24.06.2015 Views

ilya_prigogine_isabelle_stengers_alvin_tofflerbookfi-org

ilya_prigogine_isabelle_stengers_alvin_tofflerbookfi-org

ilya_prigogine_isabelle_stengers_alvin_tofflerbookfi-org

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

I<br />

I<br />

I<br />

I


ORDER OUT OF CHAOS


ORDER OUT OF CHAOS:<br />

MAN'S NEW DIALOGUE WITH NATURE<br />

A Bantam Book I April 1984<br />

New Age and the accompanying figure design as well as the<br />

statement "a search for meaning, growth and change" are<br />

trademarks of Bantam Books, Inc.<br />

All rights reserved.<br />

Copyright © 1984 by llya Prigogine and Isabelle Stengers.<br />

The foreword "Science and Change" copyright© 1984 by<br />

Alvin Tofjler.<br />

Book design by Barbara N. Cohen<br />

This book may not be reproduced in whole or in part, by<br />

mimeograph or tiny other means, without permission.<br />

For information address: Bantam Books, Inc.<br />

Library of Congress Cataloging in Publication Data<br />

Prigogine. I. (llya)<br />

Order out of chaos.<br />

1<br />

Based on the authors' Ia nouvelle alliance.<br />

Includes bibliographical references and index.<br />

I. Science-Philosophy. 2. Physics-Philosophy.<br />

3. Thermodynamics. 4. Irreversible processes.<br />

I. Stengers, Isabelle. II. Prigogine, I. (Ilya)<br />

La nouvelle alliance. Ill. Title.<br />

QI75.P8823 1984 50 83-21 403<br />

ISBN 0-553-34082-4<br />

Published simultaneously in the United States and Canada<br />

Bantam Books are published by Bantam Books.Inc.lts trademark.<br />

consisting of the words "Bantam Books" and the portrayal<br />

of a rooster, is Registered in the United States Patent<br />

and Trademark Office and in other countries. Marca Registrada.<br />

Bantam Books, Inc., 666 Fifth Avenue, New York, New<br />

York 10103.<br />

PRINTED IN THE UNITED STATES OF AMERICA<br />

FG 0987654321


ORDER OUT OF CHAOS<br />

MAN'S NEW DIALOGUE<br />

WITH NATURE<br />

llya Prigogine<br />

and<br />

Isabelle Stengers<br />

Foreword by<br />

Alvin Toffler<br />

BANTAM BOOKS<br />

TORONTO· NEW YORK • LONDON • SYDNEY


This book is dedicated to the memory of<br />

Erich Jantsch<br />

Aharon Katchalsky<br />

Pierre Resibois<br />

Leon Rosenfeld


TABLE OF CONTENTS<br />

FOREWORD: Science and Change by Alvin Toffier xi<br />

PREFACE: Man's New Dialogue with Nature<br />

xxvii<br />

INTRODUCTION: The Challenge to Science 1<br />

Book One: The Delusion of the Universal<br />

CHAPTER I: The Triumph of Reason 27<br />

1. The New Moses 27<br />

2. A Dehumanized World 30<br />

3. The Newtonian Synthesis 37<br />

4. The Experimental Dialogue 41<br />

5. The Myth at the Origin of Science 44<br />

6. The Limits of Classical Science 51<br />

CHAPTER n: The Identification of the Real 57<br />

1. Newton's Laws 57<br />

2. Motion and Change 62<br />

3. The Language of Dynamics 68<br />

4. Laplace's Demon 75<br />

CHAPTER m: The 1Wo Cultures 79<br />

1. Diderot and the Discourse of the Living 79<br />

2. Kant's Critical Ratification 86<br />

3. A Philosophy of Nature? Hegel and Bergson 89<br />

4. Process and Reality: Whitehead 93<br />

5. "Ignoramus, lgnoramibus":<br />

The Positivist's Strain 96<br />

6. A New Start 98<br />

Book Two: The Science of Complexity<br />

CHAPTER IV: Energy and the Industrial Age 103


1. Heat, the Rival of Gravitation 1(}3<br />

2. The Principle of the Conservation of Energy 107<br />

3. Heat Engines and the Arrow of Time 111<br />

4. From Technology to Cosmology 115<br />

5. The Birth of Entropy 117<br />

6. Boltzmann's Order Principle 122<br />

7. Carnot and Darwin 127<br />

CHAPTER v: The Three Stages of Thermodynamics 131<br />

1. Flux and Force 131<br />

2. Linear Thermodynamics 137<br />

3. Far from Equilibrium 140<br />

4. Beyond the Threshold of Chemical Instability 146<br />

5. The Encounter with Molecular Biology 153<br />

6. Bifurcations and Symmetry-Breaking 160<br />

7. Cascading Bifurcations and<br />

the Transitions to Chaos 167<br />

8. From Euclid to Aristotle 171<br />

CHAPTER v1: Order Through Fluctuations 177<br />

1. Fluctuations and Chemistry 177<br />

2. Fluctuations and Correlations 179<br />

3. The Amplification of Fluctuations 181<br />

4. Structural Stability 189<br />

5. Logistic Evolution 192<br />

6. Evolutionary Feedback 196<br />

7. Modelizations of Complexity 203<br />

8. An Open World 207<br />

Book Three: From Being to Becoming<br />

CHAPTER vn: Rediscovering Time 213<br />

1. A Change of Emphasis 213<br />

2. The End of Universality 217<br />

3. The Rise of Quantum Mechanics 218<br />

- 4. Heisenberg's Uncertainty Relation 222<br />

5. The Temporal Evolution of Quantum Systems 226<br />

6. A Nonequilibrium Universe 229


CHAPTER vm: The Clash of Doctrines 233<br />

1. Probability and Irreversibility 233<br />

2. Boltzmann's Breakthrough 240<br />

3. Questioning Boltzmann's Interpretation 243<br />

4. Dynamics and Thermodynamics: 1Wo Separate<br />

Worlds 247<br />

5. Boltzmann and the Arrow of Time 253<br />

CHAPTER IX: Irreversibility-the Entropy Barrier 257<br />

1. Entropy and the Arrow of Time 257<br />

2. Irreversibility as a Symmetry-Breaking Process 260<br />

3. The Limits of Classical Concepts 261<br />

4. The Renewal of Dynamics 264<br />

5. From Randomness to Irreversibility 272<br />

6. The Entropy Barrier 277<br />

7. The Dynamics of Correlations 280<br />

8. Entropy as a Selection Principle 285<br />

9. Active Matter 286<br />

coNcLusioNs: From Earth to Heaventhe<br />

Reenchantment of Nature 291<br />

1. An Open Science 291<br />

2. Time and Times 293<br />

3. The Entropy Barrier 295<br />

4. The Evolutionary Paradigm 297<br />

5. Actors and Spectators 298<br />

6. A Whirlwind in a Tu rbulent Nature 301<br />

7. Beyond Tautology 305<br />

8. The Creative Course of Time 307<br />

9. The Human Condition 311<br />

10. The Renewal of Nature 312<br />

NOTES 315<br />

INDEX 335


FOREWORD<br />

SCIENCE AND<br />

CHANGE<br />

by Alvin Toffler<br />

One of the most highly developed skills in contemporary We stern<br />

civilization is dissection: the split-up of problems into their<br />

smallest possible components. We are good at it. So good, we<br />

often f<strong>org</strong>et to put the pieces back together again.<br />

This skill is perhaps most finely honed in science. There we<br />

not only routinely break problems down into bite-sized chunks<br />

and mini-chunks, we then very often isolate each one from its<br />

environment by means of a useful trick. We say ceteris paribus-all<br />

other things being equal. In this way we can ignore<br />

the complex interactions between our problem and the rest of<br />

the universe.<br />

llya Prigogine, who won the Nobel Prize in 1977 for his<br />

work on the thermodynamics of nonequilibrium systems, is<br />

net satisfied, however, with merely taking things apart. He has<br />

spent the better part of a lifetime trying to "put the pieces<br />

back together again"-the pieces in this case being biology<br />

and physics, necessity and chance, science and humanity.<br />

Born in Russia in 1917 and raised in Belgium since the age<br />

of ten, Prigogine is a compact man with gray hair, cleanly chiseled<br />

features, and a laserlike intensity. Deeply interested in<br />

archaeology, art, and history, he brings to science a remarkable<br />

polymathic mind. He lives with his engineer-wife, Marina,<br />

and his son, Pascal, in Brussels, where a crossdisciplinary<br />

team is busy exploring the implications of his<br />

ideas in fields as disparate as the social behavior of ant colonies,<br />

diffusion reactions in chemical systems and dissipative<br />

processes in quantum field theory.<br />

He spends part of each year at the Ilya Prigogine Center for<br />

Statistical Mechanics and Thermodynamics of the University<br />

of Texas in Austin. To his evident delight and surprise, he was<br />

xi


ORDER OUT OF CHAOS<br />

xii<br />

awarded the Nobel· Prize for his work on "dissipative structures"<br />

arising out of nonlinear processes in nonequilibrium<br />

systems. The coauthor of this volume, Isabelle Stengers, is a<br />

philosopher, chemist, and historian of science who served for<br />

a time as part of Prigogine 's Brussels team. She now lives in<br />

Paris and is associated with the Musee de Ia Villette.<br />

In Order Out of Chaos they have given us a landmark-a<br />

work that is contentious and mind-energizing, a book filled<br />

with flashing insights that subvert many of our most basic assumptions<br />

and suggest fresh ways to think about them.<br />

Under the title La nouvelle alliance, its appearance in<br />

France in 1979 triggered a marvelous scientific free-for-all<br />

among prestigious intellectuals in fields as diverse as entomology<br />

and literary criticism.<br />

It is a measure of America's insularity and cultural arrogance<br />

that this book, which is either published or about to<br />

be published in twelve languages, has taken so long to cross<br />

the Atlantic. The delay carries with it a silver lining, however,<br />

in that this edition includes Prigogine 's newest findings, particularly<br />

with respect to the Second Law of thermodynamics,<br />

which he sets into a fresh perspective.<br />

For all these reasons, Order Out of Chaos is more than just<br />

another book: It is a lever for changing science itself, for compelling<br />

us to reexamine its goals, its methods, its epistemology-its<br />

world view. Indeed, this book can serve as a symbol<br />

of today's historic transformation in science-one that no informed<br />

person can afford to ignore.<br />

Some scholars picture science as driven by its own internal<br />

logic, developing according to its own laws in splendid isolation<br />

from the world around it. Yet many scientific hypotheses,<br />

theories, metaphors, and models (not to mention the choices<br />

made by scientists either to study or to ignore various problems)<br />

are shaped by economic, cultural, and political forces<br />

· operating outside the laboratory.<br />

I do not mean to suggest too neat a parallel between the nature<br />

of society and the reigning scientific world view or "paradigm."<br />

Still less would I relegate science to some "superstructure , .<br />

mounted atop a socioeconomic "base," as Marxists are wont to<br />

do. But science is not an "independent variable.'' It is an open<br />

system embedded in society and linked to it by very dense feedback<br />

loops. It is powerfully influenced by its external environ-


xiii<br />

FOREWORD: SCIENCE AND CHANGE<br />

ment, and, in a general way, its development is shaped by cultural<br />

receptivity to its dominant ideas.<br />

Take that body of ideas that came together in the seventeenth<br />

and eighteenth centuries under the heading of "classical<br />

science" or "Newtonianism." They pictured a world in<br />

which every event was determined by initial conditions that<br />

were, at least in principle, determinable with precision. It was<br />

a world in which chance played no part, in which all the pieces<br />

came together like cogs in a cosmic machine.<br />

The acceptance of this mechanistic view coincided with the<br />

rise of a factory civilization. And divine dice-shooting seems<br />

hardly enough to account for the fact that the Age of the Machine<br />

enthusiastically embraced scientific theories that pictured<br />

the entire universe as a machine.<br />

This view of the world led Laplace to his famous claim that,<br />

given enough facts, we could not merely predict the future but<br />

retrodict the past. And this image of a simple, uniform; mechanical<br />

universe not only shaped the development of science,<br />

it also spilled over into many other fields. It influenced the<br />

framers of the American Constitution to create a machine for<br />

governing, its checks and balances clicking like parts of a<br />

clock. Metternich, when he rode forth to create his balance of<br />

power in Europe, carried a copy of Laplace's writings in his<br />

baggage. And the dramatic spread of factory civilization, with<br />

its vast clanking machines, its heroic engineering breakthroughs,<br />

the rise of the railroad, and new industries such as<br />

steel, textile, and auto, seemed merely to confirm the image of<br />

the universe as an engineer's Tinkertoy.<br />

Today, however, the Age of the Machine is screeching to a<br />

halt, if ages can screech-and ours certainly seems to. And<br />

the decline of the industrial age forces us to confront the painful<br />

limitations of the machine model of reality.<br />

Of course, most of these limitations are not freshly discovered.<br />

The notion that the world is a clockwork, the planets<br />

timelessly orbiting, all systems operating deterministically in<br />

equilibrium, all subject to universal laws that an outside observer<br />

could discover-this model has come under withering<br />

fire ever since it first arose.<br />

In the early nineteenth century, thermodynamics challenged<br />

the timelessness implied in the mechanistic image of<br />

the universe. If the world was a big machine, the thermodynamicists<br />

declared, it was running down, its useful en-


ORDER OUT OF CHAOS<br />

xiv<br />

ergy leaking out. It could not go on forever, and time,<br />

therefore, took on a new meaning. Darwin's followers soon<br />

introduced a contradictory thought: The world-machine might<br />

be running down, losing energy and <strong>org</strong>anization, but biological<br />

systems, at least, were running up, becoming more, not<br />

less, <strong>org</strong>anized.<br />

By the early twentieth century, Einstein had come along to<br />

put the observer back into the system: The machine looked<br />

diffe rent-indeed, for all practical purposes it was differentdepending<br />

upon where you stood within it. But it was still a<br />

deterministic machine, and God did not throw dice. Next, the<br />

quantum people and the uncertainty folks attacked the model<br />

with pickaxes, sledgehammers, and sticks of dynamite.<br />

Nevertheless, despite all the ifs, ands, and buts, it remains<br />

fair to say, as Prigogine and Stengers do, that the machine paradigm<br />

is still the "reference point" for physics and the core<br />

model of science in general. Indeed, so powerful is its continuing<br />

influence that much of social science, and especially<br />

economics, remains under its spell.<br />

The importance of this book is not simply that it uses original<br />

arguments to challenge the Newtonian model, but also<br />

that it shows how the still valid, though much limited, claims<br />

of Newtonianism might fit compatibly into a larger scientific<br />

image of reality. It argues that the old "universal laws" are not<br />

universal at all, but apply only to local regions of reality. And<br />

these happen to be the regions to which science has devoted<br />

the most effort.<br />

Thus, in broad-stroke terms, Prigogine and Stengers argue<br />

that traditional science in the Age of the Machine tended to<br />

emphasize stability, order, uniformity, and equilibrium. It concerned<br />

itself mostly with closed systems and linear relationships<br />

in which small inputs uniformly yield small results.<br />

With the transition from an industrial society based on<br />

heavy inputs of energy, capital, and labor to a high-technology<br />

society in which information and innovation are the critical<br />

resources, it is not surprising that new scientific world models<br />

should appear.<br />

What makes the Prigoginian paradigm especially interesting<br />

is that it shifts attention to those aspects of reality that characterize<br />

today's accelerated social change: disorder, instability,<br />

diversity, disequilibrium, nonlinear relationships (in which


xv<br />

FOREWORD: SCIENCE AND CHANGE<br />

small inputs can trigger massive consequences), and temporality-a<br />

heightened sensitivity to the flows of time.<br />

The work of Ilya Prigogine and his colleagues in the socalled<br />

"Brussels school" may well represent the next revolution<br />

in science as it enters into a new dialogue not merely with<br />

nature, but with society itself.<br />

The ideas of the Brussels school , based heavily on Prigogine's<br />

work, add up to a novel, comprehensive theory of<br />

change.<br />

Summed up and simplified, they hold that while some parts of<br />

the universe may operate like machines, these are closed systems,<br />

and closed systems, at best, form only a small part of the physical<br />

universe. Most phenomena of interest to us are, in fact, open<br />

systems, exchanging energy or matter (and, one might add, information)<br />

with their environment. Surely biological and social systems<br />

are open, which means that the attempt to understand them<br />

in mechanistic terms is doomed to failure.<br />

This suggests, moreover, that most of reality, instead of<br />

being orderly, stable, and equilibria!, is seething and bubbling<br />

with change, disorder, and process.<br />

In Prigoginian terms, all systems contain subsystems,<br />

which are continually "fluctuating." At times, a single fluctuation<br />

or a combination of them may become so powerful, as a<br />

result of positive feedback, that it shatters the preexisting <strong>org</strong>anization.<br />

At this revolutionary moment-the authors call it<br />

a "singular moment" or a "bifurcation point"-it is inherently<br />

impossible to determine in advance which direction change<br />

will take: whether the system will disintegrate into "chaos" or<br />

leap to a new, more differentiated, higher level of "order" or<br />

<strong>org</strong>anization, which they call a "dissipative structure." (Such<br />

physical or chemical structures are termed dissipative because,<br />

compared with the simpler structures they replace, they<br />

require more energy to sustain them.)<br />

One of the key controversies surrounding this concept has<br />

to do with Prigogine's insistence that order and <strong>org</strong>anization<br />

can actually arise "spontaneously" out of disorder and chaos<br />

through a process of "self-<strong>org</strong>anization."<br />

To grasp this extremely powerful idea. we first need to make<br />

a distinction between systems that are in "equilibrium," sys-


ORDER OUT OF CHAOS<br />

xvi<br />

terns that are "near equilibrium," and systems that are "far<br />

from equilibrium."<br />

Imagine a primitive tribe. If its birthrate and death rate are<br />

equal, the size of the population remains stable. Assuming adequate<br />

food and other resources, the tribe forms part of a local<br />

system in ecological equilibrium.<br />

Now increase the birthrate. A few additional births (without<br />

an equivalent number of deaths) might have little effect. The<br />

system may move to a near-equilibria! state. Nothing much<br />

happens. It takes a big jolt to produce big consequences in<br />

systems that are in equilibria] or near-equilibria] states.<br />

But if the birthrate should suddenly soar, the system is<br />

pushed into a far-from-equilibrium condition, and here nonlinear<br />

relationships prevail. In this state, systems do strange<br />

things. They become inordinately sensitive to external influences.<br />

Small inputs yield huge, startling effects. The entire<br />

system may re<strong>org</strong>anize itself in ways that strike us as bizarre.<br />

Examples of such self-re<strong>org</strong>anization abound in Order Out<br />

of Chaos. Heat moving evenly through a liquid suddenly, at a<br />

certain threshold, converts into a convection current that radically<br />

re<strong>org</strong>anizes the liquid, and millions of molecules, as if on<br />

cue, suddenly form themselves into hexagonal cells.<br />

Even more spectacular are the "chemical clocks" described<br />

by Prigogine and Stengers. Imagine a million white ping-pong<br />

balls mixed at random with a million black ones, bouncing<br />

around chaotically in a tank with a glass window in it. Most of<br />

the time, the mass seen through the window would appear to<br />

be gray, but now and then, at irregular moments, the sample<br />

seen through the glass might seem black or white, depending<br />

on the distribution of the balls at that moment in the vicinity of<br />

the window.<br />

Now imagine that suddenly the window goes all white, then<br />

all black, then all white again, and on and on, changing its<br />

color completely at fixed intervals-like a clock ticking.<br />

Why do all the white balls and all the black ones suddenly<br />

<strong>org</strong>anize themselves to change color in time with one another?<br />

By all the traditional rules, this should not happen at all. Yet, if<br />

we leave ping-pong behind and look at molecules in certain chemical<br />

reactions, we find that precisely such a self-<strong>org</strong>anization or<br />

ordering can and does occur--despite what classical physics and<br />

the probability theories of Boltzmann tell us.<br />

In far-from-equilibrium situations other seemingly spon-


xvii<br />

FOREWORD: SCIENCE AND CHANGE<br />

taneous, often dramatic re<strong>org</strong>anizations of matter within time<br />

and space also take place. And if we begin thinking in terms of<br />

two or three dimensions, the number and variety of such pos·<br />

sible structures become very great.<br />

Now add to this an additional discovery. Imagine a situation<br />

in which a chemical or other reaction produces an enzyme<br />

whose presence then encourages further production of the<br />

same enzyme. This is an example of what computer scientists<br />

would call a positive-feedback loop. In chemistry it is called<br />

"auto-catalysis." Such situations are rare in in<strong>org</strong>anic chemis·<br />

try. But in recent decades the molecular biologists have found<br />

that such loops (along with inhibitory or "negative" feedback<br />

and more complicated "cross-catalytic" processes) are the<br />

very stuff of life itself. Such processes help explain how we go<br />

from little lumps of DNA to complex living <strong>org</strong>anisms.<br />

More generally, therefore, in far-from-equilibrium conditions<br />

we find that very small perturbations or fluctuations can<br />

become amplified into gigantic, structure-breaking waves.<br />

And this sheds light on all sorts of "qualitative" or "revolu·<br />

tionary" change processes. When one combines the new insights<br />

gained from studying far-from-equilibrium states and<br />

nonlinear processes, along with these complicated feedback<br />

systems, a whole new approach is opened that makes it possible<br />

to relate the so-called hard sciences to the softer sciences<br />

of life-and perhaps even to social processes as well.<br />

(Such findings have at least analogical significance for social,<br />

economic or political realities. Words like "revolution,"<br />

"economic crash," "technological upheaval ," and "paradigm<br />

shift" all take on new shades of meaning when we begin thinking<br />

of them in terms of fluctuations, feedback amplification,<br />

dissipative structures, bifurcations, and the rest of the Prigoginian<br />

conceptual vocabulary.) It is these panoramic vistas that<br />

are opened to us by Order Out of Chaos.<br />

Beyond this, there is the even more puzzling, pervasive is·<br />

sue of time.<br />

Part of today's vast revolution in both science and culture is<br />

a reconsideration of time, and it is important enough to merit a<br />

brief digression here before returning to Prigogine's role in it.<br />

Take history, for example. One of the great contributions to<br />

historiography has been Braudel's division of time into three<br />

scales-" geographical time," in which events occur over the


ORDER OUT OF CHAOS<br />

xviii<br />

course of aeons; the much shorter "social time" scale by<br />

which economies, states, and civilizations are measured; and<br />

the even shorter scale of "individual time"-the history of<br />

human events.<br />

In social science, time remains a largely unmapped terrain.<br />

Anthropology has taught us that cultures differ sharply in the<br />

way they conceive of time. For some, time is cyclical-history<br />

endlessly recurrent. For other cultures, our own included,<br />

time is a highway stretched between past and future, and people<br />

or whole societies march along it. In still other cultures,<br />

human lives are seen as stationary in time; the future advances<br />

toward us, instead of us toward it.<br />

Each society, as I've written elsewhere, betrays its own<br />

characteristic "time bias"-the degree to which it places emphasis<br />

on past, present, or future. One lives in the past. Another<br />

may be obsessed with the future.<br />

Moreover, each culture and each person tends to think in<br />

terms of "time horizons." Some of us think only of the immediate-the<br />

now. Politicians, for example, are often criticized<br />

for seeking only immediate, short-term results. Their time<br />

horizon is said to be influenced by the date of the next election.<br />

Others among us plan for the long term. These differing<br />

time horizons are an overlooked source of social and political<br />

friction-perhaps among the most important.<br />

But despite the growing recognition that cultural conceptions<br />

of time differ, the social sciences have developed little<br />

in the way of a coherent theory of time. Such a theory might<br />

reach across many disciplines, from politics to group dynamics<br />

and interpersonal psychology. It might, for example, take<br />

account of what, in Future Shock, I called "durational expectancies"-our<br />

culturally induced assumptions about how long<br />

certain processes are supposed to take.<br />

We learn very early, for example, that brushing one's teeth<br />

should last only a few minutes, not an entire morning, or that<br />

when Daddy leaves for work, he is likely to be gone approximately<br />

eight hours, or that a "mealtime" may last a few minutes<br />

or hours, but never a year. (Television, with its division of<br />

the day into fixed thirty- or sixty-minute intervals, subtly<br />

shapes our notions of duration. Thus we normally expect the<br />

hero in a melodrama to get the girl or find the money or win<br />

the war in the last five minutes. In the United States we expect


xix<br />

FOREWORD: SCIENCE AND CHANGE<br />

commercials to break in at certain intervals.) Our minds are<br />

filled with such durational assumptions. Those of children are<br />

much different from those of fully socialized adults, and here<br />

again the diffe rences are a source of conflict.<br />

Moreover, children in an industrial society are "time<br />

trained"-they learn to read the clock, and they learn to distinguish<br />

even quite small slices of time, as when their parents<br />

tell them, "You've only got three more minutes till bedtime!"<br />

These sharply honed temporal skills are often absent in<br />

slower-moving agrarian societies that require less precision in<br />

daily scheduling than our time-obsessed society.<br />

Such concepts, which fit within the social and individual<br />

time scales of Braude!, have never been systematically developed<br />

in the social sciences. Nor have they, in any significant<br />

way, been articulated with our scientific theories of time,<br />

even though they are necessarily connected with our assumptions<br />

about physical reality. And this brings us back to Prigogine,<br />

who has been fascinated by the concept of time since<br />

boyhood. He once said to me that, as a young student, he was<br />

struck by a grand contradiction in the way science viewed<br />

time, and this contradiction has been the source of his life's<br />

work ever since.<br />

In the world model constructed by Newton and his followers,<br />

time was an afterthought. A moment, whether in the<br />

present, past, or future, was assumed to be exactly like any<br />

other moment. The endless cycling of the planets-indeed,<br />

the operations of a clock or a simple machine-can, in principle,<br />

go either backward or forward in time without altering the<br />

basics of the system. For this reason, scientists refer to time in<br />

Newtonian systems as "reversible."<br />

In the nineteenth century, however, as the main focus of<br />

physics shifted from dynamics to thermodynamics and the<br />

Second Law of thermodynamics was proclaimed, time suddenly<br />

became a central concern. For, according to the Second<br />

Law, there is an inescapable loss of energy in the universe.<br />

And, if the world machine is really running down and approaching<br />

the heat death, then it follows that one moment is no<br />

longer exactly like the last. Yo u cannot run the universe backward<br />

to make up for entropy. Events over the long term cannot<br />

replay themselves. And this means that there is a directionality<br />

or, as Eddington later called it, an "arrow" in time.


ORDER OUT OF CHAOS<br />

xx<br />

The whole universe is, in fact, aging. And, in turn, if this is<br />

true, time is a one-way street. It is no longer reversible, but<br />

irreversible.<br />

In short, with the rise of thermodynamics, science split<br />

down the middle with respect to time. Worse yet, even those<br />

who saw time as irreversible soon also split into two camps.<br />

After all, as energy leaked out of the system, its ability to<br />

sustain <strong>org</strong>anized structures weakened, and these, in turn,<br />

broke down into less <strong>org</strong>anized, hence more random elements.<br />

But it is precisely <strong>org</strong>anization that gives any system<br />

internal diversity. Hence, as entropy drained the system of energy,<br />

it also reduced the differences in it. Thus the Second<br />

Law pointed toward an increasingly homogeneous-and, from<br />

the human point of view, pessimistic-future.<br />

Imagine the problems introduced by Darwin and his followers!<br />

For evolution, far from pointing toward reduced <strong>org</strong>anization<br />

and diversity, points in the opposite direction.<br />

Evolution proceeds from simple to complex, from "lower" to<br />

"higher" forms of life, from undifferentiated to differentiated<br />

structures. And, from a human point of view, all this is quite<br />

optimistic. The universe gets "better" <strong>org</strong>anized as it ages,<br />

continually advancing to a higher level as time sweeps by.<br />

In this sense, scientific views of time may be summed up as<br />

a contradiction within a contradiction.<br />

It is these paradoxes that Prigogine and Stengers set out to<br />

illuminate, asking, "What is the specific structure of dynamic<br />

systems which permits them to 'distinguish' between past and<br />

future? What is the minimum complexity involved?"<br />

The answer, for them, is that time makes its appearance<br />

with randomness: "Only when a system behaves in a sufficiently<br />

random way may the difference between past and future,<br />

and therefore irreversibility, enter its description."<br />

In classical or mechanistic science, events begin with "initial<br />

conditions," and their atoms or particles follow "world<br />

lines" or trajectories. These can be traced either backward<br />

into the past or forward into the future. This is just the opposite<br />

of certain chemical reactions, for example, in which two<br />

liquids poured into the same pot diffuse until the mixture is<br />

uniform or homogeneous. These liquids do not de-diffuse<br />

themselves. At each moment of time the mixture is different,<br />

the entire process is "time-oriented."<br />

For classical science, at least in its early stages, such pro-


xxi<br />

FOREWORD: SCIENCE AND CHANGE<br />

cesses were regarded as anomalies, peculiarities that arose<br />

from highly unlikely initial conditions.<br />

It is Prigogine and Stengers' thesis that such time-dependent,<br />

one-way processes are not merely aberrations or deviations<br />

from a world in which time is irreversible. If anything,<br />

the opposite might be true, and it is reversible time, associated<br />

with "closed systems" (if such, indeed, exist in reality), that<br />

may well be the rare or aberrant phenomenon.<br />

What is more, irreversible processes are the source of<br />

order-hence the title Order Out of Chaos. It is the processes<br />

associated with randomness, openness, that lead to higher levels<br />

of <strong>org</strong>anization, such as dissipative structures.<br />

Indeed, one of the key themes of this book is its striking<br />

reinterpretation of the Second Law of thermodynamics. For<br />

according to the authors, entropy is not merely a downward<br />

slide toward dis<strong>org</strong>anization. Under certain conditions, entropy<br />

itself becomes the progenitor of order.<br />

What the authors are proposing, therefore, is a vast synthesis<br />

that embraces both reversible and irreversible time, and<br />

shows how they relate to one another, not merely at the level of<br />

macroscopic phenomena, but at the most minute level as well.<br />

It is a breathtaking attempt at "putting the pieces back together<br />

again." The argument is complex, and at times beyond<br />

easy reach of the lay reader. But it flashes with fresh insight<br />

and suggests a coherent way to relate seemingly unconnected-even<br />

contradictory-philosophical concepts.<br />

Here we begin to glimpse, in full richness, the monumental<br />

synthesis proposed in these pages. By insisting that irreversible<br />

time is not a mere aberration, but a characteristic of much<br />

of the universe, they subvert classical dynamics. For Prigogine<br />

and Stengers, it is not a case of either/or. Of course,<br />

reversibility still applies (at least for sufficiently long times)­<br />

but in closed systems only. Irreversibility applies to the rest of<br />

the universe.<br />

Prigogine and Stengers also undermine conventional views<br />

of thermodynamics by showing that, under nonequilibrium<br />

conditions, at least, entropy may produce, rather than degrade,<br />

order, <strong>org</strong>anization-and therefore life.<br />

If this is so, then entropy, too, loses its either/or character.<br />

While certain systems run down, other systems simultaneously<br />

evolve and grow more coherent. This mutualistic,


ORDER OUT OF CHAOS<br />

xxii<br />

nonexclusive view makes it possible for biology and physics to<br />

coexist rather than merely contradict one another.<br />

Finally, yet another profound synthesis is implied-a new<br />

relationship between chance and necessity.<br />

The role of happenstance in the affairs of the universe has<br />

been debated, no doubt, since the first Paleolithic warrior accidently<br />

tripped over a rock. In the Old Te stament, God's will<br />

is sovereign, and He not only controls the orbiting planets but<br />

manipulates the will of each and every individual as He sees<br />

fit. As Prime Mover, all causality flows from Him, and all<br />

events in the universe are foreordained. Sanguinary conflicts<br />

raged over the precise meaning of predestination or free will,<br />

from the time of Augustine through the Carolingian quarrels.<br />

Wycliffe, Huss, Luther, C<strong>alvin</strong>-all contributed to the debate.<br />

No end of interpreters attempted to reconcile determinism<br />

with freedom of will. One ingenious view held that God did<br />

indeed determine the affairs of the universe, but that with respect<br />

to the free will of the individual, He never demanded a<br />

specific action. He merely preset the range of options available<br />

to the human decision-maker. Free will downstairs operated<br />

only within the limits of a menu determined upstairs.<br />

In the secular culture of the Machine Age, hard-line determinism<br />

has more or less held sway even after the challenges of<br />

Heisenberg and the "uncertaintists." Even today, thinkers<br />

such as Rene Thorn reject the idea of chance as illusory and<br />

inherently unscientific.<br />

Faced with such philosophical stonewalling, some defenders<br />

of free will, spontaneity, and ultimate uncertainty, especially<br />

the existentialists, have taken equally uncompromising stands.<br />

(For Sartre, the human being was "completely and always<br />

free," though even Sartre, in certain writings, recognized<br />

practical limitations on this freedom.)<br />

Two things seem to be happening to contemporary concepts<br />

of chance and determinism. To begin with, they are becoming<br />

more complex. As Edgar Morin, a leading French sociologistturned-epistemologist,<br />

has written:<br />

"Let us not f<strong>org</strong>et that the problem of determinism has<br />

changed over the course of a century .... In place of the idea<br />

of sovereign, anonymous, permanent laws directing all things<br />

in nature there has been substituted the idea of laws of interaction<br />

. ... There is more: the problem of determinism has be-


xxiii<br />

FOREWORD: SCIENCE AND CHANGE<br />

come that of the order of the universe. Order means that there<br />

are other things besides 'Jaws': that there are constraints, invariances,<br />

constancies, regularities in our universe . ... In<br />

place of the homogenizing and anonymous view of the old determinism,<br />

there has been substituted a diversifying and evolutive<br />

view of determinations."<br />

And as the concept of determinism has grown richer, new<br />

efforts have been made to recognize the co-presence of both<br />

chance and necessity, not with one subordinate to the· other,<br />

but as full partners in a universe that is simultaneously<br />

<strong>org</strong>anizing and de-<strong>org</strong>anizing itself.<br />

It is here that Prigogine and Stengers enter the arena. For they<br />

have taken the argument a step farther. They not only demonstrate<br />

(persuasively to me, though not to critics like the mathematician,<br />

Rene Thorn) that both determinism and chance operate, they also<br />

attempt to show how the two fit together.<br />

Thus, according to the theory of change implied in the idea<br />

of dissipative structures, when fluctuations force an existing<br />

system into a far-from-equilibrium condition and threaten its<br />

structure, it approaches a critical moment or bifurcation point.<br />

At this point, according to the authors, it is inherently impossible<br />

to determine in advance the next state of the system.<br />

Chance nudges what remains of the system down a new path<br />

of development. And once that path is chosen (from among<br />

many), determinism takes over again until the next bifurcation<br />

point is reached.<br />

Here, in short, we see chance and necessity not as irreconcilable<br />

opposites, but each playing its role as a partner in destiny.<br />

Yet another synthesis is achieved.<br />

When we bring reversible time and irreversible time, disorder<br />

and order, physics and biology, chance and necessity all<br />

into the same novel frame, and stipulate their interrelationships,<br />

we have made a grand statement-arguable, no doubt,<br />

but in this case both powerful and majestic.<br />

Yet this accounts only in part for the excitement occasioned<br />

by Order Out of Chaos. For this sweeping synthesis, as I have<br />

suggested, has strong social and even political overtones. Just<br />

as the Newtonian model gave rise to analogies in politics, diplomacy,<br />

and other spheres seemingly remote from science,<br />

so, too, does the Prigoginian model lend itself to analogical<br />

extension.


