/ω
= E −j ωtδ
. (3.91)
1
2π 0 e(t − tδ )e−j ωt ω dt 2j e

sin(ωtδ ) −at
On the other hand, the second transitory term is vt2 = − KEa a 2 +ω2
e , and its
harmonic balance is now

(1 + e−a ω )
KEaω sin(ωtδ ) π
1 π/ω
π 0 vt2 (t)e−j ωt ω dt π(a+j ω)(a 2 +ω2 )
2π/ω = E −j ωtδ
. (3.92)
1
2π 0 e(t − tδ )e−j ωt ω dt 2j e

Finally, the FOREδ describing function is obtained by adding to (3.91) and (3.92)
the base frequency response. After some simple manipulations the result is
.
π   2 
FOREδ (E, ω) j 2ω(1 + e−a ω ) δ δ j sin−1 ( δ )
=1+ ω 1− +a e E , (3.93)
FOREb (j ω) π(a 2 + ω2 ) E E

where FOREb (j ω) = j ω+a K


, and the identities sin(ωtδ ) = Eδ and cos(ωtδ ) =
/
2
1 − Eδ have been used. Note that in contrast to (3.89), FOREδ (E, ω) depends
both on the magnitude and the frequency of the input. In general, expression (3.93)
is valid for 0 ≤ Eδ ≤ 1. Obviously, for δ > E, no reset action is performed, and thus
the describing function will directly be the base frequency response.
Note that the FOREδ introduces extra phase lead for frequencies over a1 , in com-
parison with FORE for some values of δ (see Fig. 3.22). Since in addition for in-
creasing values of δ a significant phase lag is introduced, it seems that a proper
selection of δ should be in the range 0 < δ/E < 0.8.

3.4.2.2 Clegg Integrator with a Reset Band

The describing function of the Clegg integrator with a reset band, that will be re-
ferred to as CI δ , is obtained as a particular case of FOREδ for K = 1 and a = 0. The
result is
/
 j 4( 1 − ( Eδ )2 ) 
1 −1 δ
CI δ (E, ω) = 1+ ej sin ( E ) . (3.94)
jω π
Again, note that the reset band introduces extra phase lead for values 0 < δ/E <
0.8 and at every frequency (see Fig 3.23).

3.4.2.3 General Reset Compensator with Reset Band

In the case in which the general reset compensator (3.90) has a state matrix Ar with
real and distinct eigenvalues, the compensator may be implemented as a sum of
3.4 Describing Function Analysis 133

Fig. 3.22 FOREδ (E, ω). Phase for different values of δ/E

Fig. 3.23 CI d (E, ω). Phase for different values of δ/E

different FORE compensators, and thus the calculation of its describing function is
straightforward. In addition, partial reset may be implemented by considering the
base frequency response of the non-reset element.
On the other hand, although the derivation is somehow involved (see Appendix A
of [6]), a closed-form expression of the describing function can be obtained for the
134 3 Stability of Reset Control Systems

general case of full reset with a reset band. The result is:

Rδ (E, ω) = Rbase (j ω)
 
1 1 δ  π Ar −1 δ
Cr e ω Ar sin −1( E ) G , ω I + e−Ar ω · e( ω −j I )(π−sin ( E )) Br ,
δ
+
π E
(3.95)

where G is given by
 −1 −1 
G(x, ω) = (j ωI − Ar )−1 e−j sin (x) + (j ωI + Ar )−1 ej sin (x) · ω(j ωI − Ar )−1 .
(3.96)
Using (3.95)–(3.96) with δ = 0, a general closed-form expression can also be
obtained for the describing function R0 (j ω) of a full reset compensator without a
reset band. It is given by

1 1  π Ar
R0 (E, ω) = Rb (j ω) + Cr e ω Ar π F (0, ω) I + e−Ar ω · e( ω −j I )(π−π) Br (3.97)
π
and after simplification results in (note that it does not depend on E)
ω 
R0 (j ω) = Rb (j ω) + Cr (j ωI − Ar )−1
π
  π 
+ (j ωI + Ar )−1 · (j ωI − Ar )−1 eAr ω + I Br . (3.98)

Note that in general R0 (j ω) will depend only on frequency ω, and Rδ (E, ω) both
on the amplitude E and the frequency ω. Obviously, the expression for FOREδ as
given in (3.93) is a particular case of (3.95) for Rb (j ω) = j ω+a
K
, Ar = −a, Br = 1,
Cr = K. In addition, the CI δ FD expression in (3.94) is a particular case for K = 1
and a = 0.

3.4.2.4 Reset Systems with a Variable Reset Band

A variable reset band is considered as an alternative to overcome the influence of,


for example, dominant time delays over the reset action. In practical applications
(see [15, 31]), this has been shown to be a good strategy to achieve stability and
performance design specifications (see Chap. 5 for a detailed exposition).
In the fixed band case, the reset instants are determined when the error signal is
approaching zero at a fixed distance given by the reset band. A variable reset band
means that at every reset instant the band value may be different, and in general this
value will be determined by a combination of the error value and its derivative. The
derivative of a function at a chosen input value describes the best linear approxi-
mation of the function near that input value. In this way, the reset instants will be
predicted by using the error signal derivative approximation. This approximation
states that the error signal derivative at a point is equal to the slope of the tangent
3.4 Describing Function Analysis 135

Fig. 3.24 Variable reset band


surface

line to the graph of the error signal at that point. Therefore, the following expression
is obtained
de(t) −e(t)
tan α = = , (3.99)
dt h
where h is the time delay process. From this equation, the new variable reset condi-
tion is expressed as
de(t)
h + e(t) = 0. (3.100)
dt
As it can be seen, no parameter has to be fixed in this new condition, since the
time delay h can be estimated from plant tests (usually a worst-case estimate is a
good option). With this new reset condition (3.100), the reset controller (3.90) is
expressed now in the state-space as


⎪ ẋr (t) = Ar xr (t) + Br e(t) if (e(t), ė(t)) ∈
/ Bhv ,

(R) xr (t + ) = Aρ xr (t) if (e(t), ė(t)) ∈ Bhv , (3.101)



v(t) = Cr xr (t) + Dr e(t),

where the variable reset band surface Bhv is given by


 
Bhv = (e(t), ė(t)) ∈ R2 | hė(t) + e(t) = 0 . (3.102)

In this case, the reset surface is a continuous function of the error signal, as it
can be seen in Fig. 3.24. When there is no time delay, h = 0, the reset band surface
(3.102) is reduced to the reset compensator without reset band.
In the frequency domain, the describing function of the Clegg integrator with a
variable reset band may be computed, resulting in
 
1 j2 −1 
CI h (ω) = 1+ 1 + ej 2 tan (ωh) , (3.103)
jω π
136 3 Stability of Reset Control Systems

Fig. 3.25 Phase of CI h

where h is the time delay of the process. Note that in the case of the variable re-
set band the describing function only depends on the input frequency and that, in
contrast to CI, CI h gives a frequency dependent phase lead. Figure 3.25 shows the
phase of CI h for different values of h. It is clear from this figure that CI h gives an
extra phase lead over a CI up to 27◦ in the frequency interval [ 0.1
h , h ] (and up to 75
1 ◦

over an integrator) that is especially useful for systems with time delays.

3.4.3 Limit Cycle Analysis

Consider the feedback system of Fig. 2.1. The question now is to analyze the ex-
istence of limit cycles of reset control systems with a reset band by using the de-
scribing function of the reset compensator obtained above. A well-known method
is to compute possible values of E and ω such that the describing function of the
reset compensator Rδ (E, j ω) satisfies 1 + Rδ (E, j ω)P (j ω) = 0, which is usually
referred to as the harmonic balance principle. This is usually made by analyzing
the crossings of the Nyquist plot of the plant P with plots of −1/Rδ (E, j ω) vs ω
for different values of E. It is well known that this method only gives approximate
results, and it works well only if the plant P (s) filters higher order harmonics of the
periodic signal v in Fig. 2.1 (low pass condition).
Alternatively, the above procedure is equivalent to finding the crossings of the
critical locus −1/Rδ (E, ω) with the plant Nyquist plot P (j ω), and then obtaining
crossings for values of E and ω that will correspond to the amplitude and frequency
of a limit cycle. In the following, two examples are developed for the reset com-
pensators FOREδ and CI δ , respectively. Note that since, in general, their describing
3.4 Describing Function Analysis 137

Fig. 3.26 Time simulation of feedback system with FOREδ and first order plant (note that e = −y)

functions depend on the parameter δ/E, where δ/E ∈ (0, 1) defines the reset band
in relation to the compensator input amplitude, if a limit cycle exist for some δ then
there always exist limit cycles for any value of E such that δ/E ∈ (0, 1). Thus,
a normalized reset band δ/E and a describing function with amplitude E = 1 will
be used.
As a result, the critical locus −1/Rδ/E (1, ω) will be used to evaluate the cross-
ings, and then values of δ/E and ω will be obtained instead. It turns out that for each
given frequency ω, critical loci −1/Rδ/E (1, ω) are circle segments in the Nyquist
plane as δ/E goes from 0 to 1.

3.4.3.1 FOREδ with a First Order Plant

A simple analysis shows that there are no limit cycles in the case corresponding to
a feedback control system (see Fig. 2.1) with a FOREδ reset compensator having a
stable base compensator and a first order stable plant P (s). The reason is that the
after-reset state of the closed-loop system will always correspond to a zero compen-
sator output, and an alternating value of the error δ, −δ, . . . is obtained. The only
case in which a limit cycle may exist is the case in which the error signal reaches
the reset band from each one of the two values δ and −δ, and this is only possible if
the base feedback system is unstable.
In this example, a FOREδ compensator with an unstable base compensator
FOREb (s) = s−1 2
, δ = 0.1, and a first order plant P (s) = s+0.5
1
is considered. A sim-
ulation of the feedback system reveals (see Fig. 3.26) that a limit cycle exists with
amplitude E = 0.20 and frequency ω = 1 rad/s.
138 3 Stability of Reset Control Systems

Fig. 3.27 Crossings of critical loci −1/FOREδ/E (1, ω) for ω = 0.5, 1, 1.5 rad/s with Nyquist plot
of the first order plant

In Fig. 3.27, the crossings of critical loci −1/FOREδ/E (1, ω) for ω = 0.5, 1,
1.5 rad/s with P (j ω) are shown. Then, it is possible to evaluate numerically those
crossings corresponding to a limit cycle, and in addition values of δ/E and ω. The
result is that there exist limit cycles for any value of E such as δ/E = 0.68 and for
ω = 1 rad/s. In this example, the amplitude will be E = δ/0.68 = 0.1/0.68 = 0.15.
Note that the describing function method gives a very accurate prediction of the
limit cycle. In addition, a perturbation method can be used to analyze the (local)
stability of the limit cycle. It can be shown that for the limit cycle to be stable a
sufficient condition is
∂R(E, ω) ∂I (E, ω) ∂R(E, ω) ∂I (E, ω)
− >0 (3.104)
∂E ∂ω ∂ω ∂E
at (E0 , ω0 ) = (0.15, 1), where R(E, ω) = Re(1 + FOREδ (E, ω)P (j ω)) and
I (E, ω) = Im(1 + FOREδ (E, ω)P (j ω)). In this case, condition (3.104) can be
checked numerically, and as a result the limit cycle is stable.

3.4.3.2 CIδ with a Second Order Plant

Consider again the feedback system of Fig. 2.1, in this case with a Clegg integra-
tor CI δ as reset compensator with δ = 0.25 and a second order plant P (s) = s+1 s2
.
A simulation has been made first to analyze the existence of limit cycles, the result
is that there exists a limit cycle as shown in Fig. 3.28 with amplitude E = 0.5 and
frequency ω = 0.92 rad/s.
Let us consider now the use of the describing function method. Different critical
loci −1/CI δ/E (1, ω) for ω = 0.1, 0.3, 0.5, 0.92 rad/s are shown in Fig. 3.29. After a
numerical evaluation of the different crossings, the result is that there exists a limit
3.4 Describing Function Analysis 139

Fig. 3.28 Crossings of critical locus −1/CI b (E, ω) with Nyquist plot of the plant

cycle for any value of E such as δ/E = 0.82 and ω = 0.92 rad/s. For this example,
the amplitude is E = δ/0.82 = 0.3. Again, the prediction of the limit cycle is very
accurate. The deviation of the amplitude is consistent with the fact that signal v
in Fig. 3.28 may have significant higher order harmonics. In addition, condition
(3.104) is satisfied for (E0 , ω0 ) = (0.3, 0.92), and thus the limit cycle is also stable.

3.4.4 Justification of the Describing Function

As it is well known, in spite of the fact that the describing function technique gives
only approximate results, it has been found useful in many practical applications.
Although a formal justification is clearly desirable, it turns out that the several
analytical justifications that can be found in the literature are of limited use. The
first work in this direction is [14], where it is shown that a limit cycle exists, us-
ing Fig. 2.1, if R is bounded and P is able to filter enough high order harmonics
(low pass condition). Also, the works [22] and [28] develop formal results for non-
linear systems that are sector-bounded or exhibit a passivity condition, respectively.
These results are not applicable to reset control systems, and in general an analytical
treatment seems to be a hard open problem.
In the following, the low pass condition will be explored for some particular
types of reset control systems with a reset band. In addition, the harmonics structure
of the sinusoidal response of FOREδ and CI δ will be investigated. In general, the
response of FOREδ to a sinusoidal input e(t) = E sin(ω(t − tδ )) is periodic with
140 3 Stability of Reset Control Systems

Fig. 3.29 Crossings of critical locus −1/CI b (E, ω) with Nyquist plot of the second order plant

fundamental period 2π/ω (see Fig. 3.21). In addition, it is also symmetric and can
be expressed as
⎧ KE(ω cos(ωt )+a sin(ωt ))


δ δ
e−at + M(ω)E sin(ω(t − tδ ) + Φ(ω))

⎪ a 2 +ω2

⎨ if t ∈ [0, π ],
ω
v(t) = (3.105)

⎪ −y(t − πω )




if t ∈ [ πω , 2π
ω ],

where M(ω) and Φ(ω) are the magnitude and phase of | j ω+a K
|.
Now, computation of high order harmonics is relatively simple in this case since
by symmetry only odds harmonics are different from zero. Defining C(E, ω, δ) =
KEω(ω cos(ωtδ )+a sin(ωtδ ))
π(a 2 +ω2 )
(1 + e−aπ/ω ), higher harmonics have a particularly simple
form. After some computation, spectral coefficients of the Fourier series v(t) =
 ∞ j kωt are given by (see Fig. 3.30)
k=−∞ ak e

C(E,ω,δ)
for k = ±3, ±5, . . . ,
ak = a+j kω (3.106)
0 for k = ±2, ±4, . . . .
3.4 Describing Function Analysis 141

Fig. 3.30 Sinusoidal response of FOREδ for ω = 1 rad/s, K = 1, and a = −1, and approximations
with the first plus third harmonics, and with 20 harmonics

Therefore, by using the first n odd harmonics of FOREδ response as an approx-


imation, a measure of the approximation can be defined as the L2 [0, 2π ω ]-norm
of the residual term. Then, v(t) as given by (3.105) can be expressed by v(t) =

vn (t)+ v̂n (t) where vn (t) = nk=−n a2k−1 ej (2k−1)ωt , and v̂n (t) = v(t)−vn (t). Then


 ∞
 C
v̂n 2 = 2 |a2k−1 |2 = 2 , (3.107)
a2 + (2k − 1)2 ω2
k=n+1 k=n+1

where C stands for C(E, ω, δ) (since C does not depend on k, the arguments have
been dropped for simplicity). By using the ψ function (also known as the digamma
function), the norm can be computed in a closed-form as
       
C ω + ia ω − ia C ω + ia
v̂n  =
2
Ψ n+ −Ψ n+ = Im Ψ n+
i2aω 2ω 2ω aω 2ω
(3.108)
for a = 0. Figure 3.31 shows a measure of how low order harmonics represent the
sinusoidal FOREδ response. Note that in general the ratio v
v̂n 
only depends on the
142 3 Stability of Reset Control Systems

v̂n 
Fig. 3.31 Ratio (1 − v ) vs n (number of odds harmonics) for | ωa | = 1, | ωa | = 0.5, and
| ωa | = 0.25

parameters a and ω, and is given by


 1
v̂n  Im{Ψ (n + 1
+ i 2ω
a
)} 2
= 2
. (3.109)
v Im{Ψ ( 12 + i 2ω
a
)}

For the Clegg integrator with a reset band CI δ , the norm is


       
C 1+i 1−i C 1+i
v̂n 2 = Ψ n + − Ψ n + = Im Ψ n + ,
i2ω2 2 2 ω2 2
(3.110)
and in this case the ratio v
v̂n 
is constant (independent of the frequency ω), and is
given by (it has the same values as that of FOREδ for |a/ω| = 1, see Fig. 3.32)
 1      1
v̂n  Im{Ψ (n + 12 (1 + i))} 2
1 2
= = 0.8331 Im Ψ n + (1 + i) .
v Im{Ψ ( 2 (1 + i))}
1 2
(3.111)
For the example of FOREδ plus a first order plant, the low-pass condition can
be now evaluated by computing how the plant filters the harmonics of the FOREδ
sinusoidal response. In Fig. 3.33, it is shown how for different values of |a/ω| the
relative weight of the first harmonic is much more significant, and it is a good ap-
proximation of the filtered FOREδ response. For the example of CI δ with a second
order plant, the results are analogous.
Note that the harmonics structure of the FOREδ or the CI δ response is always
given as shown in Fig. 3.31, then the low-pass condition can be evaluated for a
3.4 Describing Function Analysis 143

Fig. 3.32 Sinusoidal response of CI δ for ω = 0.92 rad/s, and approximations with fundamental
component, with the first plus third harmonics, and with 20 harmonics

v̂n 
Fig. 3.33 Ratio (1 − v ) vs n (number of odds harmonics) for | ωa | = 1, | ωa | = 0.5, and
| ωa | = 0.25
144 3 Stability of Reset Control Systems

given plant by computing how the harmonics are filtered, and then evaluating how
the plant modifies the harmonics structure, for example, as shown in Fig. 3.33. This
figure shows that the relative error norm of the first harmonic approximating the
FOREδ response is around 10%, and that the relative error considering both the
first and third harmonics is around 5%, thus it may be expected that the describing
function method works well in many cases.

References
1. Aangenent, W.H.T.M., Witvoet, G., Heemels, W.P.M.H., van de Molengraft, M.J.G., Stein-
buch, M.: An LMI-based L2 gain performance analysis for reset control systems. In: Pro-
ceedings of the American Control Conference (2008)
2. Bainov, D.D., Simeonov, P.S.: Systems with Impulse Effect: Stability, Theory and Applica-
tions. Ellis Horwood, Chichester (1989)
3. Baños, A., Carrasco, J., Barreiro, A.: Reset times-dependent stability of reset control system.
In: European Control Conference, Kos, Greece (2007)
4. Baños, A., Carrasco, A., Barreiro, A.: Reset times dependent stability of reset systems with un-
stable base system. In: Proc. IEEE International Symposium on Industrial Electronics, Spain
(2007)
5. Baños, A., Dormido, S., Barreiro, A.: Stability analysis of reset control systems with reset
band. In: 3rd IFAC Conference on Analysis and Design of Hybrid Systems, Zaragoza, Spain
(2009)
6. Baños, A., Dormido, S., Barreiro, A.: Limit cycles analysis in reset systems with reset band.
Nonlinear Anal. Hybrid Syst. (2010). doi:10.1016/j.nahs.2010.07.004
7. Baños, A.: Carrasco, J., Barreiro, A.: Reset times-dependent stability of reset control systems.
IEEE Trans. Autom. Control 56(1), 217–223 (2011)
8. Bemporad, A., Bianchini, G., Brogi, F.: Passivity analysis and passification of discrete-time
hybrid systems. IEEE Trans. Autom. Control 53(4), 1004–1009 (2008)
9. Beker, O.: Analysis of reset control systems. Ph.D. Thesis, University of Massachusetts,
Amherst (2001)
10. Beker, O., Hollot, C.V., Chait, Y., Han, H.: Fundamental properties of reset control systems.
Automatica 40, 905–915 (2004)
11. Camlibel, M.K., Heemels, W.P.M.H., Schumacher, J.M.: On linear passive complementary
systems. Eur. J. Control 8(3), 220–237 (2002)
12. Carrasco, J., Baños, A., van der Schaft, A.: A passivity approach to reset control of nonlinear
systems. In: 34th Annual Conference of the IEEE Industrial Electronics Society, Orlando,
Florida, USA (2008)
13. Carrasco, J., Baños, A., van der Schaft, A.: A passivity-based approach to reset control systems
stability. Syst. Control Lett. 59(1), 18–24 (2010)
14. Bergen, A.R., Franks, R.L.: Justification of the describing function method. SIAM J. Control
9, 568–589 (1971)
15. Fernández, A., Barreiro, A., Baños, A., Carrasco, J.: Reset control for passive bilateral teleop-
eration. IEEE Trans. Ind. Electron. (2010). doi:10.1109/TIE.2010.2077610
16. Guo, Y., Wang, Y., Xie, L., Zheng, J.: Stability analysis and design of reset systems: theory
and an application. Automatica 45(2), 492–497 (2009)
17. Haddad, W.M., Chellaboina, V., Nersesov, S.G.: Impulsive and Hybrid Dynamical Systems:
Stability, Dissipativity, and Control. Princeton University Press, Princeton (2006)
18. Horowitz, I.M., Rosenbaum, P.: Nonlinear design for cost of feedback reduction in systems
with large parameter uncertainty. Int. J. Control 24(6), 977–1001 (1975)
19. Khalil, H.K.: Nonlinear Systems, 2nd edn. Prentice Hall, New York (2002)
References 145

20. Krishman, K.R., Horowitz, I.M.: Synthesis of a nonlinear feedback system with significant
plant-ignorance for prescribed system tolerances. Int. J. Control 19(4), 689–706 (1974)
21. Lozano, R., Brogliato, B., Egeland, O., Maschke, B.: Dissipative Systems Analysis and Con-
trol. Springer, London (2000)
22. Mees, A., Bergen, A.: Describing functions revisited. IEEE Trans. Autom. Control 20(4),
473–478 (1975)
23. Michel, A., Hu, B.: Towards a stability theory of general hybrid dynamical systems. Automat-
ica 35, 371–384 (1999)
24. Nešić, D., Zaccarian, L., Teel, A.R.: Stability properties of reset systems. Automatica 44(8),
2019–2026 (2008)
25. Pogromski, A., Jirstrand, M., Spangeus, P.: On stability and passivity of a class of hybrid sys-
tems. In: Proceedings of the 37th IEEE Conference on Decision and Control, Tampa, Florida,
USA (1998)
26. Rockafellar, R.T., Wets, J.-B.: Variational Analysis. Series of Comprehensive Studies in Math-
ematics, vol. 317. Springer, Berlin (1998)
27. Rugh, W.J.: Linear System Theory, 2nd edn. Prentice Hall, New York (1996)
28. Sanders, S.R.: On limit cycles and the describing function method in periodically switched
circuits. IEEE Trans. Circuits Syst. I 20(4), 473–478 (1975)
29. Van der Schaft, A.: L2 -Gain and Passivity Techniques in Nonlinear Control. Springer, Berlin
(2000)
30. Vidal, A., Baños, A.: QFT-based design of PI + CI compensators: application in process con-
trol. In: IEEE Mediterranean Conference on Control and Automation, pp. 806–811 (2008)
31. Vidal, A., Baños, A.: Reset compensation for temperature control: experimental application
on heat exchangers. Chem. Eng. J. 159(1–3), 170–181 (2010)
32. Vidyasagar, M.: Nonlinear Systems Stability. Prentice-Hall, London (1993)
33. Willems, J.C.: Dissipative dynamical systems—Part I: General theory. Arch. Ration. Mech.
Anal. 45, 321–351 (1972)
34. Wright, R.A., Kravaris, C.: On-line identification and nonlinear control of an industrial pH
process. J. Process Control 11, 361–374 (2001)
35. Zefran, M., Bullo, F., Stein, M.: A notion of passivity for hybrid systems. In: Proceedings of
the 40th IEEE Conference on Decision and Control, Orlando, FL, USA (2001)
Chapter 4
Stability of Time-Delay Reset Control Systems

4.1 Introduction

Since the early proposals for reset control, including the Clegg integrator [13] and
the first-order reset element (FORE) [20], [15], the underlying idea was to achieve
fast and robust control solutions for problems under linear limitations. The simple
nonlinear effect of resetting was proved to be useful for overcoming linear funda-
mental limitations.
As it is well known in the literature on fundamental limitations [1] (see also
Chap. 1 of this monograph), linear plants that have time-delays (and also plants
having right half-plane poles or zeros) are specially hard to control: if the delay is
significant, it is impossible to obtain a linear controller providing simultaneously a
fast and a robust behavior. One has to choose between a fast response but with poor
stability margin, or a robust stability margin, but with slow response. As a conse-
quence, since reset control is a simple nonlinear method for overcoming fundamen-
tal limitations, and since time-delay is one of the main sources of such limitations,
it is of great interest to study the problem of delayed reset systems.
A design idea to obtain a reset controller for a plant with time-delay can be
taken from the design approach proposed in [15] and [20]: first, tune the base LTI
controller so that the base LTI feedback loop becomes stable and satisfies part of
the performance specifications (typically, fast transient and large bandwidth). Then,
in a second step, a reset action is included to achieve the other part of the desired
objectives (typically, bounded overshoot and large stability margins).
However, this method should be carefully applied since it is well known [8] that
the reset action may destabilize a well designed LTI base control system, and this
fact is one of the main drawbacks for the practical application of reset control.
A solution to this problem has been reported in [8, 9] for finite-dimensional LTI
plants. The result is a frequency domain condition, the Hβ -condition (see the pre-
vious Sect. 3.2.1.1 of this monograph), that guarantees closed-loop stability of the
reset system.
Since then, a number of useful results have appeared addressing the analysis and
design of reset systems [6], [17], [18], [25]. However, none of them considered the

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 147


DOI 10.1007/978-1-4471-2250-0_4, © Springer-Verlag London Limited 2012
148 4 Stability of Time-Delay Reset Control Systems

case of time-delay plants, so it is then of great interest to study delayed reset systems.
It should be emphasized that this is extremely difficult: the combination of reset and
delay moves the problem to the field of hybrid and distributed systems. Even the
issue of existence and uniqueness of solutions may cause trouble unless special care
is taken, for example, by means of the time-regularization procedure [25].
It should be noted that stability of delay impulsive systems is a very active
research area in different areas of mathematics and systems theory, two classi-
cal monographs are [2] and [21]. In general, as reported in [2] (see the previous
Sect. 1.5.1 of Chap. 1 in this monograph), impulsive systems may be classified into
(i) systems with impulses at fixed instants, (ii) systems with impulse effect at vari-
able instants, and (iii) autonomous systems with impulse effects. Systems (ii) and
(iii) have state-dependent impulses, but there is an important difference: the reset
surface is time independent in the case (iii). Traditionally, most of the research ef-
fort has been dedicated to cases (i) and (ii), and thus case (iii) that includes reset
control systems is less developed. In addition, in the control field, the main effort
has been directed at systems without time delays [28], [18]. For the case of impul-
sive delay systems, available results in the literature are concentrated in the case (i)
(see [22], [14], [24], and the references therein), and case (ii) (see [23]).
The material in this chapter is based on the journal papers [5] and [7], and on
the preliminary conference papers [3] and [4]. Some of the results in these refer-
ences can be regarded as generalizations of the Hβ -condition for LTI plants with
time-delays. In [5], the case of delay-independent stability was considered, a stan-
dard Lyapunov–Krasovskii functional was used, and the instrumental technique for
passing between time and frequency domains was the Kalman–Yakubovich–Popov
(KYP) lemma. In [7], the delay-dependent case was addressed. As the combination
of delay-dependent LMIs with the KYP lemma may give rise to extremely large ma-
trix inequalities, which do not show much insight into the results, special care was
taken in the choice of appropriate Lyapunov functionals and passivity/frequency
domain techniques.
There are several Lyapunov–Krasovskii (LK) functionals that can be used, so
that the delay h appears explicitly in the stability conditions. Among these delay-
dependent results, the complete LK functional (Theorem 5.19 in [16]) has the ad-
vantage that it gives a sufficient but also necessary condition for stability; however,
it is a distributed functional and requires some kind of discretization for practical
computation.
Another useful result is given by Proposition 5.17 in [16], with good LMI com-
putability and which is not too conservative, in general. This functional was studied
in [4] for application to reset systems. Finally, for interpretability in terms of un-
certainty, the LK functional in [26] has the advantage that it allows for an easy
interpretability (detailed in [29]) so that the stability conditions can be understood
in terms of scaled small-gain for an uncertain plant. In the delay-dependent case,
for passing from the time-domain to the frequency-domain, in place of the involved
KYP techniques in [5], a novel interpretation was presented in [7] based on passivity
and impulsive control. In this way, all the stability conditions can also be interpreted
in terms of robust stability against certain uncertainties.
4.2 Problem Motivation and Statement 149

Fig. 4.1 Reset control loop


for a time-delay plant

The organization of the chapter is as follows. After introducing formally the prob-
lem and basic results in Sect. 4.2, the time-domain delay-independent stability of re-
set systems is studied in Sect. 4.3 with the equivalent frequency domain conditions
and interpretations presented in Sect. 4.4, followed by an example in Sect. 4.5. Then,
the delay-dependent stability results are considered in Sect. 4.6 with the frequency
domain counterpart in Sect. 4.7, including a passivity analysis of reset systems and
robustness interpretations. The chapter concludes with an illustrative example of the
delay-dependent case in Sect. 4.8.

