Synergetics Introduction and Advanced Topics, Haken, 2004
Synergetics Introduction and Advanced Topics, Haken, 2004
Synergetics
Introduction and Advanced Topics
, Springer
Professor Dr. Dr. h.c. multo Hermann Haken
Institute for Theoretical Physics I
Center of Synergetics
University of Stuttgart
PfaffenwaIdring 57/IV
70550 Stuttgart, Germany
This book appeared originally in two volumes in the series "Springer Series in Synergetics":
Synergetics, 3rd Edition (1983)
Advanced Synergetics, 1st Edition (1983)
This book is a reprint edition that comprises two titles, namely "Synergetics. An
Introduction. Nonequilibrium Phase Transitions and Self-Organization in Physics,
Chemistry and Biology" and ''Advanced Synergetics. Instability Hierarchies of Self-
Organizing Systems and Devices". The reason for this publication is two-fold:
Since synergetics is a new type of interdisciplinary field, initiated by the author
in 1969, the basic ideas developed in these volumes are of considerable theoretical
interest. But much more than this, the methods and even the concrete examples
presented in these books are still highly useful for graduate students, professors, and
even for researchers in this fascinating field. The reason lies in the following facts:
Synergetics deals with complex systems, i.e. systems that are composed of many
individual parts that are able to spontaneously form spatial, temporal or functional
structures by means of self-organization. Such phenomena occur in many fields
ranging from physics, chemistry and biology to economy and sociology. More
recent areas of application have been found in medicine and psychology, where
the great potential of the basic principles of synergetics can be unearthed. Further
applications have become possible in informatics, for instance the designing of new
types of computers, and in other fields of engineering. The central question asked by
synergetics is: Are there general principles of self-organization irrespective of the
nature of the individual parts systems are composed of?
Indeed, the original books "Synergetics" and "Advanced Synergetics" provide
the reader with a solid knowledge of basic concepts, theoretical tools and mathemat-
ical approaches to cope with the phenomenon of self-organization from a unifying
point of view. Synergetics takes into account deterministic processes as treated in
dynamic systems theory including bifurcation theory, catastrophy theory, as well
as basic notions of chaos theory and develops its own approaches. Equally well
it takes into account stochastic processes that are indispensable in many appli-
cations especially when qualitative changes of systems are treated. An important
special case refers to non-equilibrium phase-transitions that were originally treated
in physics.
The longevity of these two books derives from the timelessness of mathematics.
But also the explicit examples of applications to various fields are still of a paradig-
matic character. The wide applicability of the results of synergetics stems from the
fact that close to instability points were systems change their spatio-temporal pat-
terns or functions dramatically, similar behavior may be demonstrated in a great
VI Preface
Physics:
Spatio-temporal patterns (Staliunas, Sanchez-Morcillo, Gaul (2003», (Rosanov
(SSyn 2002» and stochastic properties of laser light, of fluids (Xu (SSyn 1997»
and plasmas, of currents in semiconductors (Scholl (2001» and of active lattices
(Nekorkin and Velarde (SSyn 2002».
Chemistry:
Formation of spatio-temporal patterns at macroscopic scales in chemical reactions.
for example the Belousov-Shabotinsky reaction (Babloyantz (1986). Nicolis (1995),
Mikhailov (SSyn 1994), Mikhailov and Loskutov (SSyn 1996», also the chemistry
of flames.
Preface VII
Computer science:
Synergetic computer for pattern recognition and decision making (Haken (SSyn
2004». This type of computer represents a genuine alternative to the by now tradi-
tional neuro-computers, to e.g. the Hopfield net.
Traffic science:
This field has become a truly interdisciplinary enterprise. Here typical synergetic
phenomena can be discovered such as phase transitions in traffic flows (Helbing
(1997».
Biology:
Morphogenesis:
Based on Turing's ideas, synergetics calculates spatial density distributions of
molecules, in particular gradients, stripes, hexagons, etc. as a function of bound-
ary and initial conditions. In initially undifferentiated omnipotent cells molecules
are produced as activators or inhibitors that diffuse between cells and react with
each other and thus can be transformed. At places of high concentration the activator
molecules switch on genes that, eventuaIly, lead to ceIl differentiation (Haken (this
book), Murray (2002».
Evolution:
By means of synergetics new kinds of analogies between biological and physical
systems have been unearthed. For instance, equations established by Eigen (Eigen
and Schuster (1979» for prebiotic, i.e. molecular evolution, turn out to be isomorphic
to specific rate equations for laser light (photons), where a specific kind of photon
wins the competition between different kinds of photons.
Population dynamics:
Resources, such as food, nesting places for birds, light intensity for plants, etc.
serve as control parameters. The numbers or densities of the individuals of species
serve as order parameters. Specific examples are provided by the Verhulst equation
or the preditor-prey relation of the Lotka-Volterra equations. Of particular interest
are dramatic changes, for instance the dying out of species under specific control
parameter values. This has influences on environmental policy. If specific control
parameters exceed critical values, the system's behavior can change dramatically.
For instance beyond a specific degree of pollution, the fish population of a lake wilI
die out.
Rhythms:
Nearly all biological systems show more or less regular periodic oscillations or fluc-
tuations. These can be imposed on the system from the outside, for instance by the
day /night cycle (circadian rhythms), seasons, etc. (exogenous), or can be produced
by the system itself (endogenous). In the foreground of synergetics research are
VIII Preface
endogenous rhythms that may proceed on quite different spatial and temporal scales
(Haken and Koepchen (SSyn 1992), Mikhailov and Calenbuhr (SSyn 2002)). Exam-
ples are cell metabolism, the circadian rhythms, brain waves in different frequency
bands (see below), cycles of menstruation, and rhythms in the cardiovascular system
(Stefanovska (2002)).
Movement science:
Rhythmical movements of humans and animals show well defined patterns of
coordination of the limbs, for instance walking, running of humans or gaits of
quadrupeds. Synergetics studies in particular transitions between movement pat-
terns, for instance the paradigmatic experiment by Kelso (Kelso (1995)). If subjects
move their index fingers parallel at a low movement frequency, suddenly at an
increased frequency an abrupt involuntary transition to a new symmetric move-
ment occurs. The control parameter is the prescribed finger movement frequency,
the order parameter is the relative phase between the index fingers (Haken (SSyn
1996)). The experimentally proven properties of a nonequilibrium phase transition
(critical fluctuations, critical slowing down, hysteresis) substantiate the concept of
self-organization and exclude that of a fixed motor program. Numerous further
coordination experiments between different limbs can be represented by the Haken-
Kelso-Bunz model (Kelso (1995), Haken (SSyn 1996)). Gaits of quadrupeds and
transitions between them were modelled in detail (Schoner et al. (1990)), see also
(Haken (SSyn 1996)). These results have lead to a paradigm change in movement
science.
Visual perception:
The recognition of patterns, e.g. of faces, is interpreted as the action of an associative
memory in accordance with usual approaches. Here incomplete data (features) with
which the system is provided from the outside are complemented by means of data
stored in the memory. A particular aspect of the synergetic approach is the idea that
pattern recognition can be conceived as pattern formation. This is not only meant as
a metaphor, but means also that specific activity patterns in the brain are established.
In pattern formation a partly ordered pattern is provided to the system, whereby
several order parameters are evoked that compete with each other dynamically. The
control parameters are so-called attention parameters that in cases without bias are
assumed to be equal. The winning order parameter imposes the total pattern on
the system according to the slaving principle. This process is also the basis of the
synergetic computer developed by Haken (Haken (SSyn 1995)).
Gestalt psychology:
As is shown in Gestalt psychology, Gestalt is conceived as an entity to which in
synergetics an order parameter with its synergetic properties (slaving principle!) can
be attached. In principle, the recognition process of Gestalt proceeds according to
the synergetic process of pattern recognition. The winning order parameter gener-
ates, according to the slaving principle, an ideal percept that is the corresponding
Preface IX
Psychology:
According to the concept of synergetics, psychological behavioral patterns are gen-
erated by self-organization of neuronal activities under specific control parameter
conditions and are represented by order parameters. In important special cases, the
order parameter dynamics can be represented as the overdamped motion of a ball in
a mountainous landscape. By means of changes of control parameters, this landscape
is deformed and allows for new equilibrium positions (stable behavioral patterns).
This leads to new approaches to psychotherapy: destabilization of unwanted behav-
ioral patterns by means of new external conditions, new cognitive influences, etc.
and measures that support the self-organization of desired behavioral patterns. This
comprises also the administration of appropriate drugs (e.g. neuroleptica), that in
the sense of Synergetics act as control parameters. The insights of synergetics have
been applied in the new field of Psychosynergetics with essential contributions by
Schiepek, Tschacher, Hansch, Ciompi and others (Schiepek (1999), Tschacher et al.
(SSyn), Hansch (2002), Ciompi (1982».
Brain theory:
Several books of the Springer Series in Synergetics are devoted to this field as well
as to experiments (KrUger (SSyn 1991), Ba§ar (SSyn 1998), Ba§ar et al. (SSyn
1983), Uhl (SSyn 1998), Tass (SSyn 1999), Haken (SSyn 1995, 2002». Also H.R.
Wilsons's fine book (1999) deserves attention in this context.
According to a proposal by Haken, the brain of humans and animals is conceived
as a synergetic, i.e. self-organizing system. This concept is supported by experiments
and models on movement coordination, visual perception, Gestalt psychology and
by EEG and MEG analysis (see below). The human brain with its 1011 neurons (and
glia cells) is a highly interconnected system with numerous feedback loops. In order
to treat it as a synergetic system, control parameters and order parameters must be
identified. While in synergetic systems of physics, chemistry and partly biology, the
control parameters are fixed from the outside, for instance by the experimentor, in
the brain and in other biological systems the control parameters can be fixed by the
system itself. In modelling them it is assumed, however, that they are practically
time-independent during the self-organization process. Such control parameters can
be, among others, the synaptic strengths between neurons that can be changed
X Preface
Sociology:
Here we may distinguish between the more psychologically and the more systems
theory oriented schools, where synergetics belongs to the second approach. We can
distinguish between a qualitative and a quantitative synergetics (For a quantitative
synergetics see Weidlich (2000)). In the previous case, a number of sociologically
relevant order parameters are identified. Examples are: the language of a nation:
After his/her birth, the baby is exposed to the corresponding language and learns
it (in the technical terms of synergetics: the baby is enslaved) and then carries on
with this language when having become an adult (circular causality). These language
order parameters may compete, where one wins (for example in the USA the English
language), they may coexist (for example in Switzerland), or they may cooperate (for
instance popular language and technical language ). Whereas in this case the action of
the slaving principle is evident, in the following examples its applicability is critically
discussed by sociologists so that instead of slaving some of them like to speak of
binding or consensualization. Corresponding order parameters are: type of state
(e.g. democracy, dictatorship), public law, rituals, corporate identity, social climate
in a company, and ethics. The latter example is particularly interesting, because order
parameters are not prescribed from the outside or ab initio, but originate through
self-organization and need not be uniquely determined.
Epistemology:
An example for order parameters (though not articulated that way) is provided by the
scientific paradigms in the sense of Thomas S. Kuhn (Kuhn (1996)), where a change
of paradigm has the properties of a nonequilibrium phase transition, such as critical
fluctuations and critical slowing down. Synergetics as a new scientific paradigm is
evidently self-referential. It explains its own origin.
Management:
The concept of self-organization is increasingly used in management theory and
management praxis (e.g. Ulrich and Probst (SSyn 1984)). Instead of fixed order
structures with many hierarchical levels, now flat organisational structures with only
few hierarchical levels are introduced. In the latter case a hierarchical level makes
its decisions by means of its distributed intelligence. For an indirect steering of
these levels by means of a higher level, specific control parameters in the sense of
synergetics must be fixed, for instance by fixing special conditions, goals, etc. The
order parameters are, for instance, the self-organized collective labor processes. In
this context, the slaving principle according to which the order parameters change
slowly, whereas the enslaved parts react quickly (adaptability), gains a new interpre-
tation. For instance, the employees that are employed for a longer time determine
the climate of labor, the style of work, etc., whereby it can also be possible that
undesired cliques are established. This trend can be counteracted by job rotation.
Development of cities:
Whereas so far the development of cities was based on the concept of city planning
with detailed plans for areas, new approaches use concepts of self-organization
XII Preface
References
1. Staliunas, K., Sanchez-Marcillo, v., Gaul, LJ.: Transverse Patterns in Nonlinear Optical
Resonators, Berlin: Springer (2003)
2. Scholl, E.: Nonlinear Spatio-Temporal Dynamics and Chaos in Semiconductors. Cam-
bridge: Cambridge University Press (2001)
3. Babloyantz, A.: Molecules, Dynamics, and Life: An Introduction to Self-Organization
of Matter. Indianapolis: Wiley (1986)
4. Nicolis, G.: Introduction to Nonlinear Science. Cambridge University Press (1995)
5. Helbing, D.: Verkehrsdynamik. Neue physikalische Modellierungskonzepte. Berlin:
Springer (1997)
6. Murray, J.D.: Mathematical Biology. Berlin: Springer (2002)
7. Eigen, M., Schuster, P.: The Hypercycle - A Principle of Natural Self-Organization.
Berlin: Springer (1979)
8. Stefanovska, A: Cardiorespiratory Interactions. Nonlinear Phenom. Complex Syst. 5,
462-469 (2002)
9. Kelso, J.AS.: Dynamic Patterns: The Self-Organization of Brain and Behavior. Cam-
bridge, MA: MIT Press (1995)
10. Schoner, G., Yiang, w.Y., Kelso, lAS.: A Synergetic Theory of Quadrupedal Gaits and
Gait Transitions. J. Theor. BioI. 142, 359-391 (1990)
11. Kohler, W.: Die physischen Gestalten in Ruhe und im stationiiren Zustand. Braun-
schweig: Vieweg (1920)
Preface XIII
12. Kohler, w.: Direction of Processes in Living Systems. Scientific Monthly 8, 29-32
(1955)
13. Schiepek, G.: Die Grundlagen der Systemischen Therapie; Theorie-Praxis-Forschung.
Gottingen: Vandenhoek and Ruprecht (1999)
14. Hansch, D.: Evolution und Lebenskunst. Grundlagen der Psychosomatik. Ein Selbst-
management-Lehrbuch. Gottingen: Vandenhoeck und Ruprecht (2002)
IS. Ciompi, L.: Affektlogik. Uber die Struktur der Psyche und ihre Entwicklung. Ein Beitrag
zur Schizophrenieforschung. Stuttgart: Klett-Cotta (1982)
16. Tass, P.A. et al.: Synchronization Tomography: A Method for Three-Dimensional Lo-
calization of Phase Synchronized Neuronal Populations in the Human Brain using
Magnetoencephalography. Phys. Rev. Letters 90, 88101-1-88101-4 (2003)
17. Wilson, H.R.: Spikes, Decisions and Actions. Dynamical Foundations of Neuroscience.
Oxford: Oxford University P.ress (1999)
18. Hebb, D.O.: The Organization of Behavior: A Neuropsychological Theory. New York:
Wiley (1949)
19. Weidlich, w.: Sociodynarnics. A Systematic Approach to Mathematical Modelling in
the Social Sciences. Amsterdam: Harwood Academic Publishers (2000)
20. Kuhn, T.S.: The Structure of Scientific Revolutions. University of Chicago Press, 3rd.
ed. (1996)
Contents
Part I An Introduction
An Introduction
Over the past years the field of synergetics has been mushrooming. An ever-
increasing number of scientific papers are published on the subject, and numerous
conferences all over the world are devoted to it. Depending on the particular
aspects of synergetics being treated, these conferences can have such varied titles
as "Nonequilibrium Nonlinear Statistical Physics," "Self-Organization," "Chaos
and Order," and others. Many professors and students have expressed the view
that the present book provides a good introduction to this new field. This is also
reflected by the fact that it has been translated into Russian, Japanese, Chinese,
German, and other languages, and that the second edition has also sold out. I
am taking the third edition as an opportunity to cover some important recent
developments and to make the book still more readable.
First, I have largely revised the section on self-organization in continuously
extended media and entirely rewritten the section on the Benard instability. Sec-
ond, because the methods of synergetics are penetrating such fields as eco-
nomics, I have included an economic model on the transition from full employ-
ment to underemployment in which I use the concept of nonequilibrium phase
transitions developed elsewhere in the book. Third, because a great many papers
are currently devoted to the fascinating problem of chaotic motion, I have added
a section on discrete maps. These maps are widely used in such problems, and can
reveal period-doubling bifurcations, intermittency, and chaos.
Instability hierarchies, which are dealt with in this book in the context of the
laser, have proved since its writing to be a widespread phenomenon in systems
driven far from thermal equilibrium. Because a general treatment of this problem
would have gone beyond the scope and level of the present monograph, however,
I have written a further book entitled Advanced Synergetics *, which can be
considered a continuation of the present text.
I wish to thank my co-worker Dr. A. Wunderlin for his assistance in revising
this book, and especially for his efforts in the Benard problem. I am also greatly
indebted to my secretary, Mrs. U. Funke, for her efficient help in reworking
several chapters.
* Springer Ser. Synergetics, Vol. 20 (Springer, Berlin, Heidelberg, New York, Tokyo \983)
Preface to the Second Edition
The publication of this second edition was motivated by several facts. First of all,
the first edition had been sold out in less than one year. It had found excellent
critics and enthusiastic responses from professors and students welcoming this
new interdisciplinary approach. This appreciation is reflected by the fact that the
book is presently translated into Russian and Japanese also.
I have used this opportunity to include some of the most interesting recent
developments. Therefore I have added a whole new chapter on the fascinating
and rapidly growing field of chaos dealing with irregular motion caused by
deterministic forces. This kind of phenomenon is presently found in quite diverse
fields ranging from physics to biology. Furthermore I have included a section on
the analytical treatment of a morphogenetic model using the order parameter
concept developed in this book. Among the further additions, there is now a com-
plete description of the onset of ultrashort laser pulses. It goes without saying that
the few minor misprints or errors of the first edition have been corrected.
I wish to thank all who have helped me to incorporate these additions.
scientists that numerous such systems show striking similarities in their behavior
when passing from the disordered to the ordered state. This strongly indicates
that the functioning of such systems obeys the same basic principles. In our book
we wish to explain such basic principles and underlying conceptions and to
present the mathematical tools to cope with them.
This book is meant as a text for students of physics, chemistry and biology
who want to learn about these principles and methods. I have tried to present
mathematics in an elementary fashion wherever possible. Therefore the knowl-
edge on an undergraduate course in calculus should be sufficient. A good deal of
important mathematical results is nowadays buried under a complicated nomen-
clature. I have avoided it as far as possible through, of course, a certain number
of technical expressions must be used. I explain them wherever they are introduced.
Incidentally, a good many of the methods can also be used for other problems,
not only for self-organizing systems. To achieve a self-contained text I included
some chapters which require some more patience or a more profound mathematical
background of the reader. Those chapters are marked by an asterisk. Some of them
contain very recent results so that they may also be profitable for research workers.
The basic knowledge required for the physical, chemical and biological
systems is, on the average, not very special. The corresponding chapters are
arranged in such a way that a student of one of these disciplines need only to read
"his" chapter. Nevertheless it is highly recommended to browse through the other
chapters just to get a feeling a how analogous all these systems are among each
other. I have called this discipline "synergetics". What we investigate is the joint
action of many subsystems (mostly of the same or of few different kinds) so as to
produce structure and functioning on a macroscopic scale. On the other hand,
many different disciplines cooperate here to find general principles governing
self-organizing systems.
I wish to thank Dr. Lotsch of Springer-Verlag who suggested writing an
extended version of my article "Cooperative phenomena in systems far from
thermal equilibrium and in nonphysical systems", in Rev. Mod. Phys. (1975). In
the course of writing the "extension", eventually a completely new manuscript
evolved. I wanted to make this field especially understandable to students of
physics, chemistry and biology. In a way, this book and my previous article have
become complementary.
It is a pleasure to thank my colleagues and friends, especially Prof.
W. Weidlich, for many fruitful discussions over the years. The assistance of my
secretary, Mrs. U. Funke, and of my coworker Dr. A. Wunderlin was an enor-
mous help for me in writing this book and I wish to express my deep gratitude to
them. Dr. Wunderlin checked the formulas very carefully, recalculating many of
them, prepared many of the figures, and made valuable suggestions how to
improve the manuscript. In spite of her extended administrative work, Mrs. U.
Funke has drawn most of the figures and wrote several versions of the manu-
script, including the formulas, in a perfect way. Her willingness and tireless efforts
encouraged me agian and again to complete this book.
7
Contents
1. Goal
1.1 Order and Disorder: Some Typical Phenomena 1
1.2 Some Typical Problems and Difficulties . 12
1.3 How We Shall Proceed . . . . . . . . . . . 15
2. Probability
2.1 Object of Our Investigations: The Sample Space 17
2.2 Random Variables 19
2.3 Probability.......... 20
2.4 Distribution . . . . . . . . . 21
2.5 Random Variables with Densities 24
2.6 Joint Probability. . . . . . . . 26
2.7 Mathematical Expectation E (X), and Moments 28
2.8 Conditional Probabilities . . . . . . . . . . 29
2.9 Independent and Dependent Random Variables 30
2.10 * Generating Functions and Characteristic Functions. 31
2.11 A Special Probability Distribution: Binomial Distribution 33
2.12 The Poisson Distribution . . . . . . . . . . . 36
2.13 The Normal Distribution (Gaussian Distribution) 37
2.14 Stirling's Formula . . 39
2.15 * Central Limit Theorem . . . . . . . . . . . . 39
3. Information
3.1 Some Basic Ideas 41
3.2 * Information Gain: An Illustrative Derivation 46
3.3 Information Entropy and Constraints 48
3.4 An Example from Physics: Thermodynamics 53
3.5* An Approach to Irreversible Thermodynamics 57
3.6 Entropy-Curse of Statistical Mechanics? 66
4. Chance
4.1 A Model of Brownian Movement . . . . . . . . . . . . 69
4.2 The Random Walk Model and Its Master Equation 75
4.3 * Joint Probability and Paths. Markov Processes. The Chapman-
Kolmogorov Equation. Path Integrals . . . . . . . . . 79
* Sections with an asterisk in the heading may be omitted during a first reading.
XII Contents
5. Necessity
5.1 Dynamic Processes. . . . . . . . . . . . . . . . . . . 105
5.2 * Critical Points and Trajectories in a Phase Plane. Once Again
Limit Cycles . . . . . . . . . . . . . . . . . 113
5.3 * Stability . . . . . . . . . . . . . . . . . . . 120
5.4 Examples and Exercises on Bifurcation and Stability 126
5.5 * Classification of Static Instabilities, or an Elementary
Approach to Thorn's Theory of Catastrophes 133
7. Self-Organization
7.1 Organization................. 191
7.2 Self-Organization . . . . . . . . . . . . . . . 194
7.3 The Role of Fluctuations: Reliability or Adaptibility?
Switching. . . . . . . . . . . . . . . . . . . 200
7.4 * Adiabatic Elimination of Fast Relaxing Variables from the
Fokker-Planck Equation . . . . . . . . . . . . . . . 202
7.5 * Adiabatic Elimination of Fast Relaxing Variables from the
Master Equation . . . . . . . . . . . . . . . . 204
7.6 Self-Organization in Continuously Extended Media. An
Outline of the Mathematical Approach . . . . . . . 205
10
Contents XIII
8. Physical Systems
8.1 Cooperative Effects in the Laser: Self-Organization and Phase
Transition . . . . . . . . . . . . . 229
8.2 The Laser Equations in the Mode Picture 230
8.3 The Order Parameter Concept. 231
8.4 The Single-Mode Laser. . . . . . . . 232
8.5 The Multimode Laser . . . . . . . . 235
8.6 Laser with Continuously Many Modes. Analogy with Super-
conductivity . . . . . . . . . . . . . . . . . . . . 237
8.7 First-Order Phase Transitions of the Single-Mode Laser . . 240
8.8 Hierarchy of Laser Instabilities and Ultrashort Laser Pulses 243
8.9 Instabilities in Fluid Dynamics: The Benard and Taylor
Problems . . . . . . . . . . . . 249
8.10 The Basic Equations . . . . . . . . . 250
8.11 The Introduction of New Variables. . . 252
8.12 Damped and Neutral Solutions (R :s;; Rc) 254
8.13 Solution Near R = Rc (Nonlinear Domain). Effective Langevin
Equations . . . . . . . . . . . . . . . . . . . . . . 258
8.14 The Fokker-Planck Equation and Its Stationary Solution . . 262
8.15 A Model for the Statistical Dynamics of the Gunn Instability
Near Threshold . . . . . . . . . . . . . 266
8.16 Elastic Stability: Outline of Some Basic Ideas 270
11
XIV Contents
12. Chaos
12.1 What is Chaos? 333
12.2 The Lorenz Model. Motivation and Realization 334
12.3 How Chaos Occurs . . . . . . . . . . . . . 336
12.4 Chaos and the Failure of the Slaving Principle . 341
12.5 Correlation Function and Frequency Distribution . 343
12.6 Discrete Maps, Period Doubling, Chaos, Intermittency 345
12
1. Goal
Why You Might Read This Book
cold
t
• ••• • • • •
• • • • •
• • • • • •
••• • • •
• Fig. 1.2. Irreversible expansion of a gas
When we have a vessel filled with gas atoms and remove the piston the gas will
fill the whole space (Fig. 1.2). The opposite process does not occur. The gas by itself
does not concentrate again in just half the volume of the vessel. When we put a
drop of ink into water, the ink spreads out more and more until finally a homo-
geneous distribution is reached (Fig. 1.3). The reverse process was never observed.
When an airplane writes words in the sky with smoke, the letters become more and
2 I. Goal
more diffuse and disappear (Fig. 1.4). In all these cases the systems develop to a
unique final state, called a state of thermal equilibrium. The original structures
disappear being replaced by homogeneous systems. When analysing these phe-
nomena on the microscopic level considering the motion of atoms or molecules one
finds that disorder is increased.
Let us conclude these examples with one for the degradation of energy. Consider
a moving car whose engine has stopped. At first the car goes on moving. From a
physicist's point of view it has a single degree of freedom (motion in one direction)
with a certain kinetic energy. This kinetic energy is eaten up by friction, converting
that energy into heat (warming up the wheels etc.). Since heat means thermal
motion of many particles, the energy of a single degree of freedom (motion of the
car) has been distributed over many degrees of freedom . On the other hand, quite
obviously, by merely heating up the wheels we cannot make a vehicle go.
In the realm of thermodynamics, these phenomena have found their proper
description. There exists a quantity called entropy which is a measure for the
degree of disorder. The (phenomenologically derived) laws of thermodynamics
state that in a closed system (i.e., a system with no contacts to the outer world) the
entropy ever increases to its maximal value.
000
000 0 0
00 0
o 0 00
gas molecules droplet crystoll iM lollic.
Fig. 1.5. Water in its different
In 0 box phases
On the other hand when we manipulate a system from the outside we can change
its degree of order. Consider for example water vapor (Fig. 1.5). At elevated tem-
perature its molecules move freely without mutual correlation. When temperature
is lowered, a liquid drop is formed, the molecules now keep a mean distance
between each other. Their motion is thus highly correlated. Finally, at still lower
14
1.1 Order and Disorder: Some Typical Phenomena 3
temperature, at the freezing point, water is transformed into ice crystals. The
molecules are now well arranged in a fixed order. The transitions between the
different aggregate states, also called phases, are quite abrupt. Though the same
kind of molecules are involved all the time, the macroscopic features of the three
phases differ drastically. Quite obviously, their mechanical, optical, electrical,
thermal properties differ wildly.
Another type of ordering occurs in ferromagnets (e.g., the magnetic needle of a
compass). When a ferromagnet is heated, it suddenly loses its magnetization. When
temperature is lowered, the magnet suddenly regains its magnetization (cf. Fig. 1.6).
What happens on a microscopic, atomic level, is this: We may visualize the magnet
as being composed of many, elementary (atomic) magnets (called spins). At elevated
temperature, the elementary magnets point in random directions (Fig. 1.7). Their
1.0 o 0
~ ,
"\ 0
08
\
€O .6
:i
E
\
~·0.4
0.2
Fig. 1.6. Magnetization of ferromagnet as
function of temperature (After C. Kittel:
Introduction to Solid State Physics. Wiley
0.2 0.4 0.6 0.8 1.0 Inc., New York 1956)
rlT.
magnetic moments, when added up, cancel each other and no macroscopic mag-
netization results. Below a critical temperature, T e , the elementary magnets are
lined up, giving rise to a macroscopic magnetization. Thus the order on the micro-
scopic level is a cause of a new feature of the material on the macroscopic level. The
change of one phase to the other one is called phase transition. A similarly dramatic
phase transition is observed in superconductors. In certain metals and alloys the
electrical resistance completely and abruptly vanishes below a certain temperature
(Fig. 1.8). This phenomenon is caused by a certain ordering of the metal electrons.
There are numerous further examples of such phase transitions which often show
striking similarities.
15
4 1. Goal
Rrsislonce
While this is a very interesting field of research, it does not give us a clue for the
explanation of any biological processes. Here order and proper functioning are not
achieved by lowering the temperature, but by maintaining a flux of energy and
matter through the system. What happens, among many other things, is this:
The energy fed into the system in the form of chemical energy, whose processing
involves many microscopic steps, eventually results in ordered phenomena on a
macroscopic scale: formation of macroscopic patterns (morphogenesis), locomo-
tion (i.e., few degrees of freedom 1) etc.
In view of the physical phenomena and thermodynamic laws we have men-
tioned above, the possibility of explaining biological phenomena, especially the
creation -of order on a macroscopic level out of chaos, seems to look rather hope-
less. This has led prominent scientists to believe that such an explanation is impos-
sible. However, let us not be discouraged by the opinion of some authorities. Let us
rather reexamine the problem from a different point of view. The example of the
car teaches us that it is possible to concentrate energy from many degrees offreedom
into a single degree of freedom. Indeed, in the car's engine the chemical energy of
gasoline is first essentially transformed into heat. In the cylinder the piston is then
pushed into a single prescribed direction whereby the transformation of energy of
many degrees of freedom into a single degree of freedom is accomplished. Two facts
are important to recall here:
1) The whole process becomes possible through a man-made machine. In it we
have established well-defined constraints.
2) We start from a situation far from thermal equilibrium. Indeed pushing the
piston corresponds to an approach to thermal equilibrium under the given
constraints.
The immediate objection against this machine as model for biological systems
lies in the fact that biological systems are self-organized, not man-made. This leads
us to the question if we can find systems in nature which operate far from thermal
equilibrium (see 2 above) and which act under natural constraints. Some systems of
this kind have been discovered quite recently, others have been known for some
while. We describe a few typical examples:
A system lying on the border line between a natural system and a man-made
device is the laser. We treat here the laser as a device, but laser action (in the micro-
wave region) has been found to take place in interstellar space. We consider the
16
1.1 Order and Disorder: Some Typical Phenomena 5
pump
I
Fig. 1.10. Photons emitted in axial direction
(a) have a much longer lifetime to in the
Fig. 1.9. Typical setup of a laser "cavity" than all other photons (b)
I1i,Id ",,"g Ih
1\ /\ C>..
\]V
f ield strength
Fig. 1.11.
Wave tracks
emitted (a)
from a lamp,
(b) from a
(b) laser
As we shall see later in this book, this analogy goes far deeper. Obviously the
laser is a system far from thermal equilibrium. As the pump energy is entering the
system, it is converted into laser light with its unique properties. Then this light
leaves the laser. The obvious question is this: What is the demon that tells the sub-
17
6 1. Goal
Diode +51
systems (i.e., the atoms) to behave in such a well organized manner? Or, in more
scientific language, what mechanisms and principles are able to explain the self-
organization of the atoms (or atomic antennas)? When the laser is pumped still
higher, again suddenly a completely new phenomenon occurs. The rod regularly
emits light flashes of extremely short duration, say 10- 12 second. Let us consider
as a second example fluid dynamics, or more specifically, the flow of a fluid round a
cylinder. At low speed the flow portrait is exhibited in Fig. 1.l3(a). At higher speed,
(e)
~"' IOO
18
1.1 Order and Disorder: Some Typical Phenomena 7
suddenly a new, static pattern appears : a pair of vortices (b). With still higher speed,
a dynamic pattern appears, the vortices are now oscillating (c). Finally at still
higher speed, an irregular pattern called turbulent flow arises (e). While we will
not treat this case in our book, the following will be given an investigation.
The convection instability (Benard instability). We consider a fluid layer heated
from below and kept at a fixed temperature from above (Fig. 1.14). At a small
temperature difference (more precisely, gradient) heat is transported by heat con-
duction and the fluid remains quiescent. When the temperature gradient reaches a
critical value, the fluid starts a macroscopic motion. Since heated parts expand,
these parts move up by buoyancy, cool, and fall back again to the bottom. Amaz-
ingly, this motion is well regulated. Either rolls (Fig. 1.15) or hexagons (Fig. 1.16)
are observed. Thus, out of a completely homogeneous state, a dynamic well-
ordered spatial pattern emerges. When the temperature gradient is further in-
creased, new phenomena occur. The rolls start a wavy motion along their axes.
Further patterns are exhibited in Fig. 1.17. Note, that the spokes oscillate tem-
porarily. These phenomena playa fundamental role for example in meteorology,
determining the motion of air and the formation of clouds (see e.g., Fig. 1.18).
19
8 l. Goal
Fig. 1.18. A typical pattern of cloud streets (After R. Scorer : Clouds of the World. Lothian Pub!.
Co., Melbourne 1972)
20
1.1 Order and Disorder: Some Typical Phenomena 9
A closely related phenomenon is the Taylor instability. Here a fluid is put be-
tween two rotating coaxial cylinders. Above a critical rotation speed, Taylor
vortices occur. In further experiments also one of the cylinders is heated. Since in a
number of cases, stars can be described as rotating liquid masses with thermal
gradients, the impact of this and related effects on astrophysics is evident. There are
numerous further examples of such ordering phenomena in physical systems far
from thermal equilibrium. However we shall now move on to chemistry.
In a number of chemical reactions, spatial, temporal or spatio-temporal pat-
terns occur. An example is provided by the Belousov-Zhabotinsky reaction. Here
Ce2(S04h, KBr0 3 , CH 2(COOHh, H 2S04 as well as a few drops of Ferroine
(redox indicator) are mixed and stirred. The resulting homogeneous mixture is then
put into a test tube, where immediately temporal oscillations occur. The solution
changes color periodically from red, indicating an excess of Ce 3 +, to blue, indicat-
ing an excess of Ce 4+ (Fig. 1.19). Since the reaction takes place in a closed system,
the system eventually approaches a homogeneous equilibrium. Further examples
of developing chemical structures are represented in Fig. 1.20. In later chapters of
this book we will treat chemical reactions under steady-state conditions, where,
nevertheless, spatio-temporal oscillations occur. It will turn out that the onset
of the occurrence of such structures is governed by principles analogous to those
governing disorder-order transitions in lasers, hydrodynamics, and other systems.
Our last class of examples is taken from biology. On quite different levels, a
pronounced spontaneous formation of structures is observed. On the global level,
we observe an enormous variety of species. What are the factors determining their
distribution and abundance? To show what kind of correlations one observes,
consider Fig. 1.21 which essentially presents the temporal oscillation in numbers
of snowshoe hares and lynx. What mechanism causes the oscillation? In evolution,
selection plays a fundamental role. We shall find that selection of species obeys the
same laws as for example, laser modes.
Let us turn to our last example. In developmental physiology it has been known
long since that a set of equal (equipotent) cells may self-organize into structures
with well-distinguished regions. An aggregation of cellular slime mold (Dictyoste-
lium disciodeum) may serve as model for cell interactions in embryo genesis.
Dictyostelium forms a multi-cellular organism by aggregation of single cells. Dur-
ing its growth phase the organism exists in the state of single amoeboid cells. Several
hours after the end of growth, these cells aggregate forming a polar body along
21
10 1. Goal
(a)
(b)
Fig. 1.20 a and b. Spirals of chemical activity in a shallow dish. Wherever two waves collide head
on, both vanish. Photographs taken by A. T. Winfree with Polaroid Sx' 70.
22
l.l Order and Disorder: Some Typical Phenomena 11
le0r--------------------------------------------------------,
140 _HARE
__ __ LyNX
~12
Z
~IOO
:>
~ eo
....
z 60
le45 1055
Fig. 1.21. Changes in the abundance of the lynx and the snowshoe hare, as indicated by the
number of pelts received by the Hudson Bay Company (After D. A. McLulich : Fluctuations in
the Numbers of Varying Hare. Univ. of Toronto Press, Toronto 1937)
Fig. 1.22. Wave pattern of <;hemotactic activity in dense cell layers of slime mold (after Gerisch
et al.)
23
12 1. Goal
which they differentiate into other spores or stalk cells the final cell types constitut-
ing the fruiting body. The single cells are capable of spontaneously emitting a
certain kind of molecules called cAMP (cyclic Adenosin 3'5'Monophosphate) in
the form of pulses into their surroundings. Furthermore cells are capable of
amplifying cAMP pulses. Thus they perform spontaneous and stimulated emission
of chemicals (in analogy to spontaneous and stimulated emission of light by laser
atoms). This leads to a collective emission of chemical pulses which migrate in the
form of concentration waves from a center causing a concentration gradient of
cAMP. The single cells can measure the direction of the gradient and migrate
towards the center with help of pseudopods. The macroscopically resulting wave
patterns (spiral or concentric circles) are depicted in Fig. 1.22 and show a striking
resemblance to the chemical concentration waves depicted in Fig. 1.20.
24
1.2 Some Typical Problems and Difficulties 13
elongat ion q
~
1 2 3 4 5 6 ~. Fig. 1.23. Point masses coupled by
springs
selects certain "typical features". Thus we shall discover that there are correlations
between the positions of adjacent atoms (Fig. 1.25). Furthermore when looking
very carefully we shall observe that the motion is periodic in time. However, in this
way we shall never discover that the natural description of the motion of the string
is by means of a spatial sine wave (cf. Fig. 1.26), unless we already know this answer
and feed it as initial condition into the computer. Now, the (spatial) sine wave is
characterized by quantities i.e., the wavelength and the amplitude which are com-
pletely unknown on the microscopic (atomic) level. The essential conclusion from
our example is this: To describe collective behavior we need entirely new con-
cepts compared to the microscopic description. The notion of wavelength and
amplitude is entirely different from that of atomic positions. Of course, when we
know the sine wave we can deduce the positions of the single atoms.
25
14 1. Goal
26
1.3 How We Shall Proceed 15
mous number of microscopic variables may describe their chemical and electrical
activities. The order-parameters are ultimately the thoughts. Both systems neces-
sitate each other. This brings us to a final remark. We have seen above that on a
macroscopic level we need concepts completely different from those on a micro-
scopic level. Consequently it will never suffice to describe only the electro-chemical
processes of the brain to describe its functioning properly. Furthermore, the
ensemble of thoughts forms again a "microscopic" system, the macroscopic order
parameters of which we do not know. To describe them properly we need new
concepts going beyond our thoughts, for us an unsolvable problem. For lack of
space we cannot dwell here on these problems which are closely tied up with deep-
lying problems in logic and which, in somewhat different shape, are well known to
mathematicians, e.g., the Entscheidungsproblem. The systems treated in our book
will be of a simpler nature, however, and no such problems will occur here.
27
16 1. Goal
2. Probability
3. Information
,t
,
I
L ____ _
II. Sociology
Table 1.1
28
2. Probability
What We Can Learn From Gambling
The objects we shall investigate in our book may be quite different. In most cases,
however, we shall treat systems consisting of very many subsystems of the same kind
or of very few kinds. In this chapter we deal with the subsystems and define a few
simple relations. A single subsystem may be among the following:
atoms plants
molecules animals
photons (light quanta) students
cells
°
coin. Denoting its tail by zero and its head by one, the sample set of the coin is
given by Q = {O, I}. Tossing a coin now means to sample or I at random. An-
other example is provided by the possible outcomes when a die is rolled. Denoting
the different faces of a die by the numbers I, 2, ... , 6 the sample set is given by
Q = {l, 2, 3,4, 5, 6}. Though we will not be concerned with games (which are,
nevertheless a very interesting subject), we shall use such simple examples to exhibit
our basic ideas. Indeed, instead of rolling a die we may do certain kinds of experi-
ments or measurements whose outcome is of a probabilistic nature. A sample point
is also called a simple event, because its sampling is the outcome of an "experiment"
(tossing a coin, etc.).
It will be convenient to introduce the following notations about sets. A collec-
tion of w's will be called a subset of Q and denoted by A, B, ... The empty set is
written as 0, the number of points in a set S is denoted by lSI. If all points of the set
A are contained in the set B, we write
A c B or B:::> A. (2.1)
18 2. Probability
A = B. (2.2)
The union
A u B = {w I WE A or WEB} (2.W
is a new set which contains all points contained either in A or B. (cf. Fig. 2.1). The
intersection
A nB= {WIWEA and wEB} (2.4)
is a set which contains those points which are both contained in A and B (cf. Fig.
2.2). The sets A and B are disjoint if (cf. Fig. 2.3).
An B = 0. (2.5)
Fig. 2.1. The union A U B of the sets A Fig. 2.2. The intersection A n B ~ of the
I:2a and B ISSSI comprises all elements of sets A E2<I and B rn
A and B. (To visualize the relation 2.3 we
represent the sets A and B by points in a
plane and not along the real axis)
1 Read the rhs of (2.3) as follows: all co's, for which w is either element of A or of B.
30
2.2 Random Variables 19
All sample points of Q which are not contained in A form a set called the comple-
ment A C of A (cf. Fig. 2.4). While the above-mentioned examples imply a countable
number of sample points, w = 1,2, ... , n (where n may be infinite), there are
other cases in which the subsystems are continuous. Think for example of an area
of a thin shell. Then this area can be further and further subdivided and there are
continuously many possibilities to select an area. If not otherwise noted, however,
we will assume that the sample space Q is discrete.
Exercises on 2.1
Hint: Take an arbitrary element w of the sets defined on the Ihs, show that it is
contained in the set on the rhs of the corresponding equations. Then do the same
with an arbitrary element w' of the rhs.
4) de Morgan's law:
a) (Ai u A2 U ... U AnY = A~ n A~ n ... n A~
b) (Ai n A2 n ... n An)e = A~ u A~ u ... u A~
Hint: Prove by complete induction.
31
20 2. Probability
em X(w) (height)
190
180
170
1===;;;========
t------.~-......----......---_.r_
160
150
This function X is called a random variable, simply because the sample point w is
picked at random: Making a velocity measurement of a single gas molecule or the
rolling of a die. Once the sample point is picked, X(w) is determined. It is, of
course, possible to ascribe several numerically valued functions XI ' X 2 , ••• to the
sample points w, e.g. molecules might be distinguished by their weights, their
velocities, their rotational energies etc. We mention a few simple facts about ran-
*
dom variables. If X and Yare random variables, then also the linear combination
aX + bY, the product XY, and the ratio X/ Y (Y 0), are also random variables.
More generally we may state the following: If cp is a function of two ordinary
variables and X and Yare random variables, then w -+ cp(X(w), Yew»~ is also a
random variable. A case of particular interest is given by the sum of random
variables
Because later on we want to treat the joint action of many subsystems (distinguished
by an index i = 1,2, ... ,n), we have often to deal with such a sum.
Examples are provided by the weight of n persons in a lift, the total firing rate
of n neurons, or the light wave built up from n wavetracks emitted from atoms. We
shall see later that such functions as (2.6) reveal whether the subsystems (persons,
neurons, atoms) act independently of each other, or in a well-organized way.
2.3 Probability
Probability theory has at least to some extent its root in the considerations of the
outcomes of games. Indeed jf one wants to present the basic ideas, it is still ad-
vantageous to resort to these examples. One of the simplest games is tossing a coin
where one can find head or tail. There are two outcomes. However, if one bets on
head there is only one favorable outcome. Intuitively it is rather obvious to define
as probability for the positive outcome the ratio between the number of positive
outcomes 1 divided by the number of all possible outcomes 2, so that we obtain
P = 1/2. When a die is thrown there are six possible outcomes so that the sample
space Q = {I, 2, 3, 4, 5, 6} . The probability to find a particular number k =
1,2,3,4,5,6 is P(k) = 1/6. Another way of reaching this result is as follows . When,
for symmetry reasons, we ascribe the same probability to all six outcomes and de-
mand that the sum of the probabilities of these outcomes must be one, we again
32
2.4 Distribution 21
obtain
Such symmetry arguments also play an important role in other examples, but in
many cases they require far-reaching analysis. First of all, in our example it is as-
sumed that the die is perfect. Each outcome depends, furthermore, on the way the
die is thrown. Thus we must assume that this is again done in a symmetric way
which after some thinking is much less obvious. In the following we shall assume
that the symmetry conditions are approximately realized. These statements can be
reformulated in the spirit of probability theory. The six outcomes are treated as
equally likely and our assumption of this "equal likelihood" is based on symmetry.
In the following we shall call P(k) a probability but when k is considered as
varying, the function P will be called a probability measure. Such probability
measures may not only be defined for single sample points w but also for subsets
A, B, etc. For example in the case of the die, a subset could consist of all even
numbers 2, 4, 6. The probability of finding an even number when throwing a die is
evidently P({2,4,6}) = 3/6 = 1/2 or in short P(A) = 1/2 where A = {2,4,6}.
We now are able to define a probability measure on the sample space Q. It is a
function on Q which satisfies three axioms:
I) For every set A c Q the value of the function is a nonnegative number P(A) ~ o.
2) For any two disjoint sets A and B, the value of the function oftheir union A u B
is equal to the sum of its values for A and B,
P(Q) = 1. (2.9)
2.4 Distribution
We again assume that the total set Q is countable. We have already seen in Section
33
22 2. Probability
em X{w) x , I
ISO
190
170
~==n:========
t-
160
ISO , ,
I , Fig. 2.6. The probability P
~~~ to find a person between
heights h) and hJ+ I (hi =
6 P(hj ~X~hj., ) 150cm, h2 =160cm, etc.)
2.3 that in this case we may assign a probability to every subset of Q. Thus we may
assign to the subset defined by (2.10) a probability which we denote by
To illustrate this definition we consider the example of a die. We ask for the
probability that, if the die is rolled once, the number of spots lies in between 2 and
5. The subset of events which have to be counted according to (2.10) is given by the
numbers of2, 3,4,5. (Note that in our example X(w) = w). Because each number
appears with the probability 1/6, the probability to find the subset 2, 3, 4, 5 is
P = 4/6 = 2/3. This example immediately lends itself to generalization. We could
equally well ask what is the probability of throwing an odd number of spots, so
that X = 1,3, 5. In that case, X does not stem from an interval but from a certain
well-defined subset, A, so that we define now quite generally
where A is a set of real numbers. As a special case of the definition (2.1 I) or (2. 12),
we may consider that the value of the random variable X is given: X = x. In this
case (2.11) reduces to
Now we come to a general rule of how to evaluate P(a ::; X ::; b). This is sug-
gested by the way we have counted the probability of throwing a die with a result-
ing number of spots lying inbetween 2 and 5. We use the fact that X(w) is countable
if Q is countable. We denote the distinct values of X(w) by VI' V2' ••• , Vm ••• and
the set {VI> V2' ••• } by Vx ' We further define that P(X = x) = 0 if x If Vx ' We may
furthermore admit that some of the v;s have zero probabilities. We abbreviate
Pn = P(X = vn) (cf. Fig. 2.7a, b). Using the axioms on page 21 we may deduce
that the probability P(a :::; X:::; b) is given by
(2.14)
34
2.4 Distribution 23
h;; h4 I (a)
hi h3 hS
P(X:: vn ) .:: Pn
P;;
P,
~
v, v;; ~ ~ I \(b)
FX(x)
If the set A consists of all real numbers X of the interval of - 00 until x we define
the so called distribution/unction of X by (cf. Fig. 2.7c)
(2.16)
The Pn's are sometimes called elementary probabilities. They have the properties
and
(2.18)
again in accordance with the axioms. The reader should be warned that "distribu-
tion" is used with two different meanings:
I II
35
24 2. Probability
1/21r 1 - -- -- - - ,
1) V~:f(~) ~ 0, (2.19)2
(2.20)
(The integral is meant as Riemann's integral). We will callfa density function and
assume that it is piecewise continuous so that J~f(e) de exists. We again consider a
random variable X defined on Q with w ---> X(w). We now describe the probability
by (cf. Fig. 2.9)
36
2.5 Random Variables with Densities 25
! IV
Plo~X",b)
lihood !), the probability of finding the needle in a certain direction in the interval
'" 0 and "'0 + d", is (J j2n )d"'. Then the probability of finding '" in the interval '" 1
and "'2
will be given by
Note that Ij(2n) stems from the normalization condition (2.20) and that in our
casef(~) = Ij(2n) is a constant. We may generalize (2.21) if A consists of a union
of intervals. We then define accordingly
(2.23)
Exercise on 2.5
b(x - x o) = 0 for x 9= Xo
and
37
26 2. Probability
{(x)
I
Vr~
em
X (w) (height) x
18()
170
190
160
§~~S~~~~~~~
ISO
Y(w) (weight)
Fig. 2.11. The random variables X (height) and Y (weight). On the rhs the probability measures
for the height (irrespective of weight) and for the weight (irrespective of height) are plotted
38
2.6 Joint Probability 27
P{heighl)
2.15 - - - - - - - -
P{weight)
//6
X
(height)
y
(weight)
Fig. 2.12. The persons are grouped according to their weights and heights
P{height) ({ X, y) f 5)
3/6 3/6
2/6 216
P{weight)
1/6
, 2/6
1/6
Fig.2.13. The joint probability P«X. Y) E S) plotted over X and Y. The subsets S are the single
squares (see for example shaded square)
random variables (X, Y) acquire. For any subset S' E S we then define the joint
probability, generalizing (2.12), as (cf. Fig. 2.13)
Using the axioms of Section 2.3, we may easily show that the probability P(X =
um) i.e. the probability that X(w) = Um irrespective of which values Yacquires, is
given by
(2.27)
39
28 2. Probability
Exercise on 2.6
I) Generalize the definitions and relations of Sections 2.4 and 2.5 to the present
case.
2) Generalize the above definitions to several random variables X,(w), XzCw), ... ,
XN(W).
where X{w) is the random variable, P is the probability of the sample point wand
the sum runs over all points of the sample set Q. If we label the sample points by
integers I, 2, ... , n we may use the definition of Section 2.4 and write:
E{ X) = In PnVn' (2.29)
(2.31)
When we put cp(X) = X' in (2.30), or (2.31), or (2.32) the mathematical expectation
40
2.8 Conditional Probabilities 29
(2.33)
IAI (2.34)
P(A) = IQI
where IAI, IQI denote the number of elements of A or Q. This rule can be generalized
if Q is countable and each point w has the weight P(w). Then
In the cases (2.34) and (2.35) we have admitted the total sample space. As an
example for (2.34) and (2.35) we just have to consider again the die as described at
the beginning of the section.
We now ask for the following probability: We restrict the sample points to a
certain subset S(w) and ask for the proportional weight of the part of A in S
relative to S. Or, in the case of the die, we admit as sample points only those
which are odd. In analogy to formula (2.35) we find for this probability
Extending the denominator and numerator in (2.36) by IIL.,Eu P(w) we may re-
write (2.36) in the form
41
30 2. Probability
other terminologies are also used, such as "knowing S", "given S", "under the
hypothesis of S".
Exercises on 2.8
1) A die is thrown. Determine the probability to obtain 2 (3) spots under the
hypothesis that an odd number is thrown.
Hint: Verify A = {2}, or = {3}, S = {t, 3, 5}. Determine A n S and use (2.37).
2) Given are the stochastic variables X and Y with probability measure
Show that
P(rn, n)
P(rn I n) = Lm P(rn, n)'
P(XI = Xl' ..• , Xn = Xn) = P(XI = XI )P(X2 = X2 ) ••• P(Xn = Xn). (2.38)
In a more general formulation which may be derived from (2.38) we state that the
variables vary independently if and only if for arbitrary countable sets S I, . . . , Sn,
the following holds:
(2.39)
42
2.10 Generating Functions and Characteristic Functions 31
it is desirable to have a measure for the degree of their independence or, in a more
positive statement, about their correlation. Because the expectation value of the
product of independent random variables factorizes (which follows from (2.38»,
a measure for the correlation will be the deviation of E(XY) from E(X)E( Y). To
avoid having large values of the random variables X, Y, with a small correlation
mimic a large correlation, one normalizes the difference
by dividing it by the standard deviations a(X) and a( Y). Thus we define the so-
called correlation
( _ E(XY) - E(X)E(Y)
P X, Y) - O"(X)a(Y) (2.41)
(2.42)
Exercise on 2.9
The random variables X, Y may both assume the values 0 and 1. Check if these
variables are statistically independent for the following joint probabilities:
a) b) c)
P(X = 0, Y = 0) = ! = ! = 1
P(X = 1, Y = 0) = ! = 0 = 0
P(X = 0, Y = I) =! =0 =0
P(X = I, Y = I) =! =! =0
We start with a special case. Consider a random variable X taking only non-
negative integer values. Examples for X are the number of gas molecules in a cell
of a given volume out of a bigger volume, or, the number of viruses in a volume of
blood out of a much larger volume. Let the probability distribution of X be given
by
43
32 2. Probability
(2.44)
We immediately obtain the coefficients aj with aid of the Taylor expansion of the
function g(z) by taking the j's derivative of g(z) and dividing it by j!
1 djg
aj=-=fj
I (2.45)
J.dz z=o
The great advantage of (2.44) and (2.45) rests in the fact that in a number of im-
portant practical cases g(z) is an explicitly defined function, and that by means of
g(z) one may easily calculate expectation values and moments. We leave it to the
reader how one may derive the expression of the first moment of the random
variables X by means of (2.44). While (2.45) allows us to determine the probability
distribution (2.43) by means of (2.44), it is also possible to derive (2.44) in terms of a
function of the random variable X, namely,
(2.46)
This can be shown as follows: For each z, ill -> ZX(ro) is a random variable and thus
according to formulas (2.30), (2.43) we obtain
(2.48)
(2.49)
The definitions (2.48) and (2.49) can be used not only for discrete values of X but
also for continuously distributed values X. We leave to the reader the detailed
formulation as an exercise.
Exercise on 2.10
44
2.11 A Special Probability Distribution: Binomial Distribution 33
distribution in the following manner: There are (Z) different sequences each having
the probability pkqn-k. Because for different subsets of events the probabilities are
45
34 2. Probability
0.2
0,1-
(a)
o 5 10
B11 (np)
,
0.2
0,1
I 1
5 10
I I k
(b)
Fig. 2.14a and b. Binomial distribution Bk(n, p) as function of k for p = 1/4 and n = 15 (a),
n = 30 (b)
(2.50)
(binomial distribution). Examples are given in Fig. 2.14. The mean value of the
binomial distribution is
the variance
46
2.11 A Special Probability Distribution: Binomial Distribution 35
We leave the proof as an exercise to the reader. For large n it is difficult to evaluate
B. On the other hand in practical cases the number of trials n may be very large.
Then (2.50) reduces to certain limiting cases which we shall discuss in subsequent
sections.
The great virtue of the concept of probability consists in its flexibility which
allows for application to completely different disciplines. As an example for the
binomial distribution we consider the distribution of viruses in a blood cell
under a microscope (cf. Fig. 2.15). We put a square grid over the cell and
• •• •
•
••
• • ••
•
• •• •
• •• Fig. 2.15. Distribution of particles under a grid
ask for the probability distribution to find a given number of particles in a square
(box). If we assume N squares and assume n = J.lN particles present, then J.l is the
average number of particles found in a box. The problem may be mapped on the
preceding one on coin tossing in the following manner: Consider a specific box and
first consider only one virus. Then we have a positive event (success) if the virus
is in the box and a negative event (failure) when it is outside the box. The prob-
ability for the positive event is apparently p = liN. We now have to distribute
J.lN = n particles into two boxes, namely into the small one under consideration
and into the rest volume. Thus we make n trials. If in a "Gedankenexperiment",
we put one virus after the other into the total volume, this corresponds exactly to
tossing a coin. The probability for each special sequence with k successes, i.e., k
viruses in the box under consideration, is given as before by pkqn-k. Since there are
again (~) different sequences, the total probability is again given by (2.50) or
using the specific form n = J.lN by
(2.53)
In practical applications one may consider N and thus also n as very large numbers,
while Jl is fixed and p tends to zero. Under the condition n -+ 00, J.l fixed, p -+ 0
one may replace (2.50) by the so-called Poisson distribution.
47
36 2. Probability
B(np)=
k ,
n(n - 1)(n - 2) ... (n - k
k!
+ 1) (fl)k( 1--fl)n(
. -n n
fl)-k
1--
n
~~ '~--------------~-----------
2 3
fl,
For k fixed but n -> 00 we find the following expressions for the factors 1, 2, 3
in (2.54). The first factor remains unchanged while the second factor yields the
exponential function
(2.55)
(2.56)
= flk! e -Il .
k
Examples are given in Fig. 2.16. Using the notation of Section 2.10 we may also
write
(2.58)
(2.59)
48
2.13 The Normal Distribution (Gaussian Distribution) 37
0,/
0,3-
O,}
0,1-
}O 40
(a) (b)
Fig. 2.16a and b. Poisson distribution 1C' •• as function of k for J1 = 2 (a) and J1 = 20 (b)
Exercise on 2.12
The normal distribution can be obtained as a limiting case of the binomial dis-
tribution, again for n -. 00 but for p and q not small, e.g., p = q = 1/2. Because
the presentation of the limiting procedure requires some more advanced mathe-
matics which takes too much space here, we do not give the details but simply
indicate the spirit. See also Section 4.1. We first introduce a new variable, u, by the
requirement that the mean value of k = np corresponds to u = 0; we further
introduce a new scale with k -. k /a where a is the variance. Because according to
formula (2.52) the variance tends to infinity with n -. 00, this scaling means that we
are eventually replacing the discrete variables k by a continuous variable. Thus
it is not surprising that we obtain instead of the original distribution function B, a
density function , q> where we use simultaneously the transformation q> = aB. Thus
more precisely we put
(2.62)
49
38 2. Probability
.
1Im._ ( ) ( ) 1 -.1..2
oo CP. u = cP u = .J2n e 2 (2.63)
cp(u) is a density function (cf. Fig. 2.17). Its integral is the normal (or Gaussian)
distribution
, Iu}
F(x) 1 JX
= -- e- t • 2 duo (2.64)
Fn -00
When several random variables are present in practical applications often the
following joint Gaussian probability density appears
(2.65)
50
2.15 Central Limit Theorem 39
Exercises on 2.13
I 1
12(n + t) < w(n) < 12n· (2.68)
Since in many practical cases, n » I, the factor exp w(n) may be safely dropped.
(2.70)
51
40 2. Probability
Subtracting n·m from Sn and dividing the difference by u.jn we obtain a new
random variable
S - n·m
y _ ---,,-n_;=_-
n - u.jn (2.71)
which has zero mean and unit variance. The central limit theorem then makes a
statement about the probability distribution of Yn in the limit n -> 00. It states
(2.72)
(2.74)
where
(2.75)
The central limit theorem plays an enormous role in the realm of synergetics in
two ways: it applies to all cases where one has no correlation between different
events but in which the outcome of the different events piles up via a sum. On the
other hand, the breakdown of the central limit theorem in the form (2.72) will
indicate that the random variables Xj are no more independent but correlated.
Later on, we will encounter numerous examples of such a cooperative behavior.
52
3. Information
How to Be Unbiased
(3.1)
we require
(3.2)
1= KlnRo (3.3)
42 3. Infonnation
where K is a constant. It can even be shown that (3.3) is the only solution to (3.2).
The constant K is still arbitrary and can be fixed by some definition. Usually the
following definition is used. We consider a so-called "binary" system which has
only two symbols (or letters). These may be the head and the tail of a coin, or
answers yes and no, or numbers 0 and 1 in a binomial system. When we form all
possible "words" (or sequences) of length n, we find R = 2n realizations. We now
want to identify I with n in such a binary system. We therefore require
which is fulfilled by
1= log2 R. (3.4a)
which reduces to the earlier definition (3.3), if R J I. A simple example for this is
given by a die. Let us define a game in which the even numbers mean gain and the
odd numbers mean loss. Then Ro = 6 and R\ = 3. In this case the information
content is the same as that of a coin with originally just two possibilities. We now
derive a more convenient expression for the information. To this end we first
consider the following example of a simplified 1 Morse alphabet with dash and dot.
We consider a word of length G which contains Nl dashes and N2 dots, with
(3.7)
We ask for the information which is obtained by the receipt of such a word. In the
spirit of information theory we must calculate the total number of words which
can be constructed out of these two symbols for fixed N l , N 2 • The consideration is
quite similar to that of Section 2. I 1 page 33. According to the ways we can dis-
tribute the dashes and dots over the N positions, there are
(3.8)
54
3.1 Some Basic Ideas 43
possibilities. Or, in other words, R is the number of messages which can be trans-
mitted by Nt dashes and N2 dots. We now want to derive the information per
symbol, i.e., i = liN. Inserting (3.8) into (3.3) we obtain
In Q! ~ Q(ln Q - 1) (3.10)
(3.11)
(3.12)
Pj = N'
N·
j = 1, 2. (3.13)
I
i = N= -K(pl lnpl + P2 ln P2)' (3.14)
This expression can be easily generalized if we have not simply two symbols but
several, such as letters in the alphabet. Then we obtain in quite similar manner as
just an expression for the information per symbol which is given by
(3.15)
Pj is the relative frequency of the occurrence of the symbols. From this interpreta-
tion it is evident that i may be used in the context of transmission of information,
etc.
Before continuing we should say a word about information used in the sense
here. It should be noted that "useful" or "useless" or "meaningful" or "meaning-
less" are not contained in the theory; e.g., in the Morse alphabet defined above
quite a number of words might be meaningless. Information in the sense used here
rather refers to scarcity of an event. Though this seems to restrict the theory con-
siderably, this theory will tum out to be extremely useful.
55
44 3. Information
The expression for the information can be viewed in two completely different
ways. On the one hand we may assume that the p;'s are given by their numerical
values, and then we may write down a number for [ by use of formula (3.3). Of
still greater importance, however, is a second interpretation; namely, to consider [
as a/unction of the p;'s; that means if we change the values of the p,'s, the value of
[changes correspondingly. To make this interpretation clear we anticipate an ap-
plication which we will treat later on in much greater detail. Consider a gas of atoms
moving freely in a box. It is then of interest to know about the spatial distribution
of the gas atoms. Note that this problem is actually identical to that of Section 2.11
but we treat it here under a new viewpoint. We again divide the container into M
cells of equal size and denote the number of particles in cell k by N k • The total
number of particles be N. The relative frequency of a particle to be found in cell k
is then given by
Nk
N = Pk' k = 1, 2, ... , M. (3.16)
Pk = 11M. (3.17)
We now want to derive this result (3.17) from the properties of information. Indeed
the information may be as follows: Before we make a measurement or obtain a
message, there are R possibilities or, in other words, Kin R is a measure of our
ignorance. Another way of looking at this is the following: R gives us the number
of realizations which are possible in principle.
Now let us look at an ensemble of M containers, each with N gas atoms. We
assume that in each container the particles are distributed according to different
distribution functions Pk' i.e.,
56
3.1 Some Basic Ideas 45
is an extremum under the constraint that the total sum of the probabilities Pi
equals unity
(3.19)
This principle will turn out to be fundamental for applications to realistic systems
in physics, chemistry, and biology and we shall come back to it later.
The problem (3.18) with (3.19) can be solved using the so-called Lagrangian
mUltipliers. This method consists in multiplying (3.19) by a still unknown para-
meter A. and adding it to the Ihs of (3.18) now requiring that the total expression
becomes an extremum. Here we are now allowed to vary the p;'s independently
of each other, not taking into account the constraint (3.19). Varying the left-hand
side of
(3.20)
-In Pi - 1 + A. = o. (3.21)
Pi = exp(A. - I) (3.22)
which is independent of the index i, i.e., the p;'s are constant. Inserting them into
(3.19) we may readily determine A. so that
M exp(A. - 1) = 1, (3.23)
1
Pi= M (3.24)
Exercises on 3.1
1) Show by means of(3.8) and the binomial theorem that R = 2N if N" N z may be
choosen arbitrarily (within (3.7)).
2) Given a container with five balls in it. These balls may be excited to equidistant
energy levels. The difference in energy between two levels is denoted by A.
Now put the energy E = SA into the system. Which is the state with the largest
number of realizations, if a state is characterized by the number of "excited"
balls? What is the total number of realizations? Compare the probabilities for
the single state and the corresponding number of realizations. Generalize the
formula for the total number of states to 6 balls which get an energy E = NA.
57
46 3. Information
N!
Zt = n k N k',. (3.25)
Taking the logarithm of Zt to the basis of two, we obtain the corresponding in-
formation. Now consider the same situation but with balls having two different
colors, black and white (cf. Fig. 3.1). Now these black balls form a subset of all
balls. The number of black balls be N', their number in box k be N~ . We ~ow
calculate the number of configurations which we find if we distinguish between
black and white balls. In each box we must now subdivide Nk into N~ and Nk - N~.
Thus we find
2 3 5
N, =5 N] =3 N3 = 3 Ns = 5 Fig. 3.1. Distribution of white
Nj =1 z
N =1 Ni =2 N5 =2 and black balls oyer boxes
N'! (N - N')!
Z2 = n N{!' nCNk _ ND! (where Il = Ilk) (3.26)
(3.27)
58
3.2 Information Gain: An Illustrative Derivation 47
N! nN~!n(Nk - ND!
Z = n Nk! . ~ (N - N')! .
(3.28)
N! N'
(N - N')! ~N (3.29)
and
(3.30)
and thus
(3.31)
(3.32)
and using
(3.33)
we obtain
In Z = "
L..k Nk N~ - "AT'
, In Nk L..k lVk In N'
N (3.34)
N' N'/N'
In Z = N' Lk N~ln ;k/N ' (3.35)
(3.36)
and
If we still divide both sides by N' (and then multiply by the constant K (3.5», we
59
48 3. Information
obtain our final formula for the information gain in the form
(3.39)
and
(3.40)
The information gain K(p',p) has the following important property, which we
will use in later chapters:
K(p',p) ~ o. (3.41)
(3.42)
60
3.3 Information Entropy and Constraints 49
knowledge available about the system. Consider an ideal gas in one dimension.
What we could measure, for instance, is the center of gravity. In this case we
would have as constraint an expression of the form
(3.43)
where q i measures the position of the cell i. M is a fixed quantity equal Q/ N, where
Q is the coordinate of the center of gravity, and N the particle number. There are,
of course, very many sets of p;'s which fulfill the relation (3.43). Thus we could
choose a set {p;} rather arbitrarily, i.e., we would favor one set against another one.
Similar to common life, this is a biased action. How may it be unbiased? When we
look again at the example of the gas atoms, then we can invoke the principle stated
in Section 3.1. With an overwhelming probability we will find those distributions
realized for which (3.42) is a maximum. However, due to (3.43) not all distributions
can be taken into account. Instead we have to seek the maximum of (3.42) under
the constraint (3.43). This principle can be generalized if we have a set of con-
straints. Let for example the variable i distinguish between different velocities. Then
we may have the constraint that the total kinetic energy mr! of the particles is fixed.
Denoting the kinetic energy of a particle with mass m and velocity Vi by li[fi =
(m/2)v71 the mean kinetic energy per particle is given by
(3.43a)
(3.44)
We further add as usual the constraint that the probability distribution is nor-
malized
(3.45)
The problem of finding the extremum of (3.42) under the constraints (3.44) and
(3.45) can be solved by using the method of Lagrange parameters Ak' k = 1,2, ... , M
(cf. 3.1). We multiply the Ihs of (3.44) by Ak and the lhs of (3.45) by (A - 1) and
take the sum of the resulting expressions. We then subtract this sum from(l/kB)S.
The factor l/kB amounts to a certain normalization of A, Ak. We then have to vary
the total sum with respect to the p;'s
(3.46)
61
50 3. Information
zero, we obtain
(3.47)
(3.48)
(3.49)
which we shall call the partition function. Inserting (3.50) into (3.49) yields
(3.51)
or
A = In Z, (3.52)
which allows us to determine A once the Ak'S are determined. To find equations for
the Ak'S, we insert (3.48) into the equations of the constraints (3.44) which lead
immediately to
(3.53)
Eq. (3.53) has a rather similar structure as (3.50). The difference between these
two expressions arises because in (3.53) each exponential function is still multiplied
by I?). However, we may easily derive the sum occurring in (3.53) from (3.50) by
differentiating (3.50) with respect to Ak • Expressing the first factor in (3.53) by Z
according to (3.51) we thus obtain
Because the lhs are prescribed (compare (3.44)) and Z is given by (3.50) which is a
62
3.3 Information Entropy and Constraints 51
function of the Ak's in a special form, (3.55) is a concise form for a set of equations
for the Ak's.
We further quote a formula which will become useful later on. Inserting (3.48)
into (3.42) yields
(3.56)
(3.57)
The maximum of the information entropy may thus be represented by the mean
values Ik and the Lagrange parameters Ak' Those readers who are acquainted with
the Lagrange equations of the first kind in mechanics will remember that the
Lagrange parameters have a physical meaning, in that case, of forces. In a similar
way we shall see later on tht the Lagrange parameters Ak have phjsical (or chemical
or biological, etc.) interpretations. By the derivation of the above formulas (i.e.,
(3.48), (3.52) with (3.42), (3.55) and (3.57» our original task to find the p's and
Smax is completed.
We now derive some further useful relations. We first investigate how the
information Smax is changed if we change the functions Ilk) and!" in (3.44). Because
S depends, according to (3.57), not only on thef's but also on Aand Ak's which are
functions of the j's, we must exercise some care in taking the derivatives with
respect to the f's. We therefore first calculate the change of A (3.52)
1
c5A == c5ln Z = Zc5Z.
Eq. (3.53) and an analogous definition of <c5/lk » allow us to write the last line as
(3.58)
Inserting this into oSmax (of (3.57» we find that the variation of the Ak's drops out
and we are left with
(3.59)
63
52 3. Information
(3.60)
(3.61)
The notation "generalized heat" will become clearer below when contact with
thermodynamics is made. In analogy to (3.55) a simple expression for the variance
of f\k) (cf. (2.33)) may be derived:
(3.62)
8,\k») 8,\k)
( ~ -~ ~ (3.63)
8a - L,i Pi 8a .
Using the p/s in the form (3.48) and using (3.51), (3.63) may be written in the form
(3.64)
(3.64) = - z1 I,;1 8Z
8a· (3.65)
1 8 In Z _ (8f~~~)
-I,;--aa - a;- . (3.66)
If there are several parameters a l present, this formula can be readily generalized
by providing on the left and right hand side the a with respect to which we differ-
entiated by this index I.
As we have seen several times, the quantity Z (3.50), or its logarithm, is very
useful (see e.g., (3.55), (3.62), (3.66)). We want to convince ourselves that In Z == A-
(cf. (3.52)) may be directly determined by a variational principle. A glance at (3.46)
reveals that (3.46) can also be interpreted in the following way: Seek the extremum
64
3.4 An Example from Physics: Thermodynamics 53
of
(3.67)
(3.68)
Now, by virtue of (3.44), (3.57) and (3.52) the extremum of (3.67) is indeed identical
with In Z. Note that the spirit of the variational principle for In Z is different from
that for S. In the former case, we had to seek the maximum of S under the con-
straints (3.44), (3.45) with fk fixed and A.k unknown. Here now, only one con-
straint, (3.68), applies and the A.k'S are assumed as given quantities. How such a
switching from one set of fixed quantities to another one can be done will become
more evident by the following example from physics, which will elucidate also many
further aspects of the foregoing.
Exercise on 3.3
To get a feeling for how much is achieved by the extremal principle, answer the
following questions:
To visualize the meaning of the index i, let us identify it with the velocity of a
particle. In a more advanced theory Pi is the occupation probability of a quantum
state i of a many-particle system. Further, we identify f~~~ with energy E and the
parameter a with the volume. Thus we put
65
54 3. Information
(3.70)
We have, in particular, called Ai = /3. With this, we may write a number of the
previous formulas in a way which can be immediately identified with relations
well known in thermodynamics and statistical mechanics. Instead of (3.48) we find
(3.71)
I
kB Smax = lnZ + /3U (3.72)
(3.73)
This equation is well known in thermodynamics and statistical physics. The first
term may be interpreted as the internal energy U, 1//3 as the absolute temperature T
multiplied by Boltzmann's constant k B • Smax is the entropy. The rhs represents the
free energy, :IF, so that (3.73) reads in thermodynamic notation
U - TS = :IF. (3.74)
By comparison we find
(3.75)
and S = Smax. Therefore we will henceforth drop the suffix "max". (3.50) reads
now
(3.76)
and is nothing but the usual partition function. A number of further identities of
thermodynamics can easily be checked by applying our above formulas.
The only problem requiring some thought is to identify independent and de-
pendent variables. Let us begin with the information entropy, Smax. In (3.57) it
appears as a function of A, Ail'S and Ik'S. However, A and A/S are themselves
determined by equations which contain the Ik'S and l\k),S as given quantities (cf.
(3.50), (3.52), (3.53)). Therefore, the independent variables are h, Ilk>, and the
dependent variables are A, Ak , and thus, by virtue of (3.57), Smax. In practice the
l\k)'S are fixed functions of i (e.g., the energy of state "i"), but still depending on
parameters ex (e.g., the volume, cf. (3.69)). Thus the truly independent variables
66
3.4 An Example from Physics: Thermodynamics 55
in our approach are !k'S (as above) and the IX'S. In conclusion we thus find S =
SU;"IX). In our example'!l = E == U, IX = V, and therefore
S = S(U,V). (3.77)
Now let us apply the general relation (3.59) to our specific model. If we vary only
the internal energy, U, but leave V unchanged, then
(3.78)
and
and therefore
or
(3.80)
According to thermodynamics, the lhs of (3.80) defines the inverse of the absolute
temperature
This yields p = Ij(kBT) as anticipated above. On the other hand, varying V but
leaving U fixed, i.e.,
(3.82)
but
(3.83)
yields in (3.59)
oS = k B ( -A1)(bEi(V)/bV)bV
or
(3.84)
bS P
bV=T (3.85)
67
56 3. Information
1 1
bS = T OU + TPOV. (3.87)
In thermodynamics the rhs is equal to dQjT where dQ is heat. This explains the
notation "generalized heat" used above after (3.61). These considerations may be
generalized to different kinds of particles whose average numbers N k , k = 1, ... m
are prescribed quantities. We therefore identify II with E, butIk'+ 1 with N k " k' =
1, ... , m (Note the shift of index !). Since each kind of particle, I, may be present
with different numbers Nt we generalize the index ito i, N I , . . . , N m and put
(3.88)
(3.90)
While the above considerations are most useful for irreversible thermodynamics
(see the following Section 3.5), in thermodynamics the role played by independent
and dependent variables is, to some extent, exchanged. It is not our task to treat
these transformations which give rise to the different thermodynamic potentials
(Gibbs, Helmholtz, etc). We just mention one important case: Instead of U, V
(and N 1 , ••• , N m ) as independent variables, one may introduce V and T =
(8Sj8U)-1 (and N 1 , ••• , N n ) as new independent variables. As an example we
68
3.5 An Approach to Irreversible Thermodynamics 57
treat the U-V case (putting formally /11' /12' ... = 0). According to (3.75) the free
energy, IF, is there directly given as function of T. The differentiation aIFlaTyields
aIF 1 1
- aT = kB In Z + TZLi Ei exp( -{JEJ.
(3.92)
Comparing this relation with (3.73), where 11{J = kBT, yields the important
relation
aIF
--=s
aT
(3.93)
S = -kBLiPilnPi (3.94a)
and
(3.94b)
(3.95a)
and
(3.95b)
We require that the sums of the f's in both subsystems are prescribed constants
(3.96)
69
58 3. Information
(e.g., the total number of particles, the total energy, etc.). According to (3.57) the
entropies may be represented as
(3.97a)
and
(3.97b)
(Here and in the following the suffix "max" is dropped). We further introduce the
sum of the entropies in the total system
(3.98)
Example: Consider the particle number and the energy as extensive variables.
Then we choose k = I to refer to the number of particles, and k = 2 to refer to the
energy. Thus
and
XI2) = E i•
Evidently, we have to identify xl k ) withflk),
(3.99)
For the rhs of (3.95) we introduce correspondingly the notation X k • In the first
part of this chapter we confine our attention to the case where the extensive variables
X k may vary, but the xl k ) are fixed. We differentiate (3.98) with respect to Xk • Using
(3.97) and
(3.100)
70
3.5 An Approach to Irreversible Thermodynamics 59
(3.102)
(3.103)
In (3.103) we have admitted that the h's are functions of time, e.g., the energy U
is a function of time. Taking now the derivative of (3.102) with respect to time and
having in mind that S depends via the constraints (3.103) onh == Xk we obtain
dS _ ~ as dXI;
dt - L..l;axl; dt . (3.104)
(3.105)
introduces the Lagrange multipliers which define the so-called intensive variables.
The second factor in (3.104) gives us the temporal change of the extensive quan-
tities XI;, e.g., the temporal change of the internal energy. Since there exist conser-
vation laws, e.g., for the energy, the decrease or increase of energy in one system
must be caused by an energy flux between different systems. This leads us to define
71
60 3. Information
(3.106)
(3.l07a)
The difference (A.k - ADkB is called a generalized thermodynamic force (or affinity)
and we thus put
(3.107b)
The motivation for this nomenclature will become evident by the examples treated
below. With (3.107b) we may cast (3.107a) into the form
(3.108)
which expresses the temporal change of the entropy SO of the total system (1 + 2)
in terms of forces and fluxes. If Fk = 0, the system is in equilibrium; if Fk # 0, an
irreversible process occurs.
Examples; Let us consider two systems separated by a diathermal wall and take
X k = U (U; internal energy). Because according to thermodynamics
(3.109)
we find
(3.l1O)
We know that a difference in temperature causes a flux of heat. Thus the generalized
forces occur as the cause of fluxes. For a second example consider X k as the mole
number. Then one derives
72
3.5 An Approach to Irreversible Thermodynamics 61
Xk = IXk' (3.112)
f~k) = f~~, •.... aJ"'" (3.113)
(Explicit examples from physics will be given below). Then according to (3.59)
we have quite generally
oS
OIX.
= '"
L...k
k A {Oil _ (Of!.~,,,,)},
B k OIX. OIX.
(3.114)
1 J J
We confine our subsequent considerations to the case whereil does not depend on
i.e., we treat il and IX as independent variables. (Example: In
IX,
(3.115)
~ oS _ '"
kB OlXj - L...k Ak{
_ «k) )
Ofi.a, . . ./0IX j}. (3.116)
For two systems, characterized by (3.98) and (3.100), we have now the additional
conditions
(3.117)
(Example: IX = volume)
The change of the entropy S(O) of the total system with respect to time, dS(O)/dt
may be found with help of (3.l07a), (3.116) and (3.117),
1 dS(O)
kB dt = Lk (Ak - AI.) d/ - Lj.k Ak
dX [(OX{k»)
O;j
(or(k»)] . dlXd:'
- AI. o~j (3.118)
dlX j _ J (3.119)
dt - i'
and
(3.120)
73
62 3. Information
Putting
PI = JI = J I
F 1; if I < I < n }
(3.122)
FI = FI - n ; J1 = i l - n if n + 1 :s; l:s; n +m
we arrive at an equation of the same structure as (3.107) but now taking into
account the parameters IX,
(3.123)
The importance of this generalization will become clear from the following example
which we already encountered in the foregoing chapter (IX = V), (i ~ i, N)
(3.124)
f (2)
~Ia
= N,
The interpretation of (3.125) is obvious from the example given in (3.110, III).
The first two terms are included in (3.107), the other term is a consequence of the
generalized (3.123).
The relation (3.125) is a well-known result in irreversible thermodynamics.
It is, however, evident that the whole formalism is applicable to quite different
disciplines, provided one may invoke there an analogue of local equilibrium.
Since irreversible thermodynamics mainly treats continuous systems we now
demonstrate how the above considerations can be generalized to such systems.
We again start from the expression of the entropy
(3.126)
We assume that S and the probability distribution Pi depend on space and time.
The basic idea is this. We divide the total continuous system into subsystems
(or the total volume in subvolumes) and assume that each subvolume is so large
that we can still apply the methods of thermodynamics but so small that the
probability distribution, or, the value of the extensive variablefmay be considered
as constant. We again assume that the p/s are determined instantaneously and
locally by the constraints (3.103) whereh(t) has now to be replaced by
h(X,f). (3.127)
74
3.5 An Approach to Irreversible Thermodynamics 63
(k)
/ i,ot:{x,t)- (3.128)
(3.129)
as determining the p;'s. We divide the entropy S by a unit volume Vo and consider
in the following the change of this entropy density
S
s = Vo' (3.130)
Xk =h(X,t), (3.131)
(3.132)
It is a simple matter to repeat all the considerations of Section 3.3 in the case in
which the f's are space- and time-dependent. One then shows immediately that one
may write the entropy density again in the form
(3.133)
(3.134)
(3.135)
75
64 3. Information
(3.134) reads
(3.136)
Using the identity AkVJk = V(AkJk) - JkVAk and rearranging terms in (3.136), we
arrive at our final relation
(3.137)
whose lhs has the typical form of a continuity equation. This leads us to the idea
that we may consider the quantity
as an adequate expression for the entropy flux. On the other hand the rhs may be
interpreted as the local production rate of entropy so that we write
(3.138)
Thus (3.137) may be interpreted as follows: The local production of entropy leads
to a temporal entropy change (first term on the lhs) and an entropy flux (second
term). The importance of this equations rests on the following: The quantities J k
and VAk can be again identified with macroscopically measurable quantities. Thus
for example if J k is the energy flux, kB VAk = VCI/T) is the thermodynamic force.
An important comment should be added. In the realm of irreversible thermo-
dynamics several additional hypotheses are made to find equations of motion for
the fluxes. The usual assumptions are I) the system is Markovian, i.e., the fluxes
at a given instant depend only on the affinities at that instant, and 2) one considers
only linear processes, in which the fluxes are proportional to the forces. That means
one assumes relations of the form
(3.139)
The coefficients Ljk are called Onsager coefficients. Consider as an example the heat
flow J. Phenomenologically it is related to temperature, T, by a gradient
J = -1<-VT, (3.140)
(3.141)
By comparison with (3.139) and (3.138) we may identify VOlT) with the affinity
76
3.5 An Approach to Irreversible Thermodynamics 65
and KT2 with the kinetic coefficient L\\. Further examples are provided by Ohm's
law in the case of electric conduction or by Fick's law in the case of diffusion.
In the foregoing we have given a rough sketch of some basic relations. There are
some important extensions. While we assumed an instantaneous relation between
p;(t) andh(t) (cf. (3.129», more recently a retarded relation is assumed by Zubarev
and others. These approaches still are subject to the limitations we shall discuss in
Section 4.11.
Exercise on 3.5
1) Since S(O), Sand S' in (3.98) can be expressed by probabilities Pi' ... according
to (3.42) it is interesting to investigate the implications of the additivity relation
(3.98) with respect to the p's. To this end, prove the theorem: Given are the en-
tropies (we omit the common factor k B )
where
Pi Ii p(i,j)
= (E.4)
pj = Ii p(i, j) (E.5)
(i.e., the joint probability p(i,j) is a product or, in other words, the events
X = i, Y = j are statistically independent).
Hint: I) to prove "if", insert (E.7) into (E.I), and use IiPi = I, IiP} = 1,
2) to prove "only if", start from (E.6), insert on the Ihs (E.l), and on the
rhs (E.2) and (E.3), which eventually leads to
Now use the property (3.41) of K(p,p'), replacing in (3.38) the index k
by the double index i, j.
2) Generalize (3.125) to several numbers, Nk> of particles.
77
66 3. Information
78
3.6 Entropy-Curse of Statistical Mechanics? 67
pletely determined. However, if we average over volumes, then the whole state gets
more and more diffuse. This approach has certain drawbacks because it appears
that the increase of entropy depends on the averaging procedure.
2) In the second approach we assume that it is impossible to determine the
time-dependence of Pi by the completely deterministic equations of mechanics.
Indeed it is impossible to neglect the interaction of a system with its surroundings.
The system is continuously subjected to fluctuations from the outer world. These can
be, if no other contacts are possible, vacuum fluctuations of the electromagnetic
field, fluctuations of the gravitational field, etc. In practice, however, such interac-
tions are much more pronounced and can be taken into account by "heatbaths",
or "reservoirs" (cf. Chap. 6). This is especially so if we treat open systems where
we often need decay constants and transition rates explicitly. Thus we shall adopt
in our book the attitude that the p/s are always, at least to some extent, spreading
which automatically leads to an increase of entropy (in a closed system), or more
generally, to a decrease of information gain (cf. Section 3.2) in an open (or closed)
system (for details cf. Exercises (1) and (2) of Section 5.3).
Lack of space does not allow us to discuss this problem in more detail but we
think the reader will learn still more by looking at explicit examples. Those will be
exhibited later in Chapter 4 in context with the master equation, in Chapter 6 in
context with the Fokker-Planck equation and Langevin equation, and in Chapters
8 to 10 by explicit examples. Furthermore the impact of fluctuations on the entropy
problem will be discussed in Section 4.10.
79
4. Chance
How Far a Drunken Man Can Walk
This chapter will serve three purposes. 1) We give a most simple example of a
stochastic process; 2) we show how a probability can be derived explicitly in this
case; 3) we show what happens if the effects of many independent events pile up.
The binomial distribution which we have derived in Section 2.11 allows us to treat
Brownian movement in a nice fashion. We first explain what Brownian movement
means: When a small particle is immersed in a liquid, one observes under a micro-
scope a zig-zag motion of this particle. In our model we assume that a particle
moves along a one-dimensional chain where it may hop from one point to one of
its two neighboring points with equal probability. We then ask what is the prob-
ability that after n steps the particle reaches a certain distance, x, from the point
where it had started. By calling a move to the right, success, and a move to the
left, failure, the final position, x, of the particle is reached after, say, s successes and
(n - s) failures. Ifthe distance for an elementary step is a, the distance, x, reached
reads
n x
s = 2 + 20· (4.2)
70 4. Chance
Denoting the transition time per elementary step by r, we have for the time t after
n steps
t = nr. (4.3)
The probability of finding the configuration with n trials out of which s are success-
ful has been derived in Section 2.11. It reads
p(x, t) == pes, n) =
tlr
( tl(2r) + xl(2a)
)(l)tl<
"2 (4.4a)
Apparently, this formula is not very handy and, indeed, difficult to evaluate exactly.
However, in many cases of practical interest we make our measurements in times t
which contain many elementary steps so that n may be considered as a large number.
Similarly we also assume that the final position x requires many steps s so that also
s may be considered as large. How to simplify (4.4a) in this limiting case requires
a few elementary thoughts in elementary calculus. (Impatient readers may quickly
pass over to the final result (4.29) which we need in later chapters). To evaluate
(4.4) for large sand n we write (4.4) in the form
n!
s!(n - s)! 2
(l)n (4.5)
and apply Stirling's formula (compare Section 2.14) to the factorials. We use
Stirling's formula in the form
(4.6)
and correspondingly for s! and (n - s)! Inserting (4.6) and the corresponding
relations into (4.4), we obtain
where
e- n
A = e se (n s) = I, (4.8)
82
4.1 A Model of Brownian Movement 71
A stems from the powers of e in (4.6) which in (4.8) cancel out. B stems from the
second factor in (4.6). The factor C stems from the last factor in (4.6), while F
stems from the last factor in (4.5). We first show how to simplify (4.9). To this end
we replace s by means of (4.2), put
x
-=~ (4.11)
2a '
and
a2
In (1 + a) = a - "2 + (4.13)
Inserting (4.12) with (4.13) into (4.9) and keeping only terms up to second order in
~, we obtain for F·B
{- ~2't"} (4.14)
B·F = exp{ -2(j2In} = exp 2a 2t .