ORDER OUT OF CHAOS·<br />

xxfv<br />

By offering rigorous ways of modeling qualitative change,<br />

for example, they shed light on the concept of revolution. By<br />

explaining how successive instabilities give rise to transformatory<br />

change, they illuminate <strong>org</strong>anization theory. They throw a<br />

fresh light, as well, on certain psychological processes-innovation,<br />

for example, which the authors see as associated with<br />

"nonaverage" behavior of the kind that arises under nonequilibrium<br />

conditions.<br />

Even more significant, perhaps, are the implications for the<br />

study of collective behavior. Prigogine and Stengers caution<br />

against leaping to genetic or sociobiological explanations for<br />

puzzling social behavior. Many things that are attributed to<br />

biological pre-wiring are not produced by selfish, determinist<br />

genes, but rather by social interactions under nonequilibrium<br />

conditions.<br />

(In one recent study, for instance, ants were divided into<br />

two categories: One consisted of hard workers, the other of<br />

inactive or "lazy" ants. One might overhastily trace such<br />

traits to genetic predisposition. Yet the study found that if the<br />

system were shattered by separating the two groups from one<br />

another, each in turn developed its own subgroups of hard<br />

workers and idlers. A significant percentage of the "lazy" ants<br />

suddenly turned into hardworking Stakhanovites.)<br />

Not surprisingly, therefore, the ideas behind this remarkable<br />

book are beginning to be researched in economics, urban<br />

studies, human geography, ecology, and many other disciplines.<br />

No one-not even its authors-can appreciate the full implications<br />

of a work as crowded with ideas as Order Out of<br />

Chaos. Each reader will no doubt come away puzzled by some<br />

passages (a few are simply too technical for the reader without<br />

scientific training); startled or stimulated by others (as their<br />

implications strike home); occasionally skeptical; yet intellectually<br />

enriched by the whole. And if one measure of a book is<br />

the degree to which it generates good questions, this one is<br />

surely successful.<br />

Here are just a couple that have haunted me.<br />

How, outside a laboratory, might one define a .. fluctuation"?<br />

What, in Prigoginian terms, does one mean by ••cause"<br />

or "effect"? And when the authors speak of molecules communicating<br />

with one another to achieve coherent, synchro-


xxv<br />

FOREWORD: SCIENCE AND CHANGE<br />

nized change, one may assume they are not anthropomorphiz·<br />

ing. But they raise for me a host of intriguing issues about<br />

whether all parts of the environment are signaling all the time,<br />

or only intermittently; about the indirect, second, and nth<br />

order communication that takes place, permitting a molecule<br />

or an <strong>org</strong>anism to respond to signals which it cannot sense for<br />

lack of the necessary receptors. (A signal sent by the environment<br />

that is undetectable by A may be received by B and converted<br />

into a different kind of signal that A is properly<br />

equipped to receive-so that B serves as a relay/converter,<br />

and A responds to an environmental change that has been signaled<br />

to it via second-order communication.)<br />

In connection with time, what do the authors make of the<br />

idea put forward by Harvard astronomer David Layzer, that<br />

we might conceive of three distinct "arrows of time"-one<br />

based on the continued expansion of the universe since the Big<br />

Bang; one based on entropy; and one based on biological and<br />

historical evolution?<br />

Another question: How revolutionary was the Newtonian<br />

revolution? Taking issue with some historians, Prigogine and<br />

Stengers point out the continuity of Newton's ideas with alchemy<br />

and religious notions of even earlier vintage . Some<br />

readers might conclude from this that the rise of Newtonianism<br />

was neither abrupt nor revolutionary. Yet, to my mind,<br />

the Newtonian breakthrough should not be seen as a linear<br />

outgrowth of these earlier ideas. Indeed, it seems to me that<br />

the theory of change developed in Order Out of Chaos argues<br />

against just such a "continuist" view.<br />

Even if Newtonianism was derivative, this doesn't mean<br />

that the intc;:rnal structure of the Newtonian world-model was<br />

actually the same or that it stood in the same relationship to its<br />

external environment.<br />

The Newtonian system arose at a time when feudalism in<br />

Western Europe was crumbling-when the social system was,<br />

so to speak, far from equilibrium. The model of the universe<br />

proposed by the classical scientists (even if partially derivative)<br />

was applied analogously to new fields and disseminated<br />

successfully, not just because of its scientific power or "rightness,"<br />

but also because an emergent industrial society based<br />

on revolutionary principles provided a particularly receptive<br />

environment for it.<br />

As suggested earlier, machine civilization, in searching for


ORDER OUT OF CHAOS<br />

.xxvi<br />

an explanation of itself in the cosmic order of things, seized<br />

upon the Newtonian model and rewarded those who further<br />

developed it. It is not only in chemical beakers that we find<br />

auto-catalysis, as the authors would be the first to contend.<br />

For these reasons, it still makes sense to me to regard the<br />

Newtonian knowledge system as, itself, a "cultural dissipative<br />

structure" born of social fluctuation.<br />

Ironically, as I've said, I believe their own ideas are central<br />

to the latest revolution in science, and I cannot help but see<br />

these ideas in relationship to the demise of the Machine Age<br />

and the rise of what I have called a "Third Wave" civilization.<br />

Applying their own terminology, we might characterize today's<br />

breakdown of industrial or "Second Wave" society as a<br />

civilizational "bifurcation," and the rise of a more differentiated,<br />

"Third Wave" society as a leap to a new "dissipative<br />

structure" on a world scale. And if we accept this analogy,<br />

might we not look upon the leap from Newtonianism to Prigoginianism<br />

in the same way? Mere analogy, no doubt. But<br />

illuminating, nevertheless.<br />

Finally, we come once more to the ever-challenging issue of<br />

chance and necessity. For if Prigogine and Stengers are right<br />

and chance plays its role at or near the point of bifurcation,<br />

after which deterministic processes take over once more until<br />

the next bifurcation, are they not embedding chance, itself,<br />

within a deterministic framework? By assigning a particular<br />

role to chance, don't they de-chance it?<br />

This question, however, I had the pleasure of discussing<br />

with Prigogine, who smiled over dinner and replied, "Yes.<br />

That would be true. But, of course, we can never determine<br />

when the next bifurcation will arise." Chance rises phoenixlike<br />

once more.<br />

Order out of Chaos is a brilliant, demanding, dazzling bookchallenging<br />

for all and richly rewarding for the attentive reader. It<br />

is a book to study, to savor, to reread-and to question yet again. It<br />

places science and humanity back in a world in which ceteris<br />

paribus is a myth-a world in which other things are seldom held<br />

steady, equal, or unchanging. In short, it projects science into<br />

today's revolutionary world of instability, disequilibrium, and turbulence.<br />

In so doing, it serves the highest creative function-it<br />

helps us create fresh order.


PREFACE<br />

MAN'S NEW DIALOGUE<br />

WITH NATURE<br />

Our vision of nature is undergoing a radical change toward the<br />

multiple, the temporal, and the complex. For a long time a<br />

mechanistic world view dominated Western science. In this<br />

view the world appeared as a vast automaton. We now understand<br />

that we live in a pluralistic world. It is true that there are<br />

phenomena that appear to us as deterministic and reversible,<br />

such as the motion of a frictionless pendulum or the motion of<br />

the earth around the sun. Reversible processes do not know<br />

any privileged direction of time. But there are also irreversible<br />

processes that involve an arrow of time. If you bring together<br />

two liquids such as water and alcohol, they tend to mix in the<br />

forward direction of time as we experience it. We never observe<br />

the reverse process, the spontaneous separation of the<br />

mixture into pure water and pure alcohol. This is therefore an<br />

irreversible process. All of chemistry involves such irreversible<br />

processes.<br />

Obviously, in addition to deterministic processes, there<br />

must be an element of probability involved in some basic processes,<br />

such as, for example, biological evolution or the evolution<br />

of human cultures. Even the scientist who is convinced of<br />

the validity of deterministic descriptions would probably hesitate<br />

to imply that at the very moment of the Big Bang, the<br />

moment of the creation of the universe as we know it, the date<br />

of the publication of this book was already inscribed in the<br />

laws of nature. In the classical view the basic processes of<br />

nature were considered to be deterministic and reversible.<br />

Processes involving randomness or irreversibility were considered<br />

only exceptions. Today we see everywhere the role of<br />

irreversible processes, of fluctuations_<br />

Although Western science has stimulated an extremely fruitxxvii


ORDER OUT OF CHAOS<br />

xxviii<br />

ful dialogue between man and nature, some of its cultural consequences<br />

have been disastrous. The dichotomy between the<br />

"two cultures" is to a large extent due to the conflict between<br />

the atemporal view of classical science and the time-oriented<br />

view that prevails in a large part of the social sciences and<br />

humanities. But in the past few decades, something very dramatic<br />

has been happening in science, something as unexpected<br />

as the birth of geometry or the grand vision of the<br />

cosmos as expressed in Newton's work. We are becoming<br />

more and more conscious of the fact that on all levels, from<br />

elementary particles to cosmology, randomness and irreversibility<br />

play an ever-increasing role. Science is rediscovering<br />

time. It is this conceptual revolution that this book sets out to<br />

describe.<br />

This revolution is proceeding on all levels, on the level of<br />

elementary particles, in cosmology, and on the level of socalled<br />

macroscopic physics, which comprises the physics and<br />

chemistry of atoms and molecules either taken individually or<br />

considered globally as, for example, in the study of liquids or<br />

gases. It is perhaps particularly on this macroscopic level that<br />

the reconceptualization of science is most easy to follow. Classical<br />

dynamics and modern chemistry are going through a period<br />

of drastic change. If one asked a physicist a few years ago<br />

what physics permits us to explain and which problems remain<br />

open, he would have answered that we obviously do not<br />

have an adequate understanding of elementary particles or of<br />

cosmological evolution but that our knowledge of things in between<br />

was pretty satisfactory. Today a growing minority, to<br />

which we belong, would not share this optimism: we have only<br />

begun to understand the level of nature on which we live, and<br />

this is the level on which we have concentrated in this book.<br />

To appreciate the reconceptualization of physics taking<br />

place today, we must put it in proper historical perspective.<br />

The history of science is far from being a linear unfolding that<br />

corresponds to a series of successive approximations toward<br />

some intrinsic truth. It is full of contradictions, of unexpected<br />

turning points. We have devoted a large portion of this book to<br />

the historical pattern followed by Western science, starting<br />

with Newton three centuries ago. We have tried to place the<br />

history of science in the frame of the history of ideas to integrate<br />

it in the evolution of Western culture during the past


xxfx<br />

MAN'S NEW DIALOGUE WITH NATURE<br />

three centuries. Only in this way can we appreciate the unique<br />

moment in which we are presently living.<br />

Our scientific heritage includes two basic questions to<br />

which till now no answer was provided. One is the relation<br />

between disorder and order. The famous law of increase of<br />

entropy -describes the world as evolving from order to disorder<br />

; still, biological or social evolution shows us the complx<br />

emerging fr om the simple. How is this possible? How can<br />

structure arise from disorder? Great progress has been realized<br />

in this question. We know now that nonequilibrium, the<br />

flow of matter and energy, may be a source of order.<br />

But there is the second question, even more basic: classical<br />

or quantum physics describes the world as reversible, as<br />

static. In this description there is no evolution, neither to<br />

order nor to disorder ; the "information," as may be defined<br />

from dynamics, remains constant in time. Therefore there is<br />

an obvious contradiction between the static view of dynamics<br />

and the evolutionary paradigm of thermodynamics. What is<br />

irreversibility? What is entropy? Few questions have been discussed<br />

more often in the course of the history of science. We<br />

begin to be able to give some answers. Order and disorder are<br />

complicated notions: the units involved in the static description<br />

of dynamics are not the same as those that have to be<br />

introduced to achieve the evolutionary paradigm as expressed<br />

by the growth of entropy. This transition leads to a new concept<br />

of matter, matter that is "active," as matter leads to irreversible<br />

processes and as irreversible processes <strong>org</strong>anize<br />

matter.<br />

The evolutionary paradigm, including the concept of entropy,<br />

has exerted a considerable fascination that goes far<br />

beyond science proper. We hope that our unification of dynamics<br />

and thermodynamics will bring out clearly the radical novelty<br />

of the entropy concept in respect to the mechanistic world<br />

view. Time and reality are closely related. For humans, reality<br />

is embedded in the flow of time. As we shall see, the irreversibility<br />

of time is itself closely connected to entropy. To make<br />

time flow backward we would have to overcome an infinite<br />

entropy barrier.<br />

Tr aditionally science has dealt with universals, humanities<br />

with particulars. The convergence of science and humanities<br />

was emphasized in the French title of our book, La Nouvelle


ORDER OUT OF CHAOS<br />

xxx<br />

Alliance, published by Gallimard, Paris, in 1979. However, we<br />

have not succeeded in finding a proper English equivalent of<br />

this title. Furthermore, the text we present here differs from<br />

the French edition, especially in Chapters VII through IX. Although<br />

the origin of structures as the result of nonequilibrium<br />

processes was already adequately treated in the French edition<br />

(as well as in the translations that followed), we had to<br />

entirely rewrite the third part, which deals with our new results<br />

concerning the roots of time as well as with the formulation<br />

of the evolutionary paradigm in the frame of the physical<br />

sciences.<br />

This is all quite recent. The reconceptualization of physics<br />

is far from being achieved. We have decided, however, to present<br />

the situation as it seems to us today. We have a feeling of<br />

great intellectual excitement: we begin to have a glimpse of the<br />

road that leads from being to becoming. As one of us has devoted<br />

most of his scientific life to this problem, he may perhaps<br />

be excused for expressing his feeling of satisfaction, of<br />

aesthetic achievement, which he hopes the reader will share.<br />

For too long there appeared a conflict between what seemed<br />

to be eternal, to be out of time, and what was in time. We see<br />

now that there is a more subtle form of reality involving both<br />

time and eternity.<br />

This book is the outcome of a collective effort in which<br />

many colleagues and friends have been involved. We cannot<br />

thank them all individually. We would like, however, to single<br />

out what we owe to Erich Jantsch, Aharon Katchalsky, Pierre<br />

Resibois, and Leon Rosenfeld, who unfo rtunately are no<br />

longer with us. We have chosen to dedicate this book to their<br />

memory.<br />

We want also to acknowledge the continuous support we<br />

have received from the Instituts Internationaux de Physique et<br />

de Chimie, founded by E. Solvay, and from the Robert A.<br />

Welch Foundation.<br />

The human race is in a period of transition. Science is likely<br />

to play an important role at this moment of demographic explosion.<br />

It is therefore more important than ever to keep open<br />

the channels of communication between science and society.<br />

The present development of We stern science has taken it outside<br />

the cultural environment of the seventeenth centu.ry, in<br />

which it was born. We believe that science today carries a uni-


xxxi<br />

MAN'S NEW DIALOGUE WITH NATURE<br />

versal message that is more acceptable to different cultural<br />

traditions.<br />

During the past decades Alvin Toffier's books have been important<br />

in bringing to the attention of the public some features<br />

of the "Third Wave" that characterizes our time. We are therefore<br />

grateful to him for having written the Foreword to the<br />

English-language version of our book. English is not our native<br />

language. We believe that to some extent ever y language<br />

provides a different way of describing the common reality in<br />

which we are embedded. Some of these characteristics will<br />

survive even the most careful translation. In any case, we are<br />

most grateful to Joseph Early, Ian MacGilvray, Carol<br />

Thurston, and especially to Carl Rubino for their help in the<br />

preparation of this English-language version. We would also<br />

like to express our deep thanks to Pamela Pape for the careful<br />

typing of the successive versions of the manuscript.


INTRODUCTION<br />

THE CHALLENGE TO<br />

SCIENCE<br />

1<br />

It is hardly an exaggeration to state that one of the greatest<br />

dates in the history of mankind was April 28, 1686, when Newton<br />

presented his Principia to the Royal Society of London. It<br />

contained the basic laws of motion together with a clear formulation<br />

of some of the fundamental concepts we still use today,<br />

such as mass, acceleration, and inertia. The greatest<br />

impact was probably made by Book III of the Principia, titled<br />

The System of the World, which included the universal law of<br />

gravitation. Newton's contemporaries immediately grasped<br />

the unique importance of his work. Gravitation became a topic<br />

of conversation both in London and Paris.<br />

Three centuries have now elapsed since Newton's Principia.<br />

Science has grown at an incredible speed, permeating the life<br />

of all of us. Our scientific horizon has expanded to truly fantastic<br />

proportions. On the microscopic scale, elementary partide<br />

physics studies processes involving physical dimensions<br />

of the order of w- ts em and times of the order of I0-22 second.<br />

On the other hand, cosmology leads us to times of the<br />

order of 1010 years, the "age of the universe." Science and<br />

technology are closer than ever. Among other factors, new<br />

biotechnologies and the progress in information techniques<br />

promise to change our lives in a radical way.<br />

Running parallel to this quantitative growth are deep qualitative<br />

changes whose repercussions reach far beyond science<br />

proper and affect the very image of nature. The great founders<br />

of Western science stressed the universality and the eternal<br />

character of natural laws. They set out to formulate general<br />

schemes that would coincide with the very ideal of rationality.


ORDER OUT OF CHAOS 2<br />

As Roger Hausheer says in his fine introduction to Isaiah<br />

Berlin's Against the Current, "They sought all-embracing<br />

schemas, universal unifying frameworks, within which everything<br />

that exists could be shown to be systematically-i.e.,<br />

logically or causally-interconnected, vast structures in which<br />

there should be no gaps left open for spontaneous, unattended<br />

developments, where everything that occurs should be, at<br />

least in principle, wholly explicable in terms of immutable<br />

general laws." t<br />

The story of this quest is indeed a dramatic one. There were<br />

moments when this ambitious program seemed near completion.<br />

A fundamental level from which all other properties of<br />

matter could be deduced seemed to be in sight. Such moments<br />

can be associated with the formulation of Bohr's celebrated<br />

atomic model, which reduced matter to simple planetary systems<br />

formed by electrons and protons. Another moment of<br />

great suspense came when Einstein hoped to condense all the<br />

laws of physics into a single "unified field theory. " Great progress<br />

has indeed been realized in the unification of some of the<br />

basic forces found in nature. Still, the fundamental level remains<br />

elusive. Wherever we look we fi nd evolution, diversification,<br />

and instabilities. Curiously, this is true on all levels,<br />

in the field of elementary particles, in biology, and in astrophysics,<br />

with the expanding universe and the formation of<br />

black holes.<br />

As we said in the Preface, our vision of nature is undergoing<br />

a radical change toward the multiple, the temporal, and the<br />

complex. Curiously, the unexpected complexity that has been<br />

discovered in nature has not led to a slowdown in the progress<br />

of science, but on the contrary to the emergence of new conceptual<br />

structures that now appear as essential to our understanding<br />

of the physical world-the world that includes us. It<br />

is this new situation, which has no precedent in the history of<br />

science, that we wish to analyze in this book.<br />

The story of the transformation of our conceptions about<br />

science and nature can hardly be separated from another<br />

story, that of the feelings aroused by science. With every new<br />

intellectual program always come new hopes, fears, and expectations.<br />

In classical science the emphasis was on time-independent<br />

laws. As we shall see, once the particular state of a<br />

system has been measured, the reversible laws of classical sci-


3 THE CHALLENGE TO SCIENCE<br />

ence are supposed to determine its future, just as they had<br />

determined its past. It is natural that this quest for an eternal<br />

truth behind changing phenomena aroused enthusiasm. But it<br />

also came as a shock that nature described in this way was in<br />

fact debased: by the very success of science, nature was<br />

shown to be an automaton, a robot.<br />

The urge to reduce the diversity of nature to a web of illusions<br />

has been present in Western thought since the time of<br />

Greek atomists. Lucretius, following his masters Democritus<br />

and Epicurus, writes that the world is "just" atoms and void<br />

and urges us to look for the hidden behind the obvious: "Still,<br />

lest you happen to mistrust my words, because the eye cannot<br />

perceive prime bodies, hear now of particles you must admit<br />

exist in the world and yet cannot be seen. "2<br />

Yet it is well known that the driving force behind the work of<br />

the Greek atomists was not to debase nature but to free men<br />

from fear, the fear of any supernatural being, of any order that<br />

would transcend that of men and nature. Again and again Lucretius<br />

repeats that we have nothing to fear, that the essence of<br />

the world is the ever-changing associations of atoms in the<br />

void.<br />

Modern science transmuted this fundamentally ethical<br />

stance into what seemed to be an established truth; and this<br />

truth, the reduction of nature to atoms and void, in turn gave<br />

rise to what Lenoble3 has called the "anxiety of modern<br />

men." How can we recognize ourselves in the random world<br />

of the atoms? Must science be defined in terms of rupture between<br />

man and nature? 'JI bodies, the firmament, the stars,<br />

the earth and its kingdoms are not equal to the lowest mind;<br />

for mind knows all this in itself and these bodies nothing. "4<br />

This "Pensee" by Pascal expresses the same feeling of alienation<br />

we find among contemporary scientists such as Jacques<br />

Monod:<br />

Man must at last finally awake from his millenary<br />

dream; and in doing so, awake to his total solitude, his<br />

fundamental isolation. Now does he at last realize that,<br />

like a gypsy, he lives on the boundary of an alien world. A<br />

world that is deaf to his music. just as indifferent to his<br />

hopes as it is to his suffering or his crimes.s


ORDER OUT OF CHAOS 4<br />

This is a paradox. A brilliant breakthrough in molecular biology,<br />

the deciphering of the genetic code, in which Monod<br />

actively participated, ends upon a tragic note. This very progress,<br />

we are told, makes us the gypsies ,0f the universe. How<br />

can we explain this situation? Is not science a way of communication,<br />

a dialogue with nature?<br />

In the past, strong distinctions were frequently made between<br />

man's world and the supposedly alien natural world. A<br />

famous passage by Vico in The New Science describes this<br />

most vividly:<br />

... in the night of thick darkness enveloping the earliest<br />

antiquity, so remote from ourselves, there shines the eternal<br />

and never failing light of a truth beyond all question:<br />

that the world of civil society has certainly been made by<br />

men, and that its principles are therefore to be found<br />

within the modifications of our own human mind.<br />

Whoever reflects on this cannot but marvel that the philosophers<br />

should have bent all their energies to the study<br />

of the world of nature, which, since God made it, He<br />

alone knows; and that they should have neglected the<br />

study of the world of nations, or civil world, which, since<br />

men had made it, men could come to know. 6<br />

Present -day research leads us farther and farther away from<br />

the opposition between man and the natural world. It will be<br />

one of the main purposes of this book to show, instead of rupture<br />

and opposition, the growing coherence of our knowledge<br />

of man and nature.<br />

In the past, the questioning of nature has taken the most diverse<br />

forms. Sumer discovered writing; the Sumerian priests<br />

speculated that the future might be written in some hidden<br />

way in the events taking place around us in the present. They<br />

even systematized this belief, mixing magical and rational elements.<br />

7 In this sense we may say that Western science, which<br />

originated in the seventeenth century, only opened a new<br />

chapter in the everlasting dialogue between man and nature.


5 THE CHALLENGE TO SCIENCI:<br />

Alexandre Koyres has defined the innovation brought about<br />

by modern science in terms of "experimentation." Modern<br />

science is based on the discovery of a new and specific form of<br />

communication with nature-that is, on the conviction that<br />

nature responds to experimental interrogation. How can we<br />

define more precisely the experimental dialogue? Experimentation<br />

does not mean merely the faithful observation of facts<br />

as they occur, nor the mere search for empirical connections<br />

between phenomena, but presupposes a systematic interaction<br />

between theoretical concepts and observation.<br />

In hundreds of different ways scientists have expressed ttieir<br />

amazement when, on determining the right question, they discover<br />

that they can see how the puzzle fits together. In this<br />

sense, science is like a two-partner game ir;t which we have to<br />

guess the behavior of a reality unrelated to our beliefs, our<br />

ambitions, or our hopes. Nature cannot be forced to say anything<br />

we want it to. Scientific investigation is not a monologue.<br />

It is precisely the risk involved that makes this game<br />

exciting.<br />

But the uniqueness of Western science is far from being exhausted<br />

by such methodological considerations. When Karl<br />

Popper discussed the normative description of scientific rationality,<br />

he was forced to admit that in the final analysis rational<br />

science owes its existence to its success; the scientific<br />

method is applicable only by virtue of the astonishing points<br />

of agreement between preconceived models and experimental<br />

results.9 Science is a risky game, but it seems to have discovered<br />

questions to which nature provides consistent answers.<br />

The success of Western science is an historical fact, unpredictable<br />

a priori, but which cannot be ignored. The surprising<br />

success of modern science has led to an irreversible transformation<br />

of our relations with nature. In this sense, the term<br />

"scientific revolution" can legitimately be used. The history of<br />

mankind has been marked by other turning points, by other<br />

singular conjunctions of circumstances leading to irreversible<br />

changes. One such crucial event is known as the "Neolithic<br />

revolution." But there, just as in the case of the "choices"<br />

marking biological evolution, we can at present only proceed<br />

by guesswork, while there is a wealth of information concerning<br />

decisive episodes in the evolution of science. The so-called


ORDER OUT OF CHAOS 6<br />

"Neolithic revolution" took thousands of years. Simplifying<br />

somewhat, we may say the scientific revolution started only<br />

three ·Centuries ago. We have what is perhaps a unique opportunity<br />

to apprehend the specific and intelligible mixture of<br />

"chance" and "necessity" marking this revolution.<br />

Science initiated a successful dialogue with nature. On the<br />

other hand, the first outcome of this dialogue was the discovery<br />

of a silent world. This is the paradox of classical science.<br />

It revealed to men a dead, passive nature, a nature that behaves<br />

as an automaton which, once programmed, continues to<br />

follow the rules inscribed in the program. In this sense the<br />

dialogue with nature isolated man from nature instead of<br />

bringing him closer to it. A triumph of human reason turned<br />

into a sad truth. It seemed that science debased everything it<br />

touched.<br />

Modern science horrified both its opponents, for whom it<br />

appeared as a deadly danger, and some of its supporters, who<br />

saw in man's solitude as "discovered" by science the price we<br />

had to pay for this new rationality.<br />

The cultural tension associated with classical science can be<br />

held at least partly responsible for the unstable position of science<br />

within society; it led to an heroic assumption of the harsh<br />

implications of rationality, but it led also to violent rejection.<br />

We shall return later to present-day antiscience movements.<br />

Let us take an earlier example-the irrationalist movement in<br />

Germany in the 1920s that formed the cultural background to<br />

quantum mechanics. to In opposition to science, which was<br />

identified with a set of concepts such as causality, determinism,<br />

reductionism, and rationality, there was a violent upsurge<br />

of ideas denied by science but seen as the embodiment<br />

of the fundamental irrationality of nature. Life, destiny, freedom,<br />

and spontaneity thus became manifestations of a shadowy<br />

underworld impenetrable to reason. Without going into<br />

the peculiar sociopolitical context to which it owed its vehement<br />

nature, we can state that this rejection illustrates the<br />

risks associated with classical science. By admitting only a<br />

subjective meaning for a set of experiences men believe to be<br />

significant, science runs the risk of transferring these into the<br />

realm of the irrational, bestowing upon them a formidable<br />

power.<br />

As Joseph Needham has emphasized, Western thought has


7 THE CHALLENGE TO SCIENCE<br />

always oscillated between the world as an automaton and a<br />

theology in which God governs the universe. This is what<br />

Needham calls the "characteristic European schizophrenia.<br />

" II In fact, these visions are connected. An automaton<br />

needs an external god.<br />

Do we really have to make this tragic choice? Must we<br />

choose between a science that leads to alienation and an antiscientific<br />

metaphysical view of nature? We think such a choice<br />

is no longer necessary, since the changes that science is undergoing<br />

today lead to a radically new situation. This recent evolution<br />

of science gives us a unique opportunity to reconsider<br />

its position in culture in general. Modern science originated in<br />

the specific context of the European seventeenth century. We<br />

are now approaching the end of the twentieth century, and it<br />

seems that some more universal message is carried by science,<br />

a message that concerns the interaction of man and nature<br />

as well as of man with man.<br />

What are the assumptions of classical science from which we<br />

believe science has freed itself today? Generally those centering<br />

around the basic conviction that at some level the world is<br />

simple and is governed by time-reversible fundamental laws.<br />

Today this appears as an excessive simplification. We may<br />

compare it to reducing buildings to piles of bricks. Ye t out of<br />

the same bricks we may construct a factory, a palace, or a<br />

cathedral. It is on the level of the building as a whole that we<br />

apprehend it as a creature of time, as a product of a culture, a<br />

society, a style. But there is the additional and obvious problem<br />

that, since there is no one to build nature, we must give to<br />

its very "bricks"-that is, to its microscopic activity-a description<br />

that accounts for this building process.<br />

The quest of classical science is in itself an illustration of a<br />

dichotomy that runs throughout the history of We stern<br />

thought. Only the immutable world of ideas was traditionally<br />

recognized as "illuminated by the sun of the intelligible," to<br />

use Plato's expression. In the same sense, only eternal laws<br />

were seen to express scientific rationality. Te mporality was<br />

looked down upon as an illusion. This is no longer true today.


ORDER OUT OF CHAOS 8<br />

We have discovered that far from being an illusion, irrevers·<br />

ibility plays an essential role in nature and lies at the origin of<br />

most processes of self-<strong>org</strong>anization. We find ourselves in a<br />

world in which reversibility and determinism apply only to<br />

limiting, simple cases, while irreversibility and randomness<br />

are the rules. .<br />

The denial of time and complexity was central to the cultural<br />

issues raised by the scientific enterprise in its classical definition.<br />

The challenge of these concepts was also decisive for the<br />

metamorphosis of science we wish to describe. In his great<br />

book The Nature of the Physical World, Arthur Eddingtonl2<br />

introduced a distinction between primary and secondary laws.<br />

"Primary laws" control the behavior of single particles, while<br />

"secondary laws" are applicable to collections of atoms or<br />

molecules. To insist on secondary laws is to emphasize that<br />

the description of elementary behaviors is not sufficient for<br />

understanding a system as a whole. An outstanding case of a<br />

secondary law is, in Eddington's view, the second law of thermodynamics,<br />

the law that introduces the "arrow of time" in<br />

physics. Eddington writes: "From the point of view of philosophy<br />

of science the conception associated with entropy<br />

must, I think, be ranked as the great contribution of the nineteenth<br />

century to scientific thought. It marked a reaction from<br />

the view that everything to which science need pay attention is<br />

discovered by a microscopic dissection of objects." 13 This<br />

trend has been dramatically amplified today.<br />

It is true that some of the greatest successes of modern science<br />

are discoveries at the microscopic level, that of molecules,<br />

atoms, or elementary particles. For example, molecular<br />

biology has been immensely successful in isolating specific<br />

molecules that play a central role in the mechanism of life. In<br />

fact, this success has been so overwhelming that for many scientists<br />

the aim of research is identified with this "microscopic<br />

dissection of objects," to use Eddington's expression. However,<br />

the second law of thermodynamics presented the first<br />

challenge to a concept of nature that would explain away the<br />

complex and reduce it to the simplicity of some hidden world.<br />

Today interest is shifting from substance to relation, to communication,<br />

to time.<br />

This change of perspective is not the result of some arbitrary<br />

decision. In physics it was forced upon us by new dis-


9 THE CHALLENGE TO SCIENCE<br />

coveries no one could have foreseen. Who would have<br />

expected that most (and perhaps all) elementary particles<br />

would prove to be unstable? Who would have expected that<br />

with the experimental confirmation of an expanding universe<br />

we could conceive of the history of the world as a whole?<br />

At the end of the twentieth century we have learned to understand<br />

better the meaning of the two great revolutions that<br />

gave shape to the physics of our time, quantum mechanics and<br />

relativity. They started as attempts to correct classical mechanics<br />

and to incorporate into it the newly found universal<br />

constants. Today the situation has changed. Quantum mechanics<br />

has given us the theoretical frame to describe the incessant<br />

transformations of particles into each other. Similarly,<br />

general relativity has become the basic theory in terms of<br />

which we can describe the thermal history of our universe in<br />

its early stages.<br />

Our universe has a pluralistic, complex character. Structures<br />

may disappear, but also they may appear. Some processes<br />

are , as far as we know, well described by deterministic<br />

equations, but others involve probabilistic processes.<br />

How then can we overcome the apparent contradiction between<br />

these concepts? We are living in a single universe. As<br />

we shall see, we are beginning to appreciate the meaning of<br />

these problems. Moreover, the importance we now give to the<br />

various phenomena we observe and describe is quite different<br />

from, even opposite to, what was suggested by classical physics.<br />

There the basic processes, as we mentioned, are considered as<br />

deterministic and reversible. Processes involving randomness<br />

or irreversibility are considered to be exceptions. Today we<br />

see everywhere the role of irreversible processes, of fluctuations.<br />

The models considered by classical physics seem to us<br />

to occur only in limiting situations such as we can create artificially<br />

by putting matter into a box and then waiting till it<br />

reaches equilibrium.<br />

The artificial may be deterministic and reversible. The natural<br />

contains essential elements of randomness and irreversibility.<br />

This leads to a new view of matter in which matter is no<br />

longer the passive substance described in the mechanistic<br />

world view but is associated with spontaneous activity. This<br />

change is so profound that, as we stated in our Preface, we can<br />

really speak about a new dialogue of man with nature.