4.2 Problem Motivation and Statement

Consider a control system formed by a plant P0 and a reset controller R (both single-
input single-output):

ẋp = Ap xp + Bp u,
(P0 ) (4.1)
y = Cp x p ,


⎪ ẋr = Ar xr + Br e,

(R) xr (t + ) = Aρ xr (t), (4.2)



v = Cr xr + Dr e,
where the second equation of the controller represents the impulsive or reset actions,
applied when a certain reset condition, to be specified later, holds. If the connection
from P0 to R (or from R to P0 ) is affected by a delay h1 (or h2 ), then it is easy to
see that the total delay h = h1 + h2 can be moved to the plant input or output.
For instance, move the delay to the plant input, as shown in Fig 4.1. As the
main interest is stability, let us suppose that the exogenous loop signals are zero
(r = d = n = 0), so that the closed loop connections are:

u(t) = v(t − h), e(t) = − y(t), (4.3)

and the closed loop system is easily described by


     
ẋp (t) x (t) xp (t − h)
=A p + Ad , (4.4)
ẋr (t) xr (t) xr (t − h)

and by impulsive dynamics


 
xr t + = Aρ xr (t).
150 4 Stability of Time-Delay Reset Control Systems

With this is mind, we generalize to the system given by the impulsive-delayed


differential equation (IDDE)

ẋ(t) = Ax(t) + Ad x(t − h) if x(t) ∈
/ M,
(IDDE) (4.5)
x(t + ) = AR x(t) if x(t) ∈ M ,

where x = (x  
p , xr ) is the global state and, from (4.1)–(4.2), the matrices A, Ad
take the form
   
Ap 0 −Bp Dr Cp Bp Cr
A= , Ad = .
−Br Cp Ar 0 0

Since the results in this chapter are independent of the precise structure of A, Ad ,
let us assume in what follows, for the sake of generality, that A, Ad are arbitrary
matrices, not necessarily coming from a control loop like the one in Fig. 4.1.
Now, to complete the system equations, it remains to introduce AR and M . Let
us suppose that only the second part of the controller states are reset to zero, that is,
Aρ = diag(1, . . . , 1, 0, . . . , 0). Then AR takes the form [9]:

AR = diag(In12 , 0n3 ), (4.6)

so, the reset action acts over the last n3 states of the state x ∈ Rn . These dimensions
are defined by:
n = n1 + n2 + n3 = n1 + n2 +n3 , (4.7)
 
 

n23 n12

where n1 (n23 ) is the number of states of the plant (controller). Accordingly, parti-
tion the state vector:
 
x = x  
1 , x2 , x3 . (4.8)

 

x
p x
r

In principle, it is considered that the reset is applied at the time t when x(t) hits
the hypersurface
 
M = x ∈ Rn : Cx = 0 . (4.9)
As it is desired that 0 = Cx defines a hypersurface, C is a row vector, C ∈ R1,n .
From the state partition (4.8), the reset condition matrix C can be partitioned as:

C = (C1 , C2 , C3 ).

There are several possibilities for C. If the IDDE system comes from a control
loop like the one in Fig. 4.1, and the reset condition is activated when the error
crosses zero, then C = (Cp , 0, 0). The most general case is an arbitrary row ma-
trix C, but it causes some difficulties if C3 = 0, that is, if the part x3 of the controller
states affected by reset have influence via C3 on the reset condition. For this reason,
several results in this chapter assume the structure C = (C1 , C2 , 0) for C.
4.2 Problem Motivation and Statement 151

In this way, (4.5), (4.6), and (4.9) define, in principle, the system that we are
going to study. However, it is known that the IDDE (4.5) may exhibit beating and
have Zeno solutions as has been remarked in [25] for the case h = 0. A certain
strategy for avoiding these pathological solutions has been analyzed in Chap. 2 of
this monograph. This problem of well-posedness of solutions affects all hybrid sys-
tems, and in particular reset systems. Note that the well-posedness issue has been
addressed in Chap. 2 for finite-dimensional reset systems. However, the existence of
the time-delay complicates the situation further because it makes the system infinite-
dimensional. In this chapter, the useful idea of time regularization (TR) will be used,
and thus the IDDE (4.5) is recast in the form (IDDE-TR):

τ̇ = 1, ẋ(t) = Ax(t) + Ad x(t − h) if (x(t) ∈


/ M ) ∨ (τ ≤ ρ),
  (4.10)
τ + = 0, x t + = AR x(t) if (x(t) ∈ M ) ∧ (τ > ρ),

where ρ is a finite positive scalar, arbitrarily small. In this way, every reset in (4.10)
will take place at least ρ seconds after the previous reset instant. That is, the reset
instants tk (k = 1, 2, . . .) satisfy tk+1 − tk > ρ (k = 1, 2, . . .). Choosing a sufficiently
small time-step ρ the behavior the IDDE and IDDE-TR systems (4.5) and (4.10) is
almost equal for most practical systems, and in this way the regularized system is
formally free of dynamic pathologies.
Now, it is important to distinguish between the (lumped) instantaneous state
x(t) ∈ Rn and the true (distributed) state given by some piecewise continuous func-
tion xt : [−h, 0] → Rn , where xt (θ ) = x(t + θ ), θ ∈ [−h, 0]. In the following,
C([−h, 0], Rn ) will stand for the set of piecewise continuous functions mapping
[−h, 0] to Rn , with a finite number of discontinuity points.
For the sake of concreteness, the piecewise continuous functions in
C([−h, 0], Rn ) will be assumed continuous from the left at the discontinuity points,
that is, x(tk ) = x(tk− ) and, after a reset, x(tk+ ) = AR x(tk ).
Then, the existence and uniqueness of solutions to (4.10) is simply guaranteed
in a constructive way. First, as the counter τ starts at τ = 0, there exists a unique
solution to the following initial value problem:

ẋ = Ax(t) + Ad x(t − h) for t ≥ 0,
(4.11)
x(θ ) = x0 (θ ) for θ ∈ [−h, 0],

where x0 (·) is the initial condition. The starting trajectory of the linear system (4.11)
is then well defined and is not affected by the first reset until some instant t1 > ρ.
Now, in a recursive way, if the trajectory is well defined until some reset instant
t = tk , then introduce the reset jump x(tk+ ) = AR x(tk ) and after the reset, solve a
new linear problem similar to (4.11). In a constructive way, the solution is
h
x(t) = e A(t−tk )
AR x(tk ) + eA(t−tk −σ ) Ad x(tk + σ − h) dσ, (4.12)
0

and this solution is valid for all t ∈ (tk , tk+1 ], up to reaching the new reset time
tk+1 > tk + ρ.
152 4 Stability of Time-Delay Reset Control Systems

For the stability analysis of (4.10), let us rewrite the system in a form that makes
explicit only the state x and implicit the time counter τ :

ẋ(t) = Ax(t) + Ad x(t − h),
(IDDE-TRx ) (4.13)
x(t + ) = AR x(t) only if x(t) ∈ M .

This should be understood as equivalent to (4.10) but we only pay attention to


x(t), driven by the continuous dynamics most of the time, and subject to reset only
if x(t) ∈ M (note that the ‘iff’ condition is not used, due to the fact that τ < ρ may
disable the reset).
Now, let us introduce a basic result to address the stability of time-delay reset
systems. It is common practice in stability analysis of time-delay systems [16] to
start with a Lyapunov–Krasovskii functional V (xt ) and force its derivative to be
negative, V̇ (xt ) < 0, along the solutions of the system.
If the system also contains jumps or resettings, then it is natural to impose that
the variation of the functional at the reset instants t = tk has to be nonpositive,
ΔV (t) = V (xt + ) − V (xt ) ≤ 0.
There are other possibilities, different from imposing V̇ < 0 and ΔV ≤ 0 as basic
stability conditions. These requirements may be conservative, in general, and within
the area of impulsive systems there is a vast amount of literature offering alternative
techniques. But, at the same time, notice that many reset design approaches (see,
e.g., [15], [20], [9], [18]) are based on the useful idea of tuning the linear system
to be fast transient, underdamped but stable, and introducing reset for adding ex-
tra damping and decreasing overshoot. Within this idea, both the continuous and
the impulsive mode must be stable, and so the basic stability principle (V̇ < 0 and
ΔV ≤ 0) is very well suited and adequate.
This principle is formalized in the following proposition taken from [7] and based
on the results in [9] for systems without time-delays.

Proposition 4.1 Let the functional V (xt ) : C([−h, 0], Rn ) → R be continuously


differentiable, positive-definite, radially unbounded and such that along the solu-
tions of the reset system
d
V (xt ) < 0, xt = 0, (4.14)
dt
ΔV = V (xt + ) − V (xt ) ≤ 0, x(t) ∈ M . (4.15)
Then, the zero equilibrium xt = 0 is globally asymptotically stable.

Proof The detailed proof is omitted for brevity, but it is quite similar to the proof of
Theorem 1 in [9], and to the proof of Proposition 3.1 in [5]. The basic idea is that,
from the decreasing behavior, V (xt ) → c ≥ 0 as t → ∞. But if we suppose that
c > 0, then xt would always lie outside a ball around the origin 0, so that necessarily
V̇ < −γ , by continuity, for some γ > 0. But hence V would become negative, and
so by contradiction c = 0. 
4.3 Delay-Independent Conditions in the Time Domain 153

4.3 Delay-Independent Conditions in the Time Domain

In this section, we study the stability of reset control systems, the results are valid
for any of the formulations: the nonregularized IDDE (4.5), the explicit regularized
IDDE-TR (4.10), and the implicit regularized IDDE-TRx (4.13). The results will be
stated for some of the models, (4.5), (4.10), or (4.13), but they are equally valid for
all of them because the regularization only affects well-posedness, and the stability
treatment is formally the same.
According to Chap. 5 in [16], the time-domain approach to stability of time-
delayed systems can be based on the Razumikhin Theorem for Lyapunov functions
or on Lyapunov–Krasovskii functionals. In any case, the resulting conditions can be
delay-independent (valid for all h > 0) or delay-dependent.
Since the reset actions complicate the treatment, compared to time-delay systems
without reset, then for the sake of simplicity let us consider Lyapunov–Krasovskii
functionals and let us start with the delay-independent case. From Sect. 5.4 in [16],
perhaps the simplest stability criterion can be obtained from the simplest Lyapunov–
Krasovskii functional:
0

V (t) = V (xt ) = xt (0) P xt (0) + x
t (θ )Qxt (θ ) dθ, (4.16)
−h

for some P , Q > 0. It is clear that V (t) ≥ 0 and that V (t) = 0 if and only if
xt (θ ) = 0 ∈ Rn , for each θ ∈ [−h, 0]. Thus the functional is positive-definite, and
also differentiable and radially unbounded. It is clear as well that the zero state
xt = 0 is a fixed point, or equilibrium point, of the reset system. The question now
is whether this equilibrium is stable or not. First, let us formalize a little the sense
of stability to be used in relation to the aforementioned quadratic functional.

Definition 4.1 The IDDE reset control system (4.5) is said to be quadratically
(asymptotically) stable when there exist P , Q > 0 such that the functional V in
(4.16) satisfies the conditions (4.14) and (4.15).

These two inequalities ensure that there exists a common Lyapunov function for
the system (4.5) without reset (the base LTI system), using (4.14), and for the reset
system xt + (0) = AR xt (0) using (4.15).

Proposition 4.2 The system (4.5) is quadratically stable if and only if there exist
matrices P , Q > 0, of size n × n, satisfying the following conditions:
  
A P + P A + Q P Ad
< 0, (4.17)
A dP −Q

and
 
Θ  A
R P AR − P Θ ≤ 0 (4.18)
for any matrix Θ with Im Θ = Ker C.
154 4 Stability of Time-Delay Reset Control Systems

Proof Condition (4.17) is commonly used in delay-independent stability of time-


delay systems [16], and easily follows from the fact that (4.16) can be written as
0
V (xt ) = x(t) P x(t) + x (t + θ )Qx(t + θ ) dθ,
−h

and thus its derivative easily gives

d
V (xt ) = x (t)P ẋ(t) + ẋ (t)P x(t) + x (t)Qx(t)
dt
− x (t − h)Qx(t − h),

so, using (4.5), the result is


    
x(t) A P + P A + Q P Ad x(t)
V̇ = ,
x(t − h) A
dP −Q x(t − h)

from which (4.17) is finally derived. It remains to prove that condition (4.18) follows
from (4.15). Since the reset action is only active when xt (0) ∈ M , and thus does
not affect the delay buffer x(t + θ ), for any θ ∈ (−h, 0), the integral part of the
Lyapunov functional (4.16) does not contribute to the jump, thus the reset jump in
the Lyapunov functional ΔV in (4.15) results in that
  
ΔV = x t (0) AR P AR − P xt (0) ≤ 0,

for every xt (0) ∈ M = Ker C, from which (4.18) directly follows. 

Note that the reset action adds condition (4.18) to the standard condition of delay-
independent stability given by the linear matrix inequality (LMI) (4.17). Since the
matrix AR has a very particular structure, this new condition can be considered as
an extra LMI, or simply interpreted as restriction over the structure of the matrix P
that is a solution of the LMI.
Let us study the structure of condition (4.18). First, notice that from (4.6) AR has
the form
AR = diag(Inρ̄ , Onρ ), (4.19)
where, for convenience, the dimensions in (4.7) have been renamed

nρ = n3 , nρ̄ = n − nρ = n1 + n2 ,

so that nρ is the number of (controller) states affected by reset and nρ̄ is the number
of (plant and controller) states not affected by reset. Now, partitioning P according
to (4.19), we have:
   
P1 P2 P1 0
A
R P A R = A 
R P P A R = , (4.20)
2 3 0 0
4.4 Delay-Independent Conditions in the Frequency Domain 155

with P1 > 0 square of size nρ̄ = n − nρ , thus (4.18) becomes


    
      θ1    0 −P2 θ1
θ1 θ2 AR P AR − P = θ1 θ2 ≤ 0, (4.21)
θ2 −P2 −P3 θ2

where Θ  = (θ1 θ2 ). Now, assume here that C has the form C = (C1 , 0) with C1
of size (1 × nρ̄ ), as typically the last nρ states do not influence the reset condition.
Notice that in this case x = (x  
1 , x2 ) belongs to Ker C = Ker(C1 , 0) if and only if
C1 x1 = 0. Furthermore, (4.21) is nonpositive when x 1 P2 x2 = 0 for all x1 , x2 such
that C1 x1 = 0. This amounts to P2 = C1 Mβ for some matrix Mβ , and hence the
next proposition easily follows.

Proposition 4.3 The system (4.5), with C = (C1 , 0) and C1 a row of length n − nρ ,
is quadratically stable if and only if there exist matrices P , Q > 0, with
 
P1 C1 Mβ
P= (4.22)
Mβ C1 P3
that satisfy (4.17), where Mβ is some (column) matrix of size nρ × 1.

4.4 Delay-Independent Conditions in the Frequency Domain


The availability of frequency domain stability conditions is important for many well-
known reasons: integration of analysis and design, treatment of uncertainty, etc. In
addition, previous stability conditions developed in reset systems [9] and [8] are in
the frequency domain, and give usually easier checkable conditions than its LMIs
counterpart. They are obtained by use of the Kalman–Yakubovich–Meyer (KYM)
lemma. In addition, there exist frequency domain approaches to the stability of de-
lay systems (see [12]) based on generalizations of the Kalman–Yakubovich–Popov
(KYP) lemma.
In this section, frequency domain conditions for time-delay reset systems will be
derived. Again the move from time to frequency domain will require some general-
ization of the KYP lemma as well. A nice, general version of KYP can be found in
[27]. We recall here this result:

Proposition 4.4 (Generalized KYP lemma) Given (A, B) controllable with


(sI − A)−1 B having no poles on the imaginary axis, and a matrix M = M  of
adequate size, the following statements are equivalent:
 ∗  
(j ωI − A)−1 B (j ωI − A)−1 B
(i) M ≤ 0, for all ω ∈ R ∪ {∞}.
I I
(4.23)
  
A P +PA PB
(ii) There exists P = P  such that M + ≤ 0. (4.24)
B P 0
The equivalence for strict inequalities holds even if (A, B) is not controllable.
156 4 Stability of Time-Delay Reset Control Systems

4.4.1 The Hβ -condition

In order to find how to use the generalized KYP lemma to obtain frequency con-
ditions for (4.17) and (4.18), it will be useful to recast the results in [8] using this
lemma. In the following, the Hβ -condition for stability of reset systems without
time-delay will be proved by using a different proof to that in [8]. It is included here
since it will reveal a basic procedure to be followed afterwards in the time-delay
case.

Proposition 4.5 Consider the IDDE system (4.5) with Ad = 0 and reset surface
 
M = x(t) ∈ Rn : Cx(t) = 0 ,

where the row vector C has the form

C = (C1 , 0),

with C1 of length nρ̄ = n − nρ (nρ is the number of reset states). Assume also that A
is Hurwitz and (A, B0 ) is controllable, where the input matrix is B0 = [0nρ̄ Inρ ].
Then, the IDDE system (4.5) is quadratically stable if and only if Hβ (j ω) =
C0 (j ωI − A)−1 B0 is strictly positive real (SPR) for some Mβ and some P3 > 0,
where B0 = [0nρ̄ Inρ ] and C0 = [Mβ C1 P3 ].

Proof The necessary and sufficient conditions for quadratic stability in this case are
as follows: First, ∃P > 0
A P + P A < 0, (4.25)
for the time derivative of the Lyapunov function V (x(t)) = x (t)P x(t)to be nega-
tive along the trajectories of the system except for the reset times, and, second,
 
x (t) AR P AR − P x(t) ≤ 0, ∀x(t) ∈ M (4.26)

for the jump ΔV to be ≤ 0 in the reset surface. Now, condition (4.26) yields, fol-
lowing (4.20) and (4.21),
 
 0 −P2
x (t) x(t) ≤ 0, ∀x(t) ∈ M , (4.27)
−P2 −P3

which gives rise, by Proposition 4.3, to a structure for P in the form


 
P1 C1 Mβ
≥0 (4.28)
Mβ C1 P3

for some column matrix Mβ of dimensions nρ ×1, and some P3 > 0. This restriction
can be finally recast as
B0 P = C0 (4.29)
for B0 = [0nρ̄ Inρ ] and C0 = [Mβ C1 P3 ].
4.4 Delay-Independent Conditions in the Frequency Domain 157

Notice that (4.25) and (4.29) are equivalent to ∃P > 0, ∃ε > 0


  
A P + P A + 2εP P B0 − C0
≤ 0. (4.30)
B0 P − C0 0

Now, (4.30) can be rewritten in the form (ii) in KYP lemma:


    
0 −C0 Aε P + P Aε P B0
∃P > 0, ∃ε > 0, + ≤ 0, (4.31)
−C0 0 B0 P 0
 

M0

where Aε = A + εI . So, using KYP, since (A, B0 ) is controllable, ∃ε > 0 such that
 ∗  
(j ωI − Aε )−1 B0 (j ωI − Aε )−1 B0
M0 ≤ 0 for all ω ∈ R ∪ {∞}.
I I
(4.32)
Notice that from the last equation (statement (i) in KYP) we recover (4.31) (state-
ment (ii) in KYP), but with ∃P = P  in place of ∃P > 0. However, since A is Hur-
witz, the existence of a symmetric P such that A P + P A = Q < 0 implies [19]
that such P is positive-definite, i.e., P > 0. Hence, the last two equations (4.31) and
(4.32) are completely equivalent.
From the last condition, putting Hβ (j ω) = C0 (j ωI − A)−1 B0 and noticing that
j ωI − Aε = j ωI − (A + εI ) = (j ω − ε)I − A, the positive real condition re-
quired for Hβ (s − ε), for some small ε > 0, is obtained. In other words (see [19]), it
amounts to the strictly positive real (SPR) requirement on the system Hβ (j ω) given
by the triple (A, B0 , C0 ), which is precisely the so-called Hβ -condition in [9]. 

Remark 4.1 It can be seen that the previous SPR condition, characterized in [19], is
equivalent to the Output Strictly Passive (OSP) condition as defined in [10]. More
precisely, since A is Hurwitz and Hβ is strictly proper, the SPR condition amounts,
following [19], to Hβ (j ω) + Hβ (j ω) > 0 for all finite ω, and for ω → ∞ to
 
lim ω2 Hβ (j ω) + Hβ (j ω) > 0.
ω→∞

On the other hand, [10] characterizes the OSP condition, with A Hurwitz and
(A, B0 ) controllable, by
 
Hβ (j ω) + Hβ (j ω) ≥ ε|Hβ (j ω)|2

for some ε > 0 small enough. By expanding Hβ (s) = H1 /s + H2 /s 2 + · · · and


putting s = j ω with ω → ∞, the equivalence between the two previous conditions,
that is, the equivalence between the SPR and OSP requirements for a strictly proper
Hβ (s) can be verified.
We will find again the OSP condition later, in Sect. 4.7, where the passage from
time to frequency domain, instead of using KYP, will be solved by an approach
based on approximating the resets by suitable Dirac delta pulses.
158 4 Stability of Time-Delay Reset Control Systems

Note that the Hβ -condition gives quadratic stability by checking if Hβ is SPR


for some Mβ and some P3 . In the original formulation [8], P3 was simplified to the
identity P3 = I , but this is not the general case. This was rectified in the subsequent
work [9]. Since the SPR condition is easily checkable specially in the scalar case
corresponding to the reset of only one state (nρ = 1), it has been proved to be an
efficient method for stability analysis of reset control systems. In the following,
application of the above procedure is extended to cope with the time-delay case.

4.4.2 The Generalized Hβ -condition

Proposition 4.6 Consider the IDDE system (4.5) under the same conditions as
in Proposition 4.5, that is, with M given by Cx = 0 and with C = (C1 , 0),
but with Ad = 0. Assume also that A is Hurwitz and (A, B1 ) controllable, with
B1 = (B0 , Ad ) and B0 = [0nρ̄ Inρ ].
Then, this IDDE system is quadratically stable if and only if for some P3 , Q > 0,
for some Mβ and for some ε > 0, the KYP lemma is satisfied with the matrices
A, M, B in the KYP frequency condition (4.23) given by

A → Aε = A + εI,
B → B1 = (B0 Ad ),

and
⎛ ⎞
Q −C0 0
M → Mε = ⎝ −C0 0 0 ⎠ (4.33)
0 0 −Q + εI
where
C0 = [Mβ C1 P3 ].

Proof The treatment of the reset-delay conditions in (4.17) and (4.18) can follow a
completely parallel manipulation to conditions in Proposition 4.5. First, notice from
Proposition 4.3 that (4.18) is equivalent to (4.22). This condition has already been
reformulated in the form
(0 I ) P = (Mβ C1 P3 ), (4.34)
 
 

B0 C0

where I and P3 inside B0 , C0 have size nρ × nρ .


Now, let us try to reproduce manipulations similar to those from (4.30) up
to (4.32). The starting point is: Quadratic stability amounts to the existence of
P , Q > 0 satisfying (4.17) and (4.34), rephrased as
  
A P + P A + Q + 2εP P Ad
≤ 0, B0 P = C0 , (4.35)
Ad P −Q + εI
4.4 Delay-Independent Conditions in the Frequency Domain 159

for some ε > 0, which, imitating (4.25)–(4.30) is equivalent to


⎛  ⎞
Aε P + P Aε + Q P Ad P B0 − C0
∃P > 0, ∃ε > 0, ⎝ AdP −Q + εI 0 ⎠ ≤ 0, (4.36)

B0 P − C0 0 0
where Aε = A + εI . Now, reproducing (4.30) → (4.31), and simultaneously in-
terchanging the second and third row-blocks and column blocks, we obtain that
∃ε > 0:
⎛ B1

⎛ 

 
Q −C0 0 ⎜ A P + P A P (B , A ) ⎟
⎝ −C0 0 0 ⎠+⎜ ⎜  
ε ε 0 d ⎟
⎟ ≤ 0. (4.37)
⎝ B ⎠
0 0 −Q + εI 0 P 0
 
A
d

The last equation is ready for KYP, reaching to the frequency condition for
quadratic stability: ∃ε > 0,
 ∗  
(j ωI − Aε )−1 B1 (j ωI − Aε )−1 B1
Mε ≤ 0 for all ω ∈ R ∪ {∞},
I I
(4.38)
where Mε , B1 are given in (4.37). This concludes the proof. 

