the essence of chaos by edward lorenz

236

Post on 21-Oct-2014

1.293 views

Category:

Technology


14 download

DESCRIPTION

 

TRANSCRIPT

The Essence of Chaos

m m m

THE ESSENCE OF CHAOS

THE ESSENCE OFCHAOS

Edward N.Lorenz

Copyright 1993 by the University of Washington Press

This book is copyright under the Berne Convention.No reproduction without permission. All rights reserved.

First published in 1993.

This edition published in the Taylor & Francis e-Library, 2005.

To purchase your own copy of this or any of Taylor & Francis orRoutledges collection of thousands of eBooks please go to

www.eBookstore.tandf.co.uk.

Paperback edition first published in the United Kingdom in 1995.

Second impression 1995.

Published in the United States by University of Washington Press as avolume in the Jessie and John Danz Lecture Series.

Published in the UK by:UCL Press Limited

1 Gunpowder SquareLondon EC4A 3DE

The name of University College London (UCL) is a registered trade markused

by UCL Press with the consent of the owner.

British Library Cataloguing in Publication DataA catalogue record for this book is available from the British Library.

ISBN 0-203-21458-7 Master e-book ISBN

ISBN 0-203-27116-5 (Adobe eReader Format)ISBN: 1-85728-454-2 PB

To my grandchildren, Nicky and Sarah

Contents

Preface vii

CHAPTER 1 Glimpses of Chaos 3

It Only Looks Random

Pinballs and Butterflies

It Aint Got Rhythm

Zeroing In on Chaos

CHAPTER 2 A Journey into Chaos 25

Chaos in Action

The Heart of Chaos

Broken Hearts

Chaos of Another Species

In and Out of Chaos

CHAPTER 3 Our Chaotic Weather 77

Prediction: A Tale of Two Fluids

Meteorology: Two Tales of One Fluid

The Unperformable Experiment

Voices from Dishpans

The Five-Million-Variable DynamicalSystem

The Consequences

Prologue

Recognition

In Limbo

Searching

The Strange Attractor

The Ubiquity of Chaos

Make Your Own Chaos

Is Randomness Chaos?

CHAPTER 5 What Else Is Chaos? 161

Nonlinearity

Complexity

Fractality

APPENDIXES

1. The Butterfly Effect 179

2. Mathematical Excursions 183

3. A Brief Dynamical-Systems Glossary 203

Bibliography 211

Index 217

vi

CHAPTER 4 Encounters with Chaos 111

Preface

IN THE SPRING OF 1990 I received an invitation from theUniversity of Washington to deliver a set of lectures, as part of theseries that had been inaugurated a generation earlier through thebenevolence and farsightedness of Jessie and John Danz. Thelectures were to be delivered before a general audience, and I wasfree to choose a subject.

Some thirty years previously, while conducting an extensiveexperiment in the theory of weather forecasting, I had come acrossa phenomenon that later came to be called chaosseeminglyrandom and unpredictable behavior that nevertheless proceedsaccording to precise and often easily expressed rules. Earlierinvestigators had occasionally encountered behavior of this sort,but usually under rather different circumstances. Often they failedto recognize what they had seen, and simply became aware thatsomething was blocking them from solving their equations orotherwise completing their studies. My situation was unique in that,as I eventually came to realize, my experiment was doomed tofailure unless I could construct a system of equations whosesolutions behaved chaotically. Chaos suddenly became somethingto be welcomed, at least under some conditions, and in the ensuingyears I found myself turning more and more toward chaos as aphenomenon worthy of study for its own sake.

It was easy to decide what topic the lectures should cover. Iaccepted the invitation, and chose as a title The Essence of Chaos.Eventually a set of three lectures took shape. The first one definedchaos and illustrated its basic properties with some simpleexamples, and ended by describing some related phenomenanonlinearity, complexity, and fractalitythat had also come to becalled chaos. The second lecture dealt with the global weather asa complicated example of a chaotic system. The final one presented

an account of our growing awareness of chaos, offered aprescription via which one could design ones own chaotic systems,and ended with some philosophical speculations. In keeping withthe anticipated make-up of the audience, I displayed nomathematical formulas, and avoided technical terms except forsome that I defined as I went along.

The present volume, with the same title, is written in the spirit ofthe Danz Lectures. It contains the same material, together withadditions written to fill in the many gaps that were inevitablypresent in a limited oral presentation. The leading lecture has beenexpanded to become Chapters 1, 2, and 5, while the second one hasbeen made into Chapter 3. The final lecture, with its historicalaccount that begins with the discovery of Neptune, proceedsthrough the work of Henri Poincar and his successors, and pausesto tell of my own involvement with chaos, has become Chapter 4.

My decision to convert the lectures into a book has beeninfluenced by my conviction that chaos, along with its manyassociated conceptsstrange attractors, basin boundaries, period-doubling bifurcations, and the likecan readily be understood andrelished by readers who have no special mathematical or otherscientific background, despite the occasionally encounteredreferences to chaos as a branch of mathematics or a new science.As in the lectures, I have presented the chaos story in nontechnicallanguage, except where, to avoid excessive repetition of lengthyphrases, I have introduced and defined a number of standard terms.I have placed the relevant mathematical equations and theirderivations in an appendix, which need not be read for anunderstanding of the main text, but which may increase thevolumes appeal to the mathematically minded reader.

Of course one cannot maintain that there is no mathematics atall in the main text, except by adopting a rather narrow view ofwhat constitutes mathematics. For example, merely noting that oneillustration shows two boards sliding thirty meters down a slope,starting ten centimeters apart and ending up ten meters apart, canbe looked upon as a mathematical observation; a verbal descriptionof what the illustration depicts is then a mathematical statement.In any event, a good deal of less simple mathematics has gone intothe production of the illustrations; most of them are end productsof mathematical developments, subsequently converted intocomputer programs. The reader nevertheless need not confront the

viii

formulas, nor the programs, to be able to absorb the messages thatthe illustrations contain.

For their aid during the preparation of this work I am indebtedto many persons. First of all I must thank the Danz Foundation,without whose sponsorship of my lectures I would never have takenthe first step. I must likewise thank the University of Washingtonfor choosing me as a lecturer. I am deeply indebted to the ClimateDynamics Program of the Atmospheric Sciences Section of theNational Science Foundation, and to the programs currentdirector, Jay Fein, for supporting my research in chaos and itsapplications to the atmosphere over many years, and, mostimmediately, for making it possible for me to write the numerouscomputer programs and to perform the subsequent computationsthat have resulted in the illustrations in this volume. I wish to thankJoel Sloman for typing and otherwise assisting with not only thefinal manuscript but also the innumerable intermediate versions,Diana Spiegel for her ever-present aid in dealing with the vagariesof our computer system, and Jane McNabb for bearing the bulk ofthe administrative burden that otherwise would have fallen on me.

Thanks go to Dave Fultz of the University of Chicago forsupplying the photographs of his dishpan experiments, and to himand the American Meteorological Society for permission toreproduce them. Thanks also go to Robert Dattore and WilburSpangler of the National Center for Atmospheric Research forpreparing and making available the lengthy tape containing themany years of recorded upper-level weather data at Singapore.

I must give special recognition to Merry Caston, who has goneover the manuscript page by page, and whose pertinent commentshave led me to incorporate a good many clarifying additions andother amendments. There are many other persons with whom I havehad brief or in some cases extensive conversations, which haveexerted their influence on the words that I have written or the ideasthat I have expressed. In this connection I must particularly mentionRobert Cornett, James Curry, Robert Devaney, Alan Faller, RobertHilborn, Philip Merilees, Tim Palmer, Bruce Street, YoshisukeUeda, J.Michael Wallace, and James Yorke. To still others who mayhave similarly influenced me without my being aware of it, and alsoto some anonymous reviewers, I can only say that their names oughtto have been included.

Finally, I am most grateful to my wife, Jane, who has suppliedmoral support throughout the preparation of this volume and has

ix

accompanied me on numerous travels in search of chaotic material,and to my children Nancy, Edward, and Cheryllawyer,economist, and psychologistwho have perfectly filled the role ofthe intelligent layperson and have subjected the manuscript invarious stages of completion to their closest scrutiny.

x

THE ESSENCE OF CHAOS

2

CHAPTER 1Glimpses of Chaos

It Only Looks Random

WORDS are not living creatures; they cannot breathe, nor walk,nor become fond of one another. Yet, like the human beings whomthey are destined to serve, they can lead unique lives. A word maybe born into a language with just one meaning, but, as it grows up,it may acquire new meanings that are related but neverthelessdistinct. Often these meanings are rather natural extensions of olderones. Early in our own lives we learn what hot and cold mean,but as we mature we discover that hot pursuit and cold comfort, orhot denials and cold receptions, are not substances or objects whosetemperatures can be measured or estimated. In other instances themore recent meanings are specializations. We learn at an equallyearly age what drink means, but if later in life someone says tous, Youve been drinking, we know that he is not suggesting thatwe have just downed a glass of orange juice. Indeed, if he tellssomeone else that we drink, he is probably implying not simply thatwe often consume alcoholic beverages, but that we drink enoughto affect our health or behavior.