(4.15)
where D is called the diffusion constant. In the following we require that when we
let a and 't" go to zero the diffusion constant D remains a finite quantity. Thus (4.14)
takes the form
We now discuss the factor C in (4.7). We shall see later that this factor is closely
related to the normalization constant. Inserting (4.3) and (4.2) into C, we find
(4.17)
(4.18)
83
72 4. Chance
Because we have in mind to let -r and a go to zero but letting D fixed we find that
-rD -+ 0
so that the term containing x 2 in (4.18) can be dropped and we are left with
(4.19)
Inserting the intermediate results A (4.8), B·F (4.16) and C (4.19) into (4.7), we
obtain
(4.20)
The occurrence of the factor -rl/2 is, of course, rather disturbing because we want
to let -r -+ O. This problem can be resolved, however, when we bear in mind that it
does not make sense to introduce a continuous variable Xl and seek a probability
to find a specific value x. This probability must necessarily tend to zero. We there-
fore have to ask what is the probability to find the particle after time t in an interval
between Xl and X2 or, in the original coordinates, between Sl and S2. We therefore
have to evaluate (compare Section 2.5)
(4.21)
If pes, n) is a slowly varying function in the interval Sl' . . . , S2' we may replace the
sum by an integral
L:~ pes, n) = I
S2
Sl
pes, n) ds. (4.22)
We now pass over to the new coordinates X, t. We note that on account of (4.2)
I
ds = 2adx. (4.23)
L:~ pes, n) = J
X2
x,
I
p(x, t)2- dx.
a
(4.24)
1
2a p(x, t) = I(x, t) (4.25)
1 Note that through (4.15) and (4.1) x becomes a continuous variable as 'I" tends to zero!
84
4.1 A Model of Brownian Movement 73
L~~ P(s, n) = J
X2
XI
f(x, t) dx. (4.26)
(4.28)
When we compare the final form (4.28) with (2.21), we may identify
{(x, I)
85
74 4. Chance
~:~
...,..._ - j n =/
I ~o
J
~ ~'\ ~
PiS, n}
4
~ tel ~'\ ~
steps
p = 1 and q = O. (4.31)
In this case, the number, s, of successes coincides with the number, n, of hops
s = n. (4.32)
P(s, n) = bs •n • (4.33)
86
4.2 The Random Walk Model and Its Master Equation 75
because the Kronecker 8 becomes Dirac's 8-function in the continuum. Note that
limt _ o and lima _ o must be taken such that v = air remains finite.
Exercise on 4.1
Calculate the first two moments for (4.29) and (4.34) and compare their time
dependence. How can this be used to check whether nerve conduction is a diffusion
or a unidirectional process?
The probability, P(m; n + 1), of finding the particle at m after n + 1 pushes when
it has been at time n, at m - 1 is given by the product w(m, m - 1)· P(m - 1; n).
A completely analogous expression can be derived when the particle comes from
m + 1. Since the particle must come either from m - 1 or m + 1 and these are
independent events, P(m; n + 1) can be written down in the form
where
Note that we did not use (4.37) explicitly so that (4.36) is also valid for the general
case, provided
87
76 4. Chance
m', n) to find the particle at step n at point m' and at step n + 1 at m. This
probability consists of two parts as we have already discussed above. The particle
must be at step n at point m' which is described by the probability P(m'; n). Then it
must pass from m' to m which is governed by the transition probability w(m, m').
Thus the joint probability can generally be written in the form
1m - m'l = 1, (4.41)
it follows that w(m, m') = 0 unless m' = m ± 1 (in our example). We return to
(4.36) which we want to transform further. We put
and shall refer to w(m, m ± 1) as transition probability per second (or per unit
time). We subtract P(m; n) from both sides of (4.36) and divide both sides by r.
Using furthermore (4.38) the following equation results
P(m; n + 1) - P(m; n)
-'---'-----'----'--'----'- = w(m, m - I)P(m - 1; n) + w(m, m + I)P(m + 1; n)
r
- (w(m + 1, m) + w(m - 1, m))P(m; n). (4.42)
In our special example the w's are both equal to If(2r). We relate the discrete
variable n with the time variable t by writing t = m and accordingly introduce now
a probability measure P by putting
Pin (4.43) is an abbreviation for the function P(m, tfr:). We now approximate the
difference on the Ihs of (4.42) by the time derivative
P(m; n + 1) - P(m; n) dP
(4.44)
r: :::::: dt
dP(m, t) _ _
dt = w(m, m - I)P(m - 1, t) + w(m, m + I)P(m + 1, t)
88
4.2 The Random Walk Model and Its Master Equation 77
Let us now consider the stationary state in which the probability P is time-
independent. How can this be reached in spite of the hopping process going on all
the time? Consider a line between m and m + I. Then P does not change, if (per
unit time) the same number of transitions occur in the right and in the left direction.
This requirement is called the principle of detailed balance (Fig. 4.3) and can be
written mathematically in the form
I
w(m, m + I) = w(m, m - I) = 2,'
where the primes denote differentiation of P with respect to m. Because the first
and second terms ex eO or e cancel, the first nonvanishing term is that of the second
derivative of P. Inserting this result into (4.42) yields
dP 1 d 2P 2
-=--e (4.49)
dt 2,dm 2 •
89
78 4. Chance
(Why have we introduced Am and Ax here? (cf. Section 4.1).) Using furthermore
d a·d
dm = dx' e = 1, (4.51)
(4.52)
which we shall refer to as diffusion equation. As we will see much later (cf. Section
6.3), this equation is a very simple example of the so-called Fokker-Planck equation.
°
The solution of (4.52) is, of course, much easier to find than that of the master
equation. The e-expansion relies on a scaling property. Indeed, letting e ----+ can be
identified with letting the length scale a go to zero. We may therefore repeat the
above steps in a more formal manner, introducing the length scale a as expansion
parameters. We put (since Ax/Am = a)
Inserting (4.53a) and (4.53b) into (4.42) or (4.47) yields, after expanding f up to
second order in a, again (4.52). The above procedure is only valid under the con-
dition that for fixed a, the function fvaries very slowly over the distance Ax =i' a.
Thus the whole procedure implies a self-consistency requirement.
Exercises on 4.2
1) Show that pes, n) (4.4) fulfils (4.36) with (4.37) if s, n are correspondingly
related to m, n. The initial condition is P(m,O) = bm,o (i.e. = I for m = 0, and
°
= for m 0). "*
2) Verify that (4.29) satisfies (4.52) with the initial condition f(x, 0) = b(x) (15:
Dirac's b-function).
3) Derive the generalization of equation (4.52) if w(m, m + 1) "*w(m, m - 1).
4) Excitation transfer between two molecules by a hopping process. Let us denote
the two molecules by the indexj = 0, I and let us denote the probability to find
the molecule j in an excited state at time t by P(j, t). We denote the transition
rate per unit time by w. The master equation for this process reads
90
4.3 Joint Probability and Paths. Markov Processes. The Chapman-Kolmogorov Equation 79
(E.5)
To explain the main ideas of this section, let us first consider the example of the
two foregoing sections : A particle hopping randomly along a chain. When we
follow up the different positions the particle occupies in a specific realization, we
may draw the plot of Fig. 4.4a. At times fl ' f2' . .• , fn the particle is found at n
specific positions m l , m 2 , • • • , m n • If we repeat the experiment, we shall find an-
other realization, say that of Fig. 4Ab. We now ask for the probability of finding
the particle at times fl' f2' •.. at the corresponding points m l , m 2 , • • • We denote
this probability by
(4.54)
m m
7
6 ·.
7
5
.,
5
~
/ "- r'\. /
·· 5
~
.,
J '\d"' 3 r'\. V I'\.!
i\.. V
;
I ·
2
/
~.
.
o o Fig. 4.4. Two paths of a Brownian
-1 I, 1; 1J ~ 15 ~ 17 -/ 1, 12 ~ ~ 15 ~ 17 particle t1 = 0, m1 = 4, ... ,
n=7
91
80 4. (]hance
Apparently Pn is a joint probability in the sense of Section 2.6. Once we know the
joint probability with respect to n different times, we can easily find other prob-
ability distributions containing a smaller number of arguments m j , I j • Thus for
example if we are interested only in the joint probability at times I I, . • • , In-I we
have to sum up (4.54) over mn
(4.55)
Another interesting probability is that for finding the particle at time In at the
position mn and at time I I at position m l , irrespective of the positions the particle
may have acquired at intermediate times. We find the probability by summing
(4.54) over all intermediate positions:
(4.56)
(4.57)
(4.59)
If the conditional probability obeys the equation (4.59) the corresponding process
is called a Markov process. In the following we want to derive some general rela-
tions for Markov processes.
The rhs of (4.59) is often referred to as transition probability. We now want to
express the joint probability (4.54) by means of the transition probability (4.59).
In a first step we multiply (4.58) by Pn - I
(4.60)
92
4.3 Joint Probability and Paths. Markov Processes. The Chapman-Kolmogorov Equation 81
(4.61)
(4.62)
and
(4.63)
Thus the joint probability of a Markov process can be obtained as a mere product
of the transition probabilities. Thus the probability of finding for example a
particle at positions mj for a time sequence tj can be found by simply following up
the path of the particle and using the individual transition probabilities from one
point to the next point.
To derive the Chapman-Kolmogorov equation we consider three arbitrary times
subject to the condition
(4.65)
(4.66)
We now use again formula (4.62) which is valid for any arbitrary time sequence.
In particular we find that for the lhs of (4.65)
(4.67)
(4.68)
Inserting (4.67) and (4.68), we find a formula in which on both sides Pt(m t , t 1 )
appears as a factor. Since this initial distribution can be chosen arbitrarily, that
relation must be valid without this factor. This yields as final result the Chapman-
Kolmogorov equation
(4.69)
93
82 4. Chance
Note that (4.69) is not so innocent as it may look because it must hold for any
time sequence between the initial time and the final time.
The relation (4.69) can be generalized in several ways. First of alI we may replace
mj by the M-dimensional vector mj' Furthermore we may let mj become a con-
(4.70)
(4.71)
for continuous variables. (Here and in the following we drop the index I of P).
We now proceed to consider infinitesimal time intervals, i.e., we put I =
t ' + T. We then form the following expression
I
- {P(m, t
T
+ T) - P(m, t)}, (4.74)
and let T go to zero or, in other words, we take the derivative of (4.72) with respect
to the time t, which yields
where p,(m, m') = lim, ... o r- 1 (pt+<.t(m, m') - ptim, m'». The discussion of Pt
requires some care. If m' #- m, P, is nothing but the number of transitions from m'
to m per second for which we put
or, in other words, w is the transition probability per second (or unit time). The
discussion of p,(m, m) with the same indices m = m' is somewhat more difficult
because it appears that no transition would occur. For a satisfactory discussion we
must go back to the definition of P according to (4.59) and consider the difference
94
4.3 Joint Probability and Paths. Markov Processes. The Chapman-Kolmogorov Equation 83
Inserting now (4.76) and (4.78) into (4.75) leaves us with the so-called master
equation
The reason for this new notation is the following: we subject the solution of (4.80)
to the initial condition:
That is we assume that a time ( equal to the initial time (' the particle is with cer-
tainty at the space point q'. Therefore p",(q, q') dq is the conditional probability to
find the particle at time ( in the interval q ... q + dq provided it has been at time
t' at point q' or, in other words, p".(q, q') is the transition probability introduced
above. For what follows we make the replacements
(4.82)
95
84 4. Chance
and
(4.83)
By inserting (4.84) into (4.80) one readily verifies that (4.84) fulfils this equation.
We further note that for r --> 0 (4.84) becomes a b-function (see Fig. 2.10). It is
now a simple matter to find an explicit expression for the joint probability (4.54).
To this end we need only to insert (4.84) (with j = 1, 2, ... ) into the general
formula (4.64) which yields
= (2~D~)-(n-I)/2
" , exp{ __l_ "'~-I
2Dr L.,J = 1
(q.J + 1 -
qj )2}P(q I, t)
1 . (4.85)
This can be interpreted as follows: (4.85) describes the probability that the
particle moves along the path qlq2q3 .... We now proceed to a continuous time
scale by replacing
r -> dt (4.86)
and
(4.87)
(4.88)
(4.89)
96
4.4 How to Use Joint Probabilities. Moments. Characteristic Function. Gaussian Processes 85
which is called a path integral. Such path integrals playa more and more important
role in statistical physics. We will come back to them later in our book.
Exercise on 4.3
Determine
for the random walk model of Section 4.1, for the initial condition P(m h t 1) = I
for m l == 0; = 0 otherwise.
(4.90)
where we now drop the index n of Pn• Using (4.90) we can define moments by
generalization of concepts introduced in Section 2.7
In this equation some ofthe v's may be equal to zero. A particularly important case
is provided by
(4.92)
where we have only the product of two powers of m at different times. When we
perform the sum over all other m's according to the definition (4.91) we find the
probability distribution which depends only on the indices I and n. Thus (4.92)
becomes
(4.93)
We mention a few other notations. Since the indicesj = 1 ... n refer to times we
can also write (4.92) in the form
(4.94)
This notation must not be interpreted that m is a given function of time. Rather,
In and 11 must be interpreted as indices. If we let the I/S become a continuous time
sequence we will drop the sufficesj and use tor t' as index so that (4.94) is written
97
86 4. Chance
in the form
(4.95)
(4.97)
(4.98)
(4.100)
(4.101)
(4.102)
We call a process Gaussian if all cumulants except the first two vanish, i.e., if
(4.103)
holds. In the case of a Gaussian process, the characteristic function thus can be
98
4.4 How to Use Joint Probabilities. Moments. Characteristic Function. Gaussian Processes 87
According to (4.101) all moments can be expressed by the first two cumulants or,
because k[ and k2 can be expressed by the first and second moment, all higher
moments can be expressed by the first two moments.
The great usefulness of the joint probability and of correlation functions for
example of the form (4.97) rests in the fact that these quantities allow us to discuss
time dependent correlations. Consider for example the correlation function
and let us assume that the mean values at the two times t and t'
vanish. If there is no correlation between q's at different times we can split the
joint probability into a product of probabilities at the two times. As a consequence
(4.105) vanishes. On the other hand if there are correlation effects then the joint
probability cannot be split into a product and (4.105) does not vanish in general.
We will exploit this formalism later on in greater detail to check how fast fluctua-
tions decrease, or, how long a coherent motion persists. If the mean values of q
do not vanish, one can replace (4.105) by
to check correlations.
Exercise on 4.4
Calculate the following moments (or correlation functions) for the diffusion process
under the initial conditions
1) P(q, t = 0) = ~(q),
99
88 4. Chance
Since the "rate in" consists of all transitions from initial points m' to m, it is
composed of the sum over the initial points. Each term of it is given by the prob-
ability to find the particle at point m', multiplied by the transition probability per
unit time to pass from m' to m. Thus we obtain
which is called the master equation (Fig. 4.5). The crux to derive a master equation
is not so much writing down the expressions (4.109) and (4.110), which are rather
obvious, but to determine the transition rates w explicitly. This can be done in two
ways. Either we can write down the w's by means of plausibility arguments. This
has been done in the example of Section 4.2. Further important examples will be
given later, applied to chemistry and sociology. Another way, however, is to derive
the w's from first principles where mainly quantum statistical methods have to be
used.
Exercise on 4.5
Why has one to assume a Markov process to write down (4.109) and (4.11O)?
100
4.6 Exact Stationary Solution of the Master Equation for Systems in Detailed Balance 89
In physics, the principle of detailed balance can be (in most cases) derived for
systems in thermal equilibrium, using microreversibility. In physical systems far
from thermal equilibrium or in nonphysical systems, it holds only in special case
(for a counter example cf. exercise 1).
Equation (4.112) represents a set of homogeneous equations which can be
solved only if certain conditions are fulfilled by the w's. Such conditions can be
derived, for instance, by symmetry considerations or in the case that w can be
replaced by differential operators. We are not concerned, however, with this
question here, but want to show how (4.112) leads to an explicit solution. In the
following we assume that pen) '# O. Then (4.112) can be written as
P(m) w(m, n)
(4.113)
pen) = wen, m)"
(4.114)
Putting
101
90 4. Chance
(4.116)
Because the solution was assumed to be unique, l[J(nN) is independent of the path
chosen. Taking a suitable limit one may apply (4.116) to continuous variables.
As an example let us consider a linear chain with nearest neighbor transitions.
Since detailed balance holds (cf. exercise 2) we may apply (4.114). Abbreviating
transition probabilities by
we find
P(m )
,11;;:::::',;;;:::::;; ;;;1, , • m
Fig. 4.6. Example of P(m) showing maxima and
minima
(4.121)
P(mo) is a maximum, if
102
4.6 Exact Stationary Solution of the Master Equation for Systems in Detailed Balance 91
(4.123)
In both cases, (4.122) and (4.123), the numbers m belong to a finite surrounding of
mo·
Exercises on 4.6
1) Verify that the process depicted in Fig. 4.7 does not allow for detailed balance.
Ls
3
N-m+1
w(m, m - I) = N Wo for 1::s; m ::s; N
103
92 4. Chance
Show that P(m) is the Poisson distribution 1I:k,,..( == ak) (2.57) with m +-+ k and
p. +-+ (1./p.
Hint: Determine P(O) by means of the normalization condition
(4.124)
The master equation represents a set of linear differential equations of first order.
To transform this equation into an ordinary algebraic equation we put
(4.126)
(4.127)
The additional index (1. arises because these algebraic equations allow for a set of
eigenvalues A. and eigenstates rp,. which we distinguish by an index (1.. Since in general
the matrix Lao,. is not symmetric the eigenvectors of the adjoint problem
"
~m
x(rt) L
III mn
= - A.«An
v(rt) (4.128)
(4.129)
In (4.129) the lhs is an abbreviation for the sum over ft. With help of the eigen-
vectors rp and X, Lm.. can be written in the form
L ... = -"
L.Jz A.«'I'.
rn(rt)x(rt)
_ . (4.130)
We now show that the matrix occurring in (4.127) can be symmetrized. We first
104
4.7 The Master Equation with Detailed Balance 93
p l / 2 (n)
L!..n = w(m, n) pl/2(m) (4.131)
for m "# n. For m = n we adopt the original form (4.125). Pen) is the stationary
solution of the master equation. It is assumed that the detailed balance condition
pl/2(m)
L~ ... = wen, m) pl/2(n)' (4.133)
Pen) pl/2(m)
(4.133) = w(m, n) P(m)' pI/2(n) (4.134)
L!n .• (4.135)
so that the symmetry is proven. To show what this symmetrization means for
(4.127), we put
which yields
I /2 -(a)
".t...ra L mil p (n) pI/2 (4.137)
CPII = - 1
Aa
-(a)
(m) CP ...
Xn(a) = p- 1/2 X
(n)
-(a)
"'
(4.139)
"L.,.m X-calLs
.. II1II
= -A.u..a
.y(a) • (4.140)
The i's may be now identified with the q,'s because the matrix L':... is symmetric.
105
94 4. Chance
This fact together with (4.136) and (4.139) yields the relation
(4.141)
Because the matrix L' is symmetric, the eigenvalues A. can be determined by the
following variational principles as can be shown by a well-known theorem of linear
algebra. The following expression must be an extremum:
(XL'i)}
-A. = Extr. { (XX) = Extr. {(XLcP)}
(xcp) • (4.142)
(4.143)
the eigenvalue A. = 0 associated with the stationary solution results. We now show
that all eigenvalues A. are nonnegative. To this end we derive the numerator in
(4.142)
(4.144)
(compare (4.139), (4.141), and (4.131» in a way which demonstrates that this
expression is nonpositive. We multiply w(m, n)P(n) ~ 0 by -lj2(X... - X..)2 ::;; 0
so that we obtain
In the second sum we exchange the indices m, n and apply the condition of detailed
balance (4.132). We then find that the second sum equals the first sum, thus that
(4.146) agrees with (4.144). Thus the variational principle (4.142) can be given the
form
1
I\.
=E
X
t
r
{t I ..... (x... - x.)2 w(m, n)p(n)} > 0'
I .. x;P(n) - .
(4.147)
106
4.8 Kirchhoff's Method of Solution of the Master Equation 9S
required by detailed balance cannot be fulfilled. Thus other methods for a solution
of the master equation are necessary. We confine our treatment to the stationary
solution in which case the master equation, (4.111), reduces to a linear algebraic
equation. One method of solution is provided by the methods of linear algebra.
However, that is a rather tedious procedure and does not use the properties inherent
in the special form of the master equation. We rather present a more elegant method
developed by Kirchhoff, originally for electrical networks. To find the stationary
solution Pen) of the master equation 2 •
2 When we use n instead of the vector, this is not a restriction because one may always rearrange
a discrete set in the form of a sequence of single numbers.
107
96 4. Chance
3 3 3
7
L/\~ 2 7 2 7 2
Fig. 4.9. The maximal trees T(G)
belonging to the graph of Fig. 4.8a
~ucn
ZV171L Fig. 4.10. The maximal trees T(G)
belonging to the graph G of Fig. 4.8b
obtain the maximal trees of Fig. 4.8b one has to drop in Fig. 4.8b either one side
and the diagonal or two sides.
We now define a directed maximal tree with index n, Tn(G). It is obtained from
T( G) by directing all edges of T( G) towards the vertex with index n. The directed
maximal trees belonging to n = I, Fig. 4.9, are then given by Fig. 4.11. After these
preliminaries we can give a recipe how to construct the stationary solution Pen).
To this end we ascribe to each directed maximal tree a numerical value called A:
A (Tn(G)): This value is obtained as product of all transition rates wen, m) whose
edges occur in Tn(G) in the corresponding direction. In the example of Fig. 4.11
we thus obtain the following different directed maximal trees
I
~/\L
T(I)
'I
2 I
r,(2)
I
2 7
T,(3}
7
2
Fig. 4.11. The directed maximal
trees Tl belonging to Fig. 4.9 and
n = I
108
4.9 Theorems about Solutions of the Master Equation 97
(It is best to read all arguments from right to left). Note that in our example Fig.
4.11 w(3, 2) = w(1, 3) = O. We now come to the last step. We define Sn as the sum
over all maximal directed trees with the same index n, i.e.,
(4.154)
Sn
Pn = ,,\,N S' (4.156)
L-'=1 I
w(1, 2)w(2, 3)
Pl=-w(~I-,2~)w-(~2,~3~)-+-w-(~2,-3~)w~(~3,~I~)-+-w~(-3,-I~)w~(l~,~2)' (4.157)
Though for higher numbers of vertices this procedure becomes rather tedious, at
least in many practical cases it allows for a much deeper insight into the construc-
tion of the solution. Furthermore it permits decomposing the problem into several
parts for example if the master equation contains some closed circles which are
only connected by a single line.
Exercise on 4.8
Consider a chain which allows only for nearest neighbor transitions. Show that
Kirchhoff's formula yields exactly the formula (4.119) for detailed balance.
109
98 4. Chance
Lm P(rn, 0) = 1, (4.159)
and
Lm P(m, t) = I. (4.161)
Thus the normalization property and the positiveness of probability are guar-
anteed for all times.
4) Inserting P(rn, t) = am exp( - At) into the master equation yields a set of linear
algebraic equations for am with eigenvalues A. These eigenvalues A have the
following properties;
a) The real part of A is nonnegative Re A :2: 0,
b) if detailed balance holds, all A's are purely real.
5) If the stationary solution is unique, then the time-dependent solution P(m, t)
for any initial distribution pOem) (so that P(m, 0) = pOem»~ tends for t -> 00
to the stationary solution. For a proof by means of the information gain com-
pare exercise 1 of Section 5.3.
110
4.10 The Meaning of Random Processes. Stationary State, Fluctuations, Recurrence Time 99
numbers 1, ... ,N, with the same probability liN, and repeat this selection process
regularly in time intervals 1:. If the number is selected, one removes the correspond-
ing ball from its present box and puts it into the other one. We are then interested
in the change of the numbers Nl and N2 in the course of time. Since N2 is given
by N - Nt, the only variable we must consider is N l . We denote the probability of
finding Nt after s steps at time t (i.e., t = S1:) by peNt, s).
We first establish an equation which gives us the change of the probability
distribution as a function of time and then discuss several important features. The
probability distribution, P, is changed in either of two ways. Either one ball is
added to box A or removed from it. Thus the total probability peNt, s) is a sum of
the probabilities corresponding to these two events. In the first case, adding a ball
A, we must start from a situation in which there are Nt - 1 balls in A. We denote
the probability that one ball is added by w(Nl' Nt - 1). If a ball is removed, we
must start from a situation in which there are Nt + 1 balls in A at "time" s - 1.
We denote the transition rate corresponding to the removal of a ball by
w(Nl, Nl + 1). Thus we find the relation
Since the probability for picking a definite number is liN and there are N2 + 1
N - Nt + 1 balls in urn B the transition rate weNt, Nt - 1) is given by
(4.163)
(4.164)
N - Nl + 1 Nt + 1
P(Nt,s) = N peNt - 1, s - 1) + ~ P(Nl + 1, s - 1).
(4.165)
If any initial distribution of balls over the urns is given we may ask to which
final distribution we will come. According to Section, 4.9 there is a unique final
stationary solution to which any initial distribution tends. This solution is given by
(4.166)
(4.167)
111
100 4. Chance
(4.168)
(4.169)
Thus, in the language of probability theory, a single event consists of the sequence
(4.169). Once this sequence is picked, there is nothing arbitrary left. On the other
hand, when talking about probability, we treat a set of events, i.e., the sample set
(sample space). In thermodynamics, we call this set "an ensemble" (including its
changes in course of time). An individual system corresponds to a sample point.
In thermodynamics, but also in other disciplines, the following question is
treated: If a system undergoes a random process for a very long time, does the tem-
poral average coincide with the ensemble average? In our book we deal with the
ensemble average if not otherwise noted. The ensemble average is defined as the
mean value of any function of random variables with respect to the joint probability
of Section 4.3.
In the following discussion the reader should always carefully distinguish
between a single system being discussed, or the whole ensemble. With this warning
in mind, let us discuss some of the main conclusions: The stationary solution is
by no means completely sharp, i.e., we do not find NI2 balls in box A and NI2
balls in box B with probability unity. Due to the selection process, there is always
a certain chance of finding a different number, N! t= N12, in box A. Thus fluctua-
tions occur. Furthermore, if we had initially a given number N!, we may show that
the system may return to this particular number after a certain time. Indeed, to
each individual process, say N! ~ N! + 1, there exists a finite probability that the
inverse process, N! ~ N! - 1, occurs (Evidently this problem can be cast into a
more rigorous mathematical formulation but this is not our concern here). Thus
the total system does not approach a unique equilibrium state, N! = N12. This
seems to be in striking contrast to what we would expect on thermodynamic
grounds, and this is the difficulty which has occupied physicists for a long time.
It is, however, not so difficult to reconcile these considerations with what we
expect in thermodynamics, namely, the approach to equilibrium. Let us consider
112
4.1 0 The Meaning of Random Processes. Stationary State, Fluctuations, Recurrence Time 101
the case that N is a very large number, which is a typical assumption in thermo-
dynamics (where one even assumes N -+ (0). Let us first discuss the stationary
distribution function. If N is very large we may convince ourselves very quickly that
it is very sharply peaked around NI = N12. Or, in other words, the distribution func-
tion effectively becomes a <5-function. Again this sheds new light on the meaning of
entropy. Since there are N!INI !Nz ! realizations of state N I , the entropy is given by
N!
S = kBln N 'N , (4.170)
l' z·
or
(4.171)
where
(4.173)
Since the probability distribution P(N I ) is strongly peaked, we shall find in all
practical cases NI = NI2 realized so that when experimentally picking up any
distribution we may expect that the entropy has acquired its maximum value where
PI = Pz = 1/2. On the other hand we must be aware that we can construct other
initial states in which we have the condition that NI is equal a given number
No =1= N12. If No is given, we possess a maximal knowledge at an initial time. If
we now let time run, P will move to a new distribution and the whole process has the
character of an irreversible process. To substantiate this remark, let us consider the
development of an initial distribution where all balls are in one box. There is a
probability equal to one that one ball is removed. The probability is still close to
unity if only a small number of balls has been removed. Thus the system tends
very quickly away from its initial state. On the other hand, the probability for
passing a ball from box A to box B becomes approximately equal to the probability
for the inverse process if the boxes are equally filled. But these transition prob-
abilities are by no means vanishing, i.e., there are still fluctuations of the particle
numbers possible in each box. If we wait an extremely long time, fro there is a finite
probability that the whole system comes back to its initial state. This recurrence
time, fro has been calculated to be
(4.174)
where fr is defined as the mean time between the appearance of two identical
macrostates.
113
102 4. Chance
Exercises on 4.10
I) Why is (12 = <N;> - <Nt >2 a measure for the fluctuations of Nt? How large
is (12?
2) Discuss Brownian movement by a single system and by an ensemble.
(4.175)
We have encountered such indices already in Section 3.3 and 3.4. In particular
we have seen that (Sect. 3.5) the entropy S is constructed from the probability
distribution, Pi' whereas the entropy of system S' is determined by a second prob-
ability distribution Pi" Furthermore we have assumed that the entropies are
additive. This additivity implies that PH' factorizes (cf. Exercise 1 on 3.5)
(4.176)
Inserting (4.176) into (4.175) shows after some inspection that (4.175) can be solved
by (4.176) only under very special assumptions, namely, if
(4.177)
114
4.11 Master Equation and Limitations of Irreversible Thermodynamics 103
(4.177) implies that there is no interaction between the two subsystems. The
hypothesis (4.176) to solve (4.175) can thus only be understood as an approxima-
tion similar to the Hartree-Fock approximation in quantum theory (compare
exercise below). Our example shows that irreversible thermodynamics is valid only
if the correlations between the two subsystems are not essential and can be taken
care of in a very global manner in the spirit of a self-consistent field approach.
Exercise on 4.11
Formulate a variational principle and establish the resulting equations for Pi'
pi, in the case that the master equation obeys detailed balance.
Hint: Make the hypothesis (4.176), transform the master equation to one with a
self-adjoint operator (according to Sect. 4.7) and vary now the resulting expres-
sion (4.142) with respect to pipi, = Xii" For those who are acquainted with the
Hartree-Fock procedure, convince yourself that the present procedure is identical
with the Hartree-Fock self-consistent field procedure in quantum mechanics.
115
5. Necessity
Old Structures Give Way to New Structures
dv
m dt = Fo· (5.1)
Let us further assume that the force Fo may be decomposed into a "driving force"
F and a friction force which we assume proportional to the velocity, v. Thus we
replace Fo by
~ o (a)
F= -k.q
--------~~------- q
(b)
The minus sign results from the fact that the elastic force tends to bring the particle
back to its equilibrium position. We express the velocity by the derivative of the
coordinate with respect to time and use a dot above q to indicate this
dq .
v = dt == q. (5.5)
We shall use this equation to draw several important general conclusions which
are valid for other systems and allow us to motivate the terminology. We mainly
consider a special case in which m is very small and the damping constant')' very
large, so that we may neglect the first term against the second term on the lhs of
(5.6). In other words, we consider the so-called overdamped motion. We further
note that by an appropriate time scale
t = ')'t' (5.7)
we may eliminate the damping constant ')'. Eq. (5.6) then acquires the form
Ii = F(q). (5.8)
Equations of this form are met in many disciplines. We illustrate this by a few
examples: In chemistry q may denote the density of a certain kind Q of molecules,
118
5.1 Dynamic Processes 107
which are created by the reaction of two other types of molecules A and B with
concentrations a and b, respectively. The production rate of the density q is de-
scribed by
q = kab (5.9)
q = kaq. (5.10)
Equations of the type (5.10) occur in biology where they describe the multiplica-
tion of cells or bacteria, or in ecology, where q can be identified with the number of a
given kind of animals. We shall come back to such examples in Section 5.4 and,
in a much more general form, later on in our book in Chapters 8 through 11. For the
moment we want to exploit the mechanical example. Here one introduces the nota-
tion of "work" which is defined by work = force times distance. Consider for
example a body of weight G (stemming from the gravitational force that the earth
exerts on the body). Thus we may identify G with the force F. When we lift this
body to a height h (== distance q) we "do work"
In general cases the force F depends on the position, q. Then (5.11) can be formu-
lated only for an infinitesimally small distance, dq, and we have instead of (5.11)
dW = F(q)dq. (5.12)
To obtain the total work over a finite distance, we have to sum up, or, rather to
integrate
W = Jql
qo
F(q)dq. (5.13)
dV
F(q) = - dq' (5.14)
Let us consider the example of the harmonic oscillator with F given by (5.4).
One readily establishes that the potential has the form
119
108 5. Necessity
when lifting a weight. This suggests interpreting the solid curve in Fig. 5.2 as the
slope of a hill. When we bring the particle to a certain point on the slope, it will
fall back down the slope and eventually come to rest at the bottom of the hill.
Due to the horizontal slope at q = 0, F(q), (5.14) vanishes and thus 4 = O. The
particle is at an equilibrium point. Because the particle returns to this equilibrium
point when we displace the particle along the slope, this position is stable.
V(q) = G·h
Now consider a slightly more complicated system, which will turn out later on
to be of fundamental importance for self-organization (though this is at present not
at all obvious). We consider the so-called anharmonic oscillator which contains a
cubic term besides a linear term in its force F.
(5.16)
(5.17)
The potential is plotted in Fig. 5.3 for two different cases, namely, for k > 0 and
Vfq). k>O
(a)
------------~~~-----------.q
Vfq). k<O
(b)
--~--~----~~~----~--~-+ q
Fig. 5.3a and b. The potential of the force
(5.16) for k > O (a) and k < 0 (b)
120
5.1 Dynamic Processes 109
4= O. (S.l8)
From Fig. S.3, it is immediately clear that we have two completely different situa-
tions corresponding to whether k > 0 or k < O. This is fully substantiated by an
algebraic discussion of (S.17) under the condition (S.18). The only solution in case
a) k > 0, kl > 0 is
q = 0, stable, (S.19)
whereas in case
b) k < 0, kl > 0,
we find three solutions, namely, q = 0 which is evidently unstable, and two stable
solutions ql,2 so that
q.
There are now two physically stable equilibrium positions, (Fig. S.4). In each of
them the particle is at rest and stays there for ever.
By means of (S.I 7) we may now introduce the concept of symmetry. If we
replace everywhere in (S.17) q by -q we obtain
(S.17a)
or, after division of both sides by -1 , we obtain the old (S.17). Thus (S.17) remains
unchanged (invariant) under the transformation
q --* -q (S.17b)
or, in other words, (S.17) is symmetric with respect to the inversion q --+ - q.
Simultaneously the potential
(S.21)
121
110 5. Necessity
°
critical slowing down. For later purposes we shall call the coordinate q now r.
When passing from k > 0 to k < the stable position r = 0 is replaced by an un-
stable position at r = 0 and a stable one at r = roo Thus we have the scheme
unstable point
stable point / (5.22)
~
stable point
Since this scheme has the form of a fork, the whole phenomenon is called "bifurca-
tion." Another example of bifurcation is provided by Fig. 5.7 below where the
two points (stable at r 1, unstable at ro) vanish when the potential curve is deformed.
b) Limit Cycles
We now proceed from our one-dimensional problem to a two-dimensional problem
as follows: We imagine that we rotate the whole potential curve V(r) around the
V-axis which gives us Fig. 5.5. We consider a case in which the particle runs along
the bottom of the valley with a constant velocity in tangential direction. We may
either use cartesian coordinates q 1 and qz or polar coordinates (radius r and angle
122
5.1 Dynamic Processes 111
cp)l. Since the angular velocity, cp, is constant, the equations of motion have the
form
r = F(r), (5.23)
cp = w.
We do not claim here that such equations can be derived for purely mechanical
systems. We merely use the fact that the interpretation of V as a mechanical
potential is extremely convenient for visualizing our results. Often the equations of
motion are not given in polar coordinates but in cartesian coordinates. The relation
between these two coordinate systems is
q! =rcoscp,
q2 = r sin cp. (5.24)
Since the particle moves along the valley, its path is a circle. Having in mind the
potential depicted in Fig. 5.5 and letting start the particle close to q = 0 we see that
it spirals away to the circle of Fig. 5.6. The point q = 0 from which the particle
spirals away is called an unstable focus. The circle which it ultimately approaches is
called a limit cycle. Since our particle 'also ends up in this cycle, if it is started from
the outer side, this limit cycle is stable. Of course, there may be other forms of the
potential in radial direction, for example that of Fig. 5.7a now allowing for a
unstablr!
focus
stable limit
cycle
Fig. 5.6. Unstable focus and limit cycle
vrr) Vfr)
123
112 5. Necessity
stable and an unstable limit cycle. When this potential is deformed, these two
cycles coalesce and the limit cycles disappear. We then have the following bifurca-
tion scheme
124
5.2 Critical Points and Trajectories in a Phase Plane. Once Again Limit Cycles 113
mq = p (p = momentum) (5.30)
(5.32)
The meaning of (5.27) and (5.28) becomes clearer when we write ql in the form
(5.34)
where
2 This "division" is done here in a formal way, but it can be given a rigorous mathematical basis,
which, however, we shall not discuss here.
125
114 5. Necessity
(5.37)
This form (and an equivalent one for q2) lends itself to the following interpretation:
If ql and q2 are given at time t, then their values can be determined at a later time
t + 't by means of the rhs of (5.37) and
(5.38)
uniquely, which can also be shown quite rigorously. Thus when at an initial time ql
and q2 are given, we may proceed from one point to the next. Repeating this pro-
cedure, we find a unique trajectory in the ql,q2-plane (Fig. 5.8). We can let this
trajectory begin at different initial values. Thus a given trajectory corresponds to
an infinity of motions differing from each other by the "phase" (or initial value)
ql(O), q2(0).
qil+r) - - T
Let us now choose other points q I' q2 not lying on this trajectory but close to it.
Through these points other trajectories pass. We thus obtain a "field" of trajec-
tories, which may be interpreted as streamlines of a fluid. An important point to
discuss is the structure of these trajectories. First it is clear that trajectories can
never cross each other because at the crossing point the trajectory must continue
in a unique way which would not be so if they split up into two or more trajectories.
The geometrical form of the trajectories can be determined by eliminating the time-
dependence from (5.27), (5.28) leading to (5.33). This procedure breaks down,
however, if simultaneously
(5.39)
for a couple q?, q~. In this case (5 .33) results in an expression % which is mean-
ingless. Such a point is called a singular (or critical) point. Its coordinates are deter-
mined by (5 .39). Due to (5.27) and (5.28), (5 .39) implies 41 = 42 = 0, i.e., the
singular point is also an equilibrium point. To determine the nature of equilibrium
126
5.2 Critical Points and Trajectories in a Phase Plane. Once Again Limit Cycles 115
(stable, unstable, neutral), we have to take into account trajectories close to the
singular point. We call each singular point asymptotically stable if all trajectories
starting sufficiently near it tend to it asymptotically for t -4 00 (compare Fig. 5.9a).
A singular point is asymptotically unstable if all trajectories which are sufficiently
close to it tend asymptotically (t -4 - 00) away from it. Interpreting trajectories as
streamlines, we call singular points which are asymptotically stable, "sinks", be-
cause the streamlines terminate at them. Correspondingly, asymptotically unstable
singular points are called "sources".
(c)
1---+---~-+---r--;---~~
127
116 5. Necessity
41 = ql'
4z = aqz, (5.40)
(5.41)
41 = -ql'
4z = -aqz (5.42)
dq2 aq2
dql = q; (5.44)
(5.45)
which can also be obtained from (5.41) and (5.43) by the elimination of time.
For a > 0 we obtain parabolic integral curves depicted in Fig. 5.9a. In this case we
have for the slope dq2/dql
dq2
-=
Caqla-I . (5.46)
dql
We now distinguish between positive and negative exponents of ql' If a > 1 we
find dq2/dql -> 0 for ql -> O. Every integral curve with exception of the qz-axis
approaches the singular point along the q I axis. If a < lone immediately establishes
that the roles of ql and q2 are interchanged. Singular points which are surrounded
by curves of the form of Fig. 5.9a are called nodes. For a = I the integral curves
are half lines converging to or radiating from the singular point. In the case of the
node every integral curve has a limiting direction at the singular point.
We now consider the case a < O. Here we find the hyperbolic curves
(5.47)
128
5.2 Critical Points and Trajectories in a Phase Plane. Once Again Limit Cycles 117
------------+------------~ ------~
(c)
-+--+-+---~f---+_+_~ Q,
For a = -1 we have ordinary hyperbolas. The curves are depicted in Fig. 5.10. Only
four trajectories tend to the singular point namely As, Ds for t --+ 00, D s, Cs for
t --+ - 00. The corresponding singular point is called saddle point.
(5.48)
(5.49)
For a> 0, by use of polar coordinates (compare (5.24», (5.48) and (5.49) acquire
the form
;- = -ar (5.50)
11 = 1, (5.51)
129
118 5. Necessity
(5.52)
(5.53)
V(r)
--------~~~------~r
(a) (c)
---+--+-+--+- -+-+-+----q,
We now turn to the general case. As mentioned above we assume that we can
expand FI and F2 around q? = 0 and q~ = 0 into a Taylor series and that we
may keep as leading terms those linear in ql and q2 , i.e.,
130
5.2 Critical Points and Trajectories in a Phase Plane. Once Again Limit Cycles 119
e = aql + /3q2
tT = "Iql + ~q2 (5.56)
(5.57)
(5.58)
The further discussion may be performed as above, if AI' A2 are real. If AI( = Ai)
are complex, inserting ¢ = 1'/* = rei'" into (5.57), (5.58) and separating the equa-
tions into their real and imaginary parts yields equations of the type (5.50), (5.51)
(with a = - Re A1> and + 1 replaced by 1m AI)'
Let us now discuss a further interesting problem, namely. How to find limit
cycles in the plane: The Poincare-Bendixson theorem. Let us pick a point qO =
(q?, qg) which is assumed to be nonsingular, and take it as initial value for the
solution of (5.27), (5.28). For later times, t > to, q(t) will move along that part of
the trajectory which starts at q(t o) = qO. We call this part half-trajectory. The
Poincare-Bendixson theorem then states: If a half-trajectory remains in a finite
domain without approaching singular points, then this half-trajectory is either a
limit cycle or approaches such a cycle. (There are also other forms of this theorem
available in the literature). We do not give a proof here, but discuss some ways
of applying it invoking the analogy between trajectories and streamlines in hydro-
dynamics. Consider a finite domain called D which is surrounded by an outer and
an inner curve (compare Fig. 5.13). In more mathematical terms: D is doubly con-
nected. If all trajectories enter the domain and there are no singular points in it or
131
120 5. Necessity
on its boundaries, the conditions of the theorem are fulfilled 3 . When we let
shrink the inner boundary to a point we see that the conditions of the Poincare-
Bendixson theorem are still fulfilled if that point is a source.
(5.59)
The case in which the forces F j can be derived from a potential V by means of
F. = __
8 V---,(~ql:-'_'_.q=n) (5.60)
J 8qj
8Fj 8Fk
for all j, k = 1, ... , n. (5.61)
8qk = 8qj
The system is in equilibrium if (5.60) vanishes. This can happen for certain points
q?, ... ,q~, but also for lines or for hypersurfaces. When we let V depend on
parameters, quite different types of hypersurfaces of equilibrium may emerge
from each other, leading to more complicated types of bifurcation. A theory of
some types has been developed by Thorn (cf. Sect. 5.5).
Exercise on 5.2
5.3* Stability
The state of a system may change considerably if it loses its stability. Simple ex-
amples were given in Section 5.1 where a deformation of the potential caused an
instability of the original state and led to a new equilibrium position of a "fictitious
particle". In later chapters we will see that such changes will cause new structures
on a macroscopic scale. For this reason it is desirable to look more closely into the
question of stability and first to define stability more exactly. Our above examples
referred to critical points where qj = O. The idea of stability can be given a much
more general form, however, which we want to discuss now.
3 For the experts we remark that Fig. 5.5 without motion in tangential direction represents a
"pathological case". The streamlines end perpendicularly on a circle. This circle represents a
line of critical points but not a limit cycle.
132
5.3 Stability 121
(5.62)
So so that this criterion can be fulfilled, uit) is called unstable. Note that this
definition of stability does not imply that the trajectories vit) which are initially
close to uit) approach it so that the distance between Uj and Vj vanishes eventually.
To take care of this situation we define asymptotic stability. The trajectory uj is
called asymptotic stable iffor all vit) which fulfil the stability criterion the condition
(5.63)
holds in addition. (cf. Fig. 5.6, where the limit cycle is asymptotically stable). So far
we have been following up the motion of the representative point (particle on its
trajectory). If we do not look at the actual time dependence of u but only at the form
of the trajectories then we may define orbital stability. We first give a qualitative
picture and then cast the formulation into a mathematical frame.
This concept is a generalization of the stability definition introduced above.
The stability consideration there started from neighboring points. Stability then
meant that moving points always remain in a certain neighborhood in the course
of time. In the case of orbital stability we do not confine ourselves to single points,
but consider a whole trajectory C. Orbital stability now means, a trajectory in a
sufficiently close neighborhood of C remains in a certain neighborhood in the
course of time. This condition does not necessarily imply that two points which
133
122 5. Necessity
belong to two different trajectories and have been close together originally, do
closely remain together at later times (cf. Fig. 5.15). In mathematical terms this
may be stated as follows: C is orbitally stable if given e > 0 there is 1'/ > 0 such that
if R is a representative point of another trajectory which is within 1'/( C) at time T
then R remains within a distance e for t > T. If no such 1'/ exists, C is unstable.
We further may define asymptotic orbital stability by the condition that C is
orbitally stable, and in addition that the distance between Rand C tends to 0 as t
goes to co. To elucidate the difference between stability and orbital stability con-
sider as an example the differential equations
e = r, (5.64a)
r= O. (5.64b)
According to (5.64b) the orbits are circles and according to (5.64a) the angular
velocity increases with the distance. Thus two points on two neighboring circles
which are initially close are at a later time far apart. Thus there is no stability.
However, the orbits remain close and are orbitally stable.
One of the most fundamental problems now is how to check whether in a given
problem the trajectories are stable or not. There are two main approaches to this
problem, a local criterion and a global criterion.
1) Local Criterion
In this we start with a given trajectory q/t) = u/t) whose asymptotic stability we
want to investigate. To this end we start with the neighboring trajectory
(5.65)
e
which differs from Uj by a small quantity j • Apparently we have asymptotic sta-
bility if e j - 0 for t - 00. Inserting (5.65) into (5.62) yields
(5.66)
Since the e's are assumed to be small, we may expand the right-hand side of (5.66)
into a Taylor series. Its lowest order term cancels against Uj on the Ihs. In the
next order we obtain
(5.67)
134
5.3 Stability 123
In general the solution of (5.67) is still a formidable problem because dF/de is still
a function of time since the u/s were functions of time. A number of cases can be
treated. One is the case in which the u/s are periodic functions. Another very
simple case is provided if the trajectory uj consists ofa singular point so that Uj =
constant. In that case we may put
(5.68)
(5.69)
~ = A~, (5.70)
(5.71)
Inserting (5.71) into (5.70), performing the differentiation with respect to time, and
dividing the equation by eAt, we are left with a set of linear homogeneous algebraic
equations for ~o. The solvability condition requires that the determinant vanishes,
i.e.,
All - A A12 A 1n
A22 - A = o. (5.72)
(5.73)
of order n. If the real parts of all A's are negative, then all possible solutions (5.71)
decay with time so that the singular point is per definition asymptotically stable.
If anyone of the real parts of the A's is positive, we can always find an initial
trajectory which departs from uP) so that the singular point is unstable. We have
so-called marginal stability if the A's have non positive real parts but one or some
of them are equal zero. The following should be mentioned for practical applica-
tions. Fortunately it is not necessary to determine the solutions of (5.73). A number
of criteria have been developed, among them the Hurwitz criterion which allows us
to check directly from the properties of A if Re A < o.
135
124 5. Necessity
Hurwitz Criterion
We quote this criterion without proof. All zeros of the polynomialf(A.) (5.73) with
real coefficients C i lie on the left half of the complex A.-plane (i.e., Re A. < 0) if and
only if the following conditions are fulfilled
C1 C2 Cn
a) - > 0 - > 0 .. · - > 0
Co ' Co ' , Co
C1 Co 0 0 0 0
C3 C2 C1 Co 0 0
Cs C4 C 3 C2 0 0
0 0 0 0 Cn - l Cn- 2
0 0 0 0 0 Cn
Without loss of generality we put the critical point into the origin.
3)
For q "# 0, Vdq) is positive in Q. (5.75)
136
5.3 Stability 125
obeying the differential equations (5.62). Thus VL becomes itself a time dependent
function. Taking the derivative of VL with respect to time we obtain
(5.76)
The great importance of VL rests on the fact that in order to check (5.77) we need
not solve the differential equations but just have to insert the Ljapunov function V L
into (5.77). We are now in a position to define Ljapunov's stability theorem.
A) Stability Theorem
If there exists in some neighborhood Q of the origin a Ljapunov function VL(q),
then the origin is stable. We furthermore have the
137
126 S. Necessity
Here it must be assumed that the solutions of (5.78) exist and are unique and that
Though the reader will rather often meet the concepts of Ljapunov functions in the
literature, there are only a few examples in which the Ljapunov function can be
determined explicitly. Thus, though the theory is very beautiful its practical ap-
plicability has been rather limited so far.
Exercise on 5.3
is a Ljapunov function for the master equation (4.111), where P'(m) is the sta-
tionary solution of (4.111) and P(m) = P(m, t) a time-dependent solution of it.
Hint: Identify P(m) with qj of this chapter (i.e., P ~ q, m _ j) and check that
(E.1) fulfils the axioms of a Ljapunov function. Change q = 0 to q = qO ==
{P'(m)}. To check (5.75) use the property (3.41). To check (5.77), use (4.111) and
the inequality In x ~ 1 - I/x. What does the result mean for P(m, t)? Show that
P'(m) is asymptotically stable.
2) If microreversibility holds, the transition probabilities w(m, m') of (4.111) are
symmetric:
138
5.4 Examples and Exercises on Bifurcation and Stability 127
differential equation. On the other hand it does not shed any light on the importance
of these considerations in other disciplines. To this end we present a few examples.
Many more will follow in later chapters. The great importance of bifurcation rests
in the fact that even a small change of a parameter, in our case the force constant
k, leads to dramatic changes of the system.
Let us first consider a simple model of a laser. The laser is a device in which
photons are produced by the process of stimulated emission. (For more details see
Sect. 8.1). For our example we need to know only a few facts. The temporal
change of photon number n, or, in other words, the photon production rate is
determined by an equation of the form
The gain stems from the so-called stimulated emission. It is proportional to the
number of photons present and to the number of excited atoms, N, (For the experts:
We assume that the ground level, where the laser emission terminates, is kept
empty) so that
G is a gain constant which can be derived from a microscopic theory but this is not
our present com,ern. The loss term comes from the escape of photons through the
endfaces of the laser. The only thing we need to assume is that the loss rate is
proportional to the number of photons present. Therefore, we have
2x = l/to, where to is the lifetime of a photon in the laser. Now an important point
comes in which renders (5.80) nonlinear. The number of excited atoms N decreases
by the emission of photons. Thus if we keep the number of excited atoms without
laser action at a fixed number No by an external pump, the actual number of excited
atoms will be reduced due to the laser process. This reduction, AN, is proportional
to the number of photons present, because all the time the photons force the atoms
to return to their groundstates. Thus the number of excited atoms has the form
Inserting (5.81), (5.82) and (5.83) into (5.80) gives us the basic laser equation in our
simplified model
(5.84)
k = 2x - GNo ~ o. (5.85)
If there is only a small number of excited atoms, No, due to the pump, k is positive,
139
128 5. Necessity
whereas for sufficiently high No, k can become negative. The change of sign occurs
at
which is the laser threshold condition. Bifurcation theory now tells us that for
k > 0 there is no laser light emission whereas for k < 0 the laser emits laser
photons. The laser functions in a completely different way when operating below
or above threshold. In later chapters we will discuss this point in much more detail
and we refer the reader who is interested in a refined theory to these chapters.
Exactly the same (5.84) or (5.80) with the terms (5.81), (5.82) and (5.83) can be
found in a completely different field, e.g., chemistry. Consider the autocatalytic
reaction between two kinds of molecules, A, B, with concentrations nand N,
respectively:
A + B -+ 2A production
A -+ C decay
The molecules A are created in a process in which the molecules themselves par-
ticipate so that their production rate is proportional to n (compare (5.81». Further-
more the production rate is proportional to N. If the supply of B-molecules is not
infinitely fast, N will decrease again by an amount proportional to the number n of
A-molecules present.
The same equations apply to certain problems of ecology and population
dynamics. If n is the number of animals of a certain kind, N is a measure for the
food supply available which is steadily renewed but only at a certain finite pace.
Many more examples will be given in Chapters 9 and 10.
Exercise
where
and no = nCO) is the initial value. Discuss the temporal behavior of this function
and show that it approaches the stationary state n = 0 or n = Ik/ k II irrespective
of the initial value no, but depending on the signs of k, k \. Discuss this dependence!
140
5.4 Examples and Exercises on Bifurcation and Stability 129
and n2 are produced. In analogy to (5.80) [with (5.81,2,3)] the rate equations read
(5.90)
(5.91)
(5.92)
Exercise
What happens if
(5.93)
Discuss the stability of (5.88) and (5.89). Does there exist a potential?
Discuss the critical point nl = n 2 = o.
Are there further critical points?
Eqs. (5.88), (5.89) and (5.90) have a simple but important interpretation in ecology.
Let n 1 , n 2 be the numbers of two kinds of species which live on the same food sup-
ply No. Under the condition (5.92) only one species survives while the other dies
out, because the species with the greater growth rate, G 1 , eats the food much more
quickly than the other species and finally eats all the food. It should be mentioned
that, as in (5.83) the food supply is not given at an initial time but is kept at a certain
rate. Coexistence of species becomes possible, however, if the food supply is at least
partly different for different species. (Compare Sect. 10.1).
Exercise
(5.94)
(5.95)
141
130 5. Necessity
(5.96)
We identify the prey fishes with the index 1. If there are no predator fishes, the prey
fishes will multiply according to the law
(5.97)
The prey fishes suffer, however, losses by being eaten by the predators. The loss
rate is proportional to the number of preys and predators
(5.98)
(5.99)
Because they live on the prey, predator multiplication rate is proportional to their
own number and to that of the prey fishes. Since predators may suffer losses by
death, the loss term is proportional to the number of predator fishes present
(5.100)
(5.101)
Exercise
(5.102)
(Hint: Put
X2 '.
n l = 2anu n2 = lXI,
-n2,• t 1 ')
= -t (5.103)
P IX 1X1
142
5.4 Examples and Exercises on Bifurcation and Stability 131
2) Determine the stationary state iiI = liz = O. Prove the following conservation
law
L= ~ (nj -
1,2 In nj) = const, ( al
az
=
=a
1) (5.104)
J
or
(5.105)
and use (5.102). From Fig. 5.17 it follows that the motion of n l , nz is periodic.
cf. Fig. 5.18. Why?
~ .
....""""
2
2 ,,
,,
,
--_ .. . - '
,- " Fig. 5.17. Two typical trajectories in the
nl-n2 phase plane of the Lotka-Volterra
model (after GoeI, N. S., S. C. Maitra,
1.5 2 2.5 :3 E. W. MontroIl: Rev. Mod. Phys. 43,
n~ -+ 231 (1971)) for fixed parameters
40.---~---r---'----r---'----r---'
30
20
10
Fig. 5.18. Time variation of
the two populations n" n2
corresponding to a trajectory
~--~--~--~--~--~--~--~+t of Fig. 5.17
2 4 6 8 10 12 14
Hint: Use Section 5.1. Are there singular points? Are the trajectories stable
or asymptotically stable?
Answer: The trajectories are stable, but not asymptotically stable. Why?
143
132 5. Necessity
I)
41 = -Aql'
42 = -M2' (S.107)
VL = q~ + q~. (S.108)
2)
41 = Aql'
(S.109)
42 = -M2'
VL = qi + q~. (S.110)
4 = (a + bi)q,
(S.111)
4* = (a - bi)q*.
with
and
a < O. (S.I13)
v= qq*. (S.1l4)
Show in the above examples that the functions (S.108), (S.1IO) and (S.114) are
Ljapunov functions. Compare the results obtained with those of Section S.2
(and identify the above results with the case of stable node, unstable saddle
point, and stable focus).
Show that the potential occurring in anharmonic oscillators has the properties
of a Ljapunov function (within a certain region, Q).
This equation which has played a fundamental role in the discussion of the perform-
ance of radio tubes has the following form:
144
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 133
with
e > O. (5.116)
q =p - e(q3/3 - q),
(5.117)
p = -q,
that the origin is the only critical point which is a source. For which e's is it an
unstable focus (node)?
Show that (5.117) allows for (at least) one limit cycle.
Hint: Use the discussion following the Poincare-Bendixson theorem on p. 119.
Draw a large enough circle around the origin, q = 0, p = 0, and show that all
trajectories enter its interior. To this end, consider the rhs of (5.117) as components
of a vector giving the local direction of the streamline passing through q, p. Now
form the scalar product of that vector and the vector q = (q, p) pointing from the
origin to that point. The sign of this scalar product tells you into which direction
the streamlines point.
A) One-Dimensional Case
We consider the potential V(q) and assume that it can be expanded into a Taylor
senes:
(5.118)
145
134 S. Necessity
provided that the expansion is taken at q = O. Since the form of the potential curve
does not change if we shift the curve by a constant amount, we may always put
c(O) = O. (5.123)
We now assume that we are dealing with a point of equilibrium (which may be
stable, unstable, or metastable)
dV =0 (5.124)
dq .
C(l) = O. (5.125)
Before going on let us make some simple but fundamental remarks on small-
ness. In what follows in this section we always assume that we are dealing with
dimensionless quantities. We now compare the smallness of different powers of q.
Choosing q = 0.1, q2 gives 0.01, i.e., q2 is only 10% of q. Choosing as a further
example q = 0.01, q2 = 0.0001, i.e., only I % of q. The same is evidently true for
consecutive powers, say qft and qft + 1. When we go from one power to the next,
choosing q sufficiently small allows us to neglect qn+l compared to qn. Therefore
we can confine ourselves in the following to the leading terms of the expansion
(5.118). The potential shows a local minimum provided
!2 d2Vl' -= c > 0
.12
(2) (5.126)
uq =0
(5.127)
As we will substantiate later on by many explicit examples, J.! may change its sign
when certain parameters of the system under consideration are changed. This turns
the stable point q = 0 into an unstable point for J.! < 0 or into a point of neutral
stability for Jl = O. In the neighborhood of such a point the behavior of V(q) is
determined by the next nonvanishing power of q. We will call a point where J.! = 0
an instability point. We first assume
146
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 135
We shall show later on in practical cases, V(q) may be disturbed either by external
causes, which in mechanical engineering may be loads, or internally by imperfec-
tions (compare the examples of Chapter 8). Let us assume that these perturbations
are small. Which of them will change the character of (5.128) the most? Very close
to q = 0 higher powers of q e.g., q4, are much smaller than (5.128), so that such a
term presents an unimportant change of (5.128). On the other hand, imperfections
or other perturbations may lead to lower powers of q than cubic so that these can
become dangerous in perturbing (5.128). Here we mean by "dangerous" that the
state of the system is changed appreciably.
The most general case would be to include all lower powers leading to
(5.129)
Adding all perturbations which change the original singularity (5.128) in a non-
trivial way are called according to Thorn "unfoldings". In order to classify all pos-
sible unfoldings of (5.128) we must do away superfluous constants. First, by an
appropriate choice of the scale of the q-axis we can choose the coefficient C(3) in
(5.128) equal to 1. Furthermore we may shift the origin of the q axis by the trans-
formation
q = q' +D (5.130)
to do away the quadratic term in (5.129). Finally we may shift the zero point of the
potential so that the constant term in (5.129) vanishes. We are thus left with the
"normal" form V(q)
This form depends on a single free parameter, u. The potentials for u = 0 and
u < 0 are exhibited in Fig. 5.19. For u -> 0 the maximum and minimum coincide
to a single turning point.
V(q }
147
136 5. Necessity
(5.132)
q4 uq2
V(q) = "4 + T + vq, (5.133)
where we have already shifted the origin of the (q - V)-coordinate system ap-
propriately. The factors 1/4 and 1/2 in the first or second term in (5.133) are chosen
in such a manner that the derivative of V acquires a simple form
dV
dq = q3 + uq + v. (5.134)
- - - - -
o II Fig. 5.21. In the u-v plane, the solid curve separates the
- ~-t-----
region with one potential minimum (right region) from the
region with two potential minima. The solid line represents
in the sense of Thom the catastrophic set co nsisting of
bifurcation points. The dotted line indicates where the two
minima have the same value. This line represents the cata-
strophic set consisting of conflict points (after Thom)
148
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 137
3) If C(4) = 0 but the next coefficient does not vanish we find as potential of the
critical point
(5.135)
qS uq3 vq2
V=-+-+-+wq (5.136)
5 3 2 '
oV
- = q4 + uq2 + vq + w = 0 (5.137)
oq
allowing for zero, two or four extrema implying zero, one or two minima. If we
change u, or v, or w or some of them simultaneously it may happen that the number
of minima changes (bifurcation points) or that their depths become equal (con-
flict points).
It turns out that such changes happen in general at certain surfaces in u-v-w
space and that these surfaces have very strange shapes. (Compare Fig. 5.22). Then,
as last example we mention the potential
(5.138)
B) Two-Dimensional Case
Expanding the potential V into a Taylor series yields
149
138 5. Necessity
(5.140)
c(O) = o. (5.141)
(5.142)
and
(5.143)
(5.144)
ql = A l1 u 1 + A 12 U 2,
(5.145)
q2 = A 21 u 1 + A 22 U 2
Vir acquires the form
(5.146)
If we apply the transformation (5.145) not only to the truncated form of V (5.144)
but to the total form (5.139) (however with c(O) = c~l) = C~I) = 0), we obtain a
new potential in the form
(5.147)
This form allows us, again in a simple way, to discuss instabilities. Those occur if
by a change of external parameters, 111 or 112 or both become = O.
For further discussion we first consider 111 = 0 and 112 > 0, or, in other words,
the system loses its stability in one coordinate. In what follows we shall denote the
coordinate in the "unstable direction" by x, in the "stable direction" by y. Thus
150
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 139
(5.148)
h« f1.2 (5.151)
151
140 5. Necessity
in y direction. For further discussion we exhibit the leading terms of (5.148) in the
form
(5.152)
where we have added and subtracted the quadratic complement and where we have
used the abbreviation
(5.154)
(5.155)
Provided the higher order terms are small enough we see that we have now found
a complete decomposition of the potential V into a term which depends only on x
and a second term which depends only on y. We now investigate in which way the
so-called higher order terms are affected by the transformation (5.154). Consider
for example the next term going with y3. It reads
(5.156)
Since yeo) ~ g(x), and g(x) is a small quantity, the higher order terms in lowest
approximation in y contribute
As (5.156) contains thus g3(X), and g is a small quantity, we find that the higher
order terms give rise to a correction to the x-dependent part of the potential V of
higher order. Of course, if the leading terms of the x-dependent part of the potential
vanish, this term may be important. However, we can evidently devise an iteration
procedure by which we repeat the procedure (5.148) to (5.155) with the higher
order terms, each term leading to a correction term of decreasing importance.
Such iteration procedure may be tedious, but we have convinced ourselves, at least
in principle, that this procedure allows us to decompose V into an x and into a y
(or y-) dependent part provided that we neglect terms of higher order within a
well-defined procedure.
We now treat the problem quite generally without resorting to that iteration
procedure. In the potential V(x, y) we put
y = yeo) + y. (5.157)
152
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 141
aVI
aji y=O
=0 (5.159)
holds. This may be considered as an equation for y(O) which is of the form
(5.161)
(5.162)
where the linear term is lacking on account of (5.159). By the above more explicit
analysis (compare (5.155)) we have seen that in the neighborhood of x = 0 the
potential retains its stability in y-direction. In other words, using the decomposition
we may be sure that (5.163) remains positive. Thus the only instability we have to
discuss is that inherent in VI(x) so that we are back with the one-dimensional case
discussed above.
We now come to the two-dimensional case, III = 0,112 = O. The first, in general
nonvanishing, terms of the potential are thus given by
(5.164)
Provided that one or several of the coefficients C(3) are unequal zero, we may stop
with the discussion of the form (5.164). Quite similar to the one-dimensional case
we may assume that certain perturbations may become dangerous to the form
(5.164). In general these are terms of a smaller power than 3. Thus the most general
unfolding of (5.164) would be given by adding to (5.164) terms of the form
(5.165)
153
142 5. Necessity
x = AllXI + A 12 x 2 + B I ,
(5.166)
y = A 2l X I + A 22 X 2 + B 2 •
We may cast (5.164) into simple normal forms quite similar to ellipses etc. This
can be achieved by a proper choice of A's. By proper choice of E's we can further
cast the quadratic or bilinear terms of (5.165) into a simpler form. If we take the
most general form (5.164) and (5.165) including the constant term Vo, there are 10
constants. On the other hand the transformation (5.166) introduces 6 constants
which together with c(O) yields 7 constants. Taking the total number of possible
constants minus the seven, we still need three constants which can be attached to
certain coefficients (5.165). With these considerations and after some analysis we
obtain three basic forms which we denote according to Thom, in the following
manner:
Hyperbolic umbilic
(5.167)
Elliptic umbilic
(5.168)
Parabolic umbilic
For the example of the parabolic umbilic a section through the potential curve
is drawn in Fig. 5.25 for different values of the parameters u, v, t. We note that for
t < 0 the parabolic umbilic goes over to the elliptic umbilic whereas for t > 0 we
obtain the hyperbolic umbilic. Again for each set of u, v, W, t, the potential curve
may exhibit different minima, one of which is the deepest (or several are simul-
taneously the deepest) representing a state of the system. Since one deepest mini-
mum may be replaced by another one by changing the parameters u, v, W, t, the
u,v,w,t-space is separated by different hyper-surfaces and thus decomposed into
subspaces. A note should be added about the occurrence of quartic terms in (5.169)
which seems to contradict our general considerations on page 134 where we stated
that only powers smaller than that of the original singularity can be important.
The reason is that the unfolding in (5.169) includes terms tx~ which have the same
power as the original terms, e.g., xi or xix 2 • If we now seek the minimum of V,
(5.170)
154
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 143
eal(,)
Fig. 5.25. The universal unfolding of the hyperbolic urnbilic (at center) surrounded by the local
potentials in regions I, 2 and 3 and at the cusp C (after Thorn)
At least one solution of this quadratic equation (5.171), however, tends to infinity
for t -+ O. This contradicts our original assumption that we are only considering
the immediate vicinity of x I = X2 = O. The zeros of (5.171), however, are bounded
to that minimum if we take quartic terms into account according to (5.169).
(5.173)
155
144 5. Necessity
and
(5.174)
(5.175)
(5.176)
(5.178)
(5.179)
has only real values /1j. Provided that all /1j > 0, the state q = 0 is stable. We now
assume that by a change of external parameters we reach a state where a certain
set of the /1'S vanishes. We number them in such a manner that they are the first,
j = 1, ... , k, so that
Thus we have now two groups of coordinates, namely, those associated with indices
1 to k which are coordinates in which the potential shows a critical behavior, while
for k + 1, ... , n the coordinates are those for which the potential remains un-
critical. To simplify the notation we denote the coordinates so as to distinguish
between these two sets by putting
(5.181)
Uk + 1> ••• , Un = Y1' •.. ,Yn-k'
156
5.5 Classification of Static Instabilities, or an Approach to Thorn's Theory 145
Our first goal is to get rid of the terms linear in Yj which can be achieved in the
following way. We introduce new coordinates Ys by
The y(O)'s are determined by the requirement that for Y s = y~O) the potential
acquires a minimum
-OVI =0 (5.184)
oYs y(O) •
where hjj' contains f.1j > 0 stemming from the first term in (5.182) and further
correction terms which depend on x I' . • . ,Xk' This can be, of course, derived in
still much more detail; but a glance at the two-dimensional case on pages 137-141
teaches us how the whole procedure works. (5.185) contains higher-order terms,
i.e., higher than second order in the ji's. Confining ourselves to the stable region
in the y-direction, we see that V is decomposed into a critical Vdepending only on
x I' . . . , Xb and a second noncritical expression depending on Y and x. Since all
the critical behavior is contained in the first term depending on k coordinates
x I, ••. , Xk' we have thus reduced the problem of instability to one which is of a
dimension k, usually much smaller than n. This is the fundamental essence of our
present discussion.
We conclude this chapter with a few general comments: If we treat, as every-
where in Chapter 5, completely deterministic processes, a system cannot jump
from a minimum to another, deeper, minimum if a potential barrier lies between.
We will come back to this question later on in Section 7.3. The usefulness of the
above considerations lies mainly in the discussion of the impact of parameter
changes on bifurcation. As will become evident later, the eigenvalues f.1j in (5.182)
are identical with the (imaginary) frequencies occurring in linearized equations for
the u's. Since (cj]l) is a real, symmetric matrix, according to linear algebra the
f.1/s are real. Therefore no oscillations occur and the critical modes are only soft
modes.
157
146 5. Necessity
Our presentation of the basic ideas of catastrophe theory has been addressed to
students and scientists of the natural sciences. In particular, we wanted to present
a classification scheme of when systems change their states. To this end we as-
sumed that the function V may be expanded into a Taylor series. For a mathema-
tician, other aspects may be more important, for instance how to prove his
theorems with minimal assumptions about V. Here, indeed, catastrophe theory
does not require the assumption that V can be expanded into a Taylor series, but
only that V can be differentiated infmitely often. Furthermore, for our presenta-
tion it was sufficient to neglect the higher-order terms because of their smallness.
In catastrophe theory it is shown that they can even be transformed away.
158
6. Chance and Necessity
Reality Needs Both
m·v = F. (6.1)
where tj is the moment a kick occurs. The effect of this kick on the change of
velocity can be determined as follows: We insert (6.2) into (6.1):
(6.3)
(6.4)
which describes that at time tj the velocity v is suddenly increased by the amount
cp/m. The total force exerted by the kicker in the course of time is obtained by
148 6. Chance and Necessity
(6.7)
in which the sequence of plus and minus signs is a random sequence in the sense
discussed earlier in this book with respect to tossing coins. Taking into account
both the continuously acting friction force due to the grass and the random kicks
of the football player, the total equation of motion of the football reads
mv = -}IV + P(t)
where
(t = }11m (6.9a)
and
1
-If' = - L
qJ
F(t) == m m
~(t - t)(± l)j. (6.9b)
How to perform the statistical average requires some thought. In one experiment,
say a football game, the particle (ball) is moved at certain times tj in a forward or
backward direction so that during this experiment the particle follows a definite
path. Compare Fig. 6.1 which shows the change of velocity during the time due to
the impulses (abrupt changes) and due to friction force (continuous decreases
inbetween the impulses). Now, in a second experiment, the times at which the
particle is moved are different. The sequence of directions may be different so that
another path arises (compare Fig. 6.2). Because the sequences of times and direc-
tions are random events, we cannot predict the single path but only averages. These
averages over the different time sequences and directions of impulses will be per-
formed below for several examples. We now imagine that we average F over the
random sequence of plus and minus signs. Since they occur with equal probability,
we immediately find
(F(t) = O. (6.10)
160
6.1 Langevin Equations: An Example 149
Fig. 6.1 . The velocity v changes due to pushes Fig. 6.2. Same as in Fig. 6.1 but with a
(random force) and friction force different realization
We further form the product of Fat a time t with F at another time t' and take the
average over the times of the pushes and their directions. We leave the evaluation
to the reader as an exercise (see end of this paragraph). Adopting a Poisson process
(compare Sect. 2.12) we find the correlation function
2
<F(t)F(t') = ~ 8(t - t') = C8(t - t'). (6.11)
m to
Equation (6.8) together with (6.10) and (6.1 1) describes a physical process which
is well known under the name of Brownian movement (Chapt. 4). Here a large
particle immersed in a liquid is pushed around by the random action of particles
of the liquid due to their thermal motion. The theory of Brownian movement
plays a fundamental role not only in mechanics but in many other parts of physics
and other disciplines, as we shall demonstrate later in this book (cf. Chapt. 8).
The differential equation (6.8) can be solved immediately by the method of varia-
tion of the constant. The solution is given by
(6.12)
In the following we neglect "switching on" effects. Therefore we neglect the last
term in (6.12), which decays rapidly.
Let us now calculate the mean kinetic energy defined by
(6.13)
Again the brackets denote averaging over all the pushes. Inserting (6.12) into (6.13)
we obtain
(6.14)
Since the averaging process over the pushes has nothing to do with the integration,
161
150 6. Chance and Necessity
we may exchange the sequence of averaging and integration and perform the
average first. Exploiting (6.11), we find for (6.14) immediately
m
- (v 2 ) = - mJt Jt d,dr' e- 2at + a «+<')b(, - ,/)C. (6.15)
2 2 0 0
Due to the b-function the double integration reduces to a single integral which can
be immediately evaluated and yields
(6.16)
Since we want to consider the stationary state we consider large times, t, so that
we can drop the exponential function in (6.16) leaving us with the result
(6.17)
Now, a fundamental point due to Einstein comes in. We may assume that the
particle immersed in the liquid is in thermal equilibrium with its surroundings.
Since the one-dimensional motion of the particle has one degree of freedom,
according to the equipartition theorem of thermodynamics it must have the mean
energy 0/2) kBT, where kB is Boltzmann's constant and Tthe absolute temperature.
A comparison of the resulting equation
m 2
- (v ) = "ikBT (6.18)
2
(6.19)
Recall that the constant C appears, in the correlation function of the fluctuating
forces (6.11). Because C contains the damping constant y, the correlation function
is intrinsically connected with the damping or, in other words, with the dissipation
of the system. (6.19) represents one of the simplest examples of a dissipation",
fluctuation theorem: The size of the fluctuation (~ C) is determined by the size
of dissipation (~ y). The essential proportionality factor is the absolute tempera-
ture. The following section will derive (6.8), not merely by plausibility arguments
but from first principles.
As we have discussed above we cannot predict a single path but only averages.
One of the most important averages is the two-time correlation function
which is a measure how fast the velocity loses its memory or, in other words, if we
have fixed the velocity at a certain time f', how long does it take the velocity to
162
6.1 Langevin Equations: An Example 151
differ appreciably from its original value? To evaluate (6.20) we insert (6.12)
(6.21)
(6.22)
(6.23)
The integrals can be immediately evaluated. The result reads, for t > t',
(6.24)
If we consider a stationary process, we may assume t and t' large but t - t' small.
This leaves us with the final result
Thus the time T after which the velocity loses its memory is T = l/ex. The case
ex = 0 leads evidently to a divergence in (6.25). The fluctuations become very large
or, in other words, we deal with critical fluctuations. Extending (6.20) we may
define higher-order correlation functions containing several times by
(6.26)
When we try to evaluate these correlation functions in the same manner as before,
we have to insert (6.21) into (6.26). Evidently we must know the correlation func-
tions
(6.27)
The evaluation of (6.27) requires additional assumptions about the F's. In many
practical cases the F's may be assumed to be Gaussian distributed (Sect. 4.4). In that
case one finds
(6.28)
<F('n)F(rn- t )·· ·F('t» = Lp <F(';.)F(,;»·· . <F(';'n_,)F(T;.J> for n even.
where Lp runs over all permutations (AI, ... , An) of (I, ... , n).
163
152 6. Chance and Necessity
Exercise on 6.1
1) Evaluate the Ihs of (6.11) by assuming for the t/s a Poisson process.
mt(tt) = kt(tt),
mitt, t z ) = kzCtt, t z ) + kt(tt)kt(t z),
m 3(tt, t z , t 3) = k3(tt' t z , t 3) + 3{k l (tl)kz{t z , t 3)}s + k t (tt)k t (t z )k t (t3),
mitt, t z , t 3, t 4) = kitt, t z , t 3, t 4) + 3{kz(tt, t Z )kz(t 3, t 4)}s
+ 4{k t (tt)k 3(t Z' t 3, t 4)}s + 6{k t (tt)k t (tz)kit 3, t 4)}s
+ kt(tt)kt(tz)kt(t3)kt(t4)' (E.1)
In the preceding chapter we have introduced the random force F and its properties
as well as the friction force by plausibility arguments. We now want to show how
both quantities can be derived consistently by a detailed physical model. We make
the derivation in a manner which shows how the whole procedure can be extended
to more general cases, not necessarily restricted to applications in physics. Instead
of a free particle we consider a harmonic oscillator, i.e., a point mass fixed to a
spring. The elongation of the point mass from its equilibrium position is called q.
Denoting Hooke's constant by k, the equation of the harmonic oscillator reads
(v = q)
m
k
= w o,
z
(6.30)
164
6.2 Reservoirs and Random Forces 153
where Wo is the frequency of the oscillator. The equation of motion (6.29) then
reads
(6.31)
Equation (6.31) can be replaced by a set of two other equations if we put, at first,
q =p (6.32)
(compare also Sect. 5.2). It is now our goal to transform the pair of equations
(6.32) and (6.33) into a pair of equations which are complex conjugate to each
other. To this end we introduce the new variable bet) and its conjugate complex
b*(t), which shall be connected with p and q by the relations
(6.34)
and
1 (_ ,-
./2 (" woq - ip/" wo) = b*. (6.35)
Multiplying (6.32) by ./wo and (6.33) by il./w o and adding the resulting equations
we find after an elementary step
Similarly the subtraction of (6.33) from (6.32) yields an equation which is just the
conjugate complex of (6.36).
After these preliminary steps we return to our original task, namely, to set up a
physically realistic model which eventually leads us to dissipation and fluctuation.
The reason why we hesitate to introduce the damping force - yv from the very
beginning is the following. All fundamental (microscopic) equations for the motion
of particles are time reversal invariant, i.e. the motion is completely reversible.
Originally there is no place in these equations for a friction force which violates
time reversal invariance. Therefore, we want to start from equations which are the
usual mechanical equations having time reversal invariance. As we have mentioned
above in connection with Brownian movement the large particle interacts with
(many) other particles in the liquid. These particles act as a "reservoir" or "heat-
bath"; they maintain the mean kinetic energy of the large particle at 1/2 kBT (per
one degree of freedom). In our model we want to mimic the effect of the "small"
particles by a set of very many harmonic oscillators, acting on the "large" particle,
which we treat as an oscillator by (6.36) and its conjugate complex. We assume that
the reservoir oscillators have different frequencies w out of a range Llw (also called
bandwidth). In analogy to the description (6.34) and (6.35), we use complex
165
154 6. Chance and Necessity
(6.37)
The coefficients gw describe the strength of the coupling between the other oscil-
lators and the one under consideration. Now the "big" oscillator reacts on all the
other oscillators which is described by
(6.38)
(Readers interested in how to obtain (6.37) and (6.38) in the usual framework of
mechanics are referred to the exercises). The solution of (6.38) consists of two
parts, namely, the solution of the homogeneous equation (where ibg w = 0) and
a particular solution of the inhomogeneous equation. One readily verifies that the
solution reads
(6.39)
where Bw(O) is the initial value, at t = 0, of the oscillator amplitude Bro. Inserting
(6.39) into (6.37) we find an equation for b
(6.40)
The only reminiscence from the B's stems from the last term in (6.40). For further
discussion, we eliminate the term -iOJob(t) by introducing the new variable 6,
(6.41)
w= OJ - OJo (6.42)
(6.43)
To illustrate the significance of this equation let us identify, for the moment,
b with the velocity v which occurred in (6.8). The integral contains b linearly which
suggests that a connection with a damping term - yb might exist. Similarly the last
term in (6.43) is a given function of time which we want to identify with a random
force F. How do we get from (6.43) to the form of (6.8)? Apparently the friction
166
6.2 Reservoirs and Random Forces 155
force in (6.43) not only depends on time, t, but also on previous times r. Under
which circumstances does this memory vanish? To this end we consider a transition
from discrete values w to continuously varying values, i.e., we replace the sum
over w by an integral
(6.44)
For time differences t - r not too short, the exponential function oscillates rapidly
for ill =1= 0 so that only contributions of ill :::::: 0 to the integral are important. Now
assume that the coupling coefficients g vary in the neighborhood of Wo (i.e., ill = 0)
only slightly. Because only small values of ill are important, we may extend the
boundaries of the integral to infinity. This allows us to evaluate the integral (6.44)
We now insert (6.45) into the integral over the time occurring in (6.43). On account
of the (i-function we may put b(-r) = bet) and put it in front of the integral. Further-
more, we observe that the (i-function contributes to the integral only with the factor
1/2 because the integration runs only to r = t and not further. Since the (i-function
is intrinsically a symmetric function, only 1/2 of the (i-function is covered. We thus
obtain
(6.46)
where", has been introduced as an abbreviation. The last term in (6.43) will now be
abbreviated by F
By this intermediate steps we may rewrite our original (6.43) in the form
The result looks very fine because (6.49) has exactly the form we have been
searching. There are, however, a few points which need a careful discussion.
In the model of the foregoing Section 6.1, we assumed that the force F is random.
How does the randomness come into our present model? A look at (6.47) reveals
that the only quantities which can introduce any randomness are the initial values
of Bro. Thus we adopt the following attitude (which closely follows information
theory). The precise initial values of the reservoir oscillators are not known except
167
156 6. Chance and Necessity
in a statistical sense; that means we know, for instance, their distribution functions.
Therefore we will use only some statistical properties of the reservoirs at the initial
time. First we may assume that the average over the amplitudes Bro vanishes be-
cause otherwise the nonvanishing part could always be subtracted as a deter-
ministic force. We furthermore assume that the Bro's are Gaussian distributed,
which can be motivated in different manners. If we consider physics, we may
assume that the reservoir oscillators are all kept at thermal equilibrium. Then the
probability distribution is given by the Boltzmann distribution (3.71) Pro =
.¥roe - Ew/kBT where Ero is the energy of oscillator co and .¥ ro == z;;; 1 is the normaliza-
tion factor. We assume, as usual for harmonic oscillators, that the energy is propor-
tional to p2 + q2 (with appropriate factors), or in our formalism, proportional
to BwB!, i.e., Ero = croB!Bro. Then we immediately find this announced Gaussian
distribution for B,B*,J(B, B*) = .¥ro exp( -croB!Bro/kBT). Unfortunately there is
no unique relation between frequency and energy, which would allow us to deter-
mine CW • (This gap can be filled only by quantum theory).
Let us now investigate if the force (6.47) has the properties (6.10), (6.11) we
expect according to the preceding section. Because we constructed the forces that
time in a completely different manner out of individual impulses, it is by no means
obvious that the forces (6.47) fulfil relations of the form (6. 11). But let us check it.
We form
<P*(t)F(t'». (6.50)
(Note that the forces are now complex quantities). Inserting (6.47) into (6.50) and
taking the average yields
(6.51)
(6.53)
The sum occurring in (6.53) is strongly reminiscent of that of the Ihs of (6.44) with
the only difference that now an additional factor, namely, the average <... ) occurs.
Evaluating (6.53) by exactly the same considerations which have led us from (6.44)
to (6.45), we now obtain
(6.54)
Using this final result (6.54) instead of (6.51) we obtain the desired correlation
function as
168
6.2 Reservoirs and Random Forces 157
(6.57)
(6.58)
(6.59)
169
158 6. Chance and Necessity
(6.60)
T =
x\TI + ... + x n Tn.
(6.61)
Xl + ... + X.
It is most important that the average temperature depends on the strength of the
dissipation (~ Xj) caused by the individual heatbath. Later on we will Ie am about
problems in which coupling of a system to reservoirs at different temperatures
indeed occurs. According to (6.59) the fluctuations should vanish when the tem-
perature approaches absolute zero. However, from quantum theory it is known
that the quantum fluctuations are important, so that a satisfactory derivation of the
fluctuating forces must include quantum theory.
~xercises on 6.2
b· = -iiJH/ob*, B = -ioH/iJB*
(and their complex conjugates).
2) Perform the integration in (6.23) for the two different cases: t > t' and t < t'.
For large values t and t', one may confirm (6.25).
170
6.3 The Fokker-Planck Equation 159
Indeed we have seen earlier in Section 2.5, Exercise that an integral over the
function b(q - qo) vanishes, if the integration interval does not contain qo, and that
it yields I if that interval contains a surrounding of qo :
J qo+£
qo-' b(q - qo) dq = I (6.64)
= 0 otherwise.
It is now our goal to derive an equation for this probability distribution P.
To this end we differentiate P with respect to time. Since on the rhs the time de-
pendence is inherent in q(/), we differentiate first the b-function with respect to q(t),
and then, using the chain rule, we multiply by 4
. d
P(q, t) = dq(t) b(q - q(t»4(t). (6.65)
d
- dq b(q - q(t»4(t). (6.66)
Now we make use of the equation of motion (6.62) replacing 4(t) by K. This yields
171
160 6. Chance and Necessity
· d
P(q, t) = - dq (K(q)P). (6.67)
where q = (ql' ... , qn). Instead of (6.68) we use the vector equation
4= K(q). (6.69)
(6.70)
where the last identity just serves a definition of the b-function of a vector. We again
take the time derivative of P. Because the time t occurs in each factor, we obtain a
sum of derivatives
· d d d
P(q, t) = - -p·(Mt) - -P·q2··· - dqnPqn (6.71)
dql dq2
· d d
P(q, t) = - -PK 1 (q) ... - -PKn(q). (6.72)
dql dqn
(6.73)
172
6.3 The Fokker-Planck Equation 161
(6.75)
and so on. Now we take the average over all these paths, introducing the function
If the probability of the occurrence of a path i is Pi' this probability distribution can
be written in the form
/dq gives us the probability of finding the particle at position q in the interval dq at
time t. Of course, it would be a very tedious task to evaluate (6.77) which would
require that we introduce a probability distribution of the pushes during the total
course of time. This can be avoided, however, by deriving directly a differential
equation for f. To this end we investigate the change of/in a time interval At
We put
q(t + At) = q(t) + Aq(t) (6.81)
and expand the ~-function with respect to powers of Aq. We now have in mind that
the motion of q is not determined by the deterministic equation (6.62) but rather by
the Langevin equation of Section 6.1. As we shall see a little later, this new situa-
tion requires that we expand up to powers quadratic in Aq. This expansion thus
yields
173
162 6. Chance and Necessity
we find Llq by integration over the time interval At. In this integration we assume
that q has changed very little but that many pushes have already occurred. We
thus obtain by integrating (6.83)
Jt+dt
t £j(t') dt' = q(t + At) - q(t) == Aq
(6.85)
d
dq {(o(q - q(t»( -yq(t)At» + ([)(q - q(t»)(AF)}. (6.86)
d
-yAt dq {(o(q - q(t»q)}. (6.87)
which using the same arguments as just now can be split into
(6.89)
When inserting Aq in the second part using (6.84) we find terms containing At 2,
terms containing AtAF, and (AF)2. We will show that <CAFf) goes with At. Since
the average of LlFvanishes, «AF)2) is the only contribution to (6.89) which is linear
174
6.3 The Fokker-Planck Equation 163
in L1t. We evaluate
<L1F(t)t1F(t) = J
t+Llt rt+Llt
t J t dt' dt" <F(t')F(t"). (6.90)
QL1t. (6.92)
d2
dq2 <<5(q - q(t)))QL1t. (6.93)
We now divide the original equation (6.82) by At and find with the results of (6.87)
and (6.93)
df d d2
dt = dq (yqf) + t Q dqzf (6.94)
after the limit At ---> 0 has been taken. This equation is the so-called Fokker-Planck
equation which describes the change of the probability distribution of a particle
during the course of time (cf. Fig. 6.4). K = - yq is called drift-coefficient, while
Q is known as diffusion coefficient. Exactly the same method can be applied to the
general case of many variables and to arbitrary forces Ki(q), i.e., not necessarily
simple friction forces. If we assume the corresponding Langevin equation in the
((q,t)
175
164 6. Chance and Necessity
form
(6.95)
(6.96)
(6.98)
f q=q(t)+e
q=q(t) -e
d
h(q) dq o(q - q(t))K(q(t)) dq (6.99)
which is obtained from the rhs of (6.66) by replacing q by means of (6.62) and by
multiplication with an arbitrary function h(q). Note that this procedure must
always be applied if a c5-function appears in a differential equation in a form like
(6.65). Partial integration of (6.99) leads to
- fq(t)+e
where the o-function can now be evaluated. On the other hand we end up with the
same result (6.100) if we start right away from
Exercises on 6.3
1) In classical mechanics the coordinates qit) and momenta Pj(t) of particles obey
176
6.4 Some Properties and Stationary Solutions of the Fokker-I1lanck Equation 165
where the Hamiltonian function H depends on all q/s and p/s: H = H(q, p).