ORDER OUT OF CHAOS 10<br />

This book deals with the conceptual transformation of science<br />

from the Golden Age of classical science to the present.<br />

To describe this transformation we could have chosen many<br />

roads. We could have studied the problems of elementary particles.<br />

We could have followed recent fascinating developments<br />

in astrophysics. These are the subjects that seem to<br />

delimit the frontiers of science. However, as we stated in our<br />

Preface, over the past years so many new feaures of nature at<br />

our level have been discovered that we decided to concentrate<br />

on this intermediate level, on problems that belong mainly to<br />

our macroscopic world, which includes atoms, molecules, and<br />

especially biomolecules. Still it is important to emphasize that<br />

the evolution of science proceeds on somewhat parallel lines at<br />

every level, be it that of elementary particles, chemistry, biology,<br />

or cosmology. On every scale self-<strong>org</strong>anization, complexity,<br />

and time play a new and unexpected role.<br />

Therefore, our aim is to examine the significance of three<br />

centuries of scientific progress from a definite viewpoint.<br />

There is certainly a subjective element in the way we have<br />

chosen our material. The problem of time is really the center<br />

of the research that one of us has been pursuing all his life.<br />

When as a young student at the University of Brussels he<br />

came into contact with physics and chemistry for the first<br />

time, he was astonished that science had so little to say about<br />

time, especially since his earlier education had centered<br />

mainly around history and archaeology. This surprise could<br />

have led him to two attitudes, both of which we find exemplified<br />

in the past: one would have been to discard the prob<br />

lem, since classical science seemed to have no place for time;<br />

and the other would have been to look for some other way of<br />

apprehending nature, in which time would play a different,<br />

more basic role. This is the path Bergson and Whitehead, to<br />

mention only two philosophers. of our century, chose. The first<br />

position would be a "positivistic" one, the second a "metaphysical"<br />

one.<br />

There was, however, a third path, which was to ask whether<br />

the simplicity of the temporal evolution traditionally consid-


11 THE CHALLENGE TO SCIENCE<br />

ered in physics and chemistry was due to the fact that attention<br />

was paid mainly to some very simplified situations, to<br />

heaps of bricks in contrast with the cathedral to which we have<br />

alluded.<br />

This book is divided into three parts. The first part deals<br />

with the triumph of classical science and the cultural consequences<br />

of this triumph. Initially, science was greeted with<br />

enthusiasm. We shall then describe the cultural polarization<br />

that occurred as a result of the existence of classical science<br />

and its astonishing success. Is this success to be accepted as<br />

such, perhaps limiting its implications, or must the scientific<br />

method itself be rejected as partial or illusory? Both choices<br />

lead to the same result-the collision between what has often<br />

been called the "two cultures," science and the humanities.<br />

These questions have played a basic role in Western thought<br />

since the formulation of classical science. Again and again we<br />

come to the problem, "How to choose?" Isaiah Berlin has<br />

rightly seen in this question the beginning of the schism between<br />

the sciences and the humanities:<br />

The specific and unique versus the repetitive and the universal,<br />

the concrete versus the abstract, perpetual movement<br />

versus rest, the inner versus the outer, quality<br />

versus quantity, culture-bound versus timeless principles,<br />

mental strife and self-transformation as a permanent condition<br />

of man versus the possibility (and desirability) of<br />

peace, order, final harmony and the satisfaction of all ra-<br />

- tional human wishes-these are some of the aspects of<br />

the contrast.14<br />

We have devoted much space to classical mechanics. Indeed,<br />

in our view this is the best vantage point from which we<br />

may contemplate the present-day transformation of science.<br />

Classical dynamics seems to express in an especially clear and<br />

striking way the static view of nature. Here time apparently is<br />

reduced to a parameter, and future and past become equivalent.<br />

It is true that quantum theory has raised many new<br />

problems not covered by classical dynamics but it has nevertheless<br />

retained a number of the conceptual positions of classical<br />

dynamics, particularly as far as time and process are<br />

concerned.


ORDER OUT OF CHAOS 12<br />

As early as at the beginning of the nineteenth century,<br />

precisely when classical science was triumphant, when the<br />

Newtonian program dominated French science and the latter<br />

dominated Europe, the first threat to the Newtonian construction<br />

loomed into sight. In the second part of our study we<br />

shall follow the development of the science of heat, this rival to<br />

Newton's science of gravity, starting from the first gauntlet<br />

thrown down when Fourier formulated the law governing the<br />

propagation of heat. It was, in fact, the first quantitative description<br />

of something inconceivable in classical dynamicsan<br />

irreversible process.<br />

The two descendants of the science of heat, the science of<br />

energy conversion and the science of heat engines, gave birth<br />

to the first "nonclassical" science-thermodynamics. The<br />

most original contribution of thermodynamics is the celebrated<br />

second law, which introduced into physics the arrow of<br />

time. This introduction was part of a more global intellectual<br />

move. The nineteenth century was really the century of evolution;<br />

biology, geology, and sociology emphasized processes of<br />

becoming, of increasing complexity. As for thermodynamics,<br />

it is based on the distinction of two types of processes: reversible<br />

processes, which are independent of the direction of time,<br />

and irreversible processes, which depend on the direction of<br />

time. We shall see examples later. It was in order to distinguish<br />

the two types of processes that the concept of entropy was<br />

introduced, since entropy increases only because of the irreversible<br />

processes.<br />

During the nineteenth century the final state of thermodynamic<br />

evolution was at the center of scientific research. This<br />

was equilibrium thermodynamics. Irreversible processes were<br />

looked down on as nuisances, as disturbances, as subjects not<br />

worthy of study. Today this situation has completely changed.<br />

We now know that far from equilibrium, new types of structures<br />

may originate spontaneously. In far-from-equilibrium<br />

conditions we may have transformation from disorder, from<br />

thermal chaos, into order. New dynamic states of matter may<br />

originate, states that reflect the interaction of a given system<br />

with its surroundings. We have called these new structures dissipative<br />

structures to emphasize the constructive role of dissipative<br />

processes in their fo rmation.<br />

This book describes some of the methods that have been


13 THE CHALLENGE TO SCIENCE<br />

developed in recent years to deal with the appearance and evolution<br />

of dissipative structures. Here we find the key words<br />

that run throughout this book like leitmotivs: nonlinearity, instability,<br />

fluctuations. They have begun to permeate our view<br />

of nature even beyond the fields of physics and chemistry<br />

proper.<br />

We cited Isaiah Berlin when we discussed the opposition between<br />

the sciences and the humanities. He opposed the specific<br />

and unique to the repetitive and the universal. The remarkable<br />

feature is that when we move from equilibrium to far-fromequilibrium<br />

conditions, we move away from the repetitive and<br />

the universal to the specific and the unique. Indeed, the laws<br />

of equilibrium are universal. Matter near equilibrium behaves<br />

in a "repetitive" way. On the other hand, far from equilibrium<br />

there appears a variety of mechanisms corresponding to the<br />

possibility of occurrence of various types of dissipative structures.<br />

For example, far from equilibrium we may witness the<br />

appearance of chemical clocks, chemical reactions which behave<br />

in a coherent, rhythmical fashion. We may also have processes<br />

of self-<strong>org</strong>anization leading to nonhomogeneous<br />

structures to nonequilibrium crystals.<br />

We would like to emphasize the unexpected character of this<br />

behavior. Every one of us has an intuitive view of how a chemical<br />

reaction takes place; we imagine molecules floating<br />

through space, colliding, and reappearing in new forms. We<br />

see chaotic behavior similar to what the atomists described<br />

when they spoke about dust dancing in the air. But in a chemical<br />

clock the behavior is quite different. Oversimplifying somewhat,<br />

we can say that in a chemical clock all molecules change<br />

their chemical identity simultaneously, at regular time intervals.<br />

If the molecules can be imagined as blue or red, we<br />

would see their change of color following the rhythm of the<br />

chemical clock reaction.<br />

Obviously such a situation can no longer be described in<br />

terms of chaotic behavior. A new type of order has appeared.<br />

We can speak of a new coherence, of a mechanism of "communication"<br />

among molecules. But this type of communication<br />

can arise only in far-from-equilibrium conditions. It is<br />

quite interesting that such communication seems to be the rule<br />

in the world of biology. It may in fact be taken as the very basis<br />

of the definition of a biological system.


ORDER OUT OF CHAOS 14<br />

In addition, the type of dissipative structure depends critically<br />

on the conditions in which the structure is formed. External<br />

fields such as the gravitational field of earth, as well as<br />

the magnetic field, may play an essential role in the selection<br />

mechanism of self-<strong>org</strong>anization.<br />

We begin to see how, starting from chemistry, we may build<br />

complex structures, complex forms, some of which may have<br />

been the precursors of life. What seems certain is that these<br />

far-from-equilibrium phenomena illustrate an essential and unexpected<br />

property of matter: physics may henceforth describe<br />

structures as adapted to outside conditions. We meet in rather<br />

simple chemical systems a kind of prebiological adaptation<br />

mechanism. To use somewhat anthropomorphic language: in<br />

equilibrium matter is "blind," but in far-from-equilibrium conditions<br />

it begins to be able to perceive, to "take into account,"<br />

in its way of functioning, differences in the external world<br />

(such as weak gravitational or electrical fields).<br />

Of course, the problem of the origin of life remains a difficult<br />

one, and we do not think a simple solution is imminent.<br />

Still, from this perspective life no longer appears to oppose the<br />

"normal" laws of physics, struggling against them to avoid its<br />

normal fate-its destruction. On the contrary, life seems to<br />

express in a specific way the very conditions in which our biosphere<br />

is embedded, incorporating the nonlinearities of chemical<br />

reactions and the far-from-equilibrium conditions imposed<br />

on the biosphere by solar radiation.<br />

We have discussed the concepts that allow us to describe the<br />

formation of dissipative structures, such as the theory of bifurcations.<br />

It is remarkable that near-bifurcations systems present<br />

large fluctuations. Such systems seem to "hesitate"<br />

among various possible directions of evolution, and the famous<br />

law of large numbers in its usual sense breaks down. A<br />

small fluctuation may start an entirely new evolution that will<br />

drastically change the whole behavior of the macroscopic system.<br />

The analogy with social phenomena, even with history, is<br />

inescapable. Far from opposing "chance" and "necessity," we<br />

now see both aspects as essential in the description of nonlinear<br />

systems far from equilibrium.<br />

The first two parts of this book thus deal with two conflicting<br />

views of the physical universe: the static view of classical<br />

dynamics, and the evolutionary view associated with entropy.


15 THE CHALLENGE TO SCIENCE<br />

A confrontation between these views has become unavoidable.<br />

For a long time this confrontation was postponed by considering<br />

irreversibility as an illusion, as an approximation; it<br />

was man who introduced time into a timeless universe. However,<br />

this solution in which irreversibility is reduced to an illusion<br />

or to approximations can no longer be accepted, since we<br />

know that irreversibility may be a source of order, of coherence,<br />

of <strong>org</strong>anization.<br />

We can no longer avoid this confrontation. It is the subject<br />

of the third part of this book. We describe traditional attempts<br />

to approach the problem of irreversibility first in classical and<br />

then in quantum mechanics. Pioneering work was done here,<br />

especially by Boltzmann and Gibbs. However, we can state<br />

that the problem was left largely unsolved. As Karl Popper<br />

relates it, it is a dramatic story: first, Boltzmann thought he<br />

had given an objective formulation to the new concept of time<br />

implied in the second law. But as a result of his controversy<br />

with Zermelo and others, he had to retreat.<br />

In the light of history-or in the darkness of history­<br />

Boltzmann was defeated, according to all accepted standards,<br />

though everybody accepts his eminence as a physicist.<br />

For he never succeeded in clearing up the status of<br />

his Ji-theorem; nor did he explain entropy increase ... .<br />

Such was the pressure that he lost faith in himself .. . . 15<br />

The problem of irreversibility still remains a subject of lively<br />

controversy. How is this possible one hundred fifty years after<br />

the discovery of the second law of thermodynamics? There are<br />

many aspects to this question, some cultural and some technical.<br />

There is a cultural component in the mistrust of time. We<br />

shall on several occasions cite the opinion of Einstein. His<br />

judgment sounds final: time (as irreversibility) is an illusion. In<br />

fact, Einstein was reiterating what Giordano Bruno had written<br />

in the sixteenth century and what had become for centuries<br />

the credo of science: "The universe is, therefore, one,<br />

infinite, immobile .... It does not move itself locally .... It<br />

does not generate itself .... It is not corruptible .... It is not<br />

alterable ... . "16 For a long time Bruno's vision dominated<br />

the scientific view of the Western world. It is therefore not<br />

surprising that the intrusion of irreversibility, coming mainly


ORDER OUT OF CHAOS 16<br />

from the engineering sciences and physical chemistry, was re·<br />

ceived with mistrust. But there are technical reasons in addi·<br />

tion to cultural ones. Every attempt to "derive" irreversibility<br />

from dynamics necessarily had to fail, because irreversibility<br />

is not a universal phenomenon. We can imagine situations that<br />

are strictly reversible, such as a pendulum in the absence of<br />

friction, or planetary motion. This failure has led to dis·<br />

couragement and to the feeling that, in the end, the whole concept<br />

of irreversibility has a subjective origin. We shall discuss<br />

all these problems at some length. Let us say here that today<br />

we can see this problem from a different point of view, since<br />

we now know that there are different classes of dynamic systems.<br />

The world is far from being homogeneous. Therefore the<br />

question can be put in different terms: What is the specific<br />

structure of dynamic systems that permits them to "distinguish"<br />

past and future? What is the minimum complexity<br />

involved?<br />

Progress has been realized along these lines. We can now be<br />

more precise about the roots of time in nature. This has farreaching<br />

consequences. The second law of thermodynamics,<br />

the law of entropy, introduced irreversibility into the macroscopic<br />

world. We now can understand its meaning on the<br />

microscopic level as well. As we shall see, the second law corresponds<br />

to a selection rule, to a restriction on initial conditions<br />

that is then propagated by the laws of dynamics.<br />

Therefore the second law introduces a new irreducible element<br />

into our description of nature. While it is consistent with<br />

dynamics, it cannot be derived from dynamics.<br />

Boltzmann already understood that probability and irreversibility<br />

had to be closely related. Only when a system behaves<br />

in a sufficiently random way may the difference between past<br />

and future, and therefore irreversibility, enter into its descrip:<br />

tion. Our analysis confirms this point of view. Indeed, what is<br />

the meaning of an arrow of time in a deterministic description<br />

of nature? If the future is already in some way contained in the<br />

present, which also contains the past, what is the meaning of<br />

an arrow of time? The arrow of time a manifestation of the<br />

fact that the future is not given, that, as the French poet Paul<br />

Valery emphasized, "time is construction." 17<br />

The experience of our everyday life manifests a radical difference<br />

between time and space. We can move from one point


17 THE CHALLENGE TO SCIENCE<br />

of space to another. However, we cannot turn time around. We<br />

cannot exchange past and future. As we shall see, this feeling<br />

of impossibility is now acquiring a precise scientific meaning.<br />

Permitted states are separated from states that are prohibited<br />

by the second law of thermodynamics by means of an infinite<br />

entropy barrier. There are other barriers in physics. One is the<br />

velocity of light, which in our present view limits the speed at<br />

which signals may be transmitted. It is essential that this barrier<br />

exist; if not, causality would fall to pieces. Similarly, the<br />

entropy barrier is the prerequisite for giving a meaning tQ communication.<br />

Imagine what would happen if our future would<br />

become the past for other people! We shall return to this later.<br />

The recent evolution of physics has emphasized the reality<br />

of time. In the process new aspects of time have been uncovered.<br />

A preoccupation with time runs all through our century.<br />

Think of Einstein, Proust, Freud, Te ilhard, Peirce, or<br />

Whitehead.<br />

One of the most surprising results of Einstein's special theory<br />

of relativity, published in 1905, was the introduction of a<br />

local time associated with each observer. However, this local<br />

time remained a reversible time. Einstein's problem both in<br />

the special and the general theories of relativity was mainly<br />

that of the "communication" between observers, the way they<br />

could compare time intervals. But we can now investigate time<br />

in other conceptual contexts.<br />

In classical mechanics time was a number characterizing the<br />

position of a point on its trajectory. But time may have a different<br />

meaning on a global level. When we look at a child and<br />

guess his or her age, this age is not located in any special part<br />

of the child's body. It is a global judgment. It has often been<br />

stated that science spatializes time. But we now discover that<br />

another point of view is possible. Consider a landscape and its<br />

evolution: villages grow, bridges and roads connect different<br />

regions and transform them. Space thus acquires a temporal<br />

dimension; following the words of geographer B. Berry, we<br />

have been led to study the "timing of space."<br />

But perhaps the most important progress is that we now<br />

may see the problem of structure, of order, from a different<br />

perspective. As we shall show in Chapter VIII, from the point<br />

of view of dynamics, be it classical or quantum, there can be<br />

no one time-directed evolution. The "information" as it can be


ORDER OUT OF CHAOS 18<br />

defined in terms of dynamics remains constant in time. This<br />

sounds paradoxical. When we mix two liquids, there would<br />

occur no "evolution" in spite of the fact that we cannot, without<br />

using some external device, undo the effect of the mixing.<br />

On the contrary, the entropy law describes the mixing as the<br />

evolution toward a "disorder," toward the most probable<br />

state. We can show now that there is no contradiction between<br />

the two descriptions, but to speak about information, or order,<br />

we have to redefine the units we are considering. The important<br />

new fact is that we now may establish precise rules to go<br />

from one type of unit to the other. In other words, we have<br />

achieved a microscopic formulation of the evolutionary paradigm<br />

expressed by the second law. As the evolutionary paradigm<br />

encompasses all of chemistry as well as essential parts of<br />

biology and the social sciences, this seems to us an important<br />

conclusion. This insight is quite recent. The process of reconceptualization<br />

occurring in physics is far from being complete.<br />

However, our intention is not to shed light on the definitive<br />

acquisitions of science, on its stable and well-established results.<br />

What we wish to do is emphasize the conceptual creativeness<br />

of scientific activity and the future prospects and<br />

new problems it raises. In any case, we now know that we are<br />

only at the beginning of this exploration. We shall not see the<br />

end of uncertainty or risk. Thus we have chosen to present<br />

things as we perceive them now, fully aware of how incomplete<br />

our answers are.<br />

Erwin Schrodinger once wrote, to the indignation of many philosophers<br />

of science:<br />

. . . there is a tendency to f<strong>org</strong>et that all science is boundl<br />

up with human culture in general, and that scientific findings,<br />

even those which at the moment appear the mostj<br />

advanced and esoteric and difficult to grasp, are meaning-!<br />

less outside their cultural context. A theoretical science!<br />

unaware that those of its constructs considered relevanti<br />

and momentou:s arc de:stined eventually to be framed inl<br />

concepts and words that have a grip on the educated com-I<br />

I<br />

I<br />

I


19 THE CHALLENGE TO SCIENCE<br />

munity and become part and parcel of the general world<br />

picture-a theoretical science, I say, where this is f<strong>org</strong>otten,<br />

and where the initiated continue musing to each<br />

other in terms that are, at best, understood by a small<br />

group of close fellow travellers, will necessarily be cut off<br />

from the rest of cultural mankind; in the long run it is<br />

bound to atrophy and ossify however virulently esoteric<br />

chat may continue within its joyfully isolated groups of<br />

experts.t8<br />

One of the main themes of this book is that of a strong interaction<br />

of the issues proper to culture as a whole and the internal<br />

conceptual problems of science in particular. We find<br />

questions about time at the very heart of science. Becoming,<br />

irreversibility-these are questions to which generations of<br />

philosophers have also devoted their lives. Today, when history-be<br />

it economic, demographic, or political-is moving at<br />

an unprecedented pace, new questions and new interests require<br />

·us to enter into new dialogues, to look for a new coherence.<br />

However, we know the progress of science has often been<br />

described in terms of rupture, as a shift away from concrete<br />

experience toward a level of abstraction that is increasingly<br />

difficult tc, grasp. We believe that this kind of interpretation is<br />

only a reflection, at the epistemological level, of the historical<br />

situation in which classical science found itself, a consequence<br />

of its inability to include in its theoretical frame vast<br />

areas of the relationship between man and his environment.<br />

There doubtless exists an abstract development of scientific<br />

theories. However, the conceptual innovations that have been<br />

decisive for the development of science are not necessarily of<br />

this type. The rediscovery of time has roots both in the internal<br />

history of science and in the social context in which science<br />

finds itself today. Discoveries such as those of unstable<br />

elementary particles or of the expanding universe clearly belong<br />

to the internal history of science, but the general interest<br />

in nonequilibrium situations, in evolving systems, may reflect<br />

our feeling that humanity as a whole is today in a transition<br />

period. Many results we shall report in Chapters V and VI, for<br />

example those on oscillating chemical reactions, could have<br />

been discovered many years ago, but the study of these non-


ORDER OUT OF CHAOS 20<br />

equilibrium problems was repressed in the cultural and ideological<br />

context of those times.<br />

We are aware that asserting this receptiveness to cultural<br />

content runs counter to the traditional conception of science.<br />

In this view science develops by freeing itself from outmoded<br />

forms of understanding nature; it purifies itself in a process<br />

that can be compared to an "ascesis" of reason. But this in<br />

turn leads to the conclusion that science should be practiced<br />

only by communities living apart, uninvolved in mundane<br />

matters. In this view, the ideal scientific community should be<br />

protected from the pressures, needs, and requirements of society.<br />

Scientific progress ought to be an essentially autonomous<br />

process that any "outside" influence, such as the<br />

scientists's participation in other cultural, social, or economic<br />

activities, would merely disturb or delay.<br />

This ideal of abstraction, of the scientist's withdrawal, finds<br />

an ally in still another ideal, this one concerning the vocation<br />

of a "true" researcher, namely, his desire to escape from<br />

worldly vicissitudes. Einstein describes the type of scientist<br />

who would find favor with the ·ngel of the Lord" should the<br />

latter be given the task of driving from the "Temple of Science"<br />

all those who are "unworthy' ' -it is not stated in what<br />

respects. They are generally<br />

... rather odd, uncommunicative, solitary fellows, who<br />

despite these common characteristics resemble one another<br />

really less than the host of the banished.<br />

What led them into the Te mple? . .. one of the strongest<br />

motives that lead men to art and science is flight<br />

from everyday life with its painful harshness and<br />

wretched dreariness, and from the fetters of one's own<br />

shifting desires. A person with a finer sensibility is driven<br />

to escape from personal existence and to the world of<br />

objective observing (Schauen) and understanding. This<br />

motive can be compared with the longing that irresistibly<br />

pulls the town-dweller away from his noisy, cramped<br />

quarters and toward the silent, high mountains, where<br />

the eye ranges freely through the still, pure air and traces<br />

the calm contours that seem to be made for eternity.<br />

With this negative motive there goes a positive one. Man<br />

seeks to form for himself, in whatever manner is suitable


21 tHE CHALLENGE TO SCIENCE<br />

for him, a simplified and. lucid image of the world (Bild<br />

der Welt), and so to overcome the world of experience by<br />

striving to replace it to some extent by this image.I9<br />

The incompatibility between the ascetic beauty sought after<br />

by science, on the one hand, and the petty swirl of worldly<br />

experience so keenly felt by Einstein, on the other, is likely to<br />

be reinforced by another incompatibility, this one openly Manichean,<br />

between science and society, or, more precisely, be·<br />

tween free human creativity and political power. In this case, it<br />

is not in an isolated community or in a temple that research<br />

would have to be carried out, but in a fortress, or else·in a<br />

madhouse, as Duerrenmatt imagined in his play The Physicists.20<br />

There, three physicists discuss the ways and means of<br />

advancing physics while at the same time safeguarding mankind<br />

from the dire consequences that result when political<br />

powers appropriate the results of its progress. The conclusion<br />

they reach is that the only possible way is that which has already<br />

been chosen by one of them; they all decide to pretend<br />

to be mad, to hide in a lunatic asylum. At the end of the play,<br />

as Fate would have it, this last refuge is discovered to be an<br />

illusion. The director of the asylum, who has been spying on<br />

her patient, steals his results and seizes world power.<br />

Duerrenmatt's play leads to a third conception of scientific<br />

activity: science progresses by reducing the complexity of reality<br />

to a hidden simplicity. What the physicist Moebius is trying<br />

to conceal in the madhouse is the fact that he has<br />

successfully solved the problem of gravitation, the unified theory<br />

of elementary particles, and, ultimately, the Principle of<br />

Universal Discovery, the source of absolute power. Of course,<br />

Duerrenmatt simplifies to make his point, yet it is commonly<br />

held that what is being sought in the "Temple of Science" is<br />

nothing less than the "formula" of the universe. The man of<br />

science, already portrayed as an ascetic, now becomes a kind<br />

of magician, a man apart, the potential holder of a universal<br />

key to all physical phenomena,. thus endowed with a potentially<br />

omnipotent knowledge. This brings us back to an issue<br />

we have already raised: it is only in a simple world (especially<br />

in the world of classical science, where complexity merely<br />

veils a fundamental simplicity) that a form of knowledge that<br />

provides a universal key can exist.


ORDER OUT OF CHAOS 22<br />

One of the problems of our time is to overcome attitudes<br />

that tend to justify and reinforce the isolation of the scientific<br />

community. We must open new channels of communication<br />

between science and society. It is in this spirit that this book<br />

has been written. We all know that man is altering his natural<br />

environment on an unprecedented scale. As Serge Moscovici<br />

puts it, he is creating a "new nature. "21 But to understand this<br />

man-made world, we need a science that is not merely a tool<br />

submissive to external interests, nor a cancerous tumor irresponsibly<br />

growing on a substrate society.<br />

1\.vo thousand years ago Chuang Tsu wrote:<br />

How [ceaselessly] Heaven revolves ! How [constantly]<br />

Earth abides at rest! Do the Sun and the Moon contend<br />

about their respective places? Is there someone presiding<br />

over and directing those things? Who binds and connects<br />

them together? Who causes and maintains them without<br />

trouble or exertion? Or is there perhaps some secret<br />

mechanism in consequence of which they cannot but be<br />

as they are ?22<br />

We believe that we are heading toward a new synthesis, a<br />

new naturalism. Perhaps we will eventually be able to combine<br />

the Western tradition, with its emphasis on experimentation<br />

and quantitative formulations, with a tradition such as the Chinese<br />

one, with its view of a spontaneous, self-<strong>org</strong>anizing<br />

world. Toward the beginning of this Introduction, we cited<br />

Jacques Monod. His conclusion was: "The ancient alliance<br />

has been destroyed; man knows at last that he is alone in the<br />

universe's indifferent immensity out of which he emerged only<br />

by chance. "23 Perhaps Monod was right. The ancient alliance<br />

has been shattered. Our role is not to lament the past. It is to<br />

try to discover in the midst of the extraordinary diversity of<br />

the sciences some unifying thread. Each great period of science<br />

has led to some model of nature. For classical science it<br />

was the clock; for nineteenth-century science, the period of<br />

the Industrial Revolution, it was an engine running down.<br />

What will be the symbol for us? What we have in mind may<br />

perhaps be expressed best by a reference to sculpture, from<br />

Indian or pre-Columbian art to our time. In some of the most


23 THE CHALLENGE TO SCIENCE<br />

beautiful manifestations of sculpture, be it in the dancing<br />

Shiva or in the miniature temples of Guerrero, there appears<br />

very clearly the search for a junction between stillness and<br />

motion, time arrested and time passing. We believe that this<br />

confrontation will give our period its uniqueness.