So it is with chaosan ancient word originally denoting acomplete lack of form or systematic arrangement, but now oftenused to imply the absence of some kind of order that ought to bepresent. Notwithstanding its age, this familiar word is not close toits deathbed, and it has recently outdone many other commonwords by acquiring several related but distinct technical meanings.

It is not surprising that, over the years, the term has often beenused by various scientists to denote randomness of one sort oranother. A recent example is provided by the penetrating bookOrder Out of Chaos, written by the Nobel Prize-winning physical

chemist Ilya Prigogine and his colleague Isabelle Stengers. Theseauthors deal with the manner in which many disorganized systemscan spontaneously acquire organization, just as a shapeless liquidmass can, upon cooling, solidify into an exquisite crystal. Ageneration or two earlier, the mathematician Norbert Wienerwould sometimes even pluralize the word, and would write abouta chaos or several chaoses when referring to systems like the hostof randomly located molecules that form a gas, or the haphazardlyarranged collection of water droplets that make up a cloud.

This usage persists, but, since the middle 1970s, the term has alsoappeared more and more frequently in the scientific literature inone or another of its recently acquired senses; one might well saythat there are several newly named kinds of chaos. In this volumewe shall be looking closely at one of them. There are numerousprocesses, such as the swinging of a pendulum in a clock, thetumbling of a rock down a mountainside, or the breaking of waveson an ocean shore, in which variations of some sort take place astime advances. Among these processes are some, perhaps includingthe rock and the waves but omitting the pendulum, whosevariations are not random but look random. I shall use the termchaos to refer collectively to processes of this sortones that appearto proceed according to chance even though their behavior is in factdetermined by precise laws. This usage is arguably the one mostoften encountered in technical works today, and scientists writingabout chaos in this sense no longer feel the need to say so explicitly.

In reading present-day accounts, we must keep in mind that oneof the other new usages may be intended. Sometimes the phenomenabeing described are things that appear to have randomarrangements in space rather than random progressions in time, likewildflowers dotting a field. On other occasions, the arrangementsor progressions are simply very intricate rather than seeminglyrandom, like the pattern woven into an oriental rug. The situationis further complicated because several other terms, notablynonlinearity, complexity, and fractality, are often used more or lesssynonymously with chaos in one or several of its senses. In a laterchapter I shall have a bit to say about these related expressions.

In his best-selling book Chaos: Making a New Science, whichdeals with chaos in several of its newer senses, James Gleick suggeststhat chaos theory may in time rival relativity and quantummechanics in its influence on scientific thought. Whether or not sucha prophecy comes true, the new science has without question

4 GLIMPSES OF CHAOS

jumped into the race with certain advantages. Systems thatpresumably qualify as examples of chaos can very often be seen andappreciated without telescopes or microscopes, and they can berecorded without time-lapse or high-speed cameras. Phenomenathat are supposedly chaotic include simple everyday occurrences,like the falling of a leaf or the flapping of a flag, as well as muchmore involved processes, like the fluctuations of climate or even thecourse of life itself.

I have said presumably and supposedly because there issomething about these phenomena that is not quite compatible withmy description of chaos as something that is random in appearanceonly. Tangible physical systems generally possess at least a smallamount of true randomness. Even the seemingly regular swingingof the pendulum in a cuckoo clock may in reality be slightlydisturbed by currents in the air or vibrations in the wall; these mayin turn be produced by people moving about in a room or trafficpassing down a nearby street. If chaos consists of things that areactually not random and only seem to be, must it exclude familiareveryday phenomena that have a bit of randomness, and beconfined to mathematical abstractions? Might not such a restrictionseverely diminish its universal significance?

An acceptable way to render the restriction unnecessary wouldbe to stretch the definition of chaos to include phenomena that areslightly random, provided that their much greater apparentrandomness is not a by-product of their slight true randomness.That is, real-world processes that appear to be behaving randomlyperhaps the falling leaf or the flapping flagshould be allowedto qualify as chaos, as long as they would continue to appearrandom even if any true randomness could somehow be eliminated.

In practice, it may be impossible to purge a real system of itsactual randomness and observe the consequences, but often we canguess what these would be by turning to theory. Most theoreticalstudies of real phenomena are studies of approximations. A scientistattempting to explain the motion of a simple swinging pendulum,which incidentally is not a chaotic system, is likely to neglect anyextraneous random vibrations and air currents, leaving suchconsiderations to the more practical engineer. Often he or she willeven disregard the clockwork that keeps the pendulum swinging,and the internal friction that makes the clockwork necessary, alongwith anything else that is inconvenient. The resulting pencil-and-paper system will be only a model, but one that is completely

GLIMPSES OF CHAOS 5

manageable. It seems appropriate to call a real physical systemchaotic if a fairly realistic model, but one with the systems inherentrandomness suppressed, still appears to behave randomly.

Pinballs and Butterflies

My somewhat colloquial definition may capture the essence ofchaos, but it would cause many mathematicians to shudder.Probably most people in other walks of life are unaware of theextent to which mathematics is dependent on definitions. Whetheror not a proposition as stated is true often depends upon just howthe words contained in it have been defined. Certainly, before onecan develop a rigorous theory dealing with some phenomenon, oneneeds an unambiguous definition of the phenomenon.

In the present instance the colloquial definition is ambiguousbecause randomness itself has two rather different definitions,although, as we shall presently see, this flaw can easily be removedby specifying the one that is intended. More serious is the simpleexpression looks random, which does not belong in a rigorousdiscussion, since things that look alike to one person often do notlook alike to another. Let us try to arrive at a working definition ofchaos while retaining the spirit of the colloquial one.

According to the narrower definition of randomness, a randomsequence of events is one in which anything that can ever happencan happen next. Usually it is also understood that the probabilitythat a given event will happen next is the same as the probabilitythat a like event will happen at any later time. A familiar example,often serving as a paradigm for randomness, is the toss of a coin.Here either heads or tails, the only two things that can ever happen,can happen next. If the process is indeed random, the probabilityof throwing heads on the next toss of any particular coin, whether50 percent or something else, is precisely the same as that ofthrowing heads on any other toss of the same coin, and it willremain the same unless we toss the coin so violently that it is bentor worn out of shape. If we already know the probability, knowingin addition the outcome of the last toss cannot improve our chancesof guessing the outcome of the next one correctly.

It is true that knowing the results of enough tosses of the samecoin can suggest to us what the probability of heads is, for that coin,if we do not know it already. If after many tosses of the coin webecome aware that heads has come up 55 percent of the time, we

6 GLIMPSES OF CHAOS

may suspect that the coin is biased, and that the probability hasbeen, is, and will be 55 percent, rather than the 50 percent that wemight have presupposed.

The coin is an example of complete randomness. It is the sort ofrandomness that one commonly has in mind when thinking ofrandom numbers, or deciding to use a random-number generator.According to the broader definition of randomness, a randomsequence is simply one in which any one of several things canhappen next, even though not necessarily anything that can everhappen can happen next. What actually is possible next will thendepend upon what has just happened. An example, which, liketossing a coin, is intimately associated with games of chance, is theshuffling of a deck of cards. The process is presumably random,because even if the shuffler should wish otherwisefor example, ifon each riffle he planned to cut the deck exactly in the middle, andthen allow a single card to fall on the table from one pile, followedby a single card from the other pile, etc.he probably could notcontrol the muscles in his fingers with sufficient precision to do so,unless he happened to be a virtuoso shuffler from a gamingestablishment. Yet the process is not completely random, if by whathappens next we mean the outcome of the next single riffle, sinceone riffle cannot change any given order of the cards in the deck toany other given order. In particular, a single riffle cannot completelyreverse the order of the cards, although a sufficient number ofsuccessive riffles, of course, can produce any order.

A deterministic sequence is one in which only one thing canhappen next; that is, its evolution is governed by precise laws.Randomness in the broader sense is therefore identical with theabsence of determinism. It is this sort of randomness that I haveintended in my description of chaos as something that looksrandom.

Tossing a coin and shuffling a deck are processes that take placein dis crete stepssuccessive tosses or riffles. For quantities thatvary continuously, such as the speed of a car on a highway, theconcept of a next event appears to lose its meaning. Nevertheless,one can still define randomness in the broader sense, and say thatit is present when more than one thing, such as more than oneprespecified speed of a car, is possible at any specified future time.Here we may anticipate that the closer the future time is to thepresent, the narrower the range of possibilitiesa car momentarilystopped in heavy traffic may be exceeding the speed limit ten

GLIMPSES OF CHAOS 7

seconds later, but not one second later. Mathematicians have foundit advantageous to introduce the concept of a completely randomcontinuous process, but it is hard to picture what such a process innature might look like.