The matrix Mε could be diagonalized in the form Mε = F  J F , with J diago-


nal with entries ±1 or 0. Thus, depending on the signature of Mε (the number of
positive, zero, and negative eigenvalues), condition (4.38) might be presented more
suitably as a small-gain frequency condition, or as a positive-real frequency condi-
tion.

4.4.3 Interpretation of the Stability Conditions

The previous conditions (4.37) and (4.38) are very useful for a quick implementa-
tion of the stability tests, however, they are based on big matrices that do not really
help in understanding the nature of the stability conditions. To clarify this, an inter-
pretation will be given in terms of the structural properties of the dynamics.
This will be achieved in three steps. In the first step, the IDDE system (4.5)
is equivalently represented by the block diagram in Fig. 4.2. The upper loop real-
izes the continuous dynamics ẋ = Ax + Ad Dh x (the first equation in (4.5)), where
Dh ≡ e−j ωh is the delay operator, by assigning

Cd = I n , Bd = Ad , Δd → Dh , (4.39)

where the delay Dh ≡ e−j ωh is treated as a particular case of an uncertainty Δd


satisfying small-gain:
Δd (j ω) ≤ 1 for all ω. (4.40)
160 4 Stability of Time-Delay Reset Control Systems

Fig. 4.2 Block diagram


representation of the IDDE
system

The impulsive or reset action is implemented by the lower loop which is zero most
of the time. It only works at the reset instants when the reset states (the last nρ
states) are switched to zero. This resetting is easily shown to be equivalent to the
application of Dirac impulses δ(·) in the following form. Consider the reset system
Rr from yr to ur defined by:

(Rr ) ur (t) = − δ(t − tk )P3−1 yr (t), (4.41)
k

with
Cr = (Mβ C1 , P3 ), Br = (0, Inρ ) ,
where tk are the reset instants characterized by Cx(tk ) = 0 and by tk ≥ tk−1 + ρ.
The equivalence stems from the fact that, around the reset instant tk , the signal ur
contains the delta function with weight:

ur (t) = −δ(t − tk )P3−1 yr (t), t ∈ (tk − ε, tk + ε), ε < ρ.

Thus, the instantaneous response of ẋ = Ax + Bd ud + Br ur = Ax + Ad x(t − h) +


Br ur to a Dirac input in ur is given by a state jump of value
tk+
   
Δx = x tk+ − x tk− = Br ur = −Br P3−1 yr (tk )
tk−
 
0
=− P3−1 (Mβ C1 , P3 )x(tk ).
I

Now, since at the reset instants


 
x12 (tk )
0 = Cx(tk ) = (C1 , 0) = C1 x12 (tk )
x3 (tk )

where the state has been partitioned making explicit the last nρ coordinates x3 (the
coordinates affected by reset),
   
    0 0
Δx = x tk+ − x tk− = = − ,
−P3−1 P3 x3 (tk ) x3 (tk )
4.4 Delay-Independent Conditions in the Frequency Domain 161

which is the precise change that produces the required reset. Thus so far we have
proved that the IDDE system can be realized in the block diagram form given by
Fig. 4.2. To complete the first step, notice that the resetting subsystem Rr : yr →
−ur satisfies
∞ 
−ur (τ )yr (τ ) dτ = yr (tk )P3−1 yr (tk ) ≥ 0, (4.42)
0 k

so, as shown in Fig. 4.2, the resetting block can be treated as a particular case,
Δr → Rr of a general uncertainty Δr : yr → −ur satisfying the positivity condition:

−ur = Δr yr ⇐⇒ −ur (τ )yr (τ ) dτ ≥ 0. (4.43)
0

In the second step, the interpretation of the upper loop condition is provided. The
quadratic stability of the continuous dynamics amounts to (4.14), V̇ < 0, and to the
first LMI in (4.35). This LMI can be treated within the KYP lemma, with M =
diag(Q, −Qε ), for Qε = Q − εI , and it is easy to show that it is equivalent to
Gd (j ω)
-
  -
- 1/2 -
-Q (j ωI − Aε ) Ad Q−1/2
−1
ε - < 1 for all ω, (4.44)

where Gd (s) is the open-loop nominal (Δd = 0) transfer function of the upper loop
in Fig. 4.2 (from ud to yd ). In this way, (4.44) is a Q-scaled strict small-gain con-
dition on the uncertain upper loop. The small-gain condition is strict due to the
presence of the small ε > 0 in Aε = A + εI and Qε = Q − εI .
It is important to notice that stability is achieved for all values of the delay 0 ≤
h < ∞. This means, in particular, stability for h = 0 (so that A + Ad is stable)
and for h → ∞ (so that A stable). Notice also that, from (4.35), A P + P A ≤
−Q − 2εP < 0, so that x P x is a Lyapunov function for the system ẋ = Ax.
In the third and final step, an interpretation of the lower loop condition is given.
As the lower loop performs the resettings, the required condition (4.15) is ΔV ≤ 0
at the reset instants. It is equivalent to the second condition in (4.35), rewritten now
as Br P = Cr . Furthermore, it amounts (given A P + P A < 0) to:
  
Aε P + P A P Br − Cr
∃ε > 0, ε ≤ 0. (4.45)
Br P − Cr 0

Recasting the previous LMI via the KYP lemma, it easily follows

∃ε > 0, 0 ≤ G∗r (j ω − ε) + Gr (j ω − ε), (4.46)

where Gr (s) = Cr (j ωI − A)−1 Br is the open-loop nominal (Δr = 0) transfer func-


tion of the lower loop in Fig. 4.2 (from ur to yr ). In this way, (4.45) is a strictly
positive real (SPR) condition on the uncertain lower loop.
Summarizing, the LMIs (4.37) and (4.38), expressed in terms of large matrices,
have been interpreted here in terms of structural properties of the IDDE system.
162 4 Stability of Time-Delay Reset Control Systems

Basically, the mechanized procedure of KYP and LMIs (computationally oriented)


corresponds to the underlying structure in Fig. 4.2, in which the delay is treated as
a small-gain uncertainty and the resetting is viewed as a positive uncertainty.
Using this framework, different conditions might be deduced, by encapsulating
the delay or the reset action within alternative uncertainty structures. Furthermore,
the reset action could be changed freely and redesigned, provided it remains within
the required sector bounds.
In the delay-dependent case, Remark 4.6 (that can be found in Sect. 4.7) will
present an interpretation in terms of robustness for some uncertain loop (see Fig. 4.6)
very similar to the one in this section. The main difference is that the delay-
dependent case will be addressed by passivity techniques applied to an equivalent
system where reset is approximated by Dirac impulses.

4.5 Example: Delay-Independent Stability

To illustrate the stability results, let us consider an example based on Example 5.7
in [16]. Consider a time-delay reset system with state x(t) = (x1 (t) x2 (t)) given
by

ẋ(t) = Ax(t) + Ad x(t − h),

with
   
−2 0 −0.25 0
A= , Ad = , (4.47)
0 −0.9 −0.25 −0.25
endowed with the reset law
 
x2 t + = 0 when x1 (t) = 0.

This system is stable irrespective of the delay in the case of no reset [16]. In
the following, it will be also shown that it is also stable if we reset the state x2 (t).
Following Proposition 4.6, in this case B0 = (0 1), and C0 = (Mβ p3 ), where p3
is simply a positive real number, p3 = 0.1, and Mβ = 0 is a possible choice, then
C0 = (0 0.1). In addition, a positive-definite matrix Q is chosen as
 
0.1 −0.05
Q= .
−0.05 0.1

Thus, using Proposition 4.6, the matrix Mε is given by (for ε = 0.01)


⎛ ⎞
Q −C0 0
Mε = ⎝ −C0 0 0 ⎠
0 0 −Q + εI
4.5 Example: Delay-Independent Stability 163

Fig. 4.3 The largest eigenvalue of H (j ω) vs ω

⎛ ⎞
0.1 −0.05 0 0 0
⎜ −0.05 0.1 −0.1 0 0 ⎟
⎜ ⎟

=⎜ 0 −0.1 0 0 0 ⎟ ⎟,
⎝ 0 0 0 −0.09 0.05 ⎠
0 0 0 0.05 −0.09

and finally, the frequency response matrix H (j ω) resulting from the KYP lemma
(4.38) is
H (j ω) = G (j ω)MG(j ω),
where
 
(j ωI − Aε )−1 B1
G(s) =
I
with Aε = A + εI and B1 = (B0 , Ad ). The matrix H (j ω) is easily shown to be
negative semidefinite for every frequency ω > 0, thus it satisfies the KYP condition
(4.38). Figure 4.3 shows the largest eigenvalue of H (j ω) versus ω (using a loga-
rithmic ω axis). Since these values are always below zero, H (j ω) ≤ 0 ∀ω > 0, and
thus the reset system given by (4.47) is stable independently of the delay.
164 4 Stability of Time-Delay Reset Control Systems

Fig. 4.4 Simulation of the delay-independent example: x1 (t) (dashed) and x2 (t) (solid)

Notice the behavior as ω → ∞ in Fig. 4.3: G(j ω) tends to (02×3 , I3×3 ) , and
thus H (j ω) tends to the 3 × 3 submatrix of Mε formed by its last three rows and
columns. This submatrix has a zero eigenvalue, and this explains the high frequency
limit in Fig. 4.3.
Figure 4.4 shows a simulation of the example with a delay h = 2 s, and with
the initial conditions x1 (θ ) = x2 (θ ) = 0 for θ ∈ [−2, 0) and x1 (0) = 10, x2 (0) = 0.
It can be seen that initially (t < h = 2) the state x1 (t) tends to zero with the sta-
ble dynamics given by A = diag(−2, −0.9). But when t = h = 2 s, the influence
of Ad x(t − h) appears and the trajectories x(t) are affected, giving rise to several
resets at the times t1 ≈ 2.1, t2 ≈ 4.6, t3 ≈ 7.3, . . .. These reset times are defined by
x1 (tk ) = 0. The simulation of this reset system for h = 2 s shows a stable behavior.
Since we have proven that this system is stable independently of the delay h, we can
as well predict stability for all fixed h ≥ 0.

4.6 Delay-Dependent Conditions in the Time Domain

This section addresses delay-dependent stability of reset systems, that is, we will
search for adequate Lyapunov–Krasovskii functionals such that, from an application
4.6 Delay-Dependent Conditions in the Time Domain 165

of the basic Proposition 4.1, the derived stability conditions depend explicitly on the
delay h. In this way, the stability conditions may hold only for some values of h. This
is more natural since many practical systems are typically stable for small values of
the delay, but unstable for large delays.
In order to gain generality, we assume here for C the structure C = (C1 , C2 , 0),
based on the state partition (4.8) x = (x   
1 , x2 , x3 ) and the corresponding dimen-
sions (4.7) n = n1 +n2 +n3 . This structure is equivalent to the one used in the delay-
independent case, with the notation equivalence nρ = n3 and nρ̄ = n−nρ = n1 +n2 .
The proposed structure C = (C1 , C2 , 0) is more general than C = (C̃1 , 0, 0), the
one adopted in [9]. For comparison with this last case, let us state the fact that there
exists an invertible transformation T , T −1 ∈ Rn×n such that



 
 
C T

 
(C1 , C2 , 0) = (C̃1 , 0, 0) diag(T12 , In3 ) . (4.48)

One possible solution is C̃1 = (1, 0, . . . , 0) and T12 with its first row equal to
(C1 , C2 ) and the remaining rows completed for full rank.
Now, the key point to obtain delay-dependent stability conditions is to choose
an appropriate Lyapunov–Krasovskii (LK) functional. There are several LK func-
tionals that can be used for generating delay-dependent conditions. The complete
LK functional (Theorem 5.19 in [16]) has the advantage that it gives a sufficient
but also necessary condition for stability, however, it is a distributed functional and
requires discretization for practical computation. Another useful result is given by
Proposition 5.17 in [16], with good LMI computability and which is generally not
too conservative. Finally, the LK functional in [26] has the advantage that the stabil-
ity conditions are interpretable in a robust stability framework, in terms of a certain
small-gain loop [29]. This LK functional takes the integral-quadratic form given by:

V (xt ) = V1 (xt ) + V2 (xt ) + V3 (xt ), (4.49)

with
1
V1 (xt ) = x (t)P x(t), (4.50)
2
0
V2 (xt ) = x (t + θ )Qx(t + θ ) dθ, (4.51)
−h
and
0 0
V3 (xt ) = G(xt+ξ ) A
d XAd G(xt+ξ ) dξ dθ, (4.52)
−h θ
where
G(xt+ξ ) = Ax(t + ξ ) + Ad x(t + ξ − h). (4.53)
In this way, by applying Proposition 4.1 to V (xt ), we will arrive at conditions
ensuring global and asymptotical stability. The requirement that d[V (xt )]/dt < 0
166 4 Stability of Time-Delay Reset Control Systems

has been already treated in [26], and it is finally expressed in an LMI form. The
requirement that ΔV (xt ) ≤ 0 when x(t) ∈ M can be addressed in a special form,
related to the one in Theorem 5 of [9]. By combining these two approaches, we
obtain the following proposition.

Proposition 4.7 The IDDE reset system (4.5) is globally and asymptotically stable,
if there exist Q > 0, V > 0, W , and

P = P  = (Pij )i,j =1,2,3 > 0, (4.54)

with
(0, 0, In3 )P = (Mβ (C1 , C2 ), P33 ), (4.55)
for some column matrix Mβ of appropriate size, and such that
⎛ ⎞
N −W Ad A A
dV h(W + 12 P )
⎜ −Q A  ⎟
⎜ d Ad V 0 ⎟ < 0, (4.56)
⎝  −V 0 ⎠
 0 0 −V

where
1 
N= (A + Ad ) P + P (A + Ad ) + W Ad + A 
d W + Q,
2
and each block “” is the transpose of the corresponding symmetric block.

Proof From Proposition 4.1, what has to be proved is that (i) dV /dt < 0 and
(ii) ΔV ≤ 0 for x ∈ M . It can be seen that (4.56) implies (i), this has been already
proved in [26]. So, it only remains to prove that (4.55) implies (ii). Actually, it will
be shown that (4.55) and (ii) are equivalent, that is, (4.55) holds iff (ii) is true. This
equivalence will be made explicit by expressing ΔV ≤ 0 for x ∈ M as a quadratic
form. First, observe from (4.48) with C = C̃T that M = Ker C = T −1 Ker C̃, that
is,
M = Ker C = T −1 Ker C̃ = T −1 Imag diag(K1 , In2 , In3 ), (4.57)
 

K
with K1 such that
Imag K1 = Ker C̃1 . (4.58)
Then, it can be seen that, after a reset jump x(t) → AR x(t), the LK functional
V = V1 + V2 + V3 only changes in its first contribution V1 because the integrals in
V2 and V3 are not affected. In this way, the condition ΔV = ΔV1 ≤ 0 for x ∈ M =
T −1 Imag K amounts to the negative semidefiniteness, P̄ ≤ 0, of

P̄ = K  T − (AR P AR − P ) T −1 K, (4.59)
 

Π(P )
4.6 Delay-Dependent Conditions in the Time Domain 167

which defines Π(P ). But, it can be seen that


 
P̄ = K  Π T − P T −1 K = K  Π(P̃ )K. (4.60)
 

This is true because AR and T −1 commute, due to their block-diagonal structure.


Now, partitioning P̃ = (P̃ij ) for i, j = 1, 2, 3, it holds that
⎛ ⎞
0 0 P̃13
Π(P̃ ) = − ⎝ 0 0 P̃23 ⎠ . (4.61)

P̃13 
P̃23 P̃33

Introducing K in (4.57), we get


⎛ ⎞
0 0 K1 P̃13
P̄ = K  Π(P̃ )K = − ⎝ 0 0 P̃23 ⎠ . (4.62)
K
P̃13 
P̃23 P̃33
1

Recall that (ii) ΔV ≤ 0 for x ∈ M is true iff P̄ ≤ 0. And this is equivalent,


imposing that the off-diagonal terms are zero, to

−P̃33 ≤ 0, P̃23 = 0, and ∃Mβ : P̃13 = Mβ C̃1 , (4.63)

as Imag K1 = Ker C̃1 , for some column matrix Mβ of appropriate size. Then, using
T  P̃ T = P , we can say that
      
(0, 0, In3 )P = P̃13 , P̃23 T12 , P̃33 = (Mβ C̃1 , 0)T12 , P̃33 , (4.64)

and using (4.48) and P̃33 = P33 , one immediately obtains (4.55) = (4.64), proving
the fact that (4.55) holds iff P̄ ≤ 0 iff ΔV ≤ 0 for x ∈ M , which concludes the
proof. 

Remark 4.2 Notice that when C2 = 0, one can take C = C̃, T = In , and P = P̃ ,
which is the case in [9]. In this way, the result here extends the previous results not
only to the delayed case, but also to a more general form of the resetting hypersur-
face M , with C2 = 0.

Remark 4.3 Usually, the term V3 in (4.52) is expressed by replacing Ax(t + ξ ) +


Ad x(t + ξ − h) by ẋ(t + ξ ) [26]. Here, this is possible only in the continuous mode
of operation. At the reset instants, the relation is no longer valid, so that the correct
expression for V3 is the one in (4.52).

Remark 4.4 Delay margin. If the delay h is not given, it is important to estimate the
set of values h ∈ Ireset for which (4.54) and (4.56) are feasible. Notice that stability
of the reset system is more restrictive than stability of the linear system, so that
Ireset ⊂ Ilin . Determining those h ∈ Ilin for which ẋ = Ax + Ad x(t − h) is stable
168 4 Stability of Time-Delay Reset Control Systems

is well known. It is known that h can have an effect of conditional stability: Ilin =
[0, η1 ) ∪ (η2 , η3 ) ∪ · · · ∪ (ηm−1 , ηm ), with the ηk easily computable [16].
With this is mind, as Ireset is bounded by Ilin , one can always define a fine mesh
of values {hk }N k=1 covering this set and check feasibility of the N related LMIs.
Furthermore, (4.56) can be written as F + hG < 0, and one can always force neg-
ative definiteness by checking Fk + hk Gk < −αk I , for some αk > 0. Then, it is
easy to show that we have a stable 0 range (hk − δk , hk + δk ) =: Ik provided that
δk = αk /Gk . Thus, the union k Ik =: Inum provides a numerically guaranteed
stable range Inum ⊂ Ireset that can be made arbitrarily close to Ireset by increasing
the number of LMIs tested.

4.7 Delay-Dependent Conditions in the Frequency Domain

This section presents frequency domain conditions related to (4.55) and (4.56)
to verify stability. One standard tool for moving from time-domain LMIs to
frequency-domain conditions is the Kalman–Yakubovich–Popov (KYP) lemma
used in Sect. 4.4. The KYP lemma could have been used here as well, however,
we prefer to present here an approach based on passivity, with useful connections
between resetting and control by impulses.
The approach is based on the notion of passivity of a system with respect to
certain storage function (the Lyapunov functional). The IDDE system will be in-
terpreted as equivalent to the base-linear subsystem (BLS) in feedback connection
with the resetting subsystem. It will be shown that the reset subsystem is passive, so
that the global system will be stable if the BLS (with certain inputs and outputs) is
strictly passive. This last condition has an easy frequency domain interpretation.
Thus, to give a frequency-domain interpretation of Proposition 4.7, note that con-
dition (4.55) can be rewritten as B  P = H or

H  = P B, (4.65)

with
B  = (O, O, In3 ), H = (Mβ C1 , Mβ C2 , P33 ). (4.66)
After this, consider the base linear system

(BLS) ẋ(t) = Ax(t) + Ad x(t − h), (4.67)

and the base linear system with inputs and outputs



ẋ(t) = Ax(t) + Ad x(t − h) + Bu(t),
(BLSio ) (4.68)
y(t) = H x(t).

Proposition 4.8 If the conditions (4.55) and (4.56) hold, then the base-linear sys-
tem BLSio is strictly passive with respect to the storage functional V (xt ) (4.49).
4.7 Delay-Dependent Conditions in the Frequency Domain 169

Proof The arguments are sketched as follows. Recall that the system BLSio is
strictly passive with respect to V (xt ), by definition, when

y  (t)u(t) > V̇BLSio (xt ), xt = 0, (4.69)

for all u(t), y(t), xt solutions to the system. By rebuilding the time derivative of the
Lyapunov–Krasovskii functional, it can be seen that
1 1
V̇BLSio = V̇BLS + x (t)P Bu(t) + u (t)B  P x(t), (4.70)
2 2
where the subindex of the V̇ denotes the system where the time derivative of V is
computed. The last expression is easy to deduce and shows that the only influence
of u in BLSio affecting V̇ = V̇1 + V̇2 + V̇3 is through the term V̇1 . The derivatives of
the remaining integral terms V2 , V3 can be written [26] as:

V̇2 = x (t)Qx(t) − x (t − h)Qx(t − h)

and
0
V̇3 = hF (t) − F (t + θ ) dθ,
−h
with
F (t) = (Ax(t) + Ad x(t − h)) A
d XAd (Ax(t) + Ad x(t − h)) .
And hence V̇2 , V̇3 depend only on past values x(t + θ ), θ ≤ 0 of the state and not
on the input u(t), so that (4.70) is true. Then, the passivity condition amounts to

y  (t)u(t) = x (t)H  u(t) > V̇BLS (xt ) + x (t)P Bu(t). (4.71)

Now, the only possibility for the previous relation to hold is (setting u = 0) that
V̇BLS (xt ) < 0 (ensured by (4.56)) and (as u(t) is completely free) that (4.65) is sat-
isfied (equivalent to (4.55)). Notice that (4.56) is only a sufficient condition for
V̇BLS (xt ) < 0, according to [26]. This is the reason why the ‘only if’ direction can-
not be proved here. 

Now, it is useful to show that the overall IDDE (4.5) is passive (and stable) by
proving that it is the negative feedback connection between BLSio and certain suit-
able subsystem R that performs the resetting actions. To show this, consider the
ideal resetting system R : y(t) → −u(t) =: v(t) with input y(t) = H x(t) and out-
put −u(t) = v(t) given by

−1
(R) v(t) = δ(t − tk )P33 y(t), (4.72)
k

where δ(t − tk ) is a Dirac impulse applied at t = tk , the instants when the reset
condition holds. This resetting system has already been used in Sect. 4.4.3. Now, we
will see more formally how R plays an important role also in the delay-dependent
case.
170 4 Stability of Time-Delay Reset Control Systems

Proposition 4.9 The IDDE system (4.5) is equivalent to the negative feedback in-
terconnection between the base linear system with inputs and outputs BLSio and the
ideal resetting subsystem R.

Proof During the continuous mode, both configurations share the same autonomous
dynamics. At the reset times tk , the IDDE system resets to zero the component x3
of the controller state x  
23 = (x2 , x3 ) (the last n3 states). At the same instants tk , the
system R behaves as follows. Notice that

y(tk ) = H x(tk ) = Mβ C1 x1 (tk ) + Mβ C2 x2 (tk ) + P33 x3 (tk ), (4.73)

and, by the reset condition Cx = C1 x1 + C2 x2 = 0,

y(tk ) = P33 x3 (tk ), (4.74)

so from (4.72) u(t) = −v(t) contains, locally around t = tk , a Dirac delta with
weight
−1
u(t) ≈ −δ(t − tk )P33 P33 x3 (tk ) = −δ(t − tk )x3 (tk ).
This is the precise input u(t) that, applied to the BLS with input matrix B, per-
forms the zero reset of x3 , the last n3 components of the state. 

Figure 4.5 illustrates the main idea of the previous proposition. The evolution
of certain reset state x3 (t) is shown in Fig. 4.5 (top). It obeys the equation ẋ3 =
f3 (x(t), x(t − h)) and resets to x3 (tk+ ) = 0 at reset instants given by Cx(tk ) = 0.
Then, this behavior is equivalently
 modeled by ẋ3 = f3 (x(t), x(t − h)) + u(t) with
the feedback law u = − k δ(t − tk )x3 (t) with corrective ideal Dirac impulses.
Indeed, this effect can be approximated
 by rectangular pulses δε (t) with width ε
and height 1/ε in the form u = − k δε (t − tk )x3 (tk ). Figure 4.5 (bottom) shows the
required train of finite pulses, for ε = 0.1. These ideas are exploited and formalized
in the following.
Now the question is: Could the stability of the IDDE system be proved with pas-
sivity techniques? From Proposition 4.7 (internal stability), one can always derive
an input–output frequency domain result, using the KYP lemma, as in Sect. 4.4.
However, it would be useful to find an independent approach based only on passiv-
ity. The main problem on the I/O stability of the loop {BLSio , −R} is that the Dirac
impulses are not signals in L2 . Then, the only possibility is to consider L2 approx-
imations of ideal impulses. So, consider the system Rε , y(t) → −u(t) =: v(t), with
input y(t) = H x(t) and output −u(t) = v(t) given by

−1
(Rε ) v(t) = δε (t − tk )P33 |y(tk )| ◦ sign(y(t)), (4.75)
k

where δε (t − tk ) is a rectangular pulse starting at t = tk , with width ε and height


1/ε, and “◦” stands for componentwise vector product.
Thus Rε acts by producing pulse trains with height proportional to the value
|y(tk )| of its inputs at the reset instants. The function ‘sign’ is included for passivity
4.7 Delay-Dependent Conditions in the Frequency Domain 171

 4.5 (Top) A reset state x3 (t). (Bottom) Equivalent resetting impulses u =


Fig.
− k δε (t − tk )x3 (tk ) (ε = 0.1)

because y(t) may vary fast enough in the interval [tk , tk + ε), changing its sign.
Connecting Rε to BLSio , from (4.66) and (4.68), and for t ∈ [tk , tk + ε), we have

ẏ(t) = H ẋ(t) = H Ax(t) + H Ad x(t − h) +H Bu(t),


 

=:ẏX (t)

thus, from H B = P33 and u(t) = −v(t) in (4.75), for t ∈ [tk , tk + ε):

|y(tk )|
ẏ(t) = H Ax(t) + H Ad x(t − h) − ◦ sign(y(t)). (4.76)
 
ε
=:ẏX (t)

To better see how Rε works, consider the scalar case with y(tk ) =: yk > 0, then
yk
ẏ(t) = ẏX (t) − sign(y(t)). (4.77)
ε
So, if ε " min(yk )/ max(|ẏX |), then the second term in (4.77) dominates the first
one, which is the objective. Thus, we make here two assumptions:
172 4 Stability of Time-Delay Reset Control Systems

(i) the state x(t) is restricted to evolve in {x : x(t) < Xmax }, and
(ii) the resetting is applied only when |y(tk )| > ymin .
For the first assumption, the bound Xmax can be taken arbitrarily large, covering
the physically meaningful values of the states. This assumption also implies that the
stability results will have a local character, and ensures that ẏX (t) is bounded.
For the second assumption, ymin can be taken arbitrarily small, equal to the res-
olution level of the quantization in digital implementation. Both assumptions are
necessary, otherwise the dominance of yk /ε and the resetting effect will be lost. If
assumptions (i) and (ii) hold, then (4.77) for yk > 0 and ε → 0 approaches
yk
ẏ(t) ≈ − for t ∈ [tk , tk + ε).
ε
Integrating this equation from y(tk ) = yk gives rise to y(tk + ε) = 0, that is, we
reset the output y(t) at the end of the pulse width. Since we are arbitrarily close
to the reset instant t ∈ [tk , tk + ε), with ε → 0, from (4.73) and (4.74) it follows
that y(t) ≈ P33 x3 (t) for t ∈ [tk , tk + ε). Thus we are resetting as well the required
part of the state x3 at the end of the pulse: x3 (tk + ε) = 0 when ε → 0. In this
way, Rε approaches the ideal reset R (4.72) when ε → 0, and the resetting action
in Proposition 4.9 is correctly applied because from (4.77) when ε → 0 the value
yk = P33 x3 (tk ) decreases linearly to y(tk + ε) = 0 in an arbitrarily fast way.
Now, the system Rε given by (4.75) maps L2e to L2e . However, Rε is not
bounded in L2 since it is possible to find signals in L2 such that their images
under Rε are not in L2 . A similar problem has been discussed in [11] in the frame-
work of sampled-data systems. The solution adopted there to overcome the prob-
lem will be also used here. Basically, the system Rε is restricted to operate in the
space F L2 of filtered L2 signals, F being an LTI system with a strictly proper
transfer function and arbitrarily large bandwidth. More precisely, Rε is restricted to
operate on F L2 = {y ∈ F L2 : |ẏ| ≤ M} where the bound M on ẏ can always
be imposed, for any fixed ε, from the discussion after (4.77). Thus, the system
Rε : F L2e ⊂ L → L is bounded, that is, R (F L  ) ⊂ L . This is a necessary
2e 2e ε 2 2
condition for Rε to be output strictly passive.