I<br />

I


BOOK ONE<br />

THE DEWSION OF<br />

THE UNIVERSAL


CHAPTER I<br />

THE TRIUMPH OF<br />

RE ASON<br />

The New Moses<br />

Nature and Natures laws lay hid in night:<br />

God said, let Newton be! and all was light.<br />

-Alexander Pope,<br />

Proposed Epitaph for Isaac 'Newton,<br />

who died in 1727<br />

There is nothing odd in the dramatic tone employedc by Pope.<br />

In the eyes of eighteenth-century England, Newton was the<br />

"new Moses" who had been shown the "tables of the law. "<br />

Poets, architects, and sculptors joined to propose monuments;<br />

a whole nation assembled to celebrate this unique event: a<br />

man had discovered the language that nature speaks-and<br />

obeys.<br />

Nature compelled, his piercing Mind obeys,<br />

And gladly shows him all her secret Ways;<br />

'Gainst Mathematicks she has no Defence,<br />

And yields t'experimental Consequence.1<br />

Ethics and politics drew upon the Newtonian episode for material<br />

on which to "ground" their arguments. Thus Desaguliers<br />

transposed the meaning of the new natural order into<br />

a political lesson: a constitutional monarchy is the best possible<br />

system of government, since the King, like the Sun, has<br />

his power limited by it.<br />

Like Ministers attending ev'ry Glance<br />

Six Worlds sweep round his Throne in Mystick Dance.<br />

27


ORDER OUT OF CHAOS 28<br />

He turns their Motion from his Devious Course,<br />

And bends their Orbits by Attractive Force;<br />

His Pow'r coerc'd by Laws, still leave them free,<br />

Directs, but not Destroys, their Liberty;2<br />

Although he himself did not encroach upon the domain of the<br />

moral sciences, Newton had no hesitation regarding the universal<br />

nature of the laws set out in his Principia. Nature is<br />

"very consonant and conformable to herself," he asserts in<br />

the celebrated Question 31 of his Opticks-and this strong and<br />

elliptical statement conceals a vast claim: combustion, fermentation,<br />

heat, cohesion, magnetism . .. there is no natural<br />

process which would not be produced by these active forcesattractions<br />

and repulsions-that govern both the motion of the<br />

stars and that of freely falling bodies.<br />

Already a national hero before his death, nearly a century later<br />

Newton was to become, mainly through the powerful influence<br />

exerted by Laplace, the symbol of the scientific revolution in<br />

Europe. Astronomers scanned a sky ruled by mathematics.<br />

The Newtonian system succeeded in overcoming all obstacles.<br />

Furthermore, it opened the way to mathematical methods<br />

by which apparent deviations could be accounted for and<br />

even be used to infer the existence of a hitherto unknown<br />

planet. The prediction of the existence of the planet Neptune<br />

was the consecration of the prophetic power inherent in the<br />

Newtonian vision.<br />

At the dawn of the nineteenth century, Newton's name<br />

tended to signify anything that claimed exemplarity. However,<br />

conflicting interpretations of his method are given. Some saw<br />

it as providing a blueprint for quantitative experimentation expressible<br />

in mathematics. For them, chemistry found its Newton<br />

in Lavoisier, who pioneered the systematic use of the<br />

balance. This was indeed a decisive step in the definition of a<br />

quantitative chemistry that took mass conservation as its<br />

Ariadne's thread. According to others, the Newtonian strategy<br />

consisted in isolating some central, specific fact and then<br />

using it as the basis for all further deductions concerning a<br />

given set of phenomena. In this perspective Newton's genius<br />

was located in his pragmatism. He did not try to explain gravitation;<br />

he took it as a fact. Similarly, each discipline should


29<br />

THE TRIUMPH OF REASON<br />

then take some central unexplained fact as its startm!; point.<br />

Physicians thus felt that they were authorized by Newton to<br />

refashion the vitalist conception and to speak of a "vital<br />

force" sui generis, the use of which would give the description<br />

of living phenomena a hoped-for systematic consistency. This<br />

is the same role that affinity, taken as the specificalJy chemical<br />

force of interaction, was calJed upon to play.<br />

Some "true Newtonians" took exception to this proliferation<br />

of forces and reasserted the universality of the explanatory<br />

power of gravitation. But it was too late. The term<br />

Newtonian was now applied to everything that dealt with a<br />

system of Jaws, with equilibrium, or even to all situations in<br />

which natural order on one side and moral, social, and political<br />

order on the other could be expressed in terms of an allembracing<br />

harmony. Romantic philosophers even discovered<br />

in the Newtonian universe an enchanted world animated by<br />

natural forces. More "orthodox" physicists saw in it a mechanical<br />

world governed by mathematics. For the positivists it<br />

meant the success of a procedure, a recipe to be identified<br />

with the very definition of science.3<br />

The rest is literature-often Newtonian literature: the harmony<br />

that reigns in the society of stars, the elective affinities<br />

and hostilities giving rise to the "social life" of chemical compounds<br />

appear as processes that can be transposed into the<br />

world of human society. No wonder that this period appears as<br />

the Golden Age of Classical Science.<br />

Thday Newtonian science still occupies a unique position.<br />

Some of the basic concepts it introduced represent a definitive<br />

acquisition that has survived all the mutations science has<br />

since undergone. However, today we know that the Golden<br />

Age of Classical Science is gone, and with it also the conviction<br />

that Newtonian rationality, even with its various conflicting<br />

interpretations, forms a suitable basis for our dialogue with<br />

nature.<br />

A central subject of this book is that of the Newtonian triumph,<br />

the continual opening up of new fields of investigation<br />

that have extended Newtonian thought right down to the pres­<br />

. ent day. It also deals with doubts and struggles that arose from<br />

this triumph. Today we are beginning to see more clearly the<br />

limits of Newtonian rationality. A more consistent conception


ORDER OUT OF CHAOS 30<br />

of science and of nature seems to be emerging. This new conception<br />

paves the way for a new unity of knowledge and culture.<br />

A Dehumanized World<br />

... May God us keep<br />

From single Vision and Newtons sleep!<br />

-William Blake,<br />

in a letter to Thomas Butts<br />

dated November 22, 1802<br />

There is no better illustration of the instability of the cultural<br />

position of Newtonian science than the introduction to a<br />

UNESCO colloquium on the relationship between science and<br />

culture:<br />

For more than a century the sector of scientific activity<br />

has been growing to such an extent within the surrounding<br />

cultural space that it seems to be replacing the totality<br />

of the culture itself. Some believe that this is merely an<br />

illusion due to its high growth rate and that the lines of<br />

force of this culture will soon reassert themselves and<br />

bring science back into the service of man. Others consider<br />

that the recent triumph of science entitles it at last<br />

to rule over the whole of culture which, moreover, would<br />

deserve to go on being known as such only because it was<br />

transmitted through the scientific apparatus. Others<br />

again, appalled by the danger of man and society being<br />

manipulated if they come under the sway of science, perceive<br />

the spectre of cultural disaster looming in the distance.4<br />

In this statement science appears as a cancer in the body of<br />

culture, a cancer whose proliferation threatens to destroy the<br />

whole of cultural life. The question is whether we can dominate<br />

science and control its development, or whether we shall<br />

be enslaved. In only one hundred fifty years, science has been


31 THE TRIUMPH OF REASON<br />

downgraded from a source of inspiration for Western culture<br />

to a threat. Not only does it threaten man's material existence,<br />

but also, more subtly, it threatens to destroy the traditions and<br />

experiences that are most deeply rooted in our cultural life. It<br />

is not just the technological fallout of one or another scientific<br />

breakthrough that is being accused, but "the spirit of science"<br />

itself.<br />

Whether the accusation refers to a global skepticism exuded<br />

by scientific culture or to specific conclusions reached<br />

through scientific theories, it is often asserted today that science<br />

is debasing our world. What for generations had been a<br />

source of joy and amazement withers at its touch. Everything<br />

it touches is dehumanized.<br />

Oddly enough, the idea of a fatal disenchantment brought<br />

about by scientific progress is an idea held not only by the<br />

critics of science but often also by those who defend or glorify<br />

it. Thus, in his book The Edge of Objectivity, historian C. C.<br />

Gillispie expresses sympathy for those who criticize science<br />

and constantly endeavor to blunt the "cutting edge of objectivity":<br />

Indeed, the renewals of the subjective approach to nature<br />

make a pathetic theme. Its ruins lie strewn like good intentions<br />

aU along the ground traversed by science, until it<br />

survives only in strange corners like Lysenkoism and anthroposophy,<br />

where nature is socialized or moralized.<br />

Such survivals are relics of the perpetual attempt to escape<br />

the consequences of western man's most characteristic<br />

and successful campaign, which must doom to<br />

conquer. So like any thrust in the face of the inevitable,<br />

romantic natural philosophy has induced every nuance of<br />

mood from desperation to heroism. At the ugliest, it is<br />

sentimental or vulgar hostility to intellect. At the noblest,<br />

it inspired Diderot's naturalistic and moralizing science,<br />

Goethe's personification of nature, the poetry of Wordsworth,<br />

and the philosophy of Alfred North Whitehead, or<br />

of any other who would find a place in science for our<br />

qualitative and aesthetic appreciation of nature. It is the<br />

science of those who would make botany of blossoms and<br />

meteorology of sunsets. 5


ORDER OUT OF CHAOS 32<br />

Thus science leads to a tragic, metaphysical choice. Man has<br />

to choose between the reassuring but irrational temptation to<br />

seek in nature a guarantee of human values, or a sign pointing<br />

to a fundamental corelatedness, and fidelity to a rationality<br />

that isolates him in a silent world.<br />

The echoes of another leitmotiv-domination-mingle with<br />

that of disenchantment. A disenchanted world is, at the same<br />

time, a world liable to control and manipulation. Any science<br />

that conceives of the world as being governed according to a<br />

universal theoretical plan that reduces its various riches to the<br />

drab applications of general laws thereby becomes an instrument<br />

of domination. And man, a stranger to the world, sets<br />

himself up as its master.<br />

This disenchantment has taken various forms in recent decades.<br />

It is outside the aim of this book to study systematically<br />

the various forms of antiscience. In Chapter III we shall present<br />

a fuller reaction of Western thought to the surprising triumph<br />

of Newtonian rationality. Here let us only note that at<br />

present there is a shift of popular attitudes to nature associated<br />

with a widespread but in our opinion erroneous belief that<br />

there exists a fundamental antagonism between science and<br />

"naturalism." To illustrate at least some of the forms antiscientific<br />

criticism has taken in recent years, we have chosen<br />

three examples. First, Heidegger, whose philosophy holds a<br />

deep fascination for contemporary thought. We shall also refer<br />

to the criticisms stated by Arthur Koestler and by the great<br />

historian of science, Alexander Koyre.<br />

Martin Heidegger directs his criticism against the very core<br />

of the scientific endeavor, which he sees as fundamentally related<br />

to a permanent aim, the domination of nature. Therefore<br />

Heidegger claims that scientific rationality is the final accomplishment<br />

of something that has been implicitly present since<br />

ancient Greece, namely, the will to dominate, which is at work<br />

in any rational discussion or enterprise, the violence lurking in<br />

all positive and communicable knowledge. Heidegger emphasizes<br />

what he·calls the technological and scientific "framing"<br />

(Geste/1), 6 which leads to the general setting to work of the<br />

world and of men.<br />

Thus Heidegger does not present a detailed analysis of any<br />

particular technological or scientific product or process. What<br />

he challenges is the essence of technology, the way each thing


33 THE TRIUMPH OF REASON<br />

is taken into account. Each theory is part of the implementation<br />

of the master plan that makes up Western history. What<br />

we call a scientific "theory" implies, following Heidegger, a<br />

way of questioning things by which they are reduced to enslavement.<br />

The scientist, like the technologist, is a toy in the<br />

hands of the will to power disguised as thirst for knowledge;<br />

his very approach to things subjects them to systematic violence.<br />

Modern physics is not experimental physics because it<br />

uses experimental devices in its questioning of nature.<br />

Rather the reverse is true. Because physics, already as<br />

pure theory, requests nature to manifest itself in terms of<br />

predictable forces, it sets up the experiments precisely<br />

for the sole purpose of asking whether and how nature<br />

follows the scheme preconceived by science. 7<br />

Similarly, Heidegger is not concerned about the fact that industrial<br />

pollution, for example, has destroyed all animal life in<br />

the Rhine. What does concern him is that the river has been<br />

put to man's service.<br />

The hydroelectric plant is set into the current of the<br />

Rhine. It sets the Rhine to supplying its hydraulic pressure,<br />

which then sets the turbines turning . ... The hydroelectric<br />

plant is not built into the Rhine river as was<br />

the old bridge that joined bank wi'th bank for hundreds of<br />

years. Rather the river is dammed up into the power<br />

plant. What the river is now, namely, a water supplier,<br />

derives from out of the essence of the power station.s<br />

The old bridge over the Rhine is valued not as a proof of<br />

soundly tested ability, of painstaking and accurate observation,<br />

but because it does not "use" the river.<br />

Heidegger's criticisms, taking the very ideal of a positive,<br />

communicable knowledge as a threat, echo some themes of<br />

the antiscience movement to which we referred in the Introduction.<br />

But the idea of an indissociable link between science<br />

and the will to dominate also permeates some apparently very<br />

different assessments of our present-day situation. For instance,<br />

under the very suggestive title "The Coming of the


ORDER OUT OF CHAOS<br />

34<br />

Golden Age, "9 Gunther Stent states that science is now reaching<br />

its limits. We are close to a point of diminishing returns,<br />

where the questions we direct to things in order to master<br />

them become more and more complicated and devoid of interest.<br />

This marks the end of progress, but it is the opportunity<br />

for humanity to stop its frantic efforts, to end the age-old<br />

struggle against nature, and to accept a static and comfortable<br />

peace. We wish to show that the relative dissociation between<br />

the scientific knowledge of an object and the possibility of<br />

mastering it, far from marking the end of science, signals a<br />

host of new perspectives and problems. Scientific understanding<br />

of the world around us is just beginning. There is yet<br />

another idea of science that we feel is potentially just as detrimental,<br />

namely, the fascination with a mysterious science<br />

that, by paths of reasoning inaccessible to common mortals,<br />

will lead to results that can, in one fell swoop, challenge the<br />

meaning of basic concepts such as time, space, causality,<br />

mind, or matter. This kind of "mystery science," the results of<br />

which are imagined to be capable of shattering the framework<br />

of any traditional conception, has actually been encouraged<br />

by the successive "revelations" of relativity and quantum mechanics.<br />

It is certainly true that some of the most imaginative<br />

steps in the past, Einstein's interpretation of gravitation as a<br />

space curvature or Dirac's antiparticles, for example, have<br />

shaken some seemingly well-established conceptions. Thus<br />

there is a very delicate balance between the readiness to imagine<br />

that science can produce anything and a kind of down-toearth<br />

realism. Today the balance is strongly shifting toward a<br />

revival of mysticism, be it in the press media or even in science<br />

itself, especially among cosmologists.IO It has even been suggested<br />

by certain physicists and popularizers of science that<br />

mysterious relationships exist between parapsychology and<br />

quantum physics. Let us cite Koestler:<br />

We have heard a whole chorus of Nobel Laureates in<br />

physics informing us that matter is dead, causality is<br />

dead, determinism is dead. If that is so, let us give them a<br />

decent burial, with a requiem of electronic music. It is<br />

time for us to draw the lesson from twentieth-century<br />

post-mechanistic science, and to get out of the strait-


36<br />

THE TRIUMPH OF REASON<br />

jacket which nineteenth-century materialism imposed on<br />

our philosophical outlook. Paradoxically, had that outlook<br />

kept abreast with modern science itself, instead of<br />

lagging a century behind it, we would have been liberated<br />

from that strait-jacket long ago . ... But once this is recognized,<br />

we might become more receptive to phenomena<br />

around us which one-sided emphasis on physical science<br />

has made us ignore; might feel the draught that is blowing<br />

through the chinks of the causal edifice; pay more attention<br />

to confluential events; include the paranormal phenomena<br />

in our concept of normality; and realise that we<br />

have been living in the "Country of the Blind." I I<br />

We do not wish to judge or condemn a priori. There may be in<br />

some of the apparently fantastic propositions we hear today<br />

some seed of new knowledge. Nevertheless, we believe that<br />

leaps into the unimaginable are far too simple escapes from<br />

the concrete complexity of our world. We do not believe we<br />

shall leave the "Country of the Blind" in a day, since conceptual<br />

blindness is not the main reason for the problems and<br />

contradictions our society has failed to solve.<br />

Our disagreement with certain criticisms or distortions of<br />

science does not mean, however, that we wish to reject aJJ criticisms.<br />

Let us take, for instance, the position of Alexander<br />

Koyre, who has made outstanding contributions to the understanding<br />

of the development of modern science. In his study of<br />

the significance and implications of the Newtonian synthesis,<br />

Koyre wrote:<br />

Ye t there is something for which Newton-or better to<br />

say not Newton alone, but modern science in generalcan<br />

still be made responsible: it is the splitting of our<br />

world in two. I have been saying that modern science<br />

broke down the barriers that separated the heavens and<br />

the earth, and that it united and unified the universe. And<br />

that is true. But, as I have said, too, it did this by substituting<br />

for our world of quality and sense perception,<br />

the world in which we live, and love, and die, another<br />

world-the world of quantity, of reified geometry, a world<br />

in which, though there is a place for everything, there is


ORDER OUT OF CHAOS 36<br />

no place for man. Thus the world of science-the real<br />

world-became estranged and utterly divorced from the<br />

world of life, which science has been unable to explainnot<br />

even to explain away by calling it "subjective."<br />

True, these worlds are everyday-and even more and<br />

more-connected by the praxis. Yet for theory they are<br />

divided by an abyss.<br />

1\vo worlds: this means two truths. Or no truth at all.<br />

This is the tragedy of the modern mind which "solved<br />

the riddle of the universe," but only to replace it by another<br />

riddle: the riddle of itself. I2<br />

However, we hear in the conclusions of Koyre the same<br />

theme expressed by Pascal and Monod-this tragic feeling of<br />

estrangement. Koyre 's criticism does not challenge scientific<br />

thinking but rather classical science based on the Newtonian<br />

perspective. We no longer have to settle for the previous dilemma<br />

of choosing between a science that reduces man to<br />

being a stranger in a disenchanted world and antiscientific,<br />

irrational protests. Koyre's criticism does not invoke the limits<br />

of a "strait-jacket" rationality but only the incapacity of classical<br />

science to deal with some fundamental aspects of the<br />

world in which we live.<br />

Our position in this book is that the science described by<br />

Koyre is no longer our science. Not because we are concerned<br />

today with new, unimaginable objects, closer to magic than to<br />

logic, but because as scientists we are now beginning to find<br />

our way toward the complex processes forming the world with<br />

which we are most familiar, the natural world in which living<br />

creatures and their societies develop. Indeed, today we are beginning<br />

to go beyond what Koyre called "the world of quantity"<br />

into the world of "qualities" and thus of "becoming."<br />

This will be the main subject of Books One and 1\vo. We believe<br />

it is precisely this transition to a new description that<br />

makes this moment in the history of science so exciting. Perhaps<br />

it is not an exaggeration to say that it is a period like the<br />

time of the Greek atomists or the Renaissance, periods in<br />

which a new view of nature was being born. But let us first<br />

return to Newtonian science, certainly one of the great moments<br />

of human history.


37 THE TRIUMPH OF REASON<br />

The Newtonian Synthesis<br />

What lay behind the enthusiasm of Newton's contemporaries,<br />

their conviction that the secret of the universe, the truth about<br />

nature, had finally been revealed? Several lines of thought,<br />

probably present from the very beginning of humanity, converge<br />

in Newton's synthesis: first of all, science as a way of<br />

acting on our environment. Newtonian science is indeed an<br />

active science; one of its sources is the knowledge of the medieval<br />

craftsmen, the knowledge of the builders of machines.<br />

This science provides the means for systematically acting on<br />

the world, for predicting and modifying the course of natural<br />

processes, for conceiving devices that can harness and exploit<br />

the forces and material resources of nature.<br />

In this sense, modern science is a continuation of the ageless.efforts<br />

of man to <strong>org</strong>anize and exploit the world in which<br />

he lives. We have very scanty knowledge about the early<br />

stages of this endeavor. However, it is possible, in retrospect,<br />

to assess the knowledge and skills required for the "Neolithic<br />

Revolution" to take place, when man gradually began to <strong>org</strong>anize<br />

his natural and social environment, using new techniques<br />

to exploit nature and to <strong>org</strong>anize his society. We still use, or<br />

have used until quite recently, Neolithic techniques-for example,<br />

animal and plant species either bred or selected, weaving,<br />

pottery, metalworking. Our social <strong>org</strong>anization was for a<br />

long time based on the same techniques of writing, geometry,<br />

and arithmetic as those required to <strong>org</strong>anize the hierarchically<br />

differentiated and structured social groups of the Neolithic<br />

city-states. Thus we cannot help acknowledging the continuity<br />

that exists between Neolithic techniques and the scientific and<br />

industrial revolutions.t3<br />

Modern science has thus extended this ancient endeavor,<br />

amplifying it and constantly speeding up its rhythm. Nevertheless,<br />

this does not exhaust the significance of science in the<br />

sense given to it by the Newtonian synthesis.<br />

In addition to the various techniques used in a given society,<br />

we find a number of beliefs and myths that seek to understand<br />

man's place in nature. Like myths and cosmologies, science's


ORDER OUT OF CHAOS<br />

38<br />

endeavor is to understand the nature of the world, the way it is<br />

<strong>org</strong>anized, and man's place in it.<br />

From our standpoint it is quite irrelevant that the early speculations<br />

of the pre-Socratics appear to be adapted from the<br />

Hesiodic myth of creation-that is, the initial polarization of<br />

Heaven and Earth, the desire aroused by Eros, the birth of the<br />

first generations of gods to fo rm the differentiated cosmic<br />

powers, discord and strife, alternating atrocities and vendettas,<br />

until stability is finally reached under the rule of Justice<br />

(dike). What does matter is that, in the space of a few generations,<br />

the pre-Socratics collected, discussed, and criticized<br />

some of the concepts we are still trying to <strong>org</strong>anize in order to<br />

understand the relation between being and becoming, or the<br />

appearance of order out of a hypothetically undifferentiated<br />

initial environment.<br />

Where does the instability of the homogeneous come from?<br />

Why does it differentiate spontaneously? Why do things exist<br />

at all? Are they the fragile and mortal result of an injustice, a<br />

disequilibrium in the static equilibrium of forces between conflicting<br />

natural powers? Or do the forces that create and drive<br />

things exist autonomously-rival powers of love and hate leading<br />

to birth, growth, decline, and dispersion? Is change an illusion<br />

or is it, on the contrary, the unceasing struggle between<br />

opposites that constitutes things? Can qualitative change be<br />

reduced to the motion in a vacuum, of atoms differing only in<br />

their forms, or do atoms themselves consist of a multitude of<br />

qualitatively different germs, each unlike the others? And last,<br />

is the harmony of the world mathematical? Are numbers the<br />

key to nature?<br />

The numerical regularities among sounds that were discovered<br />

by the Pythagoreans are still part of our present theories.<br />

The mathematical schemes worked out by the Greeks<br />

form the first body of abstract thought in European historythat<br />

is, a thought whose results are communicable and reproducible<br />

for all reasoning human beings. The Greeks<br />

achieved for the first time a form of deductive knowledge that<br />

contained a degree of certainty unaffected by convictions, expectations,<br />

or passions.<br />

The most important aspect common to Greek thought and<br />

to modern science , which contrasts with the religious and


39<br />

THE TRIUMPH OF REASON<br />

mythicaI form of inquiry, is thus the emphasis on criticaI discussion<br />

and verification.14<br />

Little is known about this pre-Socratic philosophy that grew<br />

up in the lonian cities and the colonies of Magna Graecia.<br />

Thus we can only speculate about the relationships that might<br />

have existed between the development of theoretical and cosmological<br />

hypotheses and the crafts and technological activities<br />

that tlourished in those cities. Tr adition teUs that as a<br />

result of a hostile religious and social reaction, philosophers<br />

were accused of atheism and were either exiled or put to death.<br />

This early "recall to order" may serve as a symbol of the importance<br />

of social factors in the origin, and above alI the<br />

growth, of conceptual innovations. To understand the success<br />

of modem science we also have to explain why its fóunders<br />

were as a rule not unduly persecuted and their theoretical approach<br />

repressed in favor of a form of knowledge more consistent<br />

with social anticipations and convictions.<br />

Be that as it may, from Plato and Aristotle onward, the limits<br />

were set, and thought was channeled in socially acceptable<br />

directions. ln particular, the distinction between theoretical<br />

thinking and technological activity was established. The<br />

words we still use today-machine, mechanical, engineerhave<br />

a similar meaning. They do not refer to rational knowledge<br />

but to cunning and expediency. The idea was not to leam<br />

about natural processes in order to utilize them more effectively,<br />

but to deceive nature, to "machinate" against it-that<br />

is, to work wonders and create effects extraneous to the "natural<br />

order" of things. The fields of practical manipulation and<br />

that of the rational understanding of nature were thus rigidly<br />

separated. Archimedes' status is merely that of an engineer;<br />

his mathematical analysis of the equilibrium of machines is not<br />

considered to be applicable to the world of nature, at least<br />

within the framework of traditional physics. ln contrast, the<br />

Newtonian synthesis expresses a systematic alliance between<br />

manipulation and theoretical understanding.<br />

There is a third important element that found its expression<br />

in the Newtonian revolution. There is a striking contrast,<br />

which each of us has probably experienced, between the quiet<br />

world of the stars and planets and thé ephemeral, turbulent<br />

world around uso As Mircea Eliade has emphasized, in many


ORDER OUT OF CHAOS 40<br />

ancient civilizations there is a separation between profane<br />

space and sacred space, a division of the world into an ordi·<br />

nary space that is subject to chance and degradation and a<br />

sacred one that is meaningful, independent of contingency and<br />

history. This was the very contrast Aristotle established between<br />

the world of the stars and our sub lunar world. This contrast<br />

is crucial to the way in which Aristotle evaluated the<br />

possibility of a quantitative description of nature. Since the<br />

motion of the celestial bodies is not change but a "divine"<br />

state that is eternally the same, it ntay be described by means<br />

of mathematical idealizations. Mathematical precision and<br />

rigor are not relevant to the sub lunar world. Imprecise natural<br />

processes can only be subjected to an approximate description.<br />

In any case, for an Aristotelian it is more interesting to<br />

know why a process occurs than to describe how it occurs, or<br />

rather, these two aspects are indivisible. One of the main<br />

sources of Aristotle's thinking was the observation of embryonic<br />

growth, a highly <strong>org</strong>anized process in which interlocking,<br />

although apparently independent, events participate in a<br />

process that seems to be part of some global plan. Like the developing<br />

embryo, the whole of Aristotelian nature is <strong>org</strong>anized<br />

according to final causes. The purpose of all change, if it is in<br />

keeping with the nature of things, is to realize in each being the<br />

perfection of its intelligible essence. Thus this essence, which,<br />

in the case of living creatures, is at one and the same time their<br />

final, formal, and effective cause, is the key to the understanding<br />

of nature. In this sense the "birth of modern science," the<br />

clash between the Aristotelians and Galileo, is a clash between<br />

two forms of rationality. IS<br />

In Galileo 's view the question of "why," so dear to the Aristotelians,<br />

was a very dangerous way of addressing nature, at<br />

least for a scientist. The Aristotelians, on the other hand, considered<br />

Galileo's attitude as a form of irrational fanaticism.<br />

Thus, with the coming of the Newtonian system it was a<br />

new universality that triumphed, and its emergence unified<br />

what till then had appeared as divided.


41 THE TRIUMPH OF REASON<br />

The Experimental Dialogue<br />

We have already emphasized one of the essential elements of<br />

modern science: the marriage between theory and practice,<br />

the blending of the desire to shape the world and the desire to<br />

understand it. For this to be possible, it was not enough, despite<br />

the empiricists' beliefs, merely to respect observed facts.<br />

On certain points, including even the description of mechanical<br />

motion, it was in fact Aristotelian physics that was more<br />

easily brought into contact with empirical facts. The experimental<br />

dialogue with nature discovered by modern science involves<br />

activity rather than passive observation. What must be<br />

done is to manipulate physical reality, to "stage" it in such a<br />

way that it conforms as closely as possible to a theoretical<br />

description. The phenomenon studied must be prepared and<br />

isolated until it approximates some ideal situation that may be<br />

physically unattainable but that conforms to the conceptual<br />

scheme adopted.<br />

By way of example, let us take the description of a system of<br />

pulleys, a classic since the time of Archimedes, whose reasoning<br />

has been extended by modern scientists to cover all simple<br />

machines. It is astonishing to find that the modern explanation<br />

has eliminated, on the grounds that it is irrelevant, the very<br />

thing that Aristotelian physics set out to explain, namely, the<br />

fact that, using a typical image, a stone "resists" a horse's<br />

efforts to pull it and that this resistance can be "overcome" by<br />

applying traction through a system of pulleys. Nature, according<br />

to Galileo, never gives anything away, never does something<br />

for nothing, and can never be tricked; it is absurd to<br />

think that by cunning or by using some stratagem we can make<br />

it perform extra work.I6 Since the work the horse is able to<br />

perform is the same with or without the pulleys, the effect<br />

produced must be the same. This then becomes the starting<br />

point for a mechanical explanation, which thus refers to an<br />

idealized world. In this world the "new" effect-the stone finally<br />

set in motion-is of secondary importance; and the<br />

stone's resistance is described only qualitatively, in terms of<br />

friction and heating. Instead, what is described accurately is<br />

the ideal situation, in which a relationship of equivalence links


ORDER OUT OF CHAOS 42<br />

the cause, the work done by the horse, to the effect, the mo<br />

tion of the stone. In this ideal world, the horse can, in any<br />

case, shift the stone, and the system of pulleys has the sole<br />

effect of modifying the way the pulling efforts are transmitted;<br />

instead of moving the stone over a distance L, equal to the<br />

distance it travels while pulling the rope, the horse only moves<br />

it over a distance Lin, where n depends on the number of<br />

pulleys. Like all simple machines, the pulleys form a passive<br />

device that can only transmit motion without producing it.<br />

The experimental dialogue thus corresponds to a highly specific<br />

procedure. Nature is cross-examined through experimentation,<br />

as if in a court of law, in the name of a priori principles.<br />

Nature's answers are recorded with the utmost accuracy, but<br />

relevance of those answers is assessed in terms of the very<br />

idealizations that guided the experiment. All the rest does not<br />

count as information, but is idle chatter, negligible secondary<br />

effects. It may well be that nature rejects the theoretical hypothesis<br />

in question. Nevertheless, the latter is still used as a<br />

standard against which to measure the implications and the<br />

significance of the response, whatever it may be. It is precisely<br />

this imperative way of questioning nature that Heidegger refers<br />

to in his argument against scientific rationality.<br />

For us the experimental method is truly an art-that is, it is<br />

based on special skills and not on general rules. As such there<br />

are never any guarantees of success and one always remains at<br />

the mercy of triviality or poor judgment. No methodological<br />

principle can eliminate the risk, for instance, of persisting in a<br />

blind alley of inquiry. The experimental method is the art of<br />

choosing an interesting question and of scanning all the consequences<br />

of the theoretical framework thereby implied, all<br />

the ways nature could answer in the theoretical language<br />

chosen. Amid the concrete complexity of natural phenomena,<br />

one phenomenon has to be selected as the most likely to embody<br />

the theory's implications in an unambiguous way. This<br />

phenomenon will then be abstracted from its environment and<br />

"staged" to allow the theory to be tested in a reproducible and<br />

communicable way.<br />

Although this experimental procedure was criticized right<br />

from the outset, ignored by the empiricists, and attacked by<br />

others on the grounds that it was a kind of torture, a way of<br />

putting nature on the rack, it survived all the modifications of


43<br />

THE TRIUMPH OF REASON<br />

the theoretical content of scientific descriptions and ultimately<br />

defined the new method of investigation introduced by modern<br />

science.<br />

Experimental procedure can even become a tool for purely<br />

theoretical analysis. It is then a "thought experiment," the<br />

imagining of experimental situations governed entirely by theoretical<br />

principles, which permits the exploration of the consequences<br />

of these principles in a given situation. Such<br />

thought experiments played a crucial role in Galileo's work,<br />

and today they are at the center of investigations about the<br />

consequences of the conceptual upheavals in contemporary<br />

physics, namely, relativity and quantum mechanics. One of<br />

the most famous of such thought experiments is Einstein's famous<br />

train, from which an observer can measure the velocity<br />

of propagation of a ray of light emitted along an embankment,<br />

that is, moving at a velocity c in a reference system with respect<br />

to which the train is moving at a velocity v. According to<br />

classical reasoning, the observer on the train should attribute<br />

to the light, which is traveling in the same direction as he is, a<br />

velocity of c- v. However, this classical conclusion represents<br />

precisely the absurdity that the thought experiment was designed<br />

to expose. In relativity theory, the velocity of light appears<br />

as a universal constant of nature. Whatever inertial<br />

referer.ce system is used, the velocity of light is always the<br />

same. And since then Einstein's train has gone on exploring<br />

the physical consequences of this fundamental change.<br />

The experimental method is central to the dialogue with nature<br />

established by modern science. Nature questioned in this<br />

way is, of course, simplified and occasionally mutilated. This<br />

does not deprive it of its capacity to refute most of the hypotheses<br />

we can imagine. Einstein used to say that nature says<br />

"no" to most of the questions it is asked, and occasionally<br />

"perhaps." The scientist does not do as he pleases, and he<br />

cannot force nature to say only what he wants to hear. He<br />

cannot, at least in the long run, project upon it his most cherished<br />

desires and expectations. He actually runs a greater risk<br />

and plays a more dangerous game the better his tactics succeed<br />

in encircling nature, in setting it more squarely with its<br />

back to the wall.l7 Moreover, it is true that, whether the answer<br />

is "yes" or "no," it will be expressed in the same theoretical<br />

language as the question. However, this language, too,


ORDER OUT OF CHAOS 44<br />

develops according to a complex historical process involving<br />

nature's replies in the past and its relations with other theoretical<br />

languages. In addition, new questions arise corresponding<br />

to the changing interests of each period. This sets up a complex<br />

relationship between the specific rules of the scientific<br />

game-particularly the experimental method of reasoning<br />

with nature, which places the greatest constraint on the<br />

game-and a cultural network to which, sometimes unwittingly,<br />

the scientist belongs.<br />

We believe that the experimental dialogue is an irreversible<br />

acquisition of human culture. It actually provides a guarantee<br />

that when nature is explored by man it is treated as an independent<br />

being. It forms the basis of the communicable and<br />

reproducible nature of scientific results. However partially nature<br />

is allowed to speak, once it has expressed itself, there is<br />

no further dissent: nature never lies.<br />

The Myth at the Origin of Science<br />

The dialogue between man and nature was accurately perceived<br />

by the founders of modern science as a basic step toward<br />

the intelligibility of nature. But their ambitions went even<br />

farther. Galileo, and those who came after him, conceived of<br />

science as being capable of discovering global truths about<br />

nature. Nature not only would be written in a mathematical<br />

language that can be deciphered by experimentation, but there<br />

would actually exist only one such language. Following this<br />

basic conviction, the world is seen as homogeneous, and local<br />

experimentation can reveal global truth. The simplest phenomena<br />

studied by science can thus be interpreted as the key<br />

to understanding nature as a whole; the complexity of the latter<br />

is only apparent, and its diversity can be explained in terms<br />

of the universal truth embodied, in Galileo's case, in the mathematical<br />

laws of motion.<br />

This conviction has survived centuries. In an excellent set<br />

of lectures presented on the BBC several years ago, Richard<br />

Feynman ts compared nature to a huge chess game. The complexity<br />

is only apparent; each move follows simple rules. In its<br />

early days, modern science quite possibly needed this convic-


45 THE TRIUMPH OF REASON<br />

tion of being able to reach global truth. Such a conviction<br />

added an immense value to the experimental method and, to a<br />

certain extent, inspired it. Perhaps a revolutionary conception<br />

of the world, one as all-embracing as the "biological" conception<br />

of the Aristotelian world, was necessary to throw off<br />

the yoke of tradition, to give the champions of experimentation<br />

a strength of conviction and a power of argument that enabled<br />

them to hold their own against the previous forms of rationalism.<br />

Perhaps a metaphysical conviction was needed to transmute<br />

the craftsman's and machine builder's knowledge into a<br />

new method for the rational exploration of nature. We may<br />

also wonder what the implications of the existence of this kind<br />

of "mythical" conviction are for explaining the way modern<br />

science's first developments were accepted in the social context.<br />

On this highly controversial issue, we shall restrict ourselves<br />

to a few remarks of a quite general nature for the sole<br />

purpose of pinpointing the problem-that is, the problem of a<br />

science whose advance has been felt by some as the triumph<br />

of reason, but by others as a disillusionment, as the painful<br />

discovery of the robotlike stupidity of nature.<br />

It seems hard to deny the fundamental importance of social<br />

and economic factors-particularly the development of craftsmen's<br />

techniques in the monasteries, where the residual knowledge<br />

of a destroyed world was preserved, and later in the<br />

bustling merchant cities-in the birth of experimental science,<br />

which is a systematized form of part of the craftsmen's knowledge.<br />

Moreover, a comparative analysis such as Needham'sl9 exposes<br />

the decisive importance of social structures at the close<br />

of the Middle Ages. Not only was the class of craftsmen and<br />

potential technical innovators not held in contempt, as it was<br />

in ancient Greece, but, like the craftsmen, the intellectuals<br />

were, in the main, independent of the authorities. They were<br />

free entrepreneurs, craftsmen-inventors in search of patronage,<br />

who tended to look for novelty and to exploit all the opportunities<br />

it afforded, however dangerous they may have been<br />

for the social order. On the other hand, as Needham points<br />

out, Chinese men of science were officials, bound to observe<br />

the rules of the bureaucracy. They formed an integral part of<br />

the state, whose primary objective was to keep law and order.<br />

The compass, the printing press, and gunpowder, all of which


ORDER OUT OF CHAOS 46<br />

were to contribute to undermining the foundations of medieval<br />

society and to project Europe into the modern era, were discovered<br />

much earlier in China but had a much less destabilizing<br />

effect on its society. The enterprising European merchant<br />

society appears in contrast as particularly well suited to stimulate<br />

and sustain the dynamic and innovative growth of modern<br />

science in its early stages.<br />

However, the question remains. We know that the builders<br />

of machines used mathematical concepts-gear ratios, the displacements<br />

of the various working parts, and the geometry of<br />

their relative motions. But why was mathematization not restricted<br />

to machines? Why was natural motion conceived of in<br />

the image of a rationalized machine? This question may also<br />

be asked in connection with the clock, one of the triumphs of<br />

medieval craftsmanship that was soon to set the rhythm of life<br />

in the larger medieval towns. Why did the clock almost immediately<br />

become the very symbol of world order? In this last<br />

question lies perhaps some elements of an answer. A watch is<br />

a contrivance governed by a rationality that lies outside itself,<br />

by a plan that is blindly executed by its inner workings. The<br />

clock world is a metaphor suggestive of God the Watchmaker,<br />

the rational master of a robotlike nature. At the origin of modern<br />

science, a "resonance" appears to have been set up between<br />

theological discourse and theoretical and experimental<br />

activity-a resonance that was no doubt likely to amplify and<br />

consolidate the claim that scientists were in the process of discovering<br />

the secret of the "great machine of the universe."<br />

Of course, the term resonance covers an extremely complex<br />

problem. It is not our intention to state, nor are we in any<br />

position to affirm, that religious discourse in any way determined<br />

the birth of theoretical science, or of the "world view"<br />

that happened to develop in conjunction with experimental activity.<br />

By using the term resonance-that is, mutual amplification<br />

of two discourses-we have deliberately chosen an<br />

expression that does not assume whether it was theological<br />

discourse or the "scientific myth" that came first and triggered<br />

the other.<br />

Let us note that to some philosophers the question of the<br />

"Christian origin" of Western science is not only the question<br />

of the sta bilization of the concept of nature as an automaton,<br />

but also the question of some "essential" link between experi-


47 THE TRIUMPH OF REASON<br />

mental science as such and Western civilization in its Hebraic<br />

and Greek components. For Alfred North Whitehead this link<br />

is situated at the level of instinctive conviction. Such a conviction<br />

was "needed" to inspire the "scientific faith" of the<br />

founders of modern science:<br />

I mean the inexpugnable belief that every detailed occurrence<br />

can be correlated with its antecedents in a perfectly<br />

definite manner, exemplifying general principles. Without<br />

this belief the incredible labours of scientists would be<br />

without hope. It is this instinctive conviction, vividly<br />

poised before the imagination, which is the motive power<br />

of research: that there is a secret, a secret which can be<br />

unveiled. How has this conviction been so vividly implanted<br />

in the European mind?<br />

When we compare this tone of thought in Europe with<br />

the attitude of other civilizations when left to themselves,<br />

there seems but one source for its origin. It must come<br />

from the medieval insistence on the rationality of God,<br />

conceived as with the personal energy of Jehovah and<br />

with the rationality of a Greek philosopher. Every detail<br />

was supervised and ordered: the search into nature could<br />

only result in the vindication of the faith in rationality.<br />

. Remember that I am not talking of the explicit beliefs of a<br />

few individuals. What I mean is the impress on the European<br />

mind arising from the unquestioned faith of centuries.<br />

By this I mean the instinctive tone of thought and<br />

not a mere creed of words. 2 o<br />

We will not consider this matter further. It would be out of<br />

the question to .. prove" that modern science could have originated<br />

only in Christian Europe. It is not even necessary to ask<br />

if the founders of modern science drew any real inspiration<br />

from theological arguments. Whether or not they were sincere,<br />

the important point is that those arguments made the<br />

speculations of modern science socially credible and acceptable,<br />

over a period of time varying from country to country.<br />

Religious references were still frequent in English scientific<br />

texts of the nineteenth century. Remarkably enough, in the<br />

present-day revival of interest in mysticism, the direction of<br />

the argument appears reversed. It is now science that appears<br />

to lend credibility to mystical affirmation.


ORDER OUT OF CHAOS<br />

48<br />

The question we have confronted here obviously leads to·<br />

ward a multitude of problems in which theological and scien·<br />

tific issues are inextricably bound up with the "external"<br />

history of science, that is, the description of the relationship<br />

between the form and content of scientific knowledge on the<br />

one hand, and on the other, the use to which it is put in its<br />

social, economic, and institutional context. As we have already<br />

said, the only point we are presently interested in is the<br />

very particular character and implications of scientific dis·<br />

course that was amplified by resonance with theological discourses.<br />

Needham 2t tells of the irony with which Chinese men of let·<br />

ters of the eighteenth century greeted the Jesuits' announcement<br />

of the triumphs of modern science. The idea that nature<br />

was governed by simple, knowable laws appeared to them as a<br />

perfect example of anthropocentric foolishness. Needham believes<br />

that this "foolishness" has deep cultural roots. In order<br />

to illustrate the great differences between the Western and<br />

Chinese conceptions, he cites the animal trials held in the<br />

Middle Ages. On several occasions such freaks as a cock who<br />

supposedly laid eggs were solemnly condemned to death and<br />

burned for having infringed the laws of nature, which were<br />

equated with the laws of God. Needham explains how, in<br />

China, the same cock would, in all likelihood, merely have<br />

disappeared discreetly. It was not guilty of any crime, but its<br />

freakish behavior clashed with natural and social harmony.<br />

The governor of the province or even the emperor might find<br />

himself in a delicate situation if the misbehavior of the cock<br />

became known. Needham comments that, according to a<br />

philosophic conception dominant in China, the cosmos is in<br />

spontaneous harmony and the regularity of phenomena is not<br />

due to any external authority. On the contrary, this harmony in<br />

nature, society, and the heavens originates from the equilibrium<br />

among these processes. Stable and interdependent,<br />

they resonate with each other in a kind of nonconcerted harmony.<br />

If any law were involved, it would be a law that no one,<br />

neither God nor man, had ever conceived of. Such a law would<br />

also have to be expressed in a language undecipherable by<br />

man and not be a law established by a creator conceived in our<br />

own image.