Systems that vary deterministically as time progresses, such asmathematical models of the swinging pendulum, the rolling rock,and the breaking wave, and also systems that vary with aninconsequential amount of randomnesspossibly a real pendulum,rock, or waveare technically known as dynamical systems. Atleast in the case of the models, the state of the system may bespecified by the numerical values of one or more variables. For themodel pendulum, two variablesthe position and speed of the bobwill suffice; the speed is to be considered positive or negative,according to the direction in which the bob is currently moving. Forthe model rock, the position and velocity are again required, but, ifthe model is to be more realistic, additional variables that specifythe orientation and spin are needed. A breaking wave is so intricatethat a fairly realistic model would have to possess dozens, or morelikely hundreds, of variables.

Returning to chaos, we may describe it as behavior that isdeterministic, or is nearly so if it occurs in a tangible system thatpossesses a slight amount of randomness, but does not lookdeterministic. This means that the present state completely oralmost completely determines the future, but does not appear to doso. How can deterministic behavior look random? If truly identicalstates do occur on two or more occasions, it is unlikely that theidentical states that will necessarily follow will be perceived as beingappreciably different. What can readily happen instead is thatalmost, but not quite, identical states occurring on two occasionswill appear to be just alike, while the states that follow, which neednot be even nearly alike, will be observably different. In fact, insome dynamical systems it is normal for two almost identical statesto be followed, after a sufficient time lapse, by two states bearingno more resemblance than two states chosen at random from a longsequence. Systems in which this is the case are said to be sensitivelydependent on initial conditions. With a few more qualifications, tobe considered presently, sensitive dependence can serve as anacceptable definition of chaos, and it is the one that I shall choose.

Initial conditions need not be the ones that existed when asystem was created. Often they are the conditions at the beginningof an experiment or a computation, but they may also be the ones

8 GLIMPSES OF CHAOS

at the beginning of any stretch of time that interests an investigator,so that one persons initial conditions may be anothers midstreamor final conditions.

Sensitive dependence implies more than a mere increase in thedifference between two states as each evolves with time. Forexample, there are deterministic systems in which an initialdifference of one unit between two states will eventually increaseto a hundred units, while an initial difference of a hundredth of aunit, or even a millionth of a unit, will eventually increase to ahundred units also, even though the latter increase will inevitablyconsume more time. There are other deterministic systems in whichan initial difference of one unit will increase to a hundred units, butan initial difference of a hundredth of a unit will increase only toone unit. Systems of the former sort are regarded as chaotic, whilethose of the latter sort are not considered to constitute chaos, eventhough they share some of its properties.

Because chaos is deterministic, or nearly so, games of chanceshould not be expected to provide us with simple examples, butgames that appear to involve chance ought to be able to take theirplace. Among the devices that can produce chaos, the one that isnearest of kin to the coin or the deck of cards may well be the pinballmachine. It should be an old-fashioned one, with no flippers orflashing lights, and with nothing but simple pins to disturb the freeroll of the ball until it scores or becomes dead.

One spring in the thirties, during my undergraduate years atDartmouth, a few pinball machines suddenly appeared in the localdrugstores and eating places. Soon many students were occasionallywinning, but more often losing, considerable numbers of nickels.Before long the town authorities decided that the machines violatedthe gambling laws and would have to be removed, but they wereeventually persuaded, temporarily at least, that the machines werecontests of skill rather than games of chance, and were thereforeperfectly legal.

If this was indeed so, why didnt the students perfect their skilland become regular winners? The reason was chaos. Ascounterparts of successive tosses of a coin or riffles of a deck, letthe events be successive strikes on a pin. Let the outcome of anevent consist of the particular pin that is struck, together with thedirection from the pin to the center of the ball, and the velocity ofthe ball as it leaves the pin. Note that I am using velocity in thetechnical sense, to denote speed together with direction of motion,

GLIMPSES OF CHAOS 9

just as position with respect to some reference point implies distancetogether with direction of displacement.

Suppose that two balls depart one after the other from the samepin in slightly different directions. When the balls arrive at the nextpin, their positions will be close together, compared to the distancebetween the pins, but not necessarily close, compared to thediameter of a ball. Thus, if one ball hits the pin squarely andrebounds in the direction from which it came, the other can strikeit obliquely and rebound at right angles. This is approximately whathappens in Figure 1, which shows the paths of the centers of twoballs that have left the plunger of a pinball machine at nearly equalspeeds. We see that the angle between two paths can easily increasetenfold whenever a pin is struck, until soon one ball will completelymiss a pin that the other one hits. Thus a player will need to increasehis or her control tenfold in order to strike one more pin along anintended pathway.

Of course, the pinball machine in Figure 1 is really amathematical model, and the paths of the balls have been computed.The model has incorporated the decelerating effect of friction, alongwith a further loss of energy whenever a ball bounces from a pin ora side wall, but, in a real machine, a ball will generally acquire someside spin as it hits a pin, and this will alter the manner in which itwill rebound from the next pin. It should not alter the conclusionthat the behavior is chaoticthat the path is sensitively dependenton the initial speed.

Even so, the model as it stands fails in one respect to provide aperfect example of chaos, since the chaotic behavior ceases after thelast pin is struck. If, for example, a particular ball hits only sevenpins on its downward journey, a change of a millionth of a degreein its initial direction would amplify to ten degrees, but a change ofa ten-millionth of a degree would reach only one degree. To satisfyall of the requirements for chaos, the machine would have to beinfinitely longa possibility in a model if not in realityor elsethere would have to be some other means of keeping the ball in playforever. Any change in direction, even a millionth of a millionth ofa degree, would then have the opportunity to amplify beyond tendegrees.

An immediate consequence of sensitive dependence in any systemis the impossibility of making perfect predictions, or even mediocrepredictions sufficiently far into the future. This assertionpresupposes that we cannot make measurements that are

10 GLIMPSES OF CHAOS

GLIMPSES OF CHAOS 11

Figure 1. The pinball machine. The jagged curves are the paths of the centersof two balls that have begun their journeys at nearly equal speeds. Theradius of a ball is indicated by the distance between a pin and an abruptchange in the direction of a path.

completely free of uncertainty. We cannot estimate by eye, to thenearest tenth of a degree nor probably to the nearest degree, thedirection in which a pinball is moving. This means that we cannotpredict, to the nearest ten degrees, the balls direction after one ortwo strikes on a pin, so that we cannot even predict which pin willbe the third or fourth to be struck. Sophisticated electronicequipment might measure the direction to the nearest thousandthof a degree, but this would merely increase the range ofpredictability by two or three pins. As we shall see in a later chapter,sensitive dependence is also the chief cause of our well-knownfailure to make nearly perfect weather forecasts.

I have mentioned two types of processesthose that advance stepby step, like the arrangements of cards in a deck, and those thatvary continuously, like the positions or speeds of cars on a highway.As dynamical systems, these types are by no means unrelated. Thepinball game can serve to illustrate a fundamental connectionbetween them.

Suppose that we observe 300 balls as they travel one by onethrough the machine. Let us construct a diagram containing 300points. Each point will indicate the position of the center of one ballwhen that ball strikes its first pin. Let us subsequently construct asimilar diagram for the second strike. The latter diagram may thenbe treated as a full-scale map of the former, although certainly arather distorted map. A very closely spaced cluster of points in thefirst diagram may appear as a recognizable cluster in the second.Dynamical systems that vary in discrete steps, like the pinballmachine whose events are strikes on a pin, are technically knownas mappings. The mathematical tool for handling a mapping is thedifference equation. A system of difference equations amounts to aset of formulas that together express the values of all of the variablesat the next step in terms of the values at the current step.

I have been treating the pinball game as a sequence of events, butof course the motion of a ball between strikes is as preciselygoverned by physical laws as are the rebounds when the strikesoccur. So, for that matter, is the motion of a coin while it is in theair. Why should the latter process be randomness, while the formerone is chaos? Between any two coin tosses there is humanintervention, so that the outcome of one toss fails to determine theoutcome of the next. As for the ball, the only human influence onits path occurs before the first pin is struck, unless the player has

12 GLIMPSES OF CHAOS

mastered the art of jiggling the machine without activating the tiltsign.

Since we can observe a ball between strikes, we have the optionof plotting diagrams that show the positions of the centers of the300 balls at a sequence of closely spaced times, say every fiftieth ofa second, instead of only at moments of impact. Again each diagramwill be a full-scale map of the preceding one. Now, however, theprominent features will be only slightly changed from one diagramto the next, and will appear to flow through the sequence.Dynamical systems that vary continuously, like the pendulum andthe rolling rock, and evidently the pinball machine when a ballscomplete motion is considered, are technically known as flows. Themathematical tool for handling a flow is the differential equation.A system of differential equations amounts to a set of formulas thattogether express the rates at which all of the variables are currentlychanging, in terms of the current values of the variables.