Proposition 4.10 The system Rε : F L2e  ⊂ L → L


2e 2e (that approximates the
ideal reset R by finite rectangular pulses, and approaches R when ε → 0) is an
output strictly passive (OSP) system for all ε > 0.

Proof Recall that a system Rε : y → v, satisfying


T T T
y  (t)v(t) dt ≥ β + δi y  (t)y(t) dt + δo v  (t)v(t) dt (4.78)
0 0 0

for all y, v, T and some β ≤ 0, is passive when δi , δo ≥ 0, input strictly passive


(ISP) when δi > 0, output strictly passive (OSP) when δo > 0, and strictly passive
(SP) when δi , δo > 0 (for details, see Sect. 3.3).
4.7 Delay-Dependent Conditions in the Frequency Domain 173

Let us suppose here, for simplicity, the scalar case n3 = 1 and P33 = 1. The
multivariable case requires a more careful treatment and will be reported elsewhere.
First, the L2 norm of the output is
 2
Rε y =
2
δε (t − tk )|y(tk )| sign(y(t)) dt,
k

and after some manipulation,



Rε y2 = (1/ε) |y(tk )|2 .
k

Then, the inner product %v, u& = %Rε y, y& is, by definition,
 
%Rε y, y& = δε (t − tk )|y(tk )| sign(y(t)) y(t) dt,
k

where after some work (using |y| = sign(y)y)


 tk +ε
%Rε y, y& = (1/ε) |y(tk )| |y(t)| dt.
k tk

Now, since y is a filtered L2 signal, that is, y ∈ F L2e  , |ẏ| is bounded, and one

gets
tk +ε
|y(t)| dt ≥ εα(ε)|y(tk )|,
tk

where the coefficient α(ε) > 0 can be obtained from (4.77) by taking the worst
case evolution of y(t) (least integral value |y|) as a function of the corresponding
bounds |yk | ≥ ymin and |ẏX | ≤ (H A + H Ad )Xmax .
It becomes also clear that when ε → 0 the evolution for (t, y(t)) in (4.77) is a
straight line from (tk , yk ) to (tk + ε, 0) so that |y| = εyk /2, meaning that α(ε) →
1/2 when ε → 0. Finally, comparing < Rε y, y > and Rε y2 ,

%Rε y, y& ≥ α |y(tk )|2 = αεRε y2 ≥ δRε y2 ,
k

and thus, from (4.78), Rε is output strictly passive (OSP) for some δ such that 0 <
δ ≤ αε. 

Remark 4.5 An important consequence of the last result is that the stability condi-
tion for the feedback system {BLSio , −Rε } can be translated into a frequency con-
dition on BLSio . To see this, recall that a negative feedback system {G1 , −G2 } is
finite-gain stable if any of (i)–(iii) holds [10]: (i) G1 , G2 are ISP, (ii) G1 , G2 are
OSP, (iii) G1 is SP and G2 is passive (see also Sect. 3.3 for details).
174 4 Stability of Time-Delay Reset Control Systems

Fig. 4.6 Robust stability


interpretation described in
Remark 4.6

Clearly, we are interested in case (ii), so if BLSio is OSP, then closed-loop stabil-
ity (finite-gain) is guaranteed. Recall also that a stable and controllable LTI system
G(s) is OSP if there exists some δo > 0 such that G(j ω) + G (j ω) ≥ δo G(j ω)2
(see [10]). In the scalar case, this means that G(j ω) lies inside a circle with center
s = 1/(2δo ) and radius r = 1/(2δo ). In summary, the required stability condition is
Hβ (j ω) + Hβ (j ω) ≥ δHβ (j ω)2 , δ > 0, that is, an OSP condition on
 −1
Hβ (j ω) = H j ωI − A − Ad e−j ωh B. (4.79)

So this is a generalization of the results in [9]. In the particular case when Ad → 0


(without time-delay), and when H = (Mβ C1 , Mβ C2 , P33 ) → (Mβ C1 , O, P33 ), one
gets the so-called Hβ -SPR (positive realness) condition in [9]. Strictly speaking,
this passivity analysis ensures stability of the loop {Hβ (s), R} in the I/O sense. In
general, recovering internal stability would require some mild assumptions (control-
lability, detectability, etc.) However, note that the passivity result here is presented
as an extra interpretation of an original LK analysis which directly ensures internal
asymptotic stability.

Remark 4.6 (Robustness in terms of small-gain and positive uncertainties) A sim-


ilar interpretation of stability to the one given for the delay-independent case in
Sect. 4.4.3 can be given now. One nice feature of the delay-dependent LK func-
tional (4.49) is that it allows an interpretation within a scaled small-gain framework
[26]. In the same way, the previous passivity analysis can be recast in a robust sta-
bility setting. Combining the two parts, it can be said that the IDDE system (4.13)
satisfying the conditions of Proposition 4.7 can be interpreted as shown in Fig. 4.6
(compare with Fig. 4.2).
4.8 Example: Delay-Dependent Stability 175

This figure contains two sets of blocks: the white fixed blocks and the gray uncer-
tain blocks. The set of fixed blocks is derived from the base linear system and from
some matrices appearing in the LMI stability conditions, including B, H in (4.66).
There are three uncertain blocks, the block Δ is the delay system e−hs . The block Z
is the zero-order holder system (e−hs − 1)/(hs). The third block R is the impulsive
resetting subsystem introduced above.
The proof of the equivalence is omitted for brevity, but it can be checked starting
from (13) in [29], and with some mild notation changes, as M = I + X −1 W  , with
X = 12 P . The interesting point of the relation between (4.13) and Fig. 4.6 is not only
the equivalence when Z, Δ and R are as described above, but also the important
conclusion that stability is preserved when the first two blocks are replaced by any
pair of small-gain blocks (Δ1 , Δ2 ) satisfying Δi ∞ ≤ 1, i = 1, 2. And the third
bock R can be replaced as well by any other OSP block with a stable strictly proper
transfer function Δ3 with Re[Δ3 (iω − ε)] ≥ 0 for some ε > 0.
Thus, Fig. 4.6 shows explicitly the degree of robustness achieved by the LK
stability conditions, that is, which amount of uncertainty, and in what form, can
be tolerated. This offers useful insights into theoretical and practical aspects of the
problem.

Remark 4.7 (Using alternative LK functionals and LMIs) Within the field of lin-
ear time-delay stability there is a large literature, and thus, there are many other
LK functionals alternative to (4.49). In [4], a functional taken from [16] was used.
Similarly, many other LK functionals could be combined to the passivity analysis
presented here, although it is not always possible to derive robust interpretations as
in the previous remark.

4.8 Example: Delay-Dependent Stability

Consider a feedback system according to the setup of Fig. 4.1. Here the plant has a
transfer function e−hs P0 (s) given by (the example is adapted from [8])

s +1
e−hs P0 (s) = e−hs ,
s(s + 0.2)
and the feedback compensator R is a FORE compensator with base LTI system
given by
1
FOREb (s) = .
s +1
It can be shown that the base LTI control system is stable for delays h < 0.2. In
the following, the passivity result derived in Sect. 4.7 will be used to show that the
reset control system is also stable for that range of delays. Thus, the reset action does
not destabilize the control system. This is simply made by checking if the system
given by (4.79) is OSP. A delay h = 0.15 will be chosen.
176 4 Stability of Time-Delay Reset Control Systems

Fig. 4.7 minω Re{(H (j ω))}


vs β for h = 0.15

Fig. 4.8 Re{(H (j ω))} vs ω


for h = 0.15 and β = 0.4

In this simple case, the plant has two states and the reset compensator only
one state, thus n1 = 2, n2 = 0, and n3 = 1. Using the realizations of P0 (s) and
FOREb (s):
   
−0.2 0 1  
Ap = , Bp = , Cp = 1 1 ,
1 0 0

and Ar = −1, Br = 1, Cr = 1, Dr = 0; the matrices A, Ad are


⎛ ⎞ ⎛ ⎞
−0.2 0 0 0 0 1
A=⎝ 1 0 0 ⎠, Ad = ⎝ 0 0 0⎠.
−1 −1 −1 0 0 0
4.8 Example: Delay-Dependent Stability 177

Fig. 4.9 Time response y(t) (top) and u(t) (bottom) with reset (solid) and without reset (dotted)

In addition, here Mβ = (β), P33 are scalar, in particular P33 = 1 may be chosen
without lost of generality. As a result, matrices H and B are given by (4.66):
     
H = βCp P33 = β β 1 , B = 0 0 1 .

With the above values, system (4.79) is checked to be OSP. A simple exploration
(see Fig. 4.7) shows that in fact it is OSP for values of β between 0.29 and 0.56.
Figure 4.7 shows the minimum value of Re{H (j ω)} over the ω-axis for different
values of β. In particular, Fig. 4.8 shows the positive value of Re(H (j ω)) (logarith-
mic scale) vs ω for β = 0.4.
As a result, the reset control system is guaranteed to be also stable, in this case
for h = 0.15. An identical procedure could be used to show that the reset action
never destabilizes a stable base LTI control system, and its stability limit is h = 0.2,
the same limit that the one for the base LTI control system. Figure 4.9 plots the
step response, showing the plant output y(t) and plant input u(t). Notice that the
reset instants appear actually when the error is zero, r(t) − y(t) = 0, but the cor-
rective effects (discontinuities in u(t) and derivative discontinuities in y(t)) appear
in the plant h = 0.15 later, due to the delay between the controller and the plant
(see Fig. 4.1). Notice also the damping provided by reset that improves the trade-off
between fast rise time and small overshoot.
178 4 Stability of Time-Delay Reset Control Systems

An important question is whether the reset control system might be stable even
in the case in which the base LTI system is unstable, in this case for h > 0.2. The
proposed method cannot answer this question because the stability of the base LTI
system is a necessary condition. A way to overcome this problem could be to de-
velop stability conditions based on the reset-times, as in Sect. 3.2.2. But those ideas
should be extended to time-delay, infinite-dimensional reset systems. This is a more
involved problem that requires further work.
Notice also from the time response in Fig. 4.9 that if we want to avoid the delay
of h in the reset corrective effect on the plant output y(t), then we could arrange
it by anticipating the reset actions in some way. This anticipative reset can be im-
plemented, for example, by using the concept of reset band introduced in Sect. 3.4.
Reset band is proposed in the next chapter as a design improvement and applied in
different cases as reported in Chap. 6.

References
1. Åström, K.J.: Limitations on control system performance. Eur. J. Control 6, 2–20 (2000)
2. Bainov, D.D., Simeonov, P.S.: Systems with Impulse Effect: Stability, Theory and Applica-
tions. Ellis Horwood, Chichester (1989)
3. Baños, A., Barreiro, A.: Delay-independent stability of reset systems. In: 32nd IECON Con-
ference, Paris, France (2006)
4. Baños, A., Barreiro, A.: Delay-dependent stability of reset control systems. In: Proceedings
of the American Control Conference. ACC, New York (2007)
5. Baños, A., Barreiro, A.: Delay-independent stability of reset systems. IEEE Trans. Autom.
Control 54(2), 341–346 (2009)
6. Baños, A., Carrasco, J., Barreiro, A.: Reset-times dependent stability of reset control with
unstable base systems. In: Proceedings of the IEEE Int. Symp. on Industrial Electronics (ISIE),
Vigo, Spain (2007)
7. Barreiro, A., Baños, A.: Delay-dependent stability of reset systems. Automatica 46, 216–221
(2010)
8. Beker, O.: Analysis of reset control systems. Ph.D. Thesis, University of Massachusetts,
Amherst (2001)
9. Beker, O., Hollot, C.V., Chait, Y., Han, H.: Fundamental properties of reset control systems.
Automatica 40, 905–915 (2004)
10. Carrasco, J., Baños, A., van der Schaft, A.: A passivity based approach to reset control systems
stability. Syst. Control Lett. 59, 18–24 (2010)
11. Chen, T., Francis, B.A.: Input-output stability of sampled data systems. IEEE Trans. Autom.
Control 36, 1 (1991)
12. Chen, J., Latchman, H.A.: Frequency sweeping tests for stability independent of the delay.
IEEE Trans. Autom. Control 40, 1640–1645 (1995)
13. Clegg, J.C.: A nonlinear integrator for servomechanism. Trans. AIEE, Part II 77, 41–42 (1958)
14. De la Sen, M., Luo, N.: A note on the stability of linear time-delay systems with impulsive
inputs. IEEE Trans. Circuits Syst. 50(1), 149–152 (2003)
15. Horowitz, I.M., Rosenbaum, P.: Nonlinear design for cost of feedback reduction in systems
with large parameter uncertainty. Int. J. Control 24(6), 977–1001 (1975)
16. Gu, K., Kharitonov, V.L., Chen, J.: Stability of Time-Delay Systems. Birkhäuser, Boston
(2003)
17. Guo, Y., Wang, Y., Zheng, J., Xie, L.: Stability analysis, design and application of reset control
systems. In: Proceedings of the IEEE Int. Conf. on Control and Automation, Guangzhou,
China (2007)
References 179

18. Haddad, W.M., Nersesov, S.G., Chellaboina, V.S.: Energy-based control for hybrid port-
controlled Hamiltonian systems. Automatica 39, 1425–1435 (2003)
19. Khalil, H.J.: Nonlinear Systems, 2nd edn. Prentice Hall, Upper Saddle River (1996)
20. Krishman, K.R., Horowitz, I.M.: Synthesis of a nonlinear feedback system with significant
plant-ignorance for prescribed system tolerances. Int. J. Control 19(4), 689–706 (1974)
21. Lakshmikantham, V., Bainov, D.D., Simeonov, P.S.: Theory of Impulsive Differential Equa-
tions. World Scientific, Singapore (1989)
22. Liu, X., Ballinger, G.: Uniform asymptotic stability of impulsive delay differential equations.
Comput. Math. Appl. 41, 903–915 (2001)
23. Liu, X., Wang, Q.: Stability of nontrivial solution of delay differential equations with state-
dependent impulses. Appl. Math. Comput. 174, 271–288 (2006)
24. Liu, X., Shen, X., Zhang, Y., Wang, Q.: Stability criteria for impulsive systems with time delay
and unstable system matrices. IEEE Trans. Circuits Syst. 54, 10 (2007)
25. Nešić, D., Zaccarian, L., Teel, A.R.: Stability properties of reset systems. Automatica 44, 8
(2008)
26. Park, P.: A delay-dependent stability criterion for systems with uncertain time-invariant delays.
IEEE Trans. Autom. Control 44, 876–877 (1999)
27. Rantzer, A.: On the Kalman–Yakubovich–Popov lemma. Syst. Control Lett. 28, 7–10 (1996)
28. Yang, T.: Impulsive Control Theory. Lecture Notes in Control and Information Sciences,
vol. 272. Springer, Berlin (2001)
29. Zhang, J., Knospe, C.R., Tsiotras, P.: Stability of time-delay systems: equivalence between
Lyapunov and scaled small-gain conditions. IEEE Trans. Autom. Control 46, 482–486 (2001)
Chapter 5
Design of Reset Control Systems

5.1 The PI + CI Compensator

It is well known that PID (proportional-integral-derivative) compensators are by far


the most applied form of feedback in use. As it is noted in [1], more than 90% of
all control loops are PID, in fact, most of them are PI since the derivative action in
not used very often. It is generally acknowledged that this predominance of PI/PID
compensation is due to its simplicity (three parameters to tune), and its effectiveness
in a wide range of applications: motor drives, automotive industry, flight control,
process control, etc.
In this chapter, a new type of reset compensator will be defined using as base
system a PI compensator. The idea will be to introduce a new nonlinear/hybrid PI
compensator, that will be referred to as PI + CI compensator, having in mind both
the simplicity of tuning (only a new parameter, the reset percentage, is added), and
the effectiveness from the application point of view. It will be mandatory to obtain
simple tuning rules that can improve the performance of PI + CI over PI compensa-
tion. Thus, one basic goal is to analyze the potentials of PI + CI compensation, and
identify the class of systems in which its performance is superior compared with a
well-tuned PI compensator. An important point is that although, in principle, using
the derivative term in a PID compensator may offer a better performance than a
PI + CI compensator, this has an important price in terms of cost of feedback, even
if the D term is filtered. This is the reason why the PI + CI compensator is compared
with a PI compensator, they both have a similar cost of feedback and the PI + CI
compensator, being a nonlinear/hybrid compensator, can potentially overcome PI
compensator fundamental limitations.
From the applications side, a basic goal will be to develop PI + CI tuning rules
for basic plants including first order plus deadtime (FOPDT) systems, second and
higher order plus deadtime systems, and systems with integrators. Related work
about experimental demonstrations of reset control has been developed recently. In
[22], a FORE compensator is designed and demonstrated for a tape-drive system,
exploiting the advantages of reset control to overcome fundamental limitations of
the LTI base system. The works [2] and [20] use some form of PI/PID compensation

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 181


DOI 10.1007/978-1-4471-2250-0_5, © Springer-Verlag London Limited 2012
182 5 Design of Reset Control Systems

Fig. 5.1 PI + CI controller structure

in the base system that, in contrast to PI + CI compensation, performs a full reset of


the compensator integrator term. In addition, in [23], a type of reset compensation
with reset actions at fixed instants has been developed.
The PI + CI compensator is defined by means of a base PI compensator, that
will be referred to as PIbase , in which the integral term is partly reset. Then, the
starting point is a PIbase compensator that has a transfer function depending on its
proportional gain, kp , and its integral time constant, τi , given by
 
1
PIbase (s) = kp 1 + . (5.1)
τi s
In the state space, a realization of the PIbase compensator is

⎨ ẋ(t) = e(t),
(PIbase ) (5.2)
⎩ v(t) = kp x(t) + kp e(t).
τi

Finally, the PI + CI compensator is defined by adding in parallel a Clegg inte-


grator (CI) to a PI controller (see Fig. 5.1). As a result, the PI + CI compensator
will have three terms: a proportional term P , an integral term I, and a Clegg inte-
grator (CI) term. Note that a CI term by itself does not have the characteristic of
eliminating the steady-state error in response to step disturbances by itself, thus the
integral term I is used for that purpose. The main benefit of using the CI term will
be to improve the transient response of the system, being possible in principle to get
a better response without excessive overshooting in comparison with its base PIbase .
As it will be seen later, if a well-tuned PIbase does not have a closed-loop response
with enough overshoot then the compensator PIbase may be detuned to obtain a
faster response, and then reset is performed to improve settling time and reduce
overshooting. In addition, it is expected that a PI + CI compensator will obtain better
performance in comparison with a well-tuned PI compensator in some cases. It may
be argued that the comparison should be also done with a PID compensator, but it
should be noted that in general a PID compensation would have a significatively
higher cost of feedback (sensitivity with respect to sensor noise) even in the case
5.1 The PI + CI Compensator 183

in which the term D is filtered, thus for a fair comparison only a PI compensator is
considered.
The block diagram structure of a PI + CI controller is shown in Fig. 5.1, where
the input e is the error signal, the output v is the control signal, kp is the propor-
tional gain, and τi is the integral time. The new parameter preset , the reset ratio,
is a dimensionless constant with values preset ∈ [0, 1] that accounts for the relative
weight of the CI term over the I term. Note that a PI + CI compensator is reduced
to a PI compensator if reset is eliminated, that is, preset = 0. And on the other hand,
a P + CI compensator would be obtained if preset = 1. However, as it was discussed
above, in general the reset should not be applied on the whole of the integral term
because the fundamental asymptotic property of the integral term would be lost, for
example, the steady-state error in the closed-loop step response could not disappear.
In the state space, a PI + CI compensator can be expressed by using two states,
one corresponding to the I term and the other corresponding to the CI term. Thus
the state will be xr = (xi , xir ) , where xi will correspond to the I-term state and xir
to the CI-term state. A PI + CI state space model is given simply by (4.2), where the
state space matrices are given by
 
0 0
Ar = , (5.3)
0 0
 
1
Br = , (5.4)
1
kp  
Cr = 1 − preset preset , (5.5)
τi
Dr = kp , (5.6)
and the matrix Aρ ∈ R2×2 selects the state to be reset, xir , and is given by
 
1 0
Aρ = . (5.7)
0 0

Alternatively, a (more general) PI + CI compensator can be realized by using one


state, but then the reset matrix Aρ , in this case is a scalar, will be a nonzero number
α, and then the matrix AR will be a diagonal matrix with ones corresponding to
the plant states and α = 0 corresponding to the PI + CI state. A realization of the
PI + CI with one state is:

⎪ ẋpi+ci (t) = e(t) if e(t) = 0,


xpi+ci (t + ) = (1 − rk )xpi+ci (t) if e(t) = 0, (5.8)


⎩ kp
v(t) = τi xpi+ci (t) + kp e(t),

where the previous model is obtained in the particular case of


preset xir (tk )
rk = ,
(1 − preset )xi (tk ) + preset xir (tk )
184 5 Design of Reset Control Systems

with tk , k = 1, 2, . . . , being the reset instants, and xi and xir the states of the integra-
tor and the Clegg integrator, respectively, in the two states model. In general, several
different values of rk may be used. As a result, this representation gives a particular
type of reset control system with variable reset that need a specific treatment ([17]
is an initial work in this direction).
On the other hand, as it is well known, the describing function gives good results
in practice if some low-pass filtering is assumed in the plant. A very important char-
acteristic of the CI term is that its describing function, previously given by (3.89),
does not depend on the input amplitude, only on the frequency. This property is also
inherited by the PI + CI compensator whose describing function is simply given by
 
1 − preset preset 1.62 −j 38.1◦
(PI + CI)(j ω) = kp 1 + + e
j ωτi τi ω
 
j (ωτi + π4 preset ) + 1
= kp . (5.9)
j ωτi

Note that the PI + CI describing function reduces to the PI frequency response


function whenever reset is eliminated, that is, preset = 0. In Fig. 5.2, the describing
function of a PI + CI compensator is shown for several values of the reset ratio
preset . Note that the reset action may introduce a phase lead of up to 50 degrees and
a relatively small increment of the magnitude for frequencies lower than 1/τi . This
is a basic property of the PI + CI compensator that, in comparison with its base
PI compensator, allows achieving both a bigger phase margin and a crossover gain
frequency. This means that a better performance both in terms of speed of response
and relative stability may be potentially obtained by means of PI + CI compensation,
overcoming fundamental limitations of PI compensation.
In addition, the simple form of the describing function makes it very appropriate
from a design perspective, and in fact it will serve as a basis of the tuning rules to
be developed in next sections.

5.1.1 PI + CI Tuning for First Order Plants

The simplest plant case is considered first, a first order plant, to elaborate a PI + CI
design procedure based on easy-to-apply tuning rules. The main goal is to exploit
the reset action by using the parameter preset , to obtain a significant improvement in
terms of closed-loop performance. Basically, a PI compensator is tuned in the first
place, and then a proper value of preset is chosen to obtain the desired performance
specifications. Obviously, the design of the PI compensator has to be done in such a
way that its performance can be improved by partly resetting its state. As it will be
seen below, a fast response with significant overshoot will be usually chosen for the
base PI compensation. Afterwards, CI will be used to reduce overshooting without
sacrificing speed of response.
5.1 The PI + CI Compensator 185

Fig. 5.2 PI + CI describing function for several values of preset (kp = τi = 1)

Consider the standard feedback reset system of Fig. 2.1, where the reset con-
troller R is a PI + CI compensator, and now the plant P is described by

ẋp (t) = − τ1 xp (t) + τk u(t), xp (t0 ) = xp0 ,
(P ) (5.10)
y(t) = xp (t),

where xp is the plant state and xp0 the plant initial state. In this case, the time
constant τ and the gain k are scalars. After some simple computation, the closed-
loop system (2.4) is given by the matrices
⎛ 1+kp k kp k kp k ⎞
− τ τ τi (1 − preset ) τ τi preset
A = ⎝ −1 0 0 ⎠, (5.11)
−1 0 0
 
C= 100 , (5.12)
⎛ ⎞
1 0 0
AR = ⎝0 1 0⎠ . (5.13)
0 0 0
186 5 Design of Reset Control Systems

5.1.1.1 Autonomous Closed-Loop System

Using (2.4), the time evolution of the autonomous closed-loop state is given simply
by
x(t) = eA(t−tk ) x(tk ) (5.14)
for t ∈ (tk , tk+1 ], where the state is now x = (xp , xi , xir ) , and A is given by (5.11).
In addition, the closed-loop output y is given by y = Cx = (1, 0, 0)x = xp , and by
definition after each reset time both the plant state xp (t) and the reset integral state
xir (t) are zero. Thus, taking a state after reset x = (0, xik , 0) at t = tk+ , it can be
shown after some computation that the output y(t) in the interval (tk , tk+1 ] is given
by
α 2 + β 2 −α(t−tk )
y(t) = e sin(β(t − tk ))(1 − preset )xik , (5.15)
β
where −α ± jβ are the nonzero eigenvalues of the matrix A (4.2), with α and β
being respectively
1 + kp k
α= (5.16)

and
. 
 kp k 
β = + α 2 − . (5.17)
τ τi 
k k
Note that α 2 < τpτi is assumed, otherwise the output y would not be oscillatory.
Using (5.15), the reset instants tk+1 , k = 1, 2, . . . , are given by

α 2 + β 2 −α(tk+1 −tk )
0 = y(tk+1 ) = e sin(β(tk+1 − tk ))(1 − preset )xik , (5.18)
β

and then it is not difficult to show that reset instants {t1 , t2 , . . . } are periodic with
fundamental period Δreset given by
π
Δreset = . (5.19)
β

On the other hand, the integral term state, xi (t), is not necessarily zero after the
reset instants. Its time evolution for t ∈ (tk , tk+1 ], k = 1, 2, . . . , is given by
 
1 − preset −α(t−tk )     
xi (t) = preset + e α sin β(t − tk ) + β cos β(t − tk ) xik ,
β
(5.20)
and thus the values of xi (t) at the reset instants are simply given by
 − απ 
xi,k+1 = x(tk+1 ) = preset − (1 − preset )e β xik . (5.21)
5.1 The PI + CI Compensator 187

5.1.1.2 Closed-Loop Stability

As it has been shown above, the closed-loop system consisting of a PI + CI compen-


sator and a first order plant performs reset actions periodically, with a period given
by (5.19). This fact considerably simplifies the closed-loop stability analysis since
it can be shown that a necessary and sufficient condition for internal stability of the
closed-loop system (see [5] and [21]) is simply that the matrix AR eAΔreset be Schur
stable, that is, that its eigenvalues be strictly inside the unit circle. In this case, the
matrix AR eAΔreset can be computed in closed form with some relative effort (details
are omitted here for brevity). The result is:
⎛ ⎞
− απ
−e β 0 0
⎜ − απ ⎟
AR eAΔreset = ⎝ 0 1 − (1 − preset )(1 + e β ) ∗ ⎠ , (5.22)
0 0 0

where “∗” stands for a nonzero term. Note that the nonzero eigenvalues of
− απ
AR eAΔreset are λ1 = −e β and λ2 = 1 − (1 − preset )(1 − λ1 ). Thus, AR eAΔreset
is Schur if and only if the nonzero eigenvalues of A are strictly in the left half-plane
(and thus a minimal realization of the base system is Hurwitz) independently of the
reset ratio preset , since |λ1 | < 1 if and only if α > 0, and λ2 ∈ [λ1 , 1) for values
preset ∈ [0, 1).
Thus, the closed-loop reset system is stable if and only if the closed-loop base
linear system is stable, which is a very convenient result. Unfortunately, this result
is no longer valid for high order plants and plants with deadtime, and some of the
stability results given in Chaps. 3 and 4 could be used.