We define a distribution function f(q, p; t) by f = b(q - q(t»b(p - p(t».
Show that f obeys the so-called Liouville equation
(E.l)
2) Functions f satisfying the Liouville equation E.l with J' = 0 are called con-
stants of motion.
Show a) g = H(q, p) is a constant of motion;
b) if hl(q, p) and h2(q, p) are constants of motion then also hi + h2 and
hi' h2 are constants of motion;
c) if hi, ... , hi are such constants, then any function G(h l , ... ,hi) is
also a constant of motion.
Hint : a) insert g into (E.1)
b) use the product rule of differentiation
c) use the chain rule.
4) Verify that the information entropy (3.42) satisfies (E. 1) provided the following
identifications are made:
Replace I in (3.42) now by an integral S ... dnpdnq. Why does also the coarse-
grained information entropy fulfil (E.1)?
5) Verify that (3.42) with (3.48) is solution of (E.1) provided the.t;;s are constants
of motion (cf. exercise 2).
177
166 6. Chance and Necessity
J= df )
. ( Kf-tQdq' (6.103)
(6.104)
This is the one-dimensional case of a continuity equation (cf. (6.73) and the
exercise): The temporal change of the probability density f(q) is equal to the negative
divergence of the probability current j.
2) n-Dimensional case
The Fokker-Planck equation (6.98) may be cast in the form
(6.105)
where
.
1k = [(,k f -
1 ".
2 L.,l= 1
Q af
kl aql •
(6.106)
1) One-dimension
We obtain from (6.104) by simple integration
j = const. (6.108)
178
6.4 Some Properties and Stationary Solutions of the Fokker-Planck Equation 167
Jr+_oof(q) dq = 1.
oo
(6.112)
2) n dimensions
Here, (6.107) with!" = 0 reads
(6.113)
Unfortunately, (6.113) does not always imply j = 0, even for natural boundary
conditions. However, a solution analogous to (6.110) obtains, ifthe drift coefficients
Kk(q) fulfil the so-called potential condition:
a
Kk = - -a
qk
V(q). (6.II4)
(6. II 5)
It is assumed that V(q) ensures thatf(q) vanishes for Iql --+ 00.
C) Examples
To illustrate (6.110) we treat a few special cases:
a)
K(q) = -l1.q. (6. II 7)
179
168 6. Chance and Necessity
We immediately find
(J. 2
V(q) =-q
2
which is plotted in Fig 6.5. The corresponding probability density f(q) is plotted
in the same figure. To interpretf(q) let us recall the Langevin equation correspond-
ing to (6.1 02), (6.117)
VJ
4= -rxq + F(t).
What happens to our particle with coordinate q is as follows: The random force
F(t) pushes the particle up the potential slope (which stems from the systematic
force, K(q». After each push, the particle falls down the slope. Therefore, the most
probable position is q = 0, but also other positions q are possible due to the random
pushes. Since many pushes are necessary to drive the particle far from q = 0, the
probability of finding it in those regions decreases rapidly. When we let rx become
smaller, the restoring force K becomes weaker. As a consequence, the potential
curve becomes flatter and the probability density f(q) is more spread out.
Once f(q) is known, moments (qn) = S qnf(q) dq may be calculated. In our
present case, (q) = 0, i.e., the center off(q) sits at the origin, and (q2) = (l/2)(Q/rx)
is a measure for the width of f(q) (compare Fig. 6.5).
b)
K(q) = -rxq - f3 q 3, (6.118)
rx f3
V(q) = "2q2 + 4 q4 ,
q= -rxq - f3 q 3 + F(t).
for ct < ° °
The case rx > is qualitatively the same as a), compare Fig. 6.6a. However,
a new situation arises (cf. Fig. 6.6b). While without fluctuations, the
180
6.4 Some Properties and Stationary Solutions of the Fokker-Planck Equation 169
V(q) f(q)
--.:.----"'"'''t''"''''--- q q
(0) (b)
Fig. 6.6a and b. Potential V(q) (solid line) and probability density (dashed line) for (6.118).
(a) 0( > 0, (b) 0«0
particle coordinate occupies either the left or right valley (broken symmetry,
compare Sect. 5.1), in the present casef(q) is symmetric. The "part~cle" may be
found with equal probability in both valleys. An important point should be men-
tioned, however. If the valleys are deep, and we put the particle initially at the
bottom of one of them, it may stay there for a very long time. The determination
of the time it takes to pass over to the other valley is called the "first passage time
problem".
c)
K(q) = -rxq _ yq2 _ fJ q 3, (6.119)
F()
y'q ="2rx q 2 +3Y3 fJ q4 ,
q +4
We assume l' > 0, fJ > 0 as fixed, but let rx vary from positive to negative values.
Figs. 6.7a-d exhibit the corresponding potential curves a)-d) and probability
densities. Note the pronounced jump in the probability density at q = 0 and q = q 1
when passing from Fig. c to Fig. d.
d) This and the following example illustrate the "potential case" in two dimen-
sions.
K1(q) = -rxq 1}
force, (6.120)
K 2(q) = -rxq2
181
170 6. Chance and Necessity
V,f Y,f
q q
(a) (b)
vJ V.f
--\
/ \
/ \
\
\.q ./
(e) (d)
Fig. 6.S. The potential belonging to Fig. 6.9. The distribution function belonging
(6.121) for IX > O to the potential (6.121), IX > O
182
6.4 Some Properties and Stationary Solutions of the Fokker-Planck Equation 171
(6.121)
or, in short,
We assume f3 > O. For rx > 0, potential surface and probability density f(q) are
shown in Figs. 6.8 and 6.9, respectively. For rx < 0, V and f are shown in Figs.
6.10 and 6.11, respectively. What is new compared to case b is the continuously
broken symmetry. Without fluctuating forces, the particle could sit anywhere at
the bottom of the valley in an (marginal) equilibrium position. Fluctuations drive
the particle round the valley, completely analogous to Brownian movement in one
dimension. In the stationary state, the particle may be found along the bottom with
equal probability, i.e., the symmetry is restored.
f) In the general case of a known potential V(q), a discussion in terms of Section
5.5 may be given. We leave it to the reader as an exercise to perform this "transla-
tion".
Exercise on 6.4
frq ,qp
Fig. 6.10. The potential belonging to Fig. 6.11. The distribution function belonging to
(6.121) for a < O the potential (6.121) for a < 0
183
172 6. Chance and Necessity
d
dt
Jq2 f(q) dq etc.
q,
K = -rxq
(by a simple shift of the origin of the q-coordinate we may cover also the case
K = c - rxq). We present a more or less heuristic derivation of the corresponding
solution. Because the stationary solution (6.110) has the form of a Gaussian
distribution, provided that the drift coefficients are linear, we try a hypothesis in
the form of a Gaussian distribution
where we admit that the width of the Gaussian distribution, a, the displacement, b,
and the normalization, .;V(t), are time-dependent functions. We insert (6.122) into
the time-dependent Fokker-Planck equation (6.102). After performing the differ-
entiation with respect to time t and coordinate q, we divide the resulting equation
on both its sides by
Eqs. (6.123) and (6.124) are linear differential equations for rx and f3 which can be
solved explicitly
Q
aCt) = - (1 - exp( - 2rxt))
IX
+ ao · exp( - 2rxt), (6.126)
184
6.5 Time-Dependent Solutions of the Fokker-Planck Equation 173
Eq. (6.125) looks rather grim. It is a simple matter, however, to verify that it is
solved by the hypothesis
(6.128)
which merely comes from the fact that (6.128) normalizes the distribution function
(6.122) for all times. Inserting (6.128) into (6.122) we obtain
Fig. 6.4 shows an example of (6.129). If the solution (6.129) is subject to the initial
condition a -+ 0 (i.e., ao = 0) for the initial time t -+ to == 0, (6.129) reduces to a
b-function, b(q - b o), at t = 0, or, in other words, it is a Green's function of the
Fokker-Planck equation. The same type of solutions of a Fokker-Planck equation
with linear drift and constant diffusion coefficients can be found also for many
variables q.
Inserting (6.129) into (6.130) we may evaluate the integral immediately (by passing
over to a new coordinate q = q' + bet»~ and obtain
(6.132)
(q(t)q(t'»
185
174 6. Chance and Necessity
In practical cases f(q', t') is taken as the stationary solution f(q), of the Fokker-
Planck equation, if not otherwise stated. f(q, t I q', t') is just that time-dependent
solution of the Fokker-Planck equation which reduces to the b-function b(q - q')
at time t = t'. Thus f(q, t I q', t') is a Green's function. In our present example
(6.117), we have f(q) = (rx/nQ)1/2 exp( _rxq2/Q) andf(q, t I q', t') given by (6.129)
with a o = 0 and bo = q'. With these functions we may simply evaluate (6.133)
where we put without loss of generality t ' = O.
Replacing q by q + q' exp( - rxt) we may simply perform the integrations (which
are essentially over Gaussian densities),
<q(t)q(t') = e- at <q'2)
1 Q -at
= 2;e , (6.137)
and insert it into (6.105). Performing the differentiation with respect to time and
then multiplying both sides of (6.105) by exp(At) yields
(6.139)
186
6.5 Time-Dependent Solutions of the Fokker-Planck Equation 175
combination of (6.138).
(6.140)
Even in the one-dimensional case and for rather simple K's and Q's (6.139) can be
solved only with help of computers.
J"=Lf (6.142)
(6.143)
If L were just a number, solving (6.142) would be a trivial matter andf(t) would
read
By inserting (6.144) into (6.142) and differentiating it with respect to time t, one
verifies immediately that (6.144) fulfils (6.142) even in the present case where L is
an operator. To evaluate (6.144), we define exp(Lt) just by the usual power series
expansion of the exponential function:
= 1 LVtV
eLI ,",co
L..v=ov!
_
. (6.145)
r means: apply the operator L v times on a function standing on the right of it:
187
176 6. Chance and Necessity
Leaving -r finite (but very small), we recast (6.142) into the form:
(6.148) becomes an exact solution of (6.142) in the limit -r ---+ 0, N ---+ 00 but N-r =
t - 10 , (6.148) is an alternative form to (6.144).
Exercises on 6.5
Verify that (6.129) with (6.126), (6.127), ao = 0, reduces to aD-function, D(q - qo)
for I ---+ O.
(6.149)
f(q, to + or) = .;V r: exp { - 2~-r (q - q' - -rK(q'»2} f(q', to) dq'. (6.150)
Readers, who are not so much interested in mathematical details may skip the
following section and proceed directly to formula (6.162). We will expand (6.150)
into a power series of-r, hereby proving, that (6.150) is equivalent to (6.147) up to
and including the order or, i.e., we want to show that the rhs of (6.150) may be
transformed into
d 2f
f + -r ( - dq Kf + t Q ddq2 ) ,where f = f(q, to)· (6.151)
q' = q + ~. (6.152)
188
6.6 Solution of the Fokker-Planck Equation by Path Integrals 177
ex p ( _ _
1
2Q,
e) (6.154)
gets very sharply peaked as , ~ O. More precisely, only those terms under the
integral are important for which lei < .j"T..JQ. This suggests to expand all factors
e
of (6.154) in (6.153) into a power series of and ,. Keeping the leading terms we
obtain after some rearrangements
(6.155)
where
(6.156)
Terms odd in ehave been dropped, because they vanish after integration over e.
Here
(6.161)
189
178 6. Chance and Necessity
we may establish the required identity with (6.147). We may now repeat the whole
procedure at times t2 = to + 2r, ... , tN = t2 + Nr and eventually pass over to
the limit N -> [fJ; Nr = t - to fixed. This yields an N-dimensional integral
(tN == t, to = 0)
where
We leave it to the reader to compare (6.162) with the path integral (4.85) and to
interpret (6.162) in the spirit of Chapter 4. The most probable path of the particle
is that for which 0 has a minimum, i.e.,
(6.164)
Letting r -> 0 this equation is just the equation of motion under the force K. In the
literature one often puts (l/r)(qv - qv-I) = q and replaces K(qv) by K(q). This is
correct, as long as this is only a shorthand for (6.163). Often different translations
have been used leading to misunderstandings and even to errors.
where
Exercise on 6.6
190
6.7 Phase Transition Analogy 179
Hint: Proceed as in the one-dimensional case and use (2.65), (2.66) with m --+ K
What is the most probable path?
Hint: Q and thus Q-l are positive definite.
(6.168)
h=M. (6.169)
Because we also want to treat other systems, we shall replace M by a general co-
ordinate q
In the following we assume that fF is the minimum of the free energy for a fixed
value of q. To proceed further we expand fF into a power series of q,
1
fF(q, T) = fF(O, T) + fFI(O, T)q + ... + 4! fF""(O, T)q4 + (6.171)
191
) 80 6. Chance and Necessity
due to inversion symmetry (cf. Sect. 5.1). Let us discuss this case first. We write
fF in the form
i.e., it changes its sign at the critical temperature T = Te. We therefore distinguish
between the two regions T> Te and T < Te (compare Table 6.1). For T> T e,
a > 0, and the minimum of fF (or V) lies at q = qo = o. As the entropy is con-
nected with the free energy by the formula (cf. (3.93»
s = _ ofF(q, T) (6.176)
oT '
we obtain in this region, T > Te ,
S = S = _ ofF(O, T) (6.176a)
o aT·
The second derivative of !F with respect to temperature yields the specific heat
(besides the factor T):
C = T(!:), (6.177)
(6.177a)
Now we perform the same procedure for the ordered phase T < Te , i.e., a < o.
This yields a new equilibrium value ql and a new entropy as exhibited in Table 6.1.
192
6.7 Phase Transition Analogy 181
which in the case of the potential (6.173) takes the explicit form
(6.179)
which we have met in various examples in our book in different context. For
simplicity, we have omitted a constant factor in front of offjoq. For (1 --> 0 we ob-
serve a phenomenon called critical slowing down, because the "particle" with
coordinate q falls down the potential slope more and more slowly. Furthermore
symmetry breaking occurs, which we have already encountered in Section 5.1.
The critical slowing down is associated with a soft mode (cf. Sect. 5.1). In statistical
mechanics one often includes a fluctuating force in (6.179) in analogy to Sect. 6.1.
For (1 --> 0 critical fluctuations arise: since the restoring force acts only via higher
powers of q, the fluctuations of q(t) become considerable.
We now turn to the case where the free energy has the form
q2 q3 q4
ff(q, T) = (12 + Y3" + fJ 4 (6.180)
193
182 6. Chance and Necessity
-t---~+oo::~---+. q
vase face
Fig. 6.12. Broken symmetry in visual perception. When focussing the attention to the center and
interpreting it as foreground of a picture, a vase is seen, otherwise two faces
(f3 and")' positive but (1. may change its sign according to (6.175)). When we change
the temperature T, i.e., the parameter (1., we pass through a sequence of potential
curves exhibited in Figs. 6.7a-d. Here we find the following situation.
When lowering temperature, the local minimum first remains at q = O. When
lowering temperature further we obtain the potential curve of Fig. 6.7d, i.e., now the
"particle" may fall down from q = 0 to the new (global) minimum of §' at qt.
The entropies of the two states, qo and q t, differ. This phenomenon is called "first
order phase transition" because the first derivative of §' is discontinuous. Since the
entropy S is discontinuous this transition is also referred to as a discontinuous
phase transition. When we now increase the temperature we pass through the figures
in the sequence d ---> a. It is apparent that the system stays at qt longer than it had
been before when going in the inverse direction of temperature. Quite obviously
hysteresis is present. Fig. 6. I 3. (Consider also Figs. 6.7a-d).
Our above considerations must be taken with a grain of salt. We certainly can
and will use the nomenclature connected with phase transitions as indicated above.
Indeed these considerations apply to many cases of nonequilibrium systems as we
shall substantiate in our later chapters. However, it must clearly be stated that the
Landau theory of phase transitions (Table 6.1) was found inadequate for phase
transitions of systems in thermal equilibrium. Here specific singularities of the
specific heat, etc., occur at the phase transition point which are described by so-
called critical exponents. The experimentally observed critical exponents do in
general not agree with those predicted by the Landau theory. A main reason for
this consists in an inadequate treatment of fluctuations as will transpire in the
following part. These phenomena are nowadays successfully treated by Wilson's
renormalization group techniques. We shall not enter into this discussion but
rather interpret several nonequilibrium transitions in the sense of the Landau theory
in regions where it is applicable.
In the rest of this chapter1 we want to elucidate especially what the discon-
tinuous transition of the specific heat means for nonequilibrium systems. As we
will show by explicit examples in subsequent chapters, the action of a macroscopic
system can be described by its order parameter q (or a set of them). In many cases
a measure for that action is q2. Adopting (6. I 74) as probability distribution, the
average value of q2 is defined by
1 This part is somewhat technical and can be skipped when reading this book for the first time.
194
6.7 Phase Transition Analogy 183
eillropy
d~r
.-----'001$,"9 r Fig. 6.13 a-<:. Behavior of a system at a first order
phase transition. To discuss Figs. a-<:, first have a look
at Figs. 6.7 a--d on page 170. When the potential curves
are deformed passing from Fig. 6.7a to Fig. 6.7d, we
obtain first one extremum (minimum) of V, then three,
and finally only one again.
Fig. 6.13a shows how the coordinate qo of these
extrema varies when the parameter determining the
(b) shape of the potential, V, changes. This parameter is in
Fig. 6.13 taken as (negative) temperature. Evidently,
when passing in Fig. 6.7 from a) to d), the system
stays at qo ~ 0 until the situation d) is reached, where the system jumps to a new equilibrium
value at the absolute minimum of V. On the other hand, when now passing from 6.7d to 6.7a
the system remains first at qo # 0 and jumps only in the situation 6.7b again to qo ~ O. These
jumps are evident in Fig. 6.l3c, where we have plotted the coordinate qo of the actually realized
state of the system. Fig. 6.l3b represents the corresponding variation of the entropy.
Jq2 exp( - V) dq
Jexp( - V) dq , (6.181)
where we have now absorbed Q into V, (cf. (6.116». We assume Pin the form
(6.182)
195
184 6. Chance and Necessity
2
q =-
ar- (6.183)
art.
which allows us to write (6.181) in the form
(6.185)
For the evaluation of the integral of (6.185) we assume that exp (- t/) is strongly
peaked at q = qo. If there are several peaks (corresponding to different minima
of r-(q» we assume that only one state q = qo is occupied. This is an ad hoc assump-
tion to take into account symmetry breaking in the "second-order phase transi-
tion", or to select one of the two local minima of different depth in the "first-order
phase transition". In both cases the assumption of a strongly peaked distribution
implies that we are still "sufficiently far" away from the phase transition point,
rt. = (J.c' We expand the exponent around that minimum value keeping only terms
of second order
(6.186)
Splitting the logarithm into two factors and carrying out the derivative with respect
to rt. in the first factor we obtain
(6.187)
196
6.7 Phase Transition Analogy 185
The last integral may be easily performed so that our final result reads
(6.188)
reveals that the first term on the rhs of (6.188) is proportional to the rhs of (6.176).
Therefore the entropy S may be put in parallel with the output activity (q2).
The discontinuity of (6.177) indicates a pronounced change of slope. (compare
Fig. 1.12). The second part in (6.188) stems from fluctuations. They are significant
in the neighborhood of the transition point IX = ric. Thus the Landau theory can be
interpreted as a theory in which mean values have been replaced by the most
probable values. It should be noted that at the transition point the behavior of
(q2) is not appropriately described by (6.187), i.e., the divergence inherent in the
second part of (6.188) does not really occur, but is a consequence of the method of
evaluating (6.185). For an illustration the reader should compare the behavior of
the specific heat which is practically identical with the laser output below and above
threshold.
The Landau theory of second order phase transitions is very suggestive for
finding an approximate (or sometimes even exact) stationary solution of a Fokker-
Planck equation of one or several variables q = (ql ... q.). We assume that for a
parameter IX > 0 the maximum of the stationary solution f(q) lies at q = O. To
study the behavior of f(q) at a critical value of ri, we proceed as follows:
1) We writef(q) in the form
1 1
+ 3! Illvl Pllvlqllqvql + 4! IllvAK PllvlKql'qvqlqK' (6.189)
197
186 6. Chance and Necessity
problem are provided by group theory). If the stationary solution of the Fokker-
Planck equation is unique, this symmetry requirement with respect to f(q) (or f/(q))
can be proven rigorously.
An example:
(6.190)
In the frame of a phenomenological approach the relations (6.174) and (6.178) are
generalized as follows:
Distribution function:
1Jff
q(x) = - 1Jq(x) , (6.192)
198
6.8 Phase Transition Analogy in Continuous Media: Space-Dependent Order Parameter 187
where q is now treated as function of space, x, and time, t. (Again, a constant factor
on the rhs is omitted). Inserting (6.l90) into (6.l92) yields the time-dependent
Ginzburg-Landau equation
. r 8
f = Jdnx { 8q(x) Q 8 }
(lXq(x) + pq(x) - yLlq(x» + 2" 8q(X)2 f·
3
2
(6.l95)
(6.196)
It can be easily solved by Fourier analyzing q(x, t) and F(x, t). Adopting the cor-
relation function (6.194), we may calculate the two-point correlation function
We quote one case: one dimension, equal times, t' = t, but different space points:
(q(X', t)q(x, t» = Q/(r:t.y)1/2 exp (_(1Xfy)1/2Ix' - xl). (6.199)
The factor of Ix' - xl in the exponential function has the meaning of (length)-l.
199
188 6. Chance and Necessity
We therefore put Ie = (a/y) -1/2. Since (6.199) describes the correlation between
two space points, Ie is called correlation length. Apparently
Ie --+ 00 as a --+ 0,
at least in the linearized theory. The exponent /1- in Ie OC a" is called critical exponent.
In our case, J1 = -1/2. The correlation function <q(x', t)q(x, t» has been evaluated
for the nonlinear case by a computer calculation (see end of this chapter). In
many practical cases the order parameter q(x) is a complex quantity. We denote it
by ~(x). The former equations must then be replaced by the following:
Langevin equation
Fokker-Planck equation
(6.203)
(6.204)
For equal times, t = t', the correlation functions (6.198) and (6.205) have been
determined for the nonlinear one-dimensional case (i.e., p #- 0) by using path-
integrals and a computer calculation. To treat the real q and the complex ~ in the
same way, we put
q(X)}
~(x) == lJI(x).
It can be shown that the amplitude correlation functions (6.198) and (6.205) can
be written in a good approximation, as
200
6.8 Phase Transition Analogy in Continuous Media: Space-Dependent Order Parameter 189
5 r-~-----'------,
.;:,
"'-
" "' ',',
r' mean"'"
1 field ""\',
\ '
\ \,
!..:.!.. H
-3 -2 -1 o 3 ,1t
-3 -2 -1 0 2 3 4t
Fig. 6.15. <1'1'12) versus "pump par- Fig. 6.16. Inverse correlation lengths I, and
ameter" (t- J)/M. (Compare text) 12 for real (dashed) and complex (solid) fields.
M = 2(PQ/2/0rxr,)2/3, 10 =(2r/rxo)'/2 Dashed curve: linearized theory ("mean field
(Redrawn after Scalapino et al.) theory"). (Redrawn after Scalapino et al.)
i.e., they can again be expressed by a single correlation length, 1\. <1lJ'1 2 is the >
average of lJ' over the steady state distribution. For reasons which will transpire
later we put a = C(o(l - t). (For illustration we mention a few examples. In super-
conductors, t = TlTe, where T: absolute temperature, Te: critical temperature,
c(o < 0; in lasers, t = DIDe, where D: (unsaturated) atomic inversion, Dc: critical
201
7. Self-Organization
Long-Living Systems Slave Short-Living Systems
In this chapter we come to our central topic, namely, organization and self-
organization. Before we enter into the mathematical treatment, let us briefly discuss
what we understand by these two words in ordinary life.
a) Organization
Consider, for example, a group of workers. We then speak of organization or,
more exactly, of organized behavior if each worker acts in a well-defined way on
given external orders, i.e., by the boss. It is understood that the thus-regulated
behavior results in a joint action to produce some product.
b) Self-Organization
We would call the same process as being self-organized if there are no external
orders given but the workers work together by some kind of mutual understanding,
each one doing his job so as to produce a product.
Let us now try to cast this rather vague description of what we understand by
organization or self-organization into rigorous mathematical terms. We have to
keep in mind that we have to develop a theory applicable to a large class of different
systems comprising not only the above-mentioned case of sociological systems but,
still more, physical, chemical, and biological systems.
7.1 Organization
The above-mentioned orders of the boss are the cause for the subsequent action of
the workers. Therefore we have to express causes and actions in mathematical
terms. Consider to this end an example from mechanics: Skiers on a ski lift pulled
uphill by the lift. Here the causes are the forces acting on the skiers. The action
consists in a motion of the skiers. Quite another example comes from chemistry:
Consider a set of vessels into which different chemicals are poured continuously.
This input causes a reaction, i.e., the output of new chemicals. At least in these
examples we are able to express causes and actions quantitatively, for example
by the velocity of skiers or the concentrations of the produced chemicals.
Let us discuss what kind of equations we will have for the relations between
causes and actions (effects). We confine our analysis to the case where the action
(effect), which we describe by a quantity q, changes in a small time interval At by
an amount proportional to At and to the size Fof the cause. Therefore, mathemat-
192 7. Self-Organization
q= -yq, (7.1)
where 'I is a damping constant. When an external "force" F is added, we obtain the
simple equation
where we shall neglect here and lateron transient effects. (7.3) is a simple example
for the following relation: The quantity q represents the response of the system
with respect to the applied force F(r). Apparently the value of q at time t depends
not only on the "orders" given at time t but also given in the past. In the following
we wish to consider a case in which the system reacts instantaneously i.e., in which
q(t) depends only on F(t). For further discussion we put for example
(7.4)
(7.5)
With help offormula (7.5) we can quantitatively express the condition under which
q acts instantaneously. This is the case if 'I » 15, namely
a I
q(t) ~ -e- M == -F(t), (7.6)
Y Y
or, in other words, the time constant to = 1/'1 inherent in the system must be much
shorter than the time constant t' = 1/15 inherent in the orders. We shall refer to this
assumption, which turns out to be of fundamental importance for what follows, as
"adiabatic approximation". We would have found the same result (7.6) if we had
204
7.1 Organization 193
put q = 0 in eq. (7.2) from the very beginning, i.e., if we had solved the equation
where A and B are matrices independent of qw We require that all matrix elements
of B which are linear or nonlinear functions of the F's vanish when F tends to zero.
The same is assumed for C. To secure that the system (7.8) is damped in the absence
of external forces (or, in other words, that the system is stable) we require that the
eigenvalues of the matrix A have all negative real parts
Re A. < O. (7.9)
Incidentally this guarantees the existence of the inverse of A, i.e., the determinant
is unequal zero
does not vanish, provided that the F's are small enough. Though the set of equa-
tions (7.8) is linear in qfJ.' a general solution is still a formidable task. However,
within an adiabatic elimination technique we can immediately provide an explicit
and unique solution of (7.8). To this end we assume that the F's change much
more slowly than the free system qw This allows us in exactly the same way as
discussed before to put
(7.12)
so that the differential equations (7.8) reduce to simple algebraic equations which
are solved by
(7.13)
205
194 7. Self-Organization
Note that A and B are matrices whereas C is a vector. For practical applications
an important generalization of (7.8) is given by allowing for different quantities
A, Band C for different subsystems, i.e., we make the replacements
(7.14)
in (7.13). The response qJl to the F's is unique and instantaneous and is in general a
nonlinear function of the F's.
Let us consider some further generalizations of (7.8) and discuss its usefulness.
a) (7.8) could be replaced by equations containing higher order derivatives of qJl
with respect to time. Since equations of higher order can always be reduced to a
set of equations of lower order e.g. of first order, this case is already contained in
(7.8). b) Band C could depend on F's of earlier times. This case presents no diffi-
culty here provided we still stick to the adiabatic elimination technique, but it leads
to an enormous difficulty when we treat the case of selforganization below. c) The
right hand side of (7.8) could be a nonlinear function of qw This case may appear
in practical applications. However, we can exclude it if we assume that the systems
Jl are heavily damped. In this case the q's are relatively small and the rhs of (7.8)
can be expanded in powers of q where in many cases one is allowed to keep only
linear terms. d) A very important generalization must be included later. (7.8) is a
completely causal equation, i.e., there are no fluctuations allowed for. In many
practical systems fluctuations play an important role, however.
In summary we can state the following: To describe organization quantitatively
we will use (7.8) in the adiabatic approximation. They describe a fairly wide class
of responses of physical, chemical and biological and, as we will see later on, socio-
logical systems to external causes. A note should be added for physicists. A good
deal of present days physics is based on the analysis of response functions in the
nonadiabatic domain. Certainly further progress in the theory of selforganization
(see below) will be made when such time-lag effects are taken into account.
7.2 Self-Organization
A rather obvious step to describe self-organization consists in including the external
forces as parts of the whole system. In contrast to the above-described cases,
however, we must not treat the external forces as given fixed quantities but rather
as obeying by themselves equations of motion. In the simplest case we have only
one force and one subsystem. Identifying now F with q1 and the former variable
q with qz, an explicit example of such equations is
Again we assume that the system (7.16) is damped in the absence of the system
206
7.2 Self-Organization 195
(7.15), which requires Yz > O. To establish the connection between the present case
and the former one we want to secure the validity of the adiabatic technique. To this
end we require
(7.17)
Though YI appears in (7.15) with the minus sign we shall later allow for both
Y1 ~ 0. On account of (7.17) we may solve (7.16) approximately by putting 42 =
which results in
°
(7.18)
Because (7.18) tells us that the system (7.16) follows immediately the system (7.15)
the system (7.16) is said to be slaved by the system (7.15). However, the slaved
system reacts on the system (7.15). We can substitute q2 (7.18) in (7.15). Thus we
obtain the equation
(7.19)
which we have encountered before in Section 5.1. There we have seen that two
completely different kinds of solutions occur depending on whether Y1 > or °
YI < 0. For YI > 0, q. = 0, and thus also q2 = 0, i.e., no action occurs at all.
However, if Y. < 0, the steady state solution of (7.19) reads
(7.20)
°
and consequently q2 "# according to (7.18). Thus the system, consisting of the two
° °
subsystems (7.15) and (7.16) has internally decided to produce a finite quantity qz,
i.e., nonvanishing action occurs. Since ql "# or ql = are a measure if action or
if no action occurs, we could call ql an action parameter.
For reasons which will become obvious below when dealing with complex
systems, ql describes the degree of order. This is the reason why we shall refer to
ql as "order parameter". In general we shall call variables, or, more physically
spoken, modes "order parameters" if they slave subsystems. This example lends
itself to the following generalization: We deal with a whole set of subsystems which
again are described by several variables. For all these variables we now use a single
kind of suffices running from I to n. For the moment being we assume these equa-
tions in the form
(7.21)
To proceed further, we imagine that we have arranged the indices in such a way
207
196 7. Self-Organization
that there are now two distinct groups in which i = 1, ... , m refers to modes
with small damping which can even become unstable modes (i.e., y :'S 0) and an-
other group with s = m + 1, ... , n referring to stable modes. It is understood
that the functions gj are nonlinear functions of qt, ... , qn (with no constant or
linear terms) so that in a first approximation these functions can be neglected
compared to the linear terms on the right hand side of equations (7.21). Because
where qm+ I, • . . , qn must be put equal zero in gs. Reinserting (7.23) into the first
m equations of (7.21) leads us to nonlinear equations for the q;'s alone.
(7.24)
The solutions of these equations then determine whether nonzero action of the
subsystems is possible or not. The simp1est example of (7.24) leads us back to an
equation of the type (7.19) or of e.g., the type
(7.25)
The equations (7.21) are characterized by the fact that we could group them
into two clearly distinct groups with respect to their damping, or, in other words,
into stable and (virtually) unstable variables (or "modes"). We now want to show
that self-organized behavior need not be subject to such a restriction. We rather
start now with a system in which from the very beginning the q's need not belong
to two distinct groups. We rather consider the following system of equations
(7.26)
where hj are in general nonlinear functions of the q's. We assume that the system
(7.26) is such that it allows for a time independent solution denoted by qJ. To
understand better what follows, let us have a look at the simpler system (7.21).
There, the rhs depends on a certain set of parameters, namely, the y's. Thus in a
more general case we shall also allow that the h/s on the rhs of (7.26) depend on
parameters- called (J l ' . . . , (J I' Let us first assume that these parameters are chosen
in such a way that the qO,s represent stable values. By a shift of the origin of the
208
7.2 Self-Organization 197
coordinate system of the q's we can put the qO's equal to zero. This state will be
referred to as the quiescent state in which no action occurs. In the following we put
and perform the same steps as in the stability analysis (compare Sect. 5.3 where
~(t) is now called u(t». We insert (7.27) into (7.26). Since the system is stable we may
assume that the u/s remain very small so that we can linearize the equations (7.26)
(under suitable assumptions about the h/s, which we don't formulate here ex-
plicitly). The linearized equations are written in the form
(7.28)
where the matrix element Ljj' depends on qO and simultaneously on the parameters
In short we write instead of (7.28)
0"1,0"2' ••••
Ii = Lu. (7.29)
(7.28) or (7.29) represent a set of first order differential equations with constant
coefficients. Solutions can be found as in 5.3 in the form
(7.30)
(7.31)
and u(/l)(O) are the right hand eigenvectors. The most general solution of (7.28) or
(7.29) is obtained by a superposition of (7.30)
(7.32)
(7.33)
Because we had assumed that the system is stable the real part of all eigenvalues
A/l is negative. We now require that the decomposition (7.27) satisfies the original
nonlinear equations (7.26) so that u(t) is a function still to be determined
Ii = Lu + N(u). (7.34)
We have encountered the linear part Lu in (7.28), whereas N stems from the
residual nonlinear contributions. We represent the vector u(t) as a superposition
of the right-hand eigenvectors (7.30) in the form (7.32) where, however, the ~'s are
209
198 7. Self-Organization
now unknown functions of time, and exp (A"t) is dropped. To find appropriate
equations for the time dependent amplitudes ~it) we multiply (7.34) from the
left by
v{"l(O) (7.35)
(7.36)
(7.37)
where
(7.38)
The lhs (7.37) stems from the corresponding lhs of (7.34). The first term on the
rhs of (7.37) comes from the first expression on the rhs of (7.34). Similarly, N is
transformed into g. Note that g" is a nonlinear function in the ~'s. When we now
identify~" with qj and A" with -Yj we notice that the equations (7.37) have exactly
the form of (7.21) so that we can immediately apply our previous analysis. To do
this we change the parameters 0"1' O"b' •• in such a way that the system (7.28)
becomes unstable, or in other words, that one or several of the A,,'S acquire a vanish-
ing or positive real part while the other A.'s are still connected with damped modes.
The modes ~" with Re A" ~ 0 then play the role of the order parameters which
slave all the other modes. In this procedure we assume explicitly that there are two
distinct groups of A,,'S for a given set of parameter values of 0" I, ••. ,0"••
In many practical applications (see Sect. 8.2) only one or very few modes ~"
become unstable, i.e., Re A" ~ O. If all the other modes ~" remain damped, which
again is satisfied in many practical applications, we can safely apply the adiabatic
elimination procedure. The important consequence lies in the following: Since
all the damped modes follow the order parameters adiabatically, the behavior of the
whole system is determined by the behavior of very few order parameters. Thus even
very complex systems may show a well regulated behavior. Furthermore we have
seen in previous chapters that order parameter equations may allow for bifurcation.
Consequently, complex system can operate in different "modes", which are well
defined by the behavior of the order parameters (Fig. 7.1). The above sections of
the present chapter are admittedly somewhat abstract. We therefore strongly advise
students to repeat all the individual steps by an explicit example presented in the
exercise. In practice, we often deal with a hierarchical structure, in which the
relaxation constants can be grouped so that
In this case one can apply the adiabatic elimination procedure first to the variables
210
7.2 Self-Organization 199
,,,,,.oj,, ~
subsystems
joint action
of total system (a)
(b)
output
Fig. 7.1. Typical example of a system composed of interacting subsystems each coupled to reser-
voirs (a). The reservoirs contain many degrees of freedom and we have only very limited knowl-
edge. They are treated by information theory (or in physics by thermodynamics or statistical
physics). After elimination of the reservoir variables, the "unstable" modes of the subsystems are
determined and serve as order parameters. The stable modes generate in many cases a feedback
loop selecting and stabilizing certain configurations of the order parameters (b)
connected with 'P>' leaving us with the other variables. Then we can apply this
methods to the variables connected with y(2) and so on.
There are three generalizations of our present treatment which are absolutely
necessary in quite a number of important cases of practical interest.
1) The equations (7.26) or equivalently (7.37) suffer from a principle draw back.
Assume that we have first parameters CT 1, CT 2 for the stable regime. Then all u = 0
in the stationary case, or, equivalently all ~ = 0 in that case. When we now go over
to the unstable regime, ~ = 0 remains a solution and the system will never go to
the new bifurcated states. Therefore to understand the onset of self-organization,
additional considerations are necessary. They stem from the fact that in practically
all systems fluctuations are present which push the system away from the unstable
points to new stable points with ~ =F- 0 (compare Sect. 7.3).
2) The adiabatic elimination technique must be applied with care if the Xs have
imaginary parts. In that case our above procedure can be applied only if also the
imaginary parts of Aunstable are much smaller than the real parts of Astable'
3) So far we have been considering only discrete systems, i.e., variables q depend-
ing on discrete indices j. In continuously extended media, such as fluids, or in con-
tinuous models of neuron networks, the variables q depend on space-points x in a
continuous fashion.
The points 1-3 will be fully taken into account by the method described in Sec-
tions 7.7 to 7.11.
In our above considerations we have implicitely assumed that after the adiabatic
elimination of stable modes we obtain equations for order parameters which are
now stabilized (cf. (7.19». This is a self-consistency requirement which must be
checked in the individual case. In Section 7.8 we shall present an extension of the
preseht procedure which may yield stabilized order-parameters in higher order of
a certain iteration scheme. Finally there may be certain exceptional cases where the
211
200 7. Self-Organization
damping constants ')'S of the damped modes are not large enough. Consider for
example the equation for a damped mode
If the order parameter qj can become too large so that qj - ')'s > 0, the adiabatic
elimination procedure breaks down, because ')'eff == -')'s + qj > O. In such a case,
very interesting new phenomena occur, which are used e.g., in electronics to con-
struct the so-called "universal circuit". We shall come back to this question in Sec-
tion 12.4.
Exercise on 7.2
V(q)
212
7.3 The Role of Fluctuations: Reliability or Adaptibility? Switching 201
V(q)
--~r-~------~~~~~------+--+---q
never realize that at q = q(2) there is a still more stable state. However, fluctuations
can drive the system from q(l) to q(2) by some diffusion process. Since q describes
the macroscopic performance of a system, among the new states may be those in
which the system is better adapted to its surroundings. When we allow for an
ensemble of such systems and admit competition, selection will set in (compare
Sect. lO.3). Thus the interplay between fluctuations and selection leads to an
evolution of systems.
On the other hand certain devices, e.g., tunnel diodes in electronics operate in
states described by an order parameter q with an effective potential curve e.g. that
of Fig. 7.3. By external means we can bring the tunnel diode (or other systems)
into the state q(l) and thus store information at q(1). With this state a certain
macroscopic feature of the system is connected (e.g., a certain electrical current).
Thus we can measure the state of the system from the outside, and the device
serves as a memory. However, due to fluctuations, the system may diffuse to state
q(2) thus losing its memory. Therefore the reliability of the device is lowered due to
fluctuations.
To process information we must be able to switch a device, thus bringing it
from q(l) to q(2). This can be done by letting the potential barrier become first
lower and lower (cf. Fig. 7.3). Diffusion then drives the system from q(l) to q(2).
When the potential barrier is increased again, the system is now trapped at q(2).
It may be that nature supports evolution by changing external parameters so that
the just-described switching process becomes effective in developing new species.
Therefore the size of the order parameter fluctuations is crucial for the performance
of a system, acting in two opposite ways: adaptibility and ease of switching require
large fluctuations and flat potential curves, whereas reliability requires small fluc-
tuations and deep potential valleys. How can we control the size of fluctuations?
In self-organizing systems which contain several (identical) subsystems this can be
achieved by the number of components. For a fixed, (i.e., nonfluctuating) order
parameter each subsystem, s, has a deterministic output qJs ) and a randomly
fluctuating output q?). Let us assume that the total output q(s) is additive
(7.39)
213
202 7. Self-Organization
Let us further assume that q(s) are independent stochastic variables. The total out-
put qtotal = Is q(S) can then be treated by the central limit theorem (cf. Sect. 2.15).
The total output increases with the number N of subsystems, while the fluctuations
increase only with IN. Thus we may control the reliability and adaptibility just
by the number of subsystems. Note that this estimate is based on a linear analysis
(7.39). In reality, the feedback mechanism of nonlinear equations leads to a still
more pronounced suppression of noise as we will demonstrate most explicitly in the
laser case (Chapter 8).
In conclusion, we discuss reliability of a system in the sense of stability against
malfunctioning of some of its subsystems. To illustrate a solution that self-organiz-
ing systems offer let us consider the laser (or the neuron network). Let us assume
that laser light emission occurs in a regular fashion by atoms all emitting light at
the same frequency Woo Now assume that a number of atoms get a different transi-
tion frequency WI. While in a usual lamp both lines Wo and WI appear-indicating
a malfunction-the laser (due to nonlinearities) keeps emitting only Woo This is a
consequence of the competition between different order parameters (compare
Sect. 5.4). The same behavior can be expected for neurons, when some neurons
would try to fire at a different rate. In these cases the output signal becomes merely
somewhat weaker, but it retains its characteristic features. If we allow for fluctua-
tions of the subsystems, the fluctuations remain small and are outweighed by the
"correct" macroscopic order parameter.
I by u ("unstable")
2 by s ("stable")
(7.40)
214
7.4 Adiabatic Elimination of Fast Relaxing Variables from the Fokker-Planck Equation 203
(7.42)
and
(7.43)
(7.44)
To obtain an equation for g(qu) alone we integrate (7.44) over q•. Using (7.42) we
obtain
(7.45)
(7.46)
Apparently, (7.45) contains the still unknown function h(q. I qJ. A closer inspec-
tion of (7.44) reveals that it is a reasonable assumption to determine h by the equa-
tion
(7.47)
This implies that we assume that h varies much more slowly as a function of qu
than as a function of q., or, in other words, derivates of first (second) order of h
with respect to qu can be neglected compared to the corresponding ones with
respect to q•. As we will see explicitly, this requirement is fulfilled if y. is sufficiently
215
204 7. Self-Organization
Lsh = 0, (7.48)
where Ls is the differential operator of the rhs of (7.47). As in our special example
the solution of (7.48) can be found explicitly with aid of (6.129). (Note that q. is
here treated as a constant parameter)
(7.50)
(7.51)
Provided the adiabatic condition holds, our procedure can be applied to general
functions F. and Fs (instead of those in (7.40». In one dimension, (7.48) can then
be solved explicitly (cf. Section 6.4). But also in several dimensions (see exercise)
(7.48) can be solved in a number of cases, e.g., if the potential conditions (6.114,115)
hold. In view of Sections 7.6 and 7.7 we mention that this procedure applies also
to functional Fokker-Planck equations.
Exercise on 7.4
Generalize the above procedure to a set of qu's and q;s, u = I, ... , k, s = 1, ... , I.
We put
216
7.6 Self-Organization in Continuously Extended Media. An Outline 205
and require
Inserting (7.53) into (7.52) and summing up over m. on both sides yields the still
exact equation
(7.56)
(7.57)
To derive an equation for the conditional probability H(m. I mu) we invoke the
adiabatic hypothesis namely that Xu changes much more slowly than X•. Conse-
quently we determine H from that part of (7.52), in which transitions between m~
and m. occur for fixed mu = m~: Requiring furthermore II = 0, we thus obtain
The equations (7.58) and (7.56) can be solved explicitly in a number of cases,
e.g., if detailed balance holds. The quality of this procedure can be checked by
inserting (7.53) with G and H determined from (7.56) and (7.58) into (7.52) and
making an estimate of the residual expressions. If the adiabatic hypothesis is valid,
these terms can be now taken into account as a small perturbation. Inserting the
solution H of (7.58) into (7.57) yields an explicit expression for lV, which can now
be used in (7.56). This last equation determines the probability distribution G of
the order parameters.
In this and the following sections we deal with equations of motion of continuously
extended systems containing fluctuations. We first assume external parameters
permitting only stable solutions and then linearize the equations, which define a
set of modes. When external parameters are changed, the modes becoming un-
stable are taken as order parameters. Since their relaxation time tends to infinity,
the damped modes can be eliminated adiabatically leaving us with a set of non-
linear coupled order-parameter equations. In two and three dimensions they allow
for example for hexagonal spatial structures. Our procedure has numerous practical
applications (see Chap. 8).
To explain our procedure we look at the general form of the equations which
217
206 7. Self-Organization
are used in hydrodynamics, lasers, nonlinear optics, chemical reaction models and
related problems. To be concrete we consider macroscopic variables, though in
many cases our procedure is also applicable to microscopic quantities. We denote
the physical quantities by U = (U j , U2 , ••• ) mentioning the following examples.
In lasers, U stands for the electric field strength, for the polarization of the medium,
and for the inversion density of laser active atoms. In nonlinear optics, U stands
for the field strengths of several interacting modes. In hydrodynamics, U stands
e.g., for the components of the velocity field, for the density, and for the tempera-
ture. In chemical reactions, U stands for the numbers (or densities) of molecules
participating in the chemical reaction. In all these cases, U obeys equations of the
following type
a
at U" = GlV, U) + D"V
2
Ull + Fit); f1 = 1,2, ... ,n. (7.59)
In it, G" are nonlinear functions of U and perhaps of a gradient. In most applica-
tions like lasers or hydrodynamics, G is a linear or bilinear function of U, though in
certain cases (especially in chemical reaction models) a cubic coupling term may
equally well occur. The next term in (7.59) describes diffusion (D real) or wave type
propagation (D imaginary). In this latter case the second order time derivative of
the wave equation has been replaced by the first derivative by use of the "slowly
varying amplitude approximation". The F,,(t)'s are fluctuating forces which are
caused by external reservoirs and internal dissipation and which are connected
with the damping terms occurring in (7.59).
We shall not be concerned with the derivation of equations (7.59). Rather, our
goal will it be to derive from (7.59) equations for the undamped modes which
acquire a macroscopic size and determine the dynamics of the system in the
vicinity of the instability point. These modes form a mode skeleton which grows
out from fluctuations above the instability and thus describes the "embryonic"
state of the evolving spatio-temporal structure.
U = Uo +q (7.60)
218
7.7 Generalized Ginzburg-Landau Equations for Nonequilibrium Phase Transitions 207
with
q -'
_(~I(X't)) . (7.61)
qn (x, t)
Splitting the rhs of(7.59) into a linear part, Kq, and a nonlinear part g(q) we obtain
(7.59) in the form
(7.62)
In it the matrix
(7.63)
(7.64)
(7.65)
g(2) mayor may not depend on V, and equally g(3) mayor may not depend on V
(If g(3) depends on V, the sequence of g and q must be chosen properly). First we
consider the linear homogeneous equation
(7.66)
where the matrix K is given by (7.64). To obtain the solution of (7.66) we split the
space- and time-dependent vector q into the product ofa time-dependent function
exp(At), a constant vector 0, and a space-dependent function X(x), i.e.
(7.68)
219
208 7. Self-Organization
As is well known, the solutions of (7.68) are only determined if X,,(x} obeys
proper boundary conditions. Such a condition could be that X,,(x} vanishes on the
boundary. Depending on the shape of the boundary, the solutions may take
various forms. For boundaries forming a square or cube in two or three dimen-
sions, respectively, products of sin functions are appropriate for X,,(x}. For a
circular boundary in the plane, Bessel functions of radius r multiplied by exp (icp)
can be used for x", where one chooses polar coordinates r, and cp. In the present
context we want to make as close contact as possible with the Ginzburg-Landau
theory of phase transitions of systems in thermal equilibrium. To this end we
assume that the system under consideration is infinitely extended, so that there are
no boundary conditions except that IX(x} I remains bounded for Ixl-+cx:>. The
appropriate solutions of (7.68) are then plane waves
K(ik}O = AO (7.66b)
after dividing both sides of (7.66 a} by X,,(x}. Because the matrix K whose elements
have, for instance, the form
(7.66c)
o= O(j)(ik).
By means of a formal trick we can write many of the following relations more
concisely. As we shall see below, the quantities K(ik}, Aj(ik}, and O(j)(ik} will always
be followed by exp(ikx}. In order to understand our formal trick, consider the
special case of one dimension, where kx reduces to kx and V to d/dx.
Let us further assume for simplicity that Aj(ik} can be expanded into a power
series of ik, i.e.
220
7.7 Generalized Ginzburg-Landau Equations for Nonequilibrium Phase Transitions 209
Incidentally,
is replaced by
Note that we have to retain the index k of Xk' because this index indicates by which
ik the operator V is to be replaced in OUl and Aj. As we shall see below, this
formalism is most convenient when dealing with infinitely extended media and the
corresponding plane-wave solutions. When we are dealing with finite boundaries,
however, k (or k) in general belongs to a discrete set, and in such cases it is
advantageous to retain the notation K (ik), etc. We leave it as an exercise to the
reader to formulate the final equations of Sect. 7.7 and 7.8 in that way as well
(where there are no "finite bandwidth excitations", as will be discussed below).
Because in general K is not self-adjoint, we introduce the solutions () of the
adjoint equation
(7.69)
where
(7.70)
221
210 7. Self-Organization
(7.71)
The requirements (7.67), (7.69), (7.71) fix the O's and O's besides "scaling"
factors S/V): O(j) -> OU)S/V), oU) -> O(j)Sj(V)-l. This can be used in the mode
expansion (7.72) to introduce suitable units for the ~k./S. Since ~ and 0 occur
jointly, this kind of scaling does not influence the convergence of our adiabatic
elimination procedure used below. We represent q(x) as superposition
(7.72)
Now we take a decisive step for what follows. It will be shown later by means
of explicit examples like laser or hydrodynamic instabilities that there exist finite
band width excitations of unstable modes. This suggests to build up wave packets
(as in quantum mechanics) so that we take sums over small regions of k together.
We thus obtain carrier modes with discrete wave numbers k and slowly varying
amplitudes ~k.j(X). The essentials of this procedure can be seen in a one-
dimensional example. Since k in (7.72) is a continuous variable, we write, instead
of (7.72),
(7.72 a)
Because ou> does not depend on the integration over k, we may put O(j) in front
ofthe integral. We now subdivide the integration over k into equidistant intervals
[k' - (jj2, k' + (jj2], where k' are equidistant discrete values. Recalling that Xk is an
exponential function, we then find
(7.72b)
(7.72b) = OU>.KLk,exp(ik'x) f O/ 2
~k'+Ii.jexp{ifx)df.
-0/2
.Kexp(ik'x) = xdx)
and
222
7.7 Generalized Ginzburg-Landau Equations for Nonequilibriurn Phase Transitions 211
In the following we shall drop the prime on k'. The generalization of (7.72) to three
dimensions is obvious. Thus, we replace (7.72) with our new hypothesis
But in contrast to (7.72), the sum in (7.72d) runs over a discrete set of k' s only,
and the ~' s do not only depend on time t, but are also slowly varying functions of
space x. Our first goal is to derive a general set of equations for the mode ampli-
tudes ~. To do so we insert (7.72d) into (7.62), multiply from the left hand side by
xt,(x)O(j) and integrate over a region which contains many oscillations of X" but
in which ~ changes very little. The resulting first expression on the left hand side
of (7.62) can be evaluated as follows:
Because of
(7.74)
(7.75)
223
212 7. Self-Organization
It is, of course, no problem to retain the higher-order terms, and this is indeed
required in a number of practical applications of our equations, for instance to
instabilities in fluid dynamics.
On the rhs of (7.62) we have to insert (7.72) into g (see (7.62) and (7.65», and to do
the multiplication and integration described above. Denoting the resulting func-
tion of ~ by 11, we have to evaluate
(7.77)
The explicit result will be given below, (7.87) and (7.88). Here we mention only a
few basic facts. As we shall see later the dependence of ~k.i(X) on x is important
only for the unstable modes and here again only in the term connected with 1 This
means that we make the following replacements in (7.77):
and
(7.79)
Since g (or 11) contains only quadratic and cubic terms in q, the evaluation of (7.77)
amounts to an evaluation of integrals of the form
(7.80)
where
(7.81)
and
(7.82)
where
(7.83)
Finally, the fluctuating forces FI' give rise to new fluctuating forces in the form
(7.84)
224
7.7 Generalized Ginzburg-Landau Equations for Nonequilibrium Phase Transitions 213
After these intermediate steps the basic set of equations reads as follows:
(7.85)
where
+ Lk'k"k'"
j'j"j'"
bkk'k"k'" ,j ,j' ,j" ,j"" Jkk'k·k"'~k'j'~k" j"ek"'j"" (7.86)
- (1.)
ale"'k"}}'}" - "
2 L..-p,vv'
O-(j)(k)OU')(k')OU"l(k")
Jl v v' { g,.nv'
(2) (k") + g"v'v,
(2) (k')} (7.87)
and
b""'k"k'"}}'}''}''' = ,,(3)
~JlV'V"V'" gllv'v"v'"
O-(j)(k)OU')(k')O(i")(k")OU'")(k"')
Jl v' ,," 11 m ,
(7.88)
and
An important point should be observed at this stage: Though bears two indices, e
k and u, these indices are not independent of each other. Indeed the instability
usually occurs only in a small region of k = kc (compare Fig. 7.4a-c). Thus we must
carefully distinguish between the k values at which (7.95) and (7.96) (see below)
are evaluated. Because of the "wave packets" introduced earlier, k runs over a set
of discrete values with Ikl = k c • The prime at the sums in the following formulas
indicates this restriction of the summation over k. Fig. 7.4c provides us with an
example of a situation where branchj of ~k.j can become unstable at two different
values of k. If k = 0 and k = kc are connected with a hard and soft mode, respec-
tively, parametrically modulated spatio-temporal mode patterns appear.
The basic idea of our further procedure is this. Because the undamped modes
may grow unlimited provided the nonlinear terms are neglected, we expect that the
amplitudes of the undamped modes are considerably bigger than those of the
damped modes. Since, on the other hand, close to the "phase transition" point the
relaxation time of the undamped modes tends to infinity, i.e., the real part of ~
tends to zero, the damped modes must adiabatically follow the undamped modes.
Though the amplitudes of the damped modes are small, they must not be neglected
completely. This neglect would lead to a catastrophe if in (7.86), (7.87), (7.88) the
225
214 7. Self-Organization
Re).Jo,k)
~--------------~~~~---------k
(a)
Re)..!o,k)
--~~~%~'~'-----------------------------+k
,,
(b)
ReA.{o,k)
(c)
Fig. 7.4. (a) Example of an eigenvalue A leading to an instability at k,. This dependence of A and
k corresponds qualitatively to that of the Brusselator model of chemical reactions. (cf. Section 9.4)
The dashed curve corresponds to the stable situation, the solid curve to the marginal, and the
dotted curve leads to an instability for a mode continuum near k,
(b) The situation is similar to that of Fig. 7.4a the instability occurring now at k=O. In the case
of the Brusselator model, this instability is connected with a hard mode
(c) Example of two simultaneously occurring instabilities. If the mode at k=O is a hard mode
and that at k=k" a soft mode, spatial and temporal oscillations can occur
cubic terms are lacking. As one convinces oneself very quickly, quadratic terms can
never lead to a globally stable situation. Thus cubic terms are necessary for a
stabilization. Such cubic terms are introduced even in the absence of those in the
226
7.7 Generalized Ginzburg-Landau Eq uations for N onequilibrium Phase Transitions 215
original equations by the elimination of the damped modes. To exhibit the main
features of our elimination procedure more clearly, we put for the moment
and drop all coefficients in (7.85). We assume I/;,sl « I/;,ul and, in a selfconsistent
way, /;,S oc /;,;,. Keeping in (7.85) only terms up to third order in /;'u, we obtain
(d~t - iu)(k, u) = Lk'k"u's (k', u')· (k", s) + Ln" (k', u')(k", u") u'u"
Consider now the corresponding equation for j = s. In it we keep only terms neces-
sary to obtain an equation for the unstable modes up to third order
If we adopt an iteration scheme using the inequality II;,sl « lI;,ul one readily con-
vinces oneself that 's is at least proportional to /;,; so that the only relevant terms in
(7.93) are those exhibited explicitly. We now use our second hypothesis, namely,
that the stable modes are damped much more quickly than the unstable ones which
is well fulfilled for the soft mode instability. In the case of a hard mode, we must be
careful to remove the oscillatory part of (k', u')(k", u") in (7.93). This is achieved
by keeping the time derivative in (7.93). We therefore write the solution of (7.93) in
the form
,-I
where we have used the notation of Heaviside calculus. According to it,
d
( dt - i.S) f(t) =
It -00 exp[Xs(t - r)]f(-r)d-r. (7.94a)
This definition remains valid iff is a vector and As a matrix acting in the vector
space to whichfbelongs. Using this definition and the results obtained in the next
section, we can eliminate the stable modes precisely (in the rigorous mathematical
sense). For most practical cases, however, an approximate evaluation of the ope-
rator (d/dt - As) -1 is quite sufficient; in the case of a soft mode, d/dt can be entirely
neglected, whereas in the case of a hard mode, d/dt must be replaced by the
corresponding sum of the frequencies of ~k.j(X, t) which occur in
227
216 7. Self-Organization
+ I",,,,,,,,,,
u'a"u'"
~"'u'C""'"'''''''' uu'u"u'" ~""u" ~""'u'" + Fk,u
(7.95)
F",u is defined by
. { dtd - ~s(O, k)
}-l F fi,s(X, t). (7.97)
1 Because of its importance, other forms of the slaving principle have been presented in
H. Haken: Advanced Synergetics, Springer Ser. Synergetics, Vol. 20 (Springer, Berlin, Heidel-
berg, New York, Tokyo 1983).
228
7.8 Higher-Order Contributions to Generalized Ginzburg-Landau Equations 217
and
(7.98b)
The ~'s are the expansion coefficients introduced in (7.72) which may be slowly
space dependent and also functions of time. The F's are fluctuating forces defined
in (7.84). k j are wave vectors, whereas iI, i2 , • • or m l , m 2 , ••• distinguish between
stable or unstable modes i.e.
Ij = u or s. (7.99)
Note that k and I j are not independent variables because in some regions of k the
modes may be unstable while they remain stable in other areas of k. We now
introduce the abbreviation
(7.100)
which defines the Ihs of that equation. The coefficients a and I have been given in
(7.87), (7.81). In a similar way we introduce the notation
(7.101)
where band J can be found in (7.88), (7.83). We further introduce the matrix
Am!(V' kl) 0
(
A. ~: l:,(V. k,) (7.102)
where the A's are the eigenvalues occurring in (7.85). With the abbreviations (7.98),
(7.100) to (7.102) the (7.85), if specialized to the stable modes, can be written in the
form
In the spirit of our previous approach we assume that the vector S is completely
determined by this equation. Because (7.103) is a nonlinear equation, we must
devise an iteration procedure. To this end we make the ansatz
(7.104)
where c(n) contains the components of U exactly n times. We insert (7.104) into
(7.103) and compare terms which contain exactly the same numbers of factors of
229
218 7. Self-Organization
u. Since all stable modes are damped while the unstable modes are undamped,
we are safe that the operator d/dt - As possesses an inverse. By use of this we find
the following relations
-1
(d )
e (2) (u) _
-
_
dt
_
As
.•
{Asuu· u. u + Fs},
~
(7.105)
This procedure allows us to calculate all e(n),s consecutively so that the C's are
determined uniquely. Because the modes are damped, we can neglect solutions of
the homogeneous equations provided we are interested in the stationary state or
in slowly varying states which are not affected by initial distributions of the damped
modes. Note that F is formally treated as being of the same order as u: u. This
is only a formal trick. In applications the procedure must possibly be altered to pick
out the correct orders of u and F in the final equations. Note also that 1 and
thus A may still be differential operators with respect to x but must not depend on
time t. We now direct our attention to the equations of motion for the unstable
modes. In our present notation they read
As mentioned above our procedure consists in first evaluating the stable modes as
a functional or function of the unstable modes. We now insert the expansion (7. 104)
where the C's are consecutively determined by (7.105) and (7.106) into (7.108).
Using the definition
e(O) = e(1) = 0 (7.109)
and collecting terms containing the same total number of u's, our final equation
230
7.9 Scaling Theory of Continuously Extended Nonequilibrium Systems 219
reads
( dd - Au ) U
_ _.
-
•
Au..u· u. u + 2Auus· u. L.=o C (v) + L'Io'2=0
• • n-[ •
Auss· C
(Vt}.
.C
(V2)
t '1+V2~n
(7.110)
In the general case the solution of (7.110) may be a formidable task. In practical
applications, however, in a number of cases solutions can be found (cf. Chap. 8).
In this section we want to develop a general scaling theory applicable to large classes
of phase transition-like phenomena in nonequilibrium systems. We treat an N-
component system and take fluctuations fully into account. Our present approach
is in general only applicable to the one-dimensional case, though there may be
cases in which it applies also to two and three dimensions. The basic idea of our
procedure is as follows:
We start from a situation in which the external parameters permit only stable
solutions. In this case the equations can be linearized around their steady state
values and the resulting equations allow for damped modes only. When changing
an external parameter, we eventually reach a marginal situation with one or several
modes becoming unstable. We now expand the nonlinear equations around the
marginal point with respect to powers of e, where e2 denotes the deviation of the
actual parameter from the parameter value at the marginal point. In the framework
of perturbation theory applied to the linearized equations one readily establishes
that the complex frequency depends on e2 • This leads to the idea to scale the time
with 8 2 • On the other hand, if the linearized equations contain spatial derivatives,
one may show rather simply that the corresponding changes are so that the space
coordinate r goes with e. In the following we shall therefore use these two scalings
and expand the solutions of the nonlinear equations into a superposition of solu-
tions of the linearized equations at the marginal point. Including terms up to third
order in e, we then find a self-consistent equation for the amplitude of the marginal
solution. The resulting equation is strongly reminiscent oftime-dependent Ginzburg-
Landau equations with fluctuating forces (compare Sect. 7.6, 7.7, 7.8.). As we
treat essentially the same system as in Section 7.6 we start with (7.62) which we
write as
231
220 7. Self-Organization
(7.112)
In a perturbation approach one could easily solve (7.111) by taking the inverse of
r. However, some care must be exercised because the determinant may vanish in
actual cases. We therefore write (7.111) in the form
(7.115)
Note that Un and Vn are time and space dependent functions. They may be repre-
sented by plane wave solutions. The index n distinguishes both, the k-vector and
the corresponding eigenvalue of r: n = (k, m). The vectors (7.114) and (7.115)
form a biorthogonal set with the property
(7.116)
,
D.' = D!O) + D!Z)e
, 2/ (7.117)
Kij = K{J) + KlJ)e 2 •
232
7.9 Scaling Theory of Continuously Extended Nonequilibrium Systems 221
To take into account finite band width effects, space and time are scaled simul-
taneously by
(7.119)
c) Perturbation Theory
Inserting the expansions (7.117), (7.118) and the scaling laws (7.119) in (7.113)
we obtain
(7.121)
We extract q(l) from the Ihs and bring the rest to the rhs. This yields
We will explicitly treat two different situations which may arise at the critical
point. These situations are distinguished by a different behavior of the complex
frequencies A.
I) one soft mode, that is for one and only one mode Al = 0,
2) the hard mode case in which a pair of complex conjugate modes becomes
unstable so that Re Al = 0 but 1m Al "# O.
To write the perturbation theory in the same form for both cases, we introduce
the concept of modified Green's functions. To this end we rewrite the homogeneous
problem (7.114)
(7.123)
We now define Q by
(7.124)
where I is the unity matrix. The Green's function belonging to (7.124) reads
(7.125)
233
222 7. Self-Organization
The concept of modified Green's functions is inevitable to treat the soft mode
case Al = 0, where the index 1 means (k = ko m = 1). By means of the Green's
function the solution can be written in the form
Obviously the same procedure is applicable to the left hand eigenvectors. If the
critical mode is degenerate, it is straightforward to define an appropriate, modified
Green's function. Instead of (7.126), one now gets a set of h orthogonality condi-
tions. (h is the degree of degeneracy). Each one of these conditions must be fulfilled
separately.
(7.129)
(7.130)
(7.131)
(7.l32)
".
~l D(O)f(?)(A k 2-
t i l l = O', ) = tr (Df(A 1 0 k 2 )) = 0 ,
= , (7.l33)
234
7.10 Soft-Mode Instability 223
where tr means "trace". The quantities Fii are the adjoint subdeterminants of
order (N - 1) to the elements r ii of the matrix r. Now one may construct the right-
hand and left-hand eigenvectors of the critical mode. We represent the right-hand
eigenvector as
(7.134)
(7.135)
_1
L
J+LI 2dx·e i(k-k')x In~1
N - • 1·_
an(k, l)an(k ,J) - bkk,b ji • (7.136)
-L12
Because of the critical mode which is dominating all other modes at the marginal
point the solution of (7.122) in order 8 1 is given by
(7.137)
The common factor ~(R, T) is written as a slowly varying amplitude in space and
time. In order 8 2 we find the following equation
(7.138)
Because det 1T1(1) contains differential operators, one may readily show that on ac-
count of (7.133)
(7.139)
(7.140)
235
224 7. Self-Organization
We finally discuss the balance equation of third order. The corresponding solv-
ability condition
reduces on account of
(7.143)
and
to
(7.145)
To evaluate (7.145) we calculate the single terms separately. For det JrI(2) we find
(7.146)
(7.147)
1- -_"
L..i r ii
-(0)
(AI' kJ (7.l48a)
B = Ii D\O)f'\?)(Al, kJ + 2 Li,j D!O) D~Ol...lii,jjk; (7.148b)
C = Iij (- KU) + D!2)k;b i)f'!J 1(A\, kJ. (7. 148c)
1 We have split the original g(2) into a V-independent part, g(21, and a part with the scaled V, g'(2).
236
7.10 Soft-Mode Instability 225
We now evaluate the right-hand side of (7.145). Inserting the explicit form of v 1 ,
and q(2) we obtain
9(2)
+ rm/A1,kJ(gijk
- (2)
+ gikj)[ak(k
(2) "I - *
e, I) u AlO)aj(O,l)a/O,l)gpqraq(ke,I)ar(ke' I)
Finally we discuss the fluctuating forces. We thereby assume that they are
slowly varying in time and contribute in order 8 3 • For the fluctuating force we get
(7.l52)
with
and
(7.154)
Again the fluctuating forces are Gaussian. Combining the expressions (7.147),
(7.150), (7.151) we can write the final equation for the slowly varying amplitude
~(R, T) which reads
(7.155)
237
226 7. Self-Organization
_ _I
(vn,pl "n',p') - T
J+ T 2
/
e
-iro(p-p')t
Ij=1
N - I I _
a/p,n)a/p,n)dt - (5"n,(5pp' (7.156)
-T/2
where
T
2n p and p' are mtegers.
= -,
. (7.157)
Wo
°
We assume that 1m Al 01 is practically fixed, whereas the real part of AI goes
with e2 which again serves as small expansion parameter. While the e1 balance had
defined the homogeneous problem, the e2 balance yields the equation
(7.158)
On account of
(7.159)
(7.158) has the same structure as (7.138) of the preceding paragraph. Thus the
solvability condition is fulfilled. Taking the Green's function
G -_ 1....P
\' \,1 I ipwo(t-t')I)( I (7.160)
1...." A _ ipw e "n,p vn,p'
n 0
q~2) = e In A n -
12 , ap, n)ap, n)gINak(l, l)al(I, l)e2iwot
1Wo
Inserting (7.161) into the e3 balance equation, leads on the rhs to terms of the form
where F(R, T) is the projection of the fluctuations on the unstable modes. On the
238
7.11 Hard-Mode Instability 227
(7.163)
239
8. Physical Systems
In the laser, the light field is produced by the excited atoms. We describe the field
by its electric field strength, E, which depends on space and time. We consider only
a single direction of polarization. We expand E = E(x, t) into cavity modes
a) Field Equations
(8.2)
KA is the decay constant of mode A if left alone in the cavity without laser action.
KA takes into account field losses due to semi-transparent mirrors, to scattering
centers etc. gflA is a coupling constant describing the interaction between mode A
and atom fl. g #A is proportional to the atomic dipole matrix element. FA is a stochastic
force which occurs necessarily due to the unavoidable fluctuations when dissipation
is present. Eq. (8.2) describes the temporal change of the mode amplitude b A due
to different causes: the free oscillation of the field in the cavity (~ w)), the damping
(-K)), the generation by oscillating dipole moments (-ig#AIX,J, due to fluctuations
(e.g., in the mirrors), (~ F). On the other hand, the field modes influence the atoms.
This is described by the
b) Matter Equations
1) Equationsfor the Atomic Dipole Moments
(8.3)
v is the central frequency of the atom, y its linewidth caused by the decay of the
242
8.3 The Order Parameter Concept 231
atomic dipole moment. r it) is the fluctuating force connected with the damping
constant 1'. According to (8.3), IXI' changes due to the free oscillation of the atomic
dipole moment, (- iv), due to its damping (-1') and due to the field amplitudes
( ~ b.). The factor aI' serves to establish the correct phase-relation between field and
dipole moment depending on whether light is absorbed (al' == (N2 - N 1 )1' < 0) or
emitted (al' > 0). Finally, the inversion also changes when light is emitted or
absorbed.
(8.4)
A discussion of the physical content of (8.2) to (8.4) will help us to cut the problem
down and solve it completely. Eq. (8.2) describes the temporal change of the mode
amplitude under two forces: a driving force stemming from the oscillating dipole
moments (rxl') quite analogous to the classical theory of the Hertzian dipole, and a
stochastic force F. Eqs. (8.3) and (8.4) describe the reaction of the field on the atoms.
Let us first assume that in (8.3) the inversion a I' is kept constant. Then b acts as a
driving force on the dipole moment. If the driving force has the correct phase and
is near resonance, we expect a feedback between the field and the atoms, or, in other
words, we obtain stimulated emission. This stimulation process has two opponents.
On the one hand the damping constants K and l' will tend to drive the field to zero
and, furthermore, the fluctuating forces will disturb the total emission process by
their stochastic action. Thus we expect a damped oscillation. As we shall see more
explicitly below, if we increase aI" suddenly the system becomes unstable with ex-
ponential growth of the field and correspondingly of the dipole moments. Usually
it is just a single-field mode that first becomes undamped or, in other words, that
243
232 8. Physical Systems
becomes unstable. In this instability region its internal relaxation time is apparently
very long. This makes us anticipate that the mode amplitudes, which virtually be-
come undamped, may serve as the order parameters. These slowly varying am-
plitudes now slave the atomic system. The atoms have to obey the orders of the
order parameters as described by the rhs of (8.3) and (8.4). If th~ atoms follow im-
mediately the orders of the order parameter, we may eliminate the "atomic"
variables 0(+, ex, a, adiabatically, obtaining equations for the order parameters b).
alone. These equations describe most explicitly the competition of the order
parameters among each other. The atoms will then obey that order parameter which
wins the competition. In order to learn more about this mechanism, we anticipate
that one b). has won the competition and we first confine our analysis to this
single-mode case.
We drop the index JI. in (8.2) to (8.4), assume exact resonance, W = v, and eliminate
the main time dependence by the substitutions
(8.5)
where we finally drop the tilde, ,....,. The equations we consider are then
We note that for running waves (in a single direction) the coupling coefficients gil
have the form
(8.9)
g is assumed real. Note that the field-mode amplitude b is supported via a sum of
dipole moments
(8.10)
We first determine the oscillating dipole moment from (8.7) which yields in an
elementary way
with
(8.12)
244
8.4 The Single-Mode Laser 233
We now make a very important assumption which is quite typical for many
cooperative systems (compare Section 7.2). We assume that the relaxation time
of the atomic dipole moment Q( is much smaller than the relaxation time inherent
in the order parameter b as well as in (JIt. This allows us to take b· (JIt out of the
integral in (8.11). By this adiabatic approximation we obtain
(8.13)
(8.13) tells us that the atoms obey instantaneously the order parameter. Inserting
(8.13) into (8.6) yields
(8.14)
where P is now composed of the field noise source, F, and the atomic noise sources,
r,
(8.15)
In order to eliminate the dipole moments completely, we insert (8.13) into (8.8).
A rather detailed analysis shows that one may safely neglect the fluctuating forces.
We therefore obtain immediately
(8.16)
We now again assume that the atom obeys the field instantaneously, i.e., we put
(8.17)
(8.18)
Because we shall later be mainly interested in the threshold region where the char-
acteristic laser features emerge, and in that region b + b is still a small quantity, we
replace (8.18) by the expansion
(8.19)
As we shall see immediately, laser action will start at a certain value of the inversion
do. Because in this case b + b is a small quantity, we may replace do by de in the
second term of (8.19) to the same order of approximation. We introduce the total
inversion
(8.20)
245
234 8. Physical Systems
(8.21)
(8.22a)
(compare (6.118)).
Thus we may apply the results of that discussion in particular to the critical
region, where the parameter rx changes its sign. We find that the concepts of sym-
metry breaking instability, soft mode, critical fluctuations, critical slowing down,
are immediately applicable to the single mode laser and reveal a pronounced
analogy between the laser threshold and a (second order) phase transition. (cf.
Sect. 6.7). While we may use the results and concepts exhibited in Sections 5.1, 6.4,
6.7 we may also interpret (8.22) in the terms of laser theory. If the inversion Do is
small enough, the coefficient of the linear term of (8.22) is negative. We may safely
neglect the nonlinearity, and the field is merely supported by stochastic processes
(spontaneous emission noise). Because F is (approximately) Gaussian also b is
Gaussian distributed (for the definition of a Gaussian process see Sect. 4.4). The
inverse of the relaxation time of the field amplitude b may be interpreted as the
optical line width. With increasing inversion Do the system becomes more and more
undamped. Consequently the optical line width decreases, which is a well-observed
phenomenon in laser experiments. When rx (8.22a) passes through zero, b acquires a
new equilibrium position with a stable amplitude. Because b is now to be inter-
preted as a field amplitude, this means that the laser light is completely coherent.
This coherence is only disturbed by small superimposed amplitude fluctuations
caused by F and by very small phase fluctuations.
If we consider (8.22) as an equation for a complex quantity b we may derive'the
right hand side from the potential
(8.23)
By methods described in Sections 6.3 and 6.4, the Fokker-Planck equation can
be established and readily solved yielding
2V(lbl))
feb) = ,;V exp ( --Q- , (8.24)
246
8.5 The Multimode Laser 235
03
t a--2
\
,\0
~ v\ V~ V,
01
o
/' ~ ....... '-.......
~ C 8 10
Fig. 8.1. The stationary distribution as a
function of the normalized "intensity" n.
(After H. Risken: Z. Physik 186, 85 (1965))
"
n -
where Q (compare (6.91» measures the strength of the fluctuating force. The
function (8.24) (cf. Fig. 8.1) describes the photon distribution of laser light, and
has been checked experimentally with great accuracy. So far we have seen that by
the adiabatic principle the atoms are forced to obey immediately the order para-
meter. We must now discuss in detail why only one order parameter is dominant.
If all parameters occurred simultaneously, the system could still be completely
random.
Exercise on 8.4
Verify that
f . = [ -8b(-rxb
8 82 ] f
- Plbl 2 b) + c.c. + Q8b8b*
(8.25)
247
236 8. Physical Systems
IX
I'
= i"
~).
g*1'). JI
-00
e(-iV-Y)(I-T)(b
).
(J)
I't
dor: + t 1" (8.26)
(8.27)
(8.28)
We thus obtain
(8.29)
(8.30)
where n). is the number of photons of the mode A. If we neglect for the moment
being the fluctuating forces in (8.29) we thus obtain
(8.31)
with
2yg2
= -:-:---'--'c"----,, (8.32)
(D). - vf + y2'
W
).
Igl').12 = g2. (8.33)
(8.34)
(8.35)
where Dc is the critical inversion of all atoms at threshold. To show that (8.31)-
248
8.6 Laser with Continuously Many Modes. Analogy with Superconductivity 237
(8.35) lead to the selection of modes (or order parameters), consider as example
just two modes treated in the exercise of Section 5.4. This analysis can be done
quite rigorously also for many modes and shows that in the laser system only a
single mode with smallest losses and closest to resonance survives. All the others
die out. It is worth mentioning that equations of the type (8.31) to (8.35) have been
proposed more recently in order to develop a mathematical model for evolution.
We will come back to this point in Section 10.3.
As we have seen in Section 6.4, it is most desirable to establish the Fokker-
Planck equation and its stationary solution because it gives us the overall picture of
global and local stability and the size of fluctuations. The solution of the Fokker-
Planck equation, which belongs to (8.29), with (8.35) can be found by the methods
of Section 6.4 and reads
(8.36)
where
(8.37)
The local minima of (/> describe stable or metastable states. This allows us to study
multimode configurations if some modes are degenerate.
The next example, which is slightly more involved, will allow us to make contact
with the Ginzburg-Landau theory of superconductivity. Here we assume a con-
tinuum of modes all running in one direction. Similar to the case just considered,
we expect that only modes near resonance will have a chance to participate at laser
action; but because the modes are now continuously spaced, we must take into
consideration a whole set of modes near the vicinity of resonance. Therefore we
expect (which must be proven in a self-consistent way) that only modes with
and
(8.39)
are important near laser threshold. Inserting (8.27) into (8.4), we obtain
(8.40)
249
238 8. Physical Systems
(S.42)
(S.43)
where N is the number of laser atoms. Note that we have again assumed (S.3S)
in the nonlinear part of (S.42). If
(S.44)
possesses no dispersion, i.e., Q). IX k)., the following exact solution of the cor-
responding Fokker-Planck equation holds
where
(S.46)
We do not continue the discussion of this problem here in the mode picture but
rather establish the announced analogy with the Ginzburg-Landau theory. To this
end we assume
(S.50)
250
8.6 Laser with Continuously Many Modes. Analogy with Superconductivity 239
We now replace the index A. by the wave number k and form the wave packet
'J'(x, t) = J+OO
-00 Bke+ikX-ivlkltdk. (8.51)
.
'J'(x, t) = -a'J'(x, t) +c (d
iv dx + v )2 'J'(x, t) - 2f31'J'(x, tW'J'(x, t) + F(x, t),
(8.52)
(8.53)
Eq. (8.52) is identical with the equation of the electron-pair wave function of the
Ginzburg-Landau theory of superconductivity for the one-dimensional case if the
following identifications are made:
Table 8.1
Superconductor Laser
'1', pair wave function '1', electric field strength
ex DC T-Te ex DC De-D
T temperature D total inversion
Te critical temperature Dc critical inversion
v DC Ax-component of v atomic frequency
vector potential
F(x, t) thermal fluctuations fluctuations caused by spon-
taneous emission, etc.
Note, however, that our equation holds for systems far from thermal equilibrium
where the fluctuating forces, in particular, have quite a different meaning. We may
again establish the Fokker-Planck equation and translate the solution (8.45) (8.46)
to the continuous case which yields
(8.54)
with
(8.55)
Eq. (8.55) is identical with the expression for the distribution function of the
Ginzburg-Landau theory of superconductivity if we identify (in addition to Table
8. I) cP with the free energy g; and Q with 2kBT. The analogy between systems away
251
240 8. Physical Systems
where Eo is the real amplitude and qJo a fixed phase. The frequency 0)0 is assumed in
resonance with the atomic transition and also with the field mode. ko is the cor-
responding wave number. We assume that the field strength Eo is so weak that it
practically does not influence the atomic inversion. The only laser equation that
E enters into is the equation for the atomic dipole moments. Since here the external
field plays the same role as the laser field itself, we have to replace b by b + const· E
in (8.7). This amounts to making the replacement
(8.57)
where
(8.58)
lt is now a simple matter to repeat all the iteration steps performed after (8.11).
We then see that the only effect of the change (8.57) is to make in (8.15) the following
replacements
l' . '\' n d - i'Po - 1
r -->
~
F - 1 L.." g"\7,, oEoe l' . (8.59)
(8.60)
252
8.7 First-Order Phase Transitions of the Single-Mode Laser 241
(8.61)
Again it is a simple matter to write the rhs of (8.60) and of its conjugate complex
as derivatives of a potential where we stilI add the fluctuating forces
(8.64)
(8.65)
where Vo is identical with the rotation symmetric potential of formula (8.23). The
additional cosine term destroys the rotation symmetry. If Eo = 0, the potential is
rotation symmetric and the phase can diffuse in an undamped manner giving rise
to a finite linewidth of the laser mode. This phase diffusion is no more possible if
Eo =1= 0, since the potential considered as a function of qJ has now a minimum at
qJ = 'Po + n. (8.66)
The corresponding potential is exhibited in Fig. 8.2 which really exhibits a pinning
V(r,,!)
253
242 8. Physical Systems
of l{J. Of course, there are still fluctuating forces which let l{J fluctuate around the
value (8.66).
(8.67)
While (8.67) is a reasonable approximation for not too high intensities Ib1 2 , at very
high fields (8.67) would lead to the wrong result that the loss becomes negative.
Indeed one may show that the loss constant should be replaced by
(8.68)
From it the meaning of I, becomes clear. It is that field intensity where the decrease
of the loss becomes less and less important. It is a simple matter to treat also the
variation of the inversion more exactly which we take now from (8.18). Inserting
this into (8.13) and the resulting expression into (8.6) we obtain, using as loss
constant (8.68)
+
db Idt = -
(
/(0 + I +" ,) +
IWlls b + I +
G
IWlla, b
+
+ F~+ (t) (8.69)
(8.70)
and
(8.71)
Again (8.69) and its conjugate complex possess a potential which can be found by
integration in the explicit form
For a discussion of (8.72) let us consider the special case in which I, « las. This
allows us to expand the second logarithm for not too high Ibj2, but we must keep
254
8.8 Hierarchy of Laser Instabilities and Ultrashort Laser Pulses 243
(8.73)
IJ = IJT
IJ < IJT
/---f---t---t- IJ« IJ T
is exhibited in Fig. 8.3. With change of pump parameter G the potential is de-
formed in a way similar to that of a first order phase transition discussed in Section
6.7 (compare the corresponding curves there). In particular a hysteresis effect
occurs.
(8.74)
(8.75)
(8.76)
Eq. (8.76) follows directly from Maxwell's equation, while (8.77), (8.78) are
255
244 8. Physical Systems
aE(-) ~
aE(-)/at + c-ax- + KE(-) = -2niw
0
p(-)
,
(8.80)
These equations are equivalent to (8.2), (8.3), (8.4) provided we assume there
KA = K and a set of discrete running modes. We now start with a small inversion,
Do, and investigate what happens to E; P, D, when Do is increased.
E = P = 0, D = Do, (8.83)
i.e., no laser action occurs. To check the stability of the solution (8.83), we perform
a linear stability analysis putting
(8.84)
P= beAt - ikx, (8.85)
15 - Do = ceAt- ikx, (8.86)
256
8.8 Hierarchy of Laser Instabilities and Ultrashort Laser Pulses 245
where
(8.89)
A closer investigation of (8.88), (8.89) reveals that ..1.(2) ::::: 0 for Do ::::: D thr and an
instability occurs, at first for k = O.
is now the stable solution. The rhs of (8.91) are space and time-independent con-
stants, obeying (8.80)-(8.82). This solution corresponds to that of Sections 8.4-6
in the laser domain if fluctuations are neglected. For what follows it is convenient
to normalize E, P, D with respect to (8.91), i.e., we put
(8.93)
(~at + K + C~)E
ax = KP. (8.96)
E=P=D=l. (8.97)
(8.98)
257
246 8. Physical Systems
where we abbreviate
(8.99)
3) Do ~ D thr . 2 •
(8.101)
(8.102)
with
The further steps are those of Section 7.7 and 7.8. Resolving the equations for the
stable modes up to fourth order in the amplitude of the unstable mode ~kc,u yields
the following equation:
(8.105)
(8.106)
258
8.8 Hierarchy of Laser Instabilities and Ultrashort Laser Pulses 247
(-ata+ caxa)
- R = fJR + aR 3 - bR 5 ==
av
aR
(8.107)
with
b = Re 6> 0, (8.108)
x= x/c - t, 1 = t (8.109)
we transform (8.96) into a first order differential equation with respect to time,
where x plays the role of a parameter which eventually drops out from our final
solution. Interpreting the rhs of (8.107) as a derivative of a potential function V
and the Ihs as the acceleration term of the overdamped motion of a particle, the
behavior of the amplitude can immediately be discussed by means of a potential
curve. We first discuss the case Re Ii > 0. Because the shape of this curve is quali-
tatively similar to Fig. 8.3, the reader is referred to that curve. For
°
there is only one minimum, R = 0, i.e., the continuous wave (cw) solution is stable.
For -I < d < the cw solution is still stable; but a new minimum of the potential
finally d > 0, R = °
curve indicates that there is a second new state available by a hard excitation. If
becomes unstable (or, more precisely, we have a point of
°
marginal stability) and the new stable point lies at R = R 2 • Thus the laser pulse
amplitude jumps discontinuously from R = to R = R 2 • The coefficients of (8.105)
are rather complicated expressions and are not exhibited here. Eq. (8.107) is solved
numerically; and the resulting R (and the corresponding phase) are then inserted
into (8.103). Thus the electric field strength E takes the form
where we have included terms up to the cubic harmonic. Fig. 8.4 exhibits the cor-
responding curves for the electric field strength E, the polarization P and the
inversion D for A = Ae. These results may be compared with the numerical curves
259
248 8. Physical Systems
E.P D
2.0
to
10
0.6
o
Fig. 8.4. Electric fielil strength E, polarization P and inversion D for A = Ac.
(After H. Haken, H. Ohno: Opt. Commun. 16,205 (1976))
obtained by direct computer solutions of (8.94-96) for A = 11. Very good agree-
ment is found, for example, with respect to the electric field strength.
We now discuss the case Re a < o. The shapes of the corresponding potential
curves for f3 < 0 or f3 > 0 are qualitatively given by the solid lines of Figs. 6.6a
and 6.6b, respectively, i.e., we deal now with a second order phase transition and a
soft excitation. The shapes of the corresponding field strength, polarization and
inversion resemble those of Fig. 8.4.