49 THE TRIUMPH OF REASON<br />

Needham concludes by asking the following question:<br />

In the outlook of modern science there is, of course, no<br />

residue of the notions of command and duty in the<br />

"Laws" of Nature. They are now thought of as statistical<br />

regularities, valid only in given times and places, descriptions<br />

not prescriptions, as Karl Pearson put it in a famous<br />

chapter. The exact degree of subjectivity in the formulations<br />

of scientific law has been hotly debated during the<br />

whole period from Mach to Eddington, and such questions<br />

cannot be followed further here. The problem is<br />

whether the recognition of such statistical regularities<br />

and their mathematical expression could have been<br />

reached by any other road than that which Western science<br />

actually travelled. Was perhaps the state of mind in<br />

which an egg-laying cock could be prosecuted at law necessary<br />

in a culture which should later have the property<br />

of producing a Kepler? 22<br />

It must now be stressed that scientific discourse is in no way<br />

a mere transposition of traditional religious views. Obviously<br />

the world described by classical physics is not the world of<br />

Genesis, in which God created light, heaven, earth, and the<br />

living species, the world ·where Providence has never ceased<br />

to act, spurring man on toward a history where his salvation is<br />

at stake. The world of classical physics is an atemporal world<br />

which, if created, must have been created in one fell swoop,<br />

somewhat as an engineer creates a robot before letting it function<br />

alone. In this sense, physics has indeed developed in opposition<br />

to both religion and the traditional philosophies. And<br />

yet we know that the Christian God was actually called upon<br />

to provide a basis for the world's intelligibility. In fact, one can<br />

speak here of a kind of "convergence" between the interests<br />

of theologians, who held that the world had to acknowledge<br />

God's omnipotence by its total submission to Him, and of<br />

physicists seeking a world of mathematizable processes.<br />

In any case, the Aristotelian world destroyed by modern science<br />

was unacceptable to both these theologians and physicists.<br />

This ordered, harmonious, hierarchical, and rational world<br />

was too independent, the beings inhabiting it too powerful and


ORDER OUT OF CHAOS 50<br />

active, and their subservience to the absolute sovereign too<br />

suspect and limited for the needs of many theologians. 23 On<br />

the other hand, it was too complex and qualitatively differentiated<br />

to be mathematized.<br />

The "mechanized" nature of modern science, created and<br />

ruled according to a plan that totally dominates it, but of which<br />

it is unaware , glorifies its creator, and was thus admirably<br />

suited to the needs of both theologians and the physicists. Although<br />

Leibniz had endeavored to demonstrate that mathematization<br />

is compatible with a world that can display active and<br />

qualitatively differentiated behavior, scientists and theologians<br />

joined forces to describe nature as a mindless, passive mechanics<br />

that was basically alien to freedom and the purposes<br />

of the human mind. 'i\ dull affair, soundless, scentless, colourless,<br />

merely the hurrying of matter, endless, meaningless, "24<br />

as Whitehead observes. This Christian nature, stripped of any<br />

property that permits man to identify himself with the ancient<br />

harmony of natural "becoming," leaving man alone, face to<br />

face with God, thus converged with the nature that a single<br />

language, and not the thousand mathematical voices heard by<br />

Leibniz, was sufficient to describe.<br />

Theology may also help comment on man's odd position<br />

when he laboriously deciphers the laws governing the world.<br />

Man is emphatically not part of the nature he objectively describes;<br />

he dominates it from the outside. Indeed, for Galileo,<br />

the human soul, created in God's image, is capable of grasping<br />

the intelligible truths underlying the plan of creation. It can<br />

thus gradually approach a knowledge of the world that God<br />

himself possessed intuitively, fully, and instantaneously. 25<br />

Unlike the ancient atomists, who were persecuted on the<br />

grounds of atheism, and unlike Leibniz, who was sometimes<br />

suspected of denying the existence of grace or of human freedom,<br />

modern scientists have managed to come up with a<br />

culturally acceptable definition of their enterprise. The human<br />

mind, incorporated in a body subject to the laws of nature,<br />

can, by means of experimental devices, obtain access to the<br />

vantage point from which God himself surveys the world, to<br />

the divine plan of which this world is a tangible expression.<br />

Nevertheless, the mind itself remains outside the results of its<br />

achievement_ The scientist may descr ibe as secondary<br />

qualities, not part of nature but projected onto it by the mind,


51 THE TRIUMPH OF REASON<br />

everything that goes to make up the texture of nature, such as<br />

its perfumes and its colors. The debasement of nature is parallel<br />

to the glorification of all that eludes it, God and man.<br />

The Limits of Classical Science<br />

We have tried to describe the unique historical situation in<br />

which scientific practice and metaphysical conviction were<br />

closely coupled. Galileo and those who came after him raised<br />

the same problems as the medieval builders but broke away<br />

from their empirical knowledge to assert, with the help of<br />

God, the simplicity of the world and the universality of the<br />

language the experimental method postulated and deciphered.<br />

In this way, the basic myth underlying modern science can be<br />

seen as a product of the peculiar complex which, at the close<br />

of the Middle Ages, set up conditions of resonance and reciprocal<br />

amplification among economic, political, social, religious,<br />

philosophic, and technical factors. However, the rapid<br />

decomposition of this complex left classical science stranded<br />

and isolated in a transformed culture.<br />

Classical science was born in a culture dominated by the<br />

alliance between man, situated midway between the divine<br />

order and the natural order, and God, the rational and intelligible<br />

legislator, the sovereign architect we have conceived in our<br />

own image. It has outlived this moment of cultural consonance<br />

that entitled philosophers and theologians to engage in science<br />

and that entitled scientists to decipher and express opinions<br />

on the divine wisdom and power at work in creation. With the<br />

support of religion and philosophy, scientists had come to believe<br />

their enterprise was self-sufficient, that it exhausted the<br />

possibilities of a rational approach to natural phenomena. The<br />

relationship between scientific description and natural philosophy<br />

did not, in this sense, have to be justified. It could be<br />

seen as self-evident that science and philosophy were convergent<br />

and that science was discovering the principles of an<br />

authentic natural philosophy. But, oddly enough, the selfsufficiency<br />

experienced by scientists was to outlive the departure<br />

of the medieval God and the withdrawal of the epistemological<br />

guarantee offered by theology. The originally bold bet had become<br />

the triumphant science of the eighteenth century, 26 the


ORDER OUT OF CHAOS 52<br />

science that discovered the laws governing the motion of celestial<br />

and earthly bodies, a science that df\lembert and Euler<br />

incorporated into a complete and consistent system and<br />

whose history was defined by Lagrange as a logical achievement<br />

tending toward perfection. It was the science honored by<br />

the Academies founded by absolute monarchs such as Louis<br />

XIV, Frederick II, and Catherine the Great,27 the science that<br />

made Newton a national hero. In other words, it was a successful<br />

science, convinced that it had proved that nature is<br />

transparent. "Je n'ai pas besoin de cette hypothese" was<br />

Laplace's reply to Napoleon, who had asked him God's place<br />

in his world system.<br />

The dualist implications of modern science were to survive<br />

as well as its claims. For the science of Laplace which, in<br />

many respects, is still the classical conception of science today,<br />

a description is objective to the extent to which the observer<br />

is excluded and the description itself is made from a<br />

point lying de jure outside the world, that is, from the divine<br />

viewpoint to which the human soul, created as it was in God's<br />

image, had access at the beginning. Thus classical science still<br />

aims at discovering the unique truth about the world, the one<br />

language that will decipher the whole of nature-today we<br />

would speak of the fundamental level of description from ·<br />

which everything in existence can be deduced.<br />

On this essential point let us cite Einstein, who has translated<br />

into modern terms precisely what we may call the basic<br />

myth underlying modern science:<br />

What place does the theoretical physicist's picture of the<br />

world occupy among all these possible pictures? It demands<br />

the highest possible standard of rigorous precision<br />

in the description of relations, such as only the use of<br />

mathematical language can give. In regard to his subject<br />

matter, on the other hand, the physicist has to limit himself<br />

very severely: he must content himself with describing<br />

the most simple events which can be brought within<br />

the domain of our experience; all events of a more complex<br />

order are beyond the power of the human intellect to<br />

reconstruct with the subtle accuracy and logical perfection<br />

which the theoretical physicist demands. Supreme<br />

purity, clarity, and certainty at the cost of completeness.


53 THE TRIUMPH OF REASON<br />

But what can be the attraction of getting to know such a<br />

tiny section of nature thoroughly, while one leaves everything<br />

subtler and more complex shyly and timidly alone?<br />

Does the product of such a modest effort deserve to be<br />

called by the proud name of a theory of the universe?<br />

In my belief the name is justified; for the general laws<br />

on which the structure of theoretical physics is based<br />

claim to be valid for any natural phenomenon whatsoever.<br />

With them, it ought to be possible to arrive at the<br />

description, that is to say, the theory, of every natural<br />

process, including life, by means of pure deduction, if<br />

that process of deduction were not far beyond the capacity<br />

of the human intellect. The physicist's renunciation of<br />

completeness for his cosmos is therefore not a matter of<br />

fundamental principle. 2s<br />

For some time there were those who persisted in the illusion<br />

that attraction in the form in which it is expressed in the law of<br />

gravitation would justify attributing an intrinsic animation to<br />

nature and that if it were generalized it would explain the origins<br />

of increasingly specific forms of activity, including even<br />

the interactions that compose human society. But this hope<br />

was rapidly crushed, at least partly as a consequence of the<br />

demands created by the political, economic, and institutional<br />

setting where science developed. We shall not examine this<br />

aspect of the problem, important though it is. Our point here is<br />

to emphasize that this very failure seemed to establish the<br />

consistency of the classical view and to prove that what had<br />

once been an inspiring conviction was a sad truth. In fact, the<br />

only interpretation apparently capable of rivaling this interpretation<br />

of science was henceforth the positivistic refusal of<br />

the very project of understanding the world. For example,<br />

Ernst Mach, the influential philosopher-scientist whose ideas<br />

had a great impact on the young Einstein, defined the task of<br />

scientific knowledge as arranging experience in as economical<br />

an order as possible. Science has no other meaningful goal<br />

than the simplest and most economical abstract expression of<br />

facts:<br />

Here we have a clue which strips science of all its mystery,<br />

and shows us what its power really is. With respect


ORDER OUT OF CHAOS<br />

54<br />

to specific results it yields us nothing that we could not<br />

reach in a sufficiently long time without methods . ...<br />

Just as a single human being, restricted wholly to the<br />

fruits of his own labor, could never amass a fortune, but<br />

on the contrary the accumulation of the labor of many<br />

men in the hands of one is the foundation of wealth and<br />

power, so, also, no knowledge worthy of the name can be<br />

gathered up in a single human mind limited to the span of<br />

a human life and gifted only with finite powers, except by<br />

the most exquisite economy of thought and by the careful<br />

amassment of the economically ordered experience of<br />

thousands of co-workers. 29<br />

Thus science is useful because it leads to economy of<br />

thought. There may be some element of truth in such a statement,<br />

but does it tell the whole story? How far we have come<br />

from Newton, Leibniz, and the other founders of Western science,<br />

whose ambition was to provide an intelligible frame to<br />

the physical universe! Here science leads to interesting rules<br />

of action, but no more.<br />

This brings us back to our starting point, to the idea that it is<br />

classical science, considered for a certain period of time as<br />

the very symbol of cultural unity, and not science as such that<br />

led to the cultural crisis we have described. Scientists found<br />

themselves reduced to a blind oscillation between the thunderings<br />

of .. scientific myth" and the silence of "scientific seriousness,"<br />

between affirming the absolute and global nature of<br />

scientific truth and retreating into a conception of scientific<br />

theory as a pragmatic recipe for effective intervention in natural<br />

processes.<br />

As we have already stated, we subscribe to the view that<br />

classical science has now reached its limit. One aspect of this<br />

transformation is the discovery of the limitations of classical<br />

concepts that imply that a knowledge of the world "as it is"<br />

was possible. The omniscient beings, Laplace's or Maxwell's<br />

demon, or Einstein's God, beings that play such an important<br />

role in scientific reasoning, embody the kinds of extrapolation<br />

physicists thought they were allowed to make. As randomness,<br />

complexity, and irreversibility enter into physics as objects<br />

of positive knowledge, we are moving away from this<br />

rather naive assumption of a direct connection between our


55<br />

THE TRIUMPH OF REASON<br />

description of the world and the world itself. Objectivity in<br />

theoretical physics takes on a more subtle meaning.<br />

This evolution was forced upon us by unexpected supplemental<br />

discoveries that have shown that the existence of universal<br />

constants, such as the velocity of light, limit our power<br />

to manipulate nature. (We shall discuss this unexpected situation<br />

in Chapter VII.) As a result, physicists had to introduce<br />

new mathematical tools that make the relation between perception<br />

and interpretation more complex. Whatever reality may<br />

mean, it always corresponds to an active intellectual construction.<br />

The descriptions presented by science can no longer be<br />

disentangled from our questioning activity and therefore can<br />

no longer be attributed to some omniscient being.<br />

On the eve of the Newtonian synthesis, John Donne lamented<br />

the passing of the Aristotelian cosmos destroyed by<br />

Copernicus:<br />

And new Philosophy calls all in doubt,<br />

The Element of fire is quite put out,<br />

The Sun is lost, and th'earth, and no man's wit<br />

Can well direct him where to look for it.<br />

And freely men confess that this world's spent,<br />

When in the Planets and the Firmament,<br />

They seek so many new, then they see that this<br />

Is crumbled out again to his Atomies<br />

'Tis all in Pieces, all coherence gone.JO<br />

The scattered bricks and stones of our present culture seem,<br />

as in Donne's time, capable of being rebuilt into a new "coherence."<br />

Classical science, the mythical science of a simple,<br />

passive world, belongs to the past, killed not by philosophical<br />

criticism or empiricist resignation but by the internal development<br />

of science itself.


I<br />

I


CHAPTER II<br />

THE IDENTIFICATION<br />

OFTHEREAL<br />

Newtons Laws<br />

We shall now take a closer look at the mechanistic world view<br />

as it emerged from the work of Galileo, Newton, and their<br />

successors. We wish to describe its strong points, the aspects<br />

of nature it has succeeded in clarifying, but we also want to<br />

expose its limitations.<br />

·Ever since Galileo, one of the central problems of physics<br />

has been the description of acceleration. The surprising feature<br />

was that the change undergone by the state of motion of a<br />

body could be formulated in simple mathematical terms. This<br />

seems almost trivial to us today. Still, we should remember<br />

that Chinese science, so successful in many areas, did not produce<br />

a quantitative formulation of the laws of motion. Galileo<br />

discovered that we do not need to ask for the cause of a state<br />

of motion if the motion is uniform, any more than it is necessary<br />

to ask the reason for a state of rest. Both motion and rest<br />

remain indefinitely stable unless something happens to upset<br />

them. The central problem is the change from rest to motion,<br />

and from motion to rest, as well as, more generally, all changes<br />

of velocity. How do these changes occur? The formulation of<br />

the Newtonian laws of motion made use of two converging<br />

developments: one in physics, Kepler's laws for planetary motion<br />

and Galileo's laws for falling bodies, and the other in<br />

mathematics, the formulation of differential or "infinitesimal"<br />

calculus.<br />

How can a continuously varying speed be defined? How can<br />

we describe the instantaneous changes in the various quantities,<br />

such as position, velocity, and acceleration? How can<br />

we describe the state of a body at any given instant? To answer<br />

57


ORDER OUT OF CHAOS 58<br />

these questions, mathematicians have introduced the concept<br />

of infinitesimal quantities. An infinitesimal quantity is the result<br />

of a limiting process; it is typically the variation in a quantity<br />

occurring between two successive instants when the time<br />

elapsing between these instants tends toward zero. In this way<br />

the change is broken up into an infinite series of infinitely<br />

small changes.<br />

At each instant the state of a moving body can be defined by<br />

its position r, by its velocity v, which expresses its "instantaneous<br />

tendency" to modify this position, and by its acceleration<br />

a, again its "instantaneous tendency," but now to modify<br />

its velocity. Instantaneous velocities and accelerations are limiting<br />

quantities that measure the ratio between two infinitesimal<br />

quantities: the variation of r (or v) during a temporal<br />

interval 6.t, and this interval 6.t when 6.t tends to zero. Such<br />

quantities are "derivatives with respect to time," and since<br />

Leibniz they have been written as v=drldt and a=dv/dt.<br />

Therefore, acceleration, the derivative of a derivative, a= d2r!<br />

dt2, becomes a "second derivative." The problem on which<br />

Newtonian physics concentrates is the calculation of this second<br />

derivative, that is, of the acceleration undergone at each<br />

instant by the points that form a system. The motion of each of<br />

these points over a finite interval of time can then be calculated<br />

by integration, by adding up the infinitesimal velocity<br />

changes occurring during this interval. The simplest case is<br />

when a is constant (for example, for a freely falling body a is<br />

the gravitational constant g). Generally speaking, acceleration<br />

itself varies in time, and the physicist's task is to determine<br />

precisely the nature of this variation.<br />

In Newtonian language, to study acceleration means to determine<br />

the various "forces" acting on the points in the system<br />

under examination. Newton's second law, F= ma, states that<br />

the force applied at any point is proportional to the acceleration<br />

it produces. In the case of a system of material points, the<br />

problem is more complicated, since the forces acting on a<br />

given body are determined at each instant by the relative distances<br />

between the bodies of the system, and thus vary at each<br />

instant as a result of the motion they themselves produce.<br />

A problem in dynamics is expressed in the form of a set of<br />

"differential .. equations. The instantaneous state of each of<br />

the bodies in a system is described as a point and defined by


59 THE IDENTIFICATION OF THE REAL<br />

means of its position as well as by its velocity and acceleration,<br />

that is, by the first and second derivatives of the position.<br />

At each instant, a set of forces, which is a function of the<br />

distance between the points in the system (a function of r),<br />

gives a precise acceleration to each point; the accelerations<br />

then bring about changes in the distances separating these<br />

points and therefore in the set of forces acting at the following<br />

instant.<br />

While the differential equations set up the dynamics problem,<br />

their "integration" represents the solution of this problem.<br />

It leads to the calculation of the trajectories, r(t). These<br />

trajectories contain all the information acknowledged as relevant<br />

by dynamics; it provides a complete description of the<br />

dynamic system.<br />

The description therefore implies two elements: the positions<br />

and velocities of each of the points at one instant, often<br />

called the "initial instant," and the equations of motion that<br />

relate the dynamic forces to the accelerations. The integration<br />

of the dynamic equations starting from the "initial state" unfold<br />

the succession of states, that is, the set of trajectories of<br />

its constitutive bodies.<br />

The triumph of Newtonian science is the discovery that a<br />

single force, gravity, determines both the motion of planets<br />

and comets in the sky and the motions of bodies falling toward<br />

the earth. Whatever pair of material bodies is considered, the<br />

Newtonian system implies that they are linked by the same<br />

force of attraction. Newtonian dynamics thus appears to be<br />

doubly universal. The definition of the law of gravity that describes<br />

how masses tend to approach one another contains no<br />

reference to any scale of phenomena. It can be applied equally<br />

well to the motion of atoms, of planets, or of the stars in a<br />

galaxy. Every body, whatever its size, has a mass and acts as a<br />

source of the Newtonian forces of interaction.<br />

Since gravitational forces connect any two bodies (for two<br />

bodies of mass m and m ' and separated by a distance r, the<br />

gravitational force is kmm'fr2, where k is the Newtonian force<br />

of attraction equal to 6. 67cm3g-1sec-2), the only true dynamic<br />

system is the universe as a whole. Any local dynamic<br />

system, such as our planetary system, can only be defined<br />

approximately, by neglecting forces that are small in comparison<br />

to those whose effect is being considered.


ORDER OUT OF CHAOS<br />

60<br />

It must be emphasized that whatever the dynamic system<br />

chosen, the laws of motion can always be expressed in the<br />

form F= ma. Other types of forces apart from those due to<br />

gravity may be discovered (and actually have been discovered-for<br />

instance, electric forces of attraction and repulsion)<br />

and would thereby modify the empirical content of the laws of<br />

motion. They would not, however, modify the form of those<br />

laws. In the world of dynamics, change is identified with acceleration<br />

or deceleration. The integration of the laws of motion<br />

leads to the trajectories that the particles follow. Therefore the<br />

laws of change, of time's impact on nature, are expressed in<br />

terms of the characteristics of trajectories.<br />

The basic characteristics of trajectories are lawfulness, determinism,<br />

and reversibility. We have seen that in order to calculate<br />

a trajectory we need, in addition to our knowledge of<br />

the laws of motion, an empirical definition of a single instantaneous<br />

state of the system. The general law then deduces<br />

from this "initial state" the series of states the system passes<br />

through as time progresses, just as logic deduces a conclusion<br />

from basic premises. The remarkable feature is that once the<br />

forces are known, any single state is sufficient to define the<br />

system completely, not only its future but also its past. At<br />

each instant, therefore, everything is given. Dynamics defines<br />

all states as equivalent: each of them allows all the others to be<br />

calculated along with the trajectory which connects all states,<br />

be they in the past or the future.<br />

"Everything is given." This conclusion of classical dynamics,<br />

which Bergson repeatedly emphasized, characterizes the<br />

reality that dynamics describes. Everything is given, but everything<br />

is also possible. A being who has the power to control<br />

a dynamic system may calculate the right initial state in such a<br />

way that the system "spontaneously" reaches any chosen<br />

state at some chosen time. The generality of dynamic laws is .<br />

matched by the arbitrariness of the initial conditions.<br />

The reversibility of a dynamic trajectory was explicitly<br />

stated by all the founders of dynamics. For instance, when<br />

Galilee or Huyghens described the implications of the equivalence<br />

between cause and effect postulated as the basis of<br />

their mathematization of motion, they staged thought experiments<br />

such as an elastic ball bouncing on the ground. As the


61 THE IDENTIFICATION OF THE REAL<br />

result of its instantaneous velocity inversion, such a body<br />

would return to its initial position. Dynamics assigns this<br />

property of reversibility to all dynamic changes. This early<br />

"thought experiment" illustrates a general mathematical property<br />

of dynamic equations. The structure of these equations<br />

implies that if the velocities of all the points of a system are<br />

reversed, the system will go "backward in time. " The system<br />

would retrace all the states it went through during the previous<br />

change. Dynamics defines as mathematically equivalent<br />

changes such as t-+- t, time inversion, and v-+- v, velocity<br />

reversal. What one dynamic change has achieved, another<br />

change, defined by velocity inversion, can undo, and in this<br />

way exactly restore the original conditions.<br />

This property of reversibility in dynamics leads, however, to<br />

a difficulty whose full significance was realized only with the<br />

introduction of quantum mechanics. Manipulation and measurement<br />

are essentially irreversible. Active science is thus,<br />

by definition, extraneous to the idealized, reversible world it is<br />

describing. From a more general point of view, reversibility<br />

may be taken as the very symbol of the "strangeness" of the<br />

world described by dynamics. Everyone is familiar with the<br />

absurd effects produced by projecting a film backward-the<br />

sight of a match being regenerated by its flame, broken ink<br />

pots that reassemble and return to a tabletop after the ink has<br />

poured back into them, branches that grow young again and<br />

turn into fresh shoots. In the world of classical dynamics, such<br />

events are considered to be just as likely as the normal ones.<br />

We are so accustomed to the laws of classical dynamics that<br />

are taught to us early in school that we often fail to sense the<br />

boldness of the assumptions on which they are based. A world<br />

in which all trajectories are reversible is a strange world indeed.<br />

Another astonishing assumption is that of the complete<br />

independence of initial conditions from the laws of motion. It<br />

is possible to take a stone and throw it with some initial velocity<br />

limited only by one's physical strength, but what about<br />

a complex system SlJCh as a gas formed by many particles? It<br />

is obvious that we can no longer impose arbitrary initial conditions.<br />

The initial conditions must be the outcome of the dynamic<br />

evolution itself. This is an important point to which we<br />

shall come back in the third part of this book. But whatever its


ORDER OUT OF CHAOS 62<br />

limitations, today, three centuries later, we can only admire<br />

the logical coherence and the power of the methods discovered<br />

by the founders of classical dynamics.<br />

Motion and Change<br />

Aristotle made time the measure of change. But he was fully<br />

aware of the qualitative multiplicity of change in nature. Still<br />

there is only one type of change surviving in dynamics, one<br />

"process," and that is motion. The qualitative diversity of<br />

changes in nature is reduced to the study of the relative displacement<br />

of material bodies. Time is a parameter in terms of<br />

which these displacements may be described. In this way<br />

space and time are inextricably tied together in the world of<br />

classical dynamics. (Also see Chapter IX.)<br />

It is interesting to compare dynamic change with the atomists'<br />

conception of change, which enjoyed considerable favor<br />

at the time Newton formulated his laws. Actually, it seems that<br />

not only Descartes, Gassendi, and d'Alembert, but even Newton<br />

himself believed that collisions between hard atoms were<br />

the ultimate, and perhaps the only, sources of changes of motion.t<br />

Nevertheless, the dynamic and the atomic descriptions<br />

differ radically. Indeed, the continuous nature of the acceleration<br />

described by the dynamic equations is in sharp contrast<br />

with the discontinuous, instantaneous collisions between hard<br />

particles. Newton had already noticed that, in contradiction to<br />

dynamics, an irreversible loss of motion is involved in each<br />

hard collision. The only reversible collision-that is, the only<br />

one in agreement with the laws of dynamics-is the "elastic,"<br />

momentum-conserving collision. But how can the complex<br />

property of "elasticity" be applied to atoms that are supposed<br />

to be the fundamental elements of nature?<br />

On the other hand, at a less technical level, the laws of dynamic<br />

motion seem to contradict the randomness generally<br />

attributed to collisions between atoms. The ancient philosophers<br />

had already pointed out that any natural process can be<br />

interpreted in many different ways in terms of the motion of<br />

and collisions between atoms. This was not a problem for the<br />

atomists, since their main aim was to describe a godless, law-


63<br />

THE IDENTIFICATION OF THE REAL<br />

less world in which man is free and can expect to receive neither<br />

punishment nor reward from any divine or natural order.<br />

But classical science was a science of engineers and astronomers,<br />

a science of action and prediction. Speculations based<br />

on hypothetical atoms could not satisfy its needs. In contrast,<br />

Newton's law provided a means of predicting and manipulating.<br />

Nature thus becomes law-abiding, docile, and predictable,<br />

instead of being chaotic, unruly, and stochastic. But<br />

what is the connection between the mortal, unstable world in<br />

which atoms unceasingly combine and separate, and the immutable<br />

world of dynamics governed by Newton's law, a siagle<br />

mathematical formula corresponding to an eternal truth unfolding<br />

toward a tautological future? In the twentieth century<br />

we are again witnessing the clash between lawfulness and random<br />

events, which, as Koyre has shown, had already tormented<br />

Descartes. 2 Ever since the end of the nineteenth<br />

century, with the kinetic theory of gases, the atomic chaos has<br />

reintegrated physics, and the problem of the relationship between<br />

dynamic law and statistical description has penetrated<br />

to the very core of physics. It is one of the key elements in the<br />

present renewal of dynamics (see Book III).<br />

In the eighteenth century, however, this contradiction<br />

seemed to produce a deadlock. This may partly explain the<br />

skepticism of some eighteenth-century physicists regarding<br />

the significance of Newton's dynamic description. We have already<br />

noted that collisions may lead to a loss of motion. They<br />

thereby concluded that in such nonideal cases, "energy" is<br />

not conserved but is irreversibly dissipated (see Chapter IV,<br />

section 3). Therefore, the atomists could not help considering<br />

dynamics as an idealization of limited value. Continental<br />

physicists and mathematicians such as dl\lembert, Clairaut,<br />

and Lagrange resisted the seductive charms of Newtonianism<br />

for a long time.<br />

Where do the roots of the Newtonian concept of change lie?<br />

It appears to be a synthesis3 of the science of ideal machines,<br />

where motion is transmitted without collision or friction between<br />

parts already in contact, and the science of celestial<br />

bodies interacting at a distance. We have seen that it appears<br />

as the very antithesis of atomism, which is based on the concept<br />

of random collisions. Does this, however, vindicate the<br />

view of those who believe that Newtonian dynamics repre-


ORDER OUT OF CHAOS<br />

64<br />

sents a rupture in the history of thinking, a revolutionary novelty?<br />

This is what positivist historians have claimed when they<br />

described how Newton escaped the spell of preconceived notions<br />

and had the courage to infer from the mathematical study of<br />

planetary motions and the laws of falling bodies the action of a<br />

"universal" force. We know that on the contrary the eighteenthcentury<br />

rationalists emphasized the apparent similarity between<br />

his "mathematical" forces and traditional occult<br />

qualities. Fortunately, these critics did not know the strange<br />

story behind the Newtonian forces! For behind Newton's cautious<br />

declaration-"! frame no hypotheses"-concerning the<br />

nature of the forces lurked the passion of an alchemist.4 We<br />

now know that, side by side with his mathematical studies,<br />

Newton had studied the ancient alchemists for thirty years<br />

and had carried out painstaking laboratory experiments on<br />

ways of achieving the master work, the synthesis of gold.<br />

Recently some historians have gone so far as to propose that<br />

the Newtonian synthesis of heaven and earth was the achievement<br />

of a chemist, not an astronomer. The Newtonian force<br />

"animating" matter and, in the stronger sense of the term,<br />

making up the very activity of nature would then be the inheritor<br />

of the forces Newton the chemist observed and manipulated,<br />

the chemical "affinities" forming and disrupting ever<br />

new combinations of matter.s The decisive role played by celestial<br />

orbits of course remains. Still, at the start of his intense<br />

astronomical studies-about 1679-Newton apparently expected<br />

to find new forces of attraction only in the heavens,<br />

forces similar to chemical forces and perhaps easier to study<br />

mathematically. Six years later this mathematical study produced<br />

an unexpected conclusion: the forces between the planets<br />

and those accelerating freely falling bodies are not merely<br />

similar but are the same. Attraction is not specific to each<br />

planet; it is the same-for the moon circling the earth, for the<br />

planets, and even for comets passing through the solar system.<br />

Newton set out to discover in the sky forces similar to the<br />

chemical forces: the specific affinities, different for each<br />

chemical compound and giving each compound qualitatively<br />

differentiated activities. What he actually found was a universal<br />

law, which, as he emphasized, could be applied to all phenomena-whether<br />

chemical, mechanical, or celestial in<br />

nature.


65 THE IDENTIFICATION OF THE REAL<br />

The Newtonian synthesis is thus a surprise. It is an unexpected,<br />

staggering discovery that the scientific world has commemorated<br />

by making Newton the symbol of modern science.<br />

What is particularly astonishing is that the basic code of nature<br />

appeared to have been cracked in a single creative act.<br />

For a long time this sudden loquaciousness of nature, this<br />

triumph of the English Moses, was a source of intellectual<br />

scandal for continental rationalists. Newton's work was<br />

viewed as a purely empirical discovery that could thus equally<br />

well be empirically disproved. In 1747 Euler, Clairaut, and<br />

dlembert, without doubt some of the greatest scientists of<br />

the time, came to the same conclusion: Newton was wrong. In<br />

order to describe the moon's motion, a more complex mathematical<br />

form must be given to the force of attraction, making it<br />

the sum of two terms. For the following two years, each of<br />

them believed that nature had proved Newton wrong, and this<br />

belief was a source of excitement, not of dismay. Far from considering<br />

Newton's discovery synonymous with physical science<br />

itself, physicists were blithely contemplating dropping it<br />

altogether. Dlembert went so far as to express scruples<br />

about seeking fresh evidence against Newton and giving him<br />

"le coup de pied de l'iine."6<br />

Only one courageous voice against this verdict was raised in<br />

France. In 1748, Buffon wrote:<br />

A physical law is a law only by virtue of the fact that it is<br />

easy to measure, and that the sale it represents is not<br />

only always the same, but is actually unique . ... M.<br />

Clairaut has raised an objection against Newton's system,<br />

but it is at best an objection and must not and cannot<br />

become a principle; an attempt should be ,made to overcome<br />

it and not to turn it into a theory the entire consequences<br />

of which merely rest on a calculation; for, as I<br />

have said, one may represent anything by means of calculation<br />

and achieve nothing; and if it is allowed to add<br />

one or more terms to a physical law such as that of attraction,<br />

we are only adding to arbitrariness instead of representing<br />

reality. 7<br />

Later Buffon was to announce what was to become, although<br />

for only a short time, the research program for chemistry:


ORDER OUT OF CHAOS<br />

66<br />

The laws of affinity by means of which the constituent<br />

parts of different substances separate from others to<br />

combine together to form homogeneous substances are<br />

the same as the general law governing the reciprocal action<br />

of all the celestial bodies on one another: they act in<br />

(he same way and with the same ratios of mass and distance;<br />

a globule of water, of sand or metal acts upon another<br />

globule just as the terrestrial globe acts on the<br />

moon, and if the laws of affinity have hitherto been regarded<br />

as different from those of gravity, it is because<br />

they have not been fully understood, not grasped completely;<br />

it is because the whole extent of the problem has<br />

not been taken in. The figure which, in the case of celestial<br />

bodies has little or no effect upon the law of interaction<br />

between bodies because of the great distance<br />

involved, is, on the contrary, all important when the distance<br />

is very small or zero . ... Our nephews will be<br />

able, by calculation, to gain access to this new field of<br />

knowledge [that is, to deduce the law of interaction between<br />

elementary bodies from their figures].&<br />

History was to vindicate the naturalist, for whom force was<br />

not mer;ely a mathematical artifice but the very essence of the<br />

new science of nature. The physicists were later compelled to<br />

admit their mistake. Fifty years afterward, Laplace could<br />

write his Systeme du Monde. The law of universal gravity had<br />

stood all tests successfully: the numerous cases apparently<br />

disproving it had been transformed into a brilliant demonstration<br />

of its validity. At the same time, under Buffon's influence,<br />

the French chemists rediscovered the odd analogy between<br />

physical attraction and chemical affinity.9 Despite the sarcasms<br />

of d/\lembert, Condillac, and Condorcet, whose unbending<br />

rationalism was quite incompatible with these obscure and<br />

barren "analogies," they trod Newton's path in the opposite<br />

direction-from the stars to matter.<br />

By the early nineteenth century, the Newtonian programthe<br />

reduction of all physicochemical phenomena to the action<br />

of forces (in addition to gravitational attraction, this included<br />

the repelling force of heat, which makes bodies expand and<br />

favors dissolution, as well as electric and magnetic forces)­<br />

had become the official program of Laplace's school, which


67 THE IDENTIFICATION OF THE REAL<br />

dominated the scientific world at the time when Napoleon<br />

dominated Europe. JO<br />

The early nineteenth century saw the rise of the great<br />

French ecoles and the re<strong>org</strong>anization of the universities. This<br />

is the time when scientists became teachers and professional<br />

researchers and took up the tak of training their successors.••<br />

It is also the time of the first attempts to present a synthesis of<br />

knowledge, to gather it together in textbooks and works of<br />

popularization. Science was no longer discussed in the salons;<br />

it was taught or popularized.1 2 It had become a matter of professional<br />

consensus and magistral authority. The first consensus<br />

centered around the Newtonian system: in France<br />

Buffon's confidence finally triumphed over the rational skepticism<br />

of the Enlightenment.<br />

One century after Newton's apotheosis in England, the<br />

grandiloquence of these lines written by Ampere's son echoes<br />

that of Pope's epitaph: 13<br />

Announcing the coming of science's Messiah<br />

Kepler had dispelled the clouds around the Arch.<br />

Then the Word was made man, the Word of the seeing<br />

God<br />

Whom Plato revered, and He was called Newton.<br />

He came, he revealed the principle supreme,<br />

Eternal, universal, One and unique as God Himself.<br />

The worlds were hushed, he spoke: ATTRACTION.<br />

This word was the very word of creation.*<br />

For a short time, which nevertheless left an indelible mark,<br />

science was triumphant, acknowledged and honored by<br />

powerful states and acclaimed as the possessor of a consistent<br />

conception of the world. Worshiped by Laplace, Newton became<br />

the universal symbol of this golden age. It was a happy<br />

moment, indeed, a moment in which scientists were regarded<br />

both by themselves and others as the pioneers of progress,<br />

achieving an enterprise sustained and fostered by society as a<br />

whole.<br />

What is the significance of the Newtonian synthesis today,<br />

after the advent of field theory, relativity, and quantum me-<br />

*Our translation-authors.