5.1.1.3 PI + CI Tuning Rules for First Order Plants

Following the goal of designing a PI + CI controller to improve the performance of


a previously designed PI controller, the additional degree of freedom given by the
parameter preset of the Clegg integrator will be used to reduce the overshooting of
the response. In this simple case, overshooting can be eliminated after the second
reset time if the integral state is forced to be zero at t2 , the second reset time, reach-
ing the closed-loop system steady state at that instant. Therefore, making xi (t2 ) = 0
in (5.21), the result is
− απ
e β
preset = − απ
, (5.23)
1+e β

which will relate the reset ratio, preset , with the integral time constant of the PI
controller, τi , and its proportional gain, kp , through the parameters α and β. Note
that in general the integral term state cannot be forced to be zero at the first reset
instant t1 since (5.21) is only valid when xik , k = 1, 2, . . . , is an after-reset state.
It would be only possible if the initial state were an after-reset state, but this is a
188 5 Design of Reset Control Systems

trivial case since it corresponds to a zero initial state, and thus the (autonomous)
reset system state would not depart from the origin.
For a closed-loop step response that reaches the steady state at the second cross-
ing, the tuning rules of the PI + CI controller for a first order plant can be summa-
rized as:
1. Design the PIbase compensator to obtain kp and τi for a desired oscillatory system
1+kk kk
response (( 2τ p )2 < τ τpi ) by using any tuning rules.
2. Compute α and β by using (5.16) and (5.17).
3. Finally, calculate the reset ratio, preset , by using (5.23).

5.1.1.4 Example

Consider the following plant with a unit step as a reference signal:



ẋp (t) = −0.5xp (t) + 1.5u(t),
(P ) (5.24)
y(t) = xp (t).

The following tuning rules are applied:


1. The chosen constant values for PIbase are kp = 2 and τi = 0.15. This gives a fast
response with significant overshoot (see Fig. 5.3).
2. From (5.16) and (5.17), α = 1.75 and β = 4.12.
3. Finally, by using (5.23), the reset ratio is preset = 0.21 in order to reach the
steady-state just after the second reset time. As a result, overshoot and under-
shoot of the PI response are decreased or even eliminated, without sacrificing the
speed of response.
Figure 5.3 shows step responses and control signals for both compensators PIbase
and PI + CI, respectively. Note how responses are identical up to the first reset in-
stant, and that the undershooting/overshooting is eliminated by PI + CI compensa-
tion after the second reset instant; in fact, the steady-state is reached at that instant.
This is impossible to achieve by LTI compensation.
In addition, it can be seen how this improvement can be done by simply resetting
the control signal twice (with a slight reset of the control signal, the reset ratio is
21%). Finally, note that not only is the settling time considerably reduced, but also
the first overshoot is decreased.

5.1.2 First Order Plus Deadtime (FOPDT) Systems

The focus of this section is on the use of PI + CI compensation for first order plus
deadtime systems. In this case, the presence of delays makes it impossible to obtain
explicit solutions to state space equations. Thus, a state-space approach similar to
that in the above section is problematical.
5.1 The PI + CI Compensator

Fig. 5.3 System responses and control signals for PI and PI + CI controllers
189
190 5 Design of Reset Control Systems

Internal model control (IMC) will be used to tune the PIbase compensator. The
IMC method consists of considering a known model of the system in order to get
a desired closed-loop response. For that purpose, the compensator parameters will
depend on the system parameters, and on the time constant of the desired closed-
loop response, λ. Specifically, when the system is modeled as an FOPDT system

k
P (s) = e−hs , (5.25)
τs + 1
the PI compensator is tuned by using the following rules (see [15]):
τ
kp = , (5.26)
k(λ + h)

τi = τ. (5.27)
The value of the desired closed-loop time constant λ can be chosen freely, but
from (5.27) it has to satisfy −h < λ < ∞ to get a positive and nonzero controller
gain. In general, the optimal value of λ is determined by a trade-off between:
1. Fast response and good disturbance rejection (small value of λ), and
2. Stability and robustness (large value of λ).
Among the different ways of choosing the value of λ that have been studied in
the literature (see [12–14]), the one given by S. Skogestad [14] will be used in the
following. Thus, the value λ is chosen to be equal to the time delay λ = h. This
gives a reasonably fast response with good robustness margins. In this way, the PI
proportional gain (5.26) is given by
τ
kp = . (5.28)
2kh
In the following, this method will be used as a basis to tune the PI + CI com-
pensator, but the tuning procedure will depend on whether the delay is dominant or
not.

5.1.2.1 Lag Dominant Systems

A lag dominant system is a system in which its time delay is much smaller than its
time constant, τ h, and thus time delay is not dominant over the system response.
In this case, tuning the integral time τi by using (5.27) gives a base compensator
with a small integral term, and thus it can be expected that the reset action does
not improve the linear response. On the other hand, this choice results in a long
settling time for input load disturbances. To improve the load disturbance response,
a modification of the IMC method proposed by Skogestad will be used here, where
τi is tuned by the rule [14]:
τi = 8h. (5.29)
5.1 The PI + CI Compensator 191

Table 5.1 Performance indexes


Reference Disturbance Stability margins
IAE (s) ITAE (s2 ) IAE (s) ITAE (s2 ) ϕm (°) Am (dB)

PI-IMC 2.20 5.05 5.85 237.45 61.3 9.92


PI-SIMC 3.09 14.07 2.33 67.2 52.5 9.64
PI + CI-SIMC 2.30 6.47 2.33 67.2 51 9.19

This expression, together with (5.28), is used in the Skogestad modification of


internal model control method (it will be referred to as SIMC) to tune a PI controller.
This rule produces a higher value of the integral term, and thus it may be expected
that reset can have a bigger influence.
As an example, consider the following system:
1
P (s) = e−s . (5.30)
20s + 1
Direct application of (5.26) and (5.27) gives a PI-IMC compensator with parameters
kp = 10 and τi = 20 s. On the other hand, by applying the SIMC method, a PI-
SIMC compensator is obtained with the same proportional gain, kp = 10, but with
a smaller integral time constant, τi = 8 s. Finally, these parameters will be used to
tune the PI + CI controller, PI + CI-SIMC, with reset ratio value preset = 0.55.
In Fig. 5.4, closed-loop responses and control signals are compared for the PI-
IMC, PI-SIMC, and PI + CI-SIMC compensators. It can be seen that the PI-SIMC
gives a disturbance response faster than the one given by the PI-IMC controller.
On the contrary, when the reference signal is changed, the PI-SIMC controller in-
creases the overshoot in comparison with the PI controller tuned by IMC, making
its transitory response longer. The PI + CI controller will be used to reduce this tran-
sitory response by decreasing the overshoot, and without modifying its disturbance
response, as it is shown in Fig. 5.4. As a result, it can be concluded that the PI + CI
compensator will give a good trade-off between reference tracking and disturbance
rejection, without increasing the cost of feedback (as a general rule, the cost of feed-
back for a PI + CI compensator is similar to the cost of its base compensator).
In order to get a better comparison between the different compensators, the values
of performance indexes such as the integral of absolute error (IAE) and the integral
time absolute error (ITAE) are presented in Table 5.1.
From these indexes, it can be concluded that the PI + CI controller makes the
disturbance response faster, and therefore much better, in comparison with the P-
IMC (IAEPI+CI-SIMC = 2.33 s versus IAEPI-IMC = 5.85 s), just by making the refer-
ence response slightly worse (IAEPI+CI-SIMC = 2.30 s versus IAEPI-IMC = 2.20 s).
In addition, the robustness indexes, phase and gain margins, are also computed in
Table 5.1 (by using the describing function in the case of PI + CI), and it can be
expected that the robustness properties of PI + CI are close to the values obtained
by its base compensator.
As a result, for FOPDT systems with dominant lag the PI + CI compensator
gives a good balance between disturbance rejection and reference tracking without
192

Fig. 5.4 System responses and control signals for PI-IMC, PI-SIMC, and PI + CI-SIMC controllers
5 Design of Reset Control Systems
5.1 The PI + CI Compensator 193

sacrificing robustness. After extensive simulation, it has been found that a value
of preset ∈ [0.3, 0.6] gives a good response having better rejection/worse reference
tracking characteristic for increasing values of preset . Rules for tuning a PI + CI
compensator for a lag dominant system are now summarized:
1. Tune the base PI controller, PIbase , by using
τ
kp = and τi = 8h.
2kh
2. Choose the reset ratio preset in the interval [0.3, 0.6], balancing between perfor-
mance and robustness.

5.1.2.2 Time Delay Dominant Systems

A system is said to be time-delay dominant when its time constant τ is at least of the
same order as its delay h, usually h ∈ [0.5, 1]τ . An example of this kind of system
is the following model:
1
P (s) = e−s . (5.31)
2s + 1
For this system, a PI-IMC compensator is tuned by using the IMC method (see
(5.27)–(5.28)). The following parameters are obtained: kp = 1 and τi = 2 s. At first,
a PI + CI-IMC is tuned by using as base compensator the PI-IMC and a small value
of preset , for example, preset = 0.1.
In Fig. 5.5, the closed-loop responses and the control signals are shown. It can be
seen that the reset action makes the PI + CI behavior worse for a reference change
since overshoot does not decrease and undershoot increases considerably. On the
other hand, when an input load disturbance is considered, the PI + CI controller
behaves in the same way as the PI one since the response does not cross the reference
signal, and thus there are no resets.
In general, reset action does not improve the linear response due to the fact that
the base PI compensator, tuned by the IMC method, gives a response with a low
overshoot; in such a way, the control signal reaches the steady state very fast and
with practically no oscillations. Then, when the reset is performed, the control signal
is close to its steady-state value and any reset ratio, even small valued, makes the
control signal move away from its steady state-value giving a worse response.
To solve this problem, a detuning of the PIbase compensator is performed. This
detuning consists of making the closed-loop response faster, and thus more oscil-
latory, by fixing a smaller value of the time constant of the desired closed-loop
response λ. Note that the λ value had been fixed previously to be equal to the time
delay, h. Thus, in order to make the response faster, now λ will be fixed to be two
thirds of the time delay, that is, λ = 23 h.
With this new λ value, the integral time constant of the base PI controller (5.27)
does not vary, but the proportional gain (5.28) does. By replacing λ = 23 h in (5.26),
194

Fig. 5.5 System responses and control signals for PI-IMC and PI + CI-IMC controllers
5 Design of Reset Control Systems
5.1 The PI + CI Compensator 195

Table 5.2 Performance indexes


Reference Disturbance Stability margins
IAE (s) ITAE (s2 ) IAE (s) ITAE (s2 ) ϕm (°) Am (dB)

PI-IMC 2.17 5.05 2.00 47.89 61.3 9.94


PI-des 2.10 5.01 1.67 39.46 55.6 8.36
PI + CI-des 2.15 5.03 1.67 39.46 56 8.23

this new value of the proportional gain is expressed as



kp = . (5.32)
2kh
With this modification, the detuned PI (PI-des) base parameters are now kp = 1.2
and τi = 2 s. In addition, the detuned compensator is the base for the reset com-
pensator PI + CI-des, where a small reset ratio is also considered, for example,
preset = 0.1. In Fig. 5.6, closed-loop responses and the control signals are compared
for the PI-IMC, PI-des, and PI + CI-des compensators. In addition, for a better com-
parison, the values of the performance indexes are presented in Table 5.2.
It can be seen that PI-des, in spite of having bigger overshoots and undershoots,
by sacrificing some robustness, behaves better than the PI-IMC since, when the
reference signal changes, the system response is faster, and smaller IAE and ITAE
values are obtained. When input disturbances are considered, it is seen that PI-des
also makes the system response better, by obtaining a faster response with lower
performance indexes.
However, comparing PI + CI with its base compensator PI-des, it is deduced that
the PI + CI response is slightly worse with similar robustness values. Thus, it has to
be concluded that without any modification PI + CI compensation is not appropriate
for FOPDT systems with dominant delay. This worse behavior of the reset system is
due to the detrimental effect of the dominant time delay over the reset action. In the
next section, we will discuss how the use of a reset band can improve this behavior.

5.1.3 High Order Systems

In order to be able to get tuning rules for more general systems, a simple approach
is to approximate them by an FOPDT model. For this purpose, in this subsection the
approximation given in [14] is used. This approximation was developed by using the
simple half-rule which considers that the largest neglected (denominator) time con-
stant (lag) is distributed evenly to the effective delay and the smallest retained time
constant. It will be briefly explained in the following for the sake of completeness.
A system P (s) given by
%
j (Tj 0 s + 1) −h0 s
P (s) = % e , (5.33)
i (τi0 s + 1)
196

Fig. 5.6 System responses and control signals for PI-IMC, PI-des and PI + CI-des controllers
5 Design of Reset Control Systems
5.1 The PI + CI Compensator 197

where τi0 > τi+1,0 , and Tj 0 > Tj +1,0 is approximated by an FOPDT system

k
P̃ (s) = e−hs , (5.34)
τs + 1
where the model time constant τ is calculated with the following expression:
τ20
τ = τ10 + , (5.35)
2
and the effective time delay h is computed by using
τ20 
h = h0 + + τi0 . (5.36)
2
i3

Finally, if the general system (5.33) has zeros, the numerator term (T0 s + 1) is
canceled by a neighboring denominator term (τ0 s + 1), where both T0 and τ0 are
positive and real, using the following approximations



T0
if T0  τ0  h,

⎪ τ0


T0 s + 1 ⎨ h if T0  h  τ0 ,
T0
(5.37)
τ0 s + 1 ⎪ ⎪ 1 if h  T0  τ0 ,




⎩ T0 if τ  T  5h.
τ0 0 0

If there is more than one positive numerator time constant, then one T0 should be
approximated at a time, starting with the largest T0 .
Once the general system is approximated to an FOPDT model, the tuning rules
developed in the above section can be directly applied in order to tune the PI + CI
compensator. For example, consider the following system:

(15s + 1)
P (s) = e−s . (5.38)
(25s + 1)(20s + 1)(s + 1)
Firstly, its effective time delay is computed by using (5.36), and the value
h = 1.5 s is obtained. Secondly, the zero (15s + 1) is canceled against the neigh-
boring pole (20s + 1) with (5.37). Since the condition τ0  T0  5h is fulfilled, the
zero and pole are approximated as:

15s + 1 15
≈ = 0.75. (5.39)
20s + 1 20
Finally, the approximated time constant is calculated with (5.35), so τ = 25.5 s
is obtained. Therefore, the system (5.38) is approximated by the following FOPDT
model:
0.75
P̃ (s) = e−1.5s . (5.40)
25.5s + 1
198 5 Design of Reset Control Systems

Table 5.3 Performance indexes


Reference Disturbance Stability margins
IAE (s) ITAE (s2 ) IAE (s) ITAE (s2 ) ϕm (°) Am (dB)

PI-IMC 4.52 25.42 6.29 457.98 50.6 10.31


PI-des 5.55 38.00 3.18 200.30 42.6 9.77
PI + CI-des 4.81 26.70 3.18 200.30 40.8 9.25

From this model it is seen that its time constant, τ = 25.5 s, is much bigger than
its time delay, h = 1.5 s, thus the model (5.40) is a lag dominant system. As a result,
it may be expected that reset compensation will be effective in this case to improve
performance without sacrificing robustness and cost of feedback. Firstly, a base PI-
SIMC compensator is tuned by using (5.28) and (5.29); the result is kp = 11.3 and
τi = 12 s. In addition, for the PI + CI-SIMC compensator, an initial value of the
reset ratio preset = 0.55 is chosen.
For comparison purposes, a PI-IMC compensator will be also used. Using (5.27)
and (5.28), the parameters are kp = 11.3 and τi = 25.5 s. In Fig. 5.7, closed-loop
responses and control signals are shown for PI-IMC, PI-SIMC, and PI + CI-SIMC
compensators. Both PI-SIMC and PI + CI-SIMC give a faster response than PI-
IMC, overall when input disturbances are considered. However, when the reference
signal changes, the system responses with PI-SIMC and PI + CI-SIMC exhibit more
overshoot, and therefore worse performance indexes, as it can be seen in Table 5.3.
When PI + CI is used, the PI-SIMC overshoot is reduced and better perfor-
mance indexes for references are obtained. In spite of this overshoot reduction,
the PI + CI response is slightly worse than the PI-IMC response when reference
changes are considered (IAEPI+CI-SIMC = 4.81 s versus IAEPI-IMC = 4.52 s), but it
is much better when input disturbances are considered (IAEPI+CI-SIMC = 3.18 s ver-
sus IAEPI-IMC = 6.29 s). As a conclusion, and without surprise, it can be said that
PI + CI compensation gives the best balance between performance and robustness
for higher order systems with deadtime that can be well approximated by FOPDT
systems with dominant lag.

5.1.4 Integrating Systems

Integrating systems, that is, systems with an integrator, are known to be systems
for which reset control may give significant improvement over linear control. This
property, already mentioned in the seminal work of Clegg [7], has been explicitly
shown, for example, in [6]. In the following, it will be shown how PI + CI compen-
sation can beat a well tuned PI compensator for this type of systems, and in addition,
tuning rules will be given.
In the previous systems, the reset ratio could never be fixed at its maximum
value since this value would make the system response oscillate indefinitely and
the steady-state would not be reached. In the case in which the system has one
5.1 The PI + CI Compensator

Fig. 5.7 System responses and control signals for PI-IMC, PI-SIMC, and PI + CI-SIMC controllers
199
200 5 Design of Reset Control Systems

Table 5.4 Performance indexes


Reference Disturbance Stability margins
IAE (s) ITAE (s2 ) IAE (s) ITAE (s2 ) ϕm (°) Am (dB)

PI 6.43 62.80 4.48 328.3 48.6 24.28


PI + CI 4.17 20.44 4.48 328.3 48.4 23.9

integrator, this is no longer true, and thus the reset ratio can be set at its maximum
value preset = 1.
An FOPDT system can be approximated to an integrating system when its time
constant is really large, τ → ∞:

k k −hs k  −hs
P (s) = e−hs e = e . (5.41)
τs + 1 τs s
Obviously, this integrating system behaves as a lag dominant system, τ h, in
such a way that the PI + CI controller is tuned by using the tuning rules developed
for those systems. However, in order to be able to apply these tuning rules, the
proportional gain has to be modified by using k  = τk .
As a result, PI + CI tuning rules for integrating systems like (5.41) are simply:
1. Tune the base PI controller, PIbase , by using
τ
kp = and τi = 8h.
2kh
2. Set the reset ratio to its maximum value, preset = 1.
The following example is considered to demonstrate the effectiveness of these
tuning rules:
(0.17s + 1)2
P (s) = , (5.42)
s(s + 1)2 (0.028s + 1)
which can be approximated with the following model:
1
P̃ (s) = e−1.69s . (5.43)
s
With this model, the PI + CI compensator is tuned by using the tuning rules
given above. Firstly, the PIbase parameters are computed, and the following values
are obtained: kp = 0.3 and τi = 13.5 s. Obviously, these parameters are used in the
PI + CI compensator together with a reset ratio, in this case, preset = 1.
In Fig. 5.8, the closed-loop responses and control signals are drawn for the PI
and PI + CI compensators (with the plant given by (5.41)). In addition, the values
of the performance indexes and stability margins are shown in Table 5.4.
From Fig. 5.8 and Table 5.4, it can be deduced that the reset action improves
considerably the linear system behavior since the PI + CI controller not only gives
a response with less overshoot but also reduces the settling time without modifying
5.1 The PI + CI Compensator

Fig. 5.8 System responses and control signals for PI and PI + CI controllers
201
202 5 Design of Reset Control Systems

the response speed. On the other hand, when input disturbances are considered, the
reset control system behaves in the same way as the linear one since the response
does not cross the reference and thus the reset action does not occur.
It is worthy to note that in this case the stability margins, see Table 5.4, are hardly
modified with the reset action. Therefore, it can be concluded that the PI + CI con-
troller makes the response much better without modifying its robustness properties.

5.1.5 Summary of Tuning Rules

The PI + CI reset compensator simply consists of a parallel combination of a PI


compensator and a Clegg integrator (CI). Both terms may be combined in such a
way that the PI + CI behaves like a PI base compensator until the first reset action,
thus tuning a proper base PI compensator is a natural way to tune the PI + CI com-
pensator, that will only need an extra parameter preset . In principle, PI + CI compen-
sation may have advantages for controlling systems usually found in a wide range
of applications. In particular, it has been found that PI + CI compensation gives sig-
nificant improvements in first order plus deadtime (FOPDT) systems with dominant
lag, and also in integrating systems. On the other hand, higher order systems that
can be well approximated by lag dominant FOPDT and integrating systems can also
benefit from PI + CI compensation.
Tuning rules have been developed that basically consist of using a previously
tuned PI compensator and resetting some percentage of the integral term. In general,
this partial reset of the integral term results in an improvement on the closed-loop
transient response, reducing the overshoot percentage and overall the settling time
corresponding to the design without reset, and without degrading the speed of re-
sponse as well as robustness and cost of feedback. On the one hand, the first order
case has been developed by using directly the state-space equations, and the tuning
of the PI + CI compensator results in a step response that reaches the steady-state in
a finite time (exactly at the second reset time). A summary of the tuning rules may
be found in Table 5.5. Note that, without modification, PI + CI compensation is not
recommended for systems with dominant delays. In this case, there are several mod-
ifications that will be introduced in next section, including the use of a reset band
(with fixed or variable size) and a variable reset ratio.

5.2 Design Improvements

In practice, a number of simple modifications of the original reset compensator


definition have been found useful in improving performance and robustness of the
closed-loop system. In this section, several modifications will be analyzed, includ-
ing both changes over the reset law that regulates the instants at which reset actions
are performed and in the compensator design.
5.2 Design Improvements 203

Table 5.5 Tuning rules


System kp τi preset

− απ
1+kkp 2 kk β
First order without delay ( < τ τpi
2τ ) preset = e
− απ
/ 1+e β
1+kk kk
α = 2τ p , β = τ τpi − α 2

First order with dominant lag τ


2kh 8h [0.3, 0.6]

First order with dominant delay 2kh τ –

Second and higher order Approximate by a first order plus time delay
1
Integrating systems 2kh 8h 1

Regarding the reset law, a first simple modification will consist of using a reset
band, as introduced in Sect. 3.4.2, that is, reset actions are produced when the error
signal is entering a predefined reset band. In a second stage, this reset band is con-
sidered to be time varying depending on the error signal and its time derivative; in
this way, reset actions may be more effective, for example, if different setpoints are
defined in the control problem, and moreover, not only in the transitory but also in
the steady state, adapting the bandwidth to the magnitude of the error and its time
derivative. In general, these changes over the reset law are specially effective when
significant time-delays are present, for example, in applications like process control
or teleoperation.
On the other hand, to improve robustness especially in the case in which large
plant parametric uncertainty exists, a frequency domain robust control technique
like Quantitative Feedback Theory (QFT) is useful for guaranteeing some prescribed
degree of robustness in stability and performance. Thus, the base linear compensator
can be designed using QFT with some robustness specifications. Finally, consider-
ing the case of PI + CI design, it has been found useful in some cases to use a
time varying reset percentage; in general, this is related with the use of partial reset
compensators with a time varying reset dynamics.

5.2.1 Fixed Reset Band

A reset compensator with a fixed reset band has been defined in (3.90). For a reset
compensator R with input e and output v, the reset action is performed at the instant
t when the pair (e(t), ė) is in the set Bδ .
In general, the use of a fixed reset band is expected to be a good design choice
for control of systems with significative delay h, and with a constant reference r; for
example, this is the case of regulating a process around a fixed setpoint r. In this way,
the bandwidth δ is directly related with h and r since the goal is to anticipate the
crossing for the reset action to be performed. Thus, for a closed-loop step response
204 5 Design of Reset Control Systems

Fig. 5.9 System responses and control signals for PI and PI + CI controllers

with a rise time tr , a good approximate value of δ is simply given by


δ 0.8h
≈ . (5.44)
r tr
For example, consider the integrating system given in Sect. 5.1.4. In this case,
the rise time has a value tr ≈ 2.5 s and the delay is h = 1.69 s. Thus, the value of the
reset band is δ = 0.5. Figures 5.9 and 5.10 show the step response and the control
signal for the PI + CI compensator with a reset band of δ = 0.5, respectively. Note
that in comparison with the PI + CI compensator, the step response has improved in
terms of the overshoot and undershoot.
In addition, using the describing function analysis of the Clegg integrator with
reset band given in Sect. 3.4.2, it also possible to come up with a value of δ/r
for obtaining a maximum phase lead around the frequency 1/ h. In Chap. 3, the
describing function of the Clegg integrator with a fixed reset band was computed
and given by
 / 
1 j 4 1 − ( Eδ )2
j sin−1 ( Eδ )
DCI (E, ω) =
δ
1+ e , (5.45)
jω π

where E is the error amplitude. Thus, the PI + CI describing function is simply


given by
  / 
reset 1 − ( E )
δ 2
1 j 4p −1 δ
δ
DPI+CI (E, ω) = kp 1 + 1+ ej sin ( E ) . (5.46)
j ωτi π
5.2 Design Improvements 205

Fig. 5.10 System responses and control signals for PI and PI + CI controllers

For comparison purposes, in Fig. 5.11, the Bode plot is shown for a PI compen-
sator (kp = τi = 1) and for a PI + CI compensator (kp = τi = 1 and preset = 0.50)
with several fixed reset bands ( Eδ = 0.25, 0.50, 0.75, and 1.00). It can be seen that
the controller phase is increased with the fixed reset band for some values. Specif-
ically, in this case, when preset = 0.50, the higher controller phase is obtained in
a range Eδ ∈ [0.25, 0.50]. This phase increment should make the delay effect less
influential over the responses, by giving smaller overshoots and undershoots, and
hence, smaller IAE and ITAE values.