(3 a
[y ] [¥ ]
0.001
0.04
-0.04
[em] L
Fig. 8.4a. The coefficients p, a of (8.107) as functions of laser length L. Left ordinate refers to p,
right to a. Chosen parameters: y = 2 X 10· S-I; K = O.ly; YII = 0.5y; A = 11.5. (After H.
Haken and H. Ohno, unpublished)
260
8.9 Instabilities in Fluid Dynamics: The Benard and Taylor Problems 249
The preceding treatment has shown that instability hierarchies may occur in
lasers. As we now know, such hierarchies are a typical and widespread pheno-
menon in synergetic systems. They are given detailed treatment in my book
Advanced Synergetics (see Footnote on page 216).
261
250 8. Physical Systems
cells. In a more refined theory one may then ask questions as to the probability of
fluctuations.
A closely related problem is the so-called Taylor-problem: The flow between a
long stationary outer cylinder and a concentric rotating inner cylinder takes place
along circular streamlines (Couette flow) if a suitable dimensionless measure of the
inner rotation speed (the Taylor number) is small enough. But Taylor vortices
spaced periodically in the axial direction appear when the Taylor number exceeds a
critical value.
In the following we shall explicitly treat the convection instability. The physical
quantities we have to deal with are the velocity field with components uj (j =
1,2, 3 <--> x, y, z) at space point x, y, z, the pressure p and the temperature T.
Before going into the mathematical details, we shall describe its spirit. The velocity
field, the pressure, and the temperature obey certain nonlinear equations of fluid
dynamics which may be brought into a form depending on the Rayleigh number
R which is a prescribed quantity. For small values of R, we solve by putting the
velocity components equal zero. The stability of this solution is proven by lineariz-
ing the total equations around the stationary values of u, p, T, where we obtain
damped waves. If, however, the Rayleigh number exceeds a certain critical value Ro
the solutions become unstable. The procedure is now rather similar to the one which
we have encountered in laser theory (compare Sect. 8.8). The solutions which
become unstable define a set of modes. We expand the actual field (u, T) into these
modes with unknown amplitudes. Taking now into account the nonlinearities of
the system, we obtain nonlinear equations for the mode amplitudes which quite
closely resemble those oflaser theory leading to certain stable-mode configurations.
Including thermal fluctuations we again end up with a problem defined by non-
linear deterministic forces and fluctuating forces quite in the spirit of Chapter 7.
Their interplay governs in particular the transition region, R :::0 RoO'
(8.113)
262
8.10 The Basic Equations 251
The index i = 1,2,3 corresponds to x, y, z and (Xl> X2, X3) = (x, y, z), where we
have used the convention of summing up over those expressions in which an
index, for instance j, occurs twice. The lhs represents the "substantial derivative",
where the first term may be interpreted as the local, temporal change in the
velocity field. The second term stems from the transport of matter to or from the
space point under consideration. The first term on rhs represents a force due to
the gradient of the pressure p, where (] is the density of the fluid and
(]o = (](x, y, z = 0). The second term describes an acceleration term due to gravity
(g is the gravitational acceleration constant). In this term the explicit dependence
of (] on the space point x is taken into account, and () is the Kronecker symbol.
The third term describes the friction force, where v is the kinematic viscosity.
The last term describes fluctuating forces which give rise to velocity fluctuations.
Near thermal equilibrium F(u) is connected with the fluctuating part of the stress
tensor s:
(8.114)
(8.114a)
-iJT
iJt
+ u·-iJT
J iJXj
= KAT+ F(T)
,
(8.115)
(8.116)
We assume that the heat currents h are Gaussian and possess the following
correlation functions:
(8.116a)
263
252 8. Physical Systems
Ui = 0 (i = 1,2,3) (8.117)
or at free boundaries
(8.117 a)
T= T(s) + e, (8.11Sd)
(8.l1Se)
264
8.11 The Introduction of New Variables 253
Having made these transformations, we drop the prime in u', because u' is identi-
cal with u. The continuity equation (S.112) remains unaltered:
divu = 0, (S.l12a)
a ~
OUi
~
ut
oUi _
+ Uj ~
uXj
- -
ow
~
uXi
+ rxgou3,i + V.dUi + F(u)
i, (S.l13 a)
(S.115a)
°
For the lower and upper surfaces z = and z = d respectively, we must now
require e = 0. For u the boundary conditions remain unaltered. In fluid dynam-
ics, dimensionless variables are often introduced. In accordance with conventional
procedures we therefore introduce the new scaled variables u', e', w', as well as x'
and t', by means of
d2
Xi = dxi, t = -t"
K '
(S.119)
For simplicity, we will drop the prime from now on. Then the continuity equation
(S.112) again reads
divu = 0, (S,112b)
~
OU i
~
ut
+ Uj;:)uXj
OUi _
- -
ow
~
uXi
+ pa P" F'(u)
OU3,i + aUi + i , (S.113b)
-Oe
ot
oe
+ u·-
J ox.
=.de + Ru + F'(T).
3
(S.l15b)
J
Obviously the only free dimensionless parameters occurring in these three equa-
tions are the Prandtl number
p = V/K (S.119a)
265
254 8. Physical Systems
R = cxgfJ d 4 .
(S.119b)
VK
In the context of this book we will consider the Rayleigh number to be a control
parameter which can be manipulated from the outside via {3, whereas the Prandtl
number will be assumed to be fixed.
The new fluctuating forces are connected with the former ones by
(S.1l4b)
(S.116b)
For sake of completeness we shall write down the relations between the new
fluctuating forces and the fluctuating stress tensor and the fluctuating heat cur-
rents. The corresponding formulae read
(S.114c)
(S.116c)
where s' and h' are fixed by the condition that their correlation functions remain
the same as in (S.114a) and (S.l16a) in the new variables x', t'.
Using matrix notation, the linearized equations can be written in rather concise
form:
(S.121)
266
8.12 Damped and Neutral Solutions (R ::;; R,l 255
0
0 0 0 -R-
ox
0
0 RP,1 0 0 -R-
oy
0
K= 0 0 RP,1 RP -R- (8.121 a)
oz
0 0 RP P,1 0
R.! R.! 0 0
oy oz
R 0 0 0 0
0 R 0 0 0
S= 0 0 R 0 0 (8.121 b)
0 0 0 P 0
0 0 0 0 0
(8.121c)
(8.121d)
(x,y)=x. (8.122)
(8.122 a)
267
256 8. Physical Systems
ik
- k{ In cos Inz
ik
- k i In cos Inz
In
- k2 [AI + P(k 2 + 12 n 2 )] cos Inz
(8.123c)
(8.124)
268
8.12 Damped and Neutral Solutions (R s Rcl 257
Incidentally, this equation shows the orthogonality of the q' s in the sense defined
by the scalar product. The solutions are stable as long as ..1.1 possesses a negative
real part. When we change the Rayleigh number R, ..1.1 also changes. For a given
wave number k we obtain the Rayleigh number at which ..1.1 vanishes, i.e. where
linear stability is lost. From (S.123 a) we obtain
(S.125)
(S.125 a)
In order to obtain the critical Rayleigh number and the corresponding critical
wave number k we seek the minimum of the neutral curve (S.125 a) according to
oR (S.125b)
O(k2) = O.
(S.125c)
(S.125d)
as critical Rayleigh number (under the assumption of free surfaces). For Rayleigh
numbers R close to Rc the eigenvalue of the unstable mode may be written in the
form
(S.125e)
The above analysis allows us to distinguish between the unstable and the stable
modes. Using these modes we will transform our original set of equations into a
new set of equations for the corresponding mode amplitudes. We further note that
the unstable modes are highly degenerate because linear stability analysis only
fixes the modulus of k, and the rolls described by the solutions (S.123) can be
oriented in all horizontal directions.
269
258 8. Physical Systems
(8.126)
where q was defined in (8.120), and ~k.l(t) are still unknown time-dependent
amplitudes. The sum runs over all k and l. We wish to derive equations for the
amplitudes ~ which fully contain the nonlinearities. To do this we must insert
(8.126) into the complete nonlinear equations. Therefore the equation we have to
treat reads
Sq = Kq + N(q), (S.I27)
(8.127a)
The matrices K and S were defined in (8.121 a, b), and qi(i = 1, ... ,5) are the
components of the vector q introduced in (8.120). We insert (8.126) into (8.127 a)
and use the fact that qk, leX, z) obeys (8.121 d). In order to obtain equations for the
~f s we multiply both sides of (S.127) by one of the eigenmodes qt, leX, z) and
integrate over the normalization volume. Because the corresponding calculations
are simple in principle but somewhat lengthy, we quote only the final result. We
obtain equations which have the following structurc: 1
The explicit expressions for the coefficients A are rather lengthy and will not
be presented here. We only note that one summation, e.g. over k" and I", can be
carried out from selection rules which lead to bk,k' + k" (due to conservation of wave
vectors) and a further combination of Kronecker symbols containing I, If, I". This
allows immediate evaluation of the sums over k", I".
In analogy with our treatment in Sect. 7.7 we must now consider the equations
for the unstable modes individually. Because explicit calculation and discussion
do not give us much insight into the physics, we will restrict ourselves to a verbal
1 We remind the reader that in the interest of simplicity we have already omitted a further index.
say m (m = 1,2) in the equation (8.128) corresponding to Am. see (8.125a)
270
8.13 Solution Near R = Rc (Nonlinear Domain). Effective Langevin Equations 259
description. In the first step the equations for the unstable mode amplitudes
~"c.l(t) are selected. (This corresponds to the amplitude ~".u of Sect. 7.7.) In the
lowest order of approximation in the nonlinearities we confine ourselves to the
terms ex ~"c.l'~".l.s in the equation of motion for the unstable modes ~"c.l' Here
~".l.s stands for the set of stable modes (see Sect. 7.7) which couple through the
nonlinearities and are restricted by selection rules.
Note that there is still a whole set of degenerate unstable modes which are
distinguished by different directions of the vector k c • These modes can be visuali-
zed as rolls whose axes are oriented in different horizontal directions. For the sake
of simplicity we shall consider only one definite direction in the discussion which
follows. This kind of model can be justified and related to reality if we choose a
rectangular boundary in the horizontal plane and assume that one edge is much
shorter than the other. In such a case the rolls whose axes are parallel to the
shorter edge become unstable first. We note, however, that it is by no means trivial
to extend this presentation to the general case of many directions and continuous
wave vectors, where finite bandwidth effects play an important role.
Once we have selected a specific direction k" the problem becomes one-
dimensional and we need only take into account the stable mode ~O.2' The equa-
tion for the unstable mode then reads
(8.129)
Here the coefficient of the linear term has been taken from (8.125e). In the second
step we must consider the equation of ~O.2' Here we shall restrict our attention to
terms ex I~"c.112 in the nonlinearities, i.e. to the same order of approximations as
before. We then obtain
(8.130)
We now invoke the principle of adiabatic elimination (i.e. a specific form of the
slaving principle). Accordingly we put
~O.2 = 0 (8.131)
(8.132)
which allows us to eliminate ~O.2 in the equations for the unstable modes which
now become the order parameters. The final equation reads
(8.133)
271
260 8. Physical Systems
For the sake of completeness, we will briefly discuss four important generaliza-
tions: (1) three dimensions, (2) finite bandwidth effects, (3) fluctuations, and (4)
symmetry breaking.
1. Three Dimensions. Here we have to consider all directions of the unstable
modes and, in addition, their coupling to further classes of stable modes which
were not taken into account in (8.129). As a consequence of these facts, (8.133) is
now generalized to
(8.133a)
i.e. the different directions kc are coupled through the coupling constants {3kc,k~'
where Al(k;,R - Rc) was derived in (8.125e). If required, it is, of course, possible
to include higher-order terms. In analogy with our procedure in Sect. 7.7 we make
use of the relation
(8.135)
where
(8.135a)
(8.136)
which occurs in the order parameter equations, so that (8.136) replaces ill in
(8.129). We note that the form (8.136) preserves the rotation symmetry of the
original equations. In the case where R > Rc and a definite roll pattern has
developed, e.g. along the x direction, an alternative form of (8.136) is more appro-
priate, namely
ill
4P
= Al(k;,R - Rc) - 1 + P 8x -
(8 i 8)2
J'in
2
8y2 . (8.136a)
272
8.13 Solution Near R = R, (Nonlinear Domain). Effective Langevin Equations 261
This form can be derived by taking into account the first nonvanishing bandwidth
corrections in the x and y directions.
(8.137)
(8.138)
where
(8.139)
4. Symmetry Breaking. Finally, we note that additional terms may occur if the
equations of motion of the fluid do not possess inversion symmetry with respect
to the central horizontal plane. This symmetry violation may be caused, for
instance, by a more complicated dependence of the density e on temperature than
assumed at the beginning of Sect. 8.10, or by surface tensions or heating effects. A
discussion would be rather lengthy, but the net effect can be written down in a
simple manner. It turns out that these effects can be taken into account by
additional terms on the rhs of (8.133a), namely terms which are bilinear in the
amplitudes ~kc,l:
(8.140)
(8.141)
(8.141 a)
273
262 8. Physical Systems
Exercise on 8.13
The stationary solution of the Benard problem for R < Rc yields a linear tempera-
ture profile:
1'5) = To - [3z.
In practice it may happen that the heat which is put into the system from below
cannot be completely carried away. As a consequence the temperature at the
boundaries increases at a constant rate 1'/. This may be taken into account by
changing the boundary conditions for the temperature in the following way:
Boundary conditions of
T= To - [3z + I'/r[~:
determine the change in temperature for R < Rc in this case. Show that the
symmetry of the linear temperature profile is lost, and use the elimination proce-
dure described in Chap. 7 to show that these boundary conditions give rise to a
term of the form (8.140).
(8.142)
f= %expl'J>. (8.143)
I'J> f
= ~ f{Lk C y~t.l (:x - fo :;2y ~kc.l
+ Lk c (XI ~kc.112 - t Lk~k~k~' (c5~t.l a~.l ~:~:1 (5,.~+k~+k~~O + c.c.)
- i Lkck~ [3kck~1 ~kc.1121 ~k~.l f} dx dy. (8.144)
It goes far beyond our present purpose to treat (8.144) in its whole generality. We
want to demonstrate, how such expressions (8.143) and (8.144) allow us to discuss
274
8.14 The Fokker-Planck Equation and Its Stationary Solution 263
the threshold region and the stability of various mode configurations. We neglect
the dependence of the slowly varying amplitudes ~"c.l on x, y. We first put J = O.
(8.143) and (8.144) are a suitable means for the discussion of the stability of
different mode configurations. Because qJ depends only on the absolute values of
~"c.l
(8.145)
(8.146)
The values w"c for which IfJ has an extremum are given by
oqJ
-:;,-
uW = 0, or Ct - L", p" ",w,,' = 0 (8.147)
"c
c c c c
and the second derivative tells us that the extrema are all maxima
(8.148)
(8.149)
(8.150)
We now compare the solution in which all modes participate, with the single mode
solution for which
(8.151)
holds, so that
1 Ct 2
lfJ(w) ="2 If' (8.152)
A comparison between (8.150) and (8.152) reveals that the single mode has a
greater probability than the multimode configuration. Our analysis can be gener-
275
264 8. Physical Systems
alized to different mode configurations, leading again to the result that only a
single mode is stable.
Let us discuss the form of the velocity field of such a single mode configuration,
using (8.121) to (8.125). Choosing kc in the x direction, we immediately recognize
that for example the z-component of the velocity field, U z is independent of y, and
has the form of a sine wave. Thus we obtain rolls as stable configurations.
We now come to the question how to explain the still more spectacular
hexagons. To do this we include the cubic terms in (8.144) which stem from a
spatial inhomogeneity in z-direction e.g., from non-Boussinesq terms. Admitting
for the comparison only 3 modes with amplitudes ~;, ~r, i = 1,2,3, the potential
function is given by
l/> = !X(I ~112 + I~212 + I~312) - b(~r~1~~ + c.c.) - U:::kek~ Pked ~ke.1121~k~.112
(8.153)
')- -----<
D )---
,
I ,
---< , I
\ I
\
, .
!
Fig. 8.5. Construction of hexagon from basic triangle (for
" details consult the text)
where the kc-sums run over the triangle of Fig. 8.5 which arises from the condition
kc1 + ke2 + ke3 = 0 and Iked = const. (compare (8.127». To find the extremal
values of ([J we take the derivatives of (8.153) with respect to ~i' ~r and thus obtain
six equations. Their solution is given by
(8.154a)
Using (8.154a) together with (8.121)-(8.125) we obtain for example uAx). Concen-
trating our attention on its dependence on x, y, and using Fig. 8.5, we find (with
x' = n/J2x)
(8.1S4b)
276
8.14 The Fokker-Planck Equation and Its Stationary Solution 265
(8.155)
(8.156)
(8.157)
(8.158)
(8.159)
(8.160)
2)
(8.161)
(8.162)
277
266 8. Physical Systems
.rE)
2 Readers not familiar with the basic features of semiconductors may nevertheless read this
chapter starting from (8.163)-(8.\65). This chapter then provides an example how one may
derive pulse-like phenomena from certain types of nonlinear equations.
278
8.15 A Model for the Statistical Dynamics of the Gunn Instability Near Threshold 267
time t and space point x by n and the electron current, divided by the electronic
charge e, by J, the continuity equation reads
on oj
-+-=0
ot ox . (8.163)
There are two contributions to the current J. On the one hand there is the streaming
motion of the electrons, nv(E), on the other hand this motion is superimposed by a
diffusion with diffusion constant D. Thus we write the "current" in the form
(8.164)
dE
dx = e'en - no), (8.165)
, 4ne
e =-. (8.165a)
eo
(8.166)
(8.167)
This equation immediately allows for an integration over the coordinate. The
integration constant denoted by
1
- let) (8.167a)
e
can still be a function of time. After integration and a slight rearrangement we find
279
268 8. Physical Systems
oE dE d 2E 4n
- = -e'nov(E) - v(E)- + D -2 + - let). (8.168)
ot dx dx eo
let), which has the meaning of a current density, must be determined in such a way
that the externally applied potential
u= 1: dx E(x, t) (8.169)
is constant over the whole sample. It is now our task to solve (8.168). For explicit
calculations it is advantageous to use an explicit form for veE) which, as very
often used in this context, has the form
III is the electron mobility of the lower band, B is the ratio between upper and
lower band mobility. To solve (8.168) we expand it into a Fourier series
(8.171)
The summation goes over all positive and negative integers, and the fundamental
wave number is defined by
(8.172)
Expanding veE) into a power series of E and comparing then the coefficients of the
same exponential functions in (8.168) yields the following set of equations
(8.173)
The different terms on the rhs of (8.173) have the following meaning
(8.174)
where
(8.175)
(8.176)
280
8.15 A Model for the Statistical Dynamics of the Gunn Instability Near Threshold 269
(8.177)
Eo = U/L. (8.179)
It is now our task to discuss and solve at least approximately the basic (8.173).
First neglecting the nonlinear terms we perform a linear stability analysis. We
observe that Em becomes unstable if IXm > 0. This can be the case if vb1 ) is negative
which is a situation certainly realized as we learn by a glance at Fig. 8.6. We in-
vestigate a situation in which only one mode m = ± 1 becomes unstable but the
other modes are still stable. To exhibit the essential features we confine our analysis
to the case of only 2 modes, with m = ± 1, ±2, and make the hypothesis
(8.180)
(8.181)
and
(8.182)
In realistic cases the second term in (8.184) can be neglected compared to the first
so that (8.184) reduces to (8.185). Formula (8.186) implies a similar approximation.
We now can apply the adiabatic elimination procedure described in Sections 7.1, 7.2.
Putting
(8.187)
281
270 8. Physical Systems
_ V ci (8.18S)
C z - 2" f3 + Wlc1l z,
(S.IS9)
OC I _ oCP
at - - oci'
With the abbreviation I = IClIZ, cP reads
(S.190)
When the parameter a is changed, we obtain a set of potential curves, CP(J) which
are similar to those of Fig. 8.3. Adding a fluctuating force to (S.189) we may take
into account fluctuations. By standard procedures a Fokker-Planck equation be-
longing to (S.IS9) may be established and the stable and unstable points may be
discussed. We leave it as an exercise to the reader to verify that we have again a
situation of a first order phase transition implying a discontinuous jump of the
equilibrium position and a hysteresis effect. Since C I and C z are connected with oscil-
latory terms (compare (S.171), (S.lSO)) the electric field strength shows undamped
oscillations as observed in the Gunn effect.
282
8.16 Elastic Stability: Outline of Some Basic Ideas 271
load
!
load
V (q) -
-
1 -
2 ( k
R - -R-
2 cos Ct cos q
)2 (8.191)
Since we confine our analysis to small angles q we may expand the cosine functions,
which yields
(8.192)
-P£(q). (8.195)
The potential of the total system comprising springs and load is given by the sum
of the potential energies (8.192) and (8 .195)
V(q) = V - P£ (8.196)
and thus
(8.197)
Evidently the behavior of the system is described by a simple potential. The system
283
272 8. Physical Systems
(8.198)
q = q(P). (8.199)
P = P(q). (8.200)
(8.201)
tells us if the potential has a maximum or a minimum. Note that we have to insert
(8.199) into (8.201). Thus when (8.201) changes sign from a positive to a negative
value we reach a critical load and the system breaks down. We leave the following
problem to the reader as exercise: Discuss the potential (8.197) as a function of the
parameter P and show that at a critical P the system suddenly switches from a
stable state to a different stable state. Hint: The resulting potential curves have the
form of Fig. 6.7 i.e., a first order transition occurs.
Second-order transitions can also be very easily mimicked in mechanical en-
gineering by the hinged cantilever (Fig. 8.9). We consider a rigid link of length I,
load
(8.202)
284
8.16 Elastic Stability: Outline of Some Basic Ideas 273
Just as before we can easily construct the total energy which is now
When we expand the cosine function for small values of q up to the fourth power
we obtain potential curves of Fig. 6.6. We leave the discussion of the resulting
instability which corresponds to a second order phase transition to the reader.
Note that such a local instability is introduced if PL > k. Incidentally this example
may serve as an illustration for the unfoldings introduced in Section 5.5. In practice
the equilibrium position of the spring may differ slightly from that of Fig. 8.9.
Denoting the equilibrium angle of the link without load by 8, the potential energy
of the link is now given by
Expanding the cosine function again up to 4th order we observe that the resulting
potential
(8.207)
is of the form (5.133) including now a linear term which we came across when
discussing unfoldings. Correspondingly the symmetry inherent in (8.204) is now
broken giving rise to potential curves of Fig. 5.20 depending on the sign of 8.
We leave it again as an exercise to the reader to determine the equilibrium positions
and the states of stability and instability as a function of load.
285
274 8. Physical Systems
or a limit point may be reached where the system completely loses its stability.
An important consequence of our general considerations of Section 5.5 should be
mentioned for the characterization of the instability points. It suffices to consider
only as many degrees of freedom as coefficients of the diagonal quadratic form
vanish. The coordinates of the new minima may describe completely different
mechanical configurations. As an example we mention a result obtained for thin
shells. When the point of bifurcation is reached, the shells are deformed in such a
way that a hexagonal pattern occurs. The occurrence of this pattern is a typical
post-buckling phenomenon. Exactly the same patterns are observed for example in
hydrodynamics (compare Sect. 8.13).
286
9. Chemical and Biochemical Systems
kl
A + X~2X. (9.1)
k~
276 9. Chemical and Biochemical Systems
- x x
A x Fig. 9.1. (Compare text)
The corresponding reaction rates are denoted by k 1 and k;, respectively. We further
assume that the molecule X may be converted into a molecule C by interaction with
a molecule B (Fig. 9.2)
k2
B+X~C. (9.2)
k~
Again the inverse process is admitted. The reaction rates are denoted by k2 and k~.
We denote the concentrations of the different molecules A, X, B, C as follows:
A a
X n
(9.3)
B b
C c
We assume that the concentrations of molecules A, Band C and the reaction rates
k j , kj are externally kept fixed. Therefore, what we want to study is the temporal
behavior and steady state of the concentration n. To derive an appropriate equa-
tion for n, we investigate the production rate of n. We explain this by an example.
The other cases can be treated similarly.
Let us consider the process (9.1) in the direction from left to right. The number
of molecules X produced per second is proportional to the concentration a of mole-
cules A, and to that of the molecules X, n. The proportionality factor is just the
reaction rate, k l ' Thus the corresponding production rate is a· n· k l' The complete
list for the processes 1 and 2 in the directions indicated by the arrows reads
(9.4)
The minus signs indicate the decrease of concentration n. Taking the two processes
of 1 or 2 together, we find the corresponding rates '1 and '2 as indicated above.
The total temporal variation of n, dn/dt == Ii is given by the sum of '. and '2 so
that our basic equation reads
(9.5)
288
9.2 Deterministic Processes, Without Diffusion, One Variable 277
In view of the rather numerous constants a, ... ,kl' ... appearing in (9.4) it is
advantageous to introduce new variables. By an appropriate change of units of
time t and concentration n we may put
(9.6)
(9.7)
i.e., for f3 > 1 we find no molecules of the kind X whereas for f3 < 1 a finite con-
centration, n, is maintained. This transition from "no molecules" to "molecules X
present" as f3 changes has a strong resemblance to a phase transition (cf. Section
(6.7» which we elucidate by drawing an analogy with the equation of the fer-
romagnet writing (9.8) with Ii = 0 in the form
y = n2 - (I - f3)n. (9.10)
H +--+ Y
TITc +--+ f3
H = M2 - (1 - ~) M, (9.11)
av
Ii = - an (9.12)
289
278 9. Chemical and Biochemical Systems
with
(9.l3)
We have met this type of potential at several occasions in our book and we may
leave the discussion of the equilibrium positions of n to the reader. To study the
temporal behavior we first investigate the case
a) y = O.
The equation
t = 0; n = no. (9.l5)
(9.l6)
i.e., n tends asymptotically to O. We now assume [3 of. 1. The solution of (9.14) with
(9.15) reads
A = 11 - {JI, (9.18)
11 - [31 + (1 - [3) - 2no
c = -:-:----';;-;----:-.:---------'~-=--" (9.19)
11 - [31 + (1 - [3) + 2n o'
In particular we find that the solution (9.17) tends for t --> 00 to the following
equilibrium values
0 for [3 > 1
{ (9.20)
noo = (1 - [3) for [3 < 1 .
b) ')' of. O.
In this case the solution reads
290
9.2 Deterministic Processes, Without Diffusion, One Variable 279
n
l-fl - - - - - - - - -
Fig. 9.3. Solution (9.17) for P > 1 Fig. 9.4. Solution (9.17) for P < 1 and
two different initial conditions
(9.23)
k2
B + X<2 c. (9.24)
k~
Eq. (9.23) implies a trimolecular process. Usually it is assumed that these are very
rare and practically only bimolecular processes take place. It is, however, possible
to obtain a trimolecular process from subsequent bimolecular processes, e.g.,
A + X -> Y; Y + X -> 3X, if the intermediate step takes place very quickly and
the concentration of the intermediate product can (mathematically) be eliminated
adiabatically (cf. Sect. 7.1). The rate equation of (9.23), (9.24) reads
Ii = _n 3 + 3n 2 - pn + y == <pen), (9.25)
291
280 9. Chemical and Biochemical Systems
describes a first-order phase transition (cf. Sect. 6.7). This analogy can be still
more closely exhibited by comparing the steady state equation (9.25) with Ii = 0,
with the equation of a van der Waals gas. The van der Waals equation of a real gas
reads
(9.26)
We leave it to readers who are familiar with van der Waals' equation, to exploit
this analogy to discuss the kinds of phase transition the molecular concentration n
may undergo.
a2 n
Ii = x ax2 + cp(n). (9.28)
cp(n) = an
a cP(n) (= _ aaVn)' (9.29)
n ~ n1 for z ~ + 00,
n ~ n2 for z ~ - 00. (9.30)
(9.31)
292
9.3 Reaction and Diffusion Equations 281
we invoke an analogy with an oscillator or, more generally, with a particle in the
potential field cP(n) by means of the following correspondence:
x <-> t time
cP <-> potential (9.32)
n <-> q coordinate
Note that the spatial coordinate x is now interpreted quite formally as time while
the concentration variable is now interpreted as coordinate q of a particle. The
potential cP is plotted in Fig. 9.5. We now ask under which condition is it possible
that the particle has two equilibrium positions so that when starting from one equi-
librium position for t = - 00 it will end at another equilibrium position for t -->
+ oo? From mechanics it is clear that we can meet these conditions only if the
potential heights at ql (== n 1) and q2 (== n2 ) are equal
(9.33)
Bringing (9.33) into another form and using (9.29), we find the condition
(9.34)
- 'P(n)
(9.36)
293
282 9. Chemical and Biochemical Systems
In Fig. 9.6 we have plotted l' versus n. Apparently the equilibrium condition implies
T
vdn)
n,
L---~----------~-----+n Fig. 9.6. Maxwellian construction of the
coexistence value y
that the areas in Fig. 9.6 are equal to each other. This is exactly Maxwell's con-
struction which is clearly revealed by the comparison exhibited in (9.27)
t (9.38)
This example clearly shows the fruitfulness of comparing quite different systems
in and apart from thermal equilibrium.
In this section we first consider the following reaction scheme of the "Brusselator"
A-+X
B+X-+Y+D
2X + Y -+ 3X
X-+E (9.39)
294
9.4 Reaction-Diffusion Model with Two or Three Variables: Brusselator, Oregonator 283
(9.40)
onl 2 o2nl
at = a - (b + l)nl + n ln2 + DI OX2' (9.41)
~2 ~~
at =
Z
bn l - n1n Z + Dz oxz , (9.42)
Eqs. (9.41) and (9.42) may, of course, be formulated in two or three dimensions.
One easily verifies that the stationary state of (9.41), (9.42) is given by
n? = a, (9.45)
To check whether new kinds of solutions occur, i.e., if new spatial or temporal
structures arise, we perform a stability analysis of the (9.41) and (9.42). To this end
we put
(9.46)
and linearize (9.41) and (9.42) with respect to ql' q2. The linearized equations are
(9.47)
(9.48)
295
284 9. Chemical and Biochemical Systems
(9.49)
whereas (9.44) requires qj finite for x -> ± 00. Putting, as everywhere in this book,
q = (:J, (9.50)
4 = Lq, (9.51)
(9.52)
with
Inserting (9.53) into (9.51) yields a set of homogeneous linear algebraic equations
for qo. They allow for nonvanishing solutions only if the determinant vanishes.
-D; +b- 1- A
I -b
a
-D~
2
- a2 - A
I=0 '
A = AI' (9.54)
(9. 54a)
A2 - IXA + f3 = 0, (9.55)
IX = (- D; + b - 1 - D'z - a 2 ), (9.56)
and
(9.57)
296
9.4 Reaction-Diffusion Model with Two or Three Variables: Brusselator, Oregonator 285
An instability occurs if Re (A) > o. We have in mind keeping a fixed but changing
the concentration b and looking for which b = be the solution (9.53) becomes
unstable. The solution of (9.55) reads, of course,
(9.58)
a> O. (9.62)
We skip the transformation of the inequalities (9.59), (9.60), (9.61) and (9.62) to
the corresponding quantities a, b, D;, D;, and simply quote the final result: We
find the following instability regions:
(9.63)
The left inequality stems from (9.62), the right inequality from (9.6\). The instabili-
ties occur for such a wave number first for which the smallest, b, fulfils the in-
equalities (9.63) or (9.64) for the first time. Apparently a complex A is associated
with a hard mode excitation while A real is associated with a soft mode. Since
instability (9.63) occurs for k -# 0 and real A a static spatially inhomogeneous
pattern arises. We can now apply procedures described in Sections 7.6 to 7.11.
We present the final results for two different boundary conditions. For the boundary
conditions (9.43) we put
(9.65)
297
286 9. Chemical and Biochemical Systems
where the index u refers to "unstable" in accordance with the notation of Section
7.7. The sum over j contains the stable modes which are eliminated adiabatically
leaving us, in the soft-mode case, with
(9.66)
(9.66a)
D,2
C3 = -[D' _ a2(D,2C+ 1 _ D' )f {(D~c + l)(D;c + a 2)}
2c 1c 2c
(9.66c)
where Ie is the critical value of I for which instability occurs first. A plot of (u
as a function of the parameter b is given in Fig. 5.4 (with b == k and (u == q).
Apparently at b = bc a point of bifurcation occurs and a spatially periodic structure
is established (compare Fig. 9.7). If on the other hand I is odd, the equation
for ~u reads
298
9.4 Reaction-Diffusion Model with Two or Three Variables: BrusseIator, Oregonator 287
~ ..
--+-----=::......_-- - - _ b
analogy to Section 5.1. So far we have considered only instabilities connected with
the soft mode. If there are no finite boundaries we make the following hypothesis
for q
(9.68)
The methods described in Section 7.8 allow us to derive the following equations
for eN,ke == e
a) Soft mode
(9.69)
where
(9.69d)
b) Hard mode
(9.70)
299
288 9. Chemical and Biochemical Systems
e
Note that A can become negative. In that case higher powers of must be included.
With increasing concentrations b, still more complicated temporal and spatial
structures can be expected as has been revealed by computer calculations.
The above equations may serve as model for a number of biochemical reactions
as well as a way to understand, at least qualitatively, the Belousov-Zhabotinski
reactions where both temporal and spatial oscillations have been observed. It
should be noted that these latter reactions are, however, not stationary but occur
rather as a long-lasting, transitional state after the reagents have been put together.
A few other solutions of equations similar to (9.41) and (9.42) have also been con-
sidered. Thus in two dimensions with polar coordinates in a configuration in
which a soft and a hard mode occur simultaneously, oscillating ring patterns are
found.
Let us now come to a second model, which was devised to describe some es-
sential features of the Belousov-Zhabotinski reaction. To give an idea of the
chemistry of that process we represent the following reaction scheme:
Steps (CI) and (C4) are assumed to be bimolecular processes involving oxygen
atom transfer and accompanied by rapid proton transfers; the HOBr so produced
is rapidly consumed directly or indirectly with bromination of malonic acid. Step
(C3a) is ratedetermining for the overall process of (C3a) + 2(C3b). The Ce 4 +
produced in step (C3b) is consumed in step (C5) by oxidation of bromomalonic
acid and other organic species with production of the bromide ion. The complete
chemical mechanism is considerably more complicated, but this simplified version
is sufficient to explain the oscillatory behavior of the system.
Computational Model
The significant kinetic features of the chemical mechanism can be simulated by
the model called the "Oregonator".
A+Y-+X
X+Y-+P
B + X-+2X+ Z
2X-+ Q
Z -+fY.
300
9.5 Stochastic Model for a Chemical Reaction Without Diffusion. Birth and Death Processes 289
Exercise
Verify, that the rate equations belonging to the above scheme are (in suitable units)
kl
A+X~2X (9.71)
kf
k2
B+X~C, (9.72)
k~
301
290 9. Chemical and Biochemical Systems
N N N N
Fig. 9.10. How P(N) changes Fig. 9.11. How P(N) changes due to death of
due to birth of a molecule a molecule
In a similar way we may discuss the first process in (9.72) with rate k z . Here the
number of molecules N is decreased by J ("death" of a molecule X. cf. Fig. 9.11).
If we start from the level N the rate is proportional to the probability to find that
state N occupied times the concentration of molecules b times the number of
molecules X present times the reaction rate k z . Again the proportionality factor
must be fixed later. It is indicated in the formula written below. By the same
process, however, the occupation number to level N is increased by processes start-
ing from the level N + 1. For this process we find as transition rate
N
2-7 peN + 1, t)(N + I)bk z - peN, t) vbkz V. (9.74)
It is now rather obvious how to derive the transition rates belonging to the pro-
cesses 1. We then find the scheme
N-I N
1 -> peN - I, t)V· -v-ak) - peN, t)· vak) V (9.75)
and
(N + l)N N(N - 1) ,
1<- peN + I, t)V VZ k; - peN, t) v2 k) V. (9.76)
302
9.5 Stochastic Model for a Chemical Reaction Without Diffusion. Birth and DeathProcesses 291
The rates given in (9.73)-(9.76) now occur in the master equation because they
determine the total transition rates per unit time. When we write the master equa-
tion in the general form
we find for the processes (9.71) and (9.72) the following transition probabilities per
second:
. (N - 1) ,)
weN, N - 1) = V ( ak t V + k 2c , (9.78)
weN, N + 1) = V k t
(
I (N +
v2
I)N
+ k2
beN +
V
1») . (9.79)
The scheme (9.78) and (9.79) has an intrinsic difficulty, namely that a stationary
solution of (9.77) is just P(O) = 1, peN) = 0 for N", O. For a related problem
compare Section 10.2. For this reason it is appropriate to include the spontaneous
creation of molecules X, from molecules A with the transition rate kl in (9.71) and
(9.72) as a third process
k,
A--X. (9.80)
Thus the transition rates for the processes (9.71), (9.72) and (9.80) are
(9.81)
and
weN, N + 1) = V kl
(
I (N +
v2
l)N
+ k 2b
(N +
V
1)) . (9.82)
The solution of the master equation (9.77) with the transition rates (9.78) and
(9.79) or (9.81) and (9.82) can be easily found, at least in the stationary state,
using the methods of Chapter 4. The result reads (compare (4.119»
w(v + 1, v)
peN) = P(O)· TI v=o
N-t
(9.83)
w
(
v, v + 1)'
The further discussion is very simple and can be performed as in 4.6. It turns out
that there is either an extremal value at N = 0 or at N = No '" 0 depending on the
parameter b.
In conclusion we must fix the proportionality constants which have been left
open in deriving (9.73) to (9.76). These constants may be easily found if we require
303
292 9. Chemical and Biochemical Systems
that the master equation leads to the same equation of motion for the density, n,
which we have introduced in Section 9.2, at least in the case of large numbers N.
To achieve this we derive a mean value equation for N by mUltiplying (9.77) by N
and summing up over N. After trivial manipulations we find
d
(it <N) = <weN + 1, N» - <weN - 1, N» (9.84)
where, as usual,
and
d
(it<N) {+ I)I
+ 2c - l l
= V ak l -y<N I» -} k k'l V 2 <N(N - k 2 b-y<N) .
(9.85)
304
9.5 Stochastic Model for a Chemical Reaction Without Diffusion. Birth and Death Processes 293
and
Dividing (9.92) by (9.93) and using the explicit forms (9.88)-(9.91) we find the
relation
ak J k;c
(9.94)
k'JN/V = k 2 bN/V·
Apparently both sides of (9.94) are of the form /1IN where /1 is a certain constant.
(9.94) is equivalent to the law of mass action. (According to this law,
In our case, the numerator is n·n·c, the denominator a·n·b·n). Using that (9.94) is
equal /11 N we readily verify that
/1
w(N, N - I) = N w(N - I, N) (9.95)
/1 N
P(N) = P(O) N! . (9.96)
(9.97)
Thus in the present case we, in fact, find the Poisson distribution
(9.98)
Exercises on 9.5
I) Derive the transition rates for the master equation (9.77) for the following
305
294 9. Chemical and Biochemical Systems
processes
kl
A + 2X+2 3X (E. I)
k~
k2
B+X+2C. (E.2)
k~
kJ
Aj + IjX+2 B j + (lj + I)X,j = 1, ... , k,
k~
under the requirement of detailed balance for each reaction). Show that peN) is
Poissonian.
Hint: Use wiN, N - 1) in the form
for finding the cells 1 occupied by N, molecules. In this chapter we consider only a
single kind of molecules but the whole formalism can be readily generalized to
several kinds. The number of molecules N, now changes due to two causes; namely,
due to chemical reactions as before, but now also on account of diffusion. We
describe the diffusion again as a birth and death scheme where one molecule is
annihilated in one cell and is created again in the neighboring cell. To find the total
change of P due to that process, we have to sum up over all neighboring cells
1 + a of the cell under consideration and we have to sum up over all cells 1. The
306
9.6 Stochastic Model for a Chemical Reaction with Diffusion. One Variable 295
where we may insert for PI reaction the rhs of (9.77) or any other reaction scheme.
For a nonlinear reaction scheme, (9.101) cannot be solved exactly. We therefore
employ another method, namely, we derive equations of motion for mean values
or correlation functions. Having in mind that we go from the discrete cells to a
continuum we shall replace the discrete index I by the continuous coordinate x,
I -+ x. Correspondingly we introduce a new stochastic variable, namely, the local
particle density
N,
p(x) = - (9.102)
v
1
n(x, t) = -v (N,)
1
= -;; LIN N,P( ... , N" ...).
j } (9.103)
We further introduce a correlation function for the densities at space points x and
x'
g(x, x', t) = ~
v
(N,N,,) - (N,)(N,,)
v
~- b(x - x')n(x, t), (9.104)
As a concrete example we now take the reaction scheme (9.71, 72). We assume,
however, that the back reaction can be neglected, i.e., k; = O. Multiplying the
corresponding equation (9.101) by (9.102) and taking the average (9.103) on both
sides we obtain
an(x, t) 2
-a-(- = DV n(x, t) + (Xl - x 2 )n(x, t) + x1P, (9.106)
307
296 9. Chemical and Biochemical Systems
D = D'lv. (9.107)
(9.107a)
(9.107b)
A comparison of (9.106) with (9.28), (9.5), (9.4) reveals that we have obtained
exactly the same equation as in the nonstochastic treatment. Furthermore we find
that putting k; = 0 amounts to neglecting the nonlinear term of (9.8). This is the
deeper reason why (9.106) and also (9.108) can be solved exactly. The steady state
solution of (9.108) reads
(9.109)
(9.110)
Apparently the correlation function drops off with increasing distance between x
and x'. The range of correlation is given by the inverse of the factor of Ix - xii
of the exponential. Thus the correlation length is
(9.111)
When we consider the effective reaction rates X\ and X 2 (which are proportional
to the concentrations ofmolecuIes a and b) we find that for Xl = X2 the coherence
length becomes infinite. This is quite analogous to what happens in phase transi-
tions of systems in thermal equilibrium. Indeed, we have already put the chemical
reaction models under consideration in parallel with systems undergoing a phase
transition (compare Sect. 9.2). We simply mention that one can also derive an
equation for temporal correlations. It turns out that at the transition point also the
correlation time becomes infinite. The whole process is very similar to the non-
equilibrium phase transition of the continuous-mode laser. We now study molecule
number fluctuations in small volumes and their correlation function. To this end
we integrate the stochastic density (9.102) over a volume A V where we assume
308
9.6 Stochastic Model for a Chemical Reaction with Diffusion. One Variable 297
(9.112)
It is a simple matter to calculate the variance of the stochastic variable (9.112) which
is defined as usual by
(9.113)
Using the definition (9.104) and the abbreviation Rile =r We obtain after ele-
mentary integrations
(9.114)
(9.11 5)
i.e., an ever increasing variance with increasing distance. Keeping Ie finite and
letting R -+ 00 the variance becomes proportional to the square of the correlation
length
(9.116)
For volumes with diameter small compared to the correlation length R « Ie' we
obtain
(9.117)
(9.118)
This result shows that for R -+ 00 the variance becomes independent of the distance
and would obtain from a master equation neglecting diffusion. It has been proposed
to measure such critical fluctuations by fluorescence spectroscopy which should be
much more efficient than light-scattering measurements. The divergences, which
occur at the transition point" 1 = "2' are rounded off if we take the nonlinear term
309
298 9. Chemical and Biochemical Systems
ex: n 2 , i.e., k; # 0, into account. We shall treat such a case in the next chapter
taking a still more sophisticated model.
Exercises on 9.6
= 0.
1) Derive (9.106) from (9.101) with (9.100), (9.77, 78, 79) for k't
Hint: Use exercise 3) of Section 9.5. Note that in our present dimensionless
units of space
f .= J { 0 ( D d 2p
dx - op(x) dx2(x»)
f + D (d 0)2 (p(x)f)·
dx' op(x)
Hint:
Divide the total volume into cells which still contain a number N, » 1. It is
assumed that P changes only little for neighboring cells. Expand the rhs of
(9.100) into a power series of "1" up to second order. Introduce p(x) (9.102),
and pass to the limit that I becomes a continuous variable x, hereby replacing
P by f = J{p(x)} and using the variational derivative %p(x) instead of iJ/iJN,.
(For its use see HAKEN, Quantum Field Theory of Solids, North-Holland,
Amsterdam, 1976).
A-+X
B+X-+Y+D (9.119)
2X + Y -+ 3X
X-+E,
where the concentrations of the molecules of kind A, B are externally given and
310
9.7 Stochastic Treatment of the Brusselator Close to Its Soft-Mode Instability 299
kept fixed, while the numbers of molecules of kind X and Yare assumed to be
variable. They are denoted by M, N respectively. Because we want to take into
account diffusion, we divide the space in which the chemical reaction takes place
into cells which still contain a large number of molecules (compared to unity).
We distinguish the cells by an index I and denote the numbers of molecules in cell
I by M" N,. We again introduce dimensionless constants a, b which are propor-
tional to the concentrations of the molecules of kind A, B. Extending the results of
Sections 9.5 and 6, we obtain the following master equation for the probability
distribution P( ... , M" N, ... ) which gives us the joint probability to find
M", N", ... , M" N" ... molecules in cells I', ... , I
In it v is the volume of a cell, I. The first sum takes into account the chemical reac-
tions, the second sum, containing the "diffusion constants" D;, D~, takes into ac-
count the diffusion of the two kinds of molecules. The sum over a runs over the
nearest neighboring cells of the cell I. If the numbers M" N, are sufficiently large
compared to unity and if the function P is slowly varying with respect to its argu-
ments we may proceed to the Fokker-Planck equation. A detailed analysis shows
that this transformation is justified within a weB-defined region around the soft-
mode instability. This implies in particular a» 1 and Jl == (DdD2)1/2 < 1.
To obtain the Fokker-Planck equation, we expand expressions of the type (M, + 1)
PC . .. ,M, + 1, N" ... ) etc. into a power series with respect to "1" keeping the
first three terms (cf. Sect 4.2). Furthermore we let I become a continuous index
which may be interpreted as the space coordinate x. This requires that we replace
the usual derivative by a variational derivative. Incidentally, we replace M,lv, N,lv
by the densities M(x), N(x) which we had denoted by p(x) in (9.102) and PC . .. , M"
N" ... ) by f( . .. , M(x), N(x), ... ). Since the detailed mathematics of this proce-
dure is rather lengthy we just quote the final result
311
300 9. Chemical and Biochemical Systems
The indices M(x) or N(x) indicate the variational derivative with respect to M(x)
or N(x). DI and D2 are the usual diffusion constants. The Fokker-Planck equation
(9.121) is still far too complicated to allow for an explicit solution. We therefore
proceed in several steps: We first use the results of the stability analysis of the
corresponding rate equations without fluctuations (cf. Section 9.4). According to
these considerations there exist stable spatially homogeneous and time-independent
solutions M(x) = a, N(x) = bja provided b < be. We therefore introduce new
variables qj(x) by
f .= J
dx [- {( ql + a q2
(b- 1)2 + g(ql' q2) + DI V' 2ql)f}ql(X)
- {( -bql - a 2q2 - g(ql, q2) + D/i/2q2)f}q2(1t)
+ -H1>l1(q)f}Ql(1t),Ql(1t) - {Ddq)f}Ql(1tJ,Q2(1t)
+ -HD22(q)f}Q2(1tJ,Q2(1tj + DI(V'(8j8ql(x»)2(a + ql)f
+ D2(V(8j8q2(X»)2(bja + q2)f]. (9.122)
fis now a functional of the variables qix). We have used the following abbrevia-
tions:
+ bqUa + qiq2'
g(ql' q2) = 2aqlq2 (9.123)
1>11 = 2a + 2ab + (3b + l)ql + a2q2 + 2aqlq2 + (bja)qi + qiQ2' (9.124)
1>12 = D22 = 2ab + 3bql + bqUa + a2q2 + 2aqlq2 + q;Q2' (9.125)
where
(9.127)
312
9.7 Stochastic Treatment of the Brusselator Close to Its Soft-Mode Instability 301
with
We have further
and
For a definition of I and J, see (7.81) and (7.83). For a discussion of the evolving
spatial structures, the "selection rules" inherent in I and J are important. One
readily verifies in one dimension:
1= 0 for boundary conditions (9.44), i.e., the Xk'S are plane waves and I ::::: 0 for
X" oc sin kx and k » 1.
Further
J"k'"''k''' = J'*0 only if two pairs of k's out of k, k', k", k'" satisfy:
kl = -k z = -ke
k3 = -k4 = -ke
if plane waves are used, or k = k' = k" = k'" = ke if Xk oc sin kx. We have evalu-
ated ii 1 explicitly for plane waves. The Fokker-Planck equation (9.126) then reduces
to
(9.132)
Note that for sufficiently big aJl the coefficient A becomes negative. A closer
inspection shows, that under this condition the mode with k = 0 approaches a
marginal situation which then requires to consider the modes with k = 0 (and
Ikl = 2kJ as unstable modes. We have met eqs. like (9.131) or the corresponding
Langevin equations at several instances in our book. It shows that at its soft-mode
instability the chemical reaction undergoes a nonequilibrium phase transition of
second order in complete analogy to the laser or the Benard instability (Chap. 8).
313
302 9. Chemical and Biochemical Systems
In Sections 9.2-9.4 we have met several explicit examples for equations describing
chemical processes. If we may assume spatial homogeneity, these equations have
the form
(9.133)
Equations of such a type occur also in quite different disciplines, where we have
now in mind network theory dealing with electrical networks. Here the n's have the
meaning of charges, currents, or voltages. Electronic devices, such as radios or
computers, contain networks. A network is composed of single elements (e.g.,
resistors, tunnel diodes, transistors) each of which can perform a certain function.
It can for example amplify a current or rectify it. Furthermore, certain devices can
act as memory or perform logical steps, such as "and", "or", "no". In view of the
formal analogy between a system of equations (9.133) of chemical reactions and
those of electrical networks, the question arises whether we can devise logical
elements by means of chemical reactions. In network theory and related disciplines
it is shown that for a given logical process a set of equations of the type (9.133) can
be constructed with well-defined functions Fj •
These rather abstract considerations can be easily explained by looking at our
standard example of the overdamped anharmonic oscillator whose equation was
given by
4 = rxq - fl q 3. (9.134)
In electronics this equation could describe e.g. the charge q of a tunnel diode of the
device of Fig. 7.3. We have seen in previous chapters that (9.134) allows for two
stable states q 1 = -J rx.1fl, q2 = - -J rxl fl, i.e., it describes a bistable element which
can store information. Furthermore we have discussed in Section 7.3 how we can
switch this element, e.g., by changing rx.. When we want to translate this device into
a chemical reaction, we have to bear in mind that the concentration variable, n,
is intrinsically nonnegative. However, we can easily pass from the variable q to a
positive variable n by making the replacement
(9.135)
so that both stable states lie at positive values. Introducing (9.135) into (9.134) and
rearranging this equation, we end up with
(9.136)
(9.137)
314
9.8 Chemical Networks 303
Since (9.134) allowed for a bistable state for q so does (9.136) for n. The next ques-
tion is whether (9.136) can be realized by chemical reactions. Indeed in the preceed-
ing chapters we have met reaction schemes giving rise to the first three terms in
(9.136). The last term can be realized by an adiabatic elimination process of a fast
chemical reaction with a quickly transformed intermediate state. The steps for
modeling a logical system are now rather obvious: 1) Look at the corresponding
logical elements of an electrical network and their corresponding differential
equations; 2) Translate them in analogy to the above example. There are two
main problems. One, which can be solved after some inspection, is that the im-
portant operation points must lie at positive values of n. The second problem is,
of course, one of chemistry; namely, how to find chemical processes in reality
which fulfil all the requirements with respect to the directions the processes go,
reaction constants etc.
Once the single elements are realized by chemical reactions, a whole network can
be constructed. We simply mention a typical network which consists of the follow-
ing elements: flip-flop (that is the above bistable element which can be switched),
delays (which act as memory) and the logical elements "and", "or", "no".
Our above considerations referred to spatially homogeneous reactions, but by
dividing space into cells and permitting diffusion, we can now construct coupled
logical networks. There are, of course, a number of further extensions possible,
for example, one can imagine cells separated by membranes which may be only
partly permeable for some of the reactants, or whose permeability can be switched.
Obviously, these problems readily lead to basic questions of biology.
315
10. Applications to Biology
where - {3n 2 stems from a depletion of the food resources. It is assumed that new
food is supplied only at a constant rate. The behavior of a system described by
(10.5) was discussed in detail in Section 5.4.
We now come to several species. Several basically different cases may occur:
1) Competition and coexistence
2) Predator-prey relation
3) Symbiosis
(10.6)
Things become much more complicated if different species live or try to live on the
same food supply, and/or depend on similar living conditions. Examples are pro-
vided by plants extracting phosphorous from soil, one plant depriving the other
from sunlight by its leaves, birds using the same holes to build their nests, etc.
Since the basic mathematical approach remains unaltered in these other cases,
we talk explicitly only about "food". We have discussed this case previously
(Sect. 5.4) and have shown that only one species survives which is defined as the
fittest. Here we exclude the (unstable) case that, accidentally., all growth and
decay rates coincide.
For a population to survive it is therefore vital to improve its specific rates
('J.j, {3j by adaption. Furthermore for a possible coexistence, additional food supply
is essential. Let us consider as example two species living on two "overlapping"
food supplies. This can be modelled as follows, denoting the amount of available
318
10.1 Ecology, Population-Dynamics 307
food by N1 or N 2 :
Here YjNJ is the rate of food production, and - YjNj is the decrease of food due
to internal causes (e.g. by rotting). Adopting the adiabatic elimination hypothesis
(cf. Sect. 7.1), we assume that the temporal change of the food supply may be
neglected, i.e., N 1 = N2 = O. This allow us to express N1 and N2 directly by n 1
and n2' Inserting the resulting expressions into (10.7), (10.8) leaves us with equa-
tions of the following type:
From li1 = 1i2 = 0 we obtain stationary states n~, n~. By means of a discussion
of the "forces" (i.e., the rhs of (10.11/12)) in the n 1 - n 2 plane, one may easily
discuss when coexistence is possible depending on the parameters of the system.
(cf. Fig. 10.1 a-c) This example can be immediately generalized to several kinds
of species and of food supply. A detailed discussion of coexistence becomes tedious,
however.
From our above considerations it becomes apparent why ecological "niches"
Fig. 1O.1a-c. The eqs. (10.11), (10.12) for different parameters leading to different stable configu-
rations. (a) nI = 0, n2 = C is the only stable point i.e., only one species survives. (b) nI = 0, n2
# 0 or n2 = 0, nI # 0 are two stable points, i.e., one species or the other one can survive. (c) nI
# 0, n2 # 0, the two species can coexist. If the field of arrows is made closer, one finds the trajec-
tories discussed in Section 5.2. and the points where the arrows end are sinks in the sense of that
chapter
319
308 10. Applications to Biology
are so important for survival and why surviving species are sometimes so highly
specialized. A well-known example for coexistence and competition is the distribu-
tion of flora according to different heights in mountainous regions. There, well-
defined belts of different kinds of plants are present. A detailed study of such
phenomena is performed in biogeographics.
2) Predator-Prey-Relation
The basic phenomenon is as follows:
There are two kinds of animals: Prey animals living on plants, and predators,
living on the prey. Examples are fishes in the Adriatic sea, or hares and lynxes.
The latter system has been studied in detail in nature and the theoretical predic-
tions have been substantiated. The basic Lotka-Volterra equations have been
discussed in Section 5.4. They read
(10.13)
(10.14)
where (10.13) refers to prey and (10.14) to predators. As was shown in Section 5.4
a periodic solution results: When predators become too numerous, the prey is eaten
up too quickly. Thus the food supply of the predators decreases and consequently
their population decreases. This allows for an increase of the number of prey
animals so that a greater food supply becomes available for the predators whose
number now increases again. When this problem is treated stochastically, a serious
difficulty arises: Both populations die out (cf. Sect. 10.2).
3) Symbiosis
There are numerous examples in nature where the cooperation of different species
facilitates their living. A well-known example is the cooperation of trees and bees.
This cooperation may be modelled in this way: Since the multiplication rate of one
species depends on the presence of the other, we obtain
a) nl = n z = 0, which is uninteresting,
or
b) IXI - ° + lX;nz =
1 0,
CX z - 0z + cx;nl = O.
320
10.2 Stochastic Models for a Predator-Prey System 309
It is an interesting exercise for the reader to discuss the stability properties of b).
We also leave it to the reader to convince himself that for initial values of nl and
n 2 which are large enough, an exponential explosion of the populations always
occurs.
The analogy of our rate equations stated above to those of chemical reaction
kinetics is obvious. Readers who want to treat for example (10.5) or (10.11), (10.12)
stochastically are therefore referred to those chapters. Here we treat as another
example the Lotka-Volterra system. Denoting the number of individuals of
the two species, prey and predator, by M and N, respectively, and again using the
methods of chemical reaction kinetics, we obtain as transition rates
1) Multiplication of prey
M -+ M + 1: w(M + 1, N; M, N) = 'X.1M.
2) Death rate of predator
N -+ N - 1: w(M, N - I; M, N) = 'X.2N.
M -+ M -
N -+
I}
N + 1 w(M - 1, N + 1; M, N) = (3MN.
The master equation for the probability distribution P(M, N, t) thus reads
321
310 10. Applications to Biology
Thus both species die out, even if initially both have been present. For a proof we
put P = O. Inserting (10.18) into (10.17) shows that (10.17) is indeed fulfilled.
Furthermore we may convince ourselves that all points (M, N) are connected via
at least one path with any other point (M', N'). Thus the solution is unique. Our
rather puzzling result (10.18) has a simple explanation: From the stability analysis
of the non stochastic Lotka-Volterra equations, it is known that the trajectories
have "neutral stability". Fluctuations will cause transitions from one trajectory to
a neighboring one. Once, by chance, the prey has died out, there is no hope for the
predators to survive, i.e., M = N = 0 is the only possible stationary state.
While this may indeed happen in nature, biologists have found another reason
for the survival of prey: Prey animals may find a refuge so that a certain minimum
number survives. For instance they wander to other regions where the predators
do not follow so quickly or they can hide in certain places not accessible to pre-
dators.
(10.19)
The properties of Fit) depend on both the population which was present prior to
the one described by the particular equation (10.19) and on environmental factors.
322
10.4 A Model for Morphogenesis 311
(10.20)
If the mutation rate for a special mutant is small, only that mutant survives which
has the highest gain factor 0(j and the smallest loss factor xj and is thus the "fittest".
It is possible to discuss still more complicated equations, in which the mUltiplica-
tion of a subspecies is replaced by a cycle A -+ B -+ C -+ .•. -+ A. Such cycles
have been postulated for the evolution of biomolecules. In the context of our book
it is remarkable that the occurrence of a new species due to mutation ("fluctu~ting
force") and selection ("driving force") can be put into close parallel to a second-
order, nonequilibrium phase transition (e.g., that of the laser).
323
312 10. Applications to Biology
3) d. . a2 a
IffusIOn: D a ax 2 , (10.23)
Da diffusion constant.
Furthermore it is known from other biological systems (e.g., slime mold, compare
Section 1.1) that autocatalytic processes ("stimulated emission") can take place.
They can be described-depending on the process-by the production rate
(10.24)
or
(10.25)
Finally, the effect of inhibition has to be modelled. The most direct way the
inhibitor can inhibit the action of the activator is by lowering the concentration of a.
A possible "ansatz" for the inhibition rate could be
-ah. (10.26)
Another way is to let h hinder the autocatalytic rates (10.24) or (10.25). The
higher h, the lower the production rates (10.21) or (10.25). This leads us in the case
(10.25) to
(10.27)
Apparently there is some arbitrariness in deriving the basic equations and a final
decision can only be made by detailed chemical analysis. However, selecting
typical terms, such as (10.21), (10.22), (10.23), (10.27), we obtain for the total rate
of change of a
(10.28)
Let us now turn to derive an equation for the inhibitor h. It certainly has a decay
time, i.e., a loss rate
-vh, (10.29)
(10.30)
324
IDA A Model for Morphogenesis 313
Again we may think of various generation processes. Gierer and Meinhard, whose
equations we present here, suggested (among other equations)
(10.32)
Before we represent our analytical results using the order parameter concept in
Section 10.5, we exhibit some computer solutions whose results are not restricted to
the hydra, but may be applied also to other phenomena of morphogenesis. We
simply exhibit two typical results: In Fig. 10.2 the interplay between activator and
inhibitor leads to a growing periodic structure. Fig. 10.3 shows a resulting two-
dimensional pattern of activator concentration. Obviously, in both cases the
inhibitor suppressed a second center (second head of hydra!) close to a first center
(primary head of hydra !). To derive such patterns it is essential that h diffuses more
easily than a, i.e, Dh > Da' With somewhat further developed activator-inhibitor
models, the structures of leaves, for example, can be mimicked.
325
314 10. Applications to Biology
Fig. 10.3. Results of the morphogenetic model. Left column: activator concentration plottet over
two dimensions. Right column: same for inhibitor. Rows refer to different times growing from
above to below (computer solution). (After H. Meinlwrdr. A. GiereI': J. Cell Sci. 15. 321 (1974»
, l jv k
x=Vo: x, (=l't , a'= - a,
c
(10.33)
326
10.5 Order Parameters and Morphogenesis 315
Then we have
'2
.,
a =p 11-J.l"+
,+a A' ,
a LJ 1I, (10.34)
, pc , J1.
p = vk' J1. =-, (10.36)
v
D,=Dh (10.37)
Da'
From now on we shall drop the primes. The stationary homogeneous solution of
(10.34) and (10.35) reads
1
ao = - (p+1), (10.38)
J1.
ho=aij. (10.39)
(10.40)
Eqs. (10.34) and (10.35) then can be cast in the form (compare (7.62))
where K is given by
J1.(_2_-1)+L1
p+1
K(L1) = (10.42)
2
-(p+1) -1 +DL1
J1.
and g(q) contains the nonlinearities. For the linear stability analysis we drop the
nonlinear term g(q) and make the hypothesis
327
316 10. Applications to Biology
where
2 2/1
a(k)= -(D+1)k +~-/1-1, (10.45)
p+1
(10.46)
(10.47)
(10.49)
(10.50)
328
10.5 Order Parameters and Morphogenesis 317
analysis we will focus our attention on the soft mode case. kc and Pc and Pmax
are given by
(10.51 )
(10.52)
11- 1
Pmax = 11+ 1· (10.53)
In order that the soft-mode instability occurs first we have to require Pc> Pm,x'
from which it follows
Using (10.37) and (10.36) we derive from (10.54) that the diffusion constant of the
inhibitor must be bigger than that of the activator by a certain amount. In other
words, "long range inhibition"" and "short range activation" are required for a
non oscillating pattern.
We assume a two-dimensional layer with side lengths L\ and L2 and first
adopt periodic boundary conditions. The detailed method of solution has been
described in Sections 7.6-7.8 and we repeat here only the main steps of the whole
procedure. We assume P close to Pc.
We make the hypothesis (cf. (7.72»
(10.55)
To exhibit the essential features, we neglect "small band excitations" and assume
~{ independent of x. The coefficients oj obey the equation
(10.56)
(10.57)
(10.58)
329
318 10. Applications to Biology
Inserting (10.55) into (10.41) and mUltiplying the resulting expressions from the
left with the conjugate complex of eikx and the adjoint of oj, we obtain after some
analysis the equations
The nonlinear term on the right hand side has the form
_
Jkk'k"k'''---
. ..
1
L1L2
12d
F
xe
i(k'+k"+k'''-k)x_ s:
-Ukk'+k"+k''''
•
(10.62)
We eliminate the stable modes as in Section 7.7. The great advantage of the
"slaving principle" consists in an enormous reduction of the degrees of freedom
because we keep now only the unstable modes with index k = k c ' These modes
serve as everywhere in this book as order parameters. Their cooperation or
competition determines the resulting patterns as we will show now. We introduce
a new notation by which we replace the vector kc by its modulus and the angle
cp which this vector forms with a fixed axis: ~kc -->~kc.<p' We let cp run from 0 to 7L
The resulting order parameter equations read
(10.63)
(10.64)
330
10.5 Order Parameters and Morphogenesis 319
0.00f----~--~~---+--.
In
··· ...
Fig. 10.5. d(9) is plotted as the function of 9.
In the practical calculations the region of divergence
:: l: is cut out as indicated by the bars. This procedure
-0.89 : can be justified by the buildup of wave packets
We may assume as elsewhere in this book (cf. especially Sect. 8.13) that the
eventually resulting pattern is determined by such a configuration of ~'s where the
potential V acquires a (local) minimum. Thus we have to seek such ~'s for which
(10.66)
and
(10.67)
or if the matrix
(10.69)
(10.70)
where
(10.71)
331
320 10. Applications to Biology
The angle <PI between the roll axis and a fixed axis is arbitrary, i.e., symmetry
breaking with respect to <P occurs. This configuration is locally stable for
The resulting spatial pattern can be obtained by inserting (10.69, 70) into (10.55).
In our present treatment and in the corresponding figures we neglect the impact
of the slaved modes. They would give rise to a slight sharpening of the individual
peaks. The corresponding pattern is exhibited in Fig. 10.6.
(10.73)
XI is given by
(10.75)
(10.76)
The bifurcation diagram of the solution (10.75) is shown in Fig. 10.7. The solid
line indicates the stable configuration, the dashed line the unstable configuration.
The order parameter equations allow us to determine not only the stationary
332
10.5 Order Parameters and Morphogenesis 321
I
\
" .... .....
C2
F Fig. 10.7. The amplitude Xl as a function of A.. The solid
line indicates a stable solution, the dotted line an
unstable solution
solution but also the transient. We start from a homogeneous solution on which
a small inhomogeneity of the form (10.73, 74) is superimposed and solve the
time-dependent equations (10.63). The resulting solutions of the spatial pattern are
exhibited in Figs. 10.8, 10.9, and 10.10. This shows again the usefulness of order
parameter equations.
In our next example we present the solution of the nonlinear equations within
a rectangular domain with nonflux boundary conditions (close to the instability
333
322 10. Applications to Biology
point). We now expand the wanted solution into a complete orthogonal system
with nonflux boundary conditions, i.e., with respect to functions of the form
coskxx'coskyY, where ( kk X
., , ) ~ ~(~L mn-zl)
-" ~_ . (10.77)
The procedure goes through in complete analogy to the above. k" and kl' must
be chosen so that k; +k; comes close to k;. When L t ~ L z , different 'modes
may simultaneously become unstable ("degeneracy"). For simplicity, we treat
here the case of a single unstable mode (i.e., L J =1= L z ). Its amplitude obeys the
equation
(10.78)
The solution of this time-dependent equation describes the growth of the spatial
pattern. Examples of the resulting patterns are given by Figs. 10.11 to 10.13.
r
Fig. 10.12. The activator concentration
belonging to the mode (10.77) with kx = 21[/ L 1
and ky=51[/ L 2
334
r
10.5 Order Parameters and Morphogenesis 323
(10.79)
This equation fixes a series of k-values for which (10.79) is fulfilled. The expansion
of q now reads
(10.80)
Inserting (10.80) into the original equation (10.41) leads eventually to equations
for the ~'s. The slaving principle allows us to do away the stable modes and we
thus obtain equations for the order parameter alone, since we have a discrete
sequence of k-values and for symmetry reasons we may assume that only one
mode becomes unstable first. The resulting order parameter equation has again
the form (10.78).
A marked difference occurs depending on whether the unstable mode (order
parameter) has m=mc=O or =1=0. If mc=O the right hand side of equation
(10.78) must be supplemented by a term quadratic in ~ which is absent if mc=l=O.
As is well known from phase transition theory (cf. Sect. 6.7), in the former case
(mc=O) we obtain a first-order phase transition connected with an abrupt change
of the homogeneous state into the inhomogeneous state connected with hysteresis
effects. In the latter case (mc =1= 0) we obtain a second-order phase transition,
and the pattern grows continuously out of the homogeneous state when p passes
through Pc' Some typical patterns are exhibited in Figs. 10.14 to 10.16.
In conclusion we mention that we can also take into account small band excitations
and fluctuations in complete analogy to the Benard instability (cf. Sects. 8.13
and 8.14) using the methods developed in this book. A comparison between the
figures of this section and those of Section 10.3 shows a qualitative resemblance
but no exact agreement. The reason for this lies in the fact that the analytical
335
324 10. Applications to Biology
336
10.6 Some Comments on Models of Morphogenesis 325
approach yields "pure cases", whereas the computer solution makes use of
(artificially) introduced random fluctuations. As we know, in that latter case
the analytical approach would yield a probability distribution (of patterns)
of a whole ensemble, whereas a computer solution will be equivalent to a "simple
event", i. e., a specific realization.
337
326 10. Applications to Biology
and its connection with the formation of, for instance, chemical patterns in the
brain.
Let us conclude with the following remark. In this book we have stressed the
profound analogies between quite different systems, and one is tempted to treat
biological systems in complete analogy to physical or chemical systems far from
thermal equilibrium. One important difference should be stressed, however.
While the physical and chemical systems under consideration lose their structurc
when the flux of energy and matter is switched off, much of the structure of bio-
logical systems is still preserved for an appreciable time. Thus biological systems
seem rather to combine nondissipative and dissipative structures. Furthermore,
biological systems serve certain purposes or tasks and it will be more appropriate
to consider them as functional structures. Future research will have to develop
adequate methods to cope with such functional structures. It can be hoped.
however, that the ideas and methods outlined in this book may serve as a first
step in that direction.
338
11. Sociology and Economics
(1Ll)
We are interested in the probability distribution function f(n+, n_, t). One may
easily derive the following master equation
df[n+, n; t]
dt = (n+ + l)p+ _[n+ + 1, n_ - l]f[n+ + 1, n_ - 1; t]
+ (n_ + l)p_ +[n+ - 1, n_ + l]f[n+ - 1, n_ + 1; t]
- {n+p+ _[n+, n_] + n_p_ +[n+, n_]}f[n+, n_; t]. (11.2)
328 11. Sociology and Economics
The crux of the present problem is, of course, not so much the solution of this
equation which can be done by standard methods but the determination of the
transition probability. Similar to problems in physics, where not too much is known
about the individual interaction, one may now introduce plausibility arguments to
derive p. One possibility is the following: Assume that the rate of change of the
opinion of an individual is enhanced by the group of individuals with an opposite
opinion and diminished by people of his own opinion. Assume furthermore that
there is some sort of social overall climate which facilitates the change of opinion
or makes it more difficult to form. Finally one can think of external influences on
each individual, for example, informations from abroad etc. It is not too difficult
to cast these assumptions into a mathematical form, if we think of the Ising model
of the ferromagnet. Identifying the spin direction with the opinion +, -, we are
led to put in analogy to the Ising model
+(/q + H)}
p_+[n+, n-J == p-+(q) = vexp { e
= v exp {+(kq + h)}, (I1.3)
For a quantitative treatment of (11.2) we assume the social groups big enough so
that q may be treated as a continuous parameter. Transforming (11.2) to this
continuous variable and putting
-1 { Jq
Kt(Y)d}
fslq) = cK2 (q) exp 2 _+ Kiy) Y (11.6)
with
340
11.2 Phase Transitions in Economics 329
(a) k=O
h=O
Fig. 11.1 shows a plot of the result when there is no external parameter. As one
may expect from a direct knowledge of the Ising model, there are typically two
results. The one corresponds to the high-temperature limit: on account of rather
frequent changes of opinion we find a centered distribution of opinions. If the social
climate factor e is lowered or if the coupling strength between individuals is in-
creased, two pronounced groups of opinions occur which clearly describe the
by now well-known "polarization phenomenon" of society. It should be noted that
the present model allows us to explain, at least in a qualitative manner, further
processes, for example unstable situations where the social climate parameter is
changed to a critical value. Here suddenly large groups of a certain opinion are
formed which are dissolved only slowly and it remains uncertain which group
(+ or -) finally wins. Using the considerations of Section 6.7 it is obvious again
that here concepts of phase transition theory become important, like critical
slowing down (Remember the duration of the 1968 French student revolution?),
critical fluctuations, etc. Such statistical descriptions certainly do not allow unique
predictions due to the stochastic nature of the process described. Nevertheless,
such models are certainly most valuable for understanding general features of
cooperative behavior, even that of human beings, though the behavior of an
individual may be extremely complicated and not accessible to a mathematical
description. Quite obviously the present model allows for a series of generalizations.
341
330 II. Sociology and Economics
Innovations. We have seen again and again that there are quite different regimes
with respect to the behavior of a variety of systems. On the one hand there is a
region where a lamp or a fluid layer behaves normally. In such a case their
behavior does not change qualitatively if the perturbations are not too great. On
the other hand there are particularly interesting regions where a system becomes
unstable and tends to acquire a new state, i.e. where circumstances have become
favorable for transition into a new state. When and how this transition occurs is
quite often determined by fluctuations. Precisely this behavior can also be found
in economic models. But what plays the role of fluctuations, i.e. of triggers, in
economics? One group of events which plays that role are innovations, especially
those which are based on inventions - the gasoline engine, the airplane, the
telephone, or even a new vacuum cleaner. A large group of less evident but very
important inventions are those which simplify production.
According to empirical innovation research, the initial phase starts with fun-
damental innovations which create new industrial branches, such as the invention
of the automobile. If several such fundamental innovations occur simultaneously,
they are usually followed by innovations aimed at improving production in the
new industrial branches. The growth of these branches influences other branches
of the economy so that the general economic situation gives rise to general pros-
perity. This happens in different ways, e.g. via a high level of employment and
high purchasing power.
Economic studies have further revealed that innovations aimed at the produc-
tion of new products exceeded by far the introduction of new kinds of production
processes in the European industrial countries in the late 1940s and 1950s. Then
in the sixties a shift of innovations occurred which resulted in production changes.
This shift can be characterized by the catchword "rationalization".
Considerations of profit are unquestionably a major motivation in economic
actions, and discussions of such questions are seldom free from emotion, for
example when car drivers talk of the increase in gasoline prices and the profits
gained thereby. Let us not be governed here by emotions, but instead keep in mind
342
11.2 Phase Transitions in Economics 331
that decreasing gains will ultimately result in losses and a possible drop in employ-
ment. Let us consider the economic aspects only.
To make a profit requires the sale of a sufficient number of products. Higher
salaries will diminish a company's profit and affect prices, possibly making com-
petition difficult. At the same time, an increase in production is often connected
with the introduction of new products, which, at least in the beginning, can be
expensive. Both effects, namely increased salaries and the avoidance of high initial
costs for new products, are reasons to make investments which will lead to ratio-
nalization rather than to an increase in sales. In other words, companies prefer
innovations which will improve the production process to those which will intro-
duce new kinds of products: an automobile company would rather introduce new
welding automata than a new car.
Based on empirical data, Mensch has developed a mathematical model which
is borrowed from catastrophe theory. It describes the observed transition from full
employment to underemployment. Here we will present his model in the frame-
work of synergetics and extend it correspondingly. To this end we introduce the
following quantities and relations. We start with a mean annual production X 0
(measured, for example, in dollars) and we study the behavior of the deviation X
from it, so that the actual production can be represented as X 0 + X. Furthermore,
we denote by I the annual additional investments which result in expanded
production. The annual change in X will then be given by
(11.8)
R·X. (11.9)
Thus R is beneficial for the production per article. Finally, it is known from
economics (and from many examples in the present book) that a saturation point
is reached if the deviations in X become too large. This saturation can be modeled
by the term
(11.10)
Putting the individual terms from (11.8) to (11.10) together, we obtain the funda-
mental equation
(11.11)
We are well acquainted with this kind of equation. The potential belonging to its
rhs can be interpreted as synergetic costs. Companies will try to minimize that
potential as far as possible; i.e. they will arrange the production in such a way that
V reaches its minimum. If R = 0, V has only one minimum. This is in accordance
with the basic idea of Adam Smith, the father of the theory of the free market.
According to that theory the market will always equilibrate to a single stable
343
332 II. Sociology and Economics
344
12. Chaos
\ A N\ AlIIi Po d .
If YV vv~ Fig. 12.1. Example of chaotic motion of a variable q
(versus time)
(12.1)
and similarly the temperature deviation field. We then insert these expressions
into the Navier-Stokes equations (in the Boussinesq approximation) and keep
only the three terms
After some analysis and using properly scaled variables we obtain the Lorenz
equations
X=O"Y-O"X, (12.3)
Y=-XZ+rX - Y, (12.4)
Z=X·Y-bZ. (12.5)
0"= vjx' is the Prandtl number (where v is the kinematic viscosity, x' the thermo-
metric conductivity), r=R/R c (where R is the Rayleigh number, Rc the critical
Rayleigh number), b=4n 2 /(n 2 +ki). When we put
t=b(r-()-(1] . (12.6)
346
12.2 The Lorenz Model. Motivation and Realization 335
strength E, the polarization P, and the inversion D (in properly chosen units).
We assume single mode operation by putting jJ E/ jJ x = O. In the following
we assume that E and P are real quantities, i.e., that their phases can be kept
constant, which can be proved by computer calculations. The thus resulting
equations read
E=xP-xE, (12.7)
F=yED-yP, (12.8)
Jj =YII(A+l)-YIID- YIiAEP. (12.9)
These equations are identical with those of the Lorenz model in the form (12.6)
which can be realized by the following identifications
Table 12.1
Eqs. (12.6) describe at least two instabilities which have been found in-
dependently in fluid dynamics and in lasers. For A<O (r<l) there is no laser
action (the fluid is at rest), for A~O (r~ 1) laser action (convective motion starts)
with stable, time-independent solutions ~,1'/,( occurs. As we will see in Section 12.3,
besides this well known instability a new one occurs provided
347
336 12. Chaos
that the two-level atoms used in lasers are mathematically equivalent to spins.
Therefore, such phenomena may also be obtainable in spin systems coupled to
an electromagnetic field.
q2=-fJq2+q!q~, (12.13)
q~ = d~ - q~ - q! q 2 . (12.14)
1
t=-t'·
b '
(12.15)
A linear stability analysis reveals that the stationary solution becomes unstable
for
d~ = rx 2 fJ . rx + 3 ~~ . (12.17)
rx-fJ-1
348
12.3 How Chaos Occurs 337
stem from the coherent interaction between atoms and field. As is known in
laser physics coherent interaction allows for two conservation laws, namely
energy conservation and conservation of the total length of the so-called pseudo
spin. What is of relevance for us in the following is this. The conservation laws are
equivalent to two constants of motion
pZ=q~+(q3-1)Z, (12.19)
where
(12.20)
stem from the coupling of the laser system to reservoirs and describe damping
and pumping terms, i.e., nonconservative interactions. When we first ignore
these terms in (12.12-14) the point (Q!,QZ,Q3) must move in such a way that the
349
338 12. Chaos
conservation laws (12.18) and (12.19) are obeyed. Since (12.18) describes a sphere
in Ql,Q2,Q3 space and (12.19) a cylinder, the representative point must move
on the cross section of sphere and cylinder. There are two possibilities depending
on the relative size of the diameters of sphere and cylinder. In Fig. 12.3 we have
two well-separated trajectories, whereas in Fig. 12.4 the representative point can
Fig. 12.3. The cross section of the sphere (12.18) and the Fig. 12.4. The case R < 1 + p yields
cylinder (12.19) is drawn for R> 1 + p. One obtains two a single closed trajectory
separated trajectories
move from one region of space continuously to the other region. When we include
damping and pump terms the conservation laws (12.18) and (12.19) are no longer
valid. The radii of sphere and cylinder start a "breathing" motion. When this
motion takes place, apparently both situations of Fig. 12.3 and 12.4 are accessible.
In the situation of Fig. 12.3 the representative point circles in one region of space,
while it may jump to the other region when the situation of Fig. 12.4 becomes
realized. The jump of the representative point depends very sensitively on where
it is when the jump condition is fulfilled. This explains at least intuitively the
origin of the seemingly random jumps and thus of the random motion. This
interpretation is fully substantiated by movies produced in my institute. As we
will show below, the radii of cylinder and sphere cannot grow infinitely but are
bounded. This implies that the trajectories must lie in a finite region of space.
The shape of such a region has been determined by a computer calculation and is
shown and explained in Fig. 12.5.
350
12.3 How Chaos Occurs 339
When we let start the representative point with an initial value outside of
this region, after a while it enters this region and will never leave it. In other
words, the representative point is attracted to that region. Therefore the region
itself is called attractor. We have encountered other examples of attracting regions
in earlier sections. Fig. 5.11a shows a stable focus to which all trajectories con-
verge. Similarly we have seen that trajectories can converge to a limit cycle. The
Lorenz attractor has a very strange property. When we pick a trajectory and follow
the representative point on its further way, it is as if we stick a needle through a
ball of yarn. It finds its path without hitting (or asymptotically approaching)
this trajectory. Because of this property the Lorenz attractor is called a strange
attract or.
There are other examples of strange attractors available in the mathematical
literature. We just mention for curiosity that some of these strange attractors
can be described by means of the so-called Cantor set. This set can be obtained
as follows (compare Fig. 12.6). Take a strip and cut out the middle third of it.
Fig. 12.6. Representation of a Cantor set. The whole area is first divided into three equal parts. The
right and left intervals are closed, whereas the interval in the middle is open. We remove the open
interval and succeed in dividing the two residual intervals again into three parts, and again remove
the open intervals in the middle. We proceed by again dividing the remaining closed intervals into
three parts, removing the open interval in the center, etc. If the length of the original area is one, the
Lebesgues measure of the set of all open intervals is given by I:" [ 2" - 1/3" = 1. The remaining set of
closed intervals (i.e., the Cantor set) therefore has the measure zero
Then cut out the middle third of each of the resulting parts and continue in this
way ad infinitum. Unfortunately, it is far beyond the scope of this book to present
more details here. We rather return to an analytical estimate of the size of the
Lorenz attractor.
351
340 12. Chaos
The above considerations using the conservation laws (12.18) and (12.19) suggest
introduction of the new variables R,p, and q3. The original equations (12.12-14)
transform into
~(R2)" = - 11.+ fJ - a. R2 + (a. - fJ)p2 + (2(a- fi)+ do)q3 - (1- fJ)q~ , (12.21)
The equation for q3 fixes allowed regions because the expressions under the roots
in (12.23) must be positive. This yields
(12.24)
(12.25)
(a single closed trajectory). Better estimates for Rand p can be found as follows.
We first solve (12.22) formally
(12.26)
with
(12.27)
We find an upper bound for the right hand side of (12.26) by replacing g by its
maximal value
A2
gmax(q3) = 4B· (12.2X)
This yields
(12.29)
(12.30)
352
12.4 Chaos and the Failure of the Slaving Principle 341
In an analogous way we may treat (12.21) which after some algebra yields the
estimate
1
R~( ex) ) = 4exf3(1- m{ exd~ + (ex - f3) [2(2f3 -1)d o + 1 + 4f3(ex-l)] } . (12.31)
P = 32.84
R=33.63
PL = 39.73
RL = 40.73 .
o
Fig. 12.7. Graphical representa tion o f estimates (12.30),
(12.31). Central figure: Lorenz auractor (compare F ig. 12.5 ).
Outer solid circle: estimated maximal radius of cylinder.
PM (1 2. 30). Dashed circle : Projection of cylinder (12.29)
with radius Pc constructed from the condition that at a
L---------~-------q2
o certain time q , = q2= 0 a nd q ,= d o
353
342 12. Chaos
E
,
;1
"
""
:!
., ,,
354
12.5 Correlation Function and Frequency Distribution 343
E
,
I
violated. Furthermore, it turns out that at that time the representative point
just jumps from one region to the other one in the sense discussed in Section 12.3.
Thus we see that chaotic motion occurs when the slaving principle fails and the
formerly stable mode can no longer be slaved but is destabilized.
The foregoing sections have given us at least an intuitive insight into what chaotic
motion looks like. We now want to find a more rigorous description of its pro-
perties. To this end we use the correlation function between the variable q(t)
at a time t and at a later time t + t'. We have encountered correlation functions
already in the sections on probability. There they were denoted by
( q(t)q(t + n) . (12.34)
In the present case we do not have a random process and therefore the averaging
process indicated by the brackets seems to be meaningless. However, we may
replace (12.34) by a time average in the form
where we first integrate over t and then let the time interval 2 T become very
large or, more strictly speaking, let T go to infinity. Taking purely periodic
motion as a first example, i. e., for instance
q(t)=sinw 1 t, (12.36)
we readily obtain
(12.37)
355
344 12. Chaos
That means we obtain again a periodic function (compare Fig. 12.11). One may
simply convince oneself that a motion containing several frequencies such as
(12.38)
<q(t)q(t +1"»
f--\----f--+----'';-----t'
I
/
/ Fig. 12.11. The correlation function as function of
.L __ _
time t' for the periodic function q(t), (12.36). Note
that there is no decay of the amplitude even if t' ..... <Xl
q(t) =
1
-2
f+oo C(W) eiwt dw . (12.39)
1! -00
356
12.6 Discrete Maps, Period Doubling, Chaos, Intermittency 345
other hand, chaotic motion should contain a continuous broad band of fre-
quencies. Fig. 12.14 shows an example of the intensity distribution of frequencies
of the Lorenz model. Both criteria using (12.34) and (12.39) in the way described
above are nowadays widely used, especially when the analysis is based on com-
puter solutions of the original equations of motion. For the sake of completeness
we mention a third method, namely that based on Ljapunov exponents. A de-
scription of this method is, however, beyond the scope of this book 1.
Fig. 12.14. The power spectrum lc(wW for the Lorenz attract or
(redrawn after Y. Aizawa and 1. Shimada). Their original figure
contains a set of very closely spaced points. Here the studied
'----------w variable is q2(t) of the Lorenz model as described in the text
Chaotic motion in the sense discussed above occurs in quite different disciplines.
In the last century Poincare discovered irregular motion in the three-body
problem. Chaos is also observed in electronic devices and is known to electrical
engineers. It occurs in the Gunn oscillator, whose regular spikes we discussed in
Section 8.14. More recently numerous models of chemical reactions have been
developed showing chaos. It may occur in models both with and without diffusion.
Chaos occurs also in chemical reactions when they are externally modulated,
for instance by photochemical effects. Another example are equations describing
the flipping of the earth magnetic field which again shows a chaotic motion, i.e.,
a random flipping. Certain models of population dynamics show entirely irregular
changes of populations. It seems that such models can account for certain fluctua-
tion phenomena of insect populations.
Many of the recently obtained results are based on discrete maps, which are
treated in the next section.
357
346 12. Chaos
trajectory q(t) (which may either be the solution of a differential equation or may
stem from experimental data) hits the ql axis at a sequence of points XI, X2, .... Of
course, each intersection point X n + I is determined, at least in principle, once the
previous point Xn is fixed. Therefore we may assume that a relation
Xn + I = f,.(xn) (12.40)
exists. As it turns out in many cases of practical interest, the functionf., is indepen-
dent of n, so that (12.40) reads
Xn+1 =f(xn). (12.41)
Because (12.41) maps a value Xn onto a value Xn + 1 at discrete indices n, (12.41) is
called a discrete map. Fig. 12.16 shows an explicit example of such a function 2.
Of these functions ("maps"), the "logistic map"
Xn+ I = aXn(1 - xn) (12.42)
of the interval 0 :s; x :s; 1 onto itself has probably received the greatest attention.
In it, a plays the role of a control parameter which can be varied between
O:s; a:s; 4. It is an easy matter to calculate the sequence xJ,xz, ... ,x., ... deter-
mined by (12.42) with a pocket computer. Such calculations reveal a very inter-
esting behavior of the sequence XI, ... for various values of a. For instance, for
a < 3 the sequence XI' X 2 , ... converges towards a fixed point (Fig. 12.17). When
a is increased beyond a critical value ai' a new type of behavior occurs (Fig. 12.18).
In this latter case, after a certain transition "time" no, the points X"o + 1, Xno + 2, ...
jump periodically between two values, and a motion with "period two" is reached.
When we increase a further, beyond a critical value a2, the points Xn tend to a
sequence which is only repeated after 4 steps (Fig. 12.19), so that the "motion" goes
on with a "period four". With respect to "period two", the period length has
doubled.