ORDER OUT OF CHAOS<br />

68<br />

chanics? This is a complex problem, to which we shall return.<br />

We now know that nature is not always "comfortable and consonant<br />

with herself. " At the microscopic level, the laws of<br />

classical mechanics have been replaced by those of quantum<br />

mechanics. Likewise, at the level of the universe, relativistic<br />

physics has displaced Newtonian physics. Classical physics<br />

nevertheless remains the natural reference point. Moreover, in<br />

the sense that we have defined it-that is, as the description of<br />

deterministic, reversible, static trajectories-Newtonian dynamics<br />

still may be said to form the core of physics.<br />

Of course, since Newton the formulation of classical dynamics<br />

has undergone great changes. This was a result of the<br />

work of some of the greatest mathematicians and physicists,<br />

such as Hamilton and Poincare. In brief, we may distinguish<br />

two periods. First there was a period of clarification and of<br />

generalization. During the second period, the very concepts<br />

upon which classical dynamics rests, such as initial conditions<br />

and the meaning of trajectories, have undergone a critical revision<br />

even in the fields in which (in contrast to quantum mechanics<br />

and relativity) classical dynamics remains valid. At<br />

the moment this book is being written, at the end of the twentieth<br />

century, we are still in this second period. Let us turn<br />

now to the general language of dynamics that was discovered<br />

by nineteenth-century scientists. (In Chapter IX we shall describe<br />

briefly the revival of classical dynamics in our time.)<br />

The Language of Dynarnics<br />

Today classical dynamics can be formulated in a compact and<br />

elegant way. As we shall see, all the properties of a dynamic<br />

system can be summarized in terms of a single function, the<br />

Hamiltonian. The language of dynamics presents a remarkable<br />

consistency and completeness. An unambiguous formulation<br />

can be given to each "legitimate" problem. No wonder the<br />

structure of dynamics has both fascinated and terrified the<br />

imagination since the eighteenth century.<br />

In dynamics,the same system can be studied from different<br />

points of view. In classical dynamics all these points of view<br />

are equivalent in the sense that we can go from one to another<br />

by a transformation, a change of variables. We may speak of


69 THE IDENTIFICATION OF THE REAL<br />

various equivalent representations in which the laws of dynamics<br />

are valid. These various equivalent representations<br />

form the general language of dynamics. This language can be<br />

used to make explicit the static character classical dynamics<br />

attributes to the systems it describes: for many classes of dynamic<br />

systems, time appears merely as an accident, since<br />

their description can be reduced to that of noninteracting mechanical<br />

systems. To introduce these concepts in a simple way,<br />

let us start with the principle of conservation of energy.<br />

In the ideal world of dynamics, devoid of frictions and collisions,<br />

machines have an efficiency of one-the dynamic system<br />

comprising the machine merely transmits the whole of the<br />

motion it receives. A machine receiving a certain quantity of<br />

potential energy (for example, a compressed spring, a raised<br />

weight, compressed air) can produce a motion corresponding<br />

to an "equal" quantity of kinetic energy, exactly the quantity<br />

that would be needed to restore the potential energy the machine<br />

has used in producing the motion. The simplest case is<br />

that in which the only force considered is gravity (which applies<br />

to simple machines, pulleys, levers, capstans, etc.). In<br />

this case it is easy to establish an overall relationship of equivalence<br />

between cause and effect. The height (h) through which<br />

a body falls entirely determines the velocity acquired during<br />

its fall. Whether a body of mass m falls vertically, runs down<br />

an inclined plane, or follows a roller-coaster path, the acquired<br />

velocity (v) and the kinetic energy (mv2/2) depend only on the<br />

drop in level h (v = Vfiii) and enable the body to return to its<br />

original height. The work done against the force of gravity implied<br />

in this upward motion restores the potential energy, mgh,<br />

that the system lost during the fall. Another example is the<br />

pendulum, in which kinetic energy and potential energy are<br />

continuously transformed into one another.<br />

Of course, if instead of a body falling toward the earth, we<br />

are dealing with a system of interacting bodies, the situation is<br />

less easily visualized. Still, at each instant the global variation<br />

in kinetic energy compensates for the variation in potential<br />

energy (bound to the variation in the distances between the<br />

points in the system). Here also energy is conserved in an isolated<br />

system.<br />

Potential energy (or ''potential," conventionally denoted as<br />

V), which depends on the relative positions of the particles, is


ORDER OUT OF CHAOS 70<br />

thus a generalization of the quantity that enabled builders of<br />

machines to measure the motion a machine could produce as<br />

the result of a change in its spatial configuration (for example,<br />

the change in the height of a mass m, which is part of the<br />

machine, gives it a potential energy mgh). Moreover, potential<br />

energy allows us to calculate the set of forces applied at each<br />

instant to the different points of the system to be described. At<br />

each point the derivative of the potential with respect to the<br />

space coordinate q measures the force applied at this point in<br />

the direction of that coordinate. Newton's laws of motion thus<br />

can be formulated using the potential function instead of force<br />

as the main quantity: the variation in the velocity of a point<br />

mass at each instant (or the momentum p, the product of the<br />

mass and the velocity) is measured by the derivative of the<br />

potential with respect to the coordinate q of the mass.<br />

In the nineteenth century this formulation was generalized<br />

through the introduction of a new function, the Hamiltonian<br />

(H). This function is simply the total energy, the sum of the<br />

system's potential and kinetic energy. However, this energy is<br />

no longer expressed in terms of positions and velocities, conventionally<br />

denoted by q and dq/dt, but in terms of so-called<br />

canonical variables-coordinates and momenta-for which<br />

the standard notation is q and p. In simple cases, such as with<br />

a free particle, there is a straightforward relation between velocity<br />

and momentum (p = m dqldt), but in general the relation<br />

is more complicated.<br />

A single function, the Hamiltonian, H(p, q), describes the<br />

dynamics of a system completely. All our empirical knowledge<br />

is put into the form of H. Once this function is known, we may<br />

solve, at least in principle, all possible problems. For example,<br />

the time variation of the coordinate and of the momenta is<br />

simply given by the derivatives of H in respect to p or q. This<br />

Hamiltonian formulation of dynamics is one of the greatest<br />

achievements in the history of science. It has been progressively<br />

extended to cover the theory of electricity and magnetism.<br />

It has also been used in quantum mechanics. It is true<br />

that in quantum mechanics, as we shall see later, the meaning<br />

of the Hamiltonian H had to be generalized: here it is no<br />

longer a simple function of the coordinates and momenta, but<br />

it becomes a new kind of entity, an operator. (We shall return<br />

to this question in Chapter VII.) In any case, the Hamiltonian


71 THE IDENTIFICATION OF THE REAL<br />

description is still of the greatest importance today. The equations<br />

which, through the derivatives of the Hamiltonian, give<br />

the time variation of the coordinates and momenta are the socalled<br />

canonical equations. They contain the general properties<br />

of all dynamic changes. Here we have the triumph of the<br />

mathematization of nature. All dynamic change to which classical<br />

dynamics applies can be reduced to these simple mathematical<br />

equations.<br />

Using these equations, we can verify the above-mentioned<br />

general properties implied by classical dynamics. The canonical<br />

equations are reversible: time inversion is mathematically<br />

the equivalent of velocity inversion. They are also conservative:<br />

the Hamiltonian, which expressed the system's energy in<br />

the canonical variables-coordinates and momenta-is itself<br />

conserved by the changes it brings about in the course of time.<br />

We have already noticed that there exist many points of view<br />

or "representations" in which the Hamiltonian form of the<br />

equations of motion is maintained. They correspond to various<br />

choices of coordinates and momenta. One of the basic problems<br />

of dynamics is to examine precisely how we can select<br />

the pair of canonical variables q and p to obtain as simple a<br />

description of dynamics as possible. For example, we could<br />

look for canonical variables by which the Hamiltonian is reduced<br />

to kinetic energy and depends only on the momenta<br />

(and not on the coordinates). What is remarkable is that in this<br />

case momenta become constants of motion. Indeed, as we<br />

have seen, the time variation of the momenta depends, according<br />

to the canonical equation, on the derivative of the Hamiltonian<br />

in respect to the coordinates. When this derivative<br />

vanishes, the momenta indeed become constants of motion.<br />

This is similar to what happens in a "free particle" system.<br />

What we have done when we go to a free particle system is<br />

"eliminate" the interaction through a change of representation.<br />

We will define systems for which this is possible as "integrable<br />

systems." Any integrable system may thus be<br />

represented as a set of units, each changing in isolation, quite<br />

independently of all the others, in that eternal and immutable<br />

motion Aristotle attributed to the heavenly bodies (Figure 1).<br />

We have already noted that in dynamics "everything is<br />

given." Here this means that, from the very first instant, the<br />

value of the various invariants of the motion is fixed; nothing


ORDER OUT OF CHAOS 72<br />

•<br />

><br />

•<br />

•<br />

•<br />

•<br />

•<br />

•<br />

(a )<br />

(b)<br />

Figure 1. Two representations of the same dynamic system: (a) as a set of<br />

interacting points; the interaction between the points is represented by wavy<br />

lines; (b) as a set where each point behaves independently from the others.<br />

The potential energy being eliminated, their respective · motions are not explicitly<br />

dependent on their relative positions.<br />

may "happen" or "take place." Here we reach one of those<br />

dramatic moments in the history of science when the description<br />

of nature was nearly reduced to a static picture. Indeed,<br />

through a clever change of variables, all interaction could be<br />

made to disappear. It was believed that integrable systems,<br />

reducible to free particles, were the prototype of dynamic systems.<br />

Generations of physicists and mathematicians tried hard<br />

to find for each kind of systems the "right" variables that<br />

would eliminate the interactions. One widely studied example<br />

was the three-body problem, perhaps the most important<br />

problem in the history of dynamics. The moon's motion, influenced<br />

by both the earth and the sun, is one instance of this<br />

problem. Countless attempts were made to express it in the<br />

form of an integrable system until, at the end of the nineteenth<br />

century, Bruns and Poincare showed that this was impossible.<br />

This came as a surprise and, in fact, announced the end of all<br />

simple extrapolations of dynamics based on integrable systems.<br />

The discovery of Bruns and Poincare shows that dynamic<br />

systems are not isomorphic. Simple, integrable systems<br />

can indeed be reduced to noninteracting units, but in general,<br />

interactions cannot be eliminated. Although this discovery<br />

was not clearly understood at the time, it implied the demise of<br />

the conviction that the dynamic world is homogeneous, reducible<br />

to the concept of integrable systems. Nature as an


73 THE IDENTIFICATION OF THE REAL<br />

evolving, interactive multiplicity thus resisted its reduction to<br />

a timeless and universal scheme.<br />

There were other indications pointing in the same direction.<br />

We have mentioned that trajectories correspond to deterministic<br />

laws; once an initial state is given, the dynamic laws of motion<br />

permit the calculation of trajectories at each point in the<br />

future or the past. However, a trajectory may become intrinsically<br />

indeterminate at certain singular points. For instance, a<br />

rigid pendulum may display two qualitatively different types of<br />

behavior-it may either oscillate or swing around its points of<br />

suspension. If the initial push is just enough to bring it into a<br />

vertical position with zero velocity, the direction in which it<br />

will fall, and therefore the nature of its motion, are indeterminate.<br />

An infinitesimal perturbation would be enough to set it<br />

rotating or oscillating. (This problem of the "instability" of<br />

motion will be discussed fully in Chapter IX.)<br />

It is significant that Maxwell had already stressed the importance<br />

of these singular points. After describing the explosion<br />

of gun cotton, he goes on to say:<br />

In all such cases there is one common circumstancethe<br />

system has a quantity of potential energy, which is<br />

capable of being transformed into motion, but which cannot<br />

begin to be so transformed till the system has reached<br />

a certain configuration, to attain which requires an expenditure<br />

of work, which in certain cases may be infinitesimally<br />

small, and in general bears no definite proportion to the<br />

energy developed in consequence thereof. For example,<br />

the rock loosed by frost and balanced on a singular point<br />

of the mountain-side, the little spark which kindles the<br />

great forest, the little word which sets the world a fighting,<br />

the little scruple which prevents a man from doing his<br />

will, the little spore which blights all the potatoes, the<br />

little gemmule which makes us philosophers or idiots.<br />

Every existence above a certain rank has its singular<br />

points: the higher the rank, the more of them. At these<br />

points, influences whose physical magnitude is too small<br />

to be taken account of by a finite being, may produce<br />

results of the greatest importance. All great results produced<br />

by human endeavour depend on taking advantage<br />

of these singular states when they occur. I4


ORDER OUT OF CHAOS 74<br />

This conception received no further elaboration owing to the<br />

absence of suitable mathematical techniques for identifying<br />

systems containing such singular points and the absence of the<br />

chemical and biological knowledge that today affords, as we<br />

shall see later, a deeper insight into the truly essential role<br />

played by such singular points.<br />

Be that as it may, from the time of Leibniz' monads (see the<br />

conclusion to section 4) down to the present day (for example,<br />

the stationary states of the electrons in the Bohr model-see<br />

Chapter VII), integrable systems have been the model par excellence<br />

of dynamic systems, and physicists have attempted to<br />

extend the properties of what is actually a very special class of<br />

Hamiltonian equations to cover all natural processes. This is<br />

quite understandable. The class of integrable systems is the<br />

only one that, until recently, had been thoroughly explored.<br />

Moreover, there is the fascination always associated with a<br />

closed system capable of posing all problems, provided it does<br />

not define them as meaningless. Dynamics is such a language;<br />

being complete, it is by definition coextensive with the world<br />

it is describing. It assumes that all problems, whether simple<br />

or complex, resemble one another since it can always pose<br />

them in the same general form. Thus the temptation to conclude<br />

that all problems resemble one another from the point of<br />

view of their solutions as well, and that nothing new can appear<br />

as a result of the greater or lesser complexity of the integration<br />

procedure. It is this intrinsic homogeneity that we now<br />

know to be false. Moreover, the mechanical world view was<br />

acceptable as long as all observables referred in one way or<br />

another to motion. This is no longer the case. For example ,<br />

unstable particles have an energy that can be related to motion<br />

but that also has a lifetime that is a quite different type of observable,<br />

more closely related to irreversible processes, as we<br />

shall describe them in Chapters IV and V. The necessity of<br />

introducing new observables into the theoretical sciences was,<br />

and still is today, one of the driving forces that move us beyond<br />

the mechanical world view.


75 THE IDENTIFICATION OF THE REAL<br />

Laplaces Demon<br />

Extrapolations from the dynamic description discussed above<br />

have a symbol-the demon imagined by Laplace, capable at<br />

any given instant of observing the position and velocity of<br />

each mass that forms part of the universe and of inferring its<br />

evolution, both toward the past and toward the future. Of<br />

course, no one has ever dreamed that a physicist might one<br />

day benefit from the knowledge possessed by Laplace's demon.<br />

Laplace himself only used this fiction to demonstrate the<br />

extent of our ignorance and the need for a statistical description<br />

of certain processes. The problematics of Laplace's demon<br />

are not related to the question of whether a deterministic<br />

prediction of the course of events is actually possible, but<br />

whether it is possible in principle, de jure. This possibility<br />

seems to be implied in mechanistic description, with its<br />

characteristic duality based on dynamic law and initial conditions.<br />

Indeed, the fact that a dynamic system is governed by a<br />

deterministic law, even though in practice our ignorance of the<br />

initial state precludes any possibility of deterministic predictions,<br />

allows the "objective truth" of the system as it would be<br />

seen by Laplace's demon to be distinguished from empirical<br />

limitations due to our ignorance. In the context of classical<br />

dynamics, a deterministic description may be unattainable in<br />

practice; nevertheless, it stands as a limit that defines a series<br />

of increasingly accurate descriptions.<br />

It is precisely the consistency of this duality formed by dynamic<br />

law and initial conditions that is challenged in the revival<br />

of classical mechanics, which we will describe in Chapter<br />

IX. We shall see that the motion may become so complex, the<br />

trajectories so varied, that no observation, whatever its precision,<br />

can lead us to the determination of the exact initial conditions.<br />

But at that point the duality on which classical mechanics<br />

was constructed breaks down. We can predict only the average<br />

behavior of bundles of trajectories.<br />

Modern science was born out of the breakdown of the animistic<br />

alliance with nature. Man seemed to possess a place in<br />

the Aristotelian world as both a living and a knowing creature.


ORDER OUT OF CHAOS 76<br />

The world was made to his measure. The first experimental<br />

dialogue received part of its social and philosophic justification<br />

from another alliance, this time with the rational God of<br />

Christianity. To the extent to which dynamics has become and<br />

still is the model of science, certain implications of this historical<br />

situation have persisted to our day.<br />

Science is still the prophetic announcement of a description<br />

of the world seen from a divine or demonic point of view. It is<br />

the science of Newton, the new Moses to whom the truth of<br />

the world was unveiled; it is a revealed science that seems<br />

alien to any social and historical context identifying it as the<br />

result of the activity of human society. This type of inspired<br />

discourse is found throughout the history of physics. It has accompanied<br />

each conceptual innovation, each occasion at<br />

which physics seemed at the point of unification and the prudent<br />

mask of positivism was dropped. Each time physicists<br />

repeated what Ampere's son stated so explicitly: this worduniversal<br />

attraction, energy, field theory, or elementary particles-is<br />

the word of creation. Each time-in Laplace's time,<br />

at the end of the nineteenth century, or even today-physicists<br />

announced that physics was a closed book or about to become<br />

so. There was only one final stronghold where nature continued<br />

to resist, the fall of which would leave it defenseless,<br />

conquered, and subdued by our knowledge. They were thus<br />

unwittingly repeating the ritual of the ancient faith. They were<br />

announcing the coming of the new Moses, and with him a new<br />

Messianic period in science.<br />

Some might wish to disregard this prophetic claim, this<br />

somewhat naive enthusiasm, and it is certainly true that dialogue<br />

with nature has gone on all the same, together with a<br />

search for new theoretical languages, new questions, and new<br />

answers. But we do not accept a rigid separation between the<br />

scientist's "actual" work and the way he judges, interprets,<br />

and orientates this work. To accept it would be to reduce science<br />

to an ahistorical accumulation of results and to pay no<br />

attention to what scientists are looking for, the ideal knowledge<br />

they try to attain, the reasons why they occasionally<br />

quarrel or remain unable to communicate with each other. ts<br />

Once again, it was Einstein who formulated the enigma produced<br />

by the myth of modern science. He has stated that the<br />

miracle, the only truly astonishing feature, is that science ex-


77 THE IDENTIFICATION OF THE REAL<br />

ists at all, that we find a convergence between nature and the<br />

human mind. Similarly, when, at the end of the nineteenth<br />

century, du Bois Reymond made Laplace's demon the very<br />

incarnation of the logic of modern science, he added, "Ignoramus,<br />

ignorabimus": we shall always be totally ignorant of<br />

the relationship between the world of science and the mind<br />

which knows, perceives, and creates this science.I6<br />

Nature speaks with a thousand voices, and we have only begun<br />

to listen. Nevertheless, for nearly two centuries Laplace's<br />

demon has plagued our imagination, bringing a nightmare in<br />

which all things are insignificant. If it were really true that the<br />

world is such that a demon-a being that is, after all, like us,<br />

possessing the same science, but endowed with sharper<br />

senses and greater powers of calculation-could, starting from<br />

the observation of an instantaneous state, calculate its future<br />

and past, if nothing qualitatively differentiates the simple systems<br />

we can describe from the more complex ones for which a<br />

demon is needed, then the world is nothing but an immense<br />

tautology. This is the challenge of the science we have inherited<br />

from our predecessors, the spell we have to exorcise today.


I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I


CHAPTER Ill<br />

THE TWO CULTURES<br />

Diderot and the Discourse of the Living<br />

In his interesting book on the history of the idea of progress,<br />

Nisbet writes:<br />

No single idea has been more important than, perhaps as<br />

important as, the idea of progress in Western civilization<br />

for nearly three thousand years.•<br />

There has been no stronger support for the idea of progress<br />

than the accumulation of knowledge. The grandiose spectacle<br />

of this gradual increase of knowledge is indeed a magnificent<br />

example of a successful collective human endeavor.<br />

Let us recall the remarkable discoveries achieved at the end<br />

of the eighteenth century and the beginning of the nineteenth<br />

century: the theories of heat, electricity, magnetism, and optics.<br />

It is not surprising that the idea of scientific progress,<br />

already clearly formulated in the eighteenth century, dominated<br />

the nineteenth. Still, as we have pointed out, the position<br />

of science in We stern culture remained unstable. This<br />

lends a dramatic aspect to the history of ideas from the high<br />

point of the Enlightenment.<br />

We have already stated the alternative: to accept science<br />

with what appears to be its alienating conclusions or to turn to<br />

an antiscientific metaphysics. We have also emphasized the<br />

solitude felt by modern men, the loneliness described by Pascal,<br />

Kierkegaard, or Monod. We have mentioned the antiscientific<br />

implications of Heidegger's metaphysics. Now we<br />

wish to discuss more fully some aspects of the intellectual history<br />

of the West, from Diderot, Kant, and Hegel to Whitehead<br />

and Bergson; all of them attempted to analyze and limit the<br />

scope of modern science as well as to open new perspectives<br />

79


ORDER OUT OF CHAOS<br />

80<br />

seen as radically alien to that science. Today it is usually<br />

agreed that those attempts have for the most part failed. Few<br />

would accept, for example, Kant's division of the world into<br />

phenomenal and noumenal spheres, or Bergson's "intuition"<br />

as an alternative path to a knowledge whose significance<br />

would parallel that of science. Still these attempts are part of<br />

our heritage. The history of ideas cannot be understood without<br />

reference to them.<br />

We shall also briefly discuss scientific positivism, which is<br />

based on the separation of what is true from what is scientifically<br />

useful. At the outset this positivistic view may seem to op<br />

pose clearly the metaphysical views we have mentioned, views<br />

that I. Berlin described as the "Counter-Enlightenment."<br />

However, their fundamental conclusion is the same: we must<br />

reject science as a basis for true knowledge even if at the same<br />

time we recognize its practical importance or we deny, as positivists<br />

do, the possibility of any other cognitive enterprise.<br />

We must remember all these developments to understand<br />

what is at stake To what extent is science a basis for the intelligibility<br />

of nature, including man? What is the meaning of the<br />

idea of progress today?<br />

Diderot, one of the towering figures of the Enlightenment, is<br />

certainly no representative of antiscientific thought. On the<br />

contrary, his confidence in science, in the possibilities of<br />

knowledge, was total. Yet this is the very reason why science<br />

had, following Diderot, to understand life before it could hope<br />

to achieve any coherent vision of nature.<br />

We have already mentioned that the birth of modern science<br />

was marked by the abandonment of vitalist inspiration and, in<br />

particular, of Aristotelian final causes. However, the issue of<br />

the <strong>org</strong>anization of living matter remained and became a challenge<br />

for classical science. Diderot, at the height of the Newtonian<br />

triumph, emphasizes that this problem was repressed by<br />

physics. He imagines it as haunting the dreams of physicists<br />

who cannot conceive of it while they are awake. The physicist<br />

d /\lembert is dreaming:<br />

· living point . .. No, that's wrong. Nothing at all to<br />

begin with, and then a living point. This living point is<br />

joined by another, and then another, and from these successive<br />

joinings there results a unified being, for I am a


81 THE TWO CULTURES<br />

unity, of that I am certain. . . . (As he said this he felt<br />

himself all over.) But how did this unity come about?"<br />

"Now listen, Mr. Philosopher, I can understand an aggregate,<br />

a tissue of tiny sensitive bodies, but an animal! ...<br />

A whole, a system that is a unit, an individual conscious<br />

of its own unity! I can't see it, no, I can't see it. "2<br />

In an imaginary conversation with dj\lembert, Oiderot speaks<br />

in the first person, demonstrating the inadequacy of mechanistic<br />

explanation:<br />

Look at this egg: with it you can overthrow all the schools<br />

of theology and all the churches in the world. What is this<br />

egg? An insensitive mass before the germ is put into<br />

it ... How does this mass evolve into a new <strong>org</strong>anization,<br />

into sensitivity, into life? Through heat. What will<br />

generate heat in it? Motion. What will the successive<br />

effects of motion be? Instead of answering me, sit down<br />

and let us follow out these effects with our eyes from one<br />

moment to the next. First there is a speck which moves<br />

about, a thread growing and taking colour, flesh being<br />

formed, a beak, wing-tips, eyes, feet coming into view, a<br />

yellowish substance which unwinds and turns into intestines-and<br />

you have a living creature .... Now the wall<br />

is breached and the bird emerges, walks, flies, feels pain,<br />

runs away, comes back again, complains, suffers, loves,<br />

desires, enjoys, it experiences all your affections and<br />

does all the things you do. And will you maintain, with<br />

Descartes, that it is an imitating machine pure and simple?<br />

Why, even little children will laugh at you, and philosophers<br />

will answer that if it is a machine you are one<br />

too! If, however, you admit that the only difference between<br />

you and an animal is one of <strong>org</strong>anization, you will<br />

be showing sense and reason and be acting in good faith;<br />

but then it will be concluded, contrary to what you had<br />

said, that from an inert substance arranged in a certain<br />

way and impregnated by another inert substance, subjected<br />

to heat and motion, you will get sensitivity, life,<br />

memory, consciousness, passions, thought ... Just listen<br />

to your own arguments and you will feel how pitiful


ORDER OUT OF CHAOS 82<br />

they are. You will come to feel that by refusing to entertain<br />

a simple hypothesis that explains everything-sensitivity<br />

as a property common to all matter or as a result<br />

of the <strong>org</strong>anization of matter-you are flying in the face of<br />

common sense and plunging into a chasm of mysteries,<br />

contradictions and absurdities.3<br />

In opposition to rational mechanics, to the claim that material<br />

nature is nothing but inert mass and motion, Diderot appeals<br />

to one of physics' most ancient sources of inspiration,<br />

namely, the growth, differentiation, and <strong>org</strong>anization of the<br />

embryo. Flesh forms, and so does the beak, the eyes, and the<br />

intestines; a gradual <strong>org</strong>anization occurs in biological "space,"<br />

out of an apparently homogeneous environment differentiated<br />

forms appear at exactly the right time and place through the<br />

effects of complex and coordinated processes.<br />

How can an inert mass, even a Newtonian mass animated<br />

by the forces of gravitational interaction, be the starting point<br />

for <strong>org</strong>anized active local structures? We have seen that the<br />

Newtonian system is a world system: no local configuration of<br />

·bodies can claim a particular identity; none is more than a<br />

contingent proximity between bodies connected by general relations.<br />

But Diderot does not despair. Science is only beginning; rational<br />

mechanics is merely a first, overly abstract attempt. The<br />

spectacle of the embryo is enough to refute its claims to universality.<br />

This is why Diderot compares the work of great<br />

"mathematicians" such as Euler, Bernoulli, and di\lembert to<br />

the pyramids of the Egyptians, awe-inspiring witnesses to the<br />

genius of their builders, now lifeless ruins, alone and forlorn.<br />

True science, alive and fruitful, will be carried on elsewhere."<br />

Moreover, it seems to him that this new science of <strong>org</strong>anized<br />

living matter has already begun. His friend d'Holbach is busy<br />

studying chemistry, Diderot himself has chosen medicine. The<br />

problem in chemistry as well as in medicine is to replace inert<br />

matter with active matter capable of <strong>org</strong>anizing itself and producing<br />

living beings. Diderot claims that matter has to be sensitive.<br />

Even a stone has sensation in the sense that the<br />

molecules of which it is composed actively seek certain combinations<br />

rather than others and thus are governed their likes<br />

and dislikes. The sensitivity of the whole <strong>org</strong>anism is then


83<br />

THE TWO CULTURES<br />

simply the sum of that of its parts, just as a swarm of bees with<br />

its globally coherent behavior is the result of interactions between<br />

one bee and another; and, Diderot thereby concludes,<br />

the human soul does not exist any more than the soul of the<br />

beehive does. s<br />

Diderot's vitalist protest against physics and the universal<br />

laws of motion thus stems from his rejection of any form of<br />

spiritualist dualism. Nature must be described in such a way<br />

that man's very existence becomes understandable. Otherwise,<br />

and this is what happens in the mechanistic world view,<br />

the scientific description of nature will have its counterpart in<br />

man as an automaton endowed with a soul and thereby alien to<br />

nature.<br />

The twofold basis of materialistic naturalism, at once chemical<br />

and medical, that Diderot employed to counter the physics<br />

of his time is recurrent in the eighteenth century. While biologists<br />

speculated about the animal-machine, the preexistence of<br />

germs, and the chain of living creatures-all problems close to<br />

theology6-chemists and physicians had to face directly the<br />

complexity of real processes in both chemistry and life. Chemistry<br />

and medicine were, in the late eighteenth century, privileged<br />

sciences for those who fought against the physicists'<br />

esprit de systeme in favor of a science that would take into<br />

account the diversity of natural processes. A physicist could<br />

be pure esprit, a precocious child, but a physician or a chemist<br />

must be a man of experience: he must be able to decipher the<br />

signs, to spot the clues. In this sense, chemistry and medicine<br />

are arts. They demand judgment, application, and tenacious<br />

observation. Chemistry is a madman's passion, Venel concluded<br />

in the article he wrote for Diderot's Encyclopedie, an<br />

eloquent defense of chemistry against the abstract imperialism<br />

of the Newtonians.7 To emphasize the fact that protests raised<br />

by chemists and physicians against the way physicists reduced<br />

living processes to peaceful mechanisms and the quiet unfolding<br />

of universal laws were common in Diderot's day, we invoke<br />

the eminent figure of Stahl, the father of vitalism and inventor<br />

of the first consistent chemical systematics.<br />

According to Stahl, universal laws apply to the living only in<br />

the sense that these laws condemn them to death and corruption;<br />

the matter of which living beings are composed is so frail,<br />

so easily decomposed, that if it were governed solely by the


ORDER OUT OF CHAOS<br />

84<br />

common laws of matter, it would not withstand decay or dissolution<br />

for a moment. If a living creature is to survive in spite<br />

of the general laws of physics, however short its life when it is<br />

compared to that of a stone or another inanimate object, it has<br />

to possess in itself a "principle of conservation" that maintains<br />

the harmonious equilibrium of the texture and structure<br />

of its body. The astonishing longevity of a living body in view<br />

of the extreme corruptibility of its constitutive matter is thus<br />

indicative of the action of a "natural, permanent, immanent<br />

principle," of a particular cause that is alien to the laws of<br />

inanimate matter and that constantly struggles against the constantly<br />

active corruption whose inevitability these laws imply. s<br />

To us this analysis of life sounds both near and remote. It is<br />

close to us in its acute awareness of the singularity and the precariousness<br />

of life. It is remote because, like Aristotle, Stahl<br />

defined life in static terms, in terms of conservation, not of<br />

becoming or evolution. Still, the terminology used by Stahl<br />

can be found in recent biological literature, for example,<br />

where we read that enzymes "combat" decay and allow the<br />

body to ward off the death to which it is inexorably doomed by<br />

physics. Here also, biological <strong>org</strong>anization defies the laws of<br />

nature, and the only "normal" trend is that which leads to<br />

death (see Chapter V).<br />

Indeed, Stahl's vitalism is relevant as long as the laws of<br />

physics are identified with evolution toward decay and dis<strong>org</strong>anization.<br />

Today the "vitalist principle" has been superseded<br />

by the succession of improbable mutations preserved in the<br />

genetic message "governing" the living structure. Nonetheless,<br />

some extrapolations starting from molecular biology relegate<br />

life to the confines of nature-that is, conclude life is<br />

compatible with the basic laws of physics but purely contingent.<br />

This was explicitly stated by Monod: life does not "follow<br />

from the laws of physics, it is compatible with them. Life<br />

is an event whose singularity we have to recognize."<br />

But the transition from matter to life can also be viewed in a<br />

different way. As we shall see, far from equilibrium, new self<strong>org</strong>anizational<br />

processes arise. (These questions will be studied<br />

in detail in Chapters V and VI.) In this way biological<br />

<strong>org</strong>anization begins to appear as a natural process.<br />

However, long before these recent developments, the problematics<br />

of life had been transformed. In a politically trans-


85 THE TWO CULTURES<br />

formed Europe the intellectual landscape was remodeled as<br />

the Romantic movement, closely linked with the Counter­<br />

Enlightenment, shows.<br />

Stahl criticized the metaphor of the automaton because, unlike<br />

a living being, the purpose of an automaton does not lie<br />

within itself; its <strong>org</strong>anization is imposed upon it by its maker.<br />