5.2.2 Variable Reset Band/Advanced Reset

The main drawback of a fixed reset band is that the compensator is specifically
tailored for a fixed reference (or disturbance) value. In practice, if different levels of
references and/or disturbances are present then a variable bandwidth is necessary.
For example, in Fig. 5.12, the example considered above in Fig. 5.9 is used with
different step levels; note that the PI + CI with fixed band (dashed line) adjusted for
a unit step gives poor results for the different step levels at t = 15 and t = 30 s.
A solution to this problem is to consider reset compensators with a variable reset
band, as given in Sect. 3.4.2.4. Both Figs. 5.12 and 5.13 give a comparison of fixed
and variable reset compensators. It turns out that the variable reset band is a more
efficient design choice in control practice where different levels of references and/or
disturbances may occur.
206 5 Design of Reset Control Systems

Fig. 5.11 Bode plots for PI + CI controller with several fixed reset bands

Fig. 5.12 System responses and control signals for PI and PI + CI controllers

5.2.3 Variable Reset Percentage

Until now, the reset percentage in the PI + CI controller has been given by a fixed
value, preset , which means that the PI + CI controller will always reset the same
amount of the integral term.
In the following, the reset percentage is going to be modified at each reset time so
that the PI + CI controller improves the system response. This modification consists
of making the reset percentage value variable at the reset times through an element
5.2 Design Improvements 207

Fig. 5.13 System responses and control signals for PI and PI + CI controllers

of prediction, such as the rate of change of the error signal. Specifically, the reset
percentage will be changed as a function of the error derivative through the tuning
of a new parameter, the derivative time constant, τd . In addition, a low-pass filter
will be used with the pure derivative term to avoid the amplification of the sensor
noise,
1
F (s) = , (5.47)
τf s + 1
where the filter time constant, τf , will be another parameter to be tuned. In this way,
the variable reset percentage, p̂reset , can be given through the following expression:

deF
p̂reset (t) = p̄reset − τd , (5.48)
dt
where p̄reset is a constant base value of reset percentage and eF is the error signal
filtered by using F . As it can be seen, from (5.47)–(5.48), three parameters must be
fixed to tune a PI + CI controller with variable reset percentage from a PI + CI one
with fixed reset percentage. To choose the values of these parameters, it is important
to observe that the variable reset percentage must make the reset action act strongly
over the transitory response, but weakly when the system reaches the steady-state.
In this way, the base reset percentage, p̄reset , should have a varying value in order
to avoid a high influence on the system response when the steady-state is reached.
On the other hand, the reset action should have strong influence over the transitory
response, so the derivative time constant, τd , is fixed for that purpose. Finally, the
filter time constant, τf , should be tuned to avoid the undesirable noise amplification.
In Chap. 6, an application example of the variable reset percentage technique will
be developed (see Sect. 6.3)
208 5 Design of Reset Control Systems

5.2.4 Robust Control Design Based on QFT

The control of uncertain plants is of considerable importance from both theoretical


and practical points of view. Although most feedback control systems provide some
robustness with respect to uncertainty, the development of techniques that explicitly
consider some uncertainty in the design process is a relatively recent development.
One exception to this statement is found in the work of Isaac Horowitz [9]. Since the
late 1950s, his work has focused on the control of systems with parametric uncer-
tainty. The linear case was first considered, and later work in the 1970s [10, 11] also
considered robust control of some nonlinear and/or time-varying systems with para-
metric uncertainty. In those works, Horowitz developed a technique for translating a
nonlinear and/or time-varying problem to an equivalent linear problem that, at least
in some cases, can be solved using his earlier linear robust design techniques. All
of these early ideas were then expanded and integrated by Horowitz and cowork-
ers, leading to a set of control design methods including scalar/multivariable, lin-
ear/nonlinear, time-invariant/time-varying, single-loop/multiple-loop systems, now
referred to as quantitative feedback theory (QFT) [9].
QFT is a robust control technique that is especially well suited for control prob-
lems with large plant uncertainty, and that uses frequency domain specifications.
A key step in QFT is the mapping of these specifications into allowed regions of
the Nichols plane by using plant uncertainty regions referred to as templates. These
allowed regions are relative to the open-loop gain function, and their borders are
usually referred to as boundaries. The final step is loop shaping and aimed to shape
the open-loop function by minimizing its high frequency gain. QFT works in the
frequency domain, thus plant models can be derived from transfer functions (usu-
ally with parametric uncertainty) or directly by sets of frequency responses. QFT
basically consists of several design steps:
1. Computation of templates. A template represents, at a given frequency, the un-
certainty of the plant. It is a region of the Nichols plane, with each template
point being given by the phase and magnitude of a plants set element. For a set
of working frequencies, the first design step consists of computing the templates
set.
2. Computation of boundaries. Given (robust) stability and performance specifica-
tions, each template generates a boundary. If the nominal open-loop gain avoids
the boundaries, one boundary at every working frequency, then closed-loop spec-
ifications are satisfied for all the plants considered in the template.
3. Nominal open loop shaping. Once the boundaries are computed, the next design
step is to compute (shape) the open-loop gain that fits them in some optimal way.
In general, this is a hard computational problem that usually has been solved
heuristically. Once the open-loop gain is obtained, the feedback compensator is
directly computed. In addition, a precompensator can be added to the feedback
structure if tracking specifications are to be satisfied.
A detailed description of these design steps may be found in [9]. For the case
of reset control, and in particular for PI + CI design, the novelty is simply that the
References 209

open loop gain function is described by L(s) = G(s)P (s), where the frequency
response of the feedback compensator is directly given by the describing function
of the PI + CI compensator (5.9), that is, G(j ω) = (PI + CI)(j ω). Alternatively, it
can be used to design the base compensator, guaranteeing that the response has the
desired oscillatory characteristic for the reset to be effective, and in the presence of
large plant uncertainty.
In Chap. 6, a number of examples and case studies will be developed by using
QFT.

References
1. Åström, K.J., Hägglund, T.: The future of PID control. Control Eng. Pract. 9, 1163–1175
(2001)
2. Bakkeheim, J., Smogeli, O.N., Johansen, T.A., Sorensen, A.J.: Improved transient perfor-
mance by Lyapunov-based reset of PI thruster control in extreme seas. In: Proc. IEEE Confer-
ence on Decision and Control, pp. 4052–4057 (2006)
3. Baños, A., Vidal, A.: Definition and tuning of a PI + CI reset controller. In: Proc. European
Control Conference, Kos, Greece (2007)
4. Baños, A., Vidal, A.: Design of PI + CI reset compensators for second order plants. In: Proc.
IEEE International Symposium on Industrial Electronic, Vigo, Spain (2007)
5. Baños, A., Carrasco, J., Barreiro, A.: Reset-times dependent stability of reset systems. In:
Proc. European Control Conference, Kos, Greece (2007)
6. Beker, O., Hollot, C.V., Chait, Y.: Plant with integrator: an example of reset control overcom-
ing limitations of feedback. IEEE Trans. Autom. Control 46(11), 1797–1799 (2001)
7. Clegg, J.C.: A nonlinear integrator for servomechanisms. Trans. AIEE, Part II 77, 41–42
(1958)
8. Fernández, A.F., Barreiro, A., Baños, A., Carrasco, J.: Reset control for passive bilateral tele-
operation. IEEE Trans. Ind. Electron. (2010). doi:10.1109/TIE.2010.2077610
9. Horowitz, I.M.: Quantitative Feedback Theory. QFT Press, Boulder (1992)
10. Horowitz, I.M., Rosenbaum, P.: Nonlinear design for cost of feedback reduction in systems
with large parameter uncertainty. Int. J. Control 24(6), 977–1001 (1975)
11. Krishman, K.R., Horowitz, I.M.: Synthesis of a nonlinear feedback system with significant
plant-ignorance for prescribed system tolerances. Int. J. Control 19(4), 689–706 (1976)
12. Lee, Y., Park, S., Lee, M., Brosilow, C.: PID controller tuning for desired closed-loop re-
sponses for SI/SO systems. AIChE J. 44(1), 106–115 (1998)
13. Rivera, D.E., Morari, M., Skogestad, S.: Internal model control 4, PID controller design. Ind.
Eng. Chem. Process Des. Dev. 25(1), 252–265 (1986)
14. Skogestad, S.: Simple analytic rules for model reduction and PID controller tuning. J. Process
Control 13(4), 291–309 (2003)
15. Smith, C.L., Corripio, A.B., Martin, J.: Controller tuning from simple process models. In-
strum. Technol. 22(12), 39–44 (1975)
16. Vidal, A., Baños, A.: QFT-based design of PI + CI reset compensator: applications in process
control. In: 16th IEEE Mediterranean Conference on Control and Automation, Ajaccio, France
(2008)
17. Vidal, A., Baños, A.: Stability of reset control systems with variable reset: application to
PI + CI compensation. In: European Control Conference, Budapest, Hungary (2009)
18. Vidal, A., Baños, A.: Reset compensation applied on industrial heat exchangers. In: 14th IEEE
International Conference on Emerging Technologies and Factory Automation, Mallorca, Spain
(2009)
210 5 Design of Reset Control Systems

19. Vidal, A., Baños, A., Moreno, J.C., Berenguel, M.: PI + CI compensation with variable re-
set: application on solar collector fields. In: 34th Annual Conference of the IEEE Industrial
Electronics Society, Orlando, Florida, USA (2008)
20. Wu, D., Guo, G., Wang, Y.: Reset Integral-Derivative Control for HDD servo systems. IEEE
Trans. Control Syst. Technol. 15(1), 161–167 (2007)
21. Yang, T.: Impulsive Control Theory. Lecture Notes in Control and Information Sciences,
vol. 272. Springer, Berlin (2001)
22. Zheng, Y., Chait, Y., Hollot, C.V., Steinbuch, M., Norg, M.: Experimental demonstration of
reset control design. Control Eng. Pract. 8(2), 113–120 (2000)
23. Zheng, J., Guo, Y., Fu, M., Wang, Y., Xie, L.: Development of an extended reset controller
and its experimental demonstration. IET Control Theory Appl. 2, 866–874 (2008)
Chapter 6
Application Cases

6.1 Temperature Control in a Heat Exchanger

Temperature control is a common problem in the process industry, such as power


generation, chemical plants, refrigeration, and food processing among others. This
difficult temperature control problem arises from the large process uncertainty and
the different operation conditions which are present during the whole process. For
example, in batch reactors the temperature has to be adjusted to an optimal temper-
ature profile in spite of the drastic changes at setpoints with the goal of obtaining
a specific product with desired properties [3, 15, 32]. In other applications, such as
thermal treatments in heat exchangers, the outlet temperature has to be maintained
at a particular setpoint for a wide variety of products and different operation points
in order to destroy harmful microorganisms and optimize food quality factors [5,
23, 33]. Because of these particularities, thermal control is difficult to achieve with
conventional PID controllers. Since variable operating conditions occur, the param-
eters of the PID controller require frequent retuning to adjust them to these different
operating conditions. Several works deal with the temperature control problem by
using adaptive model predictive control [1–3, 27, 32]. Other works are focused on
the use of the Quantitative Feedback Theory (QFT) to tune robust controllers which
take into account the uncertainty of the thermal process [9, 16].
QFT is a robust control technique that is especially well suited for control prob-
lems with large plant uncertainty (see Sect. 5.2.4 for a brief introduction). A family
of linear time invariant (LTI) plants obtained by local linearization may be used to
model the heat exchanger nonlinear dynamics around different operation points, and
then a compensator can be designed by using QFT [8]. The result is a robust com-
pensator that guarantees a specified worst-case behavior for all the considered oper-
ation points. In general, the robustness and performance have to be balanced due to
fundamental limitations of the resulting LTI compensator, this being the main moti-
vation for using reset compensation, in which this type of nonlinear/hybrid control is
investigated with the goal of improving performance without sacrificing robustness.
In Chap. 5, tuning rules of the PI + CI compensator for first and higher order
systems with time delay have been developed. In this section, this reset controller

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 211


DOI 10.1007/978-1-4471-2250-0_6, © Springer-Verlag London Limited 2012
212 6 Application Cases

Fig. 6.1 Pilot plant at the


University of Murcia

is tuned by using QFT and applied to the temperature control of an industrial heat
exchanger. In Sect. 6.1.1, a dynamic model of the heat exchanger is obtained for
different operation points based on experimental data. In Sect. 6.1.2, several PI + CI
compensators are designed. First, a PI + CI compensator is tuned by using QFT, and
then several design improvements such as the use of a fixed reset band and a variable
reset band, are analyzed. The different reset compensators are experimentally tested
and compared with a well designed PI compensator.

6.1.1 Process Model

A temperature control problem will be studied by using an industrial heat exchanger,


commonly used in the food industries for thermal treatments. For this purpose,
the temperature process will be modeled through pseudorandom binary sequence
(PRBS) experiments and step tests application. PI + CI compensation will be used
to fix the outlet temperature at a desired value with minimum performance indexes,
such as the integral absolute error (IAE).
The heat exchanger used in this study is part of a pilot plant which has been
designed for food industry applications, and it is a part of the pilot plant of the
Computer and Control Engineering group at the University of Murcia (see Fig. 6.1).
It contains two HRS UNICUS double tube heat exchangers, consisting of a 3 me-
ter long tube within another one in parallel flow. That is to say, the product flows
through the inner tube in parallel to the serviced fluid which flows through the an-
nulus between the inner and the outer tube. In addition, the shell and tube heat
exchanger have scraper bars fitted in each interior tube whose movement mixes the
fluid and cleans the heat exchange surface. This also keeps heat transfer high and
makes cleaning tasks easier in food industries. The physical details of this heat ex-
changer are shown in Table 6.1.
The product is pumped from a storage tank up to the heat exchanger by an indus-
trial helicoidal impeller pump. Two resistance thermometers used as thermocouples
6.1 Temperature Control in a Heat Exchanger 213

Table 6.1 Some parameters


of the industrial heat Number of inner tubes 1
exchanger Shell diameter 104 mm
Inner tube diameter 76.2 mm
Type of flow parallel
Material stainless steel
Tubes length 3000 mm

Fig. 6.2 Electropneumatic


globe valve

are placed at both the inlet and the outlet of the inner tube of the heat exchanger
for temperature measurements. The product is heated in the inner tube from the sat-
urated steam condensation at the shell side. The rate of heat transfer depends on
the temperature difference between the cold and hot fluids in such a way that the
higher this difference is, the faster the heat is transferred. Therefore, the rate of heat
transfer, q in kcal/h, is expressed through the following equation:

q = U AΔT , (6.1)

where A is the area for heat transfer in m2 , ΔT is the temperature difference in °C


and U is the overall heat transfer coefficient, kcal/°C m2 .
The saturated steam is generated in an electrical boiler of 12 kW of power which
generates 16 kg of steam per hour at a pressure which oscillates between 4.3 and
5 bar. These pressure drops will be taken into account as a perturbation, d, in the
control scheme. The control signal will be precisely the steam flow rate, which will
be controlled through the aperture (expressed in %) of an electropneumatic globe
valve placed just before the shell inlet in the heat exchanger, as seen in Fig. 6.2.
The thermocouples and the control valve are connected through a PLC to a data
acquisition system, where a SCADA is installed for monitoring and control pur-
poses, as can be seen in Fig. 6.3. The SCADA system reads both the inlet and outlet
temperatures, Tin and Tout , and from the temperature difference, ΔT = Tout − Tin ,
214 6 Application Cases

Fig. 6.3 Industrial heat


exchanger control

the control system provides the corresponding electropneumatic globe valve aper-
ture, u in %, in order to reach the desired reference temperature.
The industrial heat exchanger is modeled by using water as the product with a
flow rate of 250 L/h. The identification is done through pseudorandom binary se-
quence (PRBS) experiments and step test application. On the one hand, the heat
exchanger models are obtained by first order approximation describing the response
of the temperature along the heat exchanger to several step changes of valve aper-
ture. Specifically, this identification was done by opening the control valve from 10
up to 25% in 5% steps and then by closing the control valve up to 10% by using
again 5% steps as can be seen in Fig. 6.4.
In addition, the PRBS experiments were also done to model the heat exchanger.
These experiments [30] consist of doing different tests by using the control signal

Fig. 6.4 Heat exchanger identification


6.1 Temperature Control in a Heat Exchanger 215

Fig. 6.5 PRBS experiment

Table 6.2 Heat exchanger uncertainty


Plant Pa –b% Gain ka –b% Time constant τa –b% (s) Time delay ha –b% (s)

P10–15% 0.32 87 88
P15–20% 0.35 82 83
P20–25% 0.40 80 65
P25–20% 0.40 89 50
P20–15% 0.35 95 76
P15–10% 0.31 114 84

as the input, which will have a spectrum with significant components in some re-
gions of interest. Then, the different response data are used for getting an uncertain
parametric model. The range [10, 25]% for the control valve aperture is identified
as the working range for this study. One example of a PRBS experiment is shown in
Fig. 6.5 for a valve aperture between 10 and 15%.
With these modeling experiments, the heat exchanger can be represented by a set
of first order plus time delay (FOPTD) transfer functions given by

ka−b%
Pa−b% (s) = e−ha−b% s . (6.2)
τa−b% s + 1

For each valve aperture change, from a% to b%, a heat exchanger model is ob-
tained, as given in Table 6.2. The heat exchanger uncertainty is expressed in para-
metric form as ka−b% ∈ [0.31, 0.40], τa−b% ∈ [80, 114], and ha−b% ∈ [50, 88]. Note
that every parameter combination is not possible, in fact, the gain is increasing for
increasing valve aperture values, while the time lag and the delay are decreasing.
216 6 Application Cases

6.1.2 PI + CI Design

In the following, a PI + CI compensator is designed by using the results given in


Chap. 5. Different control experiments using the industrial heat exchanger will be
carried out with different control systems including a base PI compensator and dif-
ferent reset PI + CI compensators (with/without reset band, . . . ). The experiment
will consist of setting a fixed product flow of 250 L/h, as considered in the modeling
procedure, and opening the control valve at 10%. When the steady-state temperature
is reached at this control valve aperture, the reference temperature is increased and
decreased by step changes with a magnitude of 2 °C.

6.1.2.1 PI Compensator Design

For comparison purposes, a linear PI controller is (well-)tuned by using the Internal


Model Control (IMC) technique as described in Chap. 5. This method consists of
considering a known model of the process in order to get a desired closed-loop re-
sponse. For this purpose, the controller parameters will depend on the system param-
eters and on the time constant of the desired closed-loop response, λ. Specifically,
when the process is modeled as an FOPDT system given by (6.2), the controller
takes the structure of the PI controller with the following parameters [38]:
τ
kp = , (6.3)
k(λ + h)

τi = τ. (6.4)
The value of the desired closed-loop time constant λ can be chosen freely, but due
to (6.3) it has to satisfy −h < λ < ∞ to get a positive and nonzero controller gain.
In general, the optimal value of λ should be determined by a trade-off between:
1. Fast response and good disturbance rejection (small value of λ) and
2. Stability and robustness (large value of λ).
Between the different ways of fixing the value of λ which have been studied in
the literature [31, 35, 36], the one given in [36] is used in this work. So, the value λ
is chosen to be equal to the time delay λ = h. This gives a reasonably fast response
with good robustness margins. In this way, the PI proportional gain (6.3) is given as
τ
kp = . (6.5)
2kh
For the industrial heat exchanger modeled by (6.2) and with parameters as in
Table 6.2, a nominal model has to be considered to tune the PI controller via IMC.
In this work, the chosen model is the one in which the control valve is opened from
15% up to 20%
0.35 −83s
P15−20% = e . (6.6)
82s + 1
6.1 Temperature Control in a Heat Exchanger 217

Fig. 6.6 Responses and control signals of PI controller tuned by IMC

For this model, by using (6.4) and (6.5), the following PI parameters are obtained:
kp = 1.41 and τi = 82 s. From now on, this controller is going to be named as PI-
IMC. The experiment explained previously has been carried out with this controller.
That is, the control valve is opened at 10% until the steady-state is reached. The
difference between the outlet and inlet temperatures, ΔT = Tout − Tin , is taken as
0 °C. Then 2 °C step changes are introduced into the reference temperature from
0 °C up to 6 °C and from this value down to 0 °C again.
In Fig. 6.6, the temperature increments in the heat exchanger are shown together
with the control signal, that is, the control valve aperture. In both signals, the noise
introduced by the thermocouples and the perturbations due to the pressure drops
occurring at the electric boiler are visible. Note that the higher the control valve
is opened, the more oscillatory the response at the steady-state is. Regarding the
control system, it can be seen that the PI-IMC controller is robust for all the refer-
ence changes. In addition, fast response is obtained (rise time between 98 and 233
seconds) with a significant overshoot in some cases (between 4.5 and 26%).

6.1.2.2 Reset Compensation Design

Here the goal is to design a reset control system that guarantees both more robust and
better performance than the PI-IMC compensator for the industrial heat exchanger
modeled by (6.2) and Table 6.2. This better performance will consist of reducing
the overshoot without decreasing its speed of response. For this purpose, a PI + CI
compensator is going to be used. As was explained in Chap. 5, this compensator
is designed from a PIbase just by adding a proper reset ratio preset . Since the PI-
IMC hardly has overshoot, a faster PIbase compensator with more overshoot has
to be tuned for the PI + CI to be effective. First, Quantitative Feedback Theory
218 6 Application Cases

(QFT) will be used to design PIbase , by using design specifications about closed-
loop stability and tracking performance. In a second stage, this PI + CI compensator
will be designed choosing a proper reset ratio preset .

PIbase Tuning The PIbase compensator will be tuned with QFT by following the
design steps (briefly) explained in Chap. 5. First of all, plant templates are computed
taking into account parametric uncertainty as given in (6.2) and Table 6.2, with the
nominal plant being
0.35 −83s
P15−20% (s) = e . (6.7)
82s + 1
In this case, the templates are obtained for a wide set of working frequencies,
from 0.002 up to 0.02 rad/s.
Secondly, the boundaries are computed for each frequency from design specifi-
cations by using templates. Design specifications will be a minimum (robust) phase
margin of 24◦ and desired tracking performance, such as a rise time between 45 and
180 seconds and a maximum overshoot between 7 and 75%.
Note that a relatively small phase margin has been specified for the base PI con-
trol system. In addition, note also that the response is specified to be faster and with
more overshoot than the response obtained with the PI-IMC controller. These de-
sign specifications are considered due to the fact that the reset compensation will
improve the minimum phase margin and will reduce the overshoot of the PI com-
pensator without reducing the response speed.
The next step consists of shaping the nominal open loop. Here a PI controller is
tuned by making the open-loop gain fit both the stability and tracking boundaries.
A shaping is shown in Fig. 6.7, corresponding to a proportional gain of kp = 1.6
and an integral time constant of τi = 60 s.
Finally, once the base PI compensator is tuned, a closed-loop system analysis is
performed to ensure that the system fits the given stability and tracking specifica-
tions. This is usually referred as to compensator validation in QFT. This analysis
has to be done for a wider set of frequencies than the one used in the templates com-
putation, and in this case, this set goes from 0.00001 up to 100 rad/s with a much
bigger grid. In Fig. 6.8, the stability analysis is shown, whereas the tracking analysis
is presented in Fig. 6.9.
As it can be seen, the designed closed-loop system satisfies both stability and
tracking specifications for every frequency. Therefore, the base PI compensator de-
sign is finished.

PI + CI Tuning Once PIbase has been designed, the obtained parameters are cho-
sen as a part of the PI + CI parameters. But in addition, one more parameter, the reset
ratio preset , must be tuned. The expected result is an increment of the phase margin,
and thus a better transient response for the industrial heat exchanger. Several values
of preset are considered: 0.1, 0.2, 0.3, and 0.4.
These PIbase and PI + CI compensators are used to control the already described
heat exchanger. For this purpose, the same experiment carried out with the PI-IMC
6.1 Temperature Control in a Heat Exchanger 219

Fig. 6.7 Base PI open-loop design

Fig. 6.8 Stability analysis

controller is repeated. Figure 6.10 shows the temperature increment and the valve
aperture when the PIbase and PI + CI compensations are used.
220 6 Application Cases

Fig. 6.9 Tracking analysis

Fig. 6.10 Responses and control signals of base PI and PI + CI controllers

As it can be seen, the PIbase and the PI + CI responses are equal until the first
reset action is performed. After this reset, the PI + CI response has the same or
smaller first overshoot while the other overshoots and undershoots are increased
with respect to the base PI system, even for small values of preset . This behavior
is compatible with the results given in Chap. 5, where PI + CI is shown to behave
worse than its base compensator for delay dominant systems.
6.1 Temperature Control in a Heat Exchanger 221

Table 6.3 Performance indexes for the PI and PI + CI control systems


Controller IAE (s) ITAE (s2 ) Phase margin (°)

Base PI 3.31 × 103 12.45 × 106 41


PI + CI preset = 0.1 3.75 × 103 14.14 × 106 44
PI + CI preset = 0.2 3.28 × 103 12.09 × 106 46
PI + CI preset = 0.3 3.62 × 103 13.12 × 106 47
PI + CI preset = 0.4 3.87 × 103 14.54 × 106 47

To evaluate the tracking performance, the integral absolute error (IAE) and the
integral time absolute error (ITAE) are computed:

IAE = |e(t)| dt, (6.8)
0

ITAE = t|e(t)| dt (6.9)
0
which should be as small as possible.
These performance indexes confirm that the reset action makes the error increase,
and therefore the response gets worse. In addition, for comparison of results, phase
margins are also computed in Table 6.3 for these compensators. These values show
the extra phase lead given by the PI + CI compensator: for the PI controller the phase
margin is 41° whereas for the PI + CI reset one the phase margin can be increased
up to 47°.
The worse response under reset conditions, as shown in Fig. 6.9, is due to the
fact that the effect of the reset action takes place after the crossing of the output
with zero, due to the existence of dominant time delays in the system. In the next
section, a simple modification of the PI + CI compensator will be investigated to
significantly improve control system performance.