When the pinball game is treated as a flow instead of a mapping,and a simple enough system of differential equations is used as amodel, it may be possible to solve the equations. A completesolution will contain expressions that give the values of the variablesat any given time in terms of the values at any previous time. Whenthe times are those of consecutive strikes on a pin, the expressionswill amount to nothing more than a system of difference equations,which in this case will have been derived by solving the differentialequations. Thus a mapping will have been derived from a flow.

Indeed, we can create a mapping from any flow simply byobserving the flow only at selected times. If there are no specialevents, like strikes on a pin, we can select the times as we wishforinstance, every hour on the hour. Very often, when the flow isdefined by a set of differential equations, we lack a suitable meansfor solving themsome differential equations are intrinsicallyunsolvable. In this event, even though the difference equations ofthe associated mapping must exist as relationships, we cannot findout what they look like. For some real-world systems we even lackthe knowledge needed to formulate the differential equations; canwe honestly expect to write any equations that realistically describesurging waves, with all their bubbles and spray, being driven by agusty wind against a rocky shore?

If the pinball game is to chaos what the coin toss is to completerandomness, it has certainly not gained the popularity as a symbolfor chaos that the coin has enjoyed as a symbol for randomness.

GLIMPSES OF CHAOS 13

That distinction at present seems to be going to the butterfly, whichhas easily outdistanced any potential competitors since theappearance of James Gleicks book, whose leading chapter isentitled The Butterfly Effect.

The expression has a somewhat cloudy history. It appears to havearisen following a paper that I presented at a meeting in Washingtonin 1972, entitled Does the Flap of a Butterflys Wings in Brazil SetOff a Tornado in Texas? I avoided answering the question, butnoted that if a single flap could lead to a tornado that would nototherwise have formed, it could equally well prevent a tornado thatwould otherwise have formed. I noted also that a single flap wouldhave no more effect on the weather than any flap of any other

Figure 2. The butterfly.

14 GLIMPSES OF CHAOS

butterflys wings, not to mention the activities of other species,including our own. The paper is reproduced in its original form asAppendix 1.

The thing that has made the origin of the phrase a bit uncertainis a peculiarity of the first chaotic system that I studied in detail.Here an abbreviated graphical representation of a special collectionof states known as a strange attractor was subsequently foundto resemble a butterfly, and soon became known as the butterfly.In Figure 2 we see one butterfly; a representative of a closely relatedspecies appears on the inside cover of Gleicks book. A number ofpeople with whom I have talked have assumed that the butterflyeffect was named after this attractor. Perhaps it was.

Some correspondents have also called my attention to RayBradburys intriguing short story A Sound of Thunder, writtenlong before the Washington meeting. Here the death of a prehistoricbutterfly, and its consequent failure to reproduce, change theoutcome of a present-day presidential election.

Before the Washington meeting I had sometimes used a sea gullas a symbol for sensitive dependence. The switch to a butterfly wasactually made by the session convenor, the meteorologist PhilipMerilees, who was unable to check with me when he had to submitthe program titles. Phil has recently assured me that he was notfamiliar with Bradburys story. Perhaps the butterfly, with itsseeming frailty and lack of power, is a natural choice for a symbolof the small that can produce the great.

Other symbols have preceded the sea gull. In George R.Stewartsnovel Storm, a copy of which my sister gave me for Christmas whenshe first learned that I was to become a meteorology student, ameteorologist recalls his professors remark that a man sneezing inChina may set people to shoveling snow in New York. Stewartsprofessor was simply echoing what some real-world meteorologistshad been saying for many years, sometimes facetiously, sometimesseriously.

It Aint Got Rhythm

There are a number of heart conditions known as arrhythmias;some of them can be fatal. The heart will proceed to beat at irregularintervals and sometimes with varying intensity, instead of tickinglike a metronome. It has been conjectured that arrhythmias aremanifestations of chaos. Clearly they entail an absence of some

GLIMPSES OF CHAOS 15

order that ought to be present, but let us see how they, or anyprocesses that lack rhythm, may be related to sensitive dependence.

Precise definitions are not always convenient ones. Havingdefined chaos in terms of sensitive dependence, we may discoverthat it is difficult to determine whether certain phenomena arechaotic.

The pinball machine should present no problem. If we havewatched a ball as it rolls, and have noted its position and velocityat some initial time, it should be fairly easy for us to set a newball rolling from about the same position with about the same speedand direction, and then see whether it follows almost the same path;presumably it will not, if my analysis has been correct. If insteadour system is a flag flapping in the wind, we do not have this option.We might record the flags behavior for a while by taking high-speedphotographs, but it would be difficult to bend the flag into the shapeappearing on a selected exposure, especially with a good windblowing, and even harder to set each point of the flag moving withthe appropriate velocity.

Perhaps on two occasions we might hold the flag taut with somestout string, and let the initial moments be the times when the stringsare cut. If we then worry that subsequent differences in behaviormay result from fluctuations in the wind instead of from intrinsicproperties of the flag, we can circumvent the problem bysubstituting an electric fan. Nevertheless, we shall have introducedhighly unusual initial statesflags in a breeze do not ordinarilybecome tautand these states may proceed to evolve in a highlyunusual manner, thereby invalidating our experiment as a test forchaos.

Fortunately there is a simpler approach to systems like theflapping flag, and it involves looking for rhythm. Before we canjustify it, we must examine a special property of certain dynamicalsystems, which is known as compactness.

Suppose that on a round of golf you reach the tee of your favoritepar-three hole and drive your ball onto the green. If you shoulddecide to drive a second ball, can you make it come to rest withina foot of the first one, if this is what you wish? Presumably not;even without any wind the needed muscular control is too great,and the balls might not be equally resilient. If instead you haveseveral buckets of balls and continue to drive, you will eventuallyplace a ball within a foot of one that you have already driven,although perhaps not close to the first one. This will not be because

16 GLIMPSES OF CHAOS

your game will have improved during the day, but simply becausethe green, or even the green plus a few sand traps and a pond, cannothold an unlimited number of balls, each one of which is at least afoot from any other. The argument still holds if the critical distancebetween the centers of two balls is one centimeter, or somethingstill smaller, as long as your caddy removes each ball and marks thespot before the next drive. Of course you may need many morebuckets of balls.

The surface of the green is two-dimensionala point on it maybe specified by its distance and direction from the cupand its areais bounded. Many dynamical systems are like the ball on the green,in that their states may be specified by the values of a finite numberof quantities, each of which varies within strict bounds. If inobserving one of these systems we wait long enough, we shalleventually see a state that nearly duplicates an earlier one, simplybecause the number of possible states, no one of which closelyresembles any other one, is limited. Systems in which arbitrarilyclose repetitionscloser than any prespecified degreemusteventually occur are called compact.

For practical purposes the flag is a compact system. The bends inthe flag as it flaps often resemble smooth waves. Let us define thestate of the flag by the position and velocity of each point of a well-chosen grid, perhaps including the centers of the stars if it is anAmerican flag, instead of using every point on the flag. Two statesthat, by this definition, are nearly alike must then eventually occur,and any reasonable interpolation will indicate that the positions andvelocities of any other points on the flag will also be nearly alike inthe two instances.

Our pinball machine is not a compact system. Not only do nearrepetitions not have to occur while a single ball is in play, but theycannot, since friction is continually tending to slow the ball, andthe only way that the ball can regain or maintain its speed is to rollcloser to the base of the machine. However, we can easily visualizea modified system where near repetitions are inevitable.

Imagine a very long pinball machine; it might stand on a gentlysloping sidewalk outside a local drugstore, and extend for a cityblock. Let the playing surface be marked off into sections, say onemeter long, and let the arrangement of the pins in each section beidentical with that in any other. Except for being displaced fromeach other by one or more sections, the complete paths of two balls,occupying similar positions in different sections, and moving with

GLIMPSES OF CHAOS 17

identical velocitiesspeeds and directionswill be identical also.Thus for practical purposes the state of the system may be definedby the balls velocity, together with its position with respect to akey point in its section, such as the uppermost pin. These quantitiesall vary within limited ranges. It follows that if the city block is longenough, a near repetition of a previous state must eventually occur.

In Figure 3 we see the computed path of the center of a single ballas it works its way past eighty pins, in a machine that is not onlyelongated but very narrow; the playing space is only twice as wideas the ball. Each pin is set one-quarter of the way in from one sidewall or the other. The long machine is displayed as four columns;the right-hand column contains the first twenty pins, and eachremaining column is an extension of the one to its right. The verticalscale has been compressed, as if we were looking from a distancewith our eye just above the playing surface, so that the circularupper end appears elliptical, as would an area directly below theball, if it were shown. You may find the figure easier to study if yougive it a quarter turn counterclockwise and look at the path as agraph. Evidently the ball completely misses about a third of the pins,but the rebounds from pins somewhat outnumber those from sidewalls.