Diderot, far from situating the study of life outside the reach of<br />

science, saw it as representing the future of a science he considered<br />

to be still in its infancy. A few years later, such points<br />

of view were to be challenged.9 Mechanical change, activity as<br />

described by the laws of motion, had now become synonymous<br />

with the artificial and with death. Opposed to it,<br />

united in a complex with which we are now quite familiar, were<br />

the concepts of life, spontaneity, freedom, and spirit. This opposition<br />

was paralleled by the opposition between calculation<br />

and manipulation on the one hand, and the free speculative<br />

activity of the mind on the other. Through speculation the philosopher<br />

would reach the spiritual activity at the core of nature.<br />

As for the scientist, his concern with nature would be<br />

reduced to taking it as a set of manipulable and measurable<br />

objects; he would thus be able to take possession of nature, to<br />

dominate and control it but not understand it. Thus the intelligibility<br />

of nature would lie beyond the grasp of science.<br />

We are not concerned here with the history of philosophy<br />

but merely with emphasizing the extent to which the philosophical<br />

criticism of science had at this time become harsher,<br />

resembling certain modern forms of antiscience. It was no<br />

longer a question of refuting rather naive and shortsighted<br />

generalizations that only have to be repeated aloud-to use<br />

Diderot's language-to make even children laugh, but of refuting<br />

the type of approach that produced experimental and<br />

mathematical knowledge of nature. Scientific knowledge is not<br />

being criticized for its limitations but for its nature, and a rival<br />

knowledge, based on another approach, is being announced.<br />

Knowledge is fragmented into two opposed modes of inquiry.<br />

From a philosophical point of view, the transition from Diderot<br />

to the Romantics and, more precisely, from one of these<br />

two types of critical attitudes toward science to the other, can<br />

be found in Kant's transcendental philosophy, the essential<br />

point being that the Kantian critique identified science in general<br />

with its Newtonian realization. It thereby branded as im-


ORDER OUT OF CHAOS 86<br />

possible any opposition to classical science that was not an<br />

opposition to science itself. Any criticism against Newtonian<br />

physics must then be seen as aimed at downgrading the rational<br />

understanding of nature in favor of a different form of<br />

knowledge. Kant's approach had immense repercussions,<br />

which continue down to our day. Let us therefore summarize<br />

his point of view as presented in Critique of Pure Reason,<br />

which, in opposition to the progressist views of the Enlightenment,<br />

presents the closed and limiting conception of science<br />

we have just defined.<br />

Kants Critical Ratification<br />

How to restore order in the intellectual landscape left in disarray<br />

with the disappearance of God conceived as the rational<br />

principle that links science and nature? How could scientists<br />

ever have access to global truth when it could no longer be<br />

asserted, except metaphorically, that science deciphers the<br />

word of creation? God was now silent or at least no longer<br />

spoke the same language as human reason. Moreover, in a nature<br />

from which time was eliminated, what remained of our<br />

subjective experience? What was the meaning of freedom,<br />

destiny, or ethical values?<br />

Kant argued that there were two levels of reality: a phenomenal<br />

level that corresponds to science, and a noumenal level<br />

corresponding to ethics. The phenomenal order is created by<br />

the human mind. The noumenal level transcends man's intellect;<br />

it corresponds to a spiritual reality that supports his ethical<br />

and religious life. In a way, Kant's solution is the only one<br />

possible for those who assert both the reality of ethics and the<br />

reality of the objective world as it is expressed by classical<br />

science. Instead of God, it is now man himself who is the<br />

source of the order he perceives in nature. Kant justifies both<br />

scientific knowledge and man's alienation from the phenomenal<br />

world described by science. From this perspective we can<br />

see that Kantian philosophy explicitly spells out the philosophical<br />

content of classical science.<br />

Kant defines the subject of critical philosophy as transcendental.<br />

It is not concerned with the objects of experience but<br />

is based on the a priori fact that a systematic knowledge of


87 THE TWO CULTURES<br />

these objects is possible (this is for him proved by the existence<br />

of physics), going on to state the a priori conditions of<br />

possibility for this mode of knowledge.<br />

To do so a distinction must be made between the direct sensations<br />

we receive from the outside world and the objective,<br />

"rational" mode of knowledge. Objective knowledge is not<br />

passive; it forms its objects. When we take a phenomenon as<br />

the object of experience, we assume a priori before we actually<br />

experience it that it obeys a given set of principles. Insofar as it<br />

is perceived as a possible object of knowledge, it is the product<br />

of our mind's synthetic activity. We find ourselves in the<br />

objects of our knowledge, and the scientist himself is thus the<br />

source of the universal laws he discovers in nature.<br />

The a priori conditions of experience are also the conditions<br />

for the existence of the objects of experience. This celebrated<br />

statement sums up the "Copernican revolution" achieved by<br />

Kant's "transcendental" inquiry. The subject no longer "revolves"<br />

around its object, seeking to discover the laws by<br />

which it is governed or the language by which it may be deciphered.<br />

Now the subject itself is at the center, imposing its<br />

laws, and the world perceived speaks the language of that subject.<br />

No wonder, then, that Newtonian science is able to describe<br />

the world from an external, almost divine point of view!<br />

That all perceived phenomena are governed by the laws of<br />

our mind does not mean that a concrete knowledge of these<br />

objects is useless. According to Kant, science does not engage<br />

in a dialogue with nature but imposes its own language upon it.<br />

Still it must discover, in each case, the specific message expressed<br />

in this general language. A knowledge of the a priori<br />

concepts alone is vain and empty.<br />

From the Kantian point of view Laplace's demon, the symbol<br />

of the scientific myth, is an illusion, but it is a rational<br />

illusion. Although it is the result of a limiting process and, as<br />

such, illegitimate, it is still the expression of a legitimate conviction<br />

that is the driving force of science-the conviction<br />

that, in its entirety, nature is rightfully subjected to the laws<br />

that scientists succeed in deciphering. Wherever it goes, whatever<br />

it questions, science will always obtain, if not the same<br />

answer, at least the same kind of answer. There exists a single<br />

universal syntax that includes all possible answers.<br />

Transcendental philosophy thus ratified the physicist's


ORDER OUT OF CHAOS 88<br />

claim to have found the definitive form of all positive knowledge.<br />

At the same time, however, it secured for philosophy a<br />

dominant position in. respect to science. It was no longer necessary<br />

to look for the philosophic significance of the results of<br />

scientific activity. From the transcendental standpoint, those<br />

results cannot lead to anything really new. It is science, not its<br />

results, that is the subject of philosophy; science taken as a<br />

repetitive and closed enterprise provides a stable foundation<br />

for transcendental reflection.<br />

Therefore, while it ratifies all the claims of science, Kant's<br />

critical philosophy actually limits scientific activity to problems<br />

that can be considered both easy and futile. It condemns<br />

science to the tedious task of deciphering the monotonous language<br />

of phenomena while keeping for itself questions of human<br />

"destiny": what man may know, what he must do, what<br />

he may hope for. The world studied by science, the world accessible<br />

to positive knowledge is "only" the world of phenomena.<br />

Not only is the scientist unable to know things in themselves,<br />

but even the questions he asks are irrelevant to the real problems<br />

of mankind. Beauty, freedom, and ethics cannot be objects<br />

of positive knowledge. They belong to the noumenal<br />

world, which is the domain of philosophy, and they are quite<br />

unrelated to the phenomenal world.<br />

We can accept Kant's starting point, his emphasis on the<br />

active role man plays in scientific description. Much has already<br />

been said about experimentation as the art of choosing<br />

situations that are hypothetically governed by the law under<br />

investigation and staging them to give clear, experimental answers.<br />

For each experiment certain principles are presupposed<br />

and thus cannot be established by that experiment.<br />

However, as we have seen, Kant goes much further. He denies<br />

the diversity of possible scientific points of view, the diversity<br />

of presupposed principles. In agreement with the myth of classical<br />

science, Kant is after the unique language that science<br />

deciphers in nature, the unique set of a priori principles on<br />

which physics is based and that are thus to be identified with<br />

the categories of human understanding. Thus Kant denies the<br />

need for the scientist's active choice, the need for a selection<br />

of a problematic situation corresponding to a particular theoretical<br />

language in which definite questions may be asked and<br />

experimental answers sought.


89 THE TWO CULTURES<br />

Kant's critical ratification defines scientific endeavor as<br />

silent and systematic, closed within itself. By so doing, philosophy<br />

endorses and perpetuates the rift, debasing and surrendering<br />

the whole field of positive knowledge to science<br />

while retaining for itself the field of freedom and ethics, conceived<br />

as alien to nature.<br />

A Prlilosophy of Nature? Hegel and Bergson<br />

The Kantian truce between science and philosophy was a fragile<br />

one. Post-Kantian philosophers disrupted this truce in favor of<br />

a new philosophy of science, presupposing a new path to knowledge<br />

that was distinct from science and actually hostile to it.<br />

Speculation released from the constraints of any experimental<br />

dialogue reigned supreme, with disastrous consequences for<br />

the dialogue between scientists and philosophers. For most<br />

scientists, the philosophy of nature became synonymous with<br />

arrogant, absurd speculation riding roughshod over facts, and<br />

indeed regularly proven wrong by the facts. On the other side,<br />

for most philosophers it has become a symbol of the dangers<br />

involved in dealing with nature and in competing with science.<br />

The rift among science, philosophy, and humanistic studies<br />

was thus made greater by mutual disdain and fear.<br />

As an example of this speculative approach to nature, let us<br />

first consider Hegel. Hegel's philosophy has cosmic dimensions.<br />

In his system increasing levels of complexity are specified,<br />

and nature's purpose is the eventual self-realization of its<br />

spiritual element. Nature's history is fulfilled with the appearance<br />

of man-that is, with the coming of Spirit apprehending<br />

itself.<br />

The Hegelian philosophy of nature systematically incorporates<br />

all that is denied by Newtonian science. In particular, it<br />

rests on the qualitative difference between the simple behavior<br />

described by mechanics and the behavior of more complex entities<br />

such as living beings. It denies the possibility of reducing<br />

those levels, rejecting the idea that differences are merely apparent<br />

and that nature is basically homogeneous and simple. It<br />

affirms the existence of a hierarchy, each level of which presupposes<br />

the preceding ones.


ORDER OUT OF CHAOS<br />

90<br />

Unlike the Newtonian authors of romans de Ia matiere, of<br />

world-embracing panoramas ranging from gravitational interactions<br />

to human passions, Hegel knew perfectly well that his<br />

distinctions among levels (which, quite apart from his own interpretation,<br />

we may acknowledge as corresponding to the<br />

idea of an increasing complexity in nature and to a concept of<br />

time whose significance would be richer on each new level)<br />

ran counter to his day's mathematical science of nature. He<br />

therefore set out to limit the significance of this science, to<br />

show that mathematical description is restricted to the most<br />

trivial situations. Mechanics can be mathematized because it<br />

attributes only space-time properties to matter. · brick does<br />

not kill a man merely because it is a brick, but solely because<br />

of its acquired velocity; this means that the man is killed tzy<br />

space and time." IO The man is killed by what we call kinetic<br />

energy (mv2/2)-by an abstract quantity defining mass and velocity<br />

as interchangeable; the same murderous effect can be<br />

achieved by reducing one and increasing the other.<br />

It is precisely this interchangeability that Hegel sets as a<br />

condition for mathematization that is no longer satisfied when<br />

the mechanical level of description is abandoned for a "higher"<br />

one involving a larger spectrum of physical properties.<br />

In a sense Hegel's system provides a consistent philosophic<br />

response to the crucial problems of time and complexity.<br />

However, for generations of scientists it represented the epitome<br />

of abhorrence and contempt. In a few years, the intrinsic<br />

difficulties of Hegel's philosophy of nature were aggravated by<br />

the obsolescence of the scientific background on which his<br />

system was based, for Hegel, of course, based his rejection of<br />

the Newtonian system on the scientific conceptions of his<br />

time.11 And it was precisely those conceptions that were to fall<br />

into oblivion with astonishing speed. It is difficult to imagine a<br />

less opportune time than the beginning of the nineteenth century<br />

for seeking experimental and theoretical support for an<br />

alternative to classical science. Although this time was characterized<br />

by a remarkable extension of the experimental scope of<br />

science (see Chapter IV) and by a proliferation of theories that<br />

seemed to contradict Newtonian science, most of those theories<br />

had to be given up only a few years after their appearance.<br />

At the end of the nineteenth c:entury, when Bergson under-


91 THE TWO CULTURES<br />

took his search for an acceptable alternative to the science of<br />

his time, he turned to intuition as a form of speculative knowledge,<br />

but he presented it as quite different from that of the<br />

Romantics. He explicitly stated that intuition is unable to produce<br />

a system but produces only results that are always partial<br />

and nongeneralizable, results to be formulated with great caution.<br />

In contrast, generalization is an attribute of "intelligence,"<br />

the greatest achievement of which is classical<br />

science. Bergsonian intuition is a concentrated attention, an<br />

increasingly difficult attempt to penetrate deeper into the singularity<br />

of things. Of course, to communicate, intuition must<br />

have recourse to language-"in order to be transmitted, it will<br />

have to use ideas as a conveyance." 12 This it does with infinite<br />

patience and circumspection, at the same time accumulating<br />

images and comparisons in order to "embrace reality," 13 thus<br />

suggesting in an increasingly precise way what cannot be communicated<br />

by means of general terms and abstract ideas.<br />

Science and intuitive metaphysics "are or can become<br />

equally precise and definite. They both bear upon reality itself.<br />

But each one of them retains only half of it so that one<br />

could see in them, if one wished, two subdivisions of science<br />

or two departments of metaphysics, if they did not mark divergent<br />

directions of the activity of thought." 14<br />

The definition of these two divergent directions may also be<br />

considered as the historical consequence of scientific evolution.<br />

For Bergson, it is no longer a question of finding scientific<br />

alternatives to the physics of his time. In his view,<br />

chemistry and biology had definitely chosen mechanics as<br />

their model. The hopes that Diderot had cherished for the future<br />

of chemistry and medicine had thus been dashed. In<br />

Bergson's view, science is a whole and must therefore be<br />

judged as a whole. And this is what he does when he presents<br />

science as the product of a practical intelligence whose aim is<br />

to dominate matter and that develops by abstraction and generalization<br />

the intellectual categories needed to achieve this<br />

domination. Science is the product of our vital need to exploit<br />

the world, and its concepts are determined by the necessity of<br />

manipulating objects, of making predictions, and of achieving<br />

reproducible actions. This is why rational mechanics represents<br />

the very essence of science, its actual embodiment. The


ORDER OUT OF CHAOS 92<br />

other sciences are more vague, awkward manifestations of an<br />

approach that is all the more successful the more inert and<br />

dis<strong>org</strong>anized the terrain it explores.<br />

For Bergson all the limitations of scientific rationality can<br />

be reduced to a single and decisive one: it is incapable of understanding<br />

duration since it reduces time to a sequence of<br />

instantaneous states linked by a deterministic law.<br />

"Time is invention, or it is nothing at all." 15 Nature is<br />

change, the continual elaboration of the new, a totality being<br />

created in an essentially open process of development without<br />

any preestablished model. "Life progresses and endures in<br />

time." 16 The only part of this progression that intelligence can<br />

grasp is what it succeeds in fixing in the form of manipulable<br />

and calculable elements and in referring to a time seen as<br />

sheer juxtaposition of instants.<br />

Therefore, physics "is limited to coupling simultaneities<br />

between the events that make up this time and the positions<br />

of the mobile T on its trajectory. It detaches these<br />

events from the whole, which at every moment puts on a<br />

new form and which communicates to them something of<br />

its novelty. It considers them in the abstract, such as they<br />

would be outside of the living whole, that is to say, in a<br />

time unrolled in space. It retains only the events or systems<br />

of events that can be thus isolated without being<br />

made to undergo too profound a deformation, because<br />

only these lend themselves to the application of its<br />

method. Our physics dates from the day when it was<br />

known how to isolate such systems. " 1 7<br />

When it comes to understanding duration itself, science is<br />

powerless. What is needed is intuition, a "direct vision of the<br />

mind by the mind. "18 "Pure change, real duration, is something<br />

spiritual. Intuition is what attains the spirit, duration,<br />

pure change.I9<br />

Can we say Bergson has failed in the same way that the<br />

post-Kantian philosophy of nature failed? He has failed insofar<br />

as the metaphysics based on intuition he wished to create<br />

has not materialized. He has not failed in that, unlike Hegel,<br />

he had the good fortune to pass judgment upon science that<br />

was, on the whole, firmly established-that is, classical sci-


93 THE TWO CULTURES<br />

ence at its apotheosis, and thus identified problems which are<br />

indeed still our problems. But, like the post-Kantian critics,<br />

he identified the science of his time with science in general.<br />

He thus attributed to science de jure limitations that were only<br />

de facto. As a consequence he tried to define once and for all a<br />

statu quo for the respective domains of science and other intellectual<br />

activities. Thus the only perspective remaining open<br />

for him was to introduce a way in which antagonistic approaches<br />

could at best merely coexist.<br />

In conclusion, even if the way in which Bergson sums up the<br />

achievement of classical science is still to some extent acceptable,<br />

we can no longer accept it as a statement of the eternal<br />

limits of the scientific enterprise. We conceive of it more as a<br />

program that is beginning to be implemented by the metamorphosis<br />

science is now undergoing. In particular, we know<br />

that time linked with motion does not exhaust the meaning of<br />

time in physics. Thus the limitations Bergson criticized are<br />

beginning to be overcome, not by abandoning the scientific<br />

approach or abstract thinking but by perceiving the limitations<br />

of the concepts of classical dynamics and by discovering new<br />

formulations valid in more general situations.<br />

Process and Reality: Whitehead<br />

As we have emphasized, the element common to Kant, Hegel,<br />

and Bergson is the search for an approach to reality that is<br />

different from the approach of classical science. This is also<br />

the fundamental aim of Whitehead's philosophy, which is resolutely<br />

pre-Kantian. In his most important book, Process and<br />

Reality, he puts us back in touch with the great philosophies of<br />

the Classical Age and their quest for rigorous conceptual experimentation.<br />

Whitehead sought to understand human experience as a pro<br />

cess belonging to nature, as physical existence. This challenge<br />

led him, on the one hand, to reject the philosophic tradition<br />

that defined subjective experience in terms of consciousness,<br />

thought, and sense perception, and, on the other, to conceive<br />

of all physical existence in terms of enjoyment, feeling, urge,<br />

appetite, and yearning-that is, to cross swords with what he


ORDER OUT OF CHAOS 94<br />

calls "scientific materialism," born in the seventeenth century.<br />

Like Bergson, Whitehead was thus led to point out the<br />

basic inadequacies of the theoretical scheme developed by<br />

seventeenth-century science:<br />

The seventeenth century had finally produced a scheme<br />

of scientific thought framed by mathematicians, for the<br />

use of mathematicians. The great characteristic of the<br />

mathematical mind is its capacity for dealing with abstractions;<br />

and for eliciting from them clear-cut demonstrative<br />

trains of reasoning, entirely satisfactory so long<br />

as it is those abstractions which you want to think about.<br />

The enormous success of the scientific abstractions,<br />

yielding on the one hand matter with its simple location<br />

in space and time, on the other hand mind, perceiving,<br />

suffering, reasoning, but not interfering, has foisted on to<br />

philosophy the task of accepting them as the most concrete<br />

rendering of fact.<br />

Thereby, modern philosophy has been ruined. It has<br />

oscillated in a complex manner between three extremes.<br />

There are the dualists, who accept matter and mind as on<br />

equal basis, and the two varieties of monists, those who<br />

put mind inside matter, and those who put matter inside<br />

mind. But this juggling with abstractions can never overcome<br />

the inherent confusion introduced by the ascription<br />

of misplaced concreteness to the scientific scheme of the<br />

seventeenth century. 20<br />

However, Whitehead considered this to be only a temporary<br />

situation. Science is not doomed to remain a prisoner of confusion.<br />

We have already raised the question of whether it is possible<br />

to formulate a philosophy of nature that is not directed against<br />

science. Whitehead's cosmology is the most ambitious attempt<br />

to do so. Whitehead saw no basic contradiction between<br />

science and philosophy. His purpose was to define the<br />

conceptual field within which the problem of human experience<br />

and physical processes could be dealt with consistently<br />

and to determine the conditions under which the problem<br />

could be solved. What had to be done was to formulate: the:<br />

principles necessary to characterize all forms of existence,


95<br />

THE TWO CULTURES<br />

from that of stones to that of man. It is precisely this universality<br />

that, in Whitehead's opinion, defines his enterprise as<br />

"philosophy." While each scientific theory selects and abstracts<br />

from the world's complexity a peculiar set of relations,<br />

philosophy cannot favor any particular region of human experience.<br />

Through conceptual experimentation it must construct<br />

a consistency that can accommodate all dimensions of experience,<br />

whether they belong to physics, physiology, psychology,<br />

biology, ethics, etc.<br />

Whitehead understood perhaps more sharply than anyone<br />

else that the creative evolution of nature could never be conceived<br />

if the elements composing it were defined as permanent,<br />

individual entities that maintained their identity throughout all<br />

changes and interactions. But he also understood that to make<br />

all permanence illusory, to deny being in the name of becoming,<br />

to reject entities in favor of a continuous and ever-changing<br />

flux meant falling once again into the trap always lying in wait<br />

for philosophy-to "indulge in brilliant feats of explaining<br />

away." 21<br />

Thus for Whitehead the task of philosophy was to reconcile<br />

permanence and change, to conceive of things as processes, to<br />

demonstrate that becoming forms entities, individual identities<br />

that are born and die. It is beyond the scope of this book to<br />

give a detailed presentation of Whitehead's system. Let us<br />

only emphasize that he demonstrated the connection between<br />

a philosophy of relation-no element of nature is a permanent<br />

support for changing relations; each receives its identity from<br />

its relations with others-and a philosophy of innovating becoming.<br />

In the process of its genesis, each existent unifies the<br />

multiplicity of the world, since it adds to this multiplicity an<br />

extra set of relations. At the creation of each new entity "the<br />

many become one and are increased by one. "22<br />

In the conclusion of this book, we shall again encounter<br />

Whitehead's question of permanence and change, this time as<br />

it is raised in physics; we shall speak of entities formed by<br />

their irreversible interaction with the world. Today physics has<br />

discovered the need to assert both the distinction and interdependence<br />

between units and relations. It now recognizes that,<br />

for an interaction to be real, the "nature" of the related things<br />

must derive from these relations, while at the same time the relations<br />

must derive from the "nature" of the things (see Chap-


ORDER OUI OF CHAOS<br />

96<br />

ter X). This is the forerunner of "self-consistent" descriptions<br />

as expressed, for instance, by the "bootstrap" philosophy in<br />

elementary-particle physics, which asserts the universal connectedness<br />

of all particles. However, when Whitehead wrote<br />

Process and Reality, the situation of physics was quite different,<br />

and Whitehead's philosophy found an echo only in biology.2<br />

3<br />

Whitehead's case as well as Bergson's convince us that only<br />

an opening, a widening of science can end the dichotomy between<br />

science and philosophy. This widening of science is possible<br />

only if we revise our conception of time. To deny timethat<br />

is, to reduce it to a mere deployment of a reversible lawis<br />

to abandon the possibility of defining a conception of nature<br />

coherent with the hypothesis that nature produced living<br />

beings, particularly man. It dooms us to choosing between an<br />

antiscientific philosophy and an alienating science.<br />

"Ignoramus, lgnoramibus": The Positivists Strain<br />

Another method of overcoming the difficulties of classical rationality<br />

implied in classical science was to separate what was<br />

scientifically most fruitful from what is "true." This is another<br />

form of the Kantian cleavage. In his 1865 address "On the Goal<br />

of the Natural Sciences," Kirchoff stated that the ultimate<br />

goal of science is to reduce every phenomenon to motion, mo<br />

tion that in turn is described by theoretical mechanics. A similar<br />

statement was made by Helmholtz, a chemist, physician,<br />

physicist, and physiologist who dominated the German universities<br />

at the time when they were becoming the hub of European<br />

science. He stated: "the phenomena of nature are to be<br />

referred back to motions of material particles possessing unchangeable<br />

moving forces, which are dependent upon conditions<br />

of space alone. "2 4<br />

The aim of the natural sciences, therefore, was to reduce all<br />

observations to the laws formulated by Newton and extended<br />

by such illustrious physicists and mathematicians as Lagrange,<br />

Hamilton, and others. We were not to ask why these forces<br />

exist and enter Newton's equation. In any case, we could not<br />

"understand" matter or forces even if we used these concepts


97 THE TWO CULTURES<br />

to formulate the laws of dynamics. The why, the basic nature<br />

of forces and masses, remains hidden from us. Du Bois Reymond,<br />

as we already mentioned, expressed concisely the<br />

limitations of our knowledge: "Ignoramus, ignoramibus." Science<br />

provides no access to the mysteries of the universe. What<br />

then is science?<br />

We have already referred to Mach's influential view: Science<br />

is part of the Darwinian struggle for life. It helps us to <strong>org</strong>anize<br />

our experience. It leads to an economy of thought. Mathematical<br />

laws are nothing more than conventions useful for summarizing<br />

the results of possible experiments. At the end of the<br />

nineteenth century, scientific positivism exercised a great intellectual<br />

appeal. In France it influenced the work of eminent<br />

thinkers such as Duhem and Poincare.<br />

One more step in the elimination of "contemptible metaphysics"<br />

and we come to the Vienna school. Here science is<br />

granted jurisdiction over all positive knowledge and philosophy<br />

needed to keep this positive knowledge in .order. This<br />

meant a radical submission of all rational knowledge and questions<br />

to science. When Reichenbach, a distinguished neopositivist<br />

philosopher, wrote a book on the "direction of<br />

time," he stated:<br />

There is no other way to solve the problem of time than<br />

the way through physics. More than any other science,<br />

physics has been concerned with the nature of time. If<br />

time is objective the physicist must have discovered the<br />

fact. If there is Becoming, the physicist must know it; but<br />

if time is merely subjective and Being is timeless, the<br />

physicist must have been able to ignore time in his construction<br />

of reality and describe the world without the<br />

help of time .... It is a hopeless enterprise to search for<br />

the nature of time without studying physics. If there is a<br />

solution to the philosophical problem of time, it is written<br />

down in the equations of mathematical physics.25<br />

Reichenbach's work is of great interest to anyone wishing to<br />

see what physics has to say on the subject of time, but it is not<br />

so much a book on the philosophy of nature as an account of<br />

the way in which the problem of time challenges scientists, not<br />

philosophers.


ORDER OUT OF CHAOS 98<br />

What then is the role of philosophy? It has often been said<br />

that philosophy should become the science of science. Philosophy's<br />

objective would then be to analyze the methods of<br />

science, to axiomatize and to clarify the concepts used. Such<br />

a role would make of the former "queen of sciences" something<br />

like their housemaid. Of course, there is the possibility<br />

that this clarification of concepts would permit further progress,<br />

that philosophy understood in this way would, through<br />

the use of other methods-logic, semantics-produce new<br />

knowledge comparable to that of science proper. It is this hope<br />

that sustains the "analytic philosophy" so prevalent in Anglo­<br />

American circles. We do not want to minimize the interest of<br />

such an inquiry. However, the problems that concern us here<br />

are quite different. We do not aim to clarify or axiomatize existing<br />

knowledge but rather to close some fundamental gaps in<br />

this knowledge.<br />

A New Start<br />

In the first part of this book we described, on the one hand,<br />

dialogue with nature that classical science made possible and,<br />

on the other, the precarious cultural position of science. Is<br />

there a way out? In this chapter we have discussed some attempts<br />

to reach alternative ways of knowledge. We have also<br />

considered the positivist view, which separates science from<br />

reality.<br />

The moments of greatest excitement at scientific meetings<br />

very often occur when scientists discuss questions that are<br />

likely to have no practical utility whatsoever, no survival<br />

value-topics such as possible interpretations of quantum mechanics,<br />

or the role of the expanding universe in our concept<br />

of time. If the positivistic view, which reduces science to a<br />

symbolic calculus, was accepted , science would lose much of<br />

its appeal. Newton's synthesis between theoretical concepts<br />

and active knowledge would be shattered. We would be back<br />

to the situation familiar from the time of Greece and Rome,<br />

with an unbridgeable gap between technical, practical knowledge<br />

on one side and theoretical knowledge on the other.<br />

For the ancients. nature was a source of wisdom. Medieval<br />

nature spoke of God. In modern times nature has become so


99<br />

THE TWO CULTURES<br />

silent that Kant considered that science and wisdom, science<br />

and truth, ought to be completely separated. We have been<br />

living with this dichotomy for the past two centuries. It is time<br />

for it to come to an end. As far as science is concerned, the<br />

time is ripe for this to happen. From our present perspective,<br />

the first step toward a possible reunification of knowledge was<br />

the discovery in the nineteenth century of the theory of heat,<br />

of the laws of thermodynamics. Thermodynamics appears as<br />

the first form of a "science of complexity." This is the science<br />

we now wish to describe, from its formulation to recent developments.


I<br />

I


BOOK TWO<br />

THE SCIENCE OF<br />

COMPLEXITY


I<br />

I<br />

I<br />

I<br />

I


CHAPTER IV<br />

ENERGY AND THE<br />

INDUSTRIAL AGE<br />

Heat, the Rival of Gravitation<br />

Ignis mutat res. Ageless wisdom has always linked chemistry<br />

to the "science of fire." Fire became part of experimental science<br />

during the eighteenth century, starting a conceptual<br />

transformation that forced science to reconsider what it had<br />

previously .. rejected in the name of a mechanistic world view,<br />

topics such as irreversibility and complexity.<br />

Fire transforms matter; fire leads to chemical reactions, to<br />

processes such as melting and evaporation. Fire makes fuel<br />

burn and release heat. Out of all this common knowledge,<br />

nineteenth-century science concentrated on the single fact<br />

that combustion produces heat and that heat may lead to an<br />

increase in volume; as a result, combustion produces work.<br />

Fire leads, therefore, to a new kind of machine, the heat engine,<br />

the technological innovation on which industrial society<br />

has been founded. I<br />

It is interesting to note that Adam Smith was working on his<br />

Wealth of Nations and collecting data on the prospects and<br />

determinants of industrial growth at the same university at<br />

which James Watt was putting the finishing touches on his<br />

steam engine. Yet the only use for coal that Adam Smith could<br />

find was to provide heat for workers. In the eighteenth century,<br />

wind, water, and animals, and the simple machines<br />

driven by them, were still the only conceivable sources of<br />

power.<br />

The rapid spread of the British steam engine brought about<br />

a new interest in the mechanical effect of heat, and thermodynamics.<br />

born out of this interest, was thus not so much<br />

103


ORDER OUT OF CHAOS<br />

104<br />

concerned with the nature of heat as with heat's possibilities<br />

for producing "mechanical energy."<br />

As for the birth of the "science of complexity," we propose<br />

to date it in 1811, the year Baron Jean-Joseph Fourier, the prefect<br />

oflsere, won the prize of the French Academy of Sciences<br />

for his mathematical description of the propagation of heat in<br />

solids.<br />

The result stated by Fourier was surprisingly simple and elegant:<br />

heat flow is proportional to the gradient of temperature.<br />

It is remarkable that this simple law applies to matter, whether<br />

its state is solid, liquid, or gaseous. Moreover, it remains valid<br />

whatever the chemical composition of the body is, whether it<br />

is iron or gold. It is only the coefficient of proportionality between<br />

the heat flow and the gradient of temperature that is<br />

specific to each substance.<br />

Obviously, the universal character of Fourier's law is not<br />

directly related to dynamic interactions as expressed by Newton's<br />

law, and its formulation may thus be considered the starting<br />

point of a new type of science. Indeed, the simplicity of<br />

Fourier's mathematical description of heat propagation stands<br />

in sharp contrast to the complexity of matter considered from<br />

the molecular point of view. A solid, a gas, or a liquid are<br />

macroscopic systems formed by an immense number of molecules,<br />

and yet heat conductivity is described by a single law.<br />

Fourier formulated his result at the time when Laplace's<br />

school dominated European science. Laplace, Lagrange, and<br />

their disciples vainly joined forces to criticize Fourier's theory,<br />

but they were forced to retreat. 2 At the peak of its glory,<br />

the Laplacian dream met with its first setback. A physical theory<br />

had been created that was every bit as mathematically<br />

rigorous as the mechanical laws of motion but that remained<br />

completely alien to the Newtonian world. From this time on,<br />

mathematics, physics, and Newtonian science ceased to be<br />

synonymous.<br />

The formulation of the law of heat conduction had a lasting<br />

influence. Curiously, in France and Britain it was the starting<br />

point of different historical paths leading to our time.<br />

In France, the failure of Laplace's dream led to the positivist<br />

classification of science into the well-defined compartments<br />

introduced by Auguste Comte. The Comtean division of science<br />

has been well analyzed by Michel Serres3-heat and


105 ENERGY AND THE INDUSTRIAL AGE<br />

gravity, two universals, coexist in physics. Worse, as Comte<br />

was to state later, they are antagonistic. Gravitation acts on an<br />

inert mass that submits to it without being affected by it in any<br />

other way than by the motion it acquires or transmits. Heat<br />

transforms matter, determines changes of state, and leads to a<br />

modification of intrinsic properties. This was, in a sense, a<br />

confirmation of the protest made by the anti-Newtonian chemists<br />

of the eighteenth century and by all those who emphasized<br />

the difference between the purely spatiotemporal behavior at-<br />

. tributed to mass and the specific activity of matter. This distinction<br />

was used as a foundation for the classification of the<br />

sciences, all placed by Comte under the common sign of<br />

order-that is, of equilibrium. To the mechanical equilibrium<br />

between forces the positivist classification simply adds the<br />

concept of thermal equilibrium.<br />

In Britain, on the other hand, the theory of heat propagation<br />

did not mean giving up the attempt to unite the fields of knowledge<br />

but opened a new line of inquiry, the progressive formulation<br />

of a theory of irreversible processes.<br />

Fourier's law, when applied to an isolated body with an unhomogeneous<br />

temperature distribution, describes the gradual<br />

onset of thermal equilibrium. The effect of heat propagation is<br />

to equalize progressively the distribution of temperature until<br />

homogeneity is reached. Everyone knew that this was an irreversible<br />

process; a century before, Boerhave had stressed that<br />

heat always spread and leveled out. The science of complex<br />

phenomena-involving interaction among a large number of<br />

particles-and the occurrence of temporal asymmetry thus<br />

were linked from the outset. But heat conduction did not become<br />

the starting point of an investigation into the nature of<br />

irreversibility before it was first linked with the notion of dissipation<br />

as seen from an engineering point of view. 4<br />

Let us go into some detail about the structure of the new<br />

"science of heat" as it took shape in the early nineteenth century.<br />

Like mechanics, the science of heat implied both an original<br />

conception of the physical object and a definition of<br />

machines or engines-that is, an identification of cause and<br />

effect in a specific mode of production of mechanical work.<br />

The study of the physical processes involving heat entails<br />

defining a system, not, as in the case of dynamics, by the position<br />

and velocity of its constituents (there are some IQ23 mole-


ORDER OUT OF CHAOS<br />

106<br />

cutes in a volume of gas or a solid fragment of the order of a<br />

cm3), but by a set of macroscopic parameters such as temperature,<br />

pressure, volume, and so on. In addition, we have to<br />

take into account the boundary conditions that describe the<br />

relation of the system to its environment.<br />

Let us consider specific heat, one of the characteristic properties<br />

of a macroscopic system, as an example. The specific<br />

heat is a measure of the amount of heat required to raise the<br />

temperature of a system by one degree while its volume or<br />

pressure is kept constant. To study the specific heat-for instance,<br />

at constant volume-the system must be brought into<br />

interaction with its environment; it mu s t receive a certain<br />

amount of heat while at the same time its volume is kept constant<br />

and its pressure is allowed to vary.<br />

More generally, a system may be subjected to mechanical<br />

action (for example, either the pressure or the volume may be<br />

fixed by using a piston device), thermal action (a certain<br />

amount of heat may be given to or removed from the system,<br />

or the system itself may be brought to a given temperature<br />

through heat exchange), or chemical action (a flux of reactants<br />

and reaction products between the system and the environment).<br />

As we have already mentioned, pressure, volume,<br />

chemical composition, and temperature are the classical physicochemical<br />

parameters in terms of which the properties of<br />

macroscopic systems are defined. Thermodynamics is the science<br />

of the correlation among the variations in these properties.<br />

In comparison with dynamic objects, thermodynamic<br />

objects therefore lead to a new point of view. The aim of the<br />

theory is not to predict the changes in the system in terms of<br />

the interactions among particles; it aims instead to predict<br />

how the system will react to modifications we may impose on<br />

it from the outside.<br />

A mechanical engine gives back in the form of work the<br />

potential energy it has received from the outside world. Both<br />

cause and effect are of the same nature and, at least ideally,<br />

equivalent. In contrast, the heat engine implies material<br />

changes of states, including the transformation of the system's<br />

mechanical properties, dilatation, and expansion. The mechanical<br />

work produced must be seen as the result of a true<br />

process of transformation and not only as a transmission of<br />

movement. Thus the heat engine is not merely a passive de-


107 ENERGY AND THE INDUSTRIAL AGE<br />

vice; strictly speaking, it produces motion. This is the origin of<br />

a new problem: in order to restore the system's capacity to<br />

produce motion, the system must be brought back to its initial<br />

state. Thus a second process is needed, a second change of<br />

state that compensates for the change producing the motion.<br />

In a heat engine, this second process, which is opposite to the<br />

first, involves cooling the system until it regains its initial temperature,<br />