6.1.2.3 Reset Compensation with Reset Band

The reset action is a good choice to improve the transient response of a closed-loop
system. But there are systems whose response improvement is not as good as can
be expected. This is due to the presence of a dominant time delay. The problem in
these dominant time delay systems is that the reset action is done at reset times tk
when the error is equal to zero, but the system suffers the reset action at another
time, specifically at tk + h, where h is the time delay of the plant. This lack of
coordination between the reset times and the times when the system undergoes the
reset action can be overcome by modifying the reset condition.
The reset condition in the PI + CI controller has been usually given by the fact
that the error signal had to be equal to zero. In this section, the reset condition is
modified with the goal of doing reset some time before the crossing between the
222 6 Application Cases

Fig. 6.11 Responses and control signals of PI-IMC and PI + CI controllers with a fixed reset band

Table 6.4 Performance indexes for the PI and PI + CI control systems with a fixed reset band
Controller IAE (s) ITAE (s 2 )

PI-IMC 2.59 × 103 9.56 × 106


PI + CI preset = 0.10 and δ
E = 0.6 2.16 × 103 8.17 × 106
PI + CI preset = 0.35 and δ
E = 0.7 2.82 × 103 10.86 × 106

error signal and zero. To specify when the reset action must exactly be done, two
techniques are considered in this study: a fixed reset band and a variable reset band.

Fixed Reset Band The experiment is carried out again, but now for two PI +
CI compensators with fixed reset band. In this case, two different values of the
reset ratio are set and their corresponding fixed reset bands are computed. In one
controller, a low reset ratio is chosen and for this value, preset = 0.1, a fixed reset
band has been computed with the goal of having the maximum phase margin in the
control system. In this case, the maximum phase margin of 47° has been obtained
for a fixed reset band of Eδ = 0.6. On the other hand, for a second PI + CI controller
with a fixed reset band, a higher reset ratio value is chosen, preset = 0.35. In this
case, the control system reaches the maximum phase margin of 60° when Eδ = 0.7.
In Fig. 6.11, the heat exchanger responses and the control signals for the PI + CI
compensator and for the PI-IMC compensator are shown. In addition, their perfor-
mance indexes are compared in Table 6.4.
From Fig. 6.11 and Table 6.4, it can be said that the PI + CI compensator with
preset = 0.35 and Eδ = 0.7, in spite of having a higher phase margin, has a response
that it is not as good as that given by the PI + CI compensator with preset = 0.1 and
E = 0.6. This last PI + CI controller not only has lower IAE and ITAE values than
δ
6.1 Temperature Control in a Heat Exchanger 223

Table 6.5 Performance indexes for the PI and PI + CI control systems with a variable reset band
Controller IAE (s) ITAE (s 2 )

PI-IMC 2.59 × 103 9.56 × 106


PI + CI preset = 0.1 and h = 50 s 2.13 × 103 8.09 × 106
PI + CI preset = 0.4 and h = 50 s 2.92 × 103 10.67 × 106

the other PI + CI controller, but also has less overshoots and undershoots in such a
way that it gives the better response. In fact, this PI + CI controller response is also
much better than the PI-IMC response since it is faster and it has similar overshoots
and undershoots with lower performance indexes.
However, as it can be seen from the control signals, the control system just un-
dergoes the reset action once in each reference. Therefore, after the first and the only
reset instant, the control system behaves as the lineal base PI controller. In order to
improve further the system performance, a variable reset band will be considered in
the next section.

Variable Reset Band Here a variable reset band is considered to overcome the
influence of the dominant time delay over the reset action. For this purpose, the
approach described in Chap. 5 will be followed. In this case, the reset conditions
are also given when the error signal is approaching zero, but now the value of the
error when the reset action is done will change constantly from one reset instant to
another.
For a variable reset band to be applied, only a nominal value of the time delay has
to be fixed in the variable reset band. For the industrial heat exchanger studied in this
work, the time delay is uncertain, and several values of time delay can be chosen.
Theoretically, the reset instants are going to take place h time units before the time
the error signal crosses to zero. Therefore, in order not to degrade the response to
any reference change, the nominal time delay is taken as the minimum one, that is,
h = 50 s. In addition, due to the presence of sensor noise, the derivative part of this
reset band will be filtered by a low-pass filter with a constant time 10 times smaller
than the nominal time delay.
For comparison purposes, the same experiment explained previously is carried
out under the same operation conditions. In this case, the PI-IMC controller response
is compared to the PI + CI controller with variable reset band and two different reset
ratios, preset = 0.1 and preset = 0.4. In Fig. 6.12, both the responses and the control
signals are drawn for these three controllers. In addition, the IAE and ITAE values
are reported in Table 6.5.
From Fig. 6.12 and Table 6.5, it is deduced that the PI + CI controller with a
variable reset band and reset ratio equal to preset = 0.1 is the best option to control
the temperature increments in an industrial heat exchanger since, in comparison
with the PI controller tuned with IMC, a faster response is obtained with a similar
overshoot and undershoot. In this case, the PI + CI controller does several resets
rather than just one as with the fixed reset band. Therefore, it can be concluded that
224 6 Application Cases

Fig. 6.12 Responses and control signals of PI-IMC and PI + CI controllers with a variable reset
band

the variable reset band can be used to significantly improve the reset performance
and the linear IMC response when dominant time delays are present in the system
dynamics.

6.2 Teleoperation of a Gantry Crane

This section reports on the application of reset control to a real laboratory teleoper-
ated system consisting of a scale gantry crane teleoperated through the Internet by
means of a haptic device. The section is based on the works [21] and [22] where
a novel approach to passive teleoperation is presented that considers reset actions
appropriately applied. The motivation of these studies is to alleviate the limitations
arising from the existence of communication delays between the master and slave
sides.
Communication delays are problematic for networked controlled systems and
in particular for teleoperated systems. They give rise to a trade-off between speed
and robustness, which cannot be overcome by means of linear controllers. In or-
der to solve this problem, in this section we report a novel approach that combines
passivity-based techniques and reset control principles. In this way, it is possible to
obtain simultaneously the robust stability properties of passive control and the per-
formance improvement enabled by reset strategies. The experimental and simulation
results presented confirm the good behavior achieved with this method.
Bilateral teleoperation is an area that has motivated an enormous amount of re-
search in the last two decades [24]. Modern teleoperation is usually based on apply-
ing passivity-based control techniques [39] to teleoperated systems. Using passivity
6.2 Teleoperation of a Gantry Crane 225

tools has a number of advantages: they provide a multi-domain and modular ap-
proach to modeling, and transparent control principles based on energy and power
concepts. The passivity ideas can be adapted to the teleoperation framework, where
a human operator (actuating a master device) manipulates a remote machine (slave
device) which is possibly interacting with the environment. If some time delay ex-
ists, a communication channel transmitting velocity and force (which provides the
operator with a feeling of telepresence that is essential in applications [26, 29]) is no
longer passive. As an alternative, wave variables—which are obtained from veloci-
ties and forces via the scattering transformation, see [4, 34]—should be transmitted,
thus making the delayed line behave like an analog lossless LC line. This is called
line-passivation.
It turns out that after line-passivation, and with passive controllers at the mas-
ter and slave sides, the overall teleoperation system becomes passive. Since it is
assumed that the external elements—human operator and environment—behave in
a passive way, this means that the system is stable for all possible constant values
of the time delay. This delay-independent stability property can also be extended
to time-varying delays [18], and therefore it is possible to use the Internet as the
communication channel, as has already been shown in several applications [37, 44].
Delay-independent stability can be interpreted partly as an advantage and partly
as a drawback: on the one hand, it is an advantage because the designer does not
have to worry about the delay. On the other hand, it is a drawback because, if the
passive controllers are stable for very large delay values, they tend to provide worse
performance than other (non-passive) controllers, which are unstable for large de-
lays but stable for small nominal delays.
Realizing that the source of the problem is the time-delay, which introduces a
linear trade-off between speed and robustness, it becomes clear that using the reset-
control principle in teleoperation might enable us to combine the robustness of the
passive solutions with the fast position tracking performance provided by a proper
design of the reset action.
The section is organized as follows. Section 6.2.1 provides an overview of the
system, and the principles of passive teleoperation. Then, Sect. 6.2.2 shows how
the use of reset is capable of overcoming the fundamental limitations of linear con-
trollers, namely when time delays are present. The proposed reset strategy is pre-
sented and justified and the stability is proven from passivity properties. Finally, two
examples are presented in Sect. 6.2.3: a simulated simple robot and a real applica-
tion with a laboratory gantry crane teleoperated with a haptic device.

6.2.1 Passive Teleoperation and System Overview

The block diagram that describes the overall system is given in Fig. 6.13. It consists
of the following subsystems: a master device commanded by an operator, a mas-
ter impedance controller (represented as the parallel system of a spring Km and a
damper Bm ), the scattering transformations corresponding to master and slave sides,
226 6 Application Cases

Fig. 6.13 Teleoperation scheme

the communication channel, a slave impedance controller (Ks , Bs ) and a slave de-
vice in contact with the environment. The interchanged variables are either power
variables (forces fi and velocities ẋi , with i ∈ {h, m, s, e} and with the additional
subindex d for delayed variables); or wave variables (um , us , wm , ws ) which result
from the scattering transformation of the former.
The communication channel contains the time delays which will be considered
constant and known throughout this study. This is a frequent assumption when de-
riving theoretical results for the first time; these may be extended for time-varying
delays later. Here we restrict ourselves to constant time delays; only in the last ex-
ample of Sect. 6.2.3, and as a preliminary experimental validation, we apply the
method to unknown, time-varying delays. It is well-known that a time delay pro-
vokes a loss of passivity, which can be arranged using the scattering theory [4] or
the equivalent wave-variables [34] formulation given by the equations:
   √ −1/2  
um 2B √ −I f m
= (6.10)
ẋmd B −1 − 2B −1/2 wm

and
  √  
ws 2B −1/2 √ −I−1/2 fs
= , (6.11)
ẋsd −B −1 2B us
where the matrix B > 0 is the line impedance. The actually transmitted variables
are the wave variables:

us (t) = um (t − ΔT ), wm (t) = ws (t − ΔT ), (6.12)

where ΔT ≥ 0 is the communication delay, unknown but considered constant. This


configuration behaves as a lossless system with respect to the storage (Hamiltonian)
function

1 t  
Hc (t) = um (τ )2 + ws (τ )2 dτ, (6.13)
2 t−ΔT
that is, the integral of the power of the waves for the duration of the transmission. It
easily follows that

−fs (t)ẋsd (t) + fm (t)ẋmd (t) = Ḣc (t). (6.14)


6.2 Teleoperation of a Gantry Crane 227

This relation is true for all ΔT , so, from the point of view of passivity, we do not
have to be concerned about its actual value. The interconnection of passive subsys-
tems leads to an overall system that is itself passive [39], and hence stable behavior
is guaranteed.
As wave reflections can occur at both sides of a teleoperation scheme, there is a
need for impedance matching [34], which can be achieved if both sites are placed
under velocity control. This is one of the motivations for using the two symmet-
ric “impedance controllers” mentioned before, which are placed at each end of the
transmission channel. Thus, each side is receiving force information and providing
velocity signals.
The master side of the setup consists of a master device, its impedance controller,
and the corresponding scattering transformation. The master device interacts with
the human operator, allowing him to specify the desired movement of the plant
(velocity command), while receiving some force feedback representing information
about the operating conditions. Similarly, the slave side consists of the slave device,
its impedance controller, and the scattering transformation.
The master and slave devices will be considered as mechanical systems whose
equations can be written in the port-Hamiltonian notation as:
∂H (z)
ż = [J (z) − R(z)] + g(z)f,
∂z (6.15)
∂H (z)
ẋ = g T (z) ,
∂z
where z are the states, ẋ the output (velocities), f the input (forces), H the Hamilto-
nian function, and J, R the interconnection and dissipation structures, respectively.
Reproducing the previous equations with subindices m, s, we obtain the description
of the blocks ‘Master device’ and ‘Slave device’ in Fig. 6.13, additively including
the forces fh , fe .
Finally, the equations of the impedance controllers can be written as

fm = Km (ẋm − ẋmd ) dt + Bm (ẋm − ẋmd ),
(6.16)
fs = Ks (ẋsd − ẋs ) dt + Bs (ẋsd − ẋs ),

which are two Proportional-Integral (PI) controllers with transfer functions


 
Km
fm = + Bm em , em = ẋm − ẋmd ,
s
  (6.17)
Ks
fs = + Bs es , es = ẋsd − ẋs .
s
In this way, we are not only transmitting energy through the strings Km , Ks , but
also dissipating some of it in the dampers Bm , Bs . Consequently, artificial damping
is injected, which is a requisite to ensure a passive behavior when it is possible that
the plant interacts with the environment. The parameters of this controller are the
228 6 Application Cases

stiffness constants of the springs, Km and Ks , and the viscous frictions, Bm and Bs .
If the latter are chosen to be equal to the scattering parameter of the transmission
line, B, no wave reflections will take place [34].

6.2.2 Reset Procedure

In this section, a reset strategy for improving the performance of the impedance
controllers is presented. Recall that their transfer functions are given by (6.17), thus
defining them as Proportional-Integral (PI) controllers with respect to the velocity
error. The proposed reset procedure entails, to begin with, partially resetting the
integral value of the master PI, that is, the reset compensator is a PI + CI (see Chap.
5 for a detailed description). Thus, the overall master controller can be then written
in the form


⎪ ẋI = ẋCI = ẋm − ẋmd if c = 0,
⎨  x+   x 
+ = 0 if c = 0,
I I
(6.18)

⎪ xCI
 xI 

fm = Km [(1 − p) p] xCI + Bm em ,
with c the condition determining the reset instants (discussed later in (6.24)),
(xI , xCI ) the states of the linear and Clegg integrators, respectively, and (1 − p, p)
the parameters that multiply the integrator states.
The PI + CI controller can be implemented as two parallel integrators (see
Fig. 5.1, with kp = Bm , Km = kp /τi ), one of them being a Clegg integrator and
the other a normal one. The outputs of these two integrators are multiplied by p and
1 − p respectively, for 0 < p < 1. Thus, the modification of parameter p provides a
simple way of tuning the controller.
The reason for performing only “partial” reset is that a totally reset integrator
loses the fundamental property of eliminating the steady-state error. Since feedback
control is performed on the velocity signal, full reset leads to a loss of information
about the slave position that cannot be recovered. When this happens, position drift
appears, a well-known problem in bilateral teleoperation. However, when partial
reset is used, position information is being kept in the linear integrator (the non-
reset integrator placed in parallel to the reset one).
This can be more formally explained as follows. A steady state is achieved if
all the forces and velocities in Fig. 6.13 are zero. According to [17] and references
therein, the steady-state tracking-position error, referred to as “position drift”, is
zero if:
t
Δxm = (ẋm − ẋmd ) = 0, (6.19)
0

where t is some time at steady-state. Let us assume that the error is initially zero, that
is, Δxm (0) = 0. Now consider the master impedance controller in Fig. 6.13 replaced
6.2 Teleoperation of a Gantry Crane 229

Bm
by a PI + CI as in Fig. 5.1, with the equivalence kp = Bm , τi = Km , xi = xI , and
xir = xCI . The general equation for 0 < p < 1 is

y = fm = Bm (ẋm − ẋmd ) + Km (xI (1 − p) + xCI p), (6.20)


t t
where xI = 0 (ẋm − ẋmd ) dt and xCI = tk (ẋm − ẋmd ) dt, with tk being the last reset
time. Now, when in steady state, the controller has zero input (velocities) and zero
output (force). The problem is whether there may exist nonzero steady-state values
xI , xCI compatible with zero input and output. We will discuss three possible cases:
PI, CI, and PI + CI.
When a PI is used as a matching controller, the fact that the output is zero under
zero input yields (6.19) directly because Δxm corresponds to the state of the PI
controller. On the other hand, if a CI controller (this is, a PI + CI with p = 1) is used,
Δxm can be non-null under zero output and zero input due to the fact t that every reset
action is deleting the controller’s memory; null reset state entails tk (ẋm − ẋmd ) = 0,
where tk is the instant of the last reset action. The advantage of using a PI + CI with
p < 1 is that (6.19) is guaranteed. This is based on the fact that the reset state (xCI )
cannot maintain a non-null value under zero input. In principle, there may exist a
combination (xCI , xI ) of non-null states in (6.18) that yield zero output (fm ) with
zero input. However, a sustained zero input entails that the reset law holds and that
the reset state is made zero. Therefore, at steady-state, the linear state (xI ) has to be
zero, and since Δxm = xI , (6.19) holds and the steady-state tracking-position error
is zero, as in the linear (PI) case.
Hence, the PI + CI manages to improve the transient response while retaining
the zero steady-state error property. This entails a trade-off between transient and
steady state performance, and the parameter p offers an intuitive way of adjusting
it. This can be done by a simple trial-and-error procedure, decreasing its value from
p = 1 until the steady state performance is good enough, that is, with an overshoot
and settling time which are both small enough to be acceptable.
The option of resetting the master side controller and not the slave one is sup-
ported by several sets of simulations and experiments. In them, the effect of resetting
the master has been found to be much more important, while reset of the slave has
had little influence. The explanation is related to the fact that the signal round trip
(from the velocity specified by the operator ẋm to force feedback fm in Fig. 6.13)
is closed by the master controller forming an external loop around the plant (slave)
that contains the time-delays. On the other hand, the slave PI forms an inner loop not
affected by delay. As the expected performance improvement due to reset is more
effective under fundamental limitations (time-delays), the master controller, and not
the slave, is the one that should be reset. The value of p should be chosen so that it
is close enough to 1 to reduce overshoot, but not so high as to affect the capability
of eliminating the steady state position error.

6.2.2.1 Passivity of the PI + CI

We make use of the result that the parallel connection of two passive systems is
itself passive [39]. Since a PI + CI is the parallel connection of a PI compensator
230 6 Application Cases

and a CI compensator (see Fig. 5.1), its passivity results from the passivity of both
subsystems as follows. Without loss of generality, gains have been restructured in
order to show a simpler and easier to understand proof. First, passivity of a PI can
be proven with the classical result for linear systems [28]: a system with a trans-
fer function H (j ω) is passive if Real[H (j ω)] ≥ 0 for all ω ∈ R. The PI transfer
function is given by:
 
1
HPI (j ω) = k 1 +  , (6.21)
T jω
where T  = T
1−p and T is the integration time constant. Its real part is given by

Re[HPI (j ω)] = k ≥ 0. (6.22)

Therefore, the PI compensator is passive. On the other hand, using Proposi-


tion 3.10 (see also [14]), passivity of a full reset compensator results from the pas-
sivity of its base linear compensator. The CI base linear compensator is a linear in-
blc (j ω) = 1 . Since Real[H blc (j ω)] = 0 ≥ 0
tegrator and its transfer function is HCI jω CI
for all ω ∈ R, the linear integrator is passive, and therefore, using the above men-
tioned result, the full reset compensator CI is also passive.
Hence, it is immediately seen that the PI + CI compensator defined by (6.18) is
passive.

6.2.2.2 Reset Instants

The aforementioned passivity property is independent of the reset instants, provided


that there are a finite number of resets in any finite interval. This enables the designer
to choose the reset instants in the most convenient way for each application. In the
case of teleoperation with a known, constant delay, it is possible to profit from this
freedom by anticipating to this delay, in the following sense: the reset instants can be
chosen so that they correspond to the times when the velocity error is approaching
zero. The resulting strategy can be visualized in Fig. 6.14, where the evolution of the
velocity error, ev (t) = ẋm (t) − ẋmd (t), is plotted. By monitoring the velocity error
and its variation, the reset instants can be chosen as those satisfying the following
condition:
ev (t1 )
ėv (t1 ) = − ⇔ ev (t1 ) + ėv (t1 )ΔT = 0. (6.23)
ΔT
Hence, (6.23) defines the reset condition c = 0 in (6.18), which amounts to ap-
plying a PD control on the error signal ev (t). That is, with a slight abuse of notation,
the signal c(t) that triggers resetting in (6.18) is fixed to

c(t) = (1 + sΔT ) · ev (t). (6.24)

Thus, instead of waiting for the error to be zero, this situation is predicted and
reset is carried out in advance. This mechanism also holds if the slave interacts with
the environment: in case of a contact, the slave’s velocity is modified and the reset
instants are changed accordingly.
6.2 Teleoperation of a Gantry Crane 231

Fig. 6.14 Choosing the reset


instants

6.2.3 Examples

6.2.3.1 Example 1: Simple Robot

In order to show that the reset control action is useful in different applications, two
different teleoperated systems with two different approaches are chosen. First, we
choose a simple model of a one-dimensional robot used in many simulations of
bilateral teleoperation, consisting of a mass with friction. The “Master Device” and
“Slave Device” blocks depicted in Fig. 6.13 are then given by:
1
ẋm = (fh − fm ),
mr s + br (6.25)
1
ẋs = (fs − fe ).
mr s + br
The force exerted by the human operator, fh , is simulated as a sinusoidal force
command applied to the master device. The force fm produced by the master side
impedance controller is fed back to the master device, influencing its movement in
the same way as the operator command. Symmetrically, the force acting on the slave
device is the output of the impedance controller at the slave side.
Firstly, we will show simulation results that illustrate the potential of reset control
for eliminating overshoot. Two identical “robots” with mass mr = 1 kg and friction
br = 1 kg/s are considered at the master and slave sides. A constant time delay
of 0.5 s is present in both directions of the communications channel (round trip
delay is 1 s). The coefficients of the impedance controllers are Km = Ks = 10 kg/s2 ,
Bm = Bs = 10 kg/s, and the scattering factor is also set to b = 10 kg/s. This is
an aggressive tuning of the controllers, intended to provide fast transient response;
hence the high-gain values Km and Ks . In Fig. 6.15, the result of a maneuver in free
space is plotted, with the master trajectory as a solid line and the slave as a dashed
line. It can be noticed that the non-reset controller indeed manages to obtain this fast
transient, but at the cost of a large overshoot. A partial reset controller with p = 0.5
yields the same fast transient while achieving a 50% reduction in overshoot. Finally,
full reset (p = 1) eliminates overshoot almost completely (90–100% reduction).
However, the steady state behavior is not shown in these simulations since the
setpoint is always varying. We consider now a different maneuver, including contact
with the environment, where this steady state behavior can be evaluated. We use
the same master and slave devices as before, and place them initially at the same
232 6 Application Cases

Fig. 6.15 Teleoperation in free space: (solid lines) master device, (dashed) slave; (top) without
reset, (middle) with partial reset (p = 0.5), (bottom) full reset
6.2 Teleoperation of a Gantry Crane 233

position, xm (0) = xs (0) = 0 m; but now there is a wall in the slave location at xs =
0.5 m. A sinusoidal force with a frequency of 1 rad/s is applied by the simulated
operator at the master side (fh ), with an amplitude of 2 N during the first cycle
and of 5 N during the second and third. After 15 seconds, this force is made zero,
simulating that the operator stops holding the master device. The coefficients of the
impedance controllers are now Km = Ks = 15 kg/s2 , Bm = Bs = 20 kg/s, and the
scattering factor is also set to b = 20 kg/s. The evolution of the system is shown in
Fig. 6.16, for the no-reset, full-reset, and partial-reset cases (from top to bottom). If
no reset is performed, the plant exhibits considerable overshoot.
A reset strategy can be applied to solve this issue. Since the overall delay is
ΔT = 1, the reset instants satisfy ev + ėv = 0. In this case, a full reset strategy (see
Fig. 6.16, middle) and a partial reset strategy with p = 0.85 (Fig. 6.16, bottom)
achieve similar results as far as overshoot reduction is concerned: at the time of ≈ 3
seconds, the overshoot is eliminated; at the time of ≈ 6 seconds, it is reduced by
more than 50%; and at the time of ≈ 13 seconds, by more than 60%. In both cases,
when moving in free space, the overshoot is reduced and, due to the fact that we
are anticipating the time delay, the slave position tracks the master position more
accurately, so the shapes of the trajectories are more similar. The difference in per-
formance between both strategies is that with full reset there is an error in steady
state, which does not appear with partial reset: in this latter case, after the master de-
vice is released by the operator, this is pushed by the feedback force until it matches
the position of the slave device, thus eliminating the steady state position error. Ob-
taining simultaneously these two features (good tracking with reduced overshoot,
and no error in steady state) is possible due to the use of partial reset.
In Fig. 6.17, the force feedback to the user (fm ) is plotted for the no-reset (dashed
line) and reset cases (solid line). As expected, the use of reset decreases the magni-
tude of the force at the reset instants, resembling the physical equivalent of a charged
spring whose deformation is suddenly reduced.

6.2.3.2 Example 2: Gantry Crane

Now we apply the reset control scheme to the teleoperation of a gantry crane. A real
plant, Inteco’s 3DCrane model, was used as a slave device in the experiments. The
output to be controlled is the position of the cart moving along the x axis. Its move-
ment is affected by the oscillations of a payload hanging from a cable with constant
length. Unlike in the previous case, the master device is not similar to the slave
plant; instead, the desired velocity is directly indicated by the operator. Two dif-
ferent experiments will be carried out: in the first one, the operator is simulated in
order to obtain a completely repeatable maneuver; in the second one, the maneuver
is actually specified by a human operator.
Let us first examine the case where the operator’s command is simulated as a
velocity pulse with an amplitude of 0.2 m/s and which is 1.5 s long. The feedback
force has thus no effect on the velocity setpoint. A constant time delay of 0.25 s is
considered (round trip delay of 0.5 s). The controllers are aggressively tuned with
234 6 Application Cases

Fig. 6.16 Manoeuvre with contact: (solid lines) master device, (dashed) slave; (top) no reset,
(middle) full reset, (bottom) partial reset (p = 0.85)
6.2 Teleoperation of a Gantry Crane 235

Fig. 6.17 Feedback forces with partial reset and without reset

Fig. 6.18 Gantry crane, with and without partial reset (p = 0.95)

coefficients Km = Ks = 60 kg/s2 , Bm = Bs = b = 6 kg/s. The desired and achieved


positions are pictured in Fig. 6.18. It can be noticed that the reset action eliminates
the overshoot by 80%, while providing a fast rise time.
Next we perform a human-commanded teleoperation maneuver. The operator
handles again a SensAble PhanTom Omni haptic device, and the overall teleoper-
ation setup is depicted in Fig. 6.19. The controllers are tuned as before, but this
time the time delay is no longer constant. Instead, the use of a packet-switched net-
work such as the Internet is simulated, applying the passivity-maintaining solution
for signal reconstruction presented in [18] that was applied to crane teleoperation in
[20]. The delay follows a normal distribution, with an average value of 0.15 s for
each channel (master-to-slave and slave-to-master), a standard deviation of 0.01,
and a 5% packet loss. The reset controller is hence tuned to anticipate an expected
round trip delay of 0.3 s. In Fig. 6.20, it is shown that the overshoot is completely
eliminated.
In conclusion, this section has reported on the advantages of reset control for
overcoming fundamental limitations, namely time delays, in passive teleoperation.
236 6 Application Cases

Fig. 6.19 Setup for crane teleoperation

Fig. 6.20 Teleoperating a crane over the Internet: (solid lines) master device, (dashed) slave
(gantry crane); (top) no reset, (bottom) partial reset (p = 0.95)

The proposed method resets the state of the master impedance controller and bene-
fits from choosing the reset instants so that the system anticipates the communica-
tions channel time delay.
6.3 Control of Solar Collector Fields 237

The simulated and experimental results presented demonstrate the good perfor-
mance of the proposed scheme for different applications. The solution has shown its
usefulness even in the Internet-based teleoperation with time-varying delays.
An interesting research direction for future work is to consider the possibility of
using a time-varying parameter p, and its implications for passivity. An extension
for unknown, time-varying delays, is another natural line of research.