Ifwe increase a further and further, at a sequence of critical values (1.1 the period
doubles each time. The thus resulting Fig. 12.20 shows a sequence of "period
doubling" bifurcations. The critical values obey a simple law:
lim al+1 - (1.1. = lJ = 4.6692016 ... , (12.43)
l~oo (1.1+2 - al+ I
as was found by Grossmann and Thomae in the case of the logistic map. Feigen-
baum observed that this law has universal character because it holds for a whole
class of functions f(xn). Beyond a critical value (I.e' the "motion" of Xn becomes
chaotic, but at other values of (I. periodic windows occur. Within these windows the
motion of Xn is periodic. Interesting scaling laws have been obtained for such
maps, but a presentation of them would go beyond the scope of this book.
2 A warning must be given here. While that section of the curve of Fig. 12.16 which is above
the diagonal (dashed line) is compatible with the trajectory depicted in Fig. 12.15, the section
below the diagonal does not folJow from a sequence X n + 1 > Xn but rather from .Yn + I < X n ·
This in turn implies that the trajectory of Fig. 12.15 crosses itself, which is not possible for an
autonomous system in the plane. Indeed, the map of Fig. 12.16 can be realized only by an
autonomous system of (at least) three dimensions, and the crossing points Sn can be obtained
(for example) by a projection of the three-dimensional trajectory on the q I - q2 plane, where
the new trajectory can indeed cross itself.
358
12.6 Discrete Maps, Period Doubling, Chaos, Intermittency 347
/
/
/
/
/
/
//
/
---+----+--1--~--~-1~--~X-3- q1 /
/
/
/
/
/
/
/
/
/
L -_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ ~ __ Xn
Fig. 12.15 Example of how a trajectory Fig. 12.16 Example of mapping X. onto
q(t) (the spiral) can be connected with a x. + l ' The mapping function is a parabola,
discrete series of points Xl' X 2 , X 3 , ... see (12.42). The dashed line represents the
diagonal
L--+--~--~_+--~--~_+--~---n n
n
Fig. 12.19 Jumps of period 4 of x.
359
348 12. Chaos
1, ~.___ :._,.__........
08
07
0.6
~ 05
0.4
03
0.2
01
/
/
/ '
01 02 03 04 0.5 0.6 07 0.8 0.9 10
,/
Another important aspect of the logistic map can be rather easily explained,
however. A well-known phenomenon occurring in turbulent flows is "inter-
mittency". Times in which laminar flows occur are interrupted by times of turbu-
lent outbursts. A model for this kind of behavior is provided by an "iterated map",
which we obtain as follows: We replace n by n + 1 everywhere in (12.42) so that
we obtain
(12.44)
360
12.6 Discrete Maps, Period Doubling, Chaos, Intermittency 349
Because the mapping described by (12.42) has been applied twice, a shorthand
notation of (12.45) is
(12.46)
361
13. Some Historical Remarks and Outlook
The reader who has followed us through our book has been most probably amazed
by the profound analogies between completely different systems when they pass
through an instability. This instability is caused by a change of external parameters
and leads eventually to a new macroscopic spatio-temporal pattern of the system.
In many cases the detailed mechanism can be described as follows: close to the
instability point we may distinguish between stable and unstable collective mo-
tions (modes). The stable modes are slaved by the unstable modes and can be
eliminated. In general, this leads to an enormous reduction of the degrees of
freedom. The remaining unstable modes serve as order parameters determining the
macroscopic behavior of the system. The resulting equations for the order par-
ameters can be grouped into a few universality classes which describe the dynamics
of the order parameters. Some of these equations are strongly reminiscent of
those governing first and second order phase transitions of physical systems in
thermal equilibrium. However, new kinds of classes also occur, for instance
describing pulsations or oscillations. The interplay between stochastic and deter-
ministic "forces" ("chance and necessity") drives the systems from their old states
into new configurations and determines which new configuration is realized.
~ Unstable mOdes}
Old structure ~ slave ->
~ Stable modes
364
13. Some Historical Remarks and Outlook 353
mountain which has so far separated different disciplines, in particular the "soft"
from the "hard" sciences.
It can be hoped that synergetics will contribute to the mutual understanding and
further development of seemingly completely different sciences. How synergetics
might proceed shall be illustrated by the following example taken from philology.
Using the terminology of synergetics, languages are the order parameters slaving
the subsystems which are the human beings. A language changes only little over
the duration of the life of an individual. After his birth an individual learns a
language, i. e., he is slaved by it, and for his lifetime contributes to the survival of
the language. A number of facts about languages such as competition, fluctuations
(change of meaning of words, etc.) can now be investigated in the frame established
by synergetics.
Synergetics is a very young discipline and many surprising results are still ahead
of us. I do hope that my introduction to this field will stimulate and enable the
reader to make his own discoveries of the features of self-organizing systems.
365
References, Further Reading, and Comments
Since the field of Synergetics has ties to many disciplines, an attempt to provide a more or less
complete list of references seems hopeless. Indeed, they would fill a whole volume. We therefore
confine the references to those papers which we used in the preparation ofthis book. In addition
we quote a number of papers, articles or books which the reader might find useful for further
study. We list the references and further reading according to the individual chapters.
l. Goal
2. Probability
There are numerous good textbooks on probability. Here are some of them:
Kai Lai Chung: Elementary Probability Theory with Stochastic Processes (Springer. Berlin-Heidel-
berg-New York 1974)
W. Feller: An Introduction to Probability Theory and Its Applications, Vol. 1 (Wiley, New York
1968), Vol. 2 (Wiley, New York 1971)
R.C. Dubes: The Theory of Applied Probability (Prentice Hall, Englewood Cliffs, N.J. 1968)
Yu. V. Prokhorov, Yu. A. Rozanov: Probability Theory. In Grundlehren der mathematisehen
Wissenschajien in Einzeldarstellungen, Bd. 157 (Springer, Bcrlin-Heidelberg-J\cw York 1968)
1. L. Doob: Stochastic Processes (Wiley, New York-London 1953)
M. Loeve: Probability Theory (D. van Nostrand, Princeton, N.J.-Toronto-New York-London
1963)
R. von Mises: Mathematical Theory of Probability and Statistics (Academic Press, New York-
London 1964)
3. Information
368
References, Further Reading, and Comments 357
4. Chance
369
358 References, Further Reading, and Comments
4.4 How to Use Joint Probabilities. Moments. Characteristic Function. Gaussian Processes
Same references as on Section 4.3.
4.6 Exact Stationary Solution of the Master Equatiun for Systems in Detailed Balance
For many variables see
H. Haken: Phys. Lett. 46A, 443 (1974); Rev. Mod. Phys. 47,67 (1975),
where further discussions are given.
F or one variable see
R. Landauer: J. Appl. Phys. 33, 2209 (1962)
4.10 The Meaning of Random Processes. Statiunary State, Fluctuations, Recurrence Time
For Ehrenfesfs urn model see
P. and T. Ehrenfest: Phys. Z. 8, 311 (1907)
and also
A. Miinster: In Encyclopedia (j( Physics, ed. by S. Fliigge, Vol. III/2; Principles of Thermodynamics
and Statistics (Springer, Berlin-Gbttingen-Heidelberg 1959)
5. Necessity
370
References, Further Reading, and Comments 359
M. W. Hirsch, S. Smale: Dijf('rential Eqllalions, Dynamical Systems, and Linear Alyehra (Academic
Press, New York-London 1974)
V. V. Nemytskii, V. V. Stepanov: Qualitatire Theory oj DifJerential Equations (Princeton Univ. Press,
Princeton, N.J. 1960)
Many of the basic ideas are due to
H. Poincare: Oellvres, Vol. 1 (Gauthiers-Villars, Paris 1928)
H. Poincare: Sur I'equilibre d'une masse tluide animee d'un mouvement de rotation. Acta Math. 7
(1885)
H. Poincare: Figures d'equilibre d'une masse fluide (Paris 1903)
H. Poincare: Sur Ie probleme de trois corps et les equations de la dynamique. Acta Math. 13 (1890)
H. Poincare: Les methodes nourelles de la mechanique ("(!leste (Gauthier-Villars, Paris 1892-1899)
5.3 Stahility
1. La Salle, S. Lefshetz: Stahility by Ljapunov's Direct Method with Applications (Academic Press,
New York-London 1961)
W. Hahn: Stability of Motion. In Die Grundlehren dcr mathematischen Wissenscha/ten in Einzel-
darstel/ungen. Bd. 138 (Springer, Berlin-Heidelberg-New York 1967)
D. D. Joseph: Stahiliry of Fluid Motions I and II, Springer Tracts in Natural Philosophy, Vois.
27. 28 (Springer, Berlin-Heidelberg-New York 1976)
Exercises 5.3: F. Schlagl: Z. Phys. 243, 303 (1973)
371
360 References, Further Reading, and Comments
The solution of the n-dimensional Fokker-Planck equation with linear drift and constant diffu-
sion coefficients was given by
M. C. Wang, G. E. Uhlenbeck: Rev. Mod. Phys. 17,2 and 3 (1945)
For a short representation of the results see
H. Haken: Rev. Mod. Phys. 47,67 (1975)
The theory of phase transitions of systems in thermal equlibrium is presented, for example, in the
following books and articles
L. D. Landau, I. M. Lifshitz: In Course of Theoretical Physics, Vol. 5: Statistical Physics (Perga-
mon Press, London-Paris 1959)
R. Brout: Phase Transitions (Benjamin, New York 1965)
L. P. Kadanoff, W. Gotze, D. Hamblen, R. Hecht, E. A. S. Lewis, V. V. Palcanskas, M. Rayl,
1. Swift, D. Aspnes, 1. Kane: Rev. Mod. Phys. 39, 395 (1967)
M. E. Fischer: Repts. Progr. Phys. 30, 731 (1967)
H. E. Stanley: Introduction to Phase Transitions and Critical Phenomena. Internal. Series of
Monographs in Physics (Oxford University, New York 1971)
A. Miinster: Statistical Thermodynamics, Vol. 2 (Springer, Berlin-Heidelberg-New York and Aca-
demic Press, New York-London 1974)
372
References, Further Reading, and Comments 361
C. Domb, M. S. Green, cds.: Phase Transitions and Critical Phenomena, Vols. 1-5 (Academic
Press, London 1972··76)
The modern and powerful renormalization group technique of Wilson is reviewed by
K. G. Wilson, J. Kogut: Phys. Rep. 12 C, 75 (1974)
S.-K. Ma: Modern Theory of Critical Phenomena (Benjamin, London 1976)
The profound and detailed analogies between a second order phase transition of a system in
thermal equilibrium (for instance a superconductor) and transitions of a non-equilibrium system
were first derived in the laser-case in independent papers by
R. Graham, H. Haken: Z. Phys. 213, 420 (1968) and in particular Z. Phys. 237, 31 (1970),
who treated the continuum mode laser, and by
V DeGiorgio, M. O. Scully: Phys. Rev. A2, 1170 (1970),
who treated the single mode laser.
For further references elucidating the historical development see Section 13.
6.8 Phase Transition AnaloRY in Continuous Media: Space Dependent Order Parameter
a) References to Systems in Thermal Equilihrium
The Ginzburg-Landau theory is presented, for instance. by
N. R. Werthamer: In Superconductivity, Vo\. 1, ed. by R. D. Parks (Marcel Dekker Inc., New
York 1969) p. 321
with further references
The exact evaluation of correlation functions is due to
0.1. Scalapino, M. Sears, R. A. Ferrell: Phys. Rev. B6, 3409 (l9n)
Further papers on this evaluation are:
L. W. Gruenberg, L. Gunther: Phys. Lett. 38A, 463 (1972)
M. Nauenberg, F. Kuttner, M. Fusman: Phys. Rev. A 13, 1185 (1976)
h) References to Systems Far from Thermal Equilibrillnl (and Nonphysical Systems)
R. Graham. H. Haken: Z. Phys. 237, 31 (1970)
Furthermore the Chapter 8 and 9
7. Self-Organization
7.1 Organization
H. Haken: unpublished material
7.2 Self-Organization
A different approach to the problem of self-organization has been developed by
J. v. Neuman: Theory of Self-reproducing Automata. ed. and completed by Arthur W. Burks (Uni-
versity of Illinois Press, 1966)
7.4 Adiahatic Elimination of Fast Relaxing Variables from the Fokker-Planck Equation
H. Haken: Z. Phys. B 20, 413 (1975)
7.5 Adiabatic Elimination of Fast Relaxing Variahles from the Master Equation
H. Haken: unpublished
373
362 References, Further Reading, and Comments
8. Physical Systems
374
References, Further Reading, and Conunents 363
375
364 References, Further Reading, and Comments
8.15 A Model for the Statistical Dynamics of the Gunn Instahility Near Threshold
1. B. Gunn: Solid State Commun. I, 88 (1963)
1. B. Gunn: IBM J. Res. Develop. 8. (1964)
For a theoretical discussion of this and related effects see for instance
H. Thomas: In Synergetics. cd. by H. Haken (Teubner, Stuttgart 1973)
Here, we follow essentially
K. Nakamura: 1. Phys. Soc. Jap. 38. 46 (1975)
376
References, Further Reading, and Comments 365
Models of chemical reactions showing spatial and temporal structures were treated in numerous
publications by Prigogine and his coworkers.
P. Glansdorff, I. Prigogine: Thermodynamik Theory of Structures, Stability and Fluctuations
(Wiley, New York 1971)
with many references, and
G. Nicolis, I. Prigogine: Sel{-orlianization in Non-equilibrium Systems (Wiley, New York 1977)
Prigogine has coined the word "dissipative structures". Glansdorff and Prigogine base their work
on entropy production principles and use the excess entropy production as means to search for
the onset of an instability. The validity of such criteria has been critically investigated by
R. Landauer: Phys. Rev. A 12, 636 (1975). The Glansdorff-Prigogine approach does not give an
answer to what happens at the instability point and how to determine or classify the new evolving
structures. An important line of research by the Brussels school, namely chemical reaction
models, comes closer to the spirit of Synergetics.
A review of the statistical aspects of chemical reactions can be found in
D. McQuarry: Supplementary Review Series in Appl. Probability (Methuen, London 1967)
A detailed review over the whole field gives the
Faraday Symposium 9: Phys. Chemistry of Oscillatory Phenomena. London (1974)
Y. Schiffmann: Phys. Rep. 64, 88 (1980)
For chemical oscillations see especially
G. Nicolis. J. Portnow: Chern. Rev. 73, 365 (1973)
9.4 Reaction-Diffusion Model with Two or Three Variables; the Brusselator and the Oregonator
We give here our own nonlinear treatment (A. Wunderlin, H. Haken. unpublished) of the reac-
tion-diffusion equations of the "Brusselator" reaction, originally introduced by Prigogine and
coworkers, I.c. For related treatments see
J. F. G. Auchmuchty, G. Nicolis: Bull. Math. BioI. 37. I (1974)
Y. Kuramoto, T. Tsusuki: Progr. Theor. Phys. 52, 1399 (1974)
M. Herschkowitz-Kaufmann: Bull. Math. BioI. 37, 589 (1975)
The Belousov-Zhabotinsky reaction is described in the already cited articles by Belousov and
Zha botinsk y.
The "Oregonator" model reaction was formulated and treated by
R.1. Field,E. Koras, R. M. Noyes: J. Am. Chern. Soc. 49, 8649 (1972)
R. J. Field, R. M. Noyes: Nature 237, 390 (1972)
R.1. Field, R. M. Noyes: J. Chern Phys. 60. 1877 (1974)
R. J. Field, R. M. Noyes: 1. Am. Chern. Soc. 96, 2001 (1974)
9.5 Stochastic Model for a Chemical Reaction Without Diffusion. Birth and Death Processes.
One Variable
A first treatment of this model is due to
V. J. McNeil. D. F. Walls: 1. Stat. Phys. 10,439 (1974)
9.6 Stochastic Model for a Chemical Reaction with Diffusion. One Variable
The master equation with diffusion is derived by
H. Hakcn: Z. Phys. B20, 413 (1975)
377
366 References, Further Reading, and Comments
We essentially follow
C. H. Gardiner, K. 1. McNeil, D. F. Walls, I. S. Matheson: 1. Stat. Phys. 14. 4. 307 (1976)
Related to this chapter are the papers by
G. Nicolis, P. Aden, A. van Nypelseer: Progr. Theor. Phys. 52.1481 (1974)
M. Malek-Mansour, G. Nicolis: preprint Febr. 1975
9.7 Stochastic Treatment oJthe Brusselator Close to Irs Solt Mode Instability
We essentially follow
H. Haken: Z. Phys. 820.413 (1975)
378
References, Further Reading, and Comments 367
12. Chaos
379
368 References, Further Reading, and Comments
380
References, Further Reading, and Comments 369
381
Subject Index
384
Subject Index 373
385
374 Subject Index
386
Subject Index 375
387
Part II
Advanced Topics
394
Preface IX
395
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 What is Synergetics About? ................................ 1
1.2 Physics ................................................. 1
1.2.1 Fluids: Formation of Dynamic Patterns . . . . . . . . . . . . . . . . 1
1.2.2 Lasers: Coherent Oscillations ........................ 6
1.2.3 Plasmas: A Wealth of Instabilities .................... 7
1.2.4 Solid-State Physics: Multistability, Pulses, Chaos ....... 8
1.3 Engineering.............................................. 9
1.3.1 Civil, Mechanical, and Aero-Space Engineering:
Post-Buckling Patterns, Flutter, etc. .................. 9
1.3.2 Electrical Engineering and Electronics: Nonlinear
Oscillations ....................................... 9
1.4 Chemistry: Macroscopic Patterns ........................... 11
1.5 Biology ................................................. 13
1.5.1 Some General Remarks ............................. 13
1.5.2 Morphogenesis .................................... 13
1.5.3 Population Dynamics ............................... 14
1.5.4 Evolution......................................... 14
1.5.5 ImmuneSystem .................................... 15
1.6 Computer Sciences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.6.1 Self-Organization of Computers, in Particular
Parallel Computing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.6.2 Pattern Recognition by Machines ..................... 15
1.6.3 Reliable Systems from Unreliable Elements. . . . . . . . . . . . . 15
1.7 Economy................................................ 16
1.8 Ecology................................................. 17
1.9 Sociology ............................................... 17
1.10 What are the Common Features of the Above Examples? ....... 17
1.11 The Kind of Equations We Want to Study .................... 18
1.11.1 Differential Equations .............................. 19
1.11.2 First-Order Differential Equations .................... 19
1.11.3 Nonlinearity....................................... 19
1.11.4 Control Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.11.5 Stochasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.11.6 Many Components and the Mezoscopic Approach . . . . . . . 22
1.12 How to Visualize the Solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.13 Qualitative Changes: General Approach. . . . . . . . . . . . . . . . . . . . . . 32
1.14 Qualitative Changes: Typical Phenomena .................... 36
XII Contents
1.14.1 Bifurcation from One Node (or Focus) into Two Nodes
(or Foci) .......................................... 37
1.14.2 Bifurcation from a Focus into a Limit Cycle
(Hopf Bifurcation) ................................. 38
1.14.3 Bifurcations from a Limit Cycle ...................... 39
1.14.4 Bifurcations from a Torus to Other Tori ............... 41
1.14.5 Chaotic Attractors ................................. 42
1.14.6 Lyapunov Exponents* .............................. 42
1.15 The Impact of Fluctuations (Noise). Nonequilibrium Phase
Transitions .............................................. 46
1.16 Evolution of Spatial Patterns ......................... . . . . . . 47
1.17 Discrete Maps. The Poincare Map . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
1.18 DiscreteNoisyMaps ....................................... 56
1.19 Pathways to Self-Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
1.19.1 Self-Organization Through Change of Control
Parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
1.19.2 Self-Organization Through Change of Number of
Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
1.19.3 Self-Organization Through Transients. . . . . . . . . . . . . . . . . 58
1.20 How We Shall Proceed .................................... 58
398
Contents XIII
399
XIV Contents
400
Contents XV
401
1. Introduction
1.2 Physics
Fig. 1.2.2. (a) Schematic diagram of the formation of rolls. (b) A computer calculation of the
trajectory of a particle of the fluid within a roll. In the case shown two rolls are formed but only the
motion of a particle within the upper roll is presented. [After K. Marx: Diplom thesi s. University of
Stuttgart (1982)]
example the Taylor instability and its subsequent instabilities. In these experi-
ments (Fig. 1.2.1) the motion of a liquid between coaxial cylinders is studied.
Usually one lets the inner cylinder rotate while the outer cylinder is kept fixed,
but experiments have also been performed where both cylinders rotate. Here we
shall describe the phenomena observed with fixed outer cylinders but at various
rotation speeds of the inner cylinder. At slow rotation speeds the fluid forms
coaxial streamlines. This is quite understandable because the inner cylinder tries
to carry the fluid along with it by means of the friction between the fluid and the
cylinder. When the rotation speed is increased (which is usually measured in the
dimensioniess Taylor number), a new kind of motion occurs. The motion of the
fluid becomes organized in the form of rolls in which the fluid periodically moves
outwards and inwards in horizontal layers (Fig. 1.2.2a, b).
With further increase of the Taylor number, at a second critical value the rolls
start oscillations with one basic frequency, and at still more elevated Taylor
numbers with two fundamental frequencies. Sometimes still more complicated
frequency patterns are observed. Eventually, at still higher Taylor numbers
chaotic motion sets in. As is shown in Fig. 1.2.3, the evolving patterns can be
seen directly. Furthermore, by means of laser light scattering, the velocity distri-
bution and its Fourier spectrum have been measured (Fig. 1.2.4a - c). In par-
ticular cases with increasing Taylor number, a sequence of newly developing fre-
quencies which are just 1/2, 1/4, 1/8, 1116 of the fundamental frequency is
observed. Since half a frequency means a double period of motion, this phe-
nomenon is called period doubling. There are several features which are quite
typical for self-organizing systems. When we change an external parameter, in
404
1.2 Physics 3
Fig. 1.2.3. Instability hierarchy of the Taylor instability. (a) Formation of rolls. (b)The rolls start an
oscillation . (e) A more complicated oscillatory motion of the rolls. (d) A chaotic motion. [After H . L.
Swinney, P. R. Fenstermacher, J. P . Gollub: In Synergetics, A Workshop, ed. by H. Haken,
Springer Ser. Synergetics , Vol. 2 (Springer, Berlin, Heidelberg, New York 1977) p. 60)
the present case the rotation speed, the system can form a hierarchy of patterns,
though these patterns are not imposed on the system by some means from the
outside. The patterns can become more and more complicated in their spatial and
temporal structure.
Another standard type of experiment leads to the convection instability (or
Benard instability) and a wealth of further instabilities. Here, a fluid layer in a
vessel with a certain geometry such as cylindrical or rectangular is used. The fluid
is subject to gravity. When the fluid layer is heated from below and kept at a
certain constant temperature at its upper surface, a temperature difference is
established. Due to this temperature difference (temperature gradient), a vertical
flux of heat is created. If the temperature gradient is small, this heat transport
occurs microscopically and no macroscopic motion of the fluid is visible. When
the temperature gradient is increased further, suddenly at a critical temperature
gradient a macroscopic motion of the fluid forming pronounced patterns sets in.
For instance, the heated liquid rises along stripes, cools down at the upper sur-
face and falls down again along other stripes so that a motion in the form of rolls
occurs. The dimension of the rolls is of the order of the thickness of the fluid
layer, which in laboratory systems may range from millimeters to several centi-
meters.
The same phenomenon is observed in meteorology where cloud streets with
dimensions of several hundred meters occur. When in a rectangular geometry the
temperature gradient, which in dimensionless units is measured by the Rayleigh
number, is increased further, the rolls can start an oscillation, and at a still higher
Rayleigh number an oscillation of several frequencies can set in. Finally, an
entirely irregular motion, called turbulence or chaos, occurs.
There are still other routes from a quiescent layer to turbulence, one such
route occurring by period-doubling. When the Rayleigh number is increased, a
rather complex motion occurs in which at well-defined Rayleigh numbers the
period of motion is repeatedly doubled (Fig. 1.2.5a - c). In cylindrical containers
concentric rolls can be observed or, if the symmetry with respect to the horizontal
middle plane is violated, hexagonal patterns occur. Also, transitions between
rolls and hexagons and even the coexistence of rolls and hexagons can be
observed (Fig. 1.2.6). Other observed patterns are rolls which are oriented at rec-
405
4 1. Introduction
llT Ii~jIW,I\I,l'~1
::3
o ':-0- - - ' - -.......-...J3L--...L...--:!5
..
a.. II) ,
t lsi II.
a
o 5 10
W
IS 20
102 , -_ _ _, -_ _---,r-_ _--._ _ _--,
Plw )
II
10 1
quency of the inner cylinder, rj and ro are the
P{wl R/Rc~ 19.2 radii of the inner and outer cylinder, respec-
tively, and v is the kinematic viscosity.
(Right side) The power spectrum corre-
W I -W3 W3 i
1 sponding to the left curve . There is one fun-
101 ,I damental frequency WI but remarkably
~1 ~
many harmonics. (b) The power spectrum of
A~JI"I . • IdLMl1 the radial component o f the velocity at
R 1R c = 13.3. Note that the spectrum now
C
0 0.8 1.&
W shows three fundament a l frequencies w I'
w2' wJ and severa l linear combinations of
10 1 them. (e) These spectra illustrate the dis-
appearance of w3 at R I R c = 19.8 ± 0.1.
R/R c ' 20 .1
P(wl There is a component B at W - 0.45 in both
spectra. [After H. L. Swinney, P. R. Fen-
stermacher, J. P. Gollub: In Synergetics, A
101
Workshop, ed. by H. H aken, Springer Ser.
Synergetics, Vol. 2 (Springer, Berlin, Heidel-
berg, New York 1977) p. 60].
0 0.8 W 1&
tangles and which thus form a rectangular lattice (Fig. 1.2.7). At elevated
Rayleigh numbers a carpet-like structure is found. Many patterns can also exhibit
a number of imperfections (Fig. 1.2.8). In order to obtain most clearcut results,
in modern experiments the aspect ratio, i. e., the horizontal dimension of the
vessel divided by the vertical dimension, is taken small, i. e., of the order of
unity. At higher aspect ratios the individual transitions can very quickly follow
after one another or can even coexist. Another class of experiments, where the
fluid is heated from above, leads to the "Marengo instability". Further, in the
atmospheres of the earth and other planets where gravitation, rotation and
heating act jointly, a wealth of patterns is formed.
406
1.2 Physics 5
100,--____, -__________________- ,
~r-----------------------~
LGMAG
DB
- soc+-~~--~~--~~~--~~~
a 0.0 HZ
b -SOI.D+O-
, O~~-'-~~-H~Z~-~~~--l
5000 500,0
10,0 r-r-----------------------...,-,
Fig. 1.2.6
Fig. 1.2.7
Fig. 1.2.6. These photographs show the surface of a liquid heated from below. The coexistence of
hexagons and rolls is clearly visible. [After 1. Whitehead: private communication]
Fig. 1.2.7. Photograph of the surface of a liquid heated from below. The rectangular lattice formed
by two rectangular roll systems is clearly visible. [After J. Whitehead: In Fluctuations, Instabilities
and Phase Transitions, ed . by T. Riste (Plenum, New York 1975)]
407
6 1. Introduction
noise ->
h t '11 r }
{ co eren OSCI a IOn ->
{ periodic pulses at ff. equency
which modulate oscillation
W2}
at frequency Wj at frequency Wj
i.e.,
no oscillation -> 1 frequency -> 2 frequencies.
408
1.2 Physics 7
Eltl Eitl
I
V V V v V VV V V V
a b
Em
409
8 1. Introduction
Because of the charges and their interaction with electromagnetic fields, new
types of instabilities may occur, and indeed in plasmas a huge variety of instabili-
ties are observed. For example, Fig. 1.2.10 presents a theoretical result on the
formation of a velocity pattern of a plasma which is heated from below and sub-
jected to a constant vertical magnetic field. It is beyond the scope of this book to
list these instabilities here. Besides acquiring ordered states, a plasma can also
migrate from one instability to another or it may suffer a violent instability by
which the plasma state collapses. A study of these phenomena is of utmost im-
portance for the construction of fusion reactors and for other fields, e. g., astro-
physics.
bd
I
Fig. 1.2.11. Different operating states A, B, C of a
tunnel diode where the current I is plotted versus applied
voltage V. The states A and C are stable, whereas B is
unstable. This picture provides a typical example of a
bistable device with stability points A and C
410
1.3 Engineering 9
1.3 Engineering
411
10 1. Introduction
412
1.4 Chemistry: Macroscopic Patterns 11
In this field synergetics focuses its attention on those phenomena where macro-
scopic patterns occur. Usually, when reactants are brought together and well
stirred, a homogeneous end product arises. However, a number of reactions may
show temporal or spatial or spatio-temporal patterns. The best known example is
the Belousov-Zhabotinsky reaction. Here Ce2(S04h, KBr03' CHiCOOHh,
H 2S04 as well as a few drops of ferroine (redox indicator) are mixed and stirred.
1\1 I~
413
12 1. Introduction
~ 0
(0) (b)
-J - 1
~ -2
u
Lu
e; -3
~ -4
~
g
-J
FREQUENCY (Hz)
· '1· lee.,.
p.'
C '2C '3C ~C 's,I C5• ?
(c)
1
... "" Ie 'e,
2 3 4
I ,ieI • i •
2.0
•
1
I •
1.0 1.S T (hr)
Fig. 1.4.3a - c. Dynamics of the Belousov-Zhabotinsky reaction in a stirred fl ow reactor. Power
spectra of the bromide ion potential are shown (a) for a periodic regime, and (b) a chaotic regime for
different residence times T where T = reactor volume/flow rate. All other variables were held fixed.
The plot (c) shows the alternating sequence between periodic and chaotic regimes when the residence
rate is increased. [After] . S. Turner, J. C. Roux, W. D. McCormick, H. L. Swinney: Phys. Lett.
8SA, 9 (1981) and in Nonlinear Problems: Presence and Future, cd. by A. R. Bishop (North-Holland ,
Amsterdam 1981)]
414
1. 5 Biology 13
1.5 Biology
1.5.2 Morphogenesis
The central problem of morphogenesis can be characterized as follows. How do
the originally undifferentiated cells know where and in which way to differen-
tiate? Experiments indicate that this information is not originally given to the
individual cells but that a cell within a tissue receives information on its position
from its surroundings, whereupon it differentiates. In experiments with embryos,
transplantation of a cell from a central region of the body into the head region
causes this cell to develop into an eye. These experiments demonstrate that the
415
14 1. Introduction
cells do not receive their information how to develop from the beginning (e. g.,
through their DNA), but that they receive that information from the position
within the cell tissue. It is assumed that the positional information is provided by
a chemical "pre-pattern" which evolves similarly to the patterns described in the
section on chemistry above. These pattern formations are based on the
cooperation of reactions and diffusion of molecules.
It is further assumed that at sufficiently high local concentration of these
molecules, called morphogenes, genes are switched on, leading to a differentia-
tion of their cells. A number of chemicals have been found in hydra which are
good candidates for being activating or inhibiting molecules for the formation of
heads or feet. While a detailed theory for the development of organs, e. g., eyes,
along these lines of thought is still to be developed, simpler patterns, e. g., stripes
on furs, rings on butterfly wings, are nowadays explained by this kind of
approach.
1.5.4 Evolution
Evolution may be viewed as the formation of new macroscopic patterns (namely
new kinds of species) over and over again. Models on the evolution of biomole-
cules are based on a mathematical formulation of Darwin's principle of the
416
1.6 Computer Sciences 15
417
16 1. Introduction
v v
Fig. 1.6.3. How to visualize the buildup of a reliable system from unreliable elements. (Left side) The
potential function V of an individual element being very flat allows for quick jumps of the system
between the holding states 0 and 1. (Right side) By coupling elements together the effective potential
can be appreciably deepened so that a jump from one holding state to another becomes very
improbable
causes. It is suspected that the elements of our brain, such as neurons, are of such
a type. Nature has mastered this problem to construct reliable systems out of
these unreliable elements. When computer elements are made smaller and
smaller, they become less and less reliable. In which way can one put computer
elements together so that the total system works reliably? The methods of syn-
ergetics indicated below will allow us to devise systems which fulfill this task. Let
us exemplify here how a reliable memory can be constructed out of unreliable
elements. In order to describe the behavior of a single element, the concept of the
order parameter is used (to be explained later). Here it suffices to interpret the
order parameter by means of a particle moving in a potential (in an over damped
fashion) (Fig. 1.6.3). The two holding states 1 and 2 of the memory element are
identified with the two minima of the potential. Clearly, if the "valleys" are flat,
noise will drive the "particle" back and forth and holding a state is not possible.
However, by coupling several elements together, a potential with deeper values
can be obtained, which is obviously more reliable. Coupling elements in various
ways, several reliable holding states can be obtained. Let us now turn to phenom-
ena treated by disciplines other than the natural sciences.
1.7 Economy
418
1.10 What arc the Common Features of the Above Examples? 17
1.8 Ecology
1.9 Sociology
In all cases the systems consist of very many subsystems. When certain condi-
tions ("controls") are changed, even in a very unspecific way, the system can
develop new kinds of patterns on macroscopic scales. A system may go from a
419
18 1. Introduction
In this section we shall discuss how we can deal mathematically with the phenom-
ena described in the preceding sections and with many others. We shall keep in
mind that our approach will be applicable to problems in physics, chemistry, and
biology, and also in electrical and mechanical engineering. Further areas are
economy, ecology, and sociology. In all these cases we have to deal with systems
composed of very many subsystems for which not all information may be avail-
able. To cope with these systems, approaches based on thermodynamics or infor-
mation theory are frequently used. But during the last decade at least, it became
clear that such approaches (including some generalizations such as irreversible
thermodynamics) are inadequate to cope with physical systems driven away from
thermal equilibrium or with, say, economic processes. The reason lies in the fact
that these approaches are basically static, ultimately based on information theory
which makes guesses on the numbers of possible states. In [1], I have demon-
strated how that formalism works and where its limitations lie. In all the systems
considered here, dynamics plays the crucial role. As shall be demonstrated with
mathematical rigor, it is the growth (or decay) rates of collective "modes" that
determine which and how macroscopic states are formed. In a way, we are led to
some kind of generalized Darwinism which even acts in the inanimate world,
namely, the generation of collective modes by fluctuations, their competition and
finally the selection of the "fittest" collective mode or a combination thereof,
leading to macroscopic structures.
Clearly, the parameter "time" plays a crucial role. Therefore, we have to
study the evolution of systems in time, whereby equations which are sometimes
420
1.11 The Kind of Equations We Want to Study 19
called "evolution equations" are used. Let us study the structure of such
equations.
q= aq. (1.11.1)
Such equations are met, for instance, in chemistry, where the production rate q
of a chemical is proportional to its concentration q ("autocatalytic reaction"), or
in population dynamics, where q corresponds to the number of individuals. In
wide classes of problems one has to deal with oscillators which in their simplest
form are described by
(1.11.2)
(1.11.3)
These two equations are equivalent to (1.11.2). By means of the trick of introduc-
ing additional variables we may replace equations containing higher-order
derivatives by a set of differential equations which contain first-order derivatives
only. For a long time, equations of the form (11.1.1-4) or generalizations to
many variables were used predominantly in many fields, because these equations
are linear and can be solved by standard methods. We now wish to discuss those
additional features of equations characteristic of synergetic systems.
1.11.3 Nonlinearity
All equations of synergetics are nonlinear. Let us consider an example from
chemistry, where a chemical with concentration ql is produced in an auto-
421
20 1. Introduction
catalytic way by its interaction with a second chemical with a concentration q2.
The increase of concentration of chemical 1 is described by
(1.11.5)
Clearly, the cooperation of parts (i. e., synergetics), in our case of molecules, is
expressed by a nonlinear term. Quite generally speaking, we shall consider equa-
tions in which the right-hand side of equations of the type (1.11.5) is a nonlinear
function of the variables of the system. In general, a set of equations for several
variables qj must be considered.
(1.11.6)
the constant fJ describes the coupling between the two systems ql and q2. When
we manipulate the strength of the coupling from the outside, fJ plays the role of a
control parameter.
1.11.5 Stochasticity
A further important feature of synergetic systems is stochasticity. In other
words, the temporal evolution of these systems depends on causes which cannot
be predicted with absolute precision. These causes can be taken care of by "fluc-
tuating" forces f(t) which in their simplest form transform (1.11.1) into
q= aq + f(t) . (1.11. 7)
422
1.11 The Kind of Equations We Want to Study 21
tuating forces are given for each system under consideration. Before the advent
of quantum theory, thinking not only in physics but also in many other fields was
dominated by a purely mechanistic point of view. Namely, once the initial state
of a system is known it will be possible to predict its further evolution in time
precisely. This idea was characterized by Laplace's catch phrase, "spirit of the
world". If such a creature knows the initial states of all the individual parts of a
system (in particular all positions and velocities of its particles) and their interac-
tions, it (or he) can predict the future for ever. Three important ideas have
evolved since then.
a) Statistical Mechanics. Though it is still in principle possible to predict, for
instance, the further positions and velocities of particles in a gas, it is either un-
desirable or in practice impossible to do this explicitly. Rather, for any purpose it
will be sufficient to describe a gas by statistics, i. e., to make predictions in the
sense of probability. For instance, how probable it will be to find n particles with
a velocity v between v and v + dv. Once such a probabilistic point of view is
I
c) Chaos. There is a third, more recent, development which shows that even
without quantum fluctuations the future path of systems cannot be predicted.
Though the equations describing temporal evolution are entirely deterministic,
future development can proceed along quite different routes. This rests on the
fact that some systems are extremely sensitive in their further development to
initial conditions. This can be most easily visualized by a simple mechanical
example. When a steel ball falls on a vertical razor-blade, its future trajectory
will depend extremely sensitively on its relative position before it hits the razor-
blade. Actually, a whole industry of gambling machines is based on this and
similar phenomena.
If the impact of fluctuations on a system is taken care of by a fluctuating
force, such as in (1.11.7), we shall speak of additive noise. A randomly fluc-
423
22 1. Introduction
tuating environment may cause other types of noise, too. For instance, the
growth rate in (1.11.1) may fluctuate. In such a case, we have
q= a(t)q, (1.11.8)
424
1.12 How to Visualize the Solutions 23
In all such cases qj or the state vector q become functions of space and time
q=Df'!...q, (1.11.11)
(1.11.12)
where V is the operator (%x, %y, %z). The study of such equations is an
enormous task and we shall proceed in two steps. First we shall outline how at
least in some simple cases the solutions of these equations can be visualized, and
then we shall focus our attention on those phenomena which are of a general
nature.
425
24 1. Introduction
Fig. 1.12.1. Temporal evolution of the variables ql and q2' respectively, in the case of time I. At an
initial time 10 a specific value of ql and q2 was prescribed and then the solid curves evolved. If dif-
ferent initial values of ql' q2 are given, other curves (- - -) evolve. If a system possesses many
variables, for each variable such plots must be drawn
Fig. 1.12.2. Instead of plotting the trajectories as a function of time as in Fig. 1.12.1, we may plot the
trajectory also in the ql' q2 plane where for each time Ij the corresponding point ql (tj)' q2(t) is
plotted. If the system starts at an initial time to from different points, different trajectories evolve. In
the case of n variables the trajectories must be drawn in an n-dimensional space
----
""""..-
"
I
I ""
I
I
I
I
I
I
I
t - -(X)
Fig. 1.12.3. In general, trajectories are plotted Fig. 1.12.4. Example of a set of trajectories
for t -> + 00 and t -> - 00 if one follows the tra-
jectory
ferent ways. For instance, in two dimensions they may terminate in a node (Fig.
1.12.5) or in a focus (Fig. 1.12.6). Because the streamlines are attracted by their
endpoints, these endpoints are called attractors. In the case of a node, the tem-
poral behavior of q is given by a graph similar to Fig. 1.12.7, whereas in the case
of a focus the corresponding graph is given in Fig. 1.12.8. In the plane the only
other singular behavior of trajectories besides nodes, foci and saddle points, is a
limit cycle (Fig. 1.12.9). In the case shown in Fig. 1.12.9 the limit cycle is stable
because it attracts the neighboring trajectories. It represents an "attractor". The
corresponding temporal evolution of qt, which moves along the limit cycle, is
shown in Fig. 1.12.10 and presents an undamped oscillation. In dimensions
greater than two other kinds of attractors may also occur. An important class
consists of attractors which lie on manifolds or form manifolds. Let us explain
426
1.12 How to Visualize the Solutions 25
the concept of a manifold in some detail. The cycle along which the motion
occurs in Fig. 1.12.9 is a simple example of a manifold. Here each point on the
manifold can be mapped to a point on an interval and vice versa (Fig. 1.12.11).
q, q,
Fig. 1.12.5. Trajectories ending at a (stable) Fig. 1.12.6. Trajectories ending at a (stable)
node. If time is reversed, the trajectories start focus. If time is reversed the trajectories start
from that node and the node is now unstable from that focus and it has become an unsta-
ble focu s
q,
Fig. 1.12.9. Stable limit cycle in a Fig. 1.12.10. Temporal behavior of a variable, e. g.,
plane. The trajectories approach the q2(t), in the case of a limit cycle
limit cycle from the outside and in-
side of the limit cycle
427
26 1. Introduction
Fig. 1.12.11. One-to-one mapping of the individ- Fig. 1.12.12. One-to-one mapping of overlap-
ual points of a limit cycle onto points on a line ping pieces on a limit cycle onto overlapping
pieces on a line
The total manifold, especially, can be covered by pieces which can be mapped on
overlapping pieces along the line and vice versa (Fig. 1.12.12). Each interval on
the circle corresponds to a definite interval along the line. If we take as the total
interval 0 till 2 n we may describe each point along the circle by a point on the
qJ axis from 0 till 2 n. Because there is one-to-one mapping between points on the
limit cycle and on the interval 0 ... 2 n, we can directly use such a coordinate
(system) on the limit cycle itself. This coordinate system is independent of any co-
ordinate system of the plane, in which the limit cycle is embedded.
The limit cycle is a differentiable manifold because when we use, e. g., time as
a parameter, we assume q exists - or geometrically expressed - the limit cycle
possesses everywhere a tangent.
Of course, in general, limit cycles need not be circles but may be other closed
orbits which are repeated after a period T = 2 n/ w (w: frequency). If q exists, this
orbit forms a differentiable manifold again. In one dimension such a periodic
motion is described by
(1.12.1)
n
or in m dimensions by
428
1.12 How to Visualize the Solutions 27
2n r------------------,
2n
Fig. 1.12.15. The two-dimensional torus with local coordinates 'PI and 'P2 can be mapped one-to-one
on a square (left side of figure)
by the coordinate system ({Jl' ({J2' Because in each point ({Jl' ({J2 a tangent plane
exists, the torus forms a differentiable manifold. Such two- and higher-dimen-
sional tori are adequate means to visualize quasiperiodic motion, which takes
place at several frequencies. An example is provided by
(1.12.4)
429
28 1. Introduction
(1.12.5)
2 n: I---"'---~r---"
L -_ _ - L _ _ _L -_ _ ~_~ ~1
2n:
Fig. 1.12.16. A trajectory in a plane and its image on the torus in the case w2: WI = 3: 2. [n thi s
figure we have started the trajectory at 'PI = 0, 'P2 = O. Because of the periodicity condition we
continue that trajectory by projecting its cross section with the axis 'P2 = 271 down to the axis 'P2 = 0
(- - - ). It then continues till the point 'PI = 271, 'P2 = 71 from where we projected it on account of the
periodicity condition on 'PI = 0, 'P2 = 71 ( - - - ). From there it continues by means of the solid linc,
etc. Evidently by this construction we eventually find a closed line
2n: I--------~~
430
1.12 How to Visualize the Solutions 29
L-_'----_'----_'----_'----_L-_ <P,
2ft
Fig. 1.12.18. The case w 2 : wI = 5: 1
2 ft hrrrlTT'-TT1rrrrr--rrTT1rTT1
L-LU~UL-U~~~~~LL~ ~
2n:
Fig. 1.12.19. The case W2 : wI = ;rr: 1, the case of an irrational number. In this computer plot neither
the trajectories in the 'PI' 'P2 plane nor on the torus ever close and indeed the traj ectories fill the whole
plane or the whole toru s, respectively. Note that in this plot the computer ran only a finite time,
otherwise the square or the torus would appear entirely black
431
30 1. Introduction
stable
slable manifold
manifold
(
-- unstab te
manifotd
_e--____
-
r
Fig. 1.12.20. The stable and unstable Fig. 1.12.21. The unstable and stable manifold connect-
manifold of a saddle point (compare ed with a limit cycle
text)
(1.12.6)
(1.12.7)
where a, y > 0 and the N;'s U = 1, 2) are nonlinear functions of qj. Confining
ourselves to small deviations qj' we may safely neglect the nonlinear terms N j
which are assumed to be of the order 0 (Iq 12 ). We then notice that small devia-
tions qj are exponentially enhanced in time, meaning that ql is tangent to the
unstable manifold of the saddle. Conversely, small perturbations q2 are exponen-
tially damped in time and we may therefore conclude that direction q2 is tangent
to the stable manifold at the saddle.
But in addition, generally a third class of directions might exist along which a
perturbation is neither enhanced nor damped, i. e., the behavior along these
432
1.12 How to Visualize the Solutions 31
z z
y y y
Fig. 1.12.22. Stereoplot of the Rossler attractor. To obtain a stereoscopic impression, put a sheet of
paper between your nose and the vertical middle line. Wait until the two pictures merge to one. The
parameter values are: a = b = 0.2, c = 5.7, x(O) = y(O) = - 0.7, z(O) = 1. Axes: - 14 ... 14 for x and
y , 0 .. . 28 for z. [After O. E. Rossler: In Synergetics, A Workshop, ed. by H. Haken, Springer Ser.
Synergetics, Vol. 2 (Springer, Berlin, Heidelberg, New York 1977) p. 184]
Fig. 1.12.23. Stereoplot of the (modified) Lorenz attractor. The parameters are: a = 2.2, (] = 4,
r=80, b = 8/3, x(0) = 5.1, y(O) = -13.72, z(O) = 52. Axes: - 50 .. . 50 for x, - 70 .. . 70 for y,
0 ... 14 for z. [After O. E . Rossler: In Synergetics, A Workshop, ed. by H . Haken, Springer Ser.
Synergetics, Vol. 2 (Springer, Berlin, Heidelberg, New York 1977) p. 184]
directions is neutral. These directions are tangent to the so-called center mani-
fold. An example is provided by the limit cycle of Fig. 1.12.21. Obviously a per-
turbation tangent to the cycle can be neither enhanced nor damped in time. Fur-
thermore, we observe that in the case of the saddle (Fig. 1.12.20) the center mani-
fold is reduced to the point itself.
From recent work it appears that there may be attractors which are not mani-
folds. Such attractors have been termed "strange attractors" or "chaotic attrac-
tors". The reader should be warned of some mathematical subtleties. The notion
"strange attractor" is nowadays mostly used if certain mathematical axioms are
fulfilled and it is not known (at least at present) whether systems in nature fulfill
these axioms. Therefore we shall use the more loosely defined concept of a
chaotic attractor. Once the vector q (t) has entered such a region it will stay there
forever. But the trajectory does not lie on a manifold. Rather, the vector q(t) will
go on like a needle which we push again and again through a ball of thread.
Chaotic attractors may occur in three and higher dimensions. Examples are pre-
sented in Figs. 1.12.22 and 23.
The trajectories of a chaotic attractor may be generated by rather simple dif-
ferential equations (if the parameters are adequately chosen). The simplest
example known is the Rossler attract or. Its differential equations possess only
one nonlinearity and read
x= -y-z , (1.12.8)
y=x+ay, (1.12.9)
i = b+z(x-c), (1.12.10)
433
32 1. Introduction
The next simple (and historically earlier) example is the Lorenz attract or. The
corresponding differential equations are
x=a(y-x), (1.12.11 )
y=x(r-z)-y, (1.12.12)
z=xy-bz, (1.12.13)
where a, b, r are constant parameters. This model was derived for the convection
instability in fluid dynamics. The single mode laser is described by equations
equivalent to the Lorenz equations. A plot of a (modified) Lorenz attractor is
provided by Fig. 1.12.23. The modification consists in adding a constant a to the
rhs of (1.12.11).
1 The concept of structural stability seems to playa fundamental role in biology in a still deeper sense
than in the formation of different species by way of deformation (Fig. 1.13.1). Namely, it seems
that, say within a species, organisms exhibit a pronounced invariance of their functions against
spatial or temporal deformations. This makes it sometimes difficult to perform precise (and repro-
ducible) physical measurements on biological objects. Most probably, in such a case we have to
look for transformation groups under which the function of an organ (or animal) is left invariant.
This invariance property seems to hold for the most complicated organ, the human brain. For
example, this property enables us to recognize the letter a even if it is strongly deformed. From this
ability an art out of writing letters (in the double sense of the word) developed in China (and in old
Europe).
434
1.13 Qualitative Changes: General Approach 33
mind. In contrast to our example of the two kinds of fish (Fig. 1.13.1), we shall
not be concerned with static patterns but rather with patterns of trajectories, i. e.,
in other words, with the flows we treated in the foregoing section. As we know,
we may manipulate a system from the outside, which in mathematical form will
be done by changing certain control parameters. We want to show how the prop-
erties of a system may change dramatically even if we alter a control parameter
only slightly. In [1] a simple example for this kind of behavior was presented.
V(ql
- -r----7¥<f'----i--q
Fig. 1.13.2. Equilibrium position of a ball in tw o kind s of vase
(dashed line or solid line)
Consider a ball which slides down along the walls of a vase. If the vase has the
form of the dashed line of Fig. 1.13 .2, the ball comes to rest at q = O. If the vase,
however, is deformed as indicated by the solid line, the ball comes to rest at
q = + a or q = - a. The flow diagram of this ball is easily drawn. In the case of
the potential curve with a single minimum, we obtain Fig. 1.13.3, whereas in the
case of two minima, the flow diagram is given by Fig. 1.13.4. A related transition
from a single attractor to two attractors is provided by the self-explaining Figs.
- ---.+._0 _ _ _ - q
- - ....---- - q
Fig. 1.13.3. One-dimensional mo- Fig. 1.13.4. One-dimensional motion of a ball with
tion of a ball with coordinate q end- two points of stable equilibrium (.) and one unstable
ing at one stable point poin t ( 0 )
435
34 1. Introduction
- ./
/1
Fig. 1.13.5. Trajectories ending at a node Fig. 1.13.6. Two stable and an unstable node
1.13 .5, 6. By the way, the example of Fig. 1.13 .2 reveals the important role
played by fluctuations. If the ball is initially at q = 0, to which valley it will
eventually go entirely depends on fluctuations.
What makes the transition from Figs. 1.13.3 to 4 or from Figs. 1.13.5 to 6 so
different from the example of the two kinds of fish is the following . Let us draw
one of the two fishes on a rubber sheet. Then by mere stretching or pushing
together the rubber sheet, we may continuously proceed from one picture to the
other. However, there is no way to proceed from Fig. 1.13.5 to 6 by merely
stretching or deforming a rubber sheet continuously as it is possible in Fig.
1.13.7. In other words, there is no longer any one-to-one mapping between the
stream lines of one flow to the stream lines of the other (Fig. 1.13.8). In the
mathematical sense we shall understand by "structural instability" or "structural
changes" such situations in which a one-to-one mapping becomes impossible.
We now wish to discuss briefly how to check whether the change of a control
parameter causes structural instability. We shall come to this problem in much
greater detail later. Here the main method is illustrated by means of a most
simple example. The equation which describes the sliding motion of a ball in a
vase reads:
(1.13.1)
°
For a> 0 the solution reads q = 0, although q = still remains a solution for
°
a<O. Of course, by looking at Fig. 1.13 .2 we immediately recognize that the
position q = is now unstable. However, in many cases of practical interest we
436
1.13 Qualitative Changes: General Approach 35
may not invoke the existence of a potential curve as drawn in Fig. 1.13.2. Rather,
we must resort to another approach, namely linear stability analysis. To this end
a small time-dependent deviation u is introduced so that we write the solution q
of (1.13.1) as
q = qo + u = u. (1.13.2)
Inserting (1.13.2) into (1.13.1) and keeping only linear terms, we find the
equation (for a < 0)
U= lalu, (1.13.3)
Because I a I = - a > 0, u(t) grows exponentially. This indicates that the state
qo = 0 is unstable. In Chaps. 2, 3 linear stability analysis will be quite generally
presented. In particular, we shall study not only the case in which a constant qo
becomes unstable, but also the case in which motions on a limit cycle or on a
torus become unstable. The latter problem leads to the rather strange country of
quasiperiodic motions where still a great many discoveries can be made, to which
this book has made some contribution. After having performed the stability ana-
lysis we ask the question to which new states the system will go. In this context
two concepts are central to synergetics, namely the concept of the order para-
meter and that of slaving. Consider to this end the following two differential
equations
iII = Alql-qlq2, (1.13.5)
(1.13.6)
437
36 1. Introduction
(1.13.8)
This approach, which is often called the adiabatic approximation, allows us
to express q2 explicitly by q]. Or, in other words, q2 is slaved by q]. When we are
dealing with very many variables, which are slaved by q], we may reduce a com-
plex problem quite considerably. Instead of very many equations for the q's we
need to consider only a single equation for q], and then we may express all other
q's by q] according to the slaving principle. We shall call such a q] an order
parameter. Of course, in reality the situation may be more difficult. Then
(1.13.5, 6) must be replaced by equations which are much more involved. They
may depend on time-dependent coefficients and they may contain fluctuating
forces. Furthermore (1.13.7) and therefore (1.13.8) are just approximations.
Therefore it will be an important task to devise a general procedure by which one
may express q2 by q]. We shall show in Chap. 7 that it is indeed possible to
express q2(t) by q] (t) at the same time t, and an explicit procedure will be devised
to find the function q2(t) = f(q] (t» explicitly for a large class of stochastic non-
linear partial differential equations. There is a most important internal relation
between the loss of linear stability, the occurrence of order parameters and the
validity of the slaving principle. When we change control parameters, a system
may suffer loss oflinear stability. As can be seen from (1.13.5,6), in such a case
Re{Ad changes its sign, which means that it becomes very small. In such a situa-
tion the slaving principle applies. Thus we may expect that at points where struc-
tural changes occur, the behavior of a system is governed by the order parameters
alone. As we shall see, this connection between the three concepts allows us to
establish far-reaching analogies between the behavior of quite different systems
when macroscopic changes occur.
We assume that N does not explicitly depend on time, i. e., we are dealing with
autonomous systems. Ignoring any spatial dependence, we start with equations
of the form
We assume that for a certain value (or a range of values) of the control parameter
a, a stable solution qo(t, a) of (1.14.2) exists. To study the stability of this solu-
438
1.14 Qualitative Changes: Typical Phenomena 37
where wet) is assumed to be small. Inserting (1.14.3) into (1.14.2) and keeping
only terms linear in w, we arrive at equations of the form
w=Lw, (1.14.4)
where L depends on qo. In Sect. 1.14.6 it will be shown explicitly how L is con-
nected with N. For the moment it suffices to know that L bears the same time
dependence as qo.
Let us assume that qo is time independent. To solve (1.14.4) we make the
hypothesis
w(t) = eAtv, v constant vector, (1.14.5)
which is a linear algebraic equation for the constant vector v and the eigenvalue
A.
To elucidate the essential features of the general approach, let us assume that
the matrix L is of finite dimension n. Then there are at most n different eigen-
values Aj' which in general depend on the control parameter a. An instability
occurs if at least one eigenvalue acquires a nonnegative real part.
1.14.1 Bifurcation from One Node (or Focus) into Two Nodes (or Foci)
Let us treat the case in which Re{A) ;::: 0 just for one j, say j = 1, and Re{AJ < 0
for all other j. This example illustrates some of the essential features of our
approach, an overview of which will be presented later. Since our final goal is to
solve the fully nonlinear equation (1.14.2), we must make a suitable hypothesis
for its solution q(t). To this end we repesent q(t) in the following form
(1.14.7)
where Vj are the solutions of (1.14.6), while <j are still unknown time-dependent
coefficients. Inserting (1.14.7) into (1.14.2), we may find after some manipula-
tions (whose explanation is not important here, cf. Chap. 8) equations for t;j. Let
us take for illustration just j = 1, 2. The corresponding equations read
Here AI' A2 are the eigenvalues of (1.14.6), while Nj are nonlinear functions of t;1,
439
38 1. Introduction
¢"2 which start at least with terms quadratic (or bilinear) in ¢"t, ¢"2. Now remember
that we are close to a control parameter value a where the system loses linear
stability, i. e., where Re{At} changes sign. But this means that IAtl <t!; IA21 so that
the slaving principle applies. Therefore we may express C;2 by ¢"1' C;2 =!(¢"1), and
need solve only a single equation of the form
(1.14.10)
(1.14.11)
(1.14.12)
But when confining to real At this is precisely our former equation (1.13.1)
which describes the sliding motion of a ball in a vase with one valley (At < 0) or
two valleys (At> 0). Therefore the single node present for At < 0 is replaced by
two nodes for At> 0 (Figs. 1.13.5 and 6). Or, in other words, the single node
bifurcates into two nodes (Fig. 1.14.1). It is worth mentioning that (1.14.12) not
only describes the new equilibrium position, but also the relaxation of the system
into these positions and thus allows their stability to be verified. This example
may suffice here to give the reader a feeling how the nonlinear equation (1.14.6)
can be solved and what kind of qualitative change occurs in this case.
From now on this section will deal with classification and qualitative
description of the phenomena at instability points. The detailed mathematical
approaches for their adequate treatment will be presented in later chapters. For
instance, in Sect. 8.3 we shall discuss more complicated problems in which
several eigenvalues, which are real, become positive.
A famous example of bifurcation is the Hop! bifurcation. In this case two com-
plex conjugate eigenvalues
440
1.14 Qualitative Changes: Typical Phenomena 39
q,lt)
Fig. 1.14.4. The temporal behavior of a variable ql (t) of the limit cycle before bifurcation (left-hand
side) and after bifurcation (right-hand side). The variables ql belonging to the new stable limit cycles
are shown as solid and dashed-dotted lines, respectively. The unstable limit cycle is represented by a
dashed line
441
40 1. Introduction
--- t
This new motion of the solution vector q (I) can be visualized as that of a motion
on a torus (Figs. 1.14.5 and 6). In other words, we are now dealing with a quasi-
periodic motion. When a limit cycle becomes unstable, other phenomena may
occur also. The limit cycle can be replaced by another one where the system needs
twice the time to return to its original state or, to put it differently, the period has
doubled, or a subharmonic is generated. There are a number of systems known,
ranging from fluid dynamics to electronic systems, which pass through a
hierarchy of subsequent period doublings when the control parameter is
changed. Figures 1.14.7 -12 survey some typical results.
The phenomenon of period doubling or, in other words, of subharmonic
generation, has been known for a long time in electronics. For instance, a
number of electronic circuits can be described by the Duffing equation
(1.14.14)
Fig. 1.14.7. Projection (from the q I' q2' I space) of a trajectory on the q I ' q 2 plane of a traject ory of
the Duffing equation x + kx + x 3 = A cost!, where q l == X, q2 == X. The control parameter va lues
are fixed at k = 0.35 , A = 6.6. In this case a limit cycle occurs
Fig. 1.14.8. Solution of the same equation as in Fig. 1.14.7 but with k = 0.3 5, A = 8.0. The period
of evolution is now twice as large as that of Fi g. 1.14.7 but again a closed orbit occurs (period
doubling)
442
1.14 Qualitative Changes: Typical Phenomena 41
~:F,
Fig. 1.14.9. Same equation as in Figs. 1.14.7
and 8 with k = 0.35, A = 8.5. The period is 4
times as large as that of the original limit cycle
of Fig. 1.14.7
/ / I 1I,
I
Fig. 1.14.10
,
q It)
Fig. 1.14.10. Temporal evolution of co-
ordinate ql belonging to the case of Fig. 1.14.7
Hg. 1.14.11. Temporal evolution of the co-
ordinate q 1 after period doubling has oc-
curred. Note that two subsequent minima have
different depths
Fig. 1.14.12. Temporal evolution of ql be-
longing to Fig. 1.14.9. Close inspection of the
lower minima reveals that the cycle is repeated
only after a time 4 times as big as in Fig.
Fig. 1.14.12 1.14.10
Even this still rather simple equation describes a variety of phenomena including
period doubling and tripling. But other subharmonics may also be present. A
whole sequence of period doublings or triplings may occur, too. It seems that
here we are just at the beginning of a new development, which allows us to study
not only one or a few subsequent bifurcations, but a whole hierarchy. A warning
seems to be necessary here. While some classes of equations (certain "discrete
maps", cf. Sect. 1.17) indicate a complete sequence of period doubling, real
systems may be more complicated allowing, e. g., both for doubling and tripling
sequences or even for mixtures of them.
443
42 1. Introduction
dimensions has been studied. Here quite peculiar difficulties arise. The mathe-
matical analysis shows that a rather strange condition on the relative irrationality
of the involved basic frequencies WI, W2, ... of the system plays a crucial role.
Many later sections will be devoted to this kind of problem. Some aspects of this
problem are known in celestial mechanics, where the difficulties could be
overcome only about one or two decades ago. While in celestial mechanics
scientists deal with Hamiltonian, i. e., nondissipative, systems and are concerned
with the stability of motion, here we have to be concerned with the still more
complicated problem of dissipative, pumped systems, so that we shall deal in
particular with qualitative changes of macroscopic properties (in particular the
bifurcation from a torus into other tori).
Fig. 1.14.13. How the power spectrum [(wi reveals frequency locking. (Left-hand side) In the
unlocked state the system oscillates at two fundamental frequencies Wl and W2. (Right-hand side)
The power spectrum in case of a locked state. The two former frequencies Wl and Wz ( - - - ) have
disappeared and are replaced by a single line at frequency Wo
The results of all these studies are of great importance for many problems in
the natural sciences and other fields, because "motion on a torus" means for a
concrete system that it exerts a motion at several fundamental frequencies and
their suitable linear combinations. In many cases such a motion is caused by a
system of nonlinear oscillators which occur in nature and in technical devices
quite frequently. It is then important to know in which way such a system
changes its behavior if a control parameter is changed.
444
1.14 Qualitative Changes: Typical Phenomena 43
q=N(q). (1.14.15)
In this case the only attract or possible is a stable fixed point (a "one-dimensional
node") and the "trajectory" of this attract or is just a constant, q = qQ, denoting
the position of that fixed point. In order to prove the stability of this point, we
make a linear stability analysis introduced in Sect. 1.13. To this end we insert
into (1.14.15), linearize this equation with respect to oq, and obtain
d
-oq=Loq, (1.14.17)
dt
where L = aNlfJq Iq=qo is a constant.
The solution of (1.14.17) reads, of course,
If L is negative, the fixed point is stable. In this trivial example, we can directly
read off L from (1.14.18). But in more complicated cases to be discussed below,
only a computer solution might be available. But even in such a case, we may
derive L from (1.14.18) by a simple prescription. Namely, we form
1
-In loq(t) I
t
L = lim~lnloq(t)I. (1.14.19)
1_00 t
445
44 1. Introduction
so that f5q(t) "tests" in which way the neighboring trajectory q(t) behaves, i.e.
whether it approaches qo(t) or departs from it. In order to determine f5q(t) we
insert (1.14.20) into the nonlinear equation
(1.14.22)
A= lim~lnlf5q(t)I. (1.14.23)
t~oo t
(1.14.24)
Clearly, if Al =1= A2' the limit (1.14.23) does not exist. Therefore one has to refine
the definition of the Lyapunov exponent. Roughly speaking, one whishes to
select the biggest rate A, and one therefore replaces "lim" in (1.14.23) by the
"limes superior", or in short lim sup, so that (1.14.23) is replaced by
We shall give the precise definition of lim sup in Chap. 2 where we shall present a
theorem on the existence of Lyapunov exponents, too. Depending on different
initial values of f5q at t = to, different Lyapunov exponents may exist, but not
more than m different ones, if m is the dimension of the vector space of N (or q).
We are now able to present the criterium anounced above, which helps us to
distinguish between different kinds of attractors.
446
1.14 Qualitative Changes: Typical Phenomena 45
In one dimension, there are only stable fixed points, for which the Lyapunov
exponents A are negative ( - ). In two dimensions, the only two possible classes of
attractors are stable fixed points, or limit cycles, as is proven rigorously in
mathematics. In the case of a stable fixed point (focus) the two Lyapunov
exponents (AI, ,1.2) (which may coincide) are negative ( - , -). In the case of a
stable limit cycle, the Lyapunov exponents belonging to a motion oq transversal
to the limit cycle qo(t) is negative (stability!), whereas the Lyapunov exponent
belonging to oq in tangential direction vanishes, as we shall demonstrate in a
later section. Therefore (AI, ,1.2) = (-,0). There may be a "pathological" case,
for which (AI, ,1.2) = (-,0), but no limit cycle present, namely if there is a line of
fixed points.
Finally we discuss typical cases in three dimensions. In each case we assume
an attractor qo, i. e. Iqo Iremains bounded for t ---+ 00.
Neighboring trajectories of the limit cycle can approach the limit cycle from two
linearly independent directions transversal to the limit cycle so that (AI' ,1.2) =
( - , -), whereas the third Lyapunov exponent corresponding to a shift of the
trajectory in tangential direction is equal to zero.
stable torus.
The discussion is similar to the case of the limit cycle. (There are still some
subtleties analogous to the "pathological" case just mentioned.)
Chaos may occur, if one Lyapunov exponent becomes positive. But at any
rate, some more discussion is needed. For instance (}'1, ,1.2' ,1.3) = (+,0,0) may
mean we are dealing with an unstable torus (i. e. no attractor). If an attractor
possesses the exponents (,1.1' ,1.2' ,1.3) = (+ ,0, -), it is considered as a chaotic
attractor, (,1.1' ,1.2' ,1.3) = (+, +,0) may mean an unstable limit cycle (i. e. no
attract or) etc. Because in a chaotic attract or at least one Lyapunov exponent is
positive, neighboring trajectories depart very quickly from each other. But since
neighboring trajectories stem from initial conditions which differ but slightly we
recognize that the phenomena described by a chaotic attractor sensitively depend
on the initial conditions. It might be noteworthy that the research on Lyapunov
exponents - both what they mean for attractors and how they can be determined
- is still under its way.
In conclusion we mention the following useful theorem: If q(t) is a trajectory
which remains in a bounded region (e. g. the trajectory of an attract or) and if it
does not terminate at a fixed point, at least one of its Lyapunov exponents
vanishes. We shall present a detailed formulation of this theorem and its proof in
Sect. 2.4.
447
46 1. Introduction
Fig. 1.15.1. The distribution function f(q) Fig. 1.15.2. The distribution function f(q)
( ~ ~ - ) belonging to a node. The solid line ( - - - ) belonging to two nodes and a saddle
represents the potential in which the "particle" point (in one dimension). The solid line repre-
with coordinate q moves sents the potential in which the "particle"
moves
448
1.16 Evolution of Spatial Patterns 47
w=Lw. (1.16.1)
449
48 1. Introduction
(1.16.4)
Again we may identify order parameters c,j' for which Re{AJ ~ O. We may apply
the slaving principle and establish order parameter equations. Confining (1.16.4)
to the leading terms, i. e., the order parameters c,j == Uj alone, we obtain the
skeleton of the evolving patterns. For instance, if only one Re{Aj} ~ 0, the
"skeleton" reads
(1.16.5)
(1.16.6)
we obtain the following result: while for A1 < 0 only the solution u j = 0 and
therefore q(x, t) = const and such that the homogeneous distribution is stable,
for A1 > 0 the solution Uj =1= 0 and therefore the spatially inhomogeneous solution
(1.16.5) arises. This approach allows us to study the growth of a spatial pattern.
If several order parameters Uj =1= 0 are present, the "skeleton" is determined by
But in constrast to linear theories the u/s cannot be chosen arbitrarily. They are
determined by certain nonlinear equations derived later and which thus deter-
mine the possible "skeletons" and their growths. In higher approximation the
slaved modes contribute to the spatial pattern also. An important difference
between these transitions and phase transitions of systems in thermal equilibri-
um, where long-range order is also achieved, should be mentioned. With few
exceptions, current phase transition theory is concerned with infinitely extended
media because only here do the singularities of certain thermodynamic functions
(entropy, specific heat, etc.) become apparent. On the other hand, in nonequi-
librium phase transition, i. e., in the transitions considered here, in general the
finite geometry plays a decisive role. The evolving patterns depend on the shape
450
1.17 Discrete Maps. The Poincare Map 49
of the boundaries, for instance square, rectangular, circular, etc., and in addition
the patterns depend on the size of the system. In other words, the evolving
patterns, at least in general, bring in their own length scales which must be
matched with the given geometry.
With respect to applications to astrophysics and biology we have to study the
evolution of patterns not only in the plane or in Euclidian space, but we have to
consider that evolution also on spheres and still more complicated manifolds.
Examples are the first stages of embryonic evolution or the formation of atmos-
pheric patterns on planets, such as Jupiter. Of course, in a more model-like
fashion we may also study infinitely extended media. In this case we find phe-
nomena well known from phase transition theory and we may apply renormali-
zation group techniques to these phenomena.
Within active media such as distributed chemical reactions, interaction of
cells of a tissue, neuronal networks, etc., combined spatio-temporal patterns may
evolve also. The methods described later allow a mathematical treatment of large
classes of these phenomena.
In the previous sections a brief outline was given of how to model many processes
by means of evolution equations of the form
451
50 1. Introduction
Of course, in order to find this connection, one has to integrate (1.17.1) over a
time interval from tn' where Xn is hit, to tn+ I> where xn+ 1 is hit. Thus no simpli-
fication seems to be reached at all. The following idea has turned out to be most
useful. Let us consider (1.17.2) for all n and a fixed function f in terms of a
model! Because we no longer need to integrate (1.17.1) over continuous times,
we expect to gain more insight into the global behavior of xn-
Equation (1.17.2) is called a map because each value of Xn is mapped onto a
value of xn+l' This can be expressed in form of a graph (Fig. 1.17.2). Probably
the simplest graph which gives rise to nontrivial results is described by the
"logistic" map which reads
(1.17 .3)
452
1.17 Discrete Maps. The Poincare Map 51
However, a good insight into the behavior of the solutions of (1.17 .3) can be
gained by a very simple geometric construction explained in Figs. 1.17.4 - 9. The
reader is strongly advised to take a piece of paper and to repeat the individual
steps. As it transpires, depending on the numerical value of (1, different kinds of
behavior of xn result. For (1 < 3, Xn converges to a fixed point. For 3 < (1 < (12, Xn
converges to a periodic hopping motion with period T= 2 (Figs. 1.17.10, 11).
For (12 < (1 < (13' xn and tn big enough, xn comes back to its value for n + 4,
n + 8, ... , i. e., after a period T = 4, T = 8, ... (Figs. 1.17.12, 13). When we plot
the values of xn reached for n ..... 00 for various values of (1, Fig. 1.17.14 results.
Fig. 1.17.4. The series of Figs. 1.17.4 - 8 shows Fig. 1.17.5. Valuex 2 found in Fig. 1.17.4 serves
how one may derive the sequence X n ' now as a new initial value, obtained by project-
(n = 2, 3, 4, ... j, once one initial value Xl is ing x 2 in Fig. 1.17.4 or in the present figure on
given. Because the parabola represents the map- the diagonal straight line. Going from the
ping function, we have to go from Xl by a corresponding cross section downward, we find
vertical line until we cross the parabola to find x 2 x 2 as value on the abscissa. In order to obtain the
new value Xl we go vertically upwards to meet
the parabola
Fig. 1.17.6. To proceed from Xl to x 4 we re- Fig. 1.17.7. The construction is now rather
peat the steps done before, i. e., we first go obvious and we merely show how we may
horizontally from Xl till we hit the diagonal proceed from x 4 to Xs
straight line. Going down vertically we may
find the value Xl' but on the way we hit the
parabola and the cross section gives us x 4
453
52 1. Introduction
Xn 2 :0 4 5 6
Fig. 1.17.8. When we repeat the above steps Fig. 1.17.9. The individual values of Xl' X 2 •
many times we approach more and more a Xl' ... constructed by means of the previous
limiting point which is just the cross section figures. The resulting plot is shown where X n
of the parabola with the diagonal straight approaches more and more the dashed line
line
- ~/\=
i ~l
' - - . , , - - - , - - , - - - , - - - - - , - - , . . - - - - ---,------y------+ n
234 557 8
Fig. 1.17.12. The case of period 4 Fig. 1.17.13. The case of period 4. Plot (lfx"
versus n
454
1.17 Discrete Maps. The Poincare Map 5:1
Fi~. 1.17.14. The set of possible vailies of xn (n -> 00) (ordinate) versus control parameter 1100 - 11
(abscissa) on a logarithmic scale. Here the logistic equation is linearly transformed into the equation
xn t I = 1 - 11X~. 1100 corresponds to the critical value am. [After P. Collet, .I. P. Eckmann: In Pro-
qress in Pyhsics, Vol. 1, ed. by A. Jaffe, D. Ruelle (Birkhauser, Boston 1980)]
455
54 1. Introduction
(1.17.5)
where x n ' n = 1, 2, ... are vectors in an M-dimensional space. Here f may depend
on one or several control parameters a which allows us to study qualitative
changes of the "discrete dynamics" of xn when a is changed. A wide class of such
qualitative changes can be put in analogy to nonequilibrium phase transitions,
treated by the conventional evolution equations (1.17.1). Thus, we find critical
slowing down and symmetry breaking, applicability of the slaving principle (for
discrete maps), etc. These analogies become still closer when we treat discrete
noisy maps (see below). Equations (1.17 .5) enable many applications which so
far have been studied only partially. For instance, the state vector xn can
symbolize various spatial patterns.
Fig. 1.17.15. Poincare map belonging to the Fig. 1.17.16. The intersection points of Fig.
cross section of a two-dimensional plane with a 1.17.15
trajectory in three dimensions
(1.17.6)
and insert it into (1.17.5). We may expand f(x~ + oXn) into a power series with
respect to the components of oXn-
456
1.17 Discrete Maps. The Poincare Map 55
(1.17.7)
(1.17.8)
(1.17.9)
(1.17.10)
where the L's are multiplied with each other by the rules of matrix multiplication.
The Lyapunov exponents are defined by
Depending on different directions of oXo, different A'S may result (in multi-
dimensional maps).
For a one-dimensional map we may represent A in a rather explicit way. In
this case the L's and oXo are mere numbers so that according to (1.17.10)
Inserting this latter expression into (1.17.11) and using (1.17.8) we obtain
1 n-I
A = lim sup - L In If'(x~) I (1.17.13)
n_oo nm=O
457
56 1. Introduction
1.5
(1.18.1)
Also more complicated maps can be formulated and treated, e. g., maps for
vectors X n , and for fluctuations which depend on xn (Chap. 11).
In all the cases considered in this book the temporal, spatial or spatio-temporal
patterns evolve without being imposed on the system from the outside. We shall
call processes, which lead in this way to patterns, "self-organization". Of course,
in the evolution or functioning of complex systems, such as biological systems, a
whole hierarchy of such self-organizing processes may take place. In a way, in
this book we are considering the building blocks of self-organization. But in con-
trast to other approaches, say by molecular biology which considers individual
molecules and their interaction, we are mostly concerned with the interaction of
many molecules or many subsystems. The kind of self-organization we are
considering here can be caused by different means. We may change the global
impact of the surroundings on the system (as expressed by control parameters).
Self-organization can be caused also by a mere increase of number of
components of a system. Even if we put the same components together, entirely
new behavior can arise on the macroscopic level. Finally, self-organization can
be caused by a sudden change of control parameters when the system tries to
relax to a new state under the new conditions (constraints). This aspect offers us
458
1.19 Pathways to Self-Organization 57
a very broad view on the evolution of structures, including life in the universe_
Because the universe is in a transient state, starting from its fireball, and subject
to expansion, ordered structures can occur.
In the context of this book we are interested in rigorous mathematical for-
mulations rather than philosophical discussions. Therefore, we briefly indicate
here how these three kinds of self-organization can be treated mathematically.
ti(2) = N(2)(q(2» .
Let us assume that these equations allow stable inactive states described by
q~) = 0, j = 1, 2. (1.19.3)
(1.19.4)
(1.19.5)
We now have to deal with the total system 1 + 2 described by the equation
q = (q(1»)
(2) and (1.19.7)
q
N(1) + aK(1) )
N(q, a) = ( (2) (2)' (1.19.8)
N +aK
We havc introduced a parameter a which ranges from 0 till 1 and which plays the
459
58 1. Introduction
where v (x) describes some spatial order, and the order parameter equation
it =AU. (1.19.11)
occurs. It clearly describes some structure, but it does not tend to a new stable
state.
This approach hides a deep philosophical problem (which is inherent in all
° °
cases of self-organization), because to get the solution (1.19.12) started, some
fluctuations must be present. Otherwise the solution u == and thus q == will
persist forever.
In this book we shall focus our attention on situations where systems undergo
qualitative changes. Therefore the main line of our procedure is prescribed by
1) study of the loss of stability
2) derivation of the slaving principle
3) establishment and solution of the order parameter equations.
460
1.20 How We Shall Proceed 59
To this end we first study the loss of stability. Therefore Chaps. 2 and 3 deal
with linear differential equations. While Chap. 2 contains results well known in
mathematics (perhaps represented here, to some extent, in a new fashion),
Chap. 3 contains largely new results. This chapter might be somewhat more dif-
ficult and can be skipped during a first reading of this book.
Chap. 4 lays the basis for stochastic methods which will be used mainly in
Chap. 10. Chaps. 5 and 6 deal with coupled nonlinear oscillators and study
quasiperiodic motion. Both Chaps. 5 and 6 serve as a preparation for Chap. 8,
especially for its Sects. 8.8 -11. Chap. 6 presents an important theorem due to
Moser. In order not to overload the main text, Moser's proof of his theorem is
postponed to the appendix. Chap. 7 then resumes the main line, initiated by
Chaps. 2 and 3, and deals with the slaving principle (for nonlinear differential
Table 1.20.1. A survey on the relations between Chapters. Note that the catchwords in the boxes
don't agree with the chapter headings
r
Chap. 3 linear equations
461
60 1. Introduction
equations) without and with stochastic forces. This chapter, which contains new
results, is of crucial importance for the following Chaps. 8 and 9, because it is
shown how the number of degrees of freedom can be reduced drastically.
Chaps. 8 and 9 then deal with the main problem: the establishment and solution
of the order parameter equations. Chap. 8 deals with discrete systems while
Chap. 9 is devoted to continuously extended media. Again, a number of results
are presented here for the first time. Chap. 10 treats the impact ofjluctuations on
systems at instability points. The subsequent chapter is devoted to some general
approaches to cope with discrete noisy maps and is still in the main line of this
book.
Chap. 12 is out of that line - it illustrates that even seemingly simple ques-
tions put in dynamic systems theory cannot be answered in principle. Finally
Chap. 13 resumes the theme of the introductory chapter - what has synergetics
to do with other disciplines. In a way, this short chapter is a little excursion into
epistemology. In conclusion it is worth mentioning that Chaps. 2, - 6 are of use
also in problems when systems are away from their instability points.
462
2. Linear Ordinary Differential Equations
C= qo :=q(O). (2.1.4)
Then (2.1.2) can be written as
As shown in Sects. 1.14, 16, linear differential equations are important to deter-
mine the stability of solutions of nonlinear equations. Therefore, we shall discuss
here and in the following the time dependence of the solutions (2.1.5) for large
times t. Obviously for t > 0 the asymptotic behavior of (2.1.5) is governed by the
sign of Re{A}. For Re{A} > 0, Iqlgrows exponentially, for Re{A} = 0, q(t) is a con-
stant, and for Re{A} < 0, Iq I is exponentially damped. Here A itself is called a
characteristic exponent.
q=a(t)q, (2.1.6)
00 •
To study the asymptotic behavior of (2.1.7), we insert (2.1.8) into the integral
occurring in (2.1. 7) and obtain
464
2.1 Examples of Linear Differential Equations: The Case of a Single Variable 63
t C .
Ia(r)dr= tco+ L -._n_(e lnwt _1). (2:1.9)
o n*O Inw
With n =1= 0, the sum in (2.1.9) converges at least as well as the sum in (2.1.8).
Therefore the sum in (2.1.9) again represents a periodic function. Thus the
asymptotic behavior of (2.1.7) is governed by the coefficient Co in (2.1.8).
Depending on whether its real part is positive, zero, or negative, we find
exponential growth, a neutral solution or exponential damping, respectively.
Combining our results (2.1.7 - 9), we may write the solution of (2.1.6) in the
form
(2.1.10)
(2.1.11)
q= a(t)q (2.1.12)
a(t) is quasiperiodic, i. e., we assume that a(t) can be expanded into a multiple
Fourier series
(2.1.14)
465
64 2. Linear Ordinary Differential Equations
tion]. Or, in other words, we exclude w/ s which are rational with respect to each
other. The formal solution of (2.1.12) again has the form (2.1.7); it is now appro-
priate to evaluate the integral in the exponent of (2.1.7). If the series in (2.1.13)
converges absolutely we may integrate term by term and thus obtain
t C
Ja(r)dr = cot + L . m [exp (im· wt) -1) . (2.1.15)
° m*O 1m· w
In order that the sum in (2.1.15) has the same form as (2.1.13), i.e., that it is a
quasiperiodic function, it ought to be possible to split the sum in (2.1.15) into
\' ----"'--exp
f..,. em (.1m wt ) + canst.
0 (2.1.16)
m*O 1m w 0
const = - L . em (2.1.17)
m*O 1m 0 w
and similarly the first sum in (2.1.16) need not converge. Why is this so? The
reason lies in the fact that because the m's can take negative values there may be
combinations of m's such that
mow--O (2.1.18)
for
Iml-- 00. (2.1.19)
[Condition (2.1.18) may be fulfilled even for finite m's if the ratios of the w's are
rational numbers, in which case mow = 0 for some m = mo. But this case has
been excluded above by our assumptions on the w/s.] One might think that one
can avoid (2.1.18) if the w's are irrational with respect to each other. However,
even in such a case it can be shown mathematically that (2.1.18, 19) can be fulfil-
led. Because mow occurs in the denominator, the series (2.1.18) need not con-
verge, even if L Iem Iconverges. Thus the question arises under which conditions
m
the series (2.1.17) still converges. Because em and mow occur jointly, this condi-
tion concerns both Cm and the w/s. Loosely speaking, we are looking for such cm
which converge sufficiently quickly to zero for 1m 1-- 00 and such w/s for which
1mow 1 goes sufficiently slowly to zero for 1m I -- 00 so that (2.1.17) converges.
Both from the mathematical point of view and that of practical applications, the
condition on the w/s is more interesting so that we start with this condition.
Loosely speaking, we are looking for such w/s which are "sufficiently irra-
tional" with respect to each other. This condition can be cast into various mathe-
matical forms. A form often used is
466
2.1 Examples of Linear Differential Equations: The Case of a Single Variable 65
(2.1.21)
Here K is a constant. Though for 11m 11--- 00, (2.1.20) again goes to 0, it may tend
°
to sufficiently slowly. We shall call (2.1.20,21) the Kolmogorov-Arnold-Moser
condition or, in short, the KAM condition. When a real system is given, the ques-
tion arises whether the frequencies occurring meet the KAM condition (2.1.20,
21). While in mathematics this seems to be a reasonable question, it is proble-
matic to decide upon this question in practice. Furthermore, since systems are
subject to fluctuations, it is even highly doubtful whether a system will retain its
frequencies such that a KAM condition is fulfilled for all times. Rather, it is
meaningful to ask how probable it is that the w's fulfill that condition. This is
answered by the following mathematical theorem (which we do not prove here):
°
In the space W = (WI' W2,"" wn) the relative measure of the set of those w's
which do not satisfy the condition (2.1.20) tends to together with K. Thus for
sufficiently small K most of the w's satisfy (2.1.20).
We now discuss the second problem, namely the convergence rate of the coef-
ficients cm in (2.1.13). Because simple Fourier series can be handled more easily
than multiple Fourier series of the form (2.1.13), we try to relate (2.1.13) with
simple Fourier series. This is achieved by the following trick. We introduce
auxiliary variables <PI' ... , <Pn and replace a(t) in (2.1.13) by
a(t, <PI' <P2,···, <Pn) = Co + L cm exp (iml WI <PI + ... + imnwn <Pn + im • wt) ,
m,*,O (2.1.22)
f(x) = ~ am e imx ,
L... (2.1.23)
m=-oo
whose derivatives up to order (h -1) are continuous, and whose h'th derivative is
piecewise continuous, then
c (2.1.24)
laml~ --h'
Iml
Consider (2.1.22), and assume that its derivatives fulfill the just-mentioned con-
ditions for all <P/s. Then
(2.1.25)
467
66 2. Linear Ordinary Differential Equations
After these preparations we are able to demonstrate how the KAM condition
works. To this end we study the convergence of (2.1.17). We start with (L I
denotes m =1= 0)
(2.1.27)
e
I I
n+l
~I Cm ~ _nn+l ~I mmax (2.1.28)
L... • '" L... h h •
m 1m·W K m Imll ... lmnl
(2.1.29)
it = co' (2.1.30)
where
468
2.1 Examples of Linear Differential Equations: The Case of a Single Variable 67
O:;;;:;t<oo, (2.1.33)
so that
la(t)I:;;;:; B. (2.1.34)
We wish to study how solution (2.1.35) behaves when t goes to infinity. In partic-
ular, we shall study whether (2.1.35) grows or decays exponentially. As a first
step we form
t
In Iq(t) I = In Iq(O) 1+ Sa(r)dr, (2.1.36)
o
so that we have to discuss the behavior of the integral in (2.1.36). To this end we
first explain the notion of the supremum which is also called the "least upper
bound". If A is a set of real numbers {a}, the supremum of A is the smallest real
number b such that a < b for all a. We write "sup {A}" to denote the supremum.
A similar definition can be given for the infimum or, in other words, for the
greatest lower bound (inf {A}). If A is an infinite set of real numbers then the
symbol "lim sup {A}" denotes the infimum of all numbers b with the property
that only a finite set of numbers in A exceed b. In particular if A is a sequence
{an} then lim sup {A} is usually denoted by lim sup {an}.
n~oo
These definitions are illustrated in Fig. 2.1.1. Starting from (2.1.36) we now
form
fb}
469
68 2. Linear Ordinary Differential Equations
lim sup \
t~oo l~t afa(r)dr]. (2.1.37)
1
-In Iq(O) I..... 0 for t ..... 00 • (2.1.38)
t
t t
Sa(r)dr ~ SB dr = Bt (2.1.39)
a a
and
t
Sa(r)dr? -Bt. (2.1.40)
a
lim sup
t~oo
\~Ja(r)dr]=
II a
A (2.1.41 )
exists with
lim sup \
hoo (I
~f(t)J .. O. (2.1.44)
Thus the generalized characteristic exponents have the same significance as the
real part of the (characteristic) exponent A in the simple case of a differential
equation with constant coefficients.