We see that the path from above the first to below the seventhpin in the third column nearly duplicates the path in the same partof the second column. With the initial speed used in thecomputation, the expected close repetition of an earlier state has setin only forty pins from the start. If the width of the playing spacehad not been so drastically restricted, thus limiting the possiblepositions of the ball, we could have expected a much longer wait,but not an eternal one.

Suppose now that some compact system, perhaps an elongatedpinball machine or perhaps a flag, is not chaotic; that is, it does notexhibit sensitive dependence. Assume that the inevitable very nearrepetition of an initial state occurs after ten seconds, although itcould be after an hour or longer. From that time onward, thebehavior of the system will nearly repeat the behavior that occurredten seconds earlier, until, after ten more seconds, another nearrepetition of the initial state will occur. After another ten secondsthere will be still another near repetition, and so on, and ultimatelythe behavior may become periodic; that is, a particular pattern ofbehavior may be repeated over and over again, in this case at ten-second intervals. It is also possible that a true ten-second periodicity

18 GLIMPSES OF CHAOS

will fail to become established, because even if, say, the tenthoccurrence nearly repeats the ninth, and the ninth nearly repeatsthe eighth, etc., the tenth need not be particularly close to the first.In a case like this, the system is called almost periodic. If, then, some

GLIMPSES OF CHAOS 19

Figure 3. The elongated pinball machine, with the path of the center of asingle ball. For display the machine has been divided into four columns,with columns to the left following those to the right. The vertical scale hasbeen compressed; hence a ball striking a pin from above appears to be closerthan one striking from the side.

other compact system fails to exhibit periodicity or almost-periodicity, it presumably is chaotic.

In the case of Figure 3, we see that, if the behavior were notchaotic, the small differences between the early parts of the thirdand second columns would not amplify, and the remainder of thethird column would look almost like the remainder of the second,while the fourth column would look almost like the third. Actually,however, the ball strikes the first pin in the fourth column andmisses the second, while just the opposite happens in the thirdcolumn. The fact that the fourth column does not repeat the third,or more generally that the path proves to be neither periodic noralmost periodic, is therefore sufficient evidence that the ball isbehaving chaotically.

Returning to the flapping flag, we see that we may not need tophotograph it, or go through other involved procedures that mightreveal chaotic behavior. If near repetitions really do tend to occurafter about ten seconds rather than an hour or longer, our simplestprocedure may be just to listen to the continual puffs and plops,and to note whether they have a regular rhythm or whether theyseem to occur at random. I suspect, in fact, that the popularconception of something acting randomly is something varying withno discernible pattern, rather than something bearing the less easilydetected property of sensitive dependence. Indeed, absence ofperiodicity has sometimes been used instead of sensitive dependenceas a definition of chaos. Note, however, that if a system is notcompact, so that close repetitions need never occur, lack ofperiodicity does not guarantee that sensitive dependence is present.

Zeroing In on Chaos

Chaos, as a standard term for nonperiodic behavior, seems tohave received its big boost in 1975 with the appearance of a nowwidely quoted paper by Tien Yien Li and James Yorke of theUniversity of Maryland, bearing the terse title Period Three ImpliesChaos. In a mapping, a sequence of period three is one in whicheach state is identical with the state that occurred three steps earlier,but not with the state that occurred one step or two steps earlier;sequences with other periods are defined analogously. The authorsshowed that, for a certain class of difference equations, the existenceof a single solution of period three implies the existence of an infinitecollection of periodic solutions, in which every possible period

20 GLIMPSES OF CHAOS

periods 1, 2, 3, 4, is represented, and also an infinite collectionof nonperiodic solutions. This situation, in which virtually any typeof behavior may develop, seems to fit the nontechnical definitionof chaos, and it is not obvious that Li and Yorke intended tointroduce a new technical term.

They might as well have done so. In the ensuing years the termhas appeared with increasing frequency, and when, in 1987, itbecame the key word in the title of James Gleicks popular book,its permanence was virtually assured.

In the process of establishing itself as a scientific term, chaosalso picked up a somewhat different meaning. Li and Yorke hadused the term when referring to systems of equations that possessat least a few nonperiodic solutions, even when most solutions maybe periodic. In systems that are now called chaotic, most initialstates are followed by nonperiodic behavior, and only a special fewlead to periodicity. I shall refer to chaos in the sense of Li and Yorkeas limited chaos, calling chaos full chaos when it is necessary todistinguish it from the limited type.

You may wonder in what sense most solutions can behave inone manner when the few that behave otherwise are clearlyinfinite in number. Actually, occurrences of this sort are rathercommon. Consider, for example, a square and one of its diagonals.The number of distinct points on the diagonal is infinite, but, in anobvious sense, most points within the square lie off the diagonal.

Suppose that you decide to use the square as a dartboard, andthat you have a dart whose point is infinitely sharp, so that thereare an infinite number of different points that you might hit. If youraim is merely good enough to make it unlikely that you will missthe square altogether, your chances of striking a rather narrow bandsurrounding the diagonal are rather low, while your chances ofstriking a much narrower band are much lower, and the probabilitythat you will hit a point exactly on the diagonal is smaller than anypositive number that you can name. Mathematicians would say thatthe probability is zero. Clearly, however, a zero probability is notthe same thing as an impossibility; you are just as likely to hit anyparticular point on the diagonal as any particular point elsewhere.

In limited chaos, encountering nonperiodic behavior is analogousto striking a point on the diagonal of the square; although it ispossible, its probability is zero. In full chaos, the probability ofencountering periodic behavior is zero.

GLIMPSES OF CHAOS 21

There is a related phenomenon that, unlike full chaos or limitedchaos, has long been familiar to almost everyone, even if not by itstechnical name, unstable equilibrium. If you have ever tried to makea well-sharpened pencil stand on its point, you will in all likelihoodhave found that less than a second will elapse between the momentwhen you let go of the pencil and the time when it strikes the table.Simple calculations indicate that a pencil is several million times lesslikely to stand up for two seconds than for one, and several millionmillion times less likely to stand up for three seconds than for one,and that, in fact, if every human being who ever lived had devotedhis or her entire life to attempting to make sharpened pencils standon end, it would be highly unlikely that even one pencil would havestood up for six seconds. Of course, you may have somewhat betterluck if the points of your pencils become slightly worn.

Nevertheless, theory indicates that a pencil standing exactlyvertically will continue to stand forever; it is in equilibrium. A stateof equilibrium is one that remains unchanged as time advances. Anequilibrium is unstable if a state that differs slightly from it, suchas one that you might purposely produce by disturbing it a bit, willpresently evolve into a vastly different statea fallen pencil insteadof a standing one. It is stable if a slight initial disturbance fails tohave a large subsequent effect. The concept of equilibrium, stableor unstable, can be extended to include periodic behavior.

The vertical pencil is typical of systems in unstable equilibrium.The reason that in practice a pencil cannot be made to stand on itspoint is that neither the hand nor the eye can distinguish betweena pencil that is truly vertical and one that is tilted by perhaps a tenthof a degree, and the mathematical probability of picking the onevertical state from among the infinitely many that seem vertical iszero. Moreover, in the real world, even if unstable equilibrium couldbe precisely achieved, something would soon disturb it.

The definition of unstable equilibrium has much in common withthat of sensitive dependenceboth involve the amplification ofinitially small differences. The distinction between a system thatmerely possesses some states of unstable equilibrium and one thatis chaotic is that, in a system of the latter type, the future course ofevery state, regardless of whether it is a state of equilibrium, willdiffer more and more from the future courses of slightly differentstates.

Chaotic systems may possess states of equilibrium, which arenecessarily unstable. In the pinball machine such a state will occur

22 GLIMPSES OF CHAOS

if the ball comes to rest against a pin, directly upslope from it. Aslight disturbance will cause the ball to roll away to the left or right.

It may seem that in seeking bigger and better pinball machines Ihave overlooked the simplest example of all. Consider an objectmoving without friction on a horizontal planeeffectively a pinballmachine with no pins, no slope, and no walls. The object will movestraight ahead, at a constant speed. If we should alter its speed ordirection by even the minutest amount, its position will eventuallybe far from where it would have been without the alteration.

Should we call this chaos? Despite the sensitive dependence, thereis no irregularity or seemingly random behavior. All possible pathsare straight lines, expressible by simple mathematical formulas.Instead of recognizing a form of chaos with radically differentproperties, it would seem logical to conclude that our originaldefinition in terms of sensitive dependence lacks some essentialqualification.

Although there are a number of procedures for making thedefinition more acceptable, one approach is particularly simple. Theobject sliding without friction is an example of a dynamical systemin which certain quantities may assume any values initially, but willthen retain these values forever; these quantities are really constantsof the system. In the present example, the velocity remains constant.If two nearby objects have identical velocitiesspeeds anddirectionsand differ in their initial states only by virtue ofdiffering in their positions, they will not move apart.