6.3 Control of Solar Collector Fields


This case study is motivated by a temperature control problem that is common to
distributed collector fields in solar plants [12, 13]. In this kind of system, it is im-
portant to maintain a constant outlet temperature of the thermal fluid which flows
through the collector field, in spite of the fact that the solar radiation varies dur-
ing the day. The way to achieve a constant outlet temperature is by the adjustment
of the fluid flow which will have to change substantially during operation, making
operation conditions change constantly. Due to this variation, the temperature con-
trol problem is hard to solve in solar plants by conventional means, for example,
using PID controllers with simple rules for tuning, or even autotuning. The reason
is simple: the controller has to be designed for fixed operation conditions in the
presence of potentially large uncertainty with respect to them. Since the dynamics
is highly nonlinear and uncertain, a PID simply designed to work efficiently under
fixed operation conditions usually degrades considerably in its performance when
these conditions are changed [11, 19].
In [19], QFT (Quantitative Feedback Theory) is used for the control of a dis-
tributed collector field, ACUREX field, located at the Plataforma Solar de Almeria,
Spain. QFT is a robust control technique that is especially well suited for control
problems with large plant uncertainty. A (very) brief summary is given in Chap. 5,
see [25] for a detailed exposition by Prof. Horowitz. The basic idea of [19] consists
of using a linear time invariant (LTI) model with parametric uncertainty which was
obtained by local linearization of the distributed collector field nonlinear dynam-
ics around different operation conditions, following the method developed in [8].
The result is a robust PID compensator that guarantees a specified worst-case be-
havior for all the operation conditions considered. In general, the robustness and
performance has to be balanced due to fundamental limitations of the resulting PID
compensator, this being the main motivation of the present case study, in which reset
control is investigated with the goal of improving performance without sacrificing
robustness.
In this section, the potentials of PI + CI compensation in improving robustness
and performance of QFT-based designed base compensators will be investigated. In
particular, the tuning rules developed in Chap. 5 for systems with time delays will be
used (see also [6, 7]). First, the uncertain model for the solar field is defined through
the uncertainty of two parameters, and then the QFT design method is applied using
the uncertain model in order to get a well-tuned PI + CI compensator. Finally, the
response of the ACUREX field is improved just by modifying the controller using a
variable reset percentage.
238 6 Application Cases

Fig. 6.21 ACUREX field


diagram

Fig. 6.22 Closed-loop


system

6.3.1 The Solar Collector Field

The ACUREX field is a distributed collector system located at the Plataforma So-
lar de Almeria, Spain. This distributed solar collector field consists of a series of
parabolic mirrors that reflect solar radiation onto a pipe where thermal oil is pumped
from the bottom of a storage tank. This oil flow gets heated while circulating, and
is returned to the top of the storage tank to be used in other industrial applications
such as in thermal electrical generators and in desalination plants [11]. The diagram
of this collector field can be seen in Fig. 6.21.
In a distributed collector field, the objective of the control system is to maintain
the outlet oil temperature of the field at a desired level in spite of disturbances,
such as the irregular solar irradiance levels and inlet oil temperature changes. For
control purposes, let us consider the feedback system given by Fig. 6.22, where P
is the uncertain collector field and R is the reset compensator. In addition, a series
feedforward controller is added to the control structure in order to reject measurable
disturbances, solar irradiation I , and inlet oil temperature Tin , as much as possible.
The uncertainty of the ACUREX field has been studied by taking the feedforward
controller as part of the system model. For different operating conditions, that is, for
different ranges of oil flow, the frequency response of this system (feedforward and
plant) was obtained. From these frequency response data, it can be stated that the
field exhibits a number of antiresonance modes (frequencies at which the magnitude
of the frequency response changes abruptly) within the control bandwidth (the main
6.3 Control of Solar Collector Fields 239

one located around 0.02 rad/s) [11]. Thus, the controller tuning should be careful to
avoid exciting these antiresonance modes.

6.3.2 PI + CI Design

Since the main goal of this work is to design a robust control system that guarantees
both stable and optimal operation for the ACUREX field, QFT will be used to design
the base LTI control system. Closed-loop stability and tracking specifications are
considered. In a second stage, this base compensator will be used to design the final
reset control system.

6.3.2.1 Base PI Tuning

In the design of the base LTI compensator, a PI structure will be used to satisfy some
design specifications. After that, the PI parameters will be used to design the PI + CI
compensator, a key issue being the election of a proper reset percentage. First of all,
plant templates are computed from the sets of frequency responses obtained in [11].
In this case, the templates are obtained for a wide set of working frequencies, from
0.001 up to 0.1 rad/s.
Secondly, boundaries are computed for each frequency from design specifica-
tions by using templates. Design specifications will be a (robust) phase margin of
9°, a settling time of 15 minutes (trying to avoid the excitation of the resonance
frequencies) and a maximum overshoot of 40%. Note that a relatively small phase
margin has been specified for the base LTI control system. This is due to the fact
that the reset PI + CI compensator will be designed to improve this minimum phase
margin.
The next step consists of shaping the nominal open loop. For this purpose, the
nominal plant Gn considered is the one obtained from the frequency response at
low flow conditions, the worst operating condition, when the field is more difficult
to control:
s2
0.83(1 + 0.0084 s + 0.00842 )
0.296
Gn (s) = e−20s . (6.26)
s2 s2
(1 + 1.6
0.0045 s + 0.00452
)(1 + 0.01 s + 0.012 )
0.82

Here the PI controller is tuned by making the open-loop gain fit both the stabil-
ity and tracking boundaries. A shaping is shown in Fig. 6.23, corresponding to a
proportional gain of kp = 1.80 and an integral time constant of τi = 150 s.
Finally, once the base PI compensator is tuned, a closed-loop system analysis is
performed to ensure that the system fits the given stability and tracking specifica-
tions. This is usually referred as to compensator validation in QFT. This analysis
has to be done for a wider set of frequencies than the one used in the templates
computation, and in this case, this set goes from 0.0001 up to 0.1 rad/s with a larger
240 6 Application Cases

Fig. 6.23 Boundaries and loop gain shaping

grid. In Fig. 6.24, the stability analysis is shown, whereas the tracking analysis is
depicted in Fig. 6.25.
As it can be seen, the designed closed-loop system satisfies both stability and
tracking specifications, for every frequency, thus the base PI compensator design is
finished.

6.3.2.2 PI + CI Tuning

Once the base PI compensator has been designed, the obtained PI parameters are
chosen as part of the PI + CI parameters. But, in addition, another parameter, the
reset percentage (preset ), must be tuned. For a proper selection of this parameter, it
is important to note that for high values of preset a worse closed-loop response can
be expected. Here, the tuning rules given in Chap. 5 (see also [6, 7]) are used for
tuning this parameter for the nominal loop. The result is an increment of the phase
margin and thus a better transient response for the nominal loop, and also for rest of
possible loops derived from the uncertainty due to the extra phase lead given by reset
compensation. A value preset = 0.35 is obtained. For comparison purposes, Bode
plots of the nominal loops with the plant (6.26) and both compensators: the base
PI compensator (kp = 1.80, τi = 150), and the PI + CI compensator (kp = 1.80,
τi = 150, and preset = 0.35) are shown in Fig. 6.26. The extra phase lead given by
6.3 Control of Solar Collector Fields 241

Fig. 6.24 Validation of stability specifications

Fig. 6.25 Validation of tracking specifications

the PI + CI compensator without significant change in the magnitude, in relation to


the base PI design can be seen in the figure.
Note that, although reset compensation by PI + CI can considerably improve the
phase margin of a PI controller, these improvements need to be carefully interpreted
since in this case the closed-loop system is nonlinear, and the results have been
242 6 Application Cases

Fig. 6.26 Bode plot of both PI and PI + CI controllers

obtained by using the describing function of the PI + CI system. To simulate the


ACUREX field dynamics, a simulation software package [10] and input data from
30 November 2007 are used. In Fig. 6.27, the reset effect over the distributed col-
lector field can be seen. It shows that both systems (for PI and PI + CI controllers)
are equal until the reset action is performed.
On the other hand, the control signal is shown in Fig. 6.28. This figure shows that
the distributed collector field operates at medium-high oil flows for both compen-
sators, in such a way that the antiresonance modes are not excited as they lie at high
frequencies.
Although the responses given in Fig. 6.27 fulfill the design specifications, it can
be seen that the effect of reset over the ACUREX field leads the system to have less
overshoot than in the PI case, but more undershoot when the reference temperature
is increased. Also when the reference temperature is decreased, the reset produces
more overshoot. In what follows, a modification of the reset PI + CI compensator
will be investigated to improve the control system performance.

6.3.3 PI + CI Compensator with Variable Reset Percentage

Reset action may be a good choice to improve the transient response of a closed-
loop system, but there are systems whose response improvement is not better than
the one with the linear PI controller, in fact, the response is actually worse. This is
the case of distributed collector fields. This is due to the fact that, in this kind of
systems, the dynamic characteristics of the field are constantly changing, and the
6.3 Control of Solar Collector Fields 243

Fig. 6.27 Field responses of both PI and PI + CI controllers

Fig. 6.28 Oil flow of both PI and PI + CI controllers

reset action can be very sensitive to these changes. In fact, in the PI + CI design,
the reset percentage has been tuned by fixing a low value, preset = 0.2, but from
244 6 Application Cases

Fig. 6.29 Field responses of PI + CI with both fixed and variable reset

Fig. 6.27, it can be seen that this fixed value of reset percentage has only improved
the field response at some reset times. To extend this response improvement over all
the reset times, the reset percentage should be variable.
In the following, the reset percentage is modified at each reset time using (5.47)
and (5.48). Consider the system responses given in Fig. 6.27, that is, the reference
temperature will experience the same step changes along the simulation. The con-
trol will be done by the PI + CI controller tuned in previous section, where its pa-
rameters were kp = 0.77, τi = 120, and preset = 0.2. In addition, another PI + CI
controller is used with the same parameters, but with a variable reset percentage as
in (5.47)–(5.48), where the chosen values are p̄reset = 0.05, τd = 0.55, and τf = 2.
The systems responses are shown in Fig. 6.29.
Comparing closed-loop responses in Fig. 6.29, it can be said that closed-loop
response is considerably improved by making the reset percentage variable since
both overshoots and undershoots are strongly reduced without decreasing the speed
of response. Finally, for comparison purposes, in Fig. 6.30, closed-loop responses
are plotted for the base PI controller and the already tuned PI + CI controller with
variable reset percentage.
As can be seen in Fig. 6.30, the PI + CI controller with a variable reset per-
centage makes the response to have less overshoot in comparison to the PI, without
modifying the response speed. Therefore, it can be concluded that a variable reset
percentage improves considerably the PI response.
References 245

Fig. 6.30 Field responses of PI and PI + CI with variable reset controllers

References
1. Aguilar, R., Martínez, S.A., Rodríguez, M.G., Soto, G.: Process analysis for treatment of
industrial plating wastewater: simulation and control approach. Chem. Eng. J. 105, 139–145
(2005)
2. Alpbaz, M., Karacan, S., Cabbar, Y., Hapoǧlu, H.: Application of model predictive control
and dynamic analysis to a pilot distillation column and experimental verification. Chem. Eng.
J. 88, 163–174 (2002)
3. Altinten, A., Ketevanlioǧlu, F., Erdoǧan, S., Hapoǧlu, H., Alpbaz, M.: Self-tuning PID control
of jacketed batch polystyrene reactor using genetic algorithm. Chem. Eng. J. 138, 490–497
(2008)
4. Anderson, R.J., Spong, M.W.: Bilateral control of teleoperators with time delay. IEEE Trans.
Autom. Control 34(5), 494–501 (1989)
5. Arbaoui, M.A., Vernières-Hassimi, L., Seguin, D., Abdelghani-Idrissi, M.A.: Counter-current
tubular heat exchanger: modeling and adaptive predictive functional control. Appl. Therm.
Eng. 27, 2332–2338 (2007)
6. Baños, A., Vidal, A.: Design of PI + CI reset compensators for second order plants. In: IEEE
International Symposium on Industrial Electronics, Vigo, Spain (2007)
7. Baños, A., Vidal, A.: Definition and tuning of a PI + CI reset controller. In: Proceedings of
the European Control Conference, Kos, Greece (2007)
8. Baños, A., Yaniv, O., Montoya, F.J.: Nonlinear QFT synthesis by local linearization. Int. J.
Control 76(5), 429–436 (2003)
9. Baños, A., Garcia, P., Checa, L.: Robust control of thermal treatments in can industry. In: Pro-
ceedings of the 4th IFAC/CIGR Workshop—Control Applications in Post-Harvest and Pro-
cessing Technology (CAPPT 2006), Potsdam, Germany (2006)
10. Berenguel, M., Camacho, E.F., Rubio, F.R.: Simulation software package for the ACUREX
field. Internal Report, Departamento de Ingeniería y Automática, ESI Sevilla, Spain (1994)
11. Camacho, E.F., Berenguel, M., Rubio, F.R.: Advanced Control of Solar Plants. Springer,
Berlin (1997)
246 6 Application Cases

12. Camacho, E.F., Rubio, F.R., Berenguel, M., Valenzuela, L.: A survey on control schemes for
distributed solar collector fields. Part I: Modeling and basic control approaches. Sol. Energy
81, 1240–1251 (2007)
13. Camacho, E.F., Rubio, F.R., Berenguel, M., Valenzuela, L.: A survey on control schemes for
distributed solar collector fields. Part II: Advanced control approaches. Sol. Energy 81, 1252–
1272 (2007)
14. Carrasco, J., Baños, A., van der Schaft, A.J.: A passivity-based approach to reset control sys-
tems stability. Syst. Control Lett. 59(1), 18–24 (2010)
15. Carrizales-Martínez, G., Fermat, R., González-Alvarez, V.: Temperature control via robust
compensation of heat generation: Isoparaffin/olefin alkylation. Chem. Eng. J. 125, 89–98
(2006)
16. Cervera, J.: Ajuste automático de controladores en QFT mediante estructuras fraccionales.
Ph.D. Thesis, University of Murcia, Spain (2006)
17. Chopra, N., Spong, M.W., Ortega, R., Barabanov, N.E.: On tracking performance in bilateral
teleoperation. IEEE Trans. Robot. 22(4), 861–866 (2006)
18. Chopra, N., Berestesky, P., Spong, M.W.: Bilateral teleoperation over unreliable communica-
tion networks. IEEE Trans. Control Syst. Technol. 16(2), 304–313 (2008)
19. Cirre, C.M., Moreno, J.C., Berenguel, M.: Robust QFT control of a solar collectors field. In:
Martínez, D. (ed.) Improving Human Potential Programme—Research Results at PSA Within
the Year 2002 Access Campaign. Serie Ponencias CIEMAT, pp. 27–33 (2003)
20. Fernández, A., Barreiro, A., Raimúndez, C.: Digital passive teleoperation of a gantry crane.
In: Proc IEEE Int. Symp. Ind. Electron., Vigo, Spain (2007)
21. Fernández, A.F., Barreiro, A., Baños, A., Carrasco, J.: Reset control for passive teleoperation
applications in process control. In: 34th Annual Conference of the IEEE Industrial Electronics
Society, Orlando, Florida, USA (2008)
22. Fernández, A.F., Barreiro, A., Baños, A., Carrasco, J.: Reset control for passive bilateral tele-
operation. IEEE Trans. Ind. Electron. (2010). doi:10.1109/TIE.2010.2077610
23. Fischer, M., Nelles, O., Isermann, R.: Adaptive predictive control of a heat exchanger based
on a fuzzy model. Control Eng. Pract. 6, 259–269 (1998)
24. Hokayem, P., Spong, M.W.: Bilateral teleoperation: an historical survey. Automatica 42(12),
2035–2057 (2006)
25. Horowitz, I.M.: Quantitative Feedback Theory. QFT Press, Boulder (1992)
26. Hyodo, S., Soeda, Y., Ohnishi, K.: Verification of flexible actuator from position and force
transfer characteristic and its application to bilateral teleoperation system. IEEE Trans. Ind.
Electron. 56(1), 36–42 (2009)
27. Karacan, S., Hapoǧlu, H., Alpbaz, M.: Application of optimal adaptive generalized predictive
control to a packed distillation column. Chem. Eng. J. 84, 389–396 (2001)
28. Khalil, H.K.: Nonlinear Systems. Prentice Hall, New York (2002)
29. Khan, S., Sabanovic, A., Nergiz, A.O.: Scaled bilateral teleoperation using discrete-time
sliding-mode control. IEEE Trans. Ind. Electron. 56(9), 3609–3618 (2009)
30. Landau, I.D.: System Identification and Control Design. Prentice Hall, New York (1990)
31. Lee, Y., Park, S., Lee, M., Brosilow, C.: PID controller tuning for desired closed-loop re-
sponses for SI/SO systems. AIChE J. 44(1), 106–115 (2004)
32. Loulch, Z., Cabassud, M., Le Lann, M.V.: A new strategy for temperature control of batch
reactors: experimental application. Chem. Eng. J. 75, 11–20 (1999)
33. Maidi, A., Diaf, M., Corriou, J.P.: Optimal linear PI fuzzy controller design of a heat ex-
changer. Chem. Eng. Process. 47, 938–945 (2008)
34. Niemeyer, G., Slotine, J.-J.: Stable adaptive teleoperation. Int. J. Ocean Eng. 16(1), 152–162
(1991)
35. Rivera, D.E., Morari, M., Skogestad, S.: Internal model control. 4. PID controller design. Ind.
Eng. Chem. Process Des. Dev. 25(1), 252–265 (1986)
36. Skogestad, S.: Simple analytic rules for model reduction and PID controller tuning. J. Process
Control 13(4), 291–309 (2003)
References 247

37. Slama, T., Trevisani, A., Aubry, D., Oboe, R., Kratz, F.: Experimental analysis of an internet-
based bilateral teleoperation system with motion and force scaling using a model predictive
controller. IEEE Trans. Ind. Electron. 55(9), 3290–3299 (2008)
38. Smith, C.L., Corripio, A.B., Martin, J.: Controller tuning from simple process models. In-
strum. Technol. 22(12), 39–44 (1975)
39. van der Schaft, A.J.: L2-Gain and Passivity Techniques in Nonlinear Control. Springer, Lon-
don (2000)
40. Vidal, A., Baños, A.: QFT-based design of PI + CI reset compensator: applications in process
control. In: 16th IEEE Mediterranean Conference on Control and Automation, Ajaccio, France
(2008)
41. Vidal, A., Baños, A.: Reset compensation for temperature control: experimental applications
on heat exchangers. Chem. Eng. J. 159(1–3), 170–181 (2010)
42. Vidal, A., Baños, A., Moreno, J.C., Berenguel, M.: PI + CI compensation with variable re-
set: application on solar collector fields. In: 34th Annual Conference of the IEEE Industrial
Electronics Society, Orlando, Florida, USA (2008)
43. Vidal, A., Vidal, A., Baños, A.: Reset compensation applied on industrial heat exchangers.
In: 14th IEEE International Conference on Emerging Technologies and Factory Automation,
Mallorca, Spain (2009)
44. Wirz, R., Marin, R., Ferre, M., Barrio, J., Claver, J.M., Ortego, J.: Bidirectional transport
protocol for teleoperated robots. IEEE Trans. Ind. Electron. 56(9), 3772–3781 (2009)
Index

A E
ACUREX field, 238 Existence and uniqueness of solutions, 151
After-reset surface, 58, 73
Anticipating reset, 178, 230 F
Feedback force, 233
B Feedforward, 238
Beating, 63 First order plants, 184
First order plus deadtime (FOPDT), 188, 215
C lag dominant, 190
Clegg integrator time-delay dominant, 193
definition, 6 FORE (first order reset element)
describing function, 17, 129 definition, 1
Horowitz’s design procedure, 20 describing function, 129
Clegg integrator with advanced reset/variable Horowitz’s design procedure, 23
reset band passivity, 119
definition, 17 FORE with reset band, 131
describing function, 18, 135 describing function, 132
Clegg integrator with reset band
describing function, 132 G
Crane, 233 General reset control system, 26
full reset, 26, 117
D partial reset, 26, 120
Deadlock, 63 with reset band, 130
Delay margin, 167 General reset controller, 7
Delay-dependent stability describing function, 134
frequency domain, 168 with reset band
time domain, 164 describing function, 132
Delay-independent stability
frequency domain, 155 H
time domain, 153 H-beta condition, 27, 156
Delayed differential equation, 150 generalized, 158
Delta (Dirac) function, 160 Half-rule, 195
Detuned PI (PI-des), 195 Haptic device, 235
Dirac impulse (delta), 169 Heat exchanger, 212
Dominant eigenvalue, 61 High-order systems, 195
Dwell time, 105 Hybrid systems, 37

A. Baños, A. Barreiro, Reset Control Systems, Advances in Industrial Control, 249


DOI 10.1007/978-1-4471-2250-0, © Springer-Verlag London Limited 2012
250 Index

I Position drift, 228


IAE (integral absolute error), 221 PRBS (pseudo random binary signal), 215
Impedance controllers, 227
Impulsive control system, 32 Q
Impulsive dissipative system, 35 Quantitative Feedback Theory (QFT), 208,
Impulsive systems, 27 218, 237
autonomous systems with impulses, 30 templates, boundaries, and loop shaping,
impulses at fixed times, 29 208
impulses at variable times, 29 templates, boundaries and loop shaping,
Input strictly passive (ISP), 116 239
Input–output stability, 113
Integrating systems, 198 R
Internal model control (IMC), 190, 216 Rectangular pulses, 172
Skogestad IMC (SIMC), 190 Reduced initial condition, 61, 100
ITAE (integral time absolute error), 221 Regular reset system, 100
Reset band, 129
J Reset lag compensator, 119
Jordan form, 61 Reset percentage (ratio), 183, 240
variable, 206, 242
K Reset PID compensator, 120
Kalman–Yakubovich–Popov (KYP) lemma, Reset PII R compensator, 122
155 Reset surface, 58
Reset times, 2
L Robustness, 174
Limit cycles, 136
Linear base system, 59 S
Linear fundamental limitations, 4 SCADA, 213
frequency domain, 11 Scattering, 226
time domain, 9 Small gain, 159, 174
Linear matrix inequality (LMI), 154, 166 Solar collector fields, 237
Lyapunov stability Spectral projector, 61
definitions, 94 Stabilization by reset, 105
H-beta condition, 95 stable base system, 105
reset-times dependent conditions, 99 unstable base system, 110
reset-times independent conditions, 94 Storage functional, 168
Lyapunov–Krasovskii (LK) functionals, 153, Strictly positive real (SPR), 157, 174
165, 175
T
O Teleoperation, 224
Output strictly passive (OSP), 116, 157, 174 passive, 225
Temperature control, 211
P Time regularization, 59, 151
Passive system, 116
pH control, 125 V
PI+CI, 181 Variable reset, 184
design for a heat exchanger, 216 Variable reset band, 134
design for solar collector field, 239 Very strictly passive (VSP), 116
design for teleoperation, 228
summary of tuning rules, 202 W
variable reset band/advanced reset, 205, Well-posedness, 63
223
with fixed reset band, 203, 222 Z
Port-Hamiltonian systems, 227 Zeno solutions, 63, 67
Other titles published in this series (continued):

Soft Sensors for Monitoring and Control Real-time Iterative Learning Control
of Industrial Processes Jian-Xin Xu, Sanjib K. Panda
Luigi Fortuna, Salvatore Graziani, and Tong H. Lee
Alessandro Rizzo and Maria G. Xibilia
Deadlock Resolution in Automated
Adaptive Voltage Control in Power Systems Manufacturing Systems
Giuseppe Fusco and Mario Russo ZhiWu Li and MengChu Zhou
Advanced Control of Industrial Processes Model Predictive Control Design
Piotr Tatjewski and Implementation Using MATLAB®
Process Control Performance Assessment Liuping Wang
Andrzej W. Ordys, Damien Uduehi Predictive Functional Control
and Michael A. Johnson (Eds.) Jacques Richalet and Donal O’Donovan
Modelling and Analysis of Hybrid Fault-tolerant Flight Control
Supervisory Systems and Guidance Systems
Emilia Villani, Paulo E. Miyagi Guillaume Ducard
and Robert Valette Fault-tolerant Control Systems
Process Control Hassan Noura, Didier Theilliol,
Jie Bao and Peter L. Lee Jean-Christophe Ponsart and Abbas
Chamseddine
Distributed Embedded Control Systems
Detection and Diagnosis of Stiction
Matjaž Colnarič, Domen Verber
and Wolfgang A. Halang in Control Loops
Mohieddine Jelali and Biao Huans (Eds.)
Precision Motion Control (2nd Ed.)
Stochastic Distribution Control
Tan Kok Kiong, Lee Tong Heng
System Design
and Huang Sunan
Lei Guo and Hong Wang
Optimal Control of Wind Energy Systems Dry Clutch Control for Automotive
Julian Munteanu, Antoneta luliana Bratcu, Applications
Nicolaos-Antonio Cutululis and Emil
Pietro J. Dolcini, Carlos Canudas-de-Wit
Ceangǎ
and Hubert Béchart
Identification of Continuous-time Models Advanced Control and Supervision
from Sampled Data of Mineral Processing Plants
Hugues Gamier and Liuping Wang (Eds.) Daniel Sbárbaro and René del Villar (Eds.)
Model-based Process Supervision Active Braking Control Design for Road
Arun K. Samantaray and Belkacem Vehicles
Bouamama Sergio M. Savaresi and Mara Tanelli
Diagnosis of Process Nonlinearities and Active Control of Flexible Structures
Valve Stiction Alberto Cavallo, Giuseppe de Maria,
M.A.A. Shoukat Choudhury, Sirish L. Shah Ciro Natale and Salvatore Pirozzi
and Nina F. Thornhill
Induction Motor Control Design
Magnetic Control of Tokamak Plasmas Riccardo Marino, Patrizio Tomei
Marco Ariola and Alfredo Pironti and Cristiano M. Verrelli
Fractional-order Systems and Controls Wastewater Systems
Concepcion A. Monje, YangQuan Chen, Carlos Ocampo-Martinez
Blas M. Vinagre, Dingyu Xue and Vincente Tandem Cold Metal Rolling Mill Control
Feliu John Pitter and Marwan A. Simaan
Model Predictive Control of Wastewater
Systems
Carlos Ocampo-Martinez

Anda mungkin juga menyukai