470
2.2 Groups and Invariance 69
The very simple example of the differential equation (2.1.6) allows us to explain
some fundamental ideas about groups and invariance. Let us start with the
differential equation for the variable q(t)
The properties of a(t) (i. e., being constant, periodic, or quasiperiodic) can be
characterized by an invariance property also. Let us assume that a(t) remains un-
altered if we replace its argument
i. e., (2.2.2)
If to can be chosen arbitrarily, then (2.2.3) implies that a(t) must be time inde-
pendent. If (2.2.3) is fulfilled for a certain to (and for integer multiples of it), a(t)
is periodic. As shown in Chap. 3 our present considerations can also be extended
to quasiperiodic functions a(t), but this requires more mathematics. Now let us
subject (2.2.1) to the transformation (2.2.2), which yields
so that (2.2.4) has exactly the same appearance as (2.2.1). It is the same differen-
tial equation for a variable which we may denote by
(2.2.5)
However, it is known from the theory of linear differential equations that the
solution of (2.2.1) is unique except for a constant factor. Denoting this constant
factor by a, the relation
) (2.2.6)
must hold. We want to show that by means of relation (2.2.6) we may construct
the solution of (2.2.1) directly. To this end let us introduce the translation
operator T, defined as follows. If Tis applied to a functionf of time, we replace t
by t + to in the argument of f
471
70 2. Linear Ordinary Differential Equations
Since this relation holds for an arbitrary q(t), (2.2.8) can be written in the form
or, in other words, Tcommutes with a(t). Equation (2.2.6) can now be expressed
in the form
Tq = aq. (2.2.10)
The operations Tn, n ~ 0, form a (multiplicative) group, because they fulfill the
following axioms.
Individually, Tn, n ~ 0 are elements of the group.
1) If we mUltiply (or, more generally speaking, combine) two elements with one
another, we obtain a new element of the group. Indeed Tn. T m = T n+ m is a
new element. This relation holds because a displacement by n to and one by
m to yields a new displacement by (n + m) to.
2) A (right-sided) unity operator E exists so that Tn. E = rn.
472
2.2 Groups and Invariance 71
(2.2.11)
(2.2.12)
(2.2.13)
(2.2.14)
473
72 2. Linear Ordinary Differential Equations
T -1
--+ a -1 , (2.2.16)
The relations (2.2.16) are probably the simplest example of a group representa-
tion. The basis of that representation is formed by q and a certain number, an, is
attached to each operation Tn so that the numbers an obey the same rules as the
operators Tn as described by the above axioms (1)-(4), i.e.,
TnT m = T n+ m
TnE= Tn
(2.2.17)
T-nTn=E --+a- n a n =1
(TnTm) TI = Tn(Tm TI) --+ (an am) a l = an(a mal).
We shall see later that these concepts can be considerably generalized and will
have important applications.
Again it is our aim to get acquainted with general properties of such an equation.
First let us treat the case in which a == 0 so that (2.3.1) reduces to
q=b (2.3.2)
t
q(t) = Sb(r)dr + c, (2.3.3)
to
474
2.3 Driven Systems 73
c = q(to) . (2.3.4)
To the integral in (2.3.3) we may apply the same considerations applied in Sects.
2.1.1- 3 to the integral as it occurs, for instance, in (2.1.7). If b(t) is constant or
periodic,
and c(t) is a still unknown function, which would be a constant if we were to solve
the homogeneous equation (2.3.7) only.
Inserting (2.3.6) into (2.3.1) we readily obtain
(2.3.8)
(2.3.9)
t
c(t) = Jqol(r)b(r)dr+ a. (2.3.10)
to
Again, it is our main goal to study what types of solutions from (2.3.1) result if
certain types of time dependence of a(t) and b(t) are given. Inserting (2.3.10)
into (2.3.6) we obtain the general form of the solution of (2.3.1), namely
t
q(t) = qo(t) Jqo 1 (r) b( r) dr + aqo(t) . (2.3.11)
to
a == A. = const
b = const
'* 0 I. (2.3.12)
475
74 2. Linear Ordinary Differential Equations
or after integration
b
q(t) = -_+pe At (2.3.14)
A
with
For A> 0, the long-time behavior (t-+ 00) of (2.3.14) is dominated by the
exponential function and the solution diverges provided p does not vanish, which
could occur with a specific initial condition. For A < 0 the long-time behavior is
determined by the first term in (2.3.14), i.e., q(t) tends to the constant -blA.
We can find this solution more directly by putting q = 0 in (2.3.1) which yields
q = -b/A. (2.3.16)
The fully time-dependent solution (2.3.14) is called a transient towards the sta-
tionary solution (2.3.16). The case A = 0 brings us back to the solution (2.3.5)
discussed above.
Quite generally we may state that the solution of the inhomogeneous equation
(2.3.1) is made up of a particular solution of the inhomogeneous equation, in our
case (2.3.16), and a general solution of the homogeneous equation, in our case
p exp (At), where P is a constant which can be chosen arbitrarily. It can be fixed,
however, for instance if the value of q(t) at an initial time is given.
Let us now consider the general case in which a(t) and b(t) are constant,
periodic or quasiperiodic. In order to obtain results of practical use we assume
that in the quasiperiodic case a KAM condition is fulfilled and the Fourier series
of a(t) converges sufficiently rapidly, so that the solution of the homogeneous
equation (2.3.7) has the form
(2.3.17)
with u(t) constant, periodic, or quasiperiodic. From the explicit form of u(t) and
from the fact that the quasiperiodic function in the exponential function of
(2.3.17) is bounded, it follows that Iu lis bounded from below, so that its inverse
exists.
We first consider a particular solution of the inhomogeneous equation (2.3.1)
which by use of (2.3.11) and (2.3.17) can be written in the form
1
q(t) = eAtu(t) fe-hu-\r)b(r)dr. (2.3.18)
10
476
2.3 Driven Systems 75
(2.3.20)
r = r' + t (2.3.21)
o
q(t) = u(t) f e-h'u-I(r' + t)b(r' + t)dr' . (2.3.22)
We are now interested in what kind of temporal behavior the particular solution
of the inhomogeneous equation (2.3.22) represents. Let u(t) be quasiperiodic
with basic frequencies WI' ... , Wn • Since u - 1 exists (as stated above) and is quasi-
periodic, we may write it in the form
LJ (2.3.25)
Because evidently
477
76 2. Linear Ordinary Differential Equations
behavior of the solution q(t) of the inhomogeneous equation (2.3.1) for large tis
determined by the particular solution we have just discussed, i. e.,
The unknown variables are the q's. The coefficients aik are assumed to be given.
They may be either constants, or time-dependent functions of various types dis-
cussed below. The set of Eqs. (2.4.1) can most concisely be written in matrix
notation
(2.4.3)
(2.4.4)
478
2.4 General Theorems on Algebraic and Differential Equations 77
(2.4.5)
(2.4.6)
where Ap is an eigenvalue of L.
This theorem can be stated using the above definitions also as follows. For
each quadratic matrix L of complex numbers we may find a regular matrix S such
that
(2.4.7)
holds and i is of the form (2.4.5). Because S is regular, its determinant does not
vanish so that S -I exists.
479
78 2. Linear Ordinary Differential Equations
each tEl. The theory of linear differential equations assures us that for each toEI
and each column vector qoEEn a unique solution q(t) exists such that q(to) = qo'
The solutions of dqldt = L (t)q form an n-dimensional complex linear vector
space (compare Fig. 2.4.1). In other words, n linearly independent solutions of
(2.4.2) exist which we may label q(I)(t), ... , q(nl(t). Any solution q(t) of (2.4.2)
can be written as a linear combination
where the coefficients Cj are time independent. The qUl's are not determined
uniquely because by a transformation (2.4.10) we may go from one basis {qUl} to
another one {ij Ul}.
It is often convenient to lump the individual solution vectors q together to a
"solution matrix" Q
(2.4.11 )
(2.4.12)
we find by comparison
(2.4.13)
Our above statement on the transformations from one basis {qUl} to another can
now be given a more precise formulation.
Theorem 2.4.1. Let Q(t) be a solution matrix of dQldt = L(t)Q which is non-
singular. The set of all nonsingular matrix solutions is formed by precisely the
q,ltl
480
2.4 General Theorems on Algebraic and Differential Equations 79
matrices Q(t) C where C is any n X n constant, nonsingular matrix. For each toEI
and each complex constant matrix Qo a unique solution matrix Q(t) exists such
that Q(to) = Qo. A set of solution vectors q(I)(t), ... ,qn)(t) of dqldt=L(t)q
form a basis for the solution space if and only if they form the columns of a
solution matrix Q(t) of dQldt = L(t)Q which corresponds to a nonsingular
initial matrix Qo.
The differential system dqldt = L(t)q, where Ct < t < p = 00, is called stable
if every solution remains bounded as t -+ 00, L e.,
and some constant B. Then for every solution vector q(l)(t) of dqldt = L(t)q it
holds that
The real numbers Ai which arise in this way are called the generalized charac-
teristic exponents. There are at most n distinct generalized characteristic ex-
ponents. The differential system is stable whenever all the A'S are negative.
A special case of the "generalized characteristic exponents" is that of
Lyapunov exponents.
To this end let us consider nonlinear equations of the form
481
80 2. Linear Ordinary Differential Equations
it = L(q(t))u. (2.4.21)
(2.4.23)
we have
lul<D. (2.4.25)
As a result we obtain
(2.4.27)
482
2.5 Forward and Backward Equations: Dual Solution Spaces 81
,{ < O. (2.4.28)
According to the definition of lim sup, for any f. > 0 there exists a 10 so that for
any t > to the inequality
~ In I g I < ,{ + f. (2.4.29)
t
(2.4.30)
(2.4.31)
Therefore
N(q)-->O. (2.4.34)
q/ = const. , (2.4.35)
i. e., the trajectory terminates at a fixed point. Therefore, if the trajectory q(t)
does not terminate at a fixed point, the only remaining possibility for the
Lyapunov exponent under study is ,{ = 0, so that our theorem is proved.
For later purposes we slightly change the notation, writing w(k) instead of qU).
Equations (2.4.14, 11) read
where (2.5.1)
483
82 2. Linear Ordinary Differential Equations
(2.5.2)
We distinguish the different solution vectors by an upper index k and assume that
these solution vectors form a complete set, i. e., that they span the solution space.
The matrix L
(2.5.3)
may have any time dependence. Of course, we may represent (2.5.1) in the
somewhat more suggestive form
() ( )( ). (2.5.4)
W L w
As is known from algebra we may construct a dual space to w (kl (0) spanned by
vectors w(k'l (0) in such a way that
s _!1
Ukk' -
fork=k' (2.5.6)
ofor k =1= k' .
The scalar product ( ... ) between wand w is defined by
(W(k'lw(k l ) = L wjk'lwjkl . (2.5.7)
j
We now ask whether we may find an equation for the basis vectors, cf. (2.5.5), of
the dual space which guarantees that its solutions are for all times t ~ 0 ortho-
gonal to the basic solutions (2.5.2) of the original equations (2.5.1). To this end
we define w(k'l as the vector
- (k'l = (W(k'l
W 1 , ... , w(k'l)
n , (2.5.9)
484
2.5 Forward and Backward Equations: Dual Solution Spaces 83
(2.5.11)
L= -L. (2.5.12)
( ). (2.5.13)
L
We want to show that (2.5.5) is fulfilled for all times provided w obeys the equa-
tions (2.5.10). To this end we differentiate the scalar product (2.5.7), where we
use W(k') and W(k) at time t. We obtain
(2.5.14)
(2.5.15)
On account of (2.5.12), the rhs of (2.5.15) vanishes. This tells us that (2.5.5) is
fulfilled for all later times (or previous times when going in backward direction).
Thus we may formulate our final result by
(2.5.16)
In later sections we shall show that for certain classes of L(t) we may decompose
w(k) into
(2.5.17)
where v (k) has certain properties. In this section we use the decomposition
(2.5.17), but in an entirely arbitrary fashion so that V(k) need not be specified in
any way. All we want to show is that whenever we make a decomposition of the
form (2.5.17) and a corresponding one for W,
(2.5.18)
485
84 2. Linear Ordinary Differential Equations
the relations (2.5.16) are also fulfilled by the v's. The Ak'S are quite arbitrary.
Inserting (2.5.17 and 18) in (2.5.16) we immediately find
which due to the property of the Kronecker symbol can also be written in the
form
(2.5.20)
(2.6.1)
Since each power of a matrix is defined, (2.6.2) is defined too, and one can show
especially that the series (2.6.2) converges in the sense that each matrix element of
exp (Lt) is finite for any finite time t. By inserting (2.6.2) into (2.6.1) we may
readily convince ourselves that (2.6.1) fulfills (2.5.1) because
(2.6.3)
While this kind of solution may be useful for some purposes, we wish to derive a
more explicit form of q. Depending on different initial vectors qU)(O) different
solutions qU)(t) evolve. Provided the qU)(O),s are linearly independent, then so
are the qU) at all other times.
(2.6.4)
486
2.6 Linear Differential Equations with Constant Coefficients 85
The exponents Aj' often called "characteristic exponents", are the eigenvalues of
the matrix L. The vectors vU)(t) have the form
(2.6.5)
i. e., they are polynomials in t where the highest power mj is smaller or equal to
the degeneracy of Aj. If all A/s are different from each other, vU)(t) must then be
a constanL If several A/s coincide it might but need not happen that powers of t
occur in vJ. We shall see later how to decide by construction up to which power t
occurs in (2.6.5). Let us start to prove these statements. We start from the formal
solution of Q = L Q, namely
and leave the choice of Q(O) open. We now choose a regular matrix S so that Lis
brought to Jordan's normal form i,
(2.6.7)
Multiplying this equation by S from the left and S -I from the right we obtain
L =sis- I • (2.6.8)
(2.6.9)
Making use of the expansion of the exponential function (2.6.2), one may readily
convince oneself that (2.6.9) can be replaced by
(2.6.10)
Q(O) = S, (2.6.11)
--lITJ
L - [IJ [TI
0 J (2.6.13)
o
487
86 2. Linear Ordinary Differential Equations
In it, in the main diagonal all A/S are equal. Above this diagonal we have another
diagonal with 1'so All other matrix elements are equal to zero.
We first show that exp (it) has the same shape as (2.6.13). Using the rules of
matrix multiplication it is demonstrated that
(2.6.16)
[
[]m ~m
o
!It
0
.
l (2.6.17)
When we mUltiply both sides of (2.6.17) by tm/m! and sum up over m we obtain
which yields the same structure as (2.6.13), namely
eLI
Therefore, for our study of the form of the solution it will be sufficient to focus
our attention on each of the matrices H j which are of the form
(2.6.19)
where we have denoted the boxj by M j • For what follows we shall put
e Lt = Q(t), (2.6.20)
where we wish to use the normal form (2.6.13). Dropping the indexj, we write
the matrix (2.6.15) in the form
M=A·1+K. (2.6.21)
488
2.6 Linear Differential Equations with Constant Coefficients 87
(2.6.22)
whereas K is defined by
(2.6.23)
with mj rows and columns. Because the matrix 1 commutes with all other
matrices we may split the exponential function according to
(2.6.24)
(2.6.25)
K=O, (2.6.26)
(2.6.27)
(2.6.28)
(2.6.29)
489
88 2. Linear Ordinary Differential Equations
K = [010]
001
000
(2.6.30)
we readily obtain
K2 =
0 0 1
[ 0 0 0
J , (2.6.31)
000
K3 = (0) (2.6.32)
e K ! = 1 + tOO
[
0 1 01 J+- t 2 [00 00 10 ] = [1 t t 12-j
0 1 t
2
. (2.6.33)
000 2 000 o0 1
(2.6.34)
Ul ~ In' "'f
In order to obtain ijU) we decompose (2.6.34) into its column vectors ijU) which
yields
o 0
eA2! te A2 !
o e: 21
(2.6.36)
490
2.7 Linear Differential Equations with Periodic Coefficients 89
(2.6.37)
Evidently all these q's have the form (2.6.4) with (2.6.5), as the theorem stated at
the beginning of this section. The way in which this result was derived provides
an explicit construction method of these solutions. This procedure can be gener-
alized to the case of a time-periodic matrix L, to which case we now turn.
q= L(t)q, (2.7.1)
T: t->l+to, (2.7.2)
TL=LT. (2.7.3)
MUltiplying
Q=LQ (2.7.4)
491
90 2. Linear Ordinary Differential Equations
(2.7.8)
or due to (2.7.8)
A =SAS- 1 • (2.7.12)
This allows us to perform steps which are quite similar to those of the preceding
section. Due to (2.7.12) we have
(2.7.13)
1 Since in most cases (2.7.4) cannot be solved analytically one may resort to computer calculations.
One takes Q(O) as the unit matrix and lets the computer calculate Q(t) by standard iteration
methods until the time t = to is reached. Specializing (2.7.6) to this case we obtain
Q(to) = C, (2.7.7)
492
2.7 Linear Differential Equations with Periodic Coefficients 91
The form of this solution exhibits a strong resemblance to the form (2.6.12) of
the solution of a differential equation with constant coefficients. The only dif-
ference consists in the fact that the matrix S of (2.6.12) is now replaced by a
matrix fl, whose coefficients are periodic in time. This allows us to repeat all
former steps of Sect. 2.6 and to derive a standard form of the solution vectors.
They read
(2.7.18)
where
(2.7.19)
The characteristic exponents A.j in (2.7.18) are called Floquet exponents; they are
the eigenvalues of /1, cf. (2.7.12).
The coefficients vy)(t) are periodic functions of time with period to. For mj
we have the rule
This is the final result of this section. For readers who are interested in all details
we now turn to the question how to determine /1 in (2.7.8) if C is given. To this
end we introduce a matrix V which brings C into Jordan's normal form. From
(2.7.8) we find
(2.7.21)
(2.7.22)
(2.7.23)
493
92 2. Linear Ordinary Differential Equations
[ rsJ
o [Sjo.
lJ (2.7.24)
it is sufficient to assume that the still unknown .Ii (which stems from the still
unknown A) has the same decomposition corresponding to
(2.7.25)
[compare steps (2.6.16 to 18)]. Therefore our problem reduces to one which
refers to any of the submatrices of (2.7.24 or 25) or, in other words, we have to
solve an equation of the form
exp (Oto) =
Il
fl fl01 1 0
"~/J . (2.7.26)
where the box on the left-hand side represents a matrix still to be determined.
As C is a regular matrix,
(2.7.27)
o= A 1 + 0'1 to . (2.7.29)
l j
or
o1 0
exp(D') = 1 + fl- 1 0 ~. . (2.7.31)
o . 1
o
494
2.8 Group Theoretical Interpretation 93
Though we are dealing here with matrices, we have already seen that we may also
use functions of matrices in analogy to usual numbers by using power series
expansions. Because the logarithm can be defined by a power series expansion,
we take the logarithm of both sides of (2.7.31) which yields
(2.7.33)
Fortunately enough we need not worry about the convergence of the power series
on the rhs of (2.6.33) because it follows from formulas (2.6.28, 32) and their gen-
eralization that powers higher than a fixed number vanish. Equation (2.7.33)
gives us an explicit solution of our problem to determine the square in (2.7.29).
We have thus shown how we can explicitly calculate .It when C is given.
Results of Sects. 2.2, 7 may serve as an illustration of basic concepts of the theory
of group representations. In complete analogy to our results of Sect. 2.2, the
operator T generates an Abelian group with elements Tn, n ~ 0, n integer. But
what happens to the correspondence T -+ a we established in (2.2.16)? There our
starting point was the relation (2.2.10), i. e.,
i. e., (2.7.6) where C is a matrix. By applying T on both sides of (2.8.2) and using
(2.8.2) again we obtain
and similarly
(2.8.4)
Clearly
(2.8.5)
495
94 2. Linear Ordinary Differential Equations
Following from Theorem 2.4.1 (Sect. 2.4.3), C is a regular matrix. Thus we may
form the inverse C -1 and mUltiply both sides of (2.8.6) from the left with that
matrix. We then obtain
(2.8.7)
(2.8.8)
It is now quite obvious what the analog of relation (2.2.16) looks like:
T .... C
Tn .... C n
(2.8.9)
TO .... I
T- n -> C- n
(2.8.10)
T-nTn=E .... c-nc n =!
(TnTm) T' = yn(TmT') .... (cncm)c' = cn(cmc').
But the fundamental difference between (2.2.16, 17) on the one hand and
(2.8.9,10) on the other rests on the fact that a was a number, whereas C is a
matrix. Therefore, the abstract transformations Tn are now represented by the
matrices cn, and the multiplication of elements of the T-group is represented by
multiplication of the matrices cn. Since in mathematics one knows quite well
how to deal with matrices, one can use them to study the properties of abstract
groups (in our case the group generated by T). This is one of the basic ideas of
group representation theory.
In the foregoing section we have seen that by the transformation (2.7.14),
i. e.,
(2.8.12)
496
2.8 Group Theoretical Interpretation 95
i.e., i is decomposed into individual boxes along its diagonal [cf. (2.6.13)]. Now
we insert (2.8.11) into (2.8.2) and multiply both sides from the right by S,
(2.8.13)
This means that by a change of the basis of the solution matrix Q --+ Q, C in
(2.8.2) is transformed into
(2.8.14)
(2.8.15)
holds so that
(2.8.16)
(2.8.17)
Since the individual boxes (i. e., matrices) cannot be reduced further (due to
algebra), they are called an irreducible representation. Because the boxes are
multiplied individually if we form en, we obtain
(2.8.18)
(2.8.19)
and
(2.8.20)
497
96 2. Linear Ordinary Differential Equation<
Q=LQ (2.9.1)
into
(2.9.2)
Q=SQ, (2.9.3)
where S is a constant regular matrix, into (2.9.1), which yields after multiplying
both sides by S - j
(2.9.4)
(2.9.5)
(2.9.6)
so that
(2.9.7)
(2.9.8)
498
2.9 Perturbation Approach 97
into (2.9.7) we readily obtain after performing the differentiation with respect
to t
Q = e-tJMetJQ. (2.9.9)
'---v---'
M
(2.9.10)
where J, are the diagonal elements of the diagonal matrix J. So far all transfor-
mations were exact. Now to start with perturbation theory we assume that the
periodic part of L, i. e., L j or eventually Min (2.9.9), is a small quantity. In order
to exhibit this explicitly we introduce a small parameter e. Furthermore, we
decompose the differential equation (2.9.9) for the solution matrix into
equations for the individual solution vectors. Therefore we write
ij = eMij, (2.9.11)
where the matrix M is of the form
Mil M12
[ M21 M22 ... , (2.9.12)
Mnl ...
"'J
Mnn
and in particular
_ _ 1 2n
Mjj=O, M jj = 271: JMjjd qJ , qJ=wt. (2.9.13)
(2.9.16)
499
98 2. Linear Ordinary Differential Equations
with time-dependent vectors A (x). For e = 0 (2.9.16) reduces to the special vector
(2.9.17)
+ ...
= e [ Mll
~21 J + ... + ek + 1
Mnl
(2.9.18)
We now compare the coefficients of the same powers of e on both sides of
(2.9.18). This yields the set of differential equations discussed below.
In lowest order e, the first row of the resulting matrix equation reads
It is not necessary to choose 1) All) =1= 0 because this would alter only a normaliza-
tion. In the same order e but for the other rows with 1=1= 1 we obtain the relations
500
2.9 Perturbation Approach 99
We can solve for A)1) by integrating (2.9.21). Now it is quite important to choose
the integration constant properly, because we intend to develop a perturbation
theory which yields solutions with the form
(2.9.22)
v(t) periodic. Making a specific choice of the integration constant means that we
choose a specific initial condition. Because Mil has the form
(11)
AP)(t) = eLl/l! L cm. e imw !, (2.9.24)
m*O .,1/1 + Imw
a2 + A~2) = t MIlAP).
/~I
(2.9.25)
To study the structure of the rhs of (2.9.25) more closely, we make use of the
explicit form of Mij (2.9.14) and of A)1) (2.9.20,24).
If we decompose the periodic functions contained in M and A)I) into their
Fourier series, we encounter products of the form
For
m+m'=t:O (2.9.27)
the result is again a periodic function which truly depends on time t, whereas for
m+m' =0 (2.9.28)
a constant results. Consequently, the rhs of (2.9.25) can be written in the form
(2.9.29)
501
100 2. Linear Ordinary Differential Equations
(2.9.30)
A~2) = f
1=1
MkIAF) , (2.9.31)
(2.9.32)
or
(2.9.33)
(2.9.34)
where Pf) is periodic without a constant term and 2) is constant. Now the struc- ti
ture of the evolving equations is clear enough to treat the general case with
powers v ~ 3 of e. For the first row we obtain for v ~ 3 (A (x) == 0 for }{ ~ 0, ax = 0
for){~ 1)
Here the constant a v and the function A ~v) are still unknown. All other functions,
including A lx), 1 ;§i }{ ;§i v - 2, have been determined in the previous steps, assum-
ing that it was shown that these A's are periodic functions without constant
terms. Substituting these previously determined A's into the rhs of (2.9.35) we
readily find that the rhs has the form
(2.9.36)
502
2.9 Perturbation Approach 101
Let us now consider in the same order v the other rows of (2.9.18). Then we
obtain
In it the function A~v) is unknown. All others are determined from previous steps
and have the general form
(2.9.39)
or more concisely
(2.9.40)
(2.9.41)
j
of the solution q
PI + C 1
q= e yl
[ eLl~ll (P2 + C2 ) (2.9.42)
eLlnll(Pn + Cn)
where all terms can be constructed explicitly by an iteration procedure. From this
solution vector of (2.9.11) we can go back via (2.9.8) to the solution of (2.9.4).
We then find that the first column of the solution matrix has the form
~
iiO' ,"'," I~:~: l . (2.9.43)
lPn+CnJ
The C/s are constants, while the P/s are periodic functions (without additive
constants).
Similarly, we may determine the other column vectors by a mere interchange
of appropriate indices. In a last step we may return to Q by means of (2.9.3). As
503
102 2. Linear Ordinary Differential Equations
we have seen in Sects. 2.6 (cf. (2.6.36» and 2.7, this transformation leaves the
structure of the solution vector (2.9.22) unaltered. Our present procedure can be
generalized to the case of degeneracy of the A'S in which case we find terms in
(2.9.43) which contain not only periodic functions but also periodic functions
multiplied by finite powers of t. For practical purposes and for the lowest few
orders this procedure is quite useful. On the other hand, the convergence of this
procedure is difficult to judge because in each subsequent iteration step the
number of terms increases. In Sect. 3.9 another procedure will be introduced
which may be somewhat more involved but which converges rapidly (and even
works in the quasiperiodic case).
504
3. Linear Ordinary Differential Equations
with Quasiperiodic Coefficients *
In this section we wish to study the general form of the solution matrix Q(t) of
the differential equation
(3.1.3)
M(t, 'P) = L Mn'.n2 •.... nNexp [iw! n! (- 'Pt + t) + iW2n2( - 'P2 + t) + ... J
n,.n2····· nN (3.1.4)
104 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
which contains the phase angles <PI' ... ,<PN. In other words, we are embedding
the problem of solving (3.1.1) in the problem of solving
Tr'.[t~t+T (3.1.6)
<P ~ <P + T, T = Te ,
where T is an arbitrary shift and e is the vector (1, 1, ... , 1) in <P space. One sees
immediately that M is invariant against (3.1.6) or, in other words, that M
commutes with Tp
(3.1.7)
As a consequence of (3.1.7)
(3.1.8)
must hold (compare Theorem 2.4.1), where C(T, <p) is a matrix independent of t.
The difficulty rests on the fact that the matrix C still depends on T and <p. Using
the lhs of (3.1.9) in a more explicit form we find
After these preparatory steps we are able to formulate Theorem 3.1.1 which we
shall prove in the following.
Theorem 3.1.1. Let us make the following assumptions on (3.1.4 and 5):
1) The frequencies WI>"" wNare irrational with respect to each other (other-
wise we could choose a smaller basic set of w's).
2) M(t, <p) = M(O, <p-t) is 1j-periodic in <Pj' 1) = 2n/wj' and Ckwith respect
to <p (k ~ 0) ("C k " means as usual "k times differentiable with continuous deriva-
tives").
3) For some lfJ = lfJo the generalized characteristic exponents Aj of q(})(t, lfJo),
j = 1, ... , m, are different from each other. We choose At> A2 > ....
4) We use the decomposition
(3.1.11)
506
3.1 Formulation of the Problem and of Theorem 3.1.1 105
t~oo lI
lim sup \"..!..-ZP)] = Aj. (3.1.12)
We then require
This implies that the unit vectors u(t) keep minimum angles with each other, or,
in other words, that the u's never become collinear, i. e., they are linearly inde-
pendent.
5) Zj(t) possesses the following properties. A sequence In' In -+ 00 exists such
that
(3.1.14)
(3.1.15)
where
(In particular, the conditions (3.1.14) and (3.1.15) are fulfilled for any sequence
In if
(3.1.17)
(3.1.18)
(3.1.19)
507
106 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
q Ui = eX/vU)(t''Y
m) ,
Re{X-\
JJ
= AJ ' (3.1.20)
(3.1.21)
In Sect. 3.8 a generalization of this theorem will be presented for the case
where some of the A's coincide.
The proof of Theorem 3.1.1 will be given in Sects. 3.2 - 5,3.7 in several steps
(Sect. 3.6 is devoted to approximation methods - linked with the ideas of the
proof - for constructing the solutions q U). After stating some auxiliary
theorems ("lemmas") in Sect. 3.2, we first show how C can be brought to
triangular form (exemplified in Sect. 3.3 by the case of a 2 x 2 matrix, and in
Sect. 3.5 by the case of an m x m matrix). This implies that we can choose
qU)(t, ({J) such that Al > .1.2 > '" > Am for all rp. Then we show that the elements
of the triangular matrix can be chosen according to statements (a) and (b) of
Theorem 3.1.1 (in Sect. 3.4 by the case of a 2 x 2 matrix, and in Sect. 3.5 by the
case of a m x m matrix). Finally in Sect. 3.7 we prove assertion (c).
Proof: We start from (3.1.5) where the matrix Mis given by (3.1.4). The solu-
tion of (3.1. 5) can be expressed in the formal way
508
3.2 Auxiliary Theorems (Lemmas) 107
0 0
J
I(nl == _1_ t [fM(S, qJ)ds]n = IdSnM(Sn> qJ) ds n_ 1M(sn_l, qJ) ...
n! 0
52
Jds 1M(SI,qJ) . (3.2.3)
o
Because M is a matrix with matrix elements Mjk we can also write (3.2.3) in the
form
IJf:l==-+t[[fM(s,qJ)dSJi =
n. 0 )k
L JdSn~'I._l(Sn,qJ)
11..... 1._ 10
(3.2.5)
(3.2.6)
(3.2.7)
(3.2.8)
509
108 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
from which
(3.2.10)
Evidently the series converges for all times - 00 < t < + 00.
These considerations can be easily extended to prove the differentiability of
(3.2.1) with respect to rp. Denoting the derivative with respect to a specific rpj by a
prime
and dropping the index rpj' we obtain the following expressions for the derivative
of each matrix elementjk of a member of the series (3.2.2)
(3.2.14)
Let us consider (3.2.11) in more detail. Denoting derivatives with respect to ifJj
by a prime,
I ~ Q(t,
orpj
rp) I ::;; K exp (mMt)mq
ik
+ Iitexp f... l_o_Q(o, ifJ) I ' (3.2.16)
0 ) orpj ik
510
3.2 Auxiliary Theorems (Lemmas) 109
where the second term on the rhs can be estimated in a way analogous to (3.2.9).
Thus if M is C l with respect to ({J, cf. (3.2.13), then (3.2.11) exists for all times
- 00 < 1 < 00. Similarly, one may prove the convergence (and existence) of the
k'th derivative of Q(t, ({J) provided M is correspondingly often continuously dif-
ferentiable with respect to ({J.
Lemma 3.2.3. If the initial matrix Q(O, ({J) is periodic in ({J with periods
~ = 2nlwj, then Q(t, ({J) has the same property.
Proof' In (3.2.1, 2) each term has this periodicity property.
Lemma 3.2.4. This lemma deals with the choice of the initial matrix as unity
matrix. We denote the (nonsingular) initial matrix of the solution matrix with
({J = ({Jo by 12,
For all other ({J's we assume the same initial condition for the solution matrix for
1=0
For what follows it will be convenient to transform the solutions to such a form
that the initial solution matrix (3.2.17) is transformed into a unity matrix. To this
end we write the solution matrix in the form (omitting the argument ((J)
Inserting (3.2.19) into (3.2.20) and multiplying both sides from the left by Qo I
we obtain
(3.2.21)
Introducing
(3.2.22)
as a new solution matrix we may write (3.2.21) in the form [compare (3.1.1)]
Q(t)=MQ, (3.2.23)
511
110 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
Proof' Using the definition of Tr (3.1.6) we may write (3.1.9) in the form
When we choose t = 0 and apply (3.2.24), we obtain (3.2.25). Because Q(t, '1') is a
nonsingular matrix (cf. Theorem 2.4.1),
(3.2.27)
Lemma 3.2.6. Let us introduce the transposed matrix belonging to Q(r, '1')
(3.2.31)
We put
(3.2.33)
512
3.3 Proof of Assertion (a) of Theorem 3.1.1 111
Since all the basic ideas and essential steps can be seen by the simple case in which
M and thus Q are 2 x 2 matrices, we start with this example and put
(3.3.1)
q (I)(t, qJo - r) = /jl (r)q (I)(t + r, qJo) + /j2( r)q (2) (t+ r, qJo) (3.3.2)
q (2)(t, qJo - r) = 121 (r)q (I)(t + r, qJo) + 122 (r) q (2)(t + r, qJo) . (3.3.3)
Let us consider (3.3.2, 3) in more detail. On the lhs, q U)(t, qJo - r) is a subset of
the functions q U)(t, qJ) which are periodic in qJj and C k with respect to qJ. In par-
ticular, the functions on the lhs of (3.3.2, 3) are quasiperiodic in r. Because the
w's are assumed irrational with respect to each other, qJo- r lies dense in qJ or,
more precisely,
As a consequence of this and the C k property of q U)(t, qJ), q U)(t, qJo - r) lies dense
in the space q U)(t, qJ). On the rhs the asymptotic behavior of q U) for t --> 00 is
known. According to our assumption, q U), j = 1, 2 possess different generalized
characteristic exponents Aj. We wish to form new solutions of (3.1.5), q(1) and
q (2), which combine both these features, namely asymptotic behavior (q U) shall
possess the generalized characteristic exponent A) and quasi periodicity in the
argument qJo- r. Take AI > A2' In order to construct q(2) we multiply (3.3.2) by
a(r) and (3.3.3) by per), forming
(3.3.5)
In order that q (2) no longer contains the generalized characteristic exponent AI,
we require
513
112 3. Linear Ordinary Differential Equations with Quasiperiodic Cocfri,:icnts
(3.3.6)
which can obviously be fulfilled because the vector q (2) drops out of this equa-
tion.
By means of (3.2.33) (Lemma 3.2.6) we may transform (3.3.6) into
(3.3.10)
Namely. using (3.3.8,9) and (3.3.2,3) with (3.2.33), q(1) reads explicitly
(Here again DII (r) and D21 (r) cannot vanish simultaneously, bt::cause otherwise
the solution vectors qU)(t, 'Po- r) would become linearly dependent in contrast to
our assumption. Therefore the factor in front of q (1)(t + r, 'Po) does not vanish.)
Therefore the choice (3.3.10) secures that qI possesses the characteristic exponent
AI' In this way we have constructed two new solutions ql' q2 which are connected
with the generalized characteristic exponents Al and A2 , respectively.
When we use these new solutions vectors instead of the old ones, the matrix C
(3.1.18) appears in triangular form, as can be easily demonstrated and as will be
shown explicitly in Sect. 3.7.
514
3.4 Proof that the Elements of the Triangular Matrix C are Quasiperiodic in r 113
tions which are quasiperiodic in r. To this end we write the coefficients of a(r) in
(3.3.6) or (3.3.7) by use of (3.3.2) in the form
(3.4.1)
Similarly, we proceed with the coefficients of fJ(r) in (3.3.6). For the following it
will be sufficient to demonstrate how to deal with the coefficient of a(r) as an
example. In order to ensure that the coefficients of a(r) and fJ(r) remain
bounded for all positive times t and r, instead of (3.4.1) we form
where
(3.4.3)
q(j)(t, qJo- 7:) = exp [Zt (t+ r) - Zt (r)] u(1)(t+ r)Djt (r)
+ exp [Z2(t+ r) - z2(r)]u(2)(t+ r)Dj2 (r) , j = 1,2. (3.4.4)
(3.4.5)
(3.4.6)
515
114 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
lim SUp
t~co
[~Zj(t)]
t
= Aj. (3.4.8)
(3.4.9)
For tn sufficiently large, 0' can be chosen as small as we wish. On the other hand,
because of
(3.4.11)
we may find some 0" > 0 so that for the same sequence tn and 'I < T< T2
and 0" may be chosen arbitrarily small as tn is sufficiently large. Taking (3.4.10
and 12) together we see that (3.4.7) goes to zero for t n -> 00.
Consequently, we may replace everywhere in (3.3.7) for Tl < ,< T2
516
3.5 Construction of the Triangular Matrix C 115
JV(t, 'P)qU)(t, 'P). Because this expression is periodic and C k in 'P, so are
a(r) -+ a('P) and p(r) -+ P('P).
Let us summarize what we have achieved so far. Under assumptions (1 - 4) of
Theorem 3.1.1, we have constructed a complete set of solutions to (3.1.5),
(3.4.15)
with the following properties: ij I) is (at least) C l with respect t.o t and r, and it is
quasiperiodic in r.
Under assumptions (1 - 5) of Theorem 3.1.1 we have constructed a complete
set of solutions to (3.1.5),
(3.4.16)
with the following properties: ij I) is (at least) C l with respect to t, and 1)-
periodic and C k with respect to 'P. The set ijl)(t, 'P,) lies densely in the set
ijl)(t, 'P). In both (3.4.15, 16) the generalized characteristic exponents (for
t -+ + 00) of ij I) are given by Aj.
So far we have considered the special case in which M is a 2 x 2 matrix. Now let
us turn to the general case of an m X m matrix, again assuming that the charac-
teristic exponents are all different from each other
(3.5.1)
such that only one of these linear combinations has the generalized characteristic
exponent AI but that all other m - 1 linear combinations have at maximum a gen-
eralized characteristic exponent A2'
Furthermore, we want to show that ay) can be chosen as a quasiperiodic func-
tion in r or, more specifically, that we can embed aJI)(T) in functions aY)('P)
which are 1j-periodic and C k with respect to 'P. For simplicity, we shall drop the
index I in the following. We first introduce a normalization factor by
(3.5.3)
517
116 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
and form
(3.5.4)
(3.5.5)
. l
Let us use a sequence tn --> 00 defined by (3.4.10). We shall denote such a sequence
by Lim Sup. We readily obtain
(3.5.7)
(3.5.8)
where according to (3.5.7) all vectors C;U) are parallel, we can determine m - 1
linearly independent solution vectors
(3.5.9)
518
3.5 Construction of the Triangular Matrix C 117
Since a detailed presentation of this construction will diverge too far from our
main line, we merely indicate how this construction can be visualized. Let us con-
sider one vector component of the vector equation (3.5.8) [in which all vectors
~U) are parallel, cf. (3.5.7)]. Let us form a vector ~ = (c;£I), c;£2), ... , dm »), where k
is chosen such that ~:j: 0 which is always possible. Then the corresponding vector
component of (3.5.8) can be written as
(3.5.10)
i. e., as the scalar product between ~ and the vector a( = a(l») as introduced by
(3.5.9). In other words, we seek the m - 1 vectors a(l>, I = 2, ... , m, which span a
vector space orthogonal to the vector ~. Now, when we smoothly change the
t.
direction of we may smoothly change the vectors of the space orthogonal to ~.
(More generally, "smoothly" can be replaced by C k .) A simple example of an
algebraic construction will be given below.
For pedagogical reasons it must be added that it seems tempting to divide
(3.5.8) by the common "factor" Lim Sup {u(1)(tn+ r)}. Then the difficulty occurs
that it is not proven that Djl (r) and this factor have the required differentiability
properties with respect to r (or ffJ) individually. Therefore we must not divide
(3.5.8), which contains (3.5.7) via c;j, by Lim Sup {u(1)(tn+ r)}. If we want to get
rid of u(1), we may form the quantities In--+ oo
Lim Sup {zU)(t, ffJo- 7:) X(k)* (t, ffJo - 7:)} = 1'fjk( r) , (3.5.11)
1--+ 00
where '1jk( r) is C k with respect to r, provided M was C k with respect to ffJ. Since
ffJo-7: lies dense on the torus we can embed (3.5.11) in functions which are C k
with respect to ffJ and I;-periodic. Then (3.5.11) can be evaluated in analogy to
(3.5.6) which yields
(3.5.12)
Now U (I) drops out. In order to construct new linear combinations of (3.5.2)
which have a characteristic exponent A,2 or smaller we require
m
Eair)1'fjk(r)=O for k=l, ... ,m. (3.5.13)
j=1
519
118 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
and readily deduce from (3.5.11) that in the limit t---> 00 all these vectors are
parallel.
Furthermore we introduce (3.5.9) so that (3.5.13) can be written as
(3.5.15)
Since all 'Ik are parallel, the requirement (3.5.15) can be fulfilled by m - 1 dif-
ferent vectors a. In this way we can construct m - 1 new solutions (3.5.2) or,
more generally, we can embed
(3.5.16)
Though for some or all r some of the vectors 'Ik may vanish, on account of our
previous assumptions at least one 'Ik must be unequal to O.
As mentioned before, the a's of (3.5.15) can always be chosen in such a way
that they have the same differentiability properties with respect to parameters as
'Ik' This statement is illustrated by a simple example in the case of an even m:
If this is inserted into (3.5.15) we readily convince ourselves that (3.5.15) is ful-
filled and that a has the same differentiability properties as 'Ik' If one of the 'Ik
vanishes one can go over to another 'Ik by multiplying aj with a well-defined
factor.
Summary of Results. We have shown how to construct new solutions ij of the
original equation (3.1.5) which have the following properties. One of these
solutions has the generalized characteristic exponent Aj, all other m - 1 solutions
have the characteristic exponent ,A,2 or still smaller ones. We can now perform
precisely the same procedure with a group of the remaining m - 1 solutions and
reduce this group to one which contains only one solution with generalized
characteristic exponent ,A,2' The continuation of this procedure is then obvious.
Clearly, the resulting coefficient matrix C is triangular as is explicitly shown in
Sect. 3.7.
Since the original q U) are C k and 1j-periodic with respect to lfJ and the coef-
ficients a can be constructed in such a way that they retain this property also,
then the newly derived solutions are C k and 1j-periodic with respect to lfJ.
520
3.6 Approximation Methods. Smoothing 119
m
L la1hl 2 = O. (3.6. 1)
k=l
(3.6.2)
and we cannot fulfill (3.6.1) exactly because the vectors defined by (3.6.2) are not
completely parallel. But now we can require
Ela17kl
k=l
2 = minimum! , (3.6.3)
m
L I1J(a'1k)=Aaj' j= 1, ... ,m (3.6.4)
k=l
m
L (a* 17t)(a17k) = A lal 2 • (3.6.5)
k=l
We thus see that the eigenvalues A will be a measure as to how far we are still
away from the exact equation (3.6.1).
3.6.2 Smoothing
Since some ~k or, equivalently, some 17k> k fixed, in (3.5.15) can vanish for some
r, we have to switch to another equation stemming from another 17j (or ~). If we
have not performed the limit tn ---> 00 in Lim Sup, these two 17's may not entirely
coincide. This in turn would mean that at those switching points a becomes dis-
continuous and nondifferentiable. We want to show that we can easily smooth
the switching. To this end we introduce the new vector
(3.6.6)
afh=O. (3.6.7)
Since due to the linear independence of the solutions not all17j can vanish for the
same r, we may choose such an index j that (3.6.6) does not vanish where 17k
vanishes. We define a variable x by
521
120 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
, - '0
x= (3.6.8)
'1 - '0
which secures
and
x= 1, ,= '1' (3.6.10)
We choose '0 and '1 such that '1k is replaced by llj before 11k vanishes. On the
other hand, we choose '0 and '1 such that 111 k Iis smaller than 111 j I so that (3.6.6)
cannot vanish due to compensation of the l1's. We define {J by
{J = 0, -00 <x:::;; 0
(J=h(x), O:::;;x:::;;l (3.6.11)
(J=1, 1:::;;x<00.
In order that the "spline function" (3.6.6) is sufficiently often continuously dif-
ferentiable with respect to " we require in addition for a given n
Here and in the following the superscript (m) denotes the m'th derivative.
Since the points, at which we have to apply this smoothing procedure have
the same properties of quasi periodicity as 11k> llj' the newly constructed {J's will
also have this property. When we let t -> 00, our above procedure guarantees that
during the whole approach {J remains en and quasiperiodic.
The following short discussion shows how to construct {J in (3.6.11, 12)
explicitly. This outline is not, however, important for what follows. In order to
fulfill (3.6.12) we put
(3.6.13)
(3.6.14)
where we require
522
3.6 Approximation Methods. Smoothing 121
ao= 1. (3.6.15)
x=1+~ (3.6.16)
and put
n 1
g(~) = L bl~ . (3.6.18)
1=0
are fulfilled, where the superscript m in (3.6.20) means derivative with respect
to x.
From (3.6.19) we immediately find
g(O) = 1 (3.6.21)
and thus
bo = 1 . (3.6.22)
(3.6.24)
One may determine all other coefficients b l by use of (3.6.20) consecutively. Thus
one may indeed find a polynomial with the required properties. We still have to
convince ourselves that the thus constructed solution (3.6.13) has everywhere a
positive slope in the interval [0,1], i.e., that h(x) has no wiggles. Since h is a
polynomial of order 2n, it can have 2n -1 different extrema at most, which are
just the zeros of h (1). According to our construction n - 1 extrema coincide at
x = 0 and another n - 1 extrema coincide at x = 1. Therefore only one additional
extremum remains at most. One readily shows that
523
122 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
h'(l-c:»O (3.6.26)
hold for small positive c:. This implies that a wiggled curve must have at least two
more extrema as shown in Fig. 3.6.2, in contrast to our results predicting that
there could be at most only one additional extremum. Therefore, the case of Fig.
3.6.2 cannot apply and (3.6.13) can represent only the curve like Fig. 3.6.1. This
completes the excursion.
Fig. 3.6.1. The smoothing function P(x) Fig. 3.6.2. This figure shows that p has two extrema
in contrast to our assumption made in the text
So far we have discussed how we can bridge the jumps when '1k vanishes as a
function of T. Now we want to show that a similar procedure holds when we treat
'1k as a function of rp via embedding. Consider a rp <-+ rpo - T for which D"tl (T) 0, *
k fixed. Because qU)(t, rp) is C k , k ~ 0, a whole surrounding S(rp) of rp exists
where each point in S(rp) is approximated by sequences rp- Tn (mod T),
*
T= (11, ... , Tn), 1) = 2nlwj' and for which D"tt (Tn) 0, k fixed. In this way, the
whole torus can be covered by overlapping patches of nonvanishing Dj~ (T). On
each of these patches at least one nonvanishing vector '1k exists. Now the only
thing to do is to smooth the transition when going from one patch to a neighbor-
ing one. But the size of these patches, being defined by Dj~ (T) 0, is independent *
of t and thus fixed, whereas the misalignment of the vectors 11k tends to zero for
t --+ 00. Consequently the smoothing functions remain sufficiently often differen-
tiable for t --+ 00.
Since Q(t, rp) or Q(t, rp) (3.5.16) are solutions to (3.1.5) we may again invoke the
general relation (3.1.10). Writing down this relation for the solution vectors
explicitly we obtain (G = C T)
q- U)( t + T, rp + T ) -- ~
1... Gjk ( T, rp ) q- (k)( t, rp,
) J. -- 1, ... , m , (3.7.1)
k
524
3.7 The Triangular Matrix C and Its Reduction 123
Because the solution matrix is nonsingular we may derive from the properties of
Q(t, qJ) in Sect. 3.5 the following properties for the coefficients Gjk .
If assumptions (1- 5) of Theorem 3.1.1 hold, the coefficients Gjk(r, qJ) are
differentiable with respect to r and are 1j-periodic and C k with respect to qJ. If
the weaker assumptions (1 - 4) of Theorem 3.1.1 hold, the coefficients are of the
form G jk ( r, qJo - r), they are (at least) C l with respect to r and CO with respect to
the argument qJr( = qJo- r) and quasiperiodic in r. Now the asymptotic behavior
of the lhs of (3.7.1) must be the same as that of the rhs of (3.7.1) for t -> 00. This
implies that
(3.7.3)
The reasoning is similar to that performed above and is based on the study of the
asymptotic behavior of q(k) for t -> 00.
In the next step we want to show that it is possible to transform the q's in such
a way that eventually the matrix Gjk can be brought into diagonal form. To this
end we need to know the asymptotic behavior of q for t -> - 00. In order to
obtain such information we solve the equations (3.7.1) step by step starting with
the equation for j = m,
(3.7.4)
(3.7.5)
(3.7.6)
instead of (3.7.4). On the rhs of (3.7.6) we may replace q by the rhs of (3.7.4)
which yields
(3.7.7)
(3.7.8)
to to
Using (3.7.5) and the fact that r is infinitesimal we may replace the product over I
by the exponential function which yields
(3.7.9)
525
124 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
Eventually taking the limit r --> 0 we may replace the sum in (3.7.9) by an integral
(3.7.10)
(3.7.12)
This equation can be considered as a functional equation for ij(m). To solve this
equation we make the hypothesis
(3.7.13)
(3.7.14)
Since (3.7.13) was C k with respect to ({J and 1j-periodic in ({J, we obtain the fol-
lowing result:
(3.7.16)
ij(I)(t + r, ({J+ -r) = G l1 (r, ({J)ij (1)(t, ({J) + Gtz{ r, ({J)ij (2) (t, ({J) (3.7.17)
We want to show that we can construct a new solution to our original equation
(3.1.5) with the solution vector ij(l) so that the Eqs. (3.7.17 and 18) acquire a
526
3.7 The Triangular Matrix C and Its Reduction 125
diagonal form. We note that any linear combination of ij(l) and ij(2) is again a
solution of (3.1.5) provided the coefficients are time independent. We therefore
make the hypothesis
(3.7.19)
Since the underlined expressions are those which we want to keep finally, we
require that the rest vanishes. Making use of (3.7.18) we are then led to
(3.7.21)
We want to show that this equation can indeed be fulfilled, which is by no means
obvious because G jk are functions of both rand rp whereas h is assumed to be a
function of rp alone. In the following we shall show that the G's have such a
structure that h can indeed be chosen as a function of rp alone.
With this task in mind, we turn to a corresponding transformation of the Eqs.
(3.7.1), starting with the second last
(3.7.23)
In precisely the same way as in the case of the last equation (j = m) of (3.7.1), we
readily obtain for finite r
(3.7.24)
(3.7.25)
527
126 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
After dividing (3. 7 .26) by the exponential function on the lhs of (3.7.26) we
obtain
In order to determine the form of G m - l ,m for finite r, we first start for infini-
tesimal r's because G is differentiable with respect to r and because Gm-I,m must
vanish for r = O. According to (3.7.22) we may write
(3.7.30)
and consider (3.7.27) for the sequence r, 2 r, .... We write this sequence as
follows
N-I
y(t+Nr, rp+Nr) =y(t, rp) + 'i/(t+lr, r, rp+lr), (3.7.34)
~ '-v---' • 1=0
to to
528
3.7 The Triangular Matrix C and Its Reduction 127
(3.7.36)
~ [ t+s
K(t,to, qJ) = Jdsb(qJ+s)exp - J daam_l(qJ+s-a)
o 0
Inserting (3.7.38) into (3.7.36) and making some simple shifts of integration
variables s and a on the resulting lhs of (3.7.36), one can readily convince oneself
that (3.7.38) indeed fulfills (3.7.36). Then (3.7.38) can be written in a more
practical form as
l
with
(3.7.41)
It follows from (3.7.23) and our results on G(r, qJ) that 0m-l(qJ) is 1j-periodic
and C k with respect to qJ. If in addition we invoke assumption (6) of Theorem
3.1.1, the differentiability C k with respect to qJ and the KAM condition secure
that the integrals over OJ in (3.7.38) remain finite for a-+ ± 00 and, in particular,
529
128 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
(3.7.42)
(3.7.43)
which gives us
(3.7.44)
We immediately see that (3.7.44) has precisely the form required in (3.7.19)
where we may identify the second term on the rhs with the rhs of (3.7.44) and J
with h. Thus we can indeed choose help) or J(lp) as a function of lp alone, which
makes it possible to cast (3.7.22) into diagonal form.
We now want to study the explicit form of
Gm-l,mQ-(m) . (3.7.45)
A comparison between (3.7.22) and (3.7.36) with help of (3.7.26) indicates that
therefore we have to multiply (3.7.37) by
(3.7.46)
G m - 1,m(r, lp)
= exp [rda am-l (lp+ a)] IdS b(lp+s) exp [ldalam(lp+ a) - am-I (lp+ a)] ]-
(3.7.47)
530
3.8 The General Case: Some of the Generalized Characteristic Exponents Coincide 129
we may determine the coefficients hk(rp) in such a way that we obtain a diagonal
matrix G for ij, i. e., the solution vectors of (3.7.1) can be chosen so that
(3.7.49)
(3.7.50)
(3.7.51)
In this section we shall present two theorems. The first one deals with a reduction
of the matrix C (3.1.18) if some of the generalized characteristic exponents
coincide. The second theorem shows that C can even be diagonalized if all gen-
eralized characteristic exponents coincide, and an additional assumption on the
growth rate of Iq U) lis made. The first of these theorems is formulated as follows.
Theorem 3.8.1. Under assumptions (1, 2, 4) of Theorem 3.1.1 and by a suitable
choice of the solution vectors qU)(t, rp), the matrix C (3.1.18) can be brought to
the triangular form [cf. (3.7.1)]
(3.8.1)
531
130 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
(3.8.2)
If
(3.8.3)
we may apply our previous reduction scheme (Sects. 3.1 -7) at least up to Ak-I'
In order to treat the case in which several of the A'S coincide, we therefore
assume that AI is I times degenerate, i. e.,
(3.8.4)
We assume again that all q's are linearly independent with finite angles even if
t-+ + 00. We define a normalization factor by [cf. (3.5.3, 4)]
(3.8.5)
and form
(3.8.6)
We define
by the following description. Take a sequence t = tn' t n -+ 00, such that at least
for one of the qU)'s and some positive e
(3.8.8)
./11
k~I+1
r Djkq(k)(t+r,'Po)exp[-zk(r)] (3.8.9)
against
I
v,jl L
Djkq (k)(t + r, 'Po) exp [ - Zk( r)] , (3.tUO)
k=1
532
3.8 The General Case: Some of the Generalized Characteristic Exponents Coincide 131
I
= E Djk(r) Lim Sup {JV(t+ r) exp [Zk(t+ r) - Zk(r)] U(k)(t+ r}}. (3.8.11)
k= I I .... 00
In Lim Sup we take all such sequences tn for which at least one qU) fulfills (3.8.8).
Note that such sequences may depend on r. We now form
m .
E alr) Lim Sup {x U) (t, qJo+ r)} (3.8.12)
j= I 1.... 00
and require that the a/s are dete.rmined in such a way that (3.8.12) vanishes.
Using the explicit form of XU) and (3.8.12) we thus obtain
I m
E E alr)Djk(r) Lim Sup {,AI(t+ r) exp [Zk(t+ r) - zk(r)]u(k)(t+ r)} = 0
k=lj=1 1 .... 00 (3.8.13)
and because the u(k),S are linearly independent of each other we obtain
m
E aj(r)Djk(r) = 0 for k = 1, ... , I (3.8.15)
j=1
(3.8.16)
533
132 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
such that (3.8.12) does not vanish. Let us treat the case in which the least
reduction is possible, l' = /. If we label the vectors (3.8.16) for which (3.8.12)
remains finite by
(3.8.19)
On account of our construction, the new solutions ij (k), k = / + 1, ... , m, possess
generalized characteristic exponents A2 or smaller. When we use some arguments
of Sects. 3.1-7, we recognize that the matrix C T = (Gjk ), [cf. (3.7.1)], is reduced
to the form
m-/
(3.8.20)
m-/ 0
Again one may show that the aj(r) have the same differentiability properties as
their coefficients in (3.8.12). Ifthe coefficients in (3.8.12) are not considered as a
function of r but rather of rp (compare the previous sections), the a's acquire the
same differentiability properties. Therefore ij(k) (3.8.19) has again the same dif-
ferentiability properties as the original solution vectors qU)(t, rp). After having
reached (3.8.20), we may continue our procedure so that we can reduce the
scheme of coefficients in (3.8.20) to one of the triangular form
o
l
The question arises whether we may reduce the scheme (3.8.21) further to one in
which we have nonvanishing matrices along the main diagonal only
(3.8.21)
(3.8.22)
534
3.8 The General Case: Some of the Generalized Characteristic Exponents Coincide 133
There are various circumstances under which such reduction can be reached. This
can be achieved, for instance, if for t .... - 00 the generalized characteristic
exponents A} just obey the inverse of the corresponding relations (3.8.2). It is,
however, also possible to reduce scheme (3.8.21) to one of the form (3.8.22) if it
can be shown that the squares in (3.8.21) can be further reduced individually to
diagonal form. Even if that is not possible, further general statements can be
made on the form of the solution matrix Q(t, rp). Namely, if the solution matrices
belonging to the Qrsquares in (3.8.21) are known, the total matrix Q(t, lfJ) can be
found consecutively by the method of variation of the constant. The procedure is
analogous to that in Sect. 3.7, but ijU) must be replaced by submatrices QU), etc.
In the final part of this section we want to present a theorem in which all Ak'S
are equal.
Theorem 3.8.2. Let us make the following assumptions:
1) M(3.1.4) is at least C t with respect to rp;
2) Ak = A, k = 1, ... , m;
3) e-Anrll T;q I and eAnrll Tr-nq II are bounde~ for n .... 00 and rarbitrary, real, and
for all vectors of the solution space of Q = M(t, rp) Q, q, which are C t with
respect to rp. ( II ... II denotes the Hilbert space norm.)
provided the spectrum [of Tr acting in the space q(t, rp)] is a point spectrum.
In the case of continuous spectrum, Tr is equivalent to an operator Tr of
"scalar type", i. e., Tr can be decomposed according to
(3.8.24)
535
134 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
A= - JXE(dX) (3.8.25)
and
(3.8.26)
respectively, where E is the resolution of the identity for A (and Tr). If the
spectrum Xis a point spectrum, it follows from (3.8.26) that
536
3.9 Explicit Solution of (3.1.1) by an Iteration Procedure 135
(3.9.2)
(3.9.3)
(3.9.5)
(3.9.6)
A is a constant matrix.
To show how V and A can be constructed explicitly, we start from the
equation
(3.9.8)
Due to our assumption that the real parts of the eigenvalues of A are distinct, all
eigenvalues of A are distinct and therefore we may assume that J is of purely
diagonal form. We now put
and insert it into (3.9.7). After multiplication of (3.9.7) by C -I from the left we
obtain
Q= [J + Mo(t)] Q, (3.9.10)
537
136 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
(3.9.11)
where 1 is the unity matrix. Since we want to construct a solution of the form
(3.9.5), it would be desirable to choose U, as a quasiperiodic matrix but Q, as an
exponential function of time.
For a practical calculation it turns out, however, that we cannot construct U,
exactly. Therefore we resort to an iteration scheme in which we aim at calculating
U, approximately but retaining the requirement that U, is quasiperiodic. We then
derive a new equation for Q,. It will turn out that by an adequate choice of the
equation for U" the equation for Q, takes the same shape as equation (3.9.10)
with one major difference. In this new equation Mo is replaced by a new quasi-
periodic matrix M, whose matrix elements are smaller than those of Mo. Thus the
basic idea of the present approach is to introduce repeatedly substitutions of the
form (3.9.12).
In this way the original equation (3.9.10) is reduced to a form in which the
significance of the quasiperiodic matrix M(t) becomes smaller and smaller.
These ideas will become clearer when we perform the individual steps explicitly.
We insert hypothesis (3.9.12) into (3.9.10) which immediately yields
. .
U, Q, + [1 + U, (t)] Q, = lQ, + lU, Q, + MoQ, + MoU, Q1 . (3.9.13)
For reasons which will transpire below we add on both sides of (3.9.13)
(3.9.14)
(3.9.15)
In general, we shall define D, as the constant part of the main diagonal of the
matrix Mo. We do this here only for formal reasons because Mo was anyway con-
structed in such a way that its constant part vanishes. However, in our sub-
sequent iteration procedure such terms can arise. Next, D, can be found explicitly
via the formula
1 2n 2n Mo,11 0
D1 = ~-N- J... J M O,22 d((J1" .d((J]V. (3.9.16)
(2n) 0 0 0 Mo,mm
538
3.9 Explicit Solution of (3.1.1) by an Iteration Procedure 137
(3.9.17)
To simplify this equation we assume that U j can be chosen in such a way that the
underlined expressions in (3.9.17) cancel for an arbitrary matrix Q. Before
writing down the resulting equation for U1 , we make explicit use of the assump-
tion that U j is a quasiperiodic function, i. e., we write
(3.9.19)
(3.9.20)
(3.9.22)
(3.9.24)
(3.9.25)
539
138 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
(3.9.26)
(3.9.27)
(3.9.28)
Moex£ . (3.9.29)
(3.9.30)
J ex 1 . (3.9.31)
From (3.9.30, 31) it follows that the rhs of (3.9.20) is proportional to £. This
leads us to the conclusion
(3.9.32)
(3.9.33)
J 1 ex 1 + £. (3.9.34)
(3.9.35)
That means that we have indeed achieved an appreciable reduction of the size of
the quasiperiodic part on the rhs. It should be noted that our estimate is a super-
ficial one and that the convergence of the procedure has been proven mathema-
540
3.9 Explicit Solution of (3.1.1) by an Iteration Procedure 139
tically rigorously (see below). Our further procedure is now obvious. We shall
continue the iteration by making in a second step the hypothesis
(3.9.36)
(3.9.37)
(3.9.40)
(3.9.41)
(3.9.42)
(3.9.43)
This equation can now be solved explicitly. Its general solution reads
(3.9.44)
(3.9.45)
541
140 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
We are now left with the explicit construction of U v + j , v = 0,1, ... , which must
incorporate the proof that U can be chosen as a quasiperiodic function. Accord-
ing to (3.9.20) the equation for U reads
(3.9.46)
M-D=M' , (3.9.47)
Inserting (3.9.48, 49) into (3.9.46), we find for the Fourier coefficients the fol-
lowing relations
We now introduce the matrix elements of the matrices Un' J andM~ by means of
(3.9.51)
J. (3.9.52)
(3.9.53)
(3.9.54)
(3.9.55)
provided the denominator does not vanish. We can easily convince ourselves that
the denominator does not vanish. For k :j= j it was assumed that the real parts of
the eigenvalues of the original matrix A (or J) are distinct. Provided M is small
542
3.9 Explicit Solution of (3.1.1) by an Iteration Procedure 141
enough, Dv+l are so small that subsequent shifts of J according to (3.9.38) are so
small that (3.9.56) is fulfilled for all iterative steps. Therefore we obtain for k *j
For k =j we find
M~.(O)
JJ
= 0 • (3.9.59)
t
U v + 1 = -aJexp[(Jk-J)T1M~«(fJ+'t')dr. (3.9.61)
o
The sign a J~ .. means that t = - 00 if Jk-Jj > 0 and t = 00 if h-Jj < 0 are taken
as lower limit of the integral.
We may summarize our results as follows. We have devised an iteration
procedure by which we can construct the quasiperiodic solution matrix
V= n (1 + U
I
v=l
v ), /-+ 00 (3.9.62)
(3.9.63)
543
142 3. Linear Ordinary Differential Equations with Quasiperiodic Coefficients
544
4. Stochastic Nonlinear Differential Equations
In this chapter we shall present some of the most essential features of stochastic
differential equations. Readers interested in learning more about this subject are
referred to the book by Gardiner (cf. references).
In many problems of the natural sciences, but also in other fields, we deal
with macroscopic features of systems, e. g., with fluids, with electrical networks,
with macroscopic brain activities measured by EEG, etc. It is quite natural to
describe these macroscopic features by macroscopic quantities, for instance, in
an electrical network such variables are macroscopic electric charges and
currents. However, one must not forget that all these macroscopic processes are
the result of many, more or less coherent, microscopic processes. For instance, in
an electrical network the current is ultimately carried by the individual electrons,
or electrical brain waves are ultimately generated by individual neurons. These
microscopic degrees of freedom manifest themselves in the form of fluctuations
which can be described by adding terms to otherwise deterministic equations for
the macroscopic quantities. Because in general microscopic processes occur on a
much shorter time scale than macroscopic processes, the fluctuations represent-
ing the underworld of the individual parts of the system take place on time scales
much shorter than the macroscopic process. The theory of stochastic differential
equations treats these fluctuations in a certain mathematical idealization which
we shall discuss below.
In practical applications one must not forget to check whether such an ideal-
ization remains meaningful, and if results occur which contradict those results
one would obtain by physical reasoning, one should carefully check whether the
idealized assumptions are valid.
4.1 An Example
Let us first consider an example to illustrate the main ideas of this approach. Let
us treat a single variable q, which changes during time I. Out of the continuous
time sequence we choose a discrete set of times Ii' where we assume for simplicity
that the time intervals between Ii and Ii-I are equal to each other. The change of q
when the system goes from one state li-l to another one at time Ii will be denoted
by
Llq(t;) =q(t;) - q(ti-I). (4.1.1)
144 4. Stochastic Nonlinear Differential Equations
(4.1.2)
This impact of coherent driving forces and fluctuating forces can be described by
(4.1.3)
where Ll w(tJ = w(tJ - W(t i _ 1). Because Ll w represents the impact of micro-
scopic processes on the macroscopic process described by q, we can expect that
Ll w depends on very many degrees of freedom of the underworld. As usual in
statistical mechanics we treat these many variables by means of statistics. There-
fore we introduce a statistical average which we characterize by means of the fol-
lowing two properties.
1) We assume that the average of Ll w vanishes
Otherwise Ll w would contain a part which acts in a coherent fashion and could
be taken care of by means of K.
2) We assume that the fluctuations occur on a very short time scale. Therefore
when we consider the correlation function between .,1 w at different times ti and Ij
we shall assume that the corresponding fluctuations are uncorrelated. Therefore
we may postulate
(4.1.5)
The quantity Q is a measure for the size of the fluctuations; ()tJj expresses the sta-
tistical independence of .,1 w at different times t;, tj . The time interval Ll I enters
into (4.1.5) because we wish to define the fluctuating forces in such a way that
Brownian motion is covered as a special case. To substantiate this we treat a very
simple example, namely
K=O. (4.1.6)
In this case we may solve (4.1.3) immediately by summing both sides over Ii'
which yields
N
q(t) - q(to) = L Ll w(t), t - 10 = N Ll t . (4.1.7)
j=!
We have denoted the final time by I. We choose q(to) = O. It then follows from
(4.1.4) that
(q(t) = O. (4.1.8)
546
4.1 An Example 145
In order to get an insight as to how big the deviations of q(t) may be on the
average due to the fluctuations, we form the mean square
N N
(q2(t» = L L (,1 w(tJLI w(tj» , (4.1.9)
i~ 1 j~1
N
Q LLlt=Qt. (4.1.10)
j~1
This result is well known from Brownian motion. The mean square of the coordi-
nate of a particle undergoing Brownian motion increases linearly with time t.
This result is a consequence of the postulate (4.1.5).
We now return to (4.1.3). As it will transpire from later applications, very
often the fluctuations LI w occur jointly with the variable q, i. e., one is led to
consider equations of the form
First the system has reached q(ti- J, then fluctuations take place and carry the
system to a new state q(t;). Because the fluctuations LI w occur only after the state
q at time (i-l has been reached, q(ti-l) and ,1 w(t;) are un correlated
=0
As we shall see below, this feature is very useful when making mathematical
transformations. On the other hand, it has turned out that in many applications
this choice (4.1.12) does not represent the actual process well. Rather, the fluc-
tuations go on all the time, especially when the system moves on from one time to
the next time. Therefore the second choice, introduced by Stratonovich, requires
that in (4.1.12) q is taken at the midpoint between the times t i - I and ti. Therefore
the Stratonovich rule reads
(4.1.14)
547
146 4. Stochastic Nonlinear Differential Equations
In this section we consider the general case in which the state of the system is
described by a vector q having the components q,. We wish to study a stochastic
process in which, in contrast to (4.1.3 or 11), the time interval tends to zero. In
the generalization of (4.1.11) we consider a stochastic equation in the form
We postulate
(4.2.3)
'*
In contrast to (4.1.5) we make Q = 1 because any Q 1 could be taken care of by
an appropriate choice of g'm' For a discussion of orders of magnitude the
following observation is useful. From (4.2.3) one may guess that
(4.2.4)
(4.2.5)
m m
Working from the Ito assumption that q occurring in g, and dWm in (4.2.1) are
statistically uncorrelated, we have therefore split the average of the last term into
a product of two averages. From (4.2.2) we find
548
4.2 The Ito Differential Equation and the Ito-Fokker-Planck Equation 147
or, after a formal division of (4.2.6) by dt, we obtain the mean value equation of
Ito in the form
d
- (q/(t) > = (KM (t» > . (4.2.7)
dt
Instead of following the individual paths of q/(t) of each member of the statis-
tical ensemble, we may also ask for the probability to find the variable q / within a
given interval q/ ... q/ + dq/ at a given time t. We wish to derive an equation for
the corresponding (probability) distribution function. To this end we introduce
an arbitrary function u(q)
U = u(q), (4.2.8)
which does not explicitly depend on time. Furthermore, we form the differential
of Uj' i. e., dUj' up to terms linear in dt. To this end we have to remember that dq,
contains two parts
dq/ = dq/,l + dq/,2 with dq/,l = O(dt) and dq/,2 = O(Vdt) , (4.2.9)
where the first part stems from the coherent force K and the second part from the
fluctuating force. Therefore we must calculate dUj up to second order
8u· 1 8 2 u_
dUj = L __
1 dqk + - L 1 dqkdq/. (4.2.10)
k 8qk 2 k/ 8qk 8q,
(4.2.11)
(4.2.12)
where q is taken such that it is not correlated with dw. We now turn to the deter-
mination of the distribution functiop.. To elucidate our procedure we assume that
time is taken at"discrete values t i . We consider an individual path described by the
sequence of variables qj' Wj at time tj,j = 0,1, ... , i. The probability (density) of
finding this path is quite generally given by the joint probability
(4.2.13)
549
148 4. Stochastic Nonlinear Differential Equations
(4.2.15)
and
(4.2.16)
(4.2.17)
where we used the definition off introduced above. Interchanging the averaging
procedure and the time evolution of the system described by the differential
operator d and using (4.2.11) we readily find
2 3
(4.2.18)
550
4.2 The Ito Differential Equation and the Ito-Fokker-Planck Equation 149
d<Uj)
- -= q )-o f (q, t Iqo, to)·
Sdn qUj { (4.2.19)
dt at
(4.2.20)
(4.2.21)
where we assumed thatf (and its derivatives) vanishes at the boundaries. Term 3
vanishes because the average factorizes since q and dw are uncorrelated and
because of (4.2.2). The last term 4 of (4.2.18) can be transformed in analogy to
our procedure applied to the term (4.2.20), but using two consecutive partial in-
tegrations. This yields for a single term under the sums
When we consider the resulting terms 1, 2 and 4, which stem from (4.2.18), we
note that all of them are of the same type, namely of the form
(4.2.23)
where the bracket may contain differential operators. Because these expressions
occur on both sides of (4.2.18) and because uj was an arbitrary function, we
conclude that (4.2.18), if expressed by the corresponding terms 1, 2 and 4, must
be fulfilled, even if the integration over dnq and the factor Uj(q) are dropped.
This leads us to the equation
(4.2.24)
551
150 4. Stochastic Nonlinear Differential Equations
In this calculus q, which occurs in the functions g, is taken at the midpoint of the
time interval Ii-I . .. Ii' i. e., we have to consider expressions of the form
g'm ( q ( t+ (.
/ 2/- 1)) dwm(t;) . (4.3.1)
(Here and in the following summation will be taken over dummy indices.)
This rule can be easily extended to the case in which 9 depends explicitly on
time t, where t must be replaced according to (4.3.1) by the midpoint rule. Two
things must be observed.
1) If we introduce the definition (4.3.1) into the original Ito equation (4.2.1),
a new stochastic process is defined.
2) Nevertheless, the Ito and Stratonovich processes are closely connected with
each other and one may go from one definition to the other by a simple transfor-
mation. To this end we integrate the Ito equation (4.2.1) in a formal manner by
summing up the individual time steps and then going over to the integral. Thus
we obtain from (4.2.1)
1 1
t I
In it the Ito integral is defined by taking q in g'm at a time just before the fluctua-
tion dWm occurs. We have denoted this integral by the sign
(4.3.3)
We now consider a process which leads to precisely the same q,(t) as in (4.3.2)
but by means of the Stratonovich definition
1 1
t t
We shall show that this can be achieved by an appropriate choice of a new driving
force K and new factors g'm' In fact, we shall show that we need just a new force
K, but that we may choose g = g. Here the Stratonovich integral
(4.3.5)
552
4.3 The Stratonovich Calculus 151
is to be evaluated as the limit of the sum with the individual terms (4.3.1). There-
fore we consider
(4.3.7)
(4.3.8)
We put
+ g[q(t;-dl [ w ( tol+tO)
/- 2 / - W(t;_I) ] . (4.3.11)
We expand [lim with repect to (4.3.10) and obtain, taking into account the terms
of order dt,
553
152 4. Stochastic Nonlinear Differential Equations
+ [oglm(q(tH» Kk(q(t;-d)
Oqk
+ ~
2
02 glm (q(t;_I» g (q(/_ »g (q(t-
oqkoql kp [1 jp [I
»]. 1;-21;_1
+ Oglm(q(t;-I» g (q(1
Oqk kp [-1
» IIwp (/;-12+ t;) - W
P
(I
1-1
)]
.
(4.3.12)
The second term on the rhs occurs if glm explicitly depends on time. By means of
(4.3.12), the integral (4.3.6) acquires the form
+ ~
,;.
;=1
oglm(q(t;-d) g (q(t »
"
vqk
kp [-I
[W m (/;-1 + I;) - W (/
2 m [-1
)]215 mp
+O(Ll/Llw). (4.3.13)
An inspection of the rhs of (4.3.13) reveals that the first two sums jointly
define the ito integral. This can be seen as follows. The first sum contains a con-
tribution from dW m over the time interval from (t;-I + t;)/2 till Ii' whereas the
second sum contains the time interval from 1;-1 till (t;-1 + 1;)/2. Thus taking both
time intervals together and summing up over i the sums cover the total time inter-
val. The third sum over i is again of the Ito type. Here we must use a result from
stochastic theory according to which the square bracket squared converges
against dtl2 in probability. Thus the whole sum converges towards an ordinary
integral. These results allow us to write the connection between the Stratonovich
and Ito integrals in the form
(4.3.14)
We are now in a position to compare the ito and Stratonovich processes de-
scribed by (4.3.2 and 4) explicitly. We see that both processes lead to the same
result, as required if we make the following identifications:
Stratonovich Ito
(4.3.15)
554
5. The World of Coupled Nonlinear Oscillators
555
4.4 Langevin Equations and Fokker-Planck Equation 153
In other words, we find exactly the same result if we use a Stratonovich stochastic
equation, but use in it g and K instead of g and K of the ito equation in the way
indicated by (4.3.15). These results allow us in particular to establish a Stratono-
vich-Fokker-Planck equation. If the Stratonovich stochastic equation is written
in the form
(4.3.16)
(4.3.17)
For sake of completeness we mention the Langevin equations which are just
special cases of the Ito or Stratonovich equations, because their fluctuating
forces are independent of the variable q and of time t. Therefore the correspond-
ing Fokker-Planck equation is the same in the Ito and Stratonovich calculus.
Exercises. 1) Why do we need relations (4.2.14, 15) when evaluating the rhs of
(4.2.18)? Hint: while Uj depends on q at time t i , Kk and gkm depend on q at time
ti - 1•
2) How does the explicit expression for the joint probability (4.2.13) read?
Hint: the conditional probability Pw(w, t Iw I = 0,0) is given by
(4.4.1)
556
5.1 Linear Oscillators Coupled Together 155
irrespective of the physical nature of the oscillators. For the following it will be
important to distinguish between linear and nonlinear cases because their
behavior may be entirely different.
(5.1.1)
(5.1.2)
(5.1.3)
(5.1.4)
(5.1.5)
(5.1.6)
(5.1.7)
and the same can be done with (5.1.3). Writing the variables XI, PI> X2, P2 in the
form of a vector
(5.1.8)
we may cast the set of equations (5.1.6) and (5.1.7) and their corresponding equa-
tions with the index 2 into the form
(5.1.9)
557
156 5. The World of Coupled Nonlinear Oscillators
oscillators is entirely solved. Later we shall show that the same kind of solutions
as those derived in Sect. 2.6 holds if the variables xi are considered as con-
tinuously distributed in space.
(5.1.10)
(5.1.11)
where the terms on the right-hand side of these equations provide the coupling.
In order to discuss some features of the solutions of these equations we in-
troduce new variables. We make
(5.1.12)
(5.1.13)
(5.1.14)
and similar equations can be obtained for the second oscillator. Introducing
(5.1.15)
we may cast (5.1.13 and 14) into a single equation. To this end we multiply
(5.1.13) by i and add it to (5.1.14). With use of (5.1.15) we then obtain
(5.1.16)
(5.1.17)
ri' qJi being real, into (5.1.16), dividing both sides by exp ( - i cp), and separating
the real from the imaginary part, we obtain the two equations
(5.1.18)
(5.1.19)
558
5.1 Linear Oscillators Coupled Together 157
Similar equations can be obtained for r2 and (fJ2. These equations represent equa-
tions of motion for the radii rj and phase angles (fJj. Collecting variables in the
form of vectors
(5.1.20)
(::) = (fJ,
(::)=W' (5.1.21)
(::) = r, (5.1.22)
and writing the rhs of the equations for (fJ in the form
( : : ) = g,
(5.1.24)
(5.1.26)
(5.1.27)
559
158 5. The World of Coupled Nonlinear Oscillators
The equation for the phase angle ({Jl then acquires the form
(5.1.28)
(5.1.29)
we obtain instead of (5.1.28) a new equation for ({Jl in lowest approximation, i. e.,
({J~o>,
(5.1.30)
It indicates that in lowest approximation the effect of the nonlinear terms of the
rhs of (5.1.28) consists in a shift of frequencies. Therefore whenever nonlinear
coupling between oscillators is present we have to expect stich frequency shifts
and possibly further effects. In the following we shall deal with various classes of
nonlinearly coupled oscillators. In one class we shall study the conditions under
which the behavior of the coupled oscillators can be represented by a set of,
possibly, different oscillators which are uncoupled. We shall see that this class
plays an important role in many practical cases. On the other hand, other large
classes of quite different behavior have been found more recently. One important
class consists of solutions which describe chaotic behavior (Sect. 8.11.2)
In order to elucidate and overcome some of the major difficulties arising when
we deal with nonlinearly coupled oscillators, we shall consider a special case,
namely equations of the form (5.1.25) where we assume that r is a time-independ-
ent constant.
If there is no coupling between oscillators their phase angles obey equations
of the form
d({J
-=w. (5.2.1)
dt
560
5.2 Perturbations of Quasiperiodic Motion for Time-Independent Amplitudes 159
these oscillators and let it grow to a certain amount. In such a case we have to
deal with equations of the form
drp
-=w+ef(rp)· (5.2.2)
dt
It is assumed that f is periodic in rpl' ... , rpn and that it can be represented as a
Fourier series in the form
The small quantity e in (5.2.2) indicates the smallness of the additional forcej.
Let us assume that under the influence of coupling the solutions do not lose
their quasiperiodic character. As we have seen in the preceding section by means
of the explicit examples (5.1.28,30) the additional termf can cause a frequency
shift. On the other hand, we shall see below that we can treat the perturbation ef
in an adequate way only if the frequencies wi remain fixed. The situation will
turn out to be quite analogous to that in Sect. 2.1.3, where the convergence of
certain Fourier series depended crucially on certain irrationality conditions
between frequencies wi. Indeed, it shall turn out that we must resort precisely to
those same irrationality conditions. To this end we must secure that all the time
the basic frequencies wi remain unaltered. This can be achieved by a formal trick,
namely we not only introduce the perturbationfinto (5.2.1) but at the same time
a counterterm LI(e) which in each order of e just cancels the effect of the
frequency shift caused by f. Therefore we consider instead of (5.2.2) the equation
We shall now devise an iteration procedure to solve (5.2.5). It will turn out that
this procedure, developed by Kolmogorov and Arnold and further elaborated by
Bogolyubov and others, converges very rapidly. We first proceed in a heuristic
561
160 5. The World of Coupled Nonlinear Oscillators
(5.2.9)
Obviously we have chosen the counterterm L1 in such a way that 'P(O) retains the
old frequencies Wj. In ordinary perturbation theory we should now insert (5.2.10)
into rhs of (5.2.5) in order to obtain an improved solution tJ3(1}, i. e.,
d A(l) _
_ 'P_ = (w + L1 + f) + L J m exp [im(wt + 'Po)] . (5.2.11)
dt m*O
This solution can be found directly by integrating the rhs in (5.2.11), yielding
562
5.2 Perturbations of Quasiperiodic Motion for Time-Independent Amplitudes 161
(5.2.16)
wherejis identical with that occurring in (5.2.15). When we insert (5.2.16) into
(5.2.5) we find a rather complicated equation for rp(1) so that nothing seems to be
gained. But when rearranging the individual terms of this equation it transpires
that it has precisely the same form as (5.2.5), with one major difference. On the
rhs terms occur which can be shown to be smaller than those of equation (5.2.5),
i. e., the smallness parameter e is now replaced by e 2 • Continuing this procedure,
we find a new equation in which the smallness parameter is e 4 , then we find e 8,
etc. This indicates, at least from a heuristic point of view, that the iteration pro-
cedure is rapidly converging, as shall be proved in this section.
We insert the hypothesis (5.2.16) into both sides of (5.2.5). Of course, rp(1) is a
function of time. Therefore, when differentiating j with respect to time, by
applying the chain rule we first differentiate j with respect to rp(1) and then rp(1)
with respect to time t.
In this connection it will be advantageous to use a notation which is not too
clumsy. If v is an arbitrary vector with components VI"", un' we introduce the
following notation
(5.2.17)
where lhs is defined by rhs. When we choose v = w we readily find from (5.2.13)
the identity
(5.2.18)
It just tells us that by differentiating (5.2.13) with respect to rp(1) and multiplying
it by w, we obtain! except for the constant term (f(rp(1) = f =!o).
563
162 5. The World of Coupled Nonlinear Oscillators
After these preliminaries we may immediately write the result we obtain when
inserting (5.2.16) into (5.2.5):
This equation can be given a somewhat different form as one can check directly.
The following equation
is identical with (5.2.19) provided we make use of (5.2.18). Dividing both sides of
(5.2.20) by
(5.2.21 )
we obtain
(5.2.22)
We now wish to show that (5.2.22) has the same shape as (5.2.5). To this end we
introduce the abbreviation
(5.2.23)
and abbreviate the remaining terms on the rhs of (5.2.22) by f(1) so that
(5.2.24)
(5.2.25)
564
5.2 Perturbations of Quasiperiodic Motion for Time-Independent Amplitudes 163
which, as predicted, has the same form as (5.2.5), but where LI(l) now replaces LI
andl(l) replaces I (or ef).
We now want to convince ourselves, at least for heuristic purposes, thatl(l)
and LI(l) are smaller by an order of e than the terms in (5.2.5). Since I is
proportional to e [compare (5.2.6)] we have immediately
o{ e, (5.2.26)
OqJ
--(1-) ex
LI(I) ex e. (5.2.27)
(5.2.28)
(5.2.29)
Taking all these estimates together we immediately arrive at the result that 1(1)
goes with e2• Now, the idea is to continue this iteration procedure by making the
following substitutions
(5.2.30)
(5.2.31)
(5.2.32)
and in particular
(5.2.33)
(5.2.34)
Because the last two terms on the rhs of (5.2.25), etc., become smaller and
smaller, we expect that eventually we obtain
. (dqJ(S»)
hm - - =W (5.2.35)
S-+OOdt
565
164 5. The World of Coupled Nonlinear Oscillators
drp(2)
--=W (5.2.39)
dt
drp(3)
--=w (5.2.42)
dt
and therefore
(5.2.44)
(5.2.45)
and repeating this step once again to come back to rp we obtain our final result
566
5.3 Some Considerations on the Convergence of the Procedure 165
This result clearly indicates the construction of rp. It contains terms which
increase linearly with t. The next terms are quasiperiodic functions of time
where, for instance, the second termj(1) depends on a frequency which itself is
again a periodic function of time. Evidently, by continuing this kind of substitu-
tion we obtain a representation of rp in form of a series of termsj<S). In addition,
the arguments of thesej's appear themselves as a series. A proof of the conver-
gence of the procedure must show that the!,s converge provided the initialjis
small enough. In the subsequent section we shall study some of the most interest-
ing aspects of the convergence properties of j<s).
(5.3.1)
on the domain
IIm{rpJI ~ p - 26, (5.3.4)
the inequality
-
If(rp) I ~ -MC -2n1-1 (5.3.5)
K 6 +
C= ( n: 1)
n+l
(1 + e)n . (5.3.6)
567
166 5. The World of Coupled Nonlinear Oscillators
The function (5.3.3) is analytic on the domain (5.3.4). While the constant M is in-
troduced in (5.3.2), the constant K occurs in the inequality of the KAM condition
(5.2.14). In (5.3.4), () is a positive constant which must be suitably chosen as is
explained below. In a similar way one finds the inequality
n+2
C' = ( n: 2 ) (1 + et . (5.3.8)
We now want to show how the inequality (5.3.5,6) can be proven. The reader
who is not so interested in these details can proceed to formula (5.3.30). We
decompose/(rp) into its Fourier series. The Fourier coefficients are given by the
well-known formula
(5.3.9)
Integration over the angles e can be interpreted as an integration over the com-
plex plane where z = r exp (i e), r = 1.
Because/is an analytic function, according to the theory of functions we may
deform the path of integration as long as / remains analytic on the domain over
which the integration runs. This allows us to replace (5.3.9) by
1
- - J ... 0Sf(@+ ItP) exp [-lm(@ + 1 tP)] de! ... den •
21t 21t. . •
(5.3.10)
(271)n 0
(5.3.11)
where
(5.3.13)
568
5.3 Some Considerations on the Convergence of the Procedure 167
(5.3.15)
by taking in each term of the series the absolute value and using (5.3.15). Using
the KAM condition (5.2.14) and (5.3.14) we obtain
- M
If(rp) I ~ - L 11m Il n+ 1 exp (- 20 11m II> . (5.3.17)
Km
It is now rather easy to estimate the sum over m in the following way. We
write
11m I = z. (5.3.20)
The maximum of
vlnz - oz (5.3.21)
(5.3.22)
(5.3.23)
Taking the exponential of both sides of (5.3.23), we find after a slight rearrange-
ment
(5.3.24)
569
168 5. The World of Coupled Nonlinear Oscillators
where the sum over m can be written as the nth power of the infinite sum over a
single variable q = 1, 2, 3, ... so that the rhs of (5.3.25) is
S = (~)V ~ (1 + 2I
e 0 q~l
e - <5 q )n (5.3.26)
S (~)V_lOV (1 + e-<5)n
=
e 1-e-<5
(5.3.27)
Using
(5.3.28)
and thus
Using (5.3.30,17) we obtain the desired result (5.3.5,6). As each member of the
infinite sum (5.3.3) is analytic and this sum (5.3.3) is bounded fcf. (5.3.5)] on the
domain (5.3.4), so (5.3.3) is analytic on the domain (5.3.4).