Let us therefore amend our definition of chaos. First, for thisparticular purpose, let us refer to any quantity whose value remainsunaltered when a system evolves without our interference, but maybe altered if we introduce new initial conditions, as a virtualconstant. For an object sliding without friction on a horizontalsurface, the speed and direction are virtual constants; for one thatslides without friction inside a bowl, the speed and direction vary,but the total energy is a virtual constant. Next, let us distinguishbetween changes in initial conditions that do alter the value of atleast one virtual constant and those that do not, by calling theformer changes exterior and the latter ones interior.

We may now redefine a chaotic system as one that is sensitivelydependent on interior changes in initial conditions. Sensitivity toexterior changes will not by itself imply chaos. Concurrently, wemay wish to modify our idea as to what constitutes a singledynamical system, and decide that, if we have altered the value of

GLIMPSES OF CHAOS 23

any virtual constant, we have replaced our system by anothersystem. In that case chaos, as just redefined, will be equivalent tosensitive dependence on changes that are made within one and thesame system. For systems without virtual constants, such as theelongated pinball machine, all changes are necessarily interiorchanges, and the modified definition is the same as the previous one.For many other systems, including various objects that movewithout friction, the modified definitionsensitivity to interiorchangeswill lead to more acceptable conclusions.

Have we been glimpsing a phenomenon that can arouse enoughinterest to become the subject of an extensive scientific theory?Probably few people care whether a flag flaps chaotically orrhythmically. Chaos in the pinball game is important, andfrustrating, to anyone who is seriously trying to win. Chaos in theheartbeat, when it occurs, is of concern to all of us.

24 GLIMPSES OF CHAOS

CHAPTER 2A Journey into Chaos

Chaos in Action

THE PINBALL MACHINE is one of those rare dynamical systemswhose chaotic nature we can deduce by pure qualitative reasoning,with fair confidence that we have not wandered astray.Nevertheless, the angles in the paths of the balls that are introducedwhenever a ball strikes a pin and reboundsrather prominentfeatures of Figure 1render the system somewhat inconvenient fordetailed quantitative study. For an everyday system that will varymore smoothly, and can more easily serve to illustrate many of thebasic properties of chaotic behavior, even though it may not yieldso readily to descriptive arguments, let us consider one that stillbears some resemblance to the pinball machine. The new systemwill again be a slope, with a ball or some other object rolling orsliding down it, but there will be no pins or other obstacles to blocka smooth descent, and the slope may be of any size.

We should not expect to encounter chaos if our slope is a simpletilted plane, unless our object is an elliptical billiard ball of TheMikado fame, so let us consider a slope with a generous scatteringof smoothly rounded humps. These may even be arranged like thepins in some pinball machineperhaps the machine appearing inFigure 1but they will not have the same effect; even though anobject approaching a hump obliquely may be deflected about asmuch as a ball glancing off a pin, an object that encounters a humpstraight ahead can travel smoothly over the top instead ofrebounding. We wish to discover a system of this description thatbehaves chaotically, if, indeed, there is one.

At this point we have the usual options. We can make anexcursion to the country, and seek a slope with plenty of humps,

but without any obstacles, such as trees or boulders, that might playthe role of the pins in the machine. We can then let a soccer ball rolldown and observe its motion, but an irregularly meandering pathmay not indicate sensitive dependence, since the humps, and hencethe deflections that they produce, may not be regularly spaced.Instead of detecting chaos by its lack of rhythm, we would have tolook for sensitive dependence directly. To do this, we could releaseanother or preferably several other balls from the same point to seewhether they follow different paths, hoping that any observeddivergence will not be the product of an unobserved gust of wind.We could virtually remove the wind problem by using bowling balls,but we might be disinclined to carry several of them up the slope.

We may therefore prefer to turn to the laboratory and build ourown slope, making every hump just like every other one and spacingthem uniformly. The slope can sit on a table top, and the rollingobject can be a marble. Like the soccer ball, the marble will bedeflected by the humps, but now there will be nothing to preventits motion from varying periodically. If it follows an irregular pathanyway, we should suspect the presence of chaos. Again, we mightchoose to compare the paths of several marbles.

If we are mathematically inclined, we may wish to move from thelaboratory to the computer. Instead of physically building a slope,we can choose a mathematical formula for its topography. Insteadof watching an object as it moves down the slope, we can solve theequations that describe the manner in which its motion will vary.Let us examine in detail what can happen if our envisioned sloperesembles the one pictured in Figure 4, which shows an oblique viewof a cut-out section. The humps may remind some of us of mogulson a ski slope, and, although a computer program can handle slopesof any size with equal ease, I shall choose dimensions that are fairlytypical of ski slopes. Since some of you may live far from skiterritory, I should add that these moguls are not eastern rulers norindustrial giants who have ventured onto the snow. CompareFigure 5, where you can perhaps just detect a lone skier tackling theubiquitous moguls on the Cats Meow, a challenging slope atLoveland Basin ski area in Colorado. The moguls seem ratherregularly spaced, though not enough so to fit a simple mathematicalformula.

Since our object will travel continuously rather than in jumps,our equations will be differential equations. They will bestatements, in mathematical symbols, of one of Isaac Newtons laws

26 A JOURNEY INTO CHAOS

of motion, which simply equates the acceleration of any particle ofmatterthe rate at which its velocity is changingto the sum ofthe forces acting on the particle, divided by the mass of the particle.A complicated object can be treated as an aggregation of particles.The equations appear in a mathematical excursion in Appendix 2.

A ball or some other object rolling down a slope would acquireconsiderable spin, and to avoid the resulting mathematicalcomplications I shall assume that our object is one that will slide.It might be a flat candy bar that has fallen from a skiers pocket, oreven a loose snowboard, but not an animate skier, and I shall simplycall it a board. Its motion will be controlled by the combined actionof three forces. One of these is the pull of gravity, directed verticallydownward. Another is the resistance of friction, directed againstthe velocity. Finally there is the force that the slope exerts againstthe board, directed at right angles to the slope, and opposing theeffect of gravity to just the extent needed to keep the board slidinginstead of sinking into the slope or taking off into the air.

Laboratory models are of necessity real physical systems, eventhough they need not be like any encountered in nature, but

Figure 4. An oblique view of a section of the model ski slope.

A JOURNEY INTO CHAOS 27

mathematical models seldom duplicate any concrete systemsexactly. Real boards may be flexible, and their orientations mayvary, but in our ski-slope model we shall disregard thesepossibilities, and treat the board as if it were a single particle. It willbe convenient to choose an oversimplified law of friction, lettingthe resistance be directly proportional to the speed of the boarddoubling the speed will double the resistance. Stated otherwise, ifthe damping time is defined as the ratio of the speed to the rate atwhich friction is lowering the speed, it will be convenient to let thedamping time be constant. Its reciprocal, the coefficient of friction,will therefore be constant. It will then be easy to formulate a systemof differential equations whose solutions will reveal how the boardwill move.

Our model would be more realistic if the resistance of frictionwere made to vary also with the force of the slope against the board,so that, for example, when the board is nearly taking off,presumably because it is shooting over a mogul, the frictional effectwill be greatly reduced. Wind resistance should also be modeled,but these refinements are unlikely to have much qualitative effect.

There are different rules for laboratory and computer studies. Wetend to associate laboratory experiments with high precision, butin some instances it is possible to omit some measurementsaltogether and still obtain an answer. In a computer experimentthere is no way to begin until numerical values have been assignedto every relevant quantity.

Our model contains a number of constants. We therefore havethe option of working with any one of an infinite number ofdynamical systems, since, as we have noted, whenever we alter thevalue of any constant we produce a new system, perhaps with a newtypical behaviorobserve how oiling an aging machine, andthereby lowering its coefficient of friction, can pep up itsperformance. Systems that are formally alike except for the valuesof one or more constants are said to be members of a family ofdynamical systems. Often we refer to a whole family as simply adynamical system, when we can do so without causing confusion.

We should anticipate the possibility that only certaincombinations, if any, of values of our constantsthe average pitchof the slope, the height and spacing of the moguls, and thecoefficient of frictionwill lead to chaos. Gravity is also an essentialconstant, but its magnitude has already been determined for us aslong as we keep our experiment earthbound. In contrast to the

28 A JOURNEY INTO CHAOS

pinball model, where chaos is virtually assured, there is no rule fordetermining in advance of our computations, in the new model,what combinations of values for the constants will work. We mustdiscover them by trial and error, possibly obtaining suitable valueson the first try, possibly encountering them only after many tries,and possibly not finding them at all and concluding, perhapsincorrectly, that the model will never produce chaos.

Since I have introduced the ski slope for the purpose of illustratingthe fundamental properties of chaos, I shall choose values from asuccessful try. For convenience, let the slope face toward the south.Let its average vertical drop be 1 meter for every 4 meterssouthward. Imagine a huge checkerboard drawn on the slope, withsquares 2 meters wide and 5 meters long, and let the centers ofthe moguls be located at the centers of the dark squares, asillustrated in Figure 6, which also shows a possible path of a boarddown the slope. Let the damping time be 2 seconds. As detailed inAppendix 2, the formula selected for the topography of the slopewill place pits in the light squares of the checkerboard, as if snowhad been dug from them to build the moguls. Let each mogul rise1 meter above the pits directly to its west and east. Like the pinballmachine, the ski slope would have to be infinitely long to afford aperfect example of chaos.

We shall be encountering references to the four variables of themodel so often that I shall give them concise names. These mightas well consist of single letters, and they might as well be letters thathave served as mathematical symbols in the differential equations,or in the computer programs for solving them. Let the southwardand eastward distances of the board from some reference point, saythe center of a particular pit, be called X and Y, respectively, andlet the southward and eastward components of the velocitytherates at which X and Y are currently increasingbe called U andV. Note that when we view the slope directly from the west side, aswe nearly do in Figure 4, or as we can do by giving Figure 6 a quarterturn counterclockwise, X and Y become conventional rectangularcoordinates. Alternatively, but with some loss of mathematicalconvenience, we could have chosen the board's distance anddirection from the reference point, and its speed and direction ofmotion, as the four variables.

What I have been calling the centers of the moguls or the pits areactually the points where the slope extends farthest above or belowa simple tilted plane. These are the centers of the checkerboard

A JOURNEY INTO CHAOS 29

squares. The actual highest point in a dark square is about 1.5meters north of the center of the mogul, while the lowest point in

Figure 5. Moguls on a real-world ski slope.

30 A JOURNEY INTO CHAOS

a light square is a like distance south of the center of the pit, and Ishall call these points the high points and low points. They may be

A JOURNEY INTO CHAOS 31

Figure 6. A top view of a section of the model ski slope, with the path of asingle board sliding down it. The shaded rectangular areas of the slopeproject above a simple inclined plane, while the unshaded areas projectbelow.

detected on the forward edge of the section of the slope in Figure 4.To begin a computation, we can give the computer four numbers,

which specify numerical values of X and Y, say in meters, and Uand V, say in meters per second. In due time the computer willpresent us with more numbers, which specify the values of the samevariables at any desired later times. As our first example, let X, Y,U, and V be 0.0, 0.5, 4.0, and 2.0, implying that the board startshalf a meter due west of the center of a pit, and heads approximatelysouth-southeastward. The board will then follow the sample pathshown in Figure 6.

To see whether the board descends chaotically, let us turn toFigure 7, which shows the paths of seven boards, including the oneappearing in Figure 6, as they travel 30 meters southward. All startfrom the same west-east starting line with the same speed anddirection, but at points at 10-centimeter intervals, from 0.8 to 0.2meters west of a pit. A tendency to be deflected away from themoguls is evident. The paths soon intersect, but the states are notalike, since now the boards are heading in different directions, andsoon afterward the paths become farther apart than at first. By 10meters from the starting line, the original 0.6-meter spread has morethan doubled, and by 25 meters it has increased more than tenfold.Clearly the paths are sensitively dependent on their starting points,and the motion is chaotic.

As already noted, an essential property of chaotic behavior is thatnearby states will eventually diverge no matter how small the initialdifferences may be. In Figure 8, we let the seven boards travel 60meters down the slope, starting from points spaced only a millimeterapart, from 0.503 to 0.497 meters west of the pit. At first theseparation cannot be resolved by the picture, but by 30 meters it iseasily detectable, and subsequently it grows as rapidly as in Figure 7.The ski slope has passed another critical test. The initial separationcan be as small as we wish, provided that the slope is long enough.

Lest the paths remind some of you of ski tracks, I should hastento add that they do not follow routes that you, as knowledgeableskiers, would ordinarily choose. They might approximate paths thatyou would take if you fell and continued to slide.

The reader who looks at things as they are in nature need not beperplexed by what is meant by displacing a snowboard, not tomention a fallen skier, by 1 millimeter. We have been looking at amathematical model, which, like most models, ignores something.In this case it ignores the size of the sliding object, which,

32 A JOURNEY INTO CHAOS

presumably quite unrealistically, is assumed to slide just as it wouldif it were no larger than a pinhead, and if the slope were so smooth

A JOURNEY INTO CHAOS 33

Figure 7. The paths of seven boards starting with identical velocities frompoints spaced at 10-centimeter intervals along a west-east line. The smalldiamonds indicate the locations of the centers of the moguls.

that a pinhead could continue to slide. It would be quite possibleto write a new computer program, which could deal with sliding

Figure 8. The paths of seven boards, starting with identical velocities frompoints spaced at 1-millimeter intervals along a west-east line.

34 A JOURNEY INTO CHAOS

objects of various sizes and shapes, but which would have to bebased on a more involved mathematical analysis.

For the alternative means of detecting chaosdirectly observinga lack of periodicitylet us turn to Figure 9, which extends theoriginal path to 600 meters down the slope. The board shows anoticeable tendency to wiggle about lines that pass eithersoutheastward or south-westward between the moguls, but theswitches from one direction to the other show no sign of occurringat regular intervals. Let us see what we can infer from this behavior.

First, we are interested in periodicity, or the lack of it, as timeincreases, but Figure 9 shows only what happens as downslopedistance, i.e., X, increases. We should note, then, that since theboard always moves down the slope and its speed does not vary toogreatly, distance can serve as an approximate measure of time, with3.5 meters equaling about 1 second. Measuring time in this manneris like using a clock that sometimes runs a bit too fast or too slow,but always runs forward.

If the whole system varied periodically with time, the speed ofthe clock, which is certainly a feature of the system, would have tovary periodically, with the same period, and the variations of thesystem, as measured by this clock, would likewise be periodic.Failure to vary periodically with downslope distance thereforeimplies absence of periodicity in real time.

Next, like the pinball machine, the slope with the board coastingdown it is not a compact system, since X must increase withoutlimit, and Y may do likewise. A simple change of variables,however, will produce a system that is mathematically compact.Note that it is possible to change the variables of a tangible systemwithout altering the system itself. We would do this, for example,if we decided to express a velocity in terms of speed and directioninstead of southward and eastward components.

In the present case, we first partition the slope into 5-meter by 4-meter rectangles. Each rectangle contains an entire light square ofthe checkerboard and extends halfway through the dark squares tothe west and east, so that the center of a pit lies at the center of eachrectangle, and centers of moguls lie at the midpoints of the west andeast sides. Then, using lower-case instead of upper-case names, welet x and y, the new variables, be the southward and eastwarddistances of the board from the center of the rectangle that itcurrently occupies. Most of the time x will increase continuously,just as X does, but x will jump from 2.5 back to 2.5 whenever the

A JOURNEY INTO CHAOS 35

board enters the next rectangle down the slope, at which time y willabruptly increase or decrease by 2 meters, according to whether the

36 A JOURNEY INTO CHAOS

Figure 9. The path of a single board traveling 600 meters down the modelslope. Note that the north-south scale has been compressed, as if we werelooking at the slope from beyond the base.

board has left the previous rectangle nearer to its southwest or itssoutheast corner. Also, y will jump from 2.0 to 2.0, or from 2.0to 2.0, when the board crosses one side of a rectangle or the other,but U and V will retain their original meanings as components ofvelocity. Since the slope has the same shape in each rectangle, thevalues of x, y, U, and V at any future time can be determined fromthe present values without knowing X and Y, i.e., without knowingwhich pit the board is near. Since, like U and V, x and y vary onlyover limited ranges, the new system is compact.

We have therefore identified a compact mathematical system inwhich U and V can vary exactly as they do in Figure 9. In this figure,positive and negative values of V show up respectively assoutheastward and southwestward progressions. Since these do notappear to alternate in any regular sequence, at least one variable,V, is patently not varying periodically, and we can conclude fairlysafely, even without reference to Figure 7 or 8, that the behavior ofthe board is chaotic. The only alternative would be the unlikely onethat it is periodic with a period exceeding 600 meters in X, or aboutthree minutestoo long to be revealed by the figure.

The path down the slope in Figure 9 is actually a graph of Yagainst X; with a quarter turn it will look like a conventional graph.In effect the board draws its own graph as it slides down the slope.Many of the fundamental concepts in dynamical-systems theory canbe represented by graphs, but these graphs are not always plots ofvalues of a single variable against values of one other. Often we areinterested in the simultaneous values of several variables, and wemay wish that we could have graphs in more than two dimensions,or even in as many dimensions as the number of variables in oursystem. If this number is large, it will be utterly impossible toconstruct some of the desired graphs, but, if it is fairly small,constructions in two or three dimensions can sometimes fulfill ourneeds.

As dynamical systems go, the ski slope is fairly simple, havingonly four variables; compare even a crude model of the flappingflag. Nevertheless, most persons do not find a curve or any otherkind of graph in four-dimensional space particularly easy tovisualize.