After this small exercise we resume the main theme, namely, we wish to con-
sider the iteration procedure described in Sect. 5.2 and its convergence. Since/(1)
in (5.2.24) is built up of functions /,1 and derivatives as they appear in (5.3.2,
5, 7) estimates for /(1) can now be derived. Since these estimates are somewhat
lengthy and we shall come to this question in a more general frame in the
Appendix, we shall first give the result of the estimate. One finds that
(5.3.31)
570
5.3 Some Considerations on the Convergence of the Procedure 169
(5.3.32)
(5.3.33)
(5.3.34)
Pl = P - 2J, (5.3.35)
(5.3.36)
and so on.
Because each time Pk must be bigger than 0 the sum over the J's must not
become arbitrarily big. Rather, we have to require that Pk must be bounded from
below so that we can perform the estimates safely in each step. Therefore we have
two conflicting requirements. On the one hand, the J's must be big enough to
make M j [cf. (5.3.34)] converging, on the other hand, the J's must not become
too big but rather must converge to O. The attractive feature of the whole
approach is that J can be chosen such that in spite of this conflict a convergent
approach can be constructed. To this end we put (y < 1)
(5.3.37)
and choose y such that Pk --> p/2 for k --> 00, e. g.,
P
J=y=--«1). (5.3.39)
4+p
571
170 5. The World of Coupled Nonlinear Oscillators
We are now in a position to give explicit estimates for the subsequent M/s.
We introduce the abbreviation P = C"/K so that (5.3.33,34) reads
If (3) 1 ....,~ M 3 = P M~
02n+2 '
(5.3.42)
• 2
• 2
If (s+I)I~M
...., s+1
=p~.
o2n+2
(5.3.43)
s
M =P M; (5.3.44)
s+1
y (s+I)(2n+2)
(5.3.46)
(5.3.47)
Now we require P ~ 1, i. e., the constant K appearing in the KAM condition has
to be chosen small enough as follows from (5.3.40). It is then easy to derive for
the recursion formulas (5.3.40-43) the relation
(5.3.48)
572
5.3 Some Considerations on the Convergence of the Procedure 171
573
6. Nonlinear Coupling of Oscillators:
The Case of Persistence of Quasiperiodic Motion
To illustrate the basic idea let us start with the most simple case of a single oscil-
lator described by an amplitude r and a phase angle rp. In the case of a linear
(harmonic) oscillator the corresponding equations read
rp= w, (6.1.1)
r= O. (6.1.2)
(6.1.3)
...
Note that with increasing time ~ relaxes to 0
a b
Fig. 6.1.1. Visualization of the solution of (6.1.1,3), namely by a particle which moves in an over-
damped way in a potential shown in this figure, and which rotates with angular frequency w
As is well known [1] equations (6.1.1 and 3) can describe a particle moving in
a potential which rotates at a frequency wand whose motion in r-direction is
overdamped. As is obvious from Fig. 6.1.1, the particle will always relax with its
amplitude r(t) towards ro (6.1.4).
If we focus our attention on small deviations of r from ro we may put
(6.1.5)
(6.1.6)
In the following discussion we shall neglect the term O(e) which indicates a
function of ~ of order e
or still higher order. The solutions of the equations
(6.1.1) and (6.1.3, 6) can be easily visualized by means of the potential picture
and can be plotted according to Fig. 6.1.2, where Fig. 6.1.2a shows stationary
motion, while Fig. 6.1.2b shows how the system relaxes when initially ~ was
chosen unequal to zero.
These considerations can be extended to a set of oscillators and to the more
general case in which the number of frequencies Wt, W2, •.• , wn differs from the
number of amplitude displacements ~t, ... , ~m' In such a case equations
(6.1.1, 6) can be generalized to equations
ip(O) = W, (6.1.7)
575
174 6. Nonlinear Coupling of Oscillators: The Case of Persistence of Quasiperiodic Motion
We now introduce a nonlinear coupling between the w's and ~'s, i. e., we are
interested in equations of the form
where/and g are functions which are periodic in tpP) with period 2 n. For mathe-
matical reasons we shall further assume that they are real analytic in tp(l), ell), e,
where e is a small parameter.
As we know from the discussion in Sect. 5.2, the additional term/ may cause
a change of the basic periods 7J = 2 nl Wj. For reasons identical with those of
Sect. 5.2, we must introduce a counterterm in (6.1.9) which is a constant vector
and will guarantee that the basic periods remain the same, even if the interaction
term/ is taken into account. This leads us to consider equations of the form
q, = W + e/(rp, e, e) + A , (6.1.11)
AD=DA (6.1.14)
to d and D. Let us visualize the effect of the additional terms / and g and of the
counterterms A, d and D. In order to discuss these effects let us start with the
case n = m = 1.
We wish to consider a situation in which the quasiperiodic motion, or, in the
present special case, the periodic motion, persists. According to (6.1.7 and 8), the
unperturbed equations are given by
q,(O) = w, and (6.1.15)
They describe a uniform rotation of the angle tp(O), tp(O) = wt and a relaxation of
~(O) towards zero so that the spiral motion of Fig. 6.1.3 emerges. As a special case
576
6.1 The Problem 175
Fig. 6.1.3. Solution to .pIO) = w. ~IO) = - A,;IO) Fig. 6.1.4. The steady-state solution of Fig.
6.1.3
we obtain qJ(O) = wt, .;(0) = 0, which is the motion on the circle of Fig. 6.1.4.
These motions are perturbed when we take into account the additional terms /
and g of (6.1.9, 10) in which case we find
where/and g are 2n-periodic in qJ(l). Let us first consider the case which corre-
sponds to .;(0) = 0, i.e., the circle of Fig. 6.1.4. To visualize what happens we
adopt the adiabatic approximation from Sect. 1.13. We put ~(1) = in (6.1.18).
For small enough e we may express .;(1) uniquely by means of qJ(1), i.e.,
°
(6.1.19)
(In fact this relation may be derived even without the adiabatic approximation as
we shall demonstrate below, but for our present purpose this is not important.)
Relation (6.1.19) tells us that the circle of Fig. 6.1.4 is deformed into some other
closed curve (Fig. 6.1.5). Thus the old orbit .;(0) = 0 is deformed into the new one
(6.1.19). Inserting (6.1.19) into (6.1.17) we obtain an equation of the form
/
/ ' ----~ "-
"
h,
~I'I
\ ~o
\
\
Fig. 6.1.5. Stationary solution to (6.1.17,18), Fig. 6.1.6. Transient solution of (6.1.17, 18),
qualitative plot qualitative plot
577
176 6. Nonlinea r Coupling of Oscillators: The Case of Persistence of Quas iperiodic Motion
It tells us that the rotation speed (p(l) depends on <p(1). Therefore, at least in
general, we must expect that the length of the period of motion around the closed
orbit of Fig. 6.1.5 is different from that of Fig. 6.1.4. When we introduce a
suitable counterterm L1 in the equation
(p = w + d(<p, e) + L1 , (6.1.21)
we may secure that the period remains that of (6.1.15) (Fig. 6.1.4).
Let us now consider the analog of Fig. 6.1.3 . There we considered an initial
state which is somewhat elongated from the stationary state ~(O) = 0 and which
tends to ~(O) = 0 with increasing time. If e is small, the additional term g in
(6.1.18) will deform the spiral of Fig. 6.1.3 to that of Fig. 6.1.6, but qualitatively
~(I) will behave the same way as ~(O), i. e., it relaxes towards its stationary state.
However, the relaxation speed may change. If we wish that ~(1) relaxes with the
same speed as ~(O), at least for small enough ~(1), s, we have to add the above-men-
tioned counterterm D~.
The basic idea of the procedure we want to describe is now as follows. We
wish to introduce instead of IfJ and ~ new variables If! and X, respectively, so that
the Figs. 6.1 .5, 6 of IfJ and ~ can be transformed back into those of Figs. 6.1.4, 3,
respectively. Before we discuss how to construct such a transformation, we
indicate what happens in dimensions where n ~ 2 and/ or m ~ 2. All the essential
features can be seen with help of the case n = 2 and, depending on an adequate
interpretation, m = 2 (or 3). Let us start again with the unperturbed motion
described by (6.1.7 and 8). In the steady state, ~(O) = ~(O) =,0 and we may
visualize the solutions of (6.1.7) by plotting them with coordinates IfJb 1fJ2' The
variables can be visualized by means of a local coordinate system on the torus
(Fig. 6.1. 7).
(6.1.22)
(6.1.8) describes a relaxation of the system's motion towards the torus . To grasp
the essential changes caused by the additional term g in (6.1.10) , let us again
adopt the adiabatic approximation so that ~\l), ~il) may be expressed by IfJP), 1fJ~l)
through
578
6.1 The Problem 177
par' 01 I
tfaJectory
Clearly, the original torus is now distorted into a new surface (Fig. 6.1.8). In it,
the uniform rotation of <pfO), <p~O) is replaced by a nonuniform rotation of IPf1), IP1 1)
due to the additional term! in (6.1.9). Consequently, in general the correspond-
ing periods 11 and 0. will be changed. To retain the original periods we must
introduce the counterterms Lll' Ll 2 • Let us now consider the relaxation of ~P), ~~1)
towards the deformed surface. While originally according to our construction (or
visualization) the rectangular coordinate system ~fO), ~1°) was rotating with <pfO),
<p1°) uniformly along the torus, ~f1), ~~1) may form a new coordinate system which
is no longer orthogonal and in which the relaxation rate is different from ..1.1' ..1.2
due to the additional term gin (6.1.10). The counter term DC; allows us to correct
for the locally distorted coordinate system ~P), ~il) by making it orthogonal again
and by restoring the original relaxation rates. If the eigenvalues of A are complex
(or purely imaginary), D makes it possible to keep in particular the original A
fixed while d may correct (in the sense of an average) for the distortion of the
torus. Therefore, due to!, g and the counterterms, the trajectories are deformed
in three ways:
1) The local coordinate system <P on the torus is changed (see Fig. 6.1.9a).
2) The radial displacement has been changed or, more generally, the position of
the surface elements of the torus has been changed (Fig. 6.1.9b).
3) The orientation of the coordinate system ~ has been changed (Fig. 6.1.9c).
b <>
f~~ol Fig. 6.1.9a - c. Local changes due to the deformation of a
torus. (a) New coordinates on the torus. (b) A displacement
L ~~Ol of elements of the torus. (c) Change of the coordinate
system pointing away from the torus
c
579
178 6. Nonlinear Coupling of Oscillators: The Case of Persistence of Quasiperiodic Motion
which takes into account that the rotation speed is not uniform but changes
during the rotation. The factor 8 in front of u indicates explicitly that this term
tends to 0 and therefore ({J coincides with IJI when we let 8 tend to 0 in (6.1.11, 12).
With respect to ~, we have to take into account that the deformed trajectory
differs from the un deformed trajectory (Fig. 6.1.7) by dilation in radial direction
and by rotation of the local coordinate system. Therefore we make the following
hypothesis
where v takes care of the dilation in radial direction, whereas the second term
V· X takes care of the rotation of the local coordinate system. Because IJI is an
angular variable, we require that u, v, V are periodic in IJI with period 2 Tr.
Furthermore, these functions must be analytic in IJI, ~ and 8. We wish to show
that by a suitable choice of u, v, V and the counterterms LI, d, D, we may trans-
form (6.1.11,12) into equations for the new variables IJI and X which describe
"undistorted trajectories" (Fig. 6.1. 7) and which obey the equations
ip = w + O(X)' (6.1.26)
i=AX+ O (X 2 ). (6.1.27)
The terms 0 describe those terms which are neglected and O(X k ) denotes an
analytic function of IJI, X, 8 which vanishes with x-derivatives up to order
k - 1 ;:;, 0 for X = O. In other words, we are interested in a transformation from
(6.1.11, 12) to (6.1.26, 27), neglecting terms denoted by O. When we specifically
choose the solution
(6.1.28)
x=o,
the solution of (6.1.26) reads
IJI = wt (6.1.29)
(up to an arbitrary constant lJIo). Inserting (6.1.28,29) into (6.1.24, 25), we find
({J = wt + w(wt, 8) and (6.1.30)
580
6.2 Moser's Theorem (Theorem 6.2.1) 179
E jvwv + E kJJA JJ
n m
i for (6.2.1)
v=t /l=1
IEkJJI~ 1, (6.2.3)
(6.2.4)
We require that of the expressions (6.2.1) only finitely many vanish, namely only
for j = O. Obviously this implies that Wt, ... ,W n are rationally independent.
2) For some constants K and r with '
we require
for all integers mentioned under point (1), for which lhs of (6.2.6) does not
vanish. Condition (6.2.6) may be considered as a generalized version of the KAM
condition. Note that Ilj II is defined in the present context by
581
180 6. Nonlinear Coupling of Oscillators: The Case of Persistence of Quasiperiodic Motion
rp = VI + eu' + 0(e 2 )
Lt = eLI' + 0(e 2 )
t, = X + e(V' + V' X) + 0(e 2 ) (6.3.5)
d = ed' + 0(e 2 )
D = eD' + 0(e 2 ) •
We insert (6.3.3, 4) into (6.3.1), using (6.3.5), and keep only terms up to order e.
Since (6.3.1,2) are vector equations and we have to differentiate for instance u
with respect to rp, which itself is a vector, we must be careful with respect to the
notation. Therefore, we first use vector components. Thus we obtain
582
6.3 The Iteration Procedure 181
(6.3.7)
(6.3.8)
(6.3.9)
We now insert (6.3.3, 4) using (6.3.5) into (6.3.2) and keep only terms up to order
e, where we consider again the individual components p. Using the chain rule of
differential calculus we obtain for lhs of (6.3.2)
(6.3.13)
or in vector notation
X=AX· (6.3.14)
Furthermore we consider
(6.3.15)
583
182 6. Nonlinear Coupling of Oscillators: The Case of Persistence of Quasiperiodic Motion
v~. (6.3.16)
We equate the terms which are independent of X in (6.3.11, 12). Making use of
(6.3.13,15,16), we readily obtain (specializing to X = 0)
V~W_AV'=g(If/,O,O)+d' . (6.3.17)
In the last step we equate the terms linear in X which occur in (6.3.11, 12). We
denote the matrix, whose elements /1, v are given by
(6.3.18)
(6.3.19)
Furthermore we consider
(6.3.20)
(6.3.21)
Collecting the equations (6.3.10, 17,22) we are led to the following set of
equations (setting X = 0)
1
u~w -A' =/(If/,O,O) ,
In these equations the functions/, g, gx are, of course, given, as are the constants
wand A. In the present context, the quantities If/ can be considered as the inde-
pendent variables, whereas the functions u I, v I, V' are still to be determined. The
equations (6.3.23) can be easily brought into a form used in linear algebra. To
this end we rearrange the matrix V' into a vector just by relabeling the indices,
e. g., in the case
(6.3.24)
584
6.3 The Iteration Procedure 183
(6.3.25)
(6.3.26)
We replace the matrix g)( in the same way by the vector G and introduce the
vector
F~ [~] (6.3.27)
Following the same steps with ,1', d', D', we introduce the vector
(6.3.28)
(6.3.29)
n
L2 = L wvo/Olf/v- A, (6.3.31)
v~l
n
L3= L wvo/Olf/v-(A'-A"), (6.3.32)
v~l '--v--'
A
where A' and A" are again matrices stemming from A via rearrangement of
indices. By means of (6.3.26-29) we may cast (6.3.23) in the form
585
184 6. Nonlinear Coupling of Oscillators: The Case of Persistence of Quasiperiodic Motion
LU=F+LI. (6.3.33)
For the moment, the use of L may seem to be roundabout, but its usefulness will
become obvious below. When we insert (6.3.35) into (6.3.34) we readily obtain
(6.3.36)
(6.3.37)
(6.3.38)
In the notation of the original indices inherent in (6.3.23), the last equation
(6.3.38) reads
(6.3.39)
These equations show clearly what the eigenvalues of L are: since the equations
for Uj' OJ, VJ are uncoupled, we may treat these equations individually. From
(6.3.36) we obtain as eigenvalue:
by diagonalizing (6.3.37)
i(j,w)-A Il , (6.3.41)
and by diagonalizing (6.3.38), which can be most simply seen from (6.3.39),
(6.3.42)
586
6.3 The Iteration Procedure 185
For j =F 0 we obtain
(6.3.46)
Because the eigenvalues do not vanish, these equations can be solved uniquely for
the unknown u I, v I, V'.
We now turn to the eigenvectors Uj belonging to the vanishing eigenvalues,
obtained according to assumption (1) in Sect. 6.2 only for j = O. Thus the eigen-
vectors of the null-space obey the equations
o· U o= 0, (6.3.47)
A· Vo = 0, (6.3.48)
which follow from (6.3.36 - 38) and the condition that (6.3.40 - 42) vanish.
We now consider the inhomogeneous equations which correspond to
(6.3.43 - 45) for j = O. They read
-Ll'=!o' (6.3.50)
Avo = 0, (6.3.53)
d ' is determined uniquely by (6.3.51). Similarly, from (6.3.49) it follows that D'
is determined uniquely by (6.3.52). What happens, however, with the coefficients
uo, vo, Vo? According to linear algebra, within the limitation expressed by
(6.3.47 - 49), these null-space vectors can be chosen freely. In the following we
put them equal to zero, or in other words, we project the null-space out. Let us
summarize what we have achieved so far. We have described the first step of an
iteration procedure by which we transform (6.3.1, 2) into (6.1.26,27) by means
of the transformations (6.3.3, 4). This was done by determining u, v, V and the
counterterms LI, d, D in lowest approximation. We have shown explicitly how the
587
186 6. Nonlinear Coupling of Oscillators: The Case of Persistence of Quasiperiodic Motion
(6.3.54)
i. e., we have to deal with the problem of small divisors (Sects. 2.1.3 and 5.2). We
shall treat it anew in Appendix A.1, making use of our assumption that f and g
are analytic functions in rp and c;.
2) We have to prove the convergence of the iteration procedure as a whole.
This will shed light on the choice of the null-space vectors in each iteration step.
As it will turn out, any choice of these vectors will give an admitted solution
provided the e-sum over the null-space vectors converges. In connection with this
we shall show that there are classes of possible transformations (6.3.3,4) which
are connected with each other. We turn to a treatment of these problems in
Appendix A.
588
7. Nonlinear Equations. The Slaving Principle
The main objective of this book is the study of dramatic macroscopic changes of
systems. As seen in the introduction, this may happen when linear stability is
lost. At such a point it becomes possible to eliminate very many degrees of
freedom so that the macroscopic behavior of the system is governed by very few
degrees of freedom only. In this chapter we wish to show explicitly how to
eliminate most of the variables close to points where linear stability is lost. These
points will be called critical points. It will be our goal to describe an easily
applicable procedure covering most cases of practical importance. To this end
the essential ideas are illustrated by a simple example (Sect. 7.1), followed by a
presentation of our general procedure for nonlinear differential equations (Sects.
7.2 - 5). While the basic assumptions are stated in Sect. 7.2, the final results of
our approach are presented in Sect. 7.4 up to (and including) formula (7.4.5).
Section 7.3 and the rest of Sect. 7.4 are of a more technical nature. The rest of
this chapter is devoted to an extension of the slaving principle to discrete noisy
maps and to stochastic differential equations of the Ito (and Stratonovich) type
(Sects. 7.6-9).
7.1 An Example
u = au - us, (7.1.1)
s= -ps+ u 2 • (7.1.2)
We assume that
p>O. (7.1.4)
When we neglect the nonlinear terms u . sand u 2, respectively, we are left with
two uncoupled equations that are familiar from our linear stability analysis.
Evidently (7.1.1) represents a mode which is neutral or unstable in a linear
188 7. Nonlinear Equations. The Slaving Principle
stability analysis, whereas s represents a stable mode. This is why we have chosen
the notation u (unstable or undamped) and s (stable). We wish to show that s may
be expressed by means of u so that we can eliminate s from (7.1.1, 2). Equation
(7.1.2) can be immediately solved by integration
t
s(t) = I e- P(t-r)u 2 (r)dr, (7.1.5)
-00
where we have chosen the initial condition s = 0 for t = - 00. The integral in
(7.1.5) exists if 1u( r) 12 is bounded for all r or if this quantity diverges for r -+ - 00
less than exp ( 1pr I). This is, of course, a self-consistency requirement which has
to be checked.
We now want to show that the integral in (7.1.5) can be transformed in such a
way that s at time t becomes a function of u at the same time t only.
where we identify v with exp [- p(t- r)] and w with u 2• We thus obtain
(7.1. 7)
We now consider the case in which u varies very little so that Ii can be con-
sidered as a small quantity. This suggests that we can neglect the integral in
(7.1.7). This is the adiabatic approximation in which we obtain
(7.1.8)
Solution (7.1.8) could have been obtained from (7.1.2) by merely putting
(7.1.9)
on lhs of (7.1.2). Let us now consider under which condition we may neglect the
integral in (7.1.7) compared to the first term. To this end we consider
(7.1.10)
in (7.1. 7). We may then put (7.1.10) in front of the integral which can be
evaluated, leaving the condition
(7.1.11 )
590
7.1 An Example 189
(7.1.12)
We now use the still exact relation (7.1. 7), substituting it in it by - u . s according
to (7.1.13)
(7.1.14)
This is an integral equation for s(t) because s occurs again under the integral at
times T. To solve this equation we apply an iteration procedure which amounts to
expressing s by powers of u. In the lowest order we may express s on the rhs of
(7.1.14) by the approximate relation (7.1.8). In this way, (7.1.14) is transformed
into
(7.1.15)
(7.1.16)
(7.1.17)
We split (7.1.17) into an integral containing u 6 and another part containing still
higher-order terms
591
190 7. Nonlinear Equations. The Slaving Principle
(7.1.18)
(7.1.19)
It is now obvious that by this procedure we may again and again perform partial
integrations allowing us to express s(t) by a certain power series in u(t) at the
same time. Provided u is small enough we may expect to find a very good
approximation by keeping few terms in powers of u. Before we study the
problem of the convergence of this procedure, some exact relations must be
derived. To this end we introduce the following abbreviations into the original
equations (7.1.13, 2), namely
(7.1.21)
P=P(u). (7.1.22)
s = s(u), (7.1.23)
so that we henceforth assume that this substitution has been made. We wish to
derive a formal series expansion for s(u). We start from (7.1.5), written in the
form
t
s(u(t» = J e- fJ(t- r) P(u)rdr. (7.1.24)
As before, we integrate by parts identifying P(u) with win (7.1.6). The differen-
tiation of P(u) with respect to time can be performed in such a way that we first
differentiate P with respect to u and then u with respect to time. But according to
(7.1.20) the time derivative of u can be replaced by Q(u, s) where, at least in prin-
ciple, we can imagine that s is a function of u again. Performing this partial in-
tegration again and again, each time using (7.1.20), we are led to the relation
592
7.1 An Example 191
(~)
dt 00
= Q(U,S(U» ~
au
(7.1.26)
by which we define the left-hand side. It becomes obvious that (7.1.25) can be
considered as a formal geometric series in the operator (7.1.26) so that (7.1.25)
can be summed up to give
P (1 + ~ (:t t)
or, in the original notation,
1 1
s(t) = - P. (7.1.28)
P (1 + Q~~)
au p
Multiplying both sides by the operator
(1 +Q~~)
au p
(7.1.29)
(7.1.30)
as P(u,s)-ps
(7.1.31)
au Q(u,s)
as (7.1.32)
au
Thus we have found a first-order differential equation for s(u).
In the subsequent section we shall show how all these relations can be con-
siderably generalized to equations much more general than (7.1.1, 2). In practical
applications we shall have to evaluate sums as (7.1.25) but, of course, we shall
approximate the infinite series by a finite one. In this way we also circumvent
593
192 7. Nonlinear Equations. The Slaving Principle
difficulties which may arise with respect to the convergence of the formal series
equation (7.1. 25).
Rather, in practical cases it will be necessary to estimate the rest terms. To
this end we split the formal infinite expansion (7.1.25) up in the following way
s(t) = -
1 m( - Qa- -1)nP(u,s) + -1 L
L 00 (
-Q- -
a 1,)nP(u,s).
fJn=O au fJ fJn=m+l au fJ,
\ .....- I
r (7.1.33)
Again it is possible to sum up the rest term
r=-
1
L -Q- -
00 ( a 1 )n+m+l
P(u,s(u» , (7.1.34)
fJ n=O au fJ
which yields
1
r=-------
( a 1 )m+l P.
-Q-- (7.1.35)
fJ 1 a au fJ
( 1+ Q - - )
au fJ
can be used to estimate the rest term. The term Q introduces powers of at least
order u 3 while the differentiation repeatedly reduces the order by 1. Therefore, at
least, with each power of the bracket in front of (7.1.36), a term u 2 is introduced
in our specific example. At the same time corresponding powers 1/fJ are intro-
duced. This means that the rest term becomes very small provided u 2/ fJ is much
smaller than unity. On the other hand, due to the consecutive differentiations
introduced in (7.1.35 or 36), the number of terms increases with m!. Therefore
the rest term must be chosen in such a way that m is not too big or, in other
words, for given m one has to choose u correspondingly sufficiently small. This
procedure is somewhat different from the more conventional convergence
criteria but it shows that we can determine s as a function of u to any desired
accuracy provided we choose u small enough.
Let us now return to our explicit example (7.1.32). We wish to show that
(7.1.32) can be chosen as a starting point to express s explicitly by u. We require
that rhs of (7.1.32), which we repeat
as (7.1.37)
au -us
594
7.1 An Example 193
remains regular for simultaneous limits s-+O and u-+O. To solve (7.1.37) we
make the hypothesis
(7.1.39)
(7.1.41)
(7.1.42)
so that
(7.1.43)
2
C2= - 3 . (7.1.44)
P
We are thus led to the expansion
u2 2u 4
S=-+-3-+'" (7.1.45)
P P
which agrees with our former result (7.1.19).
Starting from (7.1.1, 2) (in which our above treatment with a = 0 is included),
we can achieve the result (7.1.37) in a much simpler way, namely writing
(7.1.1,2) in the form
.
11m --
LI t
.1t .... O
(LlU) == u. = au - us and (7.1.46)
595
194 7. Nonlinear Equations. The Slaving Principle
lim
,1t~O
(,1,1 S)
t
= Ii = _ fJs + U2 • (7.1.47)
ds - fJs + u2
(7.1.48)
du au- us
This is a relation very well known from autonomous differential equations refer-
ring to trajectories in a plane [1]. Although our series expansion seems round-
about, its usefulness is immediately evident if we are confronted with the ques-
tion in which way (7.1.48) can be generalized to nonautonomous differential
equations in several variables. We are then quickly led to methods we have out-
lined before.
Thus in the next section we wish to generalize our method in several ways. In
practical applications it will be necessary to treat equations in which the control
parameter a is unequal zero, so that we have to treat equations of the form
u = au - us. (7.1.49)
Furthermore, the equations may have a more general structure, for instance
those of the slaved variables may have the form
(7.1.51)
Also in many applications we need equations which are sets of differential equa-
tions in several variables u and s.
Other types of equations must be considered also, for instance
where cP is a function which is periodic in qJ. Equations of the type (7.1.52) con-
taining several variables qJ may also occur in applications. Last but not least, we
must deal with stochastic differential equations
(7.1.53)
in which Fs(t) is a random force. This random force renders the equations to in-
homogeneous equations and an additional difficulty arises from the fact that
596
7.2 The General Form of the Slaving Principle. Basic Equations 195
Consider a set of time dependent variables iit. ii 2, ... , S1> S2,"" f/Jj, f/J2,'" ,
lumped together into the vectors U, s, f/J, where the sign - distinguishes these
variables from those used below after some transformations. The notation u and
S will become clear later on where u
denotes the "unstable" or "undamped"
modes of a system, whereas s refers to "stable" modes. The variables f/J play the
role of phase angles, as will transpire from later chapters. We assume that these
variables obey the following set of ordinary differential equations 1
Here A~, A~' are real diagonal matrices. While the size of the elements of A~'
may be arbitrary, the size of the elements of A~ is assumed to be of order o.
Further, A~" contains the nondiagonal elements of Jordan's normal form (Sect.
2.4.2). Also the matrix As is assumed in Jordan's normal form. It can be decom-
posed into its real and imaginary parts A;, A;' , respectively
(7.2.5)
Yi ~ fJ < O. (7.2.6)
1 It is actually not difficult to extend our formalism to partial differential equations with a suitable
norm (e. g., in a Banach space)
597
196 7. Nonlinear Equations. The Slaving Principle
The quantities
(7.2.7)
are of order t5 2 or smaller and Q, P, must not contain terms linear in u or s (but
may contain u • s). The functions
Q,P,R (7.2.8)
are 2 n-periodic in qJ, the functions (7.2.7) are assumed to be bounded for
- 00 < t < + 00 for any fixed values of it, s, (p. The functions i are driving forces
which in particular may mimic stochastic forces. However, we shall assume in
this section that these forces are continuous with respect to their arguments.
In subsequent sections we shall then study stochastic forces which can be
written in the form
where Fo describes a Wiener process. In such a case the matrix G must be treated
with care because appropriate limits of the functions it, s, (p must be taken, e. g.,
in the form
(7.2.12)
We define
(7.2.13)
L Ajk k2Uk
k j k2 j j
Uk .
2
(7.2.14)
598
7.2 The General Form of the Slaving Principle. Basic Equations 197
Here Asuu may be a 2 n-periodic function in ({J and may still be a continuous func-
tion of time t
(7.2.15)
R(u,s, rp, t) = Ro(rp, t) + R j (rp, t): u + R 2(rp, t): s + terms of form P . (7.2.16)
We shall assume that Ro is of order 15 2 and R j , R2 are of order 15. We now make
the transformations 2
rp=wt+({J (7.2.18)
to the new time-dependent variables u, ({J. This transforms the original equations
(7.2.1 - 3) into
where the last equations in each row are mere abbreviations, whereas the expres-
sions in the middle column of (7.2.19 - 21) are defined as follows:
(7.2.22)
(7.2.28)
2 The transformation of a vector by means of exp (,1 u t) == exp [(iA~' + A~") tl has the following
effects (Sect. 2.6): while the diagonal matrix A~' causes a multiplication of each component of the
vector by exp (i wjt), where Wj is an element of A~' , A~" mixes vector components with coefficients
which contain finite powers of t. All integrals occurring in subsequent sections of this chapter exist,
because the powers of t are multiplied by damped exponential functions of t.
599
198 7. Nonlinear Equations. The Slaving Principle
As exemplified in Sect. 7.1, s may be expressed by u at the same time. Now let us
assume that such a relation can be established also in the general case of
(7.2.19 - 21) but with suitable generalizations. Therefore, we shall assume that s
can be expressed by u and rp and may still explicitly depend on time
Later it will be shown how s may be constructed in the form (7.3.1) explicitly,
but for the moment we assume that (7.3.1) holds, allowing us to establish a
number of important exact relations. Since s depends on time via u and rp but also
directly on time t, we find by differentiation of (7.3.1)
d a . as . as
-s=-s+u-+ rp--, (7.3.2)
dt at au arp
(7.3.3)
600
7.3 Formal Relations 199
u
Furthermore, we express and ip by the corresponding rhs of (7.2.19,21). This
enables (7.3.2, 4) to be simplified to
The bracket on the lhs represents an operator acting on s. The formal solution of
(7.3.5) can be written as
0 a a
s= ( - - A s + Q - + R -
)-1 P(u,s(u,rp,t),rp,t). (7.3.6)
at au 0'1'
At this point we remember some of the results of Sect. 7.1. There we have
seen that it is useful to express s as a series of inverse powers of p, to which As
corresponds. As will become clear in a moment, this can be achieved by making
the identifications
(7.3.7)
a a (7.3.8)
Q-+R-=B
aU 0'1'
(A + B) -1 = A -I - A -I BA -1 + ...
which is valid for operators (under suitable assumptions onA and B). In (7.3.9)
we use the definition
(7.3.10)
The proof of (7.3.9) will be left as an exercise at the end of this section.
Using (7.3.9) with the abbreviations (7.3.7, 8) we readily obtain
s= a
(--As
at
)- 1 ~ [ -
v=O
(0
Q - + Ra- ) (--As
au 0'1'
a
at
)- 1 P. ]V (7.3.11)
601
200 7. Nonlinear Equations. The Slaving Principle
In practical applications we will not extend the series in (7.3 .11) up to infinity but
rather to a finite number. Also from a more fundamental point of view the con-
vergence of the formal series expansion is not secured. On the other hand, for
practical applications we give an estimate of the rest term and we shall derive here
some useful formulas. To define the rest term we decompose s according to
(7.3.9)
m
s=A- 1 E(-BA- 1YP+r, (7.3.13)
v=o
(7.3.14)
(For the proof consult the exercises at the end of this section.)
For practical applications we note that
Q-+R-
a a (7.3.15)
au aqJ
t
S= J exp [As(t- T)]P(U,s, qJ, T)dT, where (7.3.16)
-00
S=S(U,qJ,T). (7.3.17)
g(t)h(u,s,qJ) . (7.3.18)
JV'wdt=vw-Jvw'dt. (7.3.19)
602
7.3 Formal Relations 201
Expressing the temporal derivatives it, ~ by the corresponding rhs of (7.2.19, 21),
we may write (7.3.16) in the new form
t
S = exp(Ast) S exp( - As r)P(u (t),s(u (t), rp(t), t), rp (t), r) dr
-00
t
-exp(Ast) S (0
Q-+R-0) dr'Sr' exp(-Asr)P(u(r'), ... ,r)dr.
-00 ou orp r' -00 (7.3.20)
t
S = S exp [As(t - r)]P(u (t), ... , r) dr
-00
The first part of (7.3.22) can be considered as the formal solution of the equation
o
- s = Ass + P(u,s, rp, t), (7.3.23)
ot
where u, s, rp on the rhs are considered as given parameters. Under this assump-
tion the formal solution of (7.3.23) can also be written in the form
(7.3.24)
The first term of (7.3.22) therefore provides us with the interpretation of the
inverse operator A -I. In this way we may write (7.3.22) in the form
- i exp[AsCt-r')] (Q~+R~)
-00 ou orp r'
(~-As)-Ip(u(r,),
or
... ,r')dr"
603
202 7. Nonlinear Equations. The Slaving Principle
A-I f
v=m+1
(-BA-1)"=(A+B)-I·(-BA-I)m+l. (7.3.27)
604
7.4 The Iteration Procedure 203
A comparison with (7.3.11) shows further that the operator in front of p(n-m)
can be defined as
-1
( ~-A )
( ~-A ) -1 L IT -~
[() ( ~-A ) -1 ]
dt S (m) dt S (0) all i;;,1 dt (i) dt S (0) ,
products [I = m
(7.4.6)
where we have used the abbreviation
(
d _ A)-I _ (~_ A)-1 (7.4.7)
dt S (0) at S
the relation
(7.4.8)
(dd)t (i)
f(u, qJ, t) = (Q(i+l)~ + R(i)~)f(U' qJ, t).
au aqJ
(7.4.9)
The transition from (7.3.11) to (7.4.1) with the corresponding definitions can be
easily done if we exhibit the order of each term explicitly. We have, of course,
assumed that we may express Q(i+l) and R(i) by u, qJ and t. According to our
previous assumptions in Sect. 7.2, we also have the relations
and (7.4.10)
(7.4.11)
605
204 7. Nonlinear Equations. The Slaving Principle
Its proof is of a more technical nature, so that we include it only for sake of com-
pleteness. In order to prove (7.4.12), we insert (7.4.5) on both sides of (7.4.12)
which yields
= (~_AS)-lp(n)_ (~-A
dt (0) dt
)-l nf (~)
s (0) m= 1 dt (m)
-1
n-2-m
. E ( ~- As ) p(n-m-I).
1=0 dt (I)
(7.4.13)
Obviously, the term with m = 0 cancels so that it remains to prove the relation
= -
n-2 n-2-m'
E
m' = 1
E
1=0
(dtd
--As )-1 (dtd ) (dtd
(0)
-
(m')
--As )-1
(I)
p(n-m'-l).
(7.4.14)
m'+[=m }
O~[~n-2-m' (7.4.15)
1~m ~n-2
- n-2
E Em
m=lm'=1
(dtd
--As )-1 (dtd ) (dtd )-1
(0)
-
(m')
--As
(m-m')
p(n-m). (7.4.16)
We now compare each term for fixed m on the lhs of (7.4.13) with that on the rhs
and thus are led to check the validity of the relation
(- d
dt
- As )-1
(m)
=
m'
m (d
E - -
= 1 dt
As )-1 (
(0)
- - d)
dt (m')
(d
-
dt
- As )-
(m-m')
1. (7.4.17)
606
7.5 An Estimate of the Rest Term. The Question of Differentiability 205
(~-A
dt
)-1 s (0) m'=l
~ (-~dt ) (~-A
dt
)-1
(m') s (0)
.L II
[)=m-m'
(- -dtd) (d-dt- A s)-1
(i) (0)
(7.4.18)
(~-As)-1 L II [( - ~) (~-As)-1]
dt (0) ,£i=m dt (i) dt (0)
(7.4.19)
(7.5.1)
[compare (7.1.26)]. Let us consider the case in which in P only a variable u and
not s occurs, and let us assume that
holds. Clearly
(7.5.4)
holds provided As is diagonal and IYi Imin is the smallest modulus of the negative
diagonal elements of As. We further assume Q in the form
607
206 7. Nonlinear Equations. The Slaving Principle
with (7.5.5)
Ihl<M. (7.5.6)
We readily obtain
(~)2
dt ""
P _ Ih(t) I (~)2 kuk(2k-1)U2k-2
IYi Imin
(7.5.8)
(~)m
dt ""
p -Ih(t) I (~)m k(2k-1)(3k- 2) ... (mk- (m-1»u(m+ 1)k-m.
IYilmin (7.5.9)
(7.5.10)
Similar estimates also hold for more complicated expressions for P and Q (and
also for those of R). The following conclusion can be drawn from these con-
siderations. The rest term goes with some power of u where the exponent
contains the factor m. Also all other factors go with powers of m. Thus, if u is
chosen sufficiently small these contributions would decrease more and more. On
the other hand, the occurrence of the factorial m! leads to an increase of the rest
term for sufficiently high m. In a strict mathematical sense we are therefore
dealing with a semiconvergent series. As is well known from many practical
applications in physics and other fields, such expansions can be most useful.
In this context it means that the approximation remains a very good one
provided u is sufficiently small. In such a case the first terms will give an excellent
approximation but when m becomes too big, taking higher-order rest terms into
account will rather deteriorate the result.
In conclusion, we briefly discuss the differentiability properties of s(u, rp, t)
with respect to the variables u and rp. As indicated above, we can approximate s
by a power series in u with rp- (and t-) dependent coefficients to any desired
degree of accuracy, provided we choose bin (7.4.1) small enough. If the rhs of
(7.2.19-21) are polynomials (and analytic) in u and are analytic in rp in a certain
domain, any of the finite approximations s(n) (7.4.1) to s have the same (analy-
608
7.6 Slaving Principle for Discrete Noisy Maps 207
ticity) property. Because of the properties of the rest term discussed above, with
an increasing number of differentiations a/aqJj, a/aUk> the (differentiated) rest
term may become bigger and bigger, letting the domain of u, in which the
approximation is valid, shrink, perhaps even to an empty set. This difficulty can
be circumvented by "smoothing". Namely, we consider s(n) (omitting the rest
term) as a "smoothed" approximation to the manifold s(u, qJ, t). Then s(n) pos-
sesses the just-mentioned differentiability (or analyticity) properties. If not
otherwise mentioned, we shall use s(n) later in this sense.
The introduction showed that we may deal with complex systems not only by
means of differential equations but also by means of maps using a discrete time
sequence. The study of such maps has become a modern and most lively branch
of mathematics and theoretical physics. In this section we want to show how the
slaving principle can be extended to this case. Incidentally, this will allow us to
derive the slaving principle for stochastic differential equations by taking an
appropriate limit (Sect. 7.9).
Consider a dynamic system described by a state vector q 1 which is defined at a
discrete and equidistant time sequence I. The evolution of the system from one
discrete time 1 to the next 1+ 1 is described by an equation of the general form
(7.6.1)
in which f and G are nonlinear functions of q1 and which may depend explicitly
on the index I. Here 111 is a random vector, whose probability distribution may
but need not depend on the index I. In many physical casesft is of the order O(ql)
(7.6.2)
(7.6.3)
and
(7.6.4)
609
208 7. Nonlinear Equations. The Slaving Principle
Let us assume that S can be expressed by a function of U and / alone. This allows
us to consider dP in (7.6.4) as a function of U and / alone
(7.7.1)
We first show that (7.7.2) fulfills (7.7.1). Inserting (7.7.2) into (7.7.1) we
readily obtain for the lhs
I
SI+I - SI = L (1 + Asdm)l-rn dP(u rn , urn-I"'" m)
m=-oo
I-I
- L (1 + A sdm)l-l- rn dP(u rn , urn-I"'" m). (7.7.3)
m=-oo
Taking from the first sum the member with m = / separately we obtain
SI+I - SI= dP(u{J UI_I"'" /)
I-I
+(1 + Asd/) L (l+A sdm)I-I- rn dP(u m,u m_\J···,m)
m=-oo
I-I
- L (l+A sdm)I-I- mdP(u m,u m_I, ... ,m). (7.7.4)
m=-oo
610
7.7 Formal Relations 209
According to its definition, the sum in the last term is [compare (7.7.2)]
identical with SI, so that rhs of (7.7.5) agrees with rhs of (7.7.1). To study the
convergence of the sum over m we assume that the factor of
(7.7.6)
(7.7.7)
l
is a k x k matrix given by
Mk =
o 1o 1
o
0 1
0
··.1
j . (7.7.8)
o
In its first diagonal parallel to the main diagonal, Mk is equal to 1, whereas all
other elements are O. We treat the case that I> 0 (and is, of course, an integer).
According to the binomial law
(7.7.9)
(7.7.11)
provided
1(1+A')I<I. (7.7.13)
611
210 7. Nonlinear Equations. The Slaving Principle
Decomposing A' into its real and imaginary parts A, and Ai' respectively, we
arrive at the necessary and sufficient condition
(7.7.14)
A,<O (7.7.15)
holds. Throughout this chapter we assume that conditions (7.7.14, 15) are ful-
filled.
In order to establish formal relations we introduce the abbreviation
(7.7.16)
(7.7.17)
(7.7.18)
(7.7.19)
(7.7.20)
m=-oo
(7.7.21)
(7.7.22)
612
7.7 Formal Relations 211
Using operator techniques given in (7.3.9), one readily establishes the following
identity
(7.7.27)
Because of (7.7.21), we may expect that the curly brackets in (7.7.27) possess an
analog to the explicit form of rhs of (7.7.21). This is indeed the case.
Let us assume for the moment that we may decompose dP into a sum of
expressions of the form [compare (7.3.18)]
(7.7.28)
where h is a function of the variables U alone, but does not depend explicitly on
m, whereas dg is a vector which depends on m alone. It is then simple to verify
the following relation
I
~ (1 + Asdmi-mh(um,um_I," .)dg(m)
m=-oo
I
=h(UI,UI_I,"') ~ (1 + Asdm)l-mdg(m)
m=-oo
I m-I
~ (1+A sdm)I+I-m ~ (1+A sdm)m-I-m'{"'}m,m' (7.7.29)
m= -00 m'=-oo
(7.7.30)
The essence of (7.7.29) can be expressed as follows. While lhs contains U at all
previous times, the first term on rhs contains U only at those earlier times which
are initially present in h(UI,UI_I," .). Thus, instead of an infinite regression we
have only a finite regression. From a comparison of (7.7.26) with (7.7.29) we
may derive the following definition
I
~ (1 + Asdm)l-mdP(ul, ul_I,"" m), (7.7.31)
m=-oo
613
212 7. Nonlinear Equations. The Slaving Principle
where on rhs u/, u/_!> etc., are kept fixed and the summation runs only over m.
Equation (7.7.26) with (7.7.27) may be iterated, giving
(7.7.32)
00
(7.7.34)
Since a more explicit discussion of the rest term entails many mathematical
details going far beyond the scope of this book, we shall skip the details here.
While the operator containing ,1 (I) is defined by (7.7.31), we have still to
explain in more detail how to evaluate the operator ,1(1;). It has the following
properties:
1) It acts only on u" u/_!> ....
2) For any function f of u, it is defined by (7.7.24)
(7.7.35)
where (7.7.36)
(7.7.37)
(7.7.38)
In the following we shall assume that the functions f can be considered as poly-
nomials.
614
7.7 Formal Relations 213
4) In order to treat (7.7.36) still more explicitly we use (7.7.38) and the following
rules
.d~)u'=u7-u7_1= (
VI
L
+ v2=n-l
UllU/~1).d_UI. (7.7.39)
(7.7.40)
(7.7.41)
(7.7.43)
(7.7.45)
(7.7.46)
Vo+···+vx=n
(7.7.47)
Vo+···+vx=n
615
214 7. Nonlinear Equations. The Slaving Principle
It can be rearranged to
(7.7.48)
(7.7.49)
x
= ELl_u rn + 1+ v • (7.7.51)
v=o
The sum in (7.7.50) in front of the parenthesis can again be considered as the
symmetrized derivative, while the parentheses in (7.7.50) can be expressed by
individual differences uk+ 1 - Uk·
Taking all the above formulas together we obtain again a well-defined procedure
for calculating s as a function of u and I, provided dP is prescribed as a function
of u and 1 alone. In practice, however, dP depends on s. Therefore we must
devise a procedure by which we may express s by u and 1stepwise by an inductive
process.
To this end we introduce a smallness parameter o. In general, dQ and dP
contain a non stochastic and a stochastic part according to
We assume that Au in (7.6.3) is of the order 0 and that the functions occurring in
(7.8.1, 2) can be expressed as polynomials of u and s with I· dependent coef-
ficients. The coefficients are either continuous functions (contained in Qo and
Po) or quantities describing a Wiener process [in dFo(l)] which is defined by
616
7.8 The Iteration Procedure for the Discrete Case 215
(4.2.2, 3). We assume that the constant terms which are independent of u and s
are of smallness 0 2, the coefficients of the linear terms of smallness 0, while u
and s are of order 0, 0 2, respectively. In order to devise an iteration procedure we
represent s in the form
(7.8.5)
dP= r
k=2
dP(k). (7.8.7)
We now proceed in two steps. We apply (7.7.32) with (7.7.33 and 27) on dP, but
on both sides we take only the term of the order Ok.
Starting from (7.7.45) we define g~k-k') as a function which is precisely of
order Ok-k' and dQ(k') as a function being precisely of order k'. We put
(u)(k) _
LL 'I-Ik I-I'
Igv
(k-k') - (k')
(ul,· .. ,ul'_l)dQ
,
( ... ,I+v), (7.8.8)
k' =0 v=O
which is obviously of order Ok. After these preparations the final formula is given
by
together with
(7.8.10)
C(k) = {- Asdl} I
kt +···+kr =k"=l
.n [.. ']kPt-k')dl,
'
(7.8.11)
617
216 7. Nonlinear Equations. The Slaving Principle
where
(7.8.12)
2) If I becomes a continuous variable and there are no fluctuating forces we
refer to the results of Sect. 7.4. For convenience we put dl = r. Using the nota-
tion of that section we immediately find
(7.8.14)
We are now in a position to derive the corresponding formulas for stochastic dif-
ferential equations. To introduce the latter, we consider
Although it is not difficult to include the equations for qJ [cf. Sects. 7.2, 4] and a
qrdependent dQ and dP in our analysis, to simplify our presentation they will be
omitted.
Because of the Wiener process we must be careful when taking the limit to a
continuous time sequence. Because of the jumps which take place at continuous
times we must carefully distinguish between precisely equal times and time dif-
ferences which are infinitesimally small. The fundamental regression relation to
be used stems from (7.7.29) and can be written in the form
t
J exp [AS<t- r)] h(U"U,_dt'U,_2dt, .. Jdg(r)
-00
t
= J exp[As(t- r)]dg(r)h(ut,Ut_dt, .. ·)
-00
t ,-dt
- J exp[As(t- r')]d_h(u"U,_dt,· .. ) J exp[As(r- r')dg(r')
-~ -00
(7.9.3)
618
7.9 Slaving Principle for Stochastic Differential Equations 217
with
(7.9.4)
t
J exp[As(t- r)]dg(r)h(u t ,U t -dt,U t -2dt,···)
-00
t
= J exp[As(t- r)] dg(r)h(u t ,u t ,U t -2dt,· .. )
-00
t
- J exp[As(t- r)] dg(r)[h(u t ,u t ,U t -2dt,· .. )
-00
(7.9.5)
(7.9.6)
where we shall assume the Ito-calculus, the convention to sum over indices which
occur twice, and that Wi describes a Wiener process.
It is not difficult to include in our approach the case where F u, i depends on U
and s at previous times and on time integrals over dWj over previous times (see
below). We shall further assume that dWi dWk converges in probability according
to
(7.9.8)
We use the Ito transformation rule for a function <p(u) according to w'hich
619
218 7. Nonlinear Equations. The Slaving Principle
(7.9.10)
In the next step the expression on rhs of (7.9.5) will be inserted for the stable
mode into the equation for the unstable mode. We now have to distinguish
between two cases:
a) s occurs in
Qo(u,s,t)dr. (7.9.11 )
Because (7.9.10) gives rise to terms of order O(dt), O(dw), we obtain terms of
the form dt 2 or dt dw, which can both be neglected compared to dr.
b) s occurs in
(7.9.13)
(7.9.14)
We may continue in this way which proves that we may replace everywhere
Ut-kdt, k = 1, 2, ... , by u t .
We now consider the second term on rhs of (7.9.3), i. e.,
t T-dt
S exp[As(t-r)]d_h(uT'u T_1>"') S exp[As(r-r')]dg(r'). (7.9.15)
When we use the definition of d_ of (7.9.4), we may write (7.9.15) in the form
t
S exp[As(t- r)][h(uT'uT-dt,· .. ) - h(UT-dt, UT-2dt,· .. )]
-00
T-dt
. J exp[As(r-r')]dg(r'). (7.9.16)
620
7.9 Slaving Principle for Stochastic Differential Equations 219
Using
iFu.
1
[(
St expAst-r )] [kL fjh(U r-dt,U r-2dr,···) [QOiUr-ndt,·
( .. ,r-n d t )d r
-00 n=! fjui,r-ndt
r-dt
X S exp[As(r- r')]dg(r'). (7.9.18)
-00
t fjh({u}) r
S exp[As(t- r)] QOi({u})dr S exp[As<r- r')dg(r'). (7.9.19)
-00 fj~ -00
Because we shall use the last sum in the curly brackets in the same form as it
occurs in (7.9.19), we have to deal only with the second term containing F
linearly.
We now study this term
Our goal is to replace the u's occurring in h with different time arguments by u's
which depend on a single time argument only. All u's standing in front of u r- nt
can be replaced by ur-ndt' and similarly all u's following ur-ndt can be replaced
by ur-(n+!)dt. Similarly the arguments inF can be replaced by ur-ndt. These sub-
stitutions give contributions of higher order and can therefore be rigorously
neglected in the limit dt ~ O. In this way (7.9.20) is transformed into
621
220 7. Nonlinear Equations. The Slaving Principle
St exp[A s (t-r)] ~
~
Oh({Ur-(n-lldt},Ur-ndP{Ur-(n+lldt})
" .
F.
UI'P
({u _
r ndtJ
I)
-00 n-l uU"r-ndt
r-dt
. dwp(r-ndt) S exp[A,(r- r')]dg(r') + O(dr) + O(dw). (7.9.21)
-00
t exp[A (t-r)]
S s
~
":'
lOh(Ur-ndt,{Ur-(n+lldt})
" .
F. (fu _
U"p l r ndt
})dw (r-ndt)
p
-00 n-I uU"r-ndt
r-dt
X S exp[A,(r-r')]dg(r'). (7.9.23)
(7.9.24)
622
7.9 Slaving Principle for Stochastic Differential Equations 221
We can now proceed in complete analogy to the discrete time case by devising an
iteration procedure in which we collect terms of the same order of magnitude,
i.e., the same powers of u. To this end we define the operator
(~)
dt (m)
= (u(m+l)~)
ou
(7.9.26)
and u(m+ 1) means that we have to take from (7.9.17) only those terms which
contain the m + 1 power of U (after s has been replaced by u). After these steps we
are in a position to write the final result. Accordingly, s(t) can be written as
elk) = ( ~ - As)
dt
-1
I: n[..
(O)k 1 + ... +k v =k'i=1
'kdP(k-k')
'
with (7.9.28)
(7.9.29)
623
8. Nonlinear Equations. Qualitative Macroscopic
Changes
In this and the subsequent chapter we deal with a problem central to synergetics,
namely qualitative macroscopic changes of complex systems. Though it is
possible to treat the various instabilities under the impact of noise by means of a
single approach, for pedagogical reasons we shall deal with the special cases in-
dividually. For the same reasons we first start with equations which do not
contain fluctuations (noise) and shall treat the corresponding problems only
later. The general philosophy of our approach was outlined in Sect. 1.14.
We start with probably the simplest case, namely the bifurcation of solutions
from a node. Our starting point is a set of nonlinear differential equations for the
state vector q(t), namely
q(t) = N(q(t), a) . (8.1.1)
(8.1.6)
where L depends on a but does not depend on time t. As we know from Sect. 2.6,
the solutions of (8.1.7) have the general form
(8.1.8)
If the eigenvalues Ak(a) of the matrix L(a) are nondegenerate, the vectors v are
time independent. In case of degeneracy, v(k) may depend on powers of t.
In order not to overload our presentation we shall assume in the following
that the vectors v (k) are time independent which is secured if, e. g., the eigen-
values Ak are all different from each other, i. e., nondegenerate. Since the vectors
v (k) form a complete set in the vector space of q, we may represent the desired
solution q as a superposition of the vectors V(k). Therefore, the most general
solution may be written in the form
(8.1.9)
w(t)
where the coefficients c;k(t) are still unknown time-dependent variables. It will be
our first task to determine c;k' To this end we insert (8.1.9) into (8.1.1)
625
224 8. Nonlinear Equations. Qualitative Macroscopic Changes
N(qo(a), a) = o. (8.1.12)
According to (8.1.6 or 7) and using (8.1.8), the second term can be cast into the
form
2Nf~"
The higher-order terms are defined in a similar fashion. Since we want to derive
equations for ~b we should get rid of the vectors v (k). To this end we use the dual
eigenvectors introduced in Sect. 2.5. These eigenvectors,
-(k)
V , (8.1.15)
which are in the present case as v time independent, have the property
(8.1.16)
Taking the scalar product of (8.1.10) with (8.1.15) and making use of the decom-
position (8.1.11), we find the following equations
(8.1.17)
where we have introduced the abbreviation A (2). The higher-order terms are, of
course, defined in an analogous fashion, giving the final result
(8.1.18)
We numerate the eigenvalues in such a way that the one with the biggest real pan
carries the index 1. We assume that this eigenvalue is real and nondegenerate and
that the control parameters are such that Ai (a) changes its sign from a negative to
a positive value. We shall focus our attention on a situation in which
(8.2.1)
626
8.2 A Simple Real Eigenvalue Becomes Positive 225
holds and in which the real parts of all other eigenvalues are still negative. Since
(8.2.1) in the linear stability analysis indicates that the corresponding solution is
unstable, whereas
indicates that all other solutions are still stable, we shall express this fact by a
change of notation. To this end we put for
(8.2.3)
where s refers to "stable". We must stress here that our notation may lead to
some misunderstanding which must be avoided at any rate. Note that "unstable"
and "stable" refer to the linear stability analysis only. We shall show that in
many cases, due to the nonlinearities, the mode which was formerly unstable in
the linear analysis becomes stabilized, and it will be our main purpose to explore
the new stable regions of u.
So the reader is well advised to use u and s as abbreviations distinguishing
those modes which in the linear stability analysis have eigenvalues characterized
by the properties (8.2.1, 2). After these introductory remarks we split (8.1.18)
according to the indices 1 and 2, ... , respectively, into the following equations
U - /\.uu
• - 1
+ A u(2) u 2 + A u(3) u 3 + ... + ~
IJ
A uk
(2)
u · Sk + ... , (8.2.5)
k
In this section and Sect. 8.3 we shall assume u real. Provided u is small enough
and (8.2.2) holds, we may apply the slaving principle. This allows us to express Sk
as a function of u which may be approximated by a polynomial
(8.2.7)
(8.2.8)
where C is given by
(8.2.9)
627
226 8. Nonlinear Equations. Qualitative Macroscopic Changes
(8.2.10)
(8.2.12)
For those readers who are acquainted with mechanics we note that (8.2.12) can
be visualized as the overdamped motion of a particle in a force field which has a
potential V according to the equation
.
u= ---.
av (8.2.13)
au
In the present case V is given by
u2 u4
V= -A - +
u 2
/3-,
4
(8.2.14)
which easily allows us to visualize the behavior of the solution u of the nonlinear
equation (8.2.12). An inspection of the potential curve V reveals immediately
that depending on the sign of Au two entirely different situations occur. For
Au < 0 the value u = 0 is a stationary stable solution. This solution persists for
Au> 0 but becomes unstable. The new stable solutions are given by
u = ± VAul/3 . (8.2.15)
Note that this solution becomes imaginary for negative Au and must therefore be
excluded for that branch of AU"
Taking the roots (8.2.15) and
u=O (8.2.16)
628
8.2 A Simple Real Eigenvalue Becomes Positive 227
(8.2.18)
for the individual roots, we may write (8.2.17) in the more condensed form
(8.2.19)
.
U= - - - .
av (8.2.21)
au
u= Uk + c5u, (8.2.23)
linearize (8.2.20) around uk> and study the stability by means of the linearized
equations
629
228 8. Nonlinear Equations. Qualitative Macroscopic Changes
y = aP(u) I :; : o. (8.2.26)
au u=uk
The temporal evolution in the nonlinear domain can be studied also because
(8.2.20) is a differential equation whose variables can be separated so that we im-
mediately find
du'
J
U
= (- to· (8.2.27)
Uo Auu' + feu')
Using (8.2.22) the integrand can be decomposed into partial fractions so that
the integral can be explicitly evaluated.
Exercise. Evaluate the integral in (8.2.27) in the case (8.2.19) and determine u(t).
We assume that a number M of the eigenvalues with the greatest real parts coin-
cide and that they are real
(8.3.1)
and are zero or positive while all other eigenvalues have negative real parts. In
extension of the previous section we denote the mode amplitudes which belong to
(8.3.1) by
(8.3.2)
Splitting the general system (8.1.18) into the mode amplitudes (8.3.2) and the
rest and applying the slaving principle we find the order parameter equations
(8.3.4)
In principle, these equations are of precisely the same form as the original equa-
tions (8.1.1). However, the following two points should be observed. In contrast
to the general equations (8.1.1), the number of variables of the reduced equations
(8.3.3) is in very many practical cases much smaller. Especially in complex
630
8.3 Multiple Real Eigenvalues Become Positive 229
systems, where we deal with very many degrees of freedom, an enormous reduc-
tion of the number of degrees of freedom is achieved by going from (8.1.1) to
(8.3.3). Furthermore, when eigenvalues are degenerate, symmetry properties can
often be used. Here we indicate how to cope with (8.3.3) in the special case in
which stationary solutions
(8.3.5)
exist. Note, however, that (8.3.3) may also have solutions other than (8.3.5), for
instance oscillatory solutions. We shall come back to such equations later. Here
we make a few comments for the case when (8.3.5) holds.
A rather simple case occurs when the rhs of (8.3.4) can be written as a
gradient of a potential. A class of such problems is treated by catastrophe theory
([Ref. 1, Chap. 5] and additional reading material given at the end of the present
book). We shall not dwell on that problem here further. Another class of
equations of the form (8.3.4) is provided by
(8.3.6)
(8.3.7)
(8.3.8)
Ur
-2 +-2 =
U~
Au(a). (8.3.9)
a b
631
230 8. Nonlinear Equations. Qualitative Macroscopic Changes
Our starting point is the set of differential equations (8.1.1). Again we assume
that for a certain control parameter a = ao a stable time-independent solution qo
exists. We assume that when we change the control parameter, this time-inde-
pendent solution can be extended and we again perform a stability analysis. We
first consider
w=Lw (8.4.1)
(8.4.2)
where (8.4.4)
A; ~ o. (8.4.5)
(8.4.6)
which allows us to use the slaving principle for small enough u (which is a new
notation of et).
Because (8.4.4) is a complex quantity, it is plausible that u is complex.
Exhibiting only the terms up to 3rd order in u, the order parameter equation for
u reads
In the following we shall focus our attention on important special cases for
(8.4.8) which occur repeatedly in many applications.
632
8.4 A Simple Complex Eigenvalue Crosses the Imaginary Axis. Hopf Bifurcation 231
if = 0, (8.4.13)
(8.4.14)
Equation (8.4.14) is identical with (8.2.12), with the only difference being that r
must be a positive quantity. Therefore the stationary solutions read
Because (8.4.13) holds for all times, the transient solution can be readily found
by solving (8.4.14) [1].
The stationary solution has the form
It shows that (for ro > 0) the total system undergoes an harmonic oscillation or,
when we plot u in the phase plane consisting of its real and imaginary parts as
coordinates, we immediately find a limit cycle. The transient solution tells us that
all points of the plane converge from either the outside or the inside towards this
limit cycle. An interesting modification arises if the constant b occurring in
(8.4.9) is complex
We assume
633
232 8. Nonlinear Equations. Qualitative Macroscopic Changes
Inserting (8.4.21) into (8.4.20) and separating real and imaginary parts we find
and (8.4.22)
These are two coupled differential equations, but the type of coupling is very
simple. We may first solve (8.4.22) for r(t) in the stationary or the transient state
exactly. Then we may insert the result for r2 into (8.4.23) which can now be
solved immediately. We leave the discussion of the transient case as an exercise
for the reader and write down the solution for the steady state
This is an important result because it shows that the frequency of the oscilla-
tion depends on the amplitude. Thus, in the case of Hopf bifurcation we shall not
only find new values for r, i. e., a nonzero radius of the limit cycle, but also
shifted frequencies. The present case can be generalized in a very nice way to the
case in which (8.4.20) is replaced by the more general equation
(8.4.25)
Let us assume that f and g are real functions of their arguments and are of the
form of polynomials. Then again the hypothesis (8.4.21) can be made. Inserting
it into (8.4.25) and splitting the resulting equation into its real and imaginary
parts, we obtain
(8.4.26)
(8.4.27)
Equation (8.4.26) has been discussed before [compare (8.2.20)]. The additional
equation (8.4.27) can then be solved by a simple quadrature.
634
8.5 Hopf Bifurcation, Continued 233
Let us now consider a more general case than (8.4.25) which one frequently meets
in practical applications. In it the order parameter equation has the following
form
and (8.5.3)
h = 1m {e-i(Pf}lr . (8.5.6)
(8.5.7)
(8.5.8)
(8.5.9)
(8.5.10)
635
234 8. Nonlinear Equations. Qualitative Macroscopic Changes
We assume b ' > O. While for A~ < 0 the solution, = 0 is stable, it becomes
unstable for A~ > 0, so that we make the hypothesis
, ='0 + ~ (8.5.11)
and require
(8.5.12)
(8.5.13)
Since we want to study the behavior of the solutions of (8.5.1) close to the
transition point A~ = 0, the smallness of A~ is of value. To exhibit this smallness
we introduce a smallness parameter e,
(8.5.14)
(8.5.15)
(8.5.16)
and time by
(8.5.17)
(8.5.18)
Introducing the scaling transformations (8.5.14 -18) into (8.5.13) we readily find
(8.5.19)
(8.5.20)
636
8.5 Hopf Bifurcation, Continued 235
W
This equation shows that instead of the original frequency w occurring in (8.5.4),
we now have to deal with the renormalized frequency where w
(8.5.22)
Because f, is a small quantity we recognize that on the new time scale r under con-
w
sideration, is a very high frequency. We want to show that therefore g and h in
(8.5.19) and (8.5.21), respectively, can be considered as small quantities. From
(8.5.21) we may assume that in lowest approximation exp(i<p) "'" exp(iwr) is a
very rapidly oscillating quantity.
Now consider a function of such a quantity, e. g., {h or h2 • Since (/2 and h2 are
functions of exp(i<p), they can both be presented in the form
(8.5.23)
Let us assume that g and h are small quantities. Then it is possible to make a per-
turbation expansion in just the same way as in the quasiperiodic case in Chap. 6.
As we have seen we then have to evaluate an indefinite integral over (8.5.23)
which acquires the form
2
_f,_ L ~ eimwrh2 • (8.5.24)
iw m*,O m
(8.5.25)
However, the factor f,2 in front of the sum in (8.5.24) can be arbitrarily small.
This means g2 and h2 can be treated as small quantities. In addition, according to
our assumptions above, the <p-independent terms gl and hI are still of higher
order and can again be considered as small quantities so that g and h act as small
perturbations.
In order to bring out the smallness of the terms gand h explicitly we introduce
a further scaling, namely by
(8.5.26)
h=f,lfi (8.5.27)
637
236 8. Nonlinear Equations. Qualitative Macroscopic Changes
jointly with
1'/=e'fj. (8.5.28)
(8.5.29)
(8.5.30)
respectively. Lumping all quantities which contain the smallness parameter e'
together, we may write (8.5.29, 30) in the final forms
dfj
and (8.5.31)
dr
(8.5.32)
where Wu is given by
(8.5.33)
For reasons which will transpire immediately, we rewrite (8.5.31 and 32) in the
form
(8.5.34)
(8.5.35)
where g and hare 2 n-periodic in ({J [and must not be confused with g and h in
(8.5.3,4)].
Equations (8.5.34, 35) are now in a form which allows us to make contact
with results in Sect. 6.2, where we presented a theorem of Moser that will now
enable one to cope with such equations [compare Eqs. (6.1.9, 10)].
It should be noted that the full power of that theorem is not needed here
because it deals with quasiperiodic motion whereas here we deal with periodic
motion only, which excludes the problem of small divisors. (We may treat
(8.5.34,35) more directly by perturbation expansions also.) But nevertheless we
shall use similar reasoning later on in the more complicated case of quasiperiodic
motion and therefore we present our arguments in the simpler case here.
Let us compare (8.5.34,35) with (6.1.9,10). There the aim was to introduce
counterterms D and L1 (and perhaps d) such that the quantities A and OJ were not
changed by the iteration procedure.
638
8.5 Hopf Bifurcation, Continued 237
Here our iteration is somewhat different because (8.5.34, 35) do not contain
these counterterms. Rather, we have to encounter the fact that due to the pertur-
bations g and h, A and W will be changed. It is rather simple to make contact
with the results of Sect. 6.1 nevertheless. To this end we put - Xu = Au [where
this Au must not be confused with Au in (8.4.6)] and write
Au=Au- D + D (8.5.36)
~
Ar
and similarly
Wu = Wu - L1 + L1 . (8.5.37)
~
Wr
Introducing (8.5.36, 37) into (8.5.34, 35), respectively, we cast these equations
into precisely the form required by Moser's theorem. This procedure is well
known from physics, namely from quantum field theory where the quantities
with the index u, e. g., w u , are referred to as "unrenormalized" quantities
whereas the quantities Wn Ar are referred to as "renormalized". According to
Moser's theorem we may calculate D and L1 explicitly from (8.5.34, 35) provided
we use the decomposition (8.5.36, 37). Then D and L1 become functions of An Wr
and e so that we have the relations
Ar = Au - D(An Wn e) , (8.5.38)
Wr = Wu - L1 (An Wn e) . (8.5.39)
(8.5.40)
(8.5.41)
(8.5.42)
639
238 8. Nonlinear Equations. Qualitative Macroscopic Changes
(8.5.43)
(8.5.44)
(8.5.50)
(8.5.52)
The only step still to be performed is to specialize these results to the case of
our equations (8.5.34, 35). Making the identifications
(8.5.53)
(8.5.54)
(8.5.55)
640
8.6 Frequency Locking Between Two Oscillators 239
It gives us the form of the solution close to a bifurcation point. From (8.5.58) we
may readily conclude that the solution is stable and relaxes towards a steady state
given by
(8.6.2)
(8.6.3)
(8.6.4)
where the constant frequencies wj are such that IfIj,j = 1, 2 remains bounded for
all times. We shall call wj renormalized frequency. Inserting (8.6.3), e.g., in
(8.6.1) leads us to
(8.6.5)
641
240 8. Nonlinear Equations. Qualitative Macroscopic Changes
which can be split into its real and imaginary parts, respectively,
(8.6.6)
(8.6.7)
A similar equation arises for r2' ({J2' Let us assume that Cl is a small quantity
so that in a first approximation the corresponding term can be neglected in
(8.6.6). This allows us to find in a good approximation the stationary solution
rl,o by
(8.6.8)
and (8.6.9)
(8.6.10)
(8.6.14)
'P d'P'
t=J---- (8.6.15)
'Po Q - a sin 'P'
In it 'Po denotes the initial value of 'P at time t = O. The behavior of the integral
on the rhs of (8.6.15) is entirely different depending on the size of a and Q. If
(8.6.16)
642
8.6 Frequency Locking Between Two Oscillators 241
the integrand in (8.6.15) never diverges. In particular, we may expand the de-
nominator into a power series with respect to a. In that case the integral can be
represented in the form
Therefore, aside from small pulsations we find for (8.6.15) the relation
(8.6.18)
or (8.6.19)
(8.6.20)
This means that the renormalized frequencies w} retain the same distance as the
frequencies Wj without nonlinear interaction. Quite different behavior occurs,
however, if the condition
(8.6.21)
is fulfilled. Then the integral diverges at a finite 'I' or, in other words, t = 00 is
reached for that finite '1'. Perhaps aside from transients, the solution of the dif-
ferential equation (8.6.13) reads
W:Z = w{ , (8.6.24)
i. e., if the two frequencies acquire the same size or, in other words, if they are
locked together. Also the phases ({J2 and ({Jt are locked together by means of
(8.6.22).
It should be noted that such frequency locking effects can occur in far more
general classes of equations for which (8.6.1,2) are only a very simple example.
Exercise. Treat three coupled oscillators and show that frequency locking may
occur between the frequency differences W2 - WI and W3 - W2 provided
(8.6.25)
643
242 8. Nonlinear Equations. Qualitative Macroscopic Changes
is small enough.
Take as an example the equations
}
UI = (A' + iwdul - b l u l !UI!2 + CIU~Ut
U2 = (A' + iW2)u2 - b 2u Z !U2!Z + C2U!UIU3 (8.6.26)
U3 = (A' + iW3)u3 - b3U3!U3!2 + C3U~Ut
Hint: use the same decomposition as (8.6.3, 4) and introduce as a new variable
(8.6.27)
Form a suitable linear combination of the equations for qJj which results in an
equation for (8.6.27).
In the preceding chapter we learned that a system which was originally quiescent
can start a coherent oscillation or, in other words, its trajectory in the corre-
sponding multi-dimensional space of the system lies on a limit cycle. When we
change the external control parameter further it may happen that the motion on
such a limit cycle becomes unstable and we wish to study some of the new kinds
of motion which may evolve.
We first perform a general analysis allowing us to separate the order para-
meters from the slaved mode amplitudes and then we shall discuss a number of
typical equations for the resulting order parameters. We shall confine our ana-
lysis to autonomous systems described by nonlinear evolution equations of the
form
where q (t) is, in general, a multi-dimensional vector describing the state of the
system. We assume that we have found the limit cycle solution of (8.7.1) for a
certain range of the control parameter
(8.7.2)
We note that farther away from the solution (8.7.2) other possible solutions of
(8.7.1) for the same control parameter may exist, but these shall not be consider-
ed.
We now change the value of the control parameter a and assume that qo can
be continued into this new region. Because of the assumed periodicity we require
(8.7.3)
644
8.7 Bifurcation from a Limit Cycle 243
T= T(a). (8.7.4)
In it the matrix
L = (Lij) (8.7.8)
(8.7.9)
Because N contains the periodic function qo and depends on the control para-
meter a, L depends on time t and control parameter a
L = L(t, a) (8.7.10)
According to Sect. 2.7 we know that the solutions of (8.7.7) can be written in the
form
(8.7.12)
(8.7.13)
645
244 8. Nonlinear Equations. Qualitative Macroscopic Changes
(8.7.14)
which by comparison with (8.7.7) reveals that one of the solutions of (8.7.7) is
identical with qo
(8.7.15)
Evidently WI always points along the tangent of the trajectory at each point qo
at time t. Because the solutions of (8.7.7) span a complete vector space we find
that at each time t this space can be spanned by a vector tangential to the trajec-
tory and n -1 additional vectors which are transversal to the trajectory. Since we
wish to study new trajectories evolving smoothly from the present limit cycle, we
shall construct a coordinate system based on the coordinates with respect to the
old limit cycle. We expect that due to the nonlinearities contained in (8.7.6) the
new trajectory will shift its phase with respect'to the original limit cycle, and that
it will acquire a certain distance in changing directions away from the limit cycle.
Therefore we shall construct the new solution q(t) for the evolving trajectory
using the phase angle CP(t) and deviations ~(t) away from the limit cycle. Thus,
our hypothesis reads
(8.7.16)
W
The prime at the sum means that we shall exclude the solution (8.7.15), i. e.,
k *' 1, (8.7.17)
because this direction is taken care of by the phase angle CP(t). In (8.7.16) W is a
function of ';k(t), t, and CP(t), i. e.,
1 The dot C) means derivative with respect to the total argument, I + <P, of q, v, or t of <P,
respectively.
646
8.7 Bifurcation from a Limit Cycle 245
We now use the fact that (8.7.2) fulfills (8.7.1), and also use (8.7.7) and the de-
composition (8.7.12) to obtain
(8.7.23)
In Sect. 2.5 it was shown that we may construct a bi-orthogonal set of functions
Vk (k = 1, 2, ... ), [compare (2.5.16)]. Multiplying (8.7.23) with VI and using the
orthogonality property we readily obtain
(8.7.27)
(8.7.28)
(8.7.29)
647
246 8. Nonlinear Equations. Qualitative Macroscopic Changes
We remind the reader that Fh, k = 1,2, ... ,depends on t + (/) viaq 0, qo == VI,
and Vk' Let us study this behavior in more detail. Because qo is a smooth and
periodic function we may expand it into a Fourier series
w= 2n1T. (8.7.31)
ip=w+w(/) (8.7.35)
(8.7.37)
(8.7.38)
(8.7.39)
We note that! and gk are 2 n-periodic in rp. Equation (8.7.38) can be cast into a
vector notation in the form
(8.7.40)
648
8.8 Bifurcation from a Limit Cycle: Special Cases 247
Equations (8.7.36, 40) are still exact and represent the final result of this section.
We now study special cases.
and that
For small enough ~ we may invoke the slaving principle of Chap. 7 which
allows us to do away with all the slaved variables with k = 3, 4, ... and to derive
equations for the order parameters alone,
~(t)
(fJ(t)
== u(t), real
J parameters
order
.
(8.8.3)
According to the slaving principle the original set of (8.7.36, 40) reduces to the
equations
(8.8.6)
Each coefficient gj may be decomposed again into a constant term and another
one which is 2 :n:-periodic in (fJ
(8.8.7)
so that
2rr
Jd(fJ gj,2«(fJ) = 0
o
(8.8.8)
649
248 8. Nonlinear Equations. Qualitative Macroscopic Changes
(8.8.9)
(8.8.10)
2n
jdrp!j,I(rp) =0. (8.8.11)
o
g2 = II = 0, g3 == - b <0. (8.8.12)
From Sect. 8.4 it will be clear that the resulting equations (8.8.4, 5) acquire a
form treated there, with one difference, however, which we shall explain now.
When we make the hypothesis
U = Uo + 11 (8.8.13)
and require
(8.8.14)
Uo = ± VA-ul b (S.S.15)
which are distinguished by the sign in front of the square root. We note that in
the case of Hopf bifurcations we must choose the positive sign because ro has to
be positive. Here, however, we take the two signs. Either choice of Uo takes
(S.S.4, 5) [under the assumption of (S.S.12)] to equations which we know from
Sect. S.4 so that we need not treat this case here again explicitly. We can rather
discuss the final result.
To illustrate the resulting solution we consider the case in which the original
limit cycle lies in a plane. Then the two solutions (8.S.15) mean that two limit
cycles evolve to the outer and inner side of the limit cycle. According to the ana-
lysis of Sect. 8.7, the phase motion may suffer a shift of frequency with respect to
the original limit cycle. Furthermore, the phase may have superimposed on its
steady motion oscillations with that renormalized frequency. Also deviations
from the new distance U o may occur which are in the form of oscillations at the
renormalized frequency w,. If the limit cycles lie in a higher-dimensional space,
the newly evolving two trajectories need not lie in a plane but may be deformed
curves.
650
8.8 Bifurcation from a Limit Cycle: Special Cases 249
(8.8.16)
(8.8.17)
For simplicity we assume b real though this is not essential. It is assumed that
there may be additional terms to (8.8.17) but that these terms can be treated as a
perturbation. It is assumed with (8.8.17) that the equation for qJ simply reads
qJ = w. (8.8.18)
We shall show that (8.8.17) allows for a solution which oscillates at half the
frequency or, in other words, whose period has doubled with respect to the
original limit cycle. To this end we make the hypothesis
(8.8.19)
(8.8.20)
It allows for a time-independent stable solution Yo = 0 for A~ < 0 and for solu-
tions (besides Yo = 0)
One may readily show that the solutions (8.8.21) are stable (provided A~ > 0).
Thus we have shown that a stable solution with double period exists. Of course,
in the general case one now has to study higher-order terms and show that they
can be treated as perturbations.
651
250 8. Nonlinear Equations. Qualitative Macroscopic Changes
(8.8.22)
u = ye iwtln , (8.8.23)
Yo= ° or (8.8.25)
(8.8.26)
By linear stability analysis we can show that (8.8.23) with (8.8.27) is a stable solu-
tion provided Au > O. While here we have assumed that A is real and positive, one
may readily extend this analysis to a complex A and complex b provided the real
part of A> O.
Let us briefly discuss examples in which the other remaining terms omitted in
our above analysis can be treated as perturbation. Let us consider an equation of
the form
(8.8.28)
where the perturbative terms are of order US or higher. Making the hypothesis
(8.8.29)
we find
(8.8.30)
From a comparison between this equation and (8.5.1), it transpires that further
analysis is similar to that in Sect. 8.5. Thus we must make the transformation
y = Yo + 11 (8.8.31)
with Yo
652
8.8 Bifurcation from a Limit Cycle: Special Cases 251
We now present an example indicating that there are cases in which it is not at
all evident from the beginning which terms can be considered as perturbations.
Some kind of bifurcation may occur in which, if one solution is singled out, one
class of terms can be considered as small, but if another class of solutions is
singled out, another class of terms can be considered as small. Let us consider an
example,
(8.8.33)
where Au and b are assumed real. Then in lowest approximation we may choose
or (8.8.34)
(8.8.35)
as a solution. Depending on the choice (8.8.34 or 35), (8.8.33) is cast into the
form
These equations are the same type as (8.5.3, 4), already discussed in detail (here r
is complex). There, by suitable scaling, we have put in evidence that the terms
containing exp( ± i tp) can be considered as small perturbations. Thus (8.8.33)
allows for two equivalent solutions, (8.8.34, 35) in whichy is a constant on which
small amplitude oscillations at frequencies m . wl2 are superimposed. In physics
the neglect of the term exp(± itp) compared to 1 (where tp = 2wt) is known as
"rotating wave approximation" .
Invoking the slaving principle, we have to study an order parameter equation for
u alone. In a first preparatory step we assume that this order parameter equation
has the form
653
252 8. Nonlinear Equations. Qualitative Macroscopic Changes
(8.8.41)
(8.8.42)
where the dots indicate the amplitudes of the slaved variables which are smaller
than those which are explicitly exhibited. Here V2 and its conjugate complex are
eigensolutions to (8.7.7) with (8.7.12). Inserting (8.8.41) into (8.8.43) we find
The real part of V2 and the imaginary part of V2 are linearly independent and
present two vectors which move along the original limit cycle. The solution
(8.8.44) indicates a rotating motion in the frame of these two basic vectors Re{v2}
and Im{v2}' i. e., the end point of the vector spirals around the limit cycle while
the origin of the coordinate system moves along qo. If the two frequencies W2 and
WI == W = (P == (PI are not commensurable the trajectory spans a whole torus. Thus
in our present case we have a simple example of the bifurcation of a limit cycle to
a two-dimensional torus. In most practical applications the order parameter
equation for u will not have the simple form (8.8.40) but several complications
may occur. In particular, we may have additional terms which introduce a
dependence of the constant b on ({Jl' A simple example is the equation
U = (A~ + iW2)u - bu lu 12 - C(WI t)u lu 12, (8.8.45)
({JI
and (8.8.46)
(8.8.47)
which are of precisely the same form encountered previously in the case of Hopf
bifurcation. From this analysis we know that basically the bifurcation of the limit
cycle of a torus is still present.
654
8.9 Bifurcation from a Torus (Quasiperiodic Motion) 253
Things become considerably more complicated if the equation for the order
parameter u is of the form
(8.8.48)
and that for the phase angle ({J :; ({Jt of the form
(8.8.49)
(8.8.50)
(8.8.51)
(8.8.52)
(8.8.53)
which stems from (8.8.49) and where Jt (like fJ and !2) is 2 rr-periodic in ({Jt, ({J2·
The discussion of the solutions of these equations is somewhat subtle, at least in
the general case, because a perturbation expansion with respect to fJ and Jt , J2
requires the validity of a KAM condition on Wt, W2. Since this whole discussion
will turn out to be a special case of the discussion on bifurcations from a torus to
other tori, we shall postpone it to the end of Sect. 8.10.2.
with the by now familiar notation. We assume that (8.9.1), for a certain range of
control parameters a, allows quasiperiodic solutions qo which we may write in the
form
(8.9.3)
655
254 8. Nonlinear Equations. Qualitative Macroscopic Changes
and where the m's run over all integers. As discussed in Sects. 1.12 and 8.8, such
a motion can be visualized as motion on a torus. We assume in the following that
the trajectories lie dense on that torus. We may parametrize such a torus by phase
angles lfJj , j = 1, ... , M, so that by use of these variables the vector r whose end
point lies on the torus can be written in the form
(8.9.4)
Because the trajectories lie dense on the torus 1 we may subject them to the initial
condition
(8.9.5)
This allows us to construct the solutions (8.9.2) which fulfill the initial condition
(8.9.5) explicitly by putting
(8.9.7)
(8.9.8)
The first steps of our subsequent procedure are now well known from the preced-
ing chapters. We assume that the form of the solution (8.9.6) still holds when we
increase the control parameter a further, in particular into a region where the old
torus loses its linear stability.
To study the stability we perform the stability analysis, considering the equa-
tion
(8.9.10)
1 More precisely speaking, we assume that with (8.9.2) as solution of (8.9.1) also
656
8.9 Bifurcation from a Torus (Quasiperiodic Motion) 255
is defined by
(8.9.13)
This form holds especially if the characteristic exponents (or generalized charac-
teristic exponents) are all different from each other, if L is sufficiently often dif-
ferentiable with respect to CPj' j = 1, ... ,M, and if a suitable KAM condition is
fulfilled by the w's. In this case, the Vk'S are quasiperiodic functions of time,
whose dependence on cP or tP, respectively, is of the general form of rhs of
(8.9.6). Here we need a version of that theorem in which all Ak are different from
each other except those for which Ak = 0, whose Vk will be constructed now. To
this end we start from the equation
(8.9.14)
(8.9.15)
(8.9.16)
Since lhs of (8.9.16) is quasiperiodic, we have thus found solutions in the form
(8.9.13) with AI = O. We shall denote all other solutions to (8.9.9) by
We now wish to find a construction for the solutions which bifurcate from
the old torus which becomes unstable. We do this by generalizing our considera-
657
256 8. Nonlinear Equations. Qualitative Macroscopic Changes
tions on the bifurcation of a limit cycle. We assume self-consistently that the new
torus or the new tori are not far away from the old torus, so that we may utilize
small vectors pointing from the old torus to the new torus. On the other hand, we
have to take into account the fact that when bifurcation takes place, the corre-
sponding points on the old trajectory and the new one can be separated arbi-
trarily far from each other when time elapses.
To take into account these two features we introduce the following coor-
dinate system. We utilize the phase angles as local coordinates on the old torus,
and introduce vectors v which point from each local point on the old torus to the
new point at the bifurcating new torus. The local coordinates which are trans-
versal to the old torus are provided by (8.9.17). These arguments lead us to the
following hypothesis for the bifurcating solutions (the prime on the sum means
exclusion of k = 1, 2, ... , M)
where we have expanded the rhs into powers of W. In this way, H is given by an
expression of the form
where we have indicated higher-order terms by dots. It is, of course, a simple task
to write down such higher-order terms explicitly in the way indicated by the
second-order term.
We now make use of results of Sect. 2.5, where a set of vectors v was con-
structed orthogonal on the solutions v of (8.9.13). Multiplying (8.9.21) by
we readily obtain
658
8.9 Bifurcation from a Torus (Quasiperiodic Motion) 257
. .
CPI + L L ~j(VI'
I SVjISCPI') CPI' = (vIH) , (8.9.24)
j I' '----v-----' '-v--'
ajll' iii
while by multiplication of (8.9.21) by
Vb k=M+1,oo., (8.9.25)
we obtain
. .
~k = Ak~k - L L ~j(Vk' Sv/SCP!') CPI' + (vkH)
I . (8.9.26)
j I' ~ '-v--'
ajkl' lik
Introducing the matrix K with matrix elements
and taking cP and Ii as column vectors, we may cast the set of equations (8.9.24)
into the form
(8.9.28)
(8.9.29)
where the last term of this equation is just an abbreviation. Expressing cP!' which
occurs in (8.9.26) by rhs of (8.9.29), we transform (8.9.26) into
(8.9.30)
We note that t and CP, on which the right-hand sides of (8.9.29,30) depend, occur
only in the combination
(8.9.31)
Therefore we introduce {fJj as a new variable which allows us to cast (8.9.29) into
the form
(8.9.32)
(8.9.33)
659
258 8. Nonlinear Equations. Qualitative Macroscopic Changes
1(8.9.34)
(8.9.35)
and (S.1O.1)
and put
(8.10.3)
Here u plays the role of the order parameter, while ¢b k ;;, M + 2 plays the role
of the slaved variables. Applying the slaving principle we reduce the set of equa-
tions (S.9.34 and 32) to equations of the form
(S.1O.4)
(S.10.6)
660
8.10 Bifurcation from a Torus: Special Cases 259
(8.10.9)
2][ 2][
S···Sdrpl.··drp Mh2=0. (8.10.10)
a a
In addition we assume
and (8.10.15)
u = e(ua + rO . (8.10.16)
We determine ua by
(8.10.17)
so that
Ua = ± VA-alb (8.10.18)
holds. This relation indicates that two new tori split off the old torus and keep a
mean distance ua from the old torus. Introducing in addition to (8.10.15, 16) the
new scaling of time
(8.10.19)
661
260 8. Nonlinear Equations. Qualitative Macroscopic Changes
where (8.10.23)
(8.10.24)
(8.10.25)
is of order unity, but due to its dependence on cp acts like a perturbation 0:: e.
Superficially, (8.10.20, 23) are familiar from Sects. 8.7, 8.8. The reader
should be warned, however, not to apply the previous discussion directly to the
case of (8.10.20, 23) without further reflection. The basic difference between
(8.10.20, 23) and the equations of Sects. 8.7, 8.8 is that here we are dealing with
quasiperiodic motion, while Sects. 8.7, 8.8. dealt with periodic motion only. It is
appropriate here to recall the intricacies of Moser's theorem. Before we discuss
this question any further we shall treat the case in which AM+l is complex. We
shall show that in such a case we are led to equations which again have the struc-
ture of equations (8.10.20, 23), so that we can discuss the solutions of these equa-
tions simultaneously.
and (8.10.26)
';M+I = U (8.10.27)
and assume
662
8.10 Bifurcation from a Torus: Special Cases 261
Application of the slaving principle allows the reduction of the original set of
equations (8.9.32, 34) to the two equations
Introducing the new vector fp which contains the additional phase qJM+I'
qJI ]
fp= [ : (8.10.34)
;M+I .