For example, consider a length of cord, perhaps a clothesline,with an overhand knot in it, and with its ends fastened to oppositewalls of a room. In the three-dimensional world in which we live,we cannot remove the knot without detaching one end or mutilating

A JOURNEY INTO CHAOS 37

the cord. Knot theorists have shown that in four dimensions asimilar knot can be readily removed while the ends remain fastened;effectively it is not a knot at all. This finding does not seemunreasonable, but visually I do not find it at all obvious.

It would be convenient to work with a system with fewer thanfour variables. We can produce such a system by doctoring up ourski-slope model.

Physical complexity and mathematical complexity are not thesame thing. It is quite possible to replace one system by another thatis physically more complicated but mathematically simpler. In ournew model we shall retain the original ski slope while replacing theboard by a sled. The sled will be equipped with brakes but nosteering mechanism. The driver is supposed to apply the brakes tojust the extent needed to hold the downslope speed, U, constant,while the cross-slope speed, V, may continue to vary. If the mogulsare rather high, a sled sliding over a mogul will slow down even ifthe brake is completely released, so in this case the sled must alsobe equipped with a motor, and the driver must use the acceleratorwhen the need arises.

The driver, who is unlikely to appreciate his lack of control overthe sled's direction and may not find much comfort in theknowledge that his ride may be chaotic, will probably needconsiderable practice before he can hold the southward speed evenapproximately constant. He may well opt for some specialelectronic equipment, which will observe the slope just ahead andapply the brakes or accelerator accordingly. Certainly a properlyworking sled of this sort would be physically much morecomplicated than a simple board. Designing a laboratory model ofthe sled would involve similar complications, and might requiresimilar electronic equipment. In contrast, a computer program forthe new system will be simpler than the original program, but thebig advantage will not be the modest saving in computer time butthe fact that one original variable, the southward speed U, hasbecome a constant, still called U, so that the new system has onlythree variables. These may be chosen as X, Y, and V, or instead asx, y, and V. Moreover, X, the downslope distance, now increasesuniformly with time, and the clock that used to run too fast or tooslow now works perfectly

As with the original system, we can expect that not allcombinations of values of the constants will lead to chaos. Althoughour constants now include U, the downslope speed, they no longer

38 A JOURNEY INTO CHAOS

include the coefficient of friction, whose value is controlled by thedriver when he brakes, and hence mathematically is determined byX, Y, and V. Use of the accelerator will appear in the program asnegative friction.

For our example we shall let U be 3.5 meters per seconda valueclose to the average value of U in the original examplewhile theremaining constants will retain their former values. Figure 10 showswhat happens when seven sleds, spaced at 10-centimeter intervalsalong the same west-east starting line, start with equal velocities.Although the details differ from those in Figure 7, the message isthe same; the system is chaotic.

The Heart of Chaos

An expression that has joined chaos in working its way into thescientific vocabulary, and that has aroused a fair amount of popularinterest as well, is strange attractor. Let us see what is meant byan attractor, and what one must be like to be considered strange.

If we take a look at some real-world phenomenon that has caughtour attention, we are likely to find that certain conceivable modesof behavior simply do not occur. A pendulum in a clock in goodworking order will not swing gently at times and violently at others;every swing will look like every other one. A flag in a steady breezewill never hang limp, nor will it extend itself directly into the breeze,no matter how long we wait. Subfreezing temperatures will notoccur in Honolulu, nor will relative humidities of 15 percent. Thestates of any system that do occur again and again, or areapproximated again and again, more and more closely, thereforebelong to a rather restricted set. This is the set of attractors.

When we perform a numerical experiment with a mathematicalmodel, the same situation arises. We are free to choose anymeaningful numbers as initial values of the variables, but after awhile certain numbers or combinations of numbers may fail toappear. For the sled on the slope, with the value of U, the downslopespeed, fixed at 3.5 meters per second, we can choose any initialvalue for V, the cross-slope speed, and any values for x and y withinrestricted ranges. Computations show that V will soon becomerestricted also, remaining between 5.0 and +5.0 meters per second.

Moreover, even the values of V that continue to occur will notdo so in combination with certain values of the other variables. Thesled will frequently slide almost directly over a mogul, and it will

A JOURNEY INTO CHAOS 39

often move almost directly from the northwest or northeast, butwhenever it is crossing a mogul it will be moving only from almostdue north. To the computer, this means that, if x is close to 0.0 and

40 A JOURNEY INTO CHAOS

Figure 10. The paths of seven sleds, starting with identical velocities frompoints spaced at 10-centimeter intervals along a west-east line.

y is close to 2.0 or +2.0, V will prove to be close to 0.0. Again, thestates that do manage to occur, after the disappearance of anytransient effects that may have been introduced by the choice ofinitial conditions, will form the set of attractors.

The interest in attractors that has accompanied and perhapsstimulated the recent surge of interest in chaos has been partly dueto the striking appearance of certain onesthe strange ones. Cana set of attractors, which is simply a collection of states, have anyappearance at all, let alone a strange one, particularly if thevariables happen to be invisible quantities, like atmospherictemperature and pressure? It can indeed, if by an attractor we meana graphical representation of an attractor.

As we saw earlier, we may sometimes wish that we could drawgraphs or other diagrams in a space that has as many dimensionsas the number of variables in our system. Often such a task isimpossible, but even then the concept of these diagrams can beuseful. The hypothetical multidimensional space in which such adiagram would have to be drawn is known as phase space.

In phase space, each point represents a particular state of adynamical system. The coordinates of the pointdistances inmutually perpendicular directions from some reference point, calledthe originare numerically equal to the values that the variablesassume when the state occurs. A particular solution of the equationsof a systemthat is, the set of states following and perhapspreceding a particular initial stateis represented by a curve, oftencalled an orbit, if the system is a flow. It is represented by a chainof points, also called an orbit, if the system is a mapping. Anattractor may be represented by a simple or complicatedgeometrical structure.

In the minds of many investigators, an attractor and its pictorialrepresentation in phase space are one and the same thing. In theirterminology, a point means a state, and an orbit means achronological sequence of states, so that a set of attractors can bea collection of points. When this collection consists of a singleagglomeration, it is also the attractor. When it is composed ofseveral unconnected pieces, and when no orbit passes from onepiece to another, each piece is a separate attractor.

An attractor can sometimes be a single point. A pendulum in aclock that is not wound up will eventually come to rest, hangingvertically, regardless of how it has initially been set in motion. Sincethe state of the pendulum, before it comes to rest, can be described

A JOURNEY INTO CHAOS 41

by two variables, the phase space of the pendulum must be two-dimensional; it must be a plane. On this plane, we may choose thehorizontal distance of a point from the origin to equal the horizontaldisplacement of the bob from the low point of the swing. Thevertical distance from the origin will not represent the bobs verticaldistance from anywhere; instead we may choose it to have the samenumerical value as the speed of the bobpositive when the bob ismoving to the right and negative when it moves to the left. Theattractor of the unwound clock will then be a single point on theplanethe originrepresenting the state of rest.

The attractor of a pendulum in a clock that is always kept woundwill be a closed curve, resembling an ellipse. In fact, if we measurethe speed of the bob in suitable units, perhaps miles per hour orperhaps centimeters per second, it can resemble a circle, with itscenter at the origin. Regardless of how rapidly we start thependulum swinging, it will soon acquire its normal behavior, andthen, as it swings toward the right, the point representing its statewill move to the right along the upper or positive half of the circle,crossing the top when the bob attains its greatest speed. As thependulum swings back toward the left, the point will move backalong the lower or negative half of the circle, after which it willcontinue its clockwise circuits.

Because, in this and many other systems, extremely large valuesof the variables cannot occur, except as transient conditions, pointsin the set of attractors cannot be too far removed from the origin.This means that they will occupy a rather restricted central region.They will indeed form the heart of the dynamical system.

For the sled on the slope, it is convenient to use x, y, and V ascoordinates in the three-dimensional phase space. The centralregion, containing the attractor, will then fit into a rectangular boxa box within which x extends only from 2.5 to +2.5 and yextends only from 2.0 to +2.0, since by definition x and y arelimited to these ranges, while V extends only from 5.0 to +5.0,since larger cross-slope speeds do not occur, except transiently.

There are numerous sophisticated computer programs forconstructing pictures of three-dimensional objects, although someof them may not work very well when the object is a complicatedattractor. For our system, let us adopt the simple procedure ofdisplaying cross sections of the box. The simplest of these sectionsare rectangles, parallel to one face or another. Mathematically, it iseasiest to work with rectangles on which x is constant, and on which

42 A JOURNEY INTO CHAOS

V is plotted against y. Horizontal and vertical distances from acentral point of a rectangle can then equal values of eastwarddistance and eastward speed, respectively, just as they do in thephase space of the pendulum when the clock faces south. We maythen draw a picture of the cross section of the attractor bydetermining, and marking on a chosen rectangle, the locations of alarge number of points that lie not only on the rectangle but alsoon the attractor. By looking at several parallel cross sections of th