[ 1= ~ land (8.10.38)
fM+l
663
262 8. Nonlinear Equations. Qualitative Macroscopic Changes
From a formal point of view (8.10.35, 37) have the same structure as (8.9.32),
(8.9.34) which then can be transformed into (8.10.20,23).
In precisely the same way as we discussed in Sect. 8.10.1, (8.10.35, 37) allow
for the introduction of scaled variables. It is a simple matter to formulate condi-
J
tions on g and [and thus on m(u, qJ)] which are analogs to the conditions
(8.10.7- 14). This allows us to exhibit explicitly the smallness of the perturbative
terms by which we may write (8.10.35, 37) in the form
(8.10.40)
(8.10.41)
(8.10.42)
and similarly
(8.10.43)
In this way, (8.10.40, 41) are cast into the form of the equations treated by
Moser's theorem. Let us postpone the discussion whether the conditions of this
theorem can be met, and assume that these conditions are fulfilled. Then we can
apply the procedure we described after (8.5.36,37) in Sect. 8.5. In particular we
can, at least in principle, calculate D and A so that
and (8.10.44)
(8.10.45)
In addition, we can construct the steady state and transient solutions of (8.10.40,
41) in complete analogy to the procedure following (8.5.45). Thus the transient
solutions (close to the bifurcating torus) acquire the form
qJ = wrt + cu(wrt, c), (8.10.46)
11 = Xo exp (Art) + Cv(wrt, c) + c V(wrt, c)xoexp(Art) , Ar < o. (8.10.47)
u, v, and V are 2 n-periodic in each component of wrt. The steady state solutions
are given by
qJ = wrt + cu(wrt, c) , (8.10.48)
11 = cv(wrt, c). (8.1 0.49)
664
8.10 Bifurcation from a Torus: Special Cases 263
The rest of this section will be devoted to a discussion under which circumstances
the conditions of Moser's theorem can be fulfilled, in particular under which
conditions Wr fulfills the KAM condition 1. We first turn to the question under
which hypothesis we derive (8.10.29, 30). If we use these equations or (8.10.40,
41) in the form of a model, no further specifications on Wu must be made. In this
case we find the bifurcation from a torus provided we can find for given Au and
Wu such Ar and Wr so that (8.10.44,45) are fulfilled and Wr obeys a KAM condi-
tion. If, on the other hand, (8.10.40, 41) have been derived starting from the
original autonomous equations (8.9.1) then we must take into account the hypo-
thesis we made when going from (8.9.1) to these later equations. The main hypo-
thesis involved concerned the structure ofthe functions (8.9.13), to secure that Vk
are quasiperiodic. This implied in particular that the frequencies WI, ••• , £OM
obeyed the KAM condition.
We now face a very profound problem, namely whether it is possible to check
whether or not such frequencies as wr and Wu obey a KAM condition. Maybe
with the exception of a few special cases this problem implies that we know each
number £OJ with absolute precision. This is, however, a task impossible to solve.
Further discussion depends on our attitude, whether we take the point of view of
a mathematician, or a physicist or chemist.
A mathematician would proceed as follows. In order to check whether a bi-
furcation from one torus to another can be expected to occur "in reality", we
investigate how probable it is that out of a given set of w's a specific set fulfills
the KAM condition. As is shown in mathematics, there is a high probability for
the fulfillment of such a condition provided the constant K occurring in (6.2.6) is
small enough. It should be noted that this constant K occurs jointly with the scal-
ing factor e2, as can be seen by the following argument.
We start from the KAM condition (6.2.6). The scaling of time transforms the
frequencies wrand A,u into w r le 2, A,u/e 2, respectively. This transforms the KAM
condition into
(8.10.50)
From the mathematical point of view we thus are led to the conclusion that the
probability of finding suitable combinations of w's is very high, or more precise-
ly expressed, the bifurcation from one torus to another takes place with non-
vanishing measure.
1 From a rigorous mathematical point of view we require that g andh in (8.10.40,41) are analytic in
TJ and 'fl. It should be noted that the slaving principle introduced in Chap. 7 does not guarantee this
property even if the initial equations (i. e., those before applying the slaving principle) only con-
tained right-hand sides analytic in the variables. Thus we have either to invoke "smoothing" as ex-
plained in Sect. 7.5, or to introduce the equations (8.10.40, 41) by way of a model (which appeals
to a physicist or engineer), or by making stronger assumptions on the original equations so that the
slaving principle can be given a correspondingly sharper form (which appeals to a mathematician).
A further possibility is provided by weakening the assumptions of Moser's theorem (which is also
possible in certain cases, e. g., if certain symmetries are fulfilled).
665
264 8. Nonlinear Equations. Qualitative Macroscopic Changes
From a physical point of view we can hardly think that nature discriminates
in a precise way between w's fulfilling a KAM condition and others. Rather, at a
microscopic level all the time fluctuations take place which will have an impact
on the actual frequencies. Therefore, the discussion on the bifurcation of a torus
will make sense only if we take fluctuations into account. We shall indicate a
possible treatment of such a problem in a subsequent chapter. Here we just want
to mention that fluctuations, at least in general, will let the phase angles diffuse
over the whole torus so that, at least in general, such delicate questions as to
whether a KAM condition is fulfilled or not will not playa role. Rather, a system
averages over different frequencies in one way or another.
Some other important cases must not be overlooked, however. One of them is
frequency locking which has been dicussed in Sect. 8.6. In this case a torus col-
lapses into a limit cycle. It is worth mentioning that fluctuations may play an im-
portant role in this case, too. Finally it may be noted that by a combination of the
methods developed in Sects. 8.5 -10, further phenomena, e.g. period doubling
of the motion on a torus, can easily be treated.
666
8.11 Instability Hierarchies, Scenarios, and Routes to Turbulence 265
2 This nearly trivial example illustrates another important aspect: sometimes one has to be cautious
when applying the concept of "genericity" to physics (or other fields). Here, due to symmetry, con-
servation laws, or for other reasons, the phenomena may correspond to "nongeneric" solutions
(such as Coulomb's law).
667
266 8. Nonlinear Equations. Qualitative Macroscopic Changes
sional torus is possible. We drew the conclusion that probability arguments must
be applied in order to decide whether a real system will show such a transition.
Our approach solves the puzzle of why systems may show this kind of bifurcation
despite the fact that the corresponding solutions are not "generic" in the sense of
Ruelle and Takens. Indeed, for a given system there may be regions of the con-
trol parameter or of other parameters in which a scenario of subsequent bifurca-
tions of tori is possible. However, when proceeding to higher-dimensional tori,
such transitions become more and more improbable, so that the Landau-Hopf
picture loses its validity and chaos sets in.
There are at least two further scenarios for the routes to turbulence; these are
discussed in the next sections.
668
9. Spatial Patterns
We denote the space vector (x, y, z) by x. The state of the total system is then de-
scribed by a state vector
q(x, t) (9.1.1)
These equations depend in a nonlinear fashion on q(x, t). They contain, at least
in general, spatial derivatives and they describe the impact of the surrounding by
268 9. Spatial Patterns
(9.1.5)
(9.1.6)
It takes into account that chemicals may diffuse with different diffusion con-
stants.
Another broad class of nonlinear equations of the form (9.1.3) occurs in hy-
drodynamics. The most frequently occurring nonlinearity stems from the stream-
ing term. This term arises when the streaming of a fluid is described in the usual
way by local variables. This description is obtained by a transformation of the
particle velocity from the coordinates of the streaming particles to the local coor-
dinates of a liquid. Denoting the coordinate of a single particle by x (t) and
putting
v=x (9.1.7)
dv(x(t),/) av av av av
-----= -Vx + - vy + - vz + - . (9.1.8)
dl aX ay az al
670
9.1 The Basic Differential Equations 269
At rhs of (9.1.8) we must then interpret the arguments x as that of a fixed space
point which is no longer moving along with the fluid. The right-hand side of
(9.1.8) represents the well-known streaming term, which is nonlinear in v. Of
course, other nonlinearities may also occur, for instance, when the density,
which occurs in the equations of fluid dynamics, becomes temperature dependent
and temperature itself is considered as part of the state vector q. Note that in the
equations of fluid dynamics v is part of the state vector q
q,v-+q(x,t) . (9.1.9)
q(s) = 0, (9.1.10)
(s: surface), for a generalization see below. 2) The derivative normal to the
surface must vanish. This condition is sometimes called the non flux boundary
condition
aqj(s) = 0 . (9.1.11)
an
3) Another boundary condition, which is somewhat artificial but is quite useful if
no other boundary conditions are given, is the periodic boundary condition
q (x) = q (x + L) . (9.1.12)
Here it is required that the solution is periodic along its coordinates x, y, z with
periods Lx, L y , L z' respectively. 4) It there is no boundary condition within finite
dimensions, in general one requires that
671
270 9. Spatial Patterns
where ij is a prescribed function at the surface. For instance one may require that
the concentrations of certain chemicals are kept constant at the boundary.
There are still other boundary conditions which are less obvious, namely
when q (x) is a function on a manifold, e. g., on a sphere or on a torus. Such ex-
amples arise in biology (evolution of morula, blastula, gastrula, and presumably
many other cases of biological pattern formation), as well as in astrophysics. In
such a case we require that q(x) is uniquely defined on the manifold, i. e., if we
go along a meridian we must find the same values of q after one closed path.
We assume that the solution qo can be extended into a new region of control
parameter values a but where linear stability is lost. To study the stability of the
solution of (9.1.3) we put
The right-hand side of (9.2.3) is sometimes called the linearization of (9.1.3) and
L is sometimes called the Frechet derivative. If N is a functional of qj (e. g., in the
form of integrals over functions of qj at different space points) the components
of the matrix L are given by functional derivatives
(9.2.4)
However, in order no to overload our presentation we shall not enter the problem
of how to define (9.2.3) in abstract mathematical terms but rather illustrate the
derivation of (9.2.3) by explicit examples. In the case of a reaction diffusion
equation (9.1.4), we obtain
where (9.2.5)
672
9.2 The General Method of Solution 271
We give here only one additional example of how to obtain the linearization
(9.2.3). To this end we consider as a specific term the streaming term which may
occur inN,
(9.2.8)
(}
[qo J'(x) + wJ·(x)] - [qo ,(x) + w,(x)] . (9.2.9)
• (}Xk •
Multiplying all terms with each other and keeping only the terms which are linear
in Wj we readily obtain
(9.2.10)
(9.2.11)
(9.2.12)
(9.2.13)
(9.2.15)
This equation represents a set of coupled linear ordinary differential equations
with constant coefficients. From Sect. 2.6 we know the general form of its
solution,
(9.2.16)
673
272 9. Spatial Patterns
(9.2.17)
(9.2.18)
(9.2.20)
where k is a real vector and z is periodic with periods (9.2.17). In case of other
boundary conditions in finite geometries, standing waves may be formed from
(9.2.20) in order to fulfill the adequate boundary conditions. Here it is assumed
that the boundary conditions are consistent with the requirement that L (x) is
periodic with (9.2.17).
We now wish to show how to solve equations of the type (9.1.3). We study the
bifurcation from a node or focus which are stable for a certain range of para-
674
9.3 Bifurcation Analysis for Finite Geometries 273
meter values a and which become unstable beyond a critical value a c ' We assume
that the corresponding old solution is given by
(9.3.1)
We assume that the solutions of the linearized equations (9.2.3) have the form
(9.3.2)
The essential difference between our present procedure and that of the Chap.
8 (especially Sects. 8.1-5) consists in the fact that the index k covers an infinite
set while in former chapters k was a finite set. The other difference consists in the
space dependence of q, qo and v. We insert (9.3.3) into (9.1.3) where we expand
the rhs of N into a power series of W. Under the assumption that (9.3.3) obeys
(9.1.3) we obtain
(9.3.5)
We now multiply (9.3.4) by Vk(X) and integrate over the space within the given
boundaries. According to the definition of Vk we obtain
(9.3.6)
Using in addition
(9.3.7)
with (9.3.8)
A~22'k"
675
274 9. Spatial Patterns
(9.3.11)
The spectrum of the operator L will be, at least in general, continuous if there are
no boundary conditions in infinitely extended media. Let us consider the special
case in which qo is space and time independent. We treat a linear operator L of
the form (9.2.11), where L is space and time independent. According to (9.2.13)
we may choose X in the form of plane waves
(9.4.1)
Using hypothesis (9.2.12), we have to deal again with a coupled set of linear
differential equations with constant coefficients (9.2.15). If WI is a finite dimen-
sional vector, the eigensolutions WI can be characterized by a discrete set of
indices which we call}. On the other hand, k is a continuous variable. Neglecting
676
9.4 Generalized Ginzburg-Landau Equations 275
(9.4.2)
(9.4.3)
Such a continuous spectrum can cause some difficulties when applying the
slaving principle. For this reason we resort to an approach well known to quan-
tum mechanics and which amounts to the formation of wave packets. To this end
we split k into a discrete set of vectors k' and remaining continuous parts k
k=k'+k. (9.4.5)
(9.4.6)
L J J J d3k~k'+k,j(t)eik'Xvk'+k,j(0)eik"X. (9.4.7)
k'k;-o/2 k;-O/2 k~-0!2
'~----------~v~----------~
~k',/X,t)
(9.4.8)
(9.4.9)
677
276 9. Spatial Patterns
(9.4.11)
we shall integrate over a space region which contains many oscillations of vex)
but over which ~k does not change appreciably. Under these assumptions the rela-
tions
and (9.4.12)
(9.4.13)
(9.4.14)
(9.4.15)
where AiV') yields A/ik) if A)V') is followed by exp(ik • x). The integral follow-
ing Aj in (9.4.15) can be split into the sum (9.4.9) as before, so that we obtain
(9.4.16)
~k ',j(x, t)
(9.4.17)
(9.4.18)
where formulas well known in quantum physics have been used. It follows from
(9.4.7) that ~k',j contains only small k-vectors, enabling us to expand Aj into a
power series with respect to V' . We readily obtain
Alik' +V')~kjX,t)
678
9.4 Generalized Ginzburg-Landau Equations 277
(9.4.20)
(9.4.21)
where we have made use of the result (9.4.13). After all these steps we are now in
a position to treat the nonlinear equation (9.1.3). To this end we insert hypothesis
(9.4.10) into (9.1.3), expand rhs of (9.1.3) into a power series of Nwith respect to
q, and make use of the linearized equation (9.2.3). We then readily obtain
with (9.4.22)
Except for the fact that Aj is an operator, (9.4.22) is of the same form as all the
equations we have studied in a similar context before.
To make use of the slaving principle we must distinguish between the unstable
and stable modes. To this end we distinguish between such A'S for which
hold. We must be aware of the fact that the indices j and k are not independent of
each other when we define the unstable and stable modes according to (9.4.24,
25). We shall denote those modes for which (9.4.24) holds by Ukj, whereas those
modes for which (9.4.25) is valid by Skj. Note that in the following summationsj
and k run over such restricted sets of values implicitly defined by (9.4.24, 25).
The occurrence of the continuous spectrum provokes a difficulty, because the
modes go over continuously from the slaved modes to the undamped or unslaved
modes. Because the slaving principle requires that the real part of the spectrum As
has an upper bound [cf. (9.4.25)], we make a cut between the two regions at such
C. Consequently, we also have to treat modes with a certain range of negative
real parts of Aj as unstable modes. As one may convince oneself the slaving prin-
ciple is still valid under the present conditions provided the amplitudes of Ukj are
small enough.
Since the A'S are operators, the corresponding Au and As are also operators
with respect to spatial coordinates. Some analysis shows that the slaving principle
also holds in this more general case provided the amplitudes C; are only slowly
varying functions of space so that As(k + V ) does not deviate appreciably from
AAk). After these comments it is easy to apply to formalisms of the slaving prin-
ciple. Keeping in the original equation (9.1.3) only terms of Nup to third order in
679
278 9. Spatial Patterns
Precisely speaking, the coefficients A and B may contain derivatives with respect
to spatial coordinates. However, in most cases of practical importance,
AAik + V) which occurs in the denominators may be well approximated by
As(ik) because U depends only weakly on x and the denominators are bounded
from below. If the original equation (9.1.3) contained fluctuations, the corre-
sponding fluctuating forces reappear in (9.4.26) and are added here in the form
of Fk,j' I have called these equations, which I derived some time ago, "gener-
alized Ginzburg-Landau equations" because if we use Aj in the approximation
(9.4.19) and drop the indices and sums over k and}, the equations (9.4.26) reduce
to equations originally established by Ginzburg and Landau in the case of equi-
librium phase transitions, especially in the case of superconductivity. Besides the
fact that the present equations are much more general, two points should be
stressed. The equations are derived here from first principles and they apply in
particular to systems far from thermal equilibrium.
(9.5.1)
To make contact with the A which occurs in the order parameter equations
(9.4.26), we must transform (9.5.1) by means of
(9.5.2)
680
9.5 A Simplification of Generalized Ginzburg-Landau Equations 279
(9.5.3)
In the following we shall use the order parameter equations (9.4.26) in the
form
(9.5.4)
where the Kronecker symbols under the sums ensure the conservation of wave
numbers. This is the case if we do not include boundary conditions with respect
to the horizontal plane, and use plane-wave solutions for Vk,j(X), It is evident that
the specific form of A(9.5.1) or A(9.5.3) favors those Ik .L Iwhich are close to Ikc I.
We now introduce a new function If! according to
(9.5.5)
The sum runs over the critical k vectors, which all have the same absolute value
ko but point in different horizontal directions. We now make our basic
assumption, namely that we may put
and (9.5.6)
(9.5.7)
for the k vectors, which are selected through the a-functions and the condition
that Ike 1= k o·
We mUltiply (9.5.4) by exp(ik e • x) and sum over k c . To elucidate the further
procedure we start with the thus resulting expression which stems from the last
term in (9.5.4), and therefore consider
(9.5.8)
Since
(9.5.9)
we can insert the factor
681
280 9. Spatial Patterns
L exp[i{k~ + k~' + k~"}X] Uk ' (X)Uk ,, (X)Uk ''' (X) L (\ k '+ k " + k '" (9.5.11)
k~,k~',k~" C C kc c' C c c
L 15k
k C'
kC'+ kC"+ k C'" = 1. (9.5.12)
C
(01
16
"
~j
.~ ~
,~
'~
~
~~
~:.
. ~.
.'
~
~
~
>-=..
.
~
'. ... . -
.
t'
~ ~
T 04:100; F- - 6 .22 ; N· .048
16
T ':1000 ; F· -2.06; N.ma T.25000;F' -2 .21 ; N" .0 40
o -.
o 16 16 To12600 ; Fo-6.38 ; N- .0 48
To1900 ; F· -280. ;N'0.!IO To 6400 ; F" -266 ; N a O'48
Fig. 9.la - f. Contour plots of the amplitude field I/f(x,y , r) at variou s times 1 '= r, for (a, b, e, f)
a = 0.10 and (e, d) a = 0.90, where A = 0, B = 1. The cells in (a - d) have an aspect ratio ofl6 while
the cells in (e, f) have aspect ratios of 29.2 and 19.5. The initial conditions were parallel rolls for (a , e)
and random domains of opposite sign for (b , e). The solid and dotted contours represen t positive a nd
negative values at t, t, t ,
and t\; the maximum amplitude. The contours correspond to vertical
velocity contours in optical experiments. The values of the time at which equilibrium was reached (r),
the Lyapunov functional (F) , and the Nu sselt number (N) are given. The state in (e) has not reached
equilibrium and is evolving into the equilibrium state, (f). Here equilibrium is defined to occur when
d In(F)l dr is smaller than 10 - 8 . [After H. S. Green side, W. M. Coughran, Jr., N. L. Schryer: Ph ys.
Rev. Lett. 49, 726 (1982)]
682
9.S A Simplification of Generalized Ginzburg-Landau Equations 281
Because the whole procedure yields the expression fiJ(x) for the lhs of (9.5.4), we
find the new equation
(9.5.15)
683
10. The Inclusion of Noise
In the introduction we pointed out that noise plays a crucial role especially at in-
stability points. Here we shall outline how to incorporate noise into the approach
developed in the previous chapters. In synergetics we usually start from equa-
tions at a mezoscopic level, disregarding the microscopic motion, for instance of
molecules or atoms. The equations of fluid dynamics may stand as an example
for many others. Here we are dealing with certain macroscopic quantities such as
densities, macroscopic velocities, etc. Similarly, in biological morphogenesis we
disregard individual processes below the cell level, for instance metabolic proces-
ses. On the other hand, these microscopic processes cannot be completely
neglected as they give rise to fluctuating driving forces in the equations for the
state variables q of the system under consideration. We shall not derive these
noise sources. This has to be done in the individual cases depending on the nature
of noise, whether it is of quantum mechanical origin, or due to thermal fluctua-
tions, or whether it is external noise, produced by the action of reservoirs to
which a system is coupled. Here we wish rather to outline the general approach to
deal with given noise sources. We shall elucidate our approach by explicit
examples.
Adding the appropriate fluctuating forces to the original equations we find equa-
tions of the form
which we shall call Langevin equations. If the fluctuating forces depend on the
state variables we have to use stochastic differential equations of the form
which may be treated according to the Ito or Stratonovich calculus (Chap. 4). In
the present context we wish to study the impact of fluctuating forces on the
behavior of systems close to instability points. We shall assume that F is com-
paratively small in a way that it does not change the character of the transition
appreciably. This means that we shall concentrate on those problems in which the
10.2 A Simple Example 283
q=N(q,a) (10.1.3)
possesses a solution for a given range of the control parameter a, and study as
before the stability of
(10.1.4)
The hypothesis
(10.1.5)
(10.1.6)
(10.1.7)
Again amplitudes f.k and phase angles ¢k may be introduced. Assuming that one
or several eigenvalues A's belonging to (10.1.7) acquire a positive real part, we
shall apply the slaving principle. In Sect. 7.6 we have shown that the slaving prin-
ciple can be applied to stochastic differential equations of the Langevin-Ito or
-Stratonovich type. According to the slaving principle we may reduce the original
set of equations (10.1.2) to a set of order parameter equations for the corre-
sponding f.k and ¢k' The resulting equations are again precisely of the form
(10.1.2), though, of course, the explicit form of Nand dF has changed. To illus-
trate the impact of fluctuations and how to deal with it, we shall consider a
specific example first.
(10.2.1)
(F(t» = 0, (10.2.2)
685
284 10. The Inclusion of Noise
for the fluctuating forces. If we are away from the instability point Jt = 0, in
general it suffices to solve (10.2.1) approximately by linearization. Because for
A < 0, u is a small quantity close to the stationary state, we may approximate
(10.2.1) by
u==Au+F{t). (10.2.4)
(10.2.6)
where we have neglected higher powers of". Because (10.2.6) is of the same form
as (10.2.4) and" is small, it suffices to study the solutions of (10.2.6) which can
be written in the form
,,= oSexp[-2A(t-r)]F(r)dr.
t
(10.2.7)
(We do not take care of the homogeneous solution because it drops out in the
limit performed below.) The correlation function can be easily evaluated by
means of (10.2.3) and yields in a straightforward fashion for t ~ t'
ft'
(1'/(t),,(t'» = SSexp[ -2A(t- r) - 2A(t'- r')] Qc5(r- r')drdr'
00
t'
= Jexp[ - 2A(t+ t') + 4h'] Qdr'
o
(e 4At ' 1)
=exp[-2A{t+t')]Q . (10.2.8)
4A
Assuming that t and t' are big, but t- t' remains finite, (10.2.8) reduces to the
stationary correlation function
686
10.2 A Simple Example 285
. 8 3 Q 82
f= - -[(AU - bu )f] +- --2f, where (10.2.10)
8u 2 8u
(10.2.12)
where JtI is a normalization factor. The branching of the solution of the deter-
V
ministic equation with F(t) == 0 from u = 0 for A ~ 0 into u ± = ± Alb for A > 0
is now replaced by the change of shape of the distribution function (10.2.12), in
which the single peak for A < 0 is replaced by two peaks for A> O. Some care
must be exercised in interpreting this result, because fo is a probability distri-
bution. Therefore, in reality the system may be at any point u but with given
probability (10.2.12). Clearly, for A> 0 the probability is a maximum for
u ± = ± VAlb. But, of course, at a given moment the system can be in a single
state only. From this point of view an important question arises. Let us assume
that at time t = 0 we have prepared (or measured) the system in a certain initial
state u = Uj. What is the probability of finding the system at a later time I in some
other final given state U = Uf? This question can be answered by the time-depend-
ent solution f(u, I) of the Fokker-Planck equation with the initial condition
f(u,O) = o(u - Uj). Even for Fokker-Planck equations containing several varia-
bles these solutions can be found explicitly if the drift coefficients are linear in
the variables and the diffusion coefficients constants. We shall present the results
for the one-variable Fokker-Planck equation at the end of this section and the
corresponding general theorem in Sect. 10.4.1. If, for example, the drift coef-
ficients are nonlinear, even in the case of a single variable, computer solutions
are necessary. We shall give an outline of the corresponding results in Sect. 10.3.
A further important problem is the following. Let us assume that the system
was originally prepared in the state U +. How long will it take for the system to
reach the state U _ for the first time? This is a special case of the so-called first
passage time problem which can be formally solved by a general formula. We
shall deal with the first passage time problem in the context of the more general
Chapman-Kolmogorov equation for discrete maps in Sect. 11.6. Now, let us take
up the problem of finding the time-dependent solutions of (10.2.1) in an approxi-
687
286 10. The Inclusion of Noise
u = Uo + 1] (10.2.13)
.:. 8 - Q 82 -
/= --(-2A"f) +- - - I (10.2.14)
81] 2 8,,2
(10.2.16)
(10.3.2)
688
10.3 Computer Solution of a Fokker-Planck Equation for a Complex Order Parameter 287
oWl 0 -
--+f3--[(n-r)rw]=Q
ot r or
2 [1
--0 ( r0-W)
r or or
2
- +1- 0- W
r2 orp2
-. l
(10.3.4)
[This can best be done by writing down the Fokker-Planck equation for the real
and imaginary parts of u and transforming that equation to polar coordinates
according to (10.3.3).]
In order to get rid of superfluous constants we introduce new variables or
constants by means of
(10.3.5)
(10.3.6)
;/I
Wei') = -<-exp
(~4 ~2
-!....-.. + a!....-.. ) 1
, - = Ii'exp
00 (~4 -2 )
- !....-..+a!....-.. di'.
2n 4 2 .;/10 42
(10.3.7)
In order to obtain correlation functions, e. g., of the type (10.2.8), the nonsta-
tionary solutions of the Fokker-Planck equation must be used. Since general ana-
lytical expressions are not known, either approximation methods (like a varia-
tional method) or painstaking computer calculations were performed, the latter
yielding results of great accuracy, of which those relevant are represented.
First (10.3.6) is reduced to a one-dimensional Schrodinger equation. The
hypothesis
~
W(f,rp,t)= ~
00
m-O n-
_L 00
-00
Anm [ 1
l;;exp
Vr
(i'4
8 4
l
- + ai'2- ) 'Pnm(i') exp(inrp-Anmt}~
(10.3.8)
leads to
with (10.3.9)
for the eigenfunctions 'Pnm = 'P -nm and the eigenvalues Anm = A -nm. For explicit
values of the A'S see Table 10.3.1.
689
288 10. The Inclusion of Noise
n2 'l'."
V(f)=-+~
n A2
r
'l'.
00
= (n2 _ ~4 ) _1 + a +
f2
(
-2
a 2 A2 a A4
- ) r--r+-
rA6
. (10.3.10)
4 2 4
(10.3.11)
The first five eigenfunctions, which are determined numerically, are plotted in
Fig. 10.3.2. For a plot of eigenvalues consult Fig. 10.3.3.
If follows from (10.3.9) that the Pnm's are orthogonal for different m. If they
are normalized to unity, we have
J Pnm(f) P nm ,(f)df =
00
f5 mm , . (10.3.12)
o
a=-4 a=10
VoIr) IV' (r)
80
I
60
I
\
40
\
\
20 \.- Fig. 10.3.1. The potential
Vo(f) of the Schrodinger
0 ~~S;;::cz:===:::::",--=::;==::::;::.L~a-7L--::::- equation (10.3.9) for three
l' pump parameters (solid line)
and V 1 (f) for a = 10 (broken
-20 line). [After H. Risken, H. D.
Vollmer: Z. Physik 201, 323
-40 (1967)]
690
10.3 Computer Solution of a Fokker-Planck Equation for a Complex Order Parameter 289
I
80 of the Schrodinger equation
a =10 (10.3.9) and the first five
60 eigenvalues a nd eigenfunc-
--
40 "'04- a = 10. [After H. Risken, H .
D. Vollmer: Z . Phys. 201 , 323
20
V -X ~ V~~ .~~(1967)]
o / ~
....x--
~
"--./
J.-<"F01' '"01
~'I'~ "'nn
-20
-40
o
Anm=
1
I~ exp (f,4 f'2)
---a- - 'Pnm(f')exp(-incp'). (10.3.14)
2nv f' 8 4
1 (f4 f2 f ,4 f' 2 )
G(f,cp;f',cp', f)= exp - -+a-+---a--
2n~ 8 4 8 4
00 00
2S
/
'"
20
V
~ ~2 I-----
~
kI
15
10
5
~
"-..,
1"--.. "" >:;'f~ V
~
691
290 10. The Inclusion of Noise
For the calculation of stationary two-time correlation functions the joint distri-
bution function F(f,({J;f',q/; f) is needed. F(f,({J;f',({J';f)fdfd({Jf'df'd({J' is
the probability that f(t + f), (P(t + f) lie in the interval f, ... , f + df; ({J, ... , ({J + d ({J
and that f'(t), ({J'(t) lie in the interval f', ... ,f'+df', ({J', ... ,({J'+d({J'.
Moreover, F can be expressed by means of the Green's function G(f, ({J; f', ({J'; f)
and of W(f', ({J'), the latter describing the distribution at the initial time, t,
F(f, ({J; f', ({J'; f) = G(f, ({J; f', ({J'; f) WeT', ((J'), f > 0. (10.3.16)
The correlation function of the intensity fluctuations is obtained by
(10.3.18)
For a plot of the first four matrix elements Mm see Fig. 10.3.4, and for explicit
values, Table 10.3.1. Here K(a,O) is given by
1.0
- ~ Ml
"\
0.8 ~--
"'-
0.6
0.1.
My V Fig. 10.3.4. The first four matrix cle-
--
ments M m as functions of the pump
---~ ---
o. 2
parameter a. [After H. Risken, H.
o
/ D. Vollmer: Z. Phys. 201, 323
-10 -8 -6 -I. -2 0 6 8 10 (1967)]
692
10.3 Computer Solution of a Fokker-Planck Equation for a Complex Order Parameter 291
(10.3.21)
The spectrum S(a, w) of the intensity fluctuations is given by the Fourier trans-
form of the correlation function (10.3.17)
00
.h ~ Mm
WIt -1
- = iJ -- (10.3.24)
Aeff m= I AOm
which has the same area and the same maximum value (Fig. 10.3.5).
This effective width Aeff is, however, about 25070 larger than AOI for a"" 5. The
eigenvalues and matrix elements were calculated numerically in particular for the
threshold region -10 ~ a ~ 10. Similar calculations for the correlation function
of the amplitude yield
= IIf If dff' df' dqJ dqJ' ff' exp(iqJ- i qJ')F(f, qJ; f', qJ'; f),
00
693
292 10. The Inclusion of Noise
(10.3.26)
and can be reduced to the error integral by the same substitution leading to
(10.3.21). A calculation of Yo shows that
00
1- Va = L Vm
m=1
is of the order of 20/0 near threshold and smaller farther away from threshold.
Therefore the spectral profile is nearly Lorentzian with a linewidth (in unnor-
malized units)
(10.3.28)
-""
2.0
.........
"""-
Fig. 10.3.6. The linewidth factor
0.5 a(a) = A10 as a function of the pump
parameter a. [After H. Risken: Z.
o Phys. 191, 302 (1966)]
-10 -8 -6 -4 -2 a0 _ _
2 4 6810
2"
J
00
where G is the Green's function (10.3.15) and W(f',rp',O) is the initial distribu-
tion.
After this specific example we now turn to the presentation of several useful
general theorems.
694
IDA Some Useful General Theorems on the Solutions of Fokker-Planck Equations 293
G(q,q',t) = [nndet{a(t)}l-tl2
where
(10.4.6)
C = (Cij) ,
(10.4.8)
Q= (Qu),
695
294 10. The Inclusion of Noise
and the superscript T denotes the transposed matrix. In particular the stationary
solution reads
f(q) = G(q,q ',00) = [nn det {a( 00)}1-1I2
a) Detailed Balance
We denote the set of variables q1> ... , qN by q and the set of the variables under
time reversal by
(10.4.10)
(10.4.11)
(10.4.12)
696
10.4 Some Useful General Theorems on the Solutions of Fokker-Planck Equations 295
We now formulate the principle of detailed balance. The following two defini-
tions are available.
1) The principle of detailed balance (first version)
P = P(q Iq; r, l) .
I (10.4.15)
Here and in the following we assume that the Fokker-Planck equation possesses
a unique stationary solution. One may then show directly that
Taking the derivative with respect to r on both sides of (10.4.16) and putting
'*'
r = 0 (but q q'), we obtain
2) the principle of detailed balance (second version)
It has obviously a very simple meaning. The left-hand side describes the total
transition rate out of the state q into a new state q The principle of detailed
I.
balance then requires that this transition rate is equal to the rate in the reverse
direction for q and q with reverse motion, e. g., with reverse momenta.
I
697
296 10. The Inclusion of Noise
Note that, if not otherwise stated, L may also be an integral operator. The solu-
tion of (10.4.20) is subject to the initial condition
Putting (10.4.22) into (10.4.20) and taking r = 0 on both sides, (10.4.20) acquires
the form
With these substitutions and bringing rhs to lhs, (10.4.26) acquires the form
In (10.4.29), L acts in the usual sense of an operator well known from operators
in quantum mechanics, so that Lf is to be interpreted as L (t . .. ), the points in-
698
10.4 Some Useful General Theorems on the Solutions of Fokker-Planck Equations 297
dicating an arbitrary function. So far we have seen that the condition of detailed
balance has the consequence (10.4.29).
We now demonstrate that if (10.4.29) is fulfilled the system even has the
property of the first-version principle of detailed balance (which appears to be
stronger). First we note that (10.4.29) may be iterated yielding
(10.4.31)
(10.4.32)
(10.4.33)
(10.4.34)
(10.4.35)
699
298 10. The Inclusion of Noise
(10.4.37)
(10.4.38)
(10.4.39)
If the diffusion matrix Kik possesses an inverse, (10.4.38) may be solved with
respect to the gradient of cP
(10.4.40)
(10.4.41)
(10.4.42)
Thus the conditions for detailed balance to hold are given finally by (10.4.37, 41,
42). Equation (10.4.38) or equivalently (10.4.40) then allows us to determine cP
by pure quadratures, i. e., by a line integral. Thus the stationary solution of the
Fokker-Planck equation may be determined explicitly.
10.4.3 An Example
Let us consider the following Langevin equations:
(10.4.43)
(10.4.44)
(10.4.45)
700
10.4 Some Useful General Theorems on the Solutions of Fokker-Planck Equations 299
x=wp, (10.4.46)
p= -wx, (10.4.47)
where we have chosen an appropriate scaling of the momentum p and the coor-
dinate x. As is well known from mechanics, these variables transform under time
reversal as
x=x, (10.4.48)
p= -po (10.4.49)
(10.4.50)
p = q2· (10.4.51)
Retaining this identification also for a =t= 0, we are led to postulate the following
properties of Ql, Q2,
(10.4.52)
(10.4.53)
so that el = + 1, e2 = - 1.
Using definition (10.4.34) of the irreversible drift coefficient Dl we obtain
(10.4.54)
so that
(10.4.55)
(10.4.56)
so that
(10.4.57)
701
300 10. The Inclusion of Noise
= +Wq2 (10.4.58)
and similarly
= -wQI' (10.4.59)
2a
Ai=--=-Qi, i=1,2. (10.4.60)
Q
Using (10.4.55, 57 - 59), we readily verify (10.4.42). Clearly, (10.4.60) can be
derived from the potential function
(10.4.61)
81 (10.4.62)
8t
with
(10.4.63)
Then the above conditions reduce to the following ones: Cj , Cj must have the
form
CJ = 8B18u*
J
+ /(1)
J'
(10.4.64)
702
10.5 Nonlinear Stochastic Systems Close to Critical Points: A Summary 301
(10.4.65)
(10.4.66)
uP) ar,Z) )
E ( _J_+_J_ = O. (10.4.67)
j aUj aul
As a result the stationary solution of (10.4.62) reads
where (10.4.68)
703
302 10. The Inclusion of Noise
704
11. Discrete Noisy Maps
In this chapter we shall deal with discrete noisy maps, which we got to know in
the introduction. In the first sections of the present chapter we shall study in how
far we can extend previous results on differential equations to such maps. In
Sects. 7.7-7.9 we showed how the slaving principle can be extended. We are
now going to show how an analog to the Fokker-Planck equation (Sect. 10.2 - 4)
can be found. We shall then present a path-integral solution and show how the
analog of solutions of time-dependent Fokker-Planck equations can be found.
(11.1.1)
where A is independent of the variables q k and thus jointly with the random
vector '1k represents additive noise. Here M is a function of the state vector q k
and thus jointly with 11k gives rise to mUltiplicative noise. We wish to find an
equation for the probability density at "time" k + 1 which is defined by
The average goes over the random paths of the coordinates qk' up to the index k,
due to the systematic motion and the fluctuations 11k" Therefore we transform
(11.1.3) into
(11.1. 7)
In our above treatment we have implicitly assumed that either the range of the
components of q goes from - 00 to + 00 or that the spread of the distribution
function W is small compared to the intervals on which the mapping is executed.
If these assumptions do not hold, boundary effects must be taken into account.
Since the presentation is not difficult in principle but somewhat clumsy, we show
how this problem can be treated by means ,of the one-dimensional case.
We start from the equation
b
P(q, k+ 1) = Jd~ Jdl1<5(q - /(~) -11)P(~, k) W(I1) , (11.2.1)
a
where
and (11.2.2)
q =/(0 + 11 (11.2.3)
b
JP(q,k)dx=1 (11.2.5)
a
706
11.3 Joint Probability and Transition Probability. Forward and Backward Equation 305
for all k, we must normalize W(I]) on the interval (11.2.4). This requires the in-
troduction of a normalization factor JIIwhich is explicitly dependent onj(O
b-JW
S W(I])dl] = .AI-l(f(~». (11.2.6)
a-JW
b
P(q, k+ 1) = Sd~ W(q - j(~» JII (f(~»P(~,k) , (11.2.8)
a
(11.3.2)
and D = det G. Here W is the probability distribution of the random vector 11 and
JII is a properly chosen normalization factor which secures that the fluctuations
do not kick the system away from its domain g and that probability is conserved,
i.e.,
SdnqK(q,~) = 1. (11.3.3)
'/
Here the "time" index k is related to q, whereas k' belongs to ~. Further, P 2 may
be expressed in terms of the transition probability p(q, k I~, k') and P(q, k) via
707
306 11. Discrete Noisy Maps
In the case of a stationary Markov process where the kernel (11.3.2) does not
depend on the time index k explicitly, we have furthermore
In order to derive the so-called forward equation for the transition probability p
we perform an integration over ~ in (11.3.5) to obtain
Equations (11.3.7, 1) jointly with the fact that the initial distribution may be
chosen arbitrarily yield the desired equation
holds. Subtracting (11.3.9) from (11.3.7) and using the fact that P(~,k'+1)
obeys (11.3.1) we arrive at
Renaming now the variables in the second term, i. e., Z <->~, and changing the
order of integration we obtain
Again we made use of the arbitrariness of the initial distribution P(~, k'). Equa-
tion (11.3.11) is the backward equation for the transition probability p.
Equations (11.3.8 and 11) complete the description of process (11.1.1) in terms of
probability distributions. Moments, correlation functions, etc., may now be
defined in the standard way.
708
11.5 Path Integral Solution 307
(11.4.2)
(11.4.4)
(11.4.5)
(11.4.6)
which can be considered as a path integral. To make contact with other formula-
tions we specialize W to a Gaussian distribution which we may take, without
restriction of generality, in the form
k
P(q,k+ 1) = JDqJDs TID(q/)-l exp ["'J , (11.5.4)
/=1
709
308 11. Discrete Noisy Maps
where
k
Ds = II [(1I(2n»d ns l ] ,
1=1
(11.5.5)
JdnqP(q,O) = 1. (11.6.1)
1
where P(k) is the probability of finding the system within Y without having
reached the boundary up to time k. Combining (11.6.1 and 2) we obtain the
probability that the system has reached the boundary of f during k, namely
1 - P(k). Finally, the probability that the system reaches the boundary of Y
between k and k + 1 is given by P(k) - P(k+ 1). It now becomes a simple matter
to obtain the mean first passage time by
(r) =
k=O
E(k+1)[Pck)-P(k+l)]. (11.6.3)
At this stage it should be noted that the mean first passage time not only depends
on'f/ but also on the initial distribution P(q, 0) [compare (11.6.1 )]. It is this fact
which suggests the introduction of the conditional first passage time (r(q». This
710
11.6 The Mean First Passage Time 309
is the mean first passage time for a system which has been at q at time k = 0 with
certainty. Obviously we have
00
where p(q Ie;, k) is the corresponding transition probability. The relation between
(11.6.3 and 4) is given by
In the following we shall use the fact that within .1/ the transition probability
p(q I c;,k) obeys the backward equation (11.3.11), which allows us to rewrite
(11.6.4)
(r(c;» = r
k=O
(k+ 1) Jdnq Jdnz[J(c;-z)-K(z,c;)]p(q Iz,k).
.y.y
(11.6.6)
under the integral. Using the definition of (r(q» according to (11.6.4) and
applying the backward equation again we arrive at
where R is given by
R = r
k=O
(k+1>Jd nq[jj(qlc;,k)-p(qlc;,k+2)].
.y
(11.6.9)
Jdnq
l'
r
k=O
[jj(q Ic;,k) - p(q Ic;,k+ 1)] = 1 , (11.6.11)
performing the summation over k, and replacing c; by q, our final result reads
Equation (11.6.12) contains the result announced in the beginning of this section:
We find an inhomogeneous integral equation for the conditional first passage
time (r(q» for the discrete time process (11.1.1).
711
310 11. Discrete Noisy Maps
(11.7.1)
,
W(11k) =
(- P
det-
(2n)n
)112 exp (1 _ )
- - 11kP11k ,
2
(11.7.2)
where Pdenotes a symmetric positive matrix. We note that (11.7.1) together with
(11.7.2) may be visualized as a linearized version of (11.1.1) around a fixed point.
If in addition the fluctuations are small, we are allowed to neglect the effect of
the boundaries in the case of a finite domain q) when the fixed point is far
enough from the boundaries. Using (11.7.1 and 2), the kernel (11.3.2) obviously
has the form
(11.7.3)
P(~,k) =
detB
(- -
(2n)n
)112 exp [1 __
- -(~- ~o)B(~- ~o)
2
]
, (11.7.4)
where ~o denotes the center of the probability distribution a time k and B again is
a positive symmetric matrix. Indeed, inserting both (11.7.4, 3) into (11.1.7) we
obtain
P(q,k+1)= (detB.de!p)ll2 +('dnC;exp{ ... }, (11.7.5)
(2n) -00
where
(11.7.6)
e= ~' + a (11.7.7)
712
11. 7 Linear Dynamics and Gaussian Noise 311
and choosing
(11.7.8)
Identifying
B= B k + 1; B = Bk
JII= ,Ak+l; Ak= (2n)-n12(detp)1/2 (11.7.10)
qo = qk+l; ~o = qk
and comparing (11.7.4) with (11.7.9), we immediately find the recursion relations
(11.7.11)
(11.7.12)
112
.Ai, =;v, . ( det P ) (11.7.13)
k+l 'k det(ATpA + B k )
(11.7.14)
(11. 7 .15)
(11. 7 .16)
(11. 7.17)
Finally we mention that the Gaussian case also reveals instabilities of the system.
Indeed, as soon as an eigenvalue a of the matrix A crosses Ia I = 1 due to the
variation of an external parameter, a corresponding divergence in the variance of
the probability distribution indicates the instability.
713
12. Example of an Unsolvable Problem in Dynamics
When looking back at the various chapters of this book we may note that even by
a seemingly straightforward extension of a problem (e. g., from periodic to quasi-
periodic motion) the solution of the new problem introduces qualitatively new
difficulties. Here we want to demonstrate that seemingly simple questions exist in
dynamic systems which cannot be answered even in principle. Let us consider a
dynamic system whose states are described by state vectors q. Then the system
may proceed from any given state to another one q' via transformations, A, B, C
within a time interval T,
We assume that the inverse operators A -1, B- 1, ••• exist. Obviously, A, B, C, ...
form a group. We may now study all expressions (words) formed of A, A-I, B,
B- 1, etc., e.g., BA -Ie. We may further define that for a number of specific
words, W(A,B,C ... ) = 1, e.g., BC = 1. This means that after application of C
and B any initial state q of the dynamic system is reached again. Then we may ask
the following question: given two words WI (A,B, ... ) and W2 (A,B, ... ) can we
derive a general procedure by which we can decide in finitely many steps whether
the dynamic system has reached the same end points q 1 = q2, if it started from the
same initial arbitrary point qo. That is, we ask whether
(12.2)
or (12.3)
(12.4)
This example shows clearly that some care must be exercised when problems
are formulated. This is especially so if general solutions are wanted. Rather, one
should be aware of the fact that some questions can be answered only with
respect to restricted classes of equations (or problems). It is quite possible that
such a cautious approach is necessary when dealing with self-organizing systems.
715
13. Some Comments on the Relation Between Synergetics
and Other Sciences
In the introduction we presented phenomena which are usually treated within dif-
ferent disciplines, so that close links between synergetics and other disciplines
could be established at that level. But the present book is primarily devoted to the
basic concepts and theoretical methods of synergetics, and therefore in the fol-
lowing the relations between synergetics and other disciplines are discussed at this
latter level. Since synergetics has various facets, a scientist approaching it from
his own discipline will probably notice those aspects of synergetics first which
come closest to the basic ideas of his own field. Based on my discussions with
numerous scientists, I shall describe how links can be established in this way.
Then I shall try to elucidate the basic differences between synergetics and the
other disciplines.
When physicists are dealing with synergetics, quite often thermodynamics is
brought to mind. Indeed, one of the most striking features of thermodynamics is
its universality. Its laws are valid irrespective of the different components which
constitute matter in its various forms (gases, liquids, and solids). Thermo-
dynamics achieves this universal validity by dealing with macroscopic quantities
(or "observables") such as volume, pressure, temperature, energy or entropy.
Clearly these concepts apply to large ensembles of molecules, but not to in-
dividual molecules. A closely related approach is adopted by information theory,
which attempts to make unbiased estimates about systems on which only limited
information is available. Other physicists recognize common features between
synergetics and irreversible thermodynamics. At least in the realm of physics,
chemistry, and biology, synergetics and irreversible thermodynamics deal with
systems driven away from thermal equilibrium.
Chemists and physicists are struck by the close analogy between the various
macroscopic transitions of synergetic systems and phase transitions of systems in
thermal equilibrium, such as the liquid - gas transition, the onset of ferro-
magnetism, and the occurrence of superconductivity. Synergetic systems may
undergo continuous or discontinuous transitions, and they may exhibit features
such as symmetry breaking, critical slowing down, and critical fluctuations,
which are well known in phase transition theory.
The appropriate way to cope with fluctuations, which are a necessary part of
any adequate treatment of phase transitions, is provided by statistical mechanics.
Scientists working in that field are delighted to see how typical equations of their
field, such as Langevin equations, Fokker-Planck equations and master equa-
tions, are of fundamental importance in synergetics. Electrical engineers are im-
mediately familiar with other aspects of synergetics, such as networks, positive
13. Some Comments on the Relation Between Synergetics and Other Sciences 315
and negative feedback, and nonlinear oscillations, while civil and mechanical
engineers probably consider synergetics to be a theory of static or dynamic in-
stabilities, postbuckling phenomena of solid structures, and nonlinear oscilla-
tions. Synergetics studies the behavior of systems when controls are changed;
clearly, scientists working in cybernetics may consider synergetics from the point
of view of control theory.
From a more general point of view, both dynamic systems theory and syn-
ergetics deal with the temporal evolution of systems. In particular, mathema-
ticians dealing with bifurcation theory observe that synergetics - at least in its
present stage - focusses attention on qualitative changes in the dynamics (or
statics) of a system, and in particular on bifurcations. Finally, synergetics may be
considered part of general systems theory, because in both fields scientists are
searching for the general principles under which systems act.
Quite obviously, each of the above-mentioned disciplines (and probably
many others) have good reason to consider synergetics part of them. But at the
same time in each case, synergetics introduces features, concepts, or methods
which are alien to each specific field. Thermodynamics acts at its full power only
if it deals with systems in thermal eqUilibrium, and irreversible thermodynamics
is confined to systems close to thermal equilibrium. Synergetic systems in phys-
ics, chemistry, and biology are driven far from thermal equilibrium and can
exhibit new features such as oscillations. While the concept of macroscopic
variables retains its importance in synergetics, these variables, which we have
called order parameters, are quite different in nature from those of thermody-
namics. This becomes especially clear when thermodynamics is treated with the
aid of information theory, where numbers of realizations are computed under
given constraints. In other words, information theory and thermodynamics are
static approaches, whereas synergetics deals with dynamics.
The nonequilibrium phase transitions of synergetic systems are much more
varied than phase transitions of systems in thermal equilibrium, and include
oscillations, spatial structures, and chaos. While phase transitions of systems in
thermal equilibrium are generally studied in their thermodynamic limit, where
the volume of the sample is taken to be infinite, in most nonequilibrium phase
transitions the geometry of the sample plays a crucial role, leading to quite dif-
ferent structures. Electrical engineers are quite familiar with the concepts of non-
linearity and noise, which also playa fundamental role in synergetics. But syn-
ergetics also offers other insights. Not only can synergetic processes be realized
on quite different substrates (molecules, neurons, etc.), but synergetics also deals
with spatially extended media, and the concept of phase transitions is alien to
electrical engineering. Similar points may be made with respect to mechanical en-
gineering, where in general, fluctuations are of lesser concern. Though in cyber-
netics and synergetics the concept of control is crucial, the two disciplines have
quite different goals. In cybernetics, procedures are devised for controlling a sys-
tem so that it performs in a prescribed way, whereas in synergetics we change
controls in a more or less unspecified manner and study the self-organization of
the system, i. e. the various states it acquires under the newly imposed control.
The theory of dynamic systems and its special (and probably most interesting)
branch, bifurcation theory, ignore fluctuations. But, as is shown in synergetics,
717
316 13. Some Comments on the Relation Between Synergetics and Other Sciences
fluctuations are crucial at precisely those points where bifurcations occur (and
bifurcation theory should work best in the absence of fluctuations). Or, in other
words, the transition region can be adequately dealt with only if fluctuations are
taken into account. In contrast to traditional bifurcation theory (e. g. of the
Lyapunov-Schmidt type), which derives the branching solutions alone, in syn-
ergetics we study the entire stochastic dynamics in the subspace spanned by the
time-dependent order parameters. This is necessary in order to take fluctuations
into account. At the same time our approach allows us to study the stability of
the newly evolving branches and the temporal growth of patterns. Thus there is
close contact with phase transition theory, and it is possible to introduce concepts
new to bifurcation theory, such as critical slowing down, critical fluctuations,
symmetry breaking, and the restoration of broken symmetry via fluctuations. In
addition, our methods cover bifurcation sequences within such a subspace, e. g.
period-doubling sequences and frequency locking. In most cases a number of
(noisy) components are necessary to establish a coherent state, and consequently
synergetics deals with systems composed of many components; this in turn
requires a stochastic approach.
While bifurcation theory as yet excludes fluctuations, in some of its recent de-
velopments it does consider the neighborhood of branching solutions. As experts
of dynamic systems theory and bifurcation theory will notice, this book advances
to the frontiers of modern research and offers these fields new results. One such
result concerns the form of the solutions (analogous to Floquet's theorem) of
linear differential equations with quasiperiodic coefficients, where we treat a
large class of such equations by means of embedding. Another result concerns
the bifurcation of an n-dimensional torus into other tori. Finally, the slaving
principle contains a number of important theorems as special cases, such as the
center manifold theorem, the slow manifold theorem, and adiabatic elimination
procedures.
With respect to general systems theory, synergetics seems to have entered
virgin land. By focussing its attention on situations in which the macroscopic
behavior of systems undergoes dramatic changes, it has enabled us to make
general statements and to cover large classes of systems.
In conclusion, a general remark on the relation between synergetics and
mathematics is in order. This relation is precisely the same as that between the
natural sciences and mathematics. For instance, quantum mechanics is not just
an application of the theory of matrices or of the spectral theory of linear opera-
tors. Though quantum mechanics uses these mathematical tools, is has developed
its own characteristic system of concepts. This holds a fortiori for synergetics. Its
concepts of order parameters and slaving can be applied to sciences which have
not yet been mathematized and to ones which will probably never be mathe-
matized, e. g. the theory of the development of science.
718
Appendix A: Moser's Proof of His Theorem
Lemma A.I.l
We assume that the vector F (6.3.27) (with its subvectorsJ, g, G) is a real analytic
function of period 2 n: in /fit, ... , /fin' For a given r > 0 we introduce the norm
which is finite for some positive r. Here II· .. II denotes sum over the modulus of
the individual vector components.
For any real analytic periodic function the Fourier coefficients decay ex-
ponentially. More precisely, if
then
(A.lo3)
where the integration is taken over 0 ~ /fI v ~ 2 n:. Now shifting the integration
domain into the complex to 1m {/fl v} = - p sign {i v} gives
(A.loS)
Lemma A.l.2
Let FE rYt and F be analytic in 11m {fllv} 1< r. Then the unique solution of
is analytic in the same domain. If 0 < p < r < 1, one has for U the estimate
(A. 1.7)
(A. 1.9)
the condition
The latter sum always converges and can be estimated as follows. With
a= r - p ~ 1 we have
ar+ n E (1Ij II' + 1) exp( - IIj lIa) ~ E (arllill r+ 1) exp( - IIj lIa)a n , (A.1.13)
j j
720
A.2 The Most General Solution to the Problem of Theorem 6.2.1 319
(A. 1.14)
(A.1.1S)
Theorem A.I.I. There exist unique formal expansions in powers of e for A(e),
dee), D(e) and for u(lf/, e), v(lf/, e), V(If/, e) which formally satisfy the conditions
of theorem 6.2.1 and the normalization that all coefficients in this expansion of
u, V, V belong to fl. The proof is evident from the preceding discussion. Com-
parison of coefficients leads at each step to equations of the type (6.3.33), which
by lemma A.1.2 admit a unique solution with the normalization U E fl. The more
difficult proof of the convergence of the series expansion in e is settled in Sect.
A. 3. However, in Sect. A.3 we shall not be able to ensure this normalization and
therefore we investigate now the totality of all formal solutions.
We discuss what the meaning of the arbitrary constants (of the null-space) is. To
find the most general solution, U E fl, and .AI to the problem of Moser's theorem
we let
be the unique particular solution which satisfies the normalization that all vectors
lying in the null-space vanish, i. e., U E fl. It transforms the differential equations
(6.3.1 and 2) into a system whose linearization is
If/=W
X =AX
J. (A.2.2)
721
320 Appendix A: Moser's Proof of His Theorem
which transforms the system (A.2.2) into itself will give rise to another solution
(A.2.4)
to theorem 6.2.1. Here (A.2.4) denotes the composition of q; and 0;;. For this
reason we determine the self-transformations 0;; of (A.2.2). Inserting (A.2.3) into
(A.2.2) and requiring that the differential equation in the new variables
",', X' has the same form as before, we find that [L has the shape (6.3.29) with
L j =L 2 =L 3 ]
where (A2.S)
(A.2.6)
Equation (A.2.S) means that 0 is a vector of the nUll-space, and these self-trans-
formations have the form
Ab=O, (A.2.8)
AB=BA. (A2.9)
We shall denote the group of self-transformation Q:. Thus, both °Il and (11 0 'f,'
are solutions to theorem 6.2.1. In fact, they form the most general solution
provided Lt, d, D are considered given. Indeed, one may show that these latter
quantities are uniquely determined. Since this proof is more of a technical nature
we do not present it here but rather formulate the corresponding lemma.
Lemma A.2.1
Let q; in (A.2.1) be the unique (normalized) formal transformation and let
N~ [t] (A.2.l0)
be the corresponding modifying term found in theorem ALL Thus '11 trans-
forms the system (6.3.1 and 2) into a system (6.1.26 and 27) whose linearization is
given by
722
A.3 Convergent Construction 321
'" =
i =AX
w J. (A.2.11)
Let °Il, N (where <ji = identity, N = 0 for e = 0) be any other formal expansion
with this property. Then there exists a transformation CG' E~, cf. (A.2.7), such
that q;; = q; 0 CG' and N = N.
asn---+ oo , (A.3.1)
we obtain a factor (rn-1 - rn) -, = O(n 2,) going from the determination of the co-
efficients of e n - 1 to those of en. This leads then to the series I(n! )2, en, in-
dicating the divergence of the series. Still, the convergence of the series can be
established by using an alternate construction which converges more rapidly
(Sects. 5.2 and 5.3). This idea is from Kolmogorov, Moser's proof here being a
generalization and sharpened version of that approach.
We shall describe an iteration method in which at each step the linear equa-
tions of lemma A.l.2 have to be solved but the precision increases with an
exponent 3/2 so that the previous series can be replaced by
ip=a+/ J (A.3.2)
~=b+B~+g ,
where a = (a1' ... ,an) vary freely, while b = (b 1, ... ,b m ) and the m by m matrix
B are restricted by
723
322 Appendix A: Moser's Proof of His Theorem
Here w = (Wt, ... , wn) and the eigenvalues At, ... , Am of the matrix A satisfy
the condition
(A.3.4)
(A.3.S)
except the finitely many (j,k) = (O,k) for which the lhs of (A.3.4) vanishes. The
number r is chosen > n - 1 to ensure the existence of such w, A.
In the following it will be decisive to consider a, b, B as variables [under the
linear restrictions (A.3.3)] and we shall specify the complex domain of these
variables as well as that of lp, ~, by
(A.3.6)
and we require with a fixed constant Co ~ 1, that q ~ coK. In the following, all
constants which depend on n, m, co, r only will be denoted by Cl> C2, C3,' •••
Theorem A.3.t. Constants 150 , c depending on co' n, m, r only exist such that if
15 < 150 and
and a = ii, b = 6, B = 13 in ':t such that (A.3.2) for this choice of a = ii, b = 6,
B = 13 is transformed by (A.3.8) into a system
(A.3.9)
In particular
724
A.3 Convergent Construction 323
A1,···,A m ·
Moreover, the ii, 6, fJ lie in
lii- W 1 161
---+~+ IB-AI<cruKo<q
- (A.3.11)
r s
c) The remainder of this section is devoted to the proof of this theorem. First
we observe that replacing ~ by s~ we can assume s = 1. Similarly, replacing t by
At, multiplying a, b, B, W, A,J, g by A -I we can normalize K to K = 1. However,
r cannot be replaced by 1 by stretching the 'P variable, since the angular variables
'PI' ... ,'Pn where chosen of period 21L Therefore, we take s = K = 1, r ~ 1,
q ~ Co ~ 1.
In the following construction the transformation 0/1 will be built up as an
infinite product of transformations
where each o/1v refines the previous approximation further. We denote the given
family of differential equations symbolically by ff = ,~o, and §j denotes the
system obtained by transforming Yo by the coordinate transformation 1&'0' etc.
Hence ,:7;, is transformed by °llv into .:7;,+ 1, and Yo by 0/10 0 0/11 0 •.• 0 (flv into
,:7;,+1'
It will be decisive in the following proof to describe precisely the domain ~v of
validity of the transformation and the differential equations. In particular, we
mention that the transformation °llv will involve a change of the parameters a, b,
B, as well as a transformation of the variables 'P, ~. To make this change of
variables 111v clear we drop the subscript v and write o/1v in the form
o 'IIm{III}I<r
c/'v+I' 'Y v+I' Ivl<s
A v+l, la-wl+Jl!l+IB_AI<qv+I,
rv+1 sv+1 (A.3.15)
725
324 Appendix A: Moser's Proof of His Theorem
where the sequence rv, sv, qv will be chosen in such a manner that f/lv maps 0'v+ 1
into
(A.3.16)
and such that the family of differential equations ,CJ~ is mapped into a system
')'~+1 which approximates (A.3.9) to a higher degree than ,)'~. We shall drop the
index v and write JI~ in the form (A.3.2) and ,7,,+1 in Greek letters
rp=a+(/I
-
X=P+BX+ S ,
J'
III 9tv+1 (A.3.17)
(A.3.18)
(A.3.19)
and construct 1iv in such a manner that it maps 2v+ 1 into 0'v and that for the
transformed system 3'\,+1 we have the corresponding error estimate
1411 lsi (J 5:
--+--<rv+lvv+l-qv+21ll 7 v +I'
_ • (
(A.3.20)
rv+1 sv+1
(A.3.21)
726
A.3 Convergent Construction 325
d) If the above statement and the estimates (A.3.20, 21) are established, the
above theorem follows readily as we will now show. Since <71v maps 91v + 1 into 9:1v ,
the composite transformation litv = <710 0 <711 0 ... 0 <71v maps 9:1v + 1 into 010 and
can be estimated by
(A.3.22)
uniformly and the transformation ~oo is analytic in 9:100' Moreover, the above
inequality implies (A.3.12) for X = 0 if c is chosen >2c4 • Since it is independ-
ent of X and v depends linearly on X, it remains to estimate ovloX. The term
1 + ovloX is the product of the corresponding terms 1 + oVvloX and, since
I oVvloX I < c4 b v , leads to the estimate
(A.3.24)
as was to be proven.
Finally, the determination of ii, 6, iJ is obtained as the image of ~oo from the
last three components in (A.3.14). Since °llv maps 9:1v+ 1 into 9:1v, the images
(A.3.26)
(A.3.27)
(A.3.28)
727
326 Appendix A: Moser's Proof of His Theorem
if c50 is chosen small enough, proving (A.3.11) for c > 2C4' One readily verifies
that the conditions of the theorem imply those of (A.3.20) for v = 0 so that the
induction can be started.
e) This reduces the proof of theorem A.3.1 to the construction of 11= Ilv
and the proof of the estimates (A.3.20, 21).
For this purpose, we truncate!, g to its linear part
!o =!(rp,O,a,b,B) J (A.3.29)
go = g(rp, O,a,b,B) + gt;(rp, O,a, b,B) ~
and break up ifo,go) into its components in uf, ;J1 (as in Sect. A.2), which we
denote by if" ,g,j') and (/.!t, g@). Then the transformationflvwill be obtained by
solving the linearized equations
u",m=!rfI(lf/,a,b,B) J (A.3.30)
v",m + vxAX - Av = g@ (""a,b,B)
As we saw in the previous section, these equations can be uniquely solved if we
require that
(A.3.31)
a = a +!JV(a,b,B) J (A.3.32)
P+BX=b+Bx+gh(a,b,B;x) .
f) The relations (A.3.30 - 32) define the transformation 011 and we proceed to
verify that it maps1lv+l = 1l+ into 1lv = 1l. (To simplify the notation, we denote
quantities referring to v + 1, such as sv+ 1 by s +, and those referring to v without
subscript.) For this purpose we have to check that (A.3.32) can be inverted for
(A.3.33)
and that the solution a , b, B falls into 1l, see (A.3.6). We explain the argument
for the first equation ignoring the dependence on b, B.
We use the implicit function theorem: In I a - m I < rq we have through
(A.3.19,18)
(A.3.34)
and using Cauchy's estimate in the sphere Ia - m I< R = c4q + < rq /2, we find
with c2 ~ 1. (A.3.35)
728
A.3 Convergent Construction 327
The last relation can be achieved by taking C4 > 2 C2. In the sphere Ia - w I< R we
also have from (A.3.33)
(A.3.36)
and the standard implicit function theorem guarantees the unique existence of an
analytic solution a = a(a) of the equation
a-w=a-w-!JV. (A.3.37)
(A.3.38)
(A.3.40)
We may choose the same constant C4 as before by enlarging the previous one if
necessary. This proves the first half of (A.3.21), the second part having been
verified already above.
g) Having found the transformation 0/1 we transform ff,; [or (A.3.2)] into
new variables (A.3 .17) and estimate the remainder terms CP, E to complete the in-
duction proof.
For this purpose we introduce the following symbolic notation:
W= ('fI+U)
X+v
(A.3.42)
729
328 Appendix A: Moser's Proof of His Theorem
W' = (1 + 01 + )
VIfI
Ulfl
VIfI •
(A.3.43)
Finally let
(A.3.44)
Then the transformation equations which express thatfl takes (A.3.2) into
(A.3.17) take the form
where on the left side we have matrix multiplication and on the rhs the circle 0
indicates composition, i. e., substitution of ({J by 'fI + u, etc.
We compare these equations with those satisfied by solving (A.3.30, 32). In
the present notation these equations take the form
(A.3.47)
with
Fo= ( f ) (A.3.49)
g+g~X
and the arguments are 'fI, X, a, b, B. Subtracting (A.3.48) from (A.3.45) we find
after a short calculation for (/J
where
730
A.3 Convergent Construction 329
We proceed to estimate the three error terms. The second one is due to lineariza-
tion of F, the third is due to evaluation of F - or Fo - at a displaced argument
and the first is due to the fact that a was replaced by ro, etc., when we solved
(A.3.30).
h) To estimate the quantity
(A.3.54)
(the + in I fIJ I+ refers to the new domain § + as well as to r +, S +; for the success
of this approach it is essential that the norm be changed during the iteration) it
suffices to estimate the corresponding terms II 1+, III 1+, IIII 1+. This follows from
the fact that the Jacobian W' is close to one. But since the two components fIJ, E
are scaled differently one has to show that even
(A.3.55)
is close to the identity. For the diagonal elements this is obvious from (A.3.21),
and for the remaining term we have by (A.3.21) and by Cauchy's estimate
(A.3.56)
(A.3.57)
(A.3.58)
731
330 Appendix A: Moser's Proof of His Theorem
(A.3.59)
The estimation of these terms is now straightforward, but we shall show how the
various scale factors enter.
i) First we shall show
(A.3.60)
(A.3.61)
where we used (A.3.15 and 21). The other terms can be handled similarly, but the
term (B - A) v requires special attention. It does not suffice to use the estimate
IB-A 1< q from (A.3.52), but rather
(A.3.62)
since by (A.3.57) the second term in the brackets can be made small by choice of
<50 , Hence
(A.3.65)
Similarly
(A.3.66)
hence
(A.3.67)
732
A.4 Proof of Theorem 6.2.1 331
(A.3.69)
Combining the estimates (A.3.60, 67, 68) for I, II, III we have
(A.3.70)
Clearly, the optimal choice for s is found by making both terms in the bracket
equal. This agrees approximatively with our choice, as is seen by (A.3.S7). Thus,
with (A.3.18) we get
(A.3.71)
(A.3.72)
which was claimed in (A.3.20). This completes the proof of the theorem.
a) In this section we show that the series expansion constructed in Sect. 6.3
actually converges. For this purpose we make use of the existence theorem A.3.1
of the previous section and establish the convergence of some power series solu-
tion. This solution, however, may not agree with that found by formal expansion
(in theorem 6.2.1), since in Sect. 6.3 the normalization was imposed on the
factors °llv and not on the products 01/1 0 01/2 0 ... 0 'ltv. But with the help of
lemma A.2.1 we shall be able to establish the convergence of the unique solution
described in theorem 6.2.1.
For the first step we consider the equations (6.3.1 and 2) and assume that!, g
are real analytic in a fixed domain
733
332 Appendix A: Moser's Proof of His Theorem
where
(AA.2)
(AAA)
This gives a lower bound for the radius of convergence for the solutions u (1/1, e),
v( 'fI, X,e), Ii(e), b (e), B(e) which are analytic in 11m {I/I} I< r/2, Ie I< eo (v is linear
in X).
Moreover, for e = 0 the solutions constructed in Sect. A.3 reduce to
U = v = 0, Ii = ro, b = 0, B = A as one sees by setting 0 = 0 there. Hence u, v, Ii;
ro, b, B; A can be expanded into power series, without constant terms, which
converge in IeI< eo·
This proves the existence of one analytic solution
U = u(l/I, e)
v = v('fI, ~,e)
A=Ii(e)-ro (AA.5)
d = bee)
D =B(e)-A
to our problem and theorem 6.2.1 is established. In fact, as was stated at the end
of Sect. A.2, the power series for A, d, D are uniquely determined and inde-
pendent of normalizations.
b) We now turn to the proof of the convergence of the series expansion con-
structed in Sect. 6.3. It was normalized by the condition
(AA.6)
which is not necessarily satisfied for the solution (AA.5). We denote the trans-
formation given by (AA.5) by
734
A.4 Proof of Theorem 6.2.1 333
By lemma A.2.1 the most general solution is given in the form OIl 0 '6', where
'6' E~, while..1, d, D is independent of the normalization. Therefore it suffices to
show that a convergent series expansion for '6' can be found such that the expan-
sion of OIl 0 '6' agrees with that found in theorem A.lolo Since ~ is finite dimen-
sional, this assertion follows from the implicit function theorem, as we now
show. With
rp=lfI+a+U(IfI+a,e) ] (A.4.9)
~ =X+b +BX+ v(lfI+a,x+b+BX,e)
( a +U ) rJll (A.4.10)
b+v+BX E.
is fulfilled.
We decompose
(: )
into its components in.AI and fit and denote by P,A/ the projection of our function
space into .AI.
Then we try to determine a, b, B in such a manner that
P
,A/
(a +
b+BX+v
U ) _
-
(a
b+BX
) + (u)
P
,A/ v
(A.4.11)
735
334 Appendix A: Moser's Proof of His Theorem
(A.4.12)
736
Bibliography and Comments
Since the field of synergetics has ties to many disciplines, an attempt to provide a more or less
complete list of references seems hopeless. Indeed, such a list would fill a whole volume. We therefore
confine the references to those works which we used in the preparation of this book. In addition, we
quote a number of papers, articles, or books which the reader might find useful for further study. We
list the references and further reading material according to the individual chapters.
1. Introduction
1.1 What is Synergetics About?
H. Haken: Synergetics, An Introduction, 3rd ed. (Springer, Berlin, Heidelberg, New York 1983)
This reference is referred to in the present book as [I]
H. Haken, R. Graham: Synergetik - Die Lehre vom Zusammenwirken. Umschau 6, 191 (1971)
H. Haken (ed.): Synergetics (Proceedings of a Symposium on Synergetics, Elmau 1972) (Teubner,
Stuttgart 1973)
H. Haken (ed.): Cooperative Effects, Progress in Synergetics (North-Holland, Amsterdam 1974)
H. Haken: Cooperative effects in systems far from thermal equilibrium and in nonphysical system.
Rev. Mod. Phys. 47, 67 (1975)
A further source of references is the Springer Series in Synergetics, whose individual volumes are
listed in the front matter of this book.
For a popularisation see
H. Haken: Erfolgsgeheimnisse der Natur (Deutsche Verlagsanstalt, Stuttgart 1981)
1.2 Physics
The modern treatment of phase transitions of systems in thermal equilibrium rests on the renormali-
zation group approach:
K. G. Wilson: Phys. Rev. 84,3174; 3184 (1971)
K. G. Wilson, M. E. Fisher: Phys. Rev. Lett. 28, 248 (1972)
F. J. Wegener: Phys. Rev. 85, 4529 (1972); 86, 1891 (1972)
T. W. Burkhardt, J. M. J. van Leeuwen (eds.): Real-Space Renormalization, Topics Curro Phys.,
Vol. 30 (Springer, Berlin, Heidelberg, New York 1982)
Books and reviews on the subject are, for example,
K. G. Wilson, J. Kogut: Phys. Rep. 12C, 75 (1974)
C. Domb, M. S. Green (eds.): Phase Transitions and Critical Phenomena. Internat. Series of Mono-
graphs in Physics, VoIs. I - 6 (Academic, London 1972 - 76)
S. K. Ma: Modern Theory of Critical Phenomena (Benjamin, Reading, MA 1976)
Benard Instability
H. Bimard: Rev. Gen. Sci. Pures Appl. 11, 1261, 1309 (1900)
Lord Rayleigh: Philos. Mag. 32, 529 (1916)
For more recent theoretical studies on linear stability see, e. g.
S. Chandrasekhar: Hydrodynamic and Hydromagnetic Stability (Clarendon, Oxford 1961)
For nonlinear treatments see
A. Schluter, D. Lortz, F. Busse: J. Fluid Mech. 23, 129 (1965)
F. H. Busse: J. Fluid Mech. 30, 625 (1967)
A. C. Newell, J. A. Whitehead: J. Fluid Mech. 38, 279 (1969)
R. C. DiPrima, H. Eckhaus, L. A. Segel: J. Fluid Mech. 49, 705 (1971)
F. H. Busse: J. Fluid Mech. 52, 1 (1972)
F. H. Busse: Rep. Prog. Phys. 41, 1929 (1978)
Nonlinearity and fluctuations are treated in
H. Haken: Phys. Lett. 46A, 193 (1973); and, in particular, Rev. Mod. Phys. 47, 67 (1975); and
H. Haken: Synergetics, SpringerSer. Synergetics, Vol. 1, 3rd. ed. (Springer, Berlin, Heidelberg, New
York 1983)
R. Graham: Phys. Rev. Lett. 31, 1479 (1973); Phys. Rev. 10, 1762 (1974)
For recent experiments see also
G. Ahlers, R. Behringer: Phys. Rev. Lett. 40, 712 (1978)
G. Ahlers, R. Walden: Phys. Rev. Lett. 44, 445 (1981)
P. Berge: In Dynamical Critical Phenomena and Related Topics, ed. by C. P. Enz, Lecture Notes
Phys., Vol. 104 (Springer, Berlin, Heidelberg, New York 1979) p. 288
F. H. Busse, R. M. Clever: J. Fluid Mech. 102, 75 (1981)
M. Giglio, S. Musazzi, U. Perini: Phys. Rev. Lett. 47, 243 (1981)
E. L. Koschmieder, S. G. Pallas: Int. J. Heat Mass Transfer 17, 991 (1974)
J. P. GolJub, S. W. Benson: J. Fluid Mech. 100,449 (1980)
J. Maurer, A. Libchaber: J. Phys. Paris Lett. 39, 369 (1978); 40, 419 (1979); 41, 515 (1980)
G. Pfister, I. Rehberg: Phys. Lett. 83A, 19 (1981)
H. Haken (ed.): Chaos and Order in Nature, Springer Ser. Synergetics, Vol. 11 (Springer, Berlin,
Heidelberg, New York 1981), see in particular contributions by A. Libchaber and S. Fauve, E. O.
Schulz-DuBois et aI., P. Berge, F. H. Busse
H. Haken (ed.): Evolution oj Order and Chaos, Springer Ser. Synergetics, Vol. 17 (Springer, Berlin,
Heidelberg, New York 1982)
H. L. Swinney, J. P. Gollub (eds.): Hydrodynamic Instabilities and the Transition to Turbulence,
Topics Appl. Phys., Vol. 45 (Springer, Berlin, Heidelberg, New York 1981)
Texts and Monographs on Hydrodynamics
L. D. Landau, E. M. Lifshitz: Course oj Theoretical Physics, Vol. 6 (Pergamon, London, New York
1959)
Chia-Shun-Yih: Fluid Mechanics (University Press, Cambridge 1970)
C. C. Lin: Hydrodynamic Stability (University Press, Cambridge 1967)
D. D. Joseph: Stability oj Fluid Motions, Springer Tracts Nat. Phil., Vols. 27, 28 (Springer, Berlin,
Heidelberg, New York 1976)
Metereology
R. Scorer: Clouds oj the World (Lothian, Melbourne 1972)
738
Bibliography and Comments 337
Tunnel Diodes
C. Zener: Proc. R. Soc. London 145, 523 (1934)
L. Esaki: Phys. Rev. 109, 603 (1958)
R. Landauer: J. Appl. Phys. 33, 2209 (1962)
R. Landauer, J. W. F. Woo: In Synergetics, ed. by H. Haken (Teubner, Stuttgart 1973) p. 97
739
338 Bibliography and Comments
Thermoelastic Instabilities
C. E. Bottani, G. Caglioti, P. M. Ossi: 1. Phys. F. 11,541 (1981)
C. Caglioti, A. F. Milone (eds.): Mechanical and Thermal Behaviour oj Metallic Materials. Proc. Int.
School of Physics Enrico Fermi (North-Holland, Amsterdam 1982)
Crystal Growth
1. S. Langer: In Fluctuations, Instabilities and Phase Transitions, ed. by T. Riste (Plenum, New York
1975) p. 82
1. S. Langer: Rev. Mod. Phys. 52, I (1980)
1.3 Engineering
1.3.1 Civil, Mechanical, and Aero-Space Engineering: Post-Buckling Patterns, Flutter etc.
1. M. T. Thompson, G. W. Hunt: A General Theory oj Elastic Stability (Wiley, London 1973)
K. Huseyn: Nonlinear Theory oj Elastic Stability (Nordhoff, Leyden 1975)
D. O. Brush, B. D. Almroth: Buckling oj Bars, Plates and Shells (McGraw-Hili, New York 1975)
1.5 Biology
740
Bibliography and Comments 339
1.5.1 Morphogenesis
A. M. Turing: Philos. Trans. R. Soc. London B237, 37 (1952)
L. Wolpert: J. Theor. BioI. 25, 1 (1969)
A. Gierer, H. Meinhardt: Kybernetik 12, 30 (1972); J. Cell. Sci. IS, 321 (1974)
H. Haken, H. Olbrich: J. Math. BioI. 6, 317 (1978)
J. P. Murray: J. Theor. BioI. 88, 161 (1981)
C. Berding, H. Haken: J. Math. BioI. 14, 133 (1982)
1.5.4 Evolution
M. Eigen: Naturwissenschaften 58, 465 (1971)
M. Eigen, P. Schuster: Naturwissenschaften 64, 541 (1977); 65, 7 (1978); 65, 341 (1978)
W. Ebeling, R. Feistel: Physik der Selbstorganisation und Evolution (Akademie-Verlag, Berlin 1982)
741
340 Bibliography and Comments
1.7 Economy
1.8 Ecology
Ch. J. Krebs: Ecology. The Experimental Analysis oj Distribution and A bundance (Harper and Row,
New York 1972)
R. E. Rickleps: Ecology (Nelson, London 1973)
1.9 Sociology
742
Bibliography and Comments 341
1.11.5 Stochasticity
J. L. Doob: Stochastic Processes (Wiley, New York 1953)
M. Loeve: Probability Theory (van Nostrand, Princeton 1963)
R. von Mises: Mathematical Theory of Probability and Statistics (Academic, New York 1964)
Yu. V. Prokhorov, Yu. A. Rozanov: Probability Theory, Grundlehren der mathematischen Wissen-
schaften in Einzeldarstellungen, Vol. 157 (Springer, Berlin, Heidelberg, New York 1968)
R. C. Dubes: The Theory of Applied Probability (Prentice Hall, Englewood Cliffs, NJ 1968)
W. Feller: An Introduction to Probability Theory and Its Applications, Vol. I (Wiley, New York
1971)
Kai Lai Chung: Elementary Probability Theory with Stochastic Processes (Springer, Berlin, Heidel-
berg, New York 1974)
T. Hida: Brownian Motion, Applications of Mathematics, Vol. II (Springer, Berlin, Heidelberg,
New York 1980)
Statistical Mechanics
L. D. Landau, E. M. Lifshitz: In Course of Theoretical Physics, Vol. 5 (Pergamon, London 1952)
R. Kubo: Thermodynamics (North-Holland, Amsterdam 1968)
D. N. Zubarev: Non-Equilibrium Statistical Thermodynamics (Consultants Bureau, New York 1974)
Quantum Fluctuations
H. Haken: Laser Theory, in Encylopedia of Physics, Vol. XXV /2c, Light and Matter Ie (Springer,
Berlin, Heidelberg, New York 1970) and reprint edition Laser Theory (Springer, Berlin,
Heidelberg, New York 1983)
with many further references.
Chaos
H. Haken: Synergetics, Springer Ser. Synergetics, Vol. I, 3rd. ed. (Springer, Berlin, Heidelberg, New
York 1978)
H. Haken (ed.): Chaos and Order in Nature, Springer Ser. Synergetics, Vol. II (Springer, Berlin,
Heidelberg, New York 1981)
H. Haken: Order and Chaos, Springer Ser. Synergetics, Vol. 17 (Springer, Berlin, Heidelberg, New
York 1982)
H. Haken: Synergetics, Springer Ser. Synergetics, Vol. 1, 3rd. ed. (Springer, Berlin, Heidelberg, New
York 1983)
D'Arcy W. Thompson: On Growth and Form (Cambridge University Press, London 1961)
743
342 Bibliography and Comments
See H. Haken: Synergetics, Springer Ser. Synergetics, Vol. I, 3rd. ed. (Springer, Berlin, Heidelberg,
New York 1983) and references cited in later chapters. Here References are presented for
Discrete Maps are treated in Chap. 11. The Poincare map is discussed, e. g., in
R. Abraham, J. E. Marsden: Foundations oj Mechanics (Benjamin/Cummings, Reading, MA 1978)
744
Bibliography and Comments 343
745
344 Bibliography and Comments
746
Bibliography and Comments 345
Further reading
l. Moser: "Nearly Integrable and Integrable Systems", in Topics in Nonlinear Dynamics, ed. by S.
lorna (AlP Conf. Proc. 46, I 1978)
M. V. Berry: "Regular and Irregular Motion", in Topics in Nonlinear Dynamics, ed. by S . .lorna
(AlP Conf. Proc. 46, 16 1978)
Here, by an appropriate decomposition of the variables into rapidly oscillating parts and slowly
varying amplitudes, the atomic variables were expressed by the field modes (order parameters)
Other procedures are given in
H. Haken: Z. Phys. 820, 413 (1975); 821,105 (1975); 822, 69 (1975); 823, 388 (l975) and
H. Haken: Z. Phys. 829, 61 (1978); 830, 423 (1978)
The latter procedures are based on rapidly converging continued fractions, at the expense that the
slaved variables depend on the order parameters (unstable modes) at previous times (in higher order
approximation). These papers included fluctuations of the Langevin type.
In a number of special cases (in particular, if the fluctuations are absent), relations can be established
to other theorems and procedures, developed in mathematics, theoretical physics, or other disci-
plines.
Relations between the slaving principle and the center manifold theorem (and related theorems) are
studied by
A. Wunderlin, H. Haken: Z. Phys. 844, 135 (I981)
For the center manifold theorem, see
V. A. Pliss: Izv. Akad. Nauk SSSR., Mat. Ser. 28, 1297 (I964)
A. Kelley: In Transversal Mappings and Flows, ed. by R. Abraham, J. Robbin (Benjamin, New York
1967)
In contrast to the center manifold theorem, the slaving principle contains fluctuations, includes the
surrounding of the center manifold, and provides a construction of s(u, tp, I).
747
346 Bibliography and Comments
While this field seems to have been more or less dormant for a while (with the exception of
bifurcation theory in fluid dynamics), the past decade has seen a considerable increase of interest as
reflected by recent texts. We mention in particular
D. H. Sattinger: Topics in Stability and Bifurcation Theory, Lecture Notes Math., Vol. 309
(Springer, Berlin, Heidelberg, New York 1972)
D. H. Sattinger: Group Theoretic Methods in Bifurcation Theory, Lecture Notes Math., Vol. 762
(Springer, Berlin, Heidelberg, New York 1980)
G. Iooss: Bifurcation of Maps and Applications, Lecture Notes, Mathematical Studies (North-
Holland, Amsterdam 1979)
G. Iooss, D. D. Joseph: Elementary Stability and Bifurcation Theory (Springer, Berlin, Heidelberg,
New York 1980)
These authors deal in an elegant fashion with "static" bifurcation theory.
The branching of oscillatory solutions was first treated in the classical paper
E. Hopf: Abzweigung einer periodischen L6sung eines Differentialsystems. Berichte der Mathema-
tisch-Physikalischen Klasse der Sachsischen Akademie der Wissenschaften zu Leipzig XCIV, I
(1942)
For recent treatments see
J. Marsden, M. McCracken: The Hopf Bifurcation and Its Applications. Lecture Notes Appl. Math.
Sci., Vol. 18 (Springer, Berlin, Heidelberg, New York 1976)
D. D. Joseph: Stability of Fluids Motion. Springer Tracts Natural Philos., Vols. 27, 28 (Springer,
Berlin, Heidelberg, New York 1976)
A. S. Monin, A. M. Yaglom: Statistical Fluid Mechanics, Vol. I (MIT Press, Cambridge, MA 1971)
H. Haken: Synergetics, Springer Ser. Synergetics, Vol. 1, 3rd. ed. (Springer, Berlin, Heidelberg, New
York 1983)
748
Bibliography and Comments 347
9. Spatial Patterns
9.1 The Basic Differential Equations
749
348 Bibliography and Comments
D. D. Joseph: Stability oj Fluid Motions I and II, Springer Tracts Natural Philos., Vols. 27, 2R
(Springer, Berlin, Heidelberg, New York 1976)
Reaction Diffusion equations are treated, e. g., in
P. C. Fife: In Dynamics oj Synergetic Systems, Springer Ser. Synergetics, Vol. 6, ed. by H. Haken
(Springer, Berlin, Heidelberg, New York 1980) p. 97, with further references
J. S. Turner: Adv. Chern. Phys. 29, 63 (1975)
J. W. Turner: Trans. NY Acad. Sci. 36, 800 (1974), Bull. Cl. Sci. Acad. Belg. 61, 293 (1975)
Y. Schiffmann: Phys. Rep. 64, 87 (1980)
Compare also Sect. 1.5.2
H. Haken: Synergetics, Springer Ser. Synergetics, Vol. 1, 3rd. ed. (Springer, Berlin, Heidelberg, New
York 1983)
H. Haken: Synergetics, Springer Ser. Synergetics, Vol. 1, 3rd. cd. (Springer, Berlin, Heidelberg, New
York 1983)
See also the references cited in Sect. 9.1
H. Haken: Synergetics, Springer Ser. Synergetics, Vol. I, 3rd. ed. (Springer, Berlin, Heidelberg, New
York 1983)
750
Bibliography and Comments 349
/0.4.2 Exact Stationary Solution of the Fokker-Planck Equation for Systems in Detailed Balance
We follow essentially
R. Graham, H. Haken: Z. Phys. 248, 289 (1971)
H. Risken: Z. Phys. 251, 231 (1972)
For related work see
H. Haken: Z. Phys. 219, 246 (1969)
H. Haken: Rev. Mod. Phys. 47, 67 (1975)
R. Graham: Z. Phys. B40, 149 (1980)
Compare (H. Haken: Synergetics, Springer Ser. Synergetics, Vol. I, 3rd. ed. (Springer, Berlin,
Heidelberg, New York 1983)), where also another approach is outlined. That approach starts right
away from the master equation or Fokker-Planck equation and eliminates the slaved variables from
these equations.
Appendix
J. Moser: Convergent Series Expansions for Quasi-Periodic Motions. Math. Ann. 169, 136 (1967)
751
Subject Index
754
Subject Index 353
755
354 Subject Index
756
Subject Index 355
757
356 Subject Index
758
Part III
, Springer