pressure, and volume.<br />

The problem of the efficiency of heat engines, of the ratio<br />

between the work done and the heat that must be supplied to<br />

the system to produce the two mutually compensating processes,<br />

is the very point at which the concept of irreversible<br />

process was introduced into physics. We shall return to the<br />

importance of Fourier's law in this context. Let us first describe<br />

the essential role played by the principle of energy conservation.<br />

The Principle of the Conservation of Energy<br />

We have already emphasized the central place of energy in<br />

classical dynamics. The Hamiltonian (the sum of the kinetic<br />

and potential energies) is expressed in terms of canonical variables-coordinates<br />

and momenta-and leads to changes in<br />

these variables while itself remaining constant throughout the<br />

motion. Dynamic change merely modifies the respective importance<br />

of potential and kinetic energy, conserving their totality.<br />

The early nineteenth century was characterized by unprecedented<br />

experimental ferment. 5 Physicists realized that motion<br />

does more than bring about changes in the relative position of<br />

bodies in space. New processes identified in the laboratories<br />

gradually formed a network that ultimately linked all the new<br />

fields of physics with other, more traditional branches, such as<br />

mechanics. One of these connections was accidentally discovered<br />

by Galvani. Before him, only static electric charges<br />

were known. Galvani, using a frog's body, set up the first experimental<br />

electric current. Volta soon recognized that the<br />

··galvanic" contractions in the frog were actually the effect of<br />

an electric current passing through it. In 1800, Volta con-


ORDER OUT OF CHAOS<br />

108<br />

structed a chemical battery; electricity could thus be produced<br />

by chemical reactions. Then came electrolysis: electric<br />

current can modify chemical affinities and produce chemical<br />

reactions. But this current can also produce light and heat;<br />

and, in 1820, Oersted discovered the magnetic effects produced<br />

by electrical currents. In 1822, Seebeck showed that,<br />

inversely, heat could produce electricity and, in 1834, how<br />

matter could be cooled by electricity. Then, in 183 1, Faraday<br />

induced an electric current by means of magnetic effects. A<br />

whole network of new effects was gradually uncovered. The<br />

scientific horizon was expanding at an unprecedented rate.<br />

In 1847 a decisive step was taken by Joule: the links among<br />

chemistry, the science of heat, electricity, magnetism, and biology<br />

were recognized as a "conversion." The idea of conversion,<br />

which postulates that "something" is quantitatively<br />

conserved while it is qualitatively transformed, generalizes<br />

what occurs during mechanical motion. As we have seen, total<br />

energy is conserved while potential energy is converted into<br />

kinetic energy, or vice versa. Joule defined a general equivalent<br />

for physicochemical transformations, thus making it<br />

possible to measure the quantity conserved. This quantity was<br />

later6 to become known as "energy." He established the first<br />

equivalence by measuring the mechanical work required to<br />

raise the temperature of a given quantity of water by one degree.<br />

A unifying element had been discovered in the middle of<br />

a bewildering variety of new discoveries. The conservation of<br />

energy, throughout the various transformations undergone by<br />

physical, chemical, and biological systems, was to provide a<br />

guiding principle in the exploration of the new processes.<br />

No wonder that the principle of the conservation of energy<br />

was so important to nineteenth-century physicists. For many<br />

of them it meant the unification of the whole of nature. Joule<br />

expressed this conviction in an English context:<br />

Indeed the phenomena of nature, whether mechanical,<br />

chemical, or vital, consist almost entirely in a continual<br />

conversion of attraction through space, living force<br />

(N.B., kinetic energy) and heat into one another. Thus it<br />

is that order is maintained in the universe-nothing is deranged,<br />

nothing ever lost, but the entire machinery, com-


109<br />

ENERGY AND THE INDUSTRIAL AGE<br />

plicated as it is, works smoothly and harmoniously. And<br />

though, as in the awful vision of Ezekiel, "wheel may be<br />

in the middle of wheel," and everything may appear complicated<br />

and involved in the apparent confusion and intricacy<br />

of an almost ·endless variety of causes, effects,<br />

conversions, and arrangements, yet is the most perfect<br />

regularity preserved-the whole being governed by the<br />

sovereign will of God. 7<br />

The case of the Germans Helmholtz, Mayer, and Lie big-all<br />

three belonging to a culture that would have rejected Joule's<br />

conviction on the grounds of strictly positivist practice-is<br />

even more striking. At the time of their discoveries, none of<br />

the three was, strictly speaking, a physicist. On the other<br />

hand, all of them were interested in the physiology of respiration.<br />

This had become, since Lavoisier, a model problem in<br />

which the functioning of a living being could be described in<br />

precise physical and chemical terms, such as the combustion<br />

of oxygen, the release of heat, and muscular work. It was thus<br />

a question that would attract physiologists and chemists hostile<br />

to Romantic speculation and eager to contribute to experimental<br />

science. However, judging from the account of how<br />

these three scientists came to the conclusion that respiration,<br />

and then the whole of nature, was governed by some fundamental<br />

"equivalence," we may state that the German philosophic<br />

tradition had imbued them with a conception that was<br />

quite alien to a positivist position: without hesitation they all<br />

concluded that the whole of nature, in each of its details, is<br />

ruled by this single principle of conservation.<br />

The case of Mayer is the most remarkable.s As a young doctor<br />

working in the Dutch colonie s in Java, he noticed the bright<br />

red color of the venous blood of one of his patients. This led<br />

him to conclude that, in a warm, tropical climate, the inhabitants<br />

need to burn less oxygen to maintain body temperature;<br />

this results in the bright color of their blood. Mayer went on to<br />

establish the balance between oxygen consumption, which is<br />

the source of energy, and the energy consumption involved in<br />

maintaining body temperature despite heat losses and manual<br />

work. This was quite a leap, since the color of the blood could<br />

as well be due to the patient's ··taziness. ,, But Mayer went


ORDER OUT OF CHAOS 110<br />

further and concluded that the balance between oxygen consumption<br />

and heat loss was merely the particular manifestation<br />

of the existence of an indestructible "force" underlying all<br />

phenomena.<br />

This tendency to see natural phenomena as the products of<br />

an underlying reality that remains constant throughout its<br />

transformations is strikingly reminiscent of Kant. Kant's influence<br />

can also be recognized in another idea held by some<br />

physiologists, the distinction between vitalism as philosophical<br />

speculation and the problem of scientific methodology. For<br />

those physiologists, even if there was a "vital" force underlying<br />

the function of living beings, the object of physiology<br />

would nonetheless be purely physicochemical in nature. From<br />

the two points of view mentioned, Kantianism, which ratified<br />

the systematic form taken by mathematical physics during the<br />

eighteenth century, can also be identified as one of the roots of<br />

the renewal of physics in the nineteenth century.9<br />

Helmholtz quite openly acknowledged Kant's influence.<br />

For Helmholtz, the principle of the conservation of energy was<br />

merely the embodiment in physics of the general a priori requirement<br />

on which all science is based-the postulate that<br />

there is a basic invariance underlying natural transformations:<br />

The problem of the sciences is, in the first place, to seek<br />

the laws by which the particular processes of nature may<br />

be referred to, and deduced from, general rules.<br />

We are justified, and indeed impelled in this proceeding,<br />

by the conviction that every change in nature must<br />

have a sufficient cause. The proximate causes to which<br />

we refer phenomena may, in themselves, be either variable<br />

or invariable; in the former case the above conviction<br />

impels us to seek for causes to account for the<br />

change, and thus we proceed until we at length arrive at<br />

final causes which are unchangeable, and which therefore<br />

must, in all cases where the exterior conditions are<br />

the same, produce the same invariable effects. The final<br />

aim of the theoretic natural sciences is therefore to discover<br />

the ultimate and unchangeable causes of natural<br />

phenomena. tO<br />

With the principle of the conservation of energy, the idea of


111 ENERGY AND THE INDUSTRIAL AGE<br />

a new golden age of physics began to take shape, an age that<br />

would lead to the ultimate generalization of mechanics.<br />

The cultural implications were far-reaching, and they included<br />

a conception of society and men as energy-transforming<br />

engines. But energy conversion cannot be the whole story. It<br />

represents the aspects of nature that are peaceful and controllable,<br />

but below there must be another, more "active" level.<br />

Nietzsche was one of those who detected the echo of creations<br />

and destructions that go far beyond mere conservation or conversion.<br />

Indeed, only difference, such as a difference of temperature<br />

or of potential energy, can produce results that are<br />

also differences.11 Energy conversion is merely the destruction<br />

of a difference, together with the creation of another difference.<br />

The power of nature is thus concealed by the use of<br />

equivalences. However, there is another aspect of nature that<br />

involves the boilers of steam engines, chemical transformations,<br />

life and death, and that goes beyond equivalences and<br />

conservation of energy.12 Here we reach the most original contribution<br />

of thermodynamics, the concept of irreversibility.<br />

Heat Engines and the Arrow of Time<br />

When we compare mechanical devices to thermal engines, for<br />

example, to the red-hot boilers of locomotives, we can see at a<br />

glance the gap between the classical age and nineteenthcentury<br />

technology. Still, physicists first thought that this gap<br />

could be ignored, that thermal engines could be described like<br />

mechanical ones, neglecting the crucial fact that fuel used by<br />

the steam engine disappears forever. But such complacency<br />

soon became impossible. For classical mechanics the symbol<br />

of nature was the clock; for the Industrial Age, it became a<br />

reservoir of energy that is always threatened with exhaustion.<br />

The world is burning like a furnace; energy, although being<br />

conserved, also is being dissipated.<br />

The original formulation of the second law of thermodynamics,<br />

which would lead to the first quantitative expression<br />

of irreversibility, was made by Sadi Carnot in 1824, before<br />

the general formulation of the principle of ;onservation of energy<br />

by Mayer (1842) and Helmholtz (1847). Carnot analyzed


ORDER OUT OF CHAOS 112<br />

the heat engine, closely following the work of his father, Lazare<br />

Carnot, who had produced an influential description of me ..<br />

chanical engines.<br />

The description of mechanical engines assumes motion as a<br />

given. In modern language this corresponds to conservation of<br />

energy and momentum. Motion is merely converted and<br />

transferred to other bodies. But the analogy between mechani ..<br />

cal and thermal engines was a natural one for Sadi Carnot,<br />

since he assumed, with most of the scientists of his time, that<br />

heat as well as mechanical energy are conserved.<br />

Water falling from one level to another can drive a mill. Sim ..<br />

ilarly, Sadi Carnot assumed two sources, one of which gives<br />

heat to the engine system, and the other, at a different tern ..<br />

perature, which absorbs the heat given by the former. It is the<br />

motion of the heat through the engine, between the two<br />

sources at different temperatures-that is, the driving force of<br />

fire-that will make the engine work.<br />

Carnot repeated his father's questions. l3 Which machine<br />

will have the highest efficiency? What are the sources of loss?<br />

What are the processes whereby heat propagates without producing<br />

work? Lazare Carnot had concluded that in order to<br />

obtain maximum efficiency from a mechanical machine it<br />

must be built and made to function to reduce to a minimum<br />

shocks, friction, or discontinuous changes of speed-in short,<br />

all that is caused by the sudden contact of bodies moving at<br />

different speeds. In doing so he had merely applied the physics<br />

of his time: only continuous phenomena are conservative;<br />

all abrupt changes in motion cause an irreversible loss of the<br />

"living force. " Similarly, the ideal heat engine, instead of having<br />

to avoid all contacts between bodies moving at different<br />

speeds, will have to avoid all contact between bodies having<br />

different temperatures.<br />

The cycle therefore has to be designed so that no temperature<br />

change results from direct heat flow between two bodies<br />

at different temperatures. Since such flows have no mechanical<br />

effect, they would merely lead to a loss of efficiency.<br />

The ideal Carnot cycle is thus a rather tricky device that<br />

achieves the paradoxical result of a heat transfer between two<br />

sources at different temperatures without any contact between<br />

bodies of different temperatures. It is divided into four phases.<br />

During each of the two isothermal phases, the system is in


..<br />

...<br />

113 ENERGY AND THE INDUSTRIAL AGE<br />

contact with one of the two heat sources and is kept at the<br />

temperature of this source. When in contact with the hot<br />

source, it absorbs heat and expands; when in contact with the<br />

cold source, it loses heat and contracts. The two isothermal<br />

phases are linked up by two phases in which the system is<br />

isolated from the sources-that is, heat no longer enters or<br />

leaves the system, but the temperature of the latter changes as<br />

a result, respectively, of expansion and compression. The volume<br />

continues to change until the system has passed from the<br />

temperature of one source to that of the other.<br />

p<br />

'<br />

'<br />

'<br />

...<br />

..<br />

..<br />

..<br />

...<br />

',Q<br />

...<br />

..<br />

...<br />

..<br />

... .........<br />

.. _<br />

_ _<br />

T<br />

H<br />

---- T<br />

c L<br />

v<br />

Figure 2. Pressure-volume diagram of the Carnot cycle: a thermodynamic<br />

engine, functioning between two sources, one "hot" at temperature TH, the<br />

other "cold" at temperature TL. Between state a and state b, there is an<br />

isothermal change: The system, kept at temperature TH, absorbs heat and<br />

expands. Between b and c, the system is kept expanding while in thermal<br />

isolation; its temperature goes down from TH to TL. Those two steps produce<br />

mechanical energy. Between c and d, there is a second isothermal change:<br />

the system is compressed and releases heat while being kept at temperature<br />

TL. Between d and a, the system, again isolated, is compressed while its<br />

temperature increases to temperature TH.


ORDER OUT OF CHAOS 114<br />

It is quite remarkable that this description of an ideal thermal<br />

engine does not mention the irreversible processes that<br />

are at the basis of its realization. No mention is made of the<br />

furnace in which the coal is burning. The model is only concerned<br />

with the effect of the combustion, which permits the<br />

maintenance of the temperature difference between the two<br />

sources.<br />

In 1850, Clausius described the Carnot cycle from the new<br />

perspective provided by the conservation of energy. He discovered<br />

that the need for two sources and the formula for theoretical<br />

efficiency stated by Carnot express a specific problem<br />

with heat engines: the need for a process compensating for<br />

conversion (in the present instance, cooling by contact with a<br />

cold source) to restore the engine to its initial mechanical and<br />

thermal conditions. Balance relations expressing energy conversion<br />

are now joined by new equivalence relations between<br />

the effects of two processes on the state of the system, heat<br />

flux between the sources and conversion of heat into work. A<br />

new science, thermodynamics, which linked mechanical and<br />

thermal effects, came into being.<br />

The work of Clausius explicitly demonstrated that we cannot<br />

use without restriction the seemingly inexhaustible energy<br />

reservoir that nature provides. Not all energy-conserving processes<br />

are possible. An energy difference, for instance, cannot<br />

be created without the destruction of an at least equivalent<br />

energy difference. Thus in the ideal Carnot cycle, the price for<br />

the work produced is paid by the heat, which is transferred<br />

from one source to the other. The outcome, as expressed by<br />

the mechanical work produced on one side, and the transfer of<br />

heat on the other, is linked by an equivalence. This equivalence<br />

is valid in both directions. By working in reverse, the<br />

same machine can restore the initial temperature difference<br />

while consuming the work produced. No heat engine can be<br />

constructed using a single source of heat.<br />

Clausius was no more concerned than Carnot with the<br />

losses whereby all real engines have an efficiency lower than<br />

the ideal value predicted by the theory. His description, like<br />

that of Carnot, corresponds to an idealization. It leads to the'<br />

definition of the limit nature imposes on the yield of thermal<br />

engines.


115 ENERGY AND THE INDUSTRIAL AGE<br />

However, since the eighteenth century, the status of idealizations<br />

had changed. Based as it was on the principle of the conservation<br />

of energy, the new science claimed to describe not<br />

only idealizations, but also nature itself, including "losses."<br />

This raised a new problem, whereby irreversibility entered<br />

physics. How does one describe what happens in a real engine?<br />

How does one include losses in the energy balance?<br />

How do they reduce efficiency? These questions paved the<br />

way to the second law of thermodynamics.<br />

From Technology to Cosmology<br />

As we have seen, the question raised by Carnot and Clausius<br />

led to a description of ideal engines that was based on conservation<br />

and compensation. In addition, it provided an opportunity<br />

for presenting new problems, such as the dissipation of<br />

energy. William Thomson, who had great respect for Fourier's<br />

work, was quick to grasp the importance of the problem, and<br />

in 1852 he was the first to formulate the second law of thermodynamics.<br />

It was Fourier's heat propagation that Carnot had identified<br />

as a possible cause for the power losses in a heat engine. Carnot's<br />

cycle, no longer the ideal cycle but the "real" cycle, thus<br />

became the point of convergence of the two universalities discovered<br />

in the nineteenth century-energy conversion and<br />

heat propagation. The combination of these two discoveries<br />

led Thomson to formulate his new principle: the existence in<br />

nature of a universal tendency toward the degradation of mechanical<br />

energy. Note the word "universal," which has obvious<br />

cosmological connotations.<br />

The world of Laplace was eternal, an ideal perpetual-motion<br />

machine. Since Thomson's cosmology is not merely a reflection<br />

of the new ideal heat engine but also incorporates the consequences<br />

of the irreversible propagation of heat in a world in<br />

which energy is conserved. This world is described as an engine<br />

in which heat is converted into motion only at the price of some<br />

irreversible waste and useless dissipation. Effect-producing<br />

differences in nature progressively diminish. The world uses


ORDER OUT OF CHAOS 116<br />

up its differences as it goes from one conversion to another<br />

and tends toward a final state of thermal equilibrium, "heat<br />

death." In accordance with Fourier's law, in the end there will<br />

no longer be any differences of temperature to produce a mechanical<br />

effect.<br />

Thomson thus made a dizzy leap from engine technology to<br />

cosmology. Hs formulation of the second law was couched in<br />

the scientific terminology of his time: the conservation of energy,<br />

engines, and Fourier's law. It is clear, moreover, that the<br />

part played by the cultural context was important. It is generally<br />

accepted that the problem of time took on a new importance<br />

during the nineteenth century. Indeed, the essential role<br />

of time began to be noticed in all fields-in geology, in biology,<br />

in language, as well as in the study of human social evolution<br />

and ethics. But it is interesting that the specific form in which<br />

time was introduced in physics, as a tendency toward homogeneity<br />

and death, reminds us more of ancient mythological and<br />

religious archetypes than of the progressive complexification<br />

and diversification described by biology and the social sciences.<br />

The return of these ancient themes can be seen as a<br />

cultural repercussion of the social and economic upheavals of<br />

the time. The rapid transformation of the technological mode<br />

of interaction with nature, the constantly accelerating pace of<br />

change experienced by the nineteenth century, produced a<br />

deep anxiety. This anxiety is still with us and takes various<br />

forms, from the repeated proposals for a "zero growth" society<br />

or for a moratorium on scientific research to the<br />

announcement of "scientific truths" concerning our disintegrating<br />

universe. Present knowledge in astrophysics is still<br />

scanty and very problematic, since in this field gravitational<br />

effects play an essential role and problems imply the simultaneous<br />

use of thermodynamics and relativity. Yet most texts<br />

in this field are unanimous in predicting final doom. The conclusion<br />

of a recent book reads:<br />

The unpalatable truth appears to be that the inexorable<br />

disintegration of the universe as we know it seems assured,<br />

the <strong>org</strong>anization which sustains all ordered activity,<br />

frem men to galaxies, is slowly but inevitably<br />

running down, and may even be overtaken by total gravitational<br />

collapse into oblivion.t4


117 ENERGY AND THE INDUSTRIAL AGE<br />

Others are more optimistic. In an excellent article on the<br />

energy of the universe, Freeman Dyson has written:<br />

It is conceivable however that life may have a larger role<br />

to play than we have yet imagined. Life may succeed<br />

against all of the odds in molding the universe to its own<br />

purpose. And the design of the inanimate universe may<br />

not be as detached from the potentialities of life and intelligence<br />

as scientists of the twentieth century have tended<br />

to suppose.15<br />

In spite of the important progress made by Hawking and others,<br />

our knowledge of large-scale transformations in our universe<br />

remains inadequate.<br />

The Birth of Entropy<br />

In 1865, it was Clausius' turn to make the leap from technology<br />

to cosmology. At the outset he merely reformulated his<br />

earlier conclusions, but in doing so he introduced a new concept,<br />

entropy. His first goal was to distinguish clearly between<br />

the concepts of conservation and of reversibility. Unlike mechanical<br />

transformations, where reversibility and conservation<br />

coincide, a physicochemical transformation may conserve energy<br />

even though it cannot be reversed. This is true, for instance,<br />

in the case of friction, in which motion is converted<br />

into heat, or in the case of heat conduction as it was described<br />

by Fourier.<br />

We are already familiar with energy, which is a function of<br />

the state of a system-that is, a function dependent only on<br />

the value of the parameters (pressure, volume, temperature)<br />

by which that state may be defined.t6 But we must go beyond<br />

the principle of energy conservation and find a way to express<br />

the distinction between "useful" exchanges of energy in the<br />

Carnot cycle and "dissipated" energy that is irreversibly<br />

wasted.<br />

This is precisely the role of Clausius' new function, entropy,<br />

generally denoted by S.<br />

Apparently Clausius merely wished to express in a new form


ORDER OUT OF CHAOS 118<br />

the obvious requirement that an engine return to its initial<br />

state at the end of its cycle. The first definition of entropy is<br />

centered on conservation: at the end of each cycle, whether<br />

ideal or not, the function of the system's state, entropy, returns<br />

to its initial value. But the parallel between entropy and energy<br />

ends as soon as we abandon idealizations.t7<br />

Let us consider the variation of the entropy dS over a short<br />

time interval dt. The situation is quite different for ideal and<br />

real engines. In the first case, dS may be expressed completely<br />

in terms of the exchanges between the engine and its environment.<br />

We can set up experiments in which heat is given up by<br />

the system instead of flowing into the system. The corresponding<br />

change in entropy would simply have its sign<br />

changed. This kind of contribution to entropy, which we shall<br />

call deS, is therefore reversible in the sense that it can have<br />

either a positive or a negative sign. The situation is drastically<br />

different in a real engine. Here, in addition to reversible exchanges,<br />

we have irreversible processes inside the system,<br />

such as heat losses, friction, and so on. These produce an entropy<br />

increase or "entropy production" inside the system.<br />

This increase of entropy, which we shall call diS, cannot<br />

change its sign through a reversal of the heat exchange with<br />

the outside world. Like all irreversible processes (such as heat<br />

conduction), entropy production always proceeds in the same<br />

direction. In other words, diS can only be positive or vanish in<br />

the absence of irreversible processes. Note that the positive<br />

sign of diS is chosen merely by convention; it could just as<br />

well have been negative. The point is that the variation is monotonous,<br />

that entropy production cannot change its sign as<br />

time goes on.<br />

The notations deS and diS have been chosen to remind the<br />

reader that the first term refers to exchanges (e) with the outside<br />

world, while the second refers to the irreversible processes<br />

inside (i) the system. The entropy variation dS is<br />

therefore the sum of the two terms deS and diS, which have<br />

quite different physical meanings.18<br />

To grasp the peculiar feature of this decomposition of entropy<br />

variation into two parts, it is useful to apply our formulation<br />

to energy. Let us denote energy by E and variation over a<br />

short time dt by dE. Of course, we would still write that dE is<br />

equal to the sum of a term deE due to the exchanges of energy


119 ENERGY AND THE INDUSTRIAL AGE<br />

and a term diE linked to the "internal production" of energy.<br />

However, the principle of the conservation of energy states<br />

that energy is never "produced" but only transferred from one<br />

place to another. The variation in energy dE is then reduced to<br />

deE. On the other hand, if we take a nonconserved quantity,<br />

such as the quantity of hydrogen molecules contained in a vessel,<br />

this quantity may vary both as the result of adding hydrogen<br />

to the vessel or through chemical reactions occurring<br />

inside the vessel. But in this case, the sign of the "production"<br />

is not determined. Depending on the circumstances, we can<br />

produce or destroy hydrogen molecules by transferring hydrogen<br />

atoms to other chemical components. The peculiar feature<br />

of the second law is the fact that the production term diS is<br />

always positive. The production of entropy expresses the occurrence<br />

of irreversible changes inside the system.<br />

Clausius was able to express quantitatively the entropy flow<br />

deS in terms of the heat received (or given up) by the system.<br />

In a world dominated by the concepts of reversibility and conservation,<br />

this was his main concern. Regarding the irreversible<br />

processes involved in entropy production, he merely stated<br />

the existence of the inequality diS/dt>O. Even so, important<br />

progress had been made, for, if we leave the Carnot cycle and<br />

consider other thermodynamic systems, the distinction between<br />

entropy flow and entropy production can still be made.<br />

For an isolated system .that has no exchanges with its environment,<br />

the entropy flow is, by definition, zero. Only the production<br />

term remains, and the system's entropy can only<br />

increase or remain constant. Here, then, it is no longer a question<br />

of irreversible transformations considered as approximations<br />

of reversible transformations; increasing entropy corresponds<br />

to the spontaneous evolution of the system. Entropy thus becomes<br />

an "indicator of evolution," or an "arrow of time," as<br />

Eddington aptly called it. For all isolated systems, the future<br />

is the direction of increasing entropy.<br />

What system would be better "isolated" than the universe<br />

as a whole? This concept is the basis of the cosmological formulation<br />

of the two laws of thermodynamics given by Clausius<br />

in 1865:<br />

Die Energie der Welt ist konstant.<br />

Die Entropie der Welt strebt einem Maximum zu.19<br />

The statement that the entropy of an isolated system in-


ORDER OUT OF CHAOS<br />

120<br />

creases to a maximum goes far beyond the technological problem<br />

that gave rise to thermodynamics. Increasing entropy is<br />

no longer synonymous with loss but now refers to the natural<br />

processes within the system. These are the processes that ultimately<br />

lead the system to thermodynamic "equilibrium"<br />

corresponding to the state of maximum entropy.<br />

In Chapter I we emphasized the element of surprise involved<br />

in the discovery of Newton's universal laws of dynamics.<br />

Here also the element of surprise is apparent. When Sadi<br />

Carnot formulated the laws of ideal thermal engines, he was<br />

far from imagining that his work would lead to a conceptual<br />

revolution in physics.<br />

Reversible transformations belong to classical science in the<br />

sense that they define the possibility of acting on a system, of<br />

controlling it. The dynamic object could be controlled through<br />

its initial conditions. Similarly, when defined in terms of its<br />

reversible transformations, the thermodynamic object may be<br />

controlled through its boundary conditions: any system in<br />

thermodynamic equilibrium whose temperature, volume, or<br />

pressure are gradually changed passes through a series of<br />

equilibrium states, and any reversal of the manipulation leads<br />

to a return to its initial state. The reversible nature of such<br />

change and controlling the object through its boundary conditions<br />

are interdependent processes. In this context irreversibility<br />

is "negative"; it appears in the form of "uncontrolled"<br />

changes that occur as soon as the system eludes control. But<br />

inversely, irreversible processes may be considered as the last<br />

remnants of the spontaneous and intrinsic activity displayed<br />

by nature when experimental devices are employed to harness<br />

it.<br />

Thus the "negative" property of dissipation shows that, unlike<br />

dynamic objects, thermodynamic objects can only be partially<br />

controlled. Occasionally they "break loose" into<br />

spontaneous change.<br />

All changes are not equivalent for a thermodynamic system.<br />

This is the meaning of the expression dS =deS+ diS.<br />

Spontaneous change toward equilibrium diS is different from<br />

the change deS, which is determined and controlled by a modification<br />

of the boundary conditions (for example, ambient<br />

temperature). For an isolated system, equilibrium appears as


121 ENERGY AND THE INDUSTRIAL AGE<br />

an ••attractor" of nonequilibrium states. Our initial assertion<br />

may thus be generalized by saying that evolution toward an<br />

attractor state differs from all other changes, especially from<br />

changes determined by boundary conditions.<br />

Max Planck often emphasized the difference between the<br />

two types of change found in nature. Nature, wrote Planck,<br />

seems to "favor" certain states. The irreversible increase in<br />

entropy diS/dt describes a system's approach to a state which<br />

"attracts" it, which the system prefers and from which it will<br />

not move of its own "free will." "From this point of view, Nature<br />

does not permit processes whose final states she finds<br />

less attractive than their initial states. Reversible processes are<br />

limiting cases. In them, Nature has an equal propensity for<br />

initial and final states; this is why the passage between them<br />

can be made in both directions. "20<br />

How foreign such language sounds when compared with the<br />

language of dynamics! In dynamics, a system changes according<br />

to a trajectory that is given once and for all, whose starting<br />

point is never f<strong>org</strong>otten (since initial conditions determine the<br />

trajectory for all time). However, in an isolated system all nonequilibrium<br />

situations produce evolution toward the same kind<br />

of equilibrium state. By the time equilibrium has been<br />

reached, the system has f<strong>org</strong>otten its initial conditions-that<br />

is, the way it had been prepared.<br />

Thus specific heat or the compressibility of a system in<br />

equilibrium are properties independent of the way the system<br />

has been set up. This fortunate circumstance greatly simplifies<br />

the study of the physical states of matter. Indeed, complex<br />

systems consist of an immense number of particles.* From the<br />

dynamic standpoint it is practically impossible to reproduce<br />

any state of such systems in view of the infinite variety of dynamic<br />

states that may occur.<br />

We are now confronted with two baically different descriptions:<br />

dynamics, which applies to the world of motion, and<br />

*Physical chemistry often employs Avogadro's number-that is, the number<br />

of molecules in a "mole" of matter (a mole always contains the same<br />

number of particles, the number of atoms contained in one gram of hydro<br />

gen). This number is of the order of 6.1023, and it is the characteristic order<br />

of magnitude of the number of particles forming systems governed by the<br />

laws of classical thermodynamics.


ORDER OUT OF CHAOS 122<br />

thermodynamics, the science of complex systems with its intrinsic<br />

direction of evolution toward increasing entropy. This<br />

dichotomy immediately raises the question of how these descriptions<br />

are related, a problem that has been debated since<br />

the laws of thermodynamics were formulated.<br />

Boltzmanns Order Principle<br />

The second law of thermodynamics contains two fundamental<br />

elements: ( 1) a "negative " one that expresses the impossibility<br />

of certain processes (heat flows from the hot source to the cold<br />

and not vice versa) and (2) a "positive," constructive one. The<br />

second is a consequence of the first; it is the impossibility of<br />

certain processes that permits us to introduce a function, entropy,<br />

which increases uniformly for isolated systems. Entropy<br />

behaves as an attractor for isolated systems.<br />

How could the formulations of thermodynamics be reconciled<br />

with dynamics? At the end of the nineteenth century,<br />

most scientists seemed to think this was impossible. The principles<br />

of thermodynamics were new laws forming the basis of a<br />

new science that could not be reduced to traditional physics.<br />

Both the qualitative diversity of energy and its tendency toward<br />

dissipation had to be accepted as new axioms. This was<br />

the argument of the "energeticists" as opposed to the "atomists,"<br />

who refused to abandon what they considered to be the<br />

essential mission of physics-to reduce the complexity of natural<br />

phenomena to the simplicity of elementary behavior expressed<br />

by the laws of motion.<br />

Actually, the problems of the transition from the microscopic<br />

to the macroscopic level were to prove exceptionally<br />

fruitful for the development of physics as a whole. Boltzmann<br />

was the first to take up the challenge. He felt that new concepts<br />

had to be developed to extend the physics of trajectories<br />

to cover the situation described by thermodynamics. Following<br />

in Maxwell's footsteps, Boltzmann sought this conceptual<br />

innovation in the theory of probability.<br />

That probability could play a role in the description of complex<br />

phenomena was not surprising: Maxwell himself appears


123 ENERGY AND THE INDUSTRIAL AGE<br />

to have been influenced by the work of Quetelet, the inventor<br />

of the "average" man in sociology. The innovation was to introduce<br />

probability in physics not as a means of approximation<br />

but rather as an explanatory principle, to use it to show<br />

that a system could display a new type of behavior by virtue of<br />

its being composed of a large population to which the laws of<br />

probability could be applied.<br />

Let us consider a simple example of the application of the<br />

concept of probability in physics. An ensemble composed of<br />

N particles is contained in a box divided into two equal compartments.<br />

The problem is to find the probability of the<br />

various possible distributions of particles between the compartments-that<br />

is, the probability of finding N1 particles in<br />

the first compartment (and N2 = N-N1 in the second).<br />

Using combinatorial analysis, it is easy to calculate the<br />

number of ways in which each different distribution of N particles<br />

can be achieved. Thus if N = 8, there is only one way of<br />

placing the eight particles in a single half. There are, however,<br />

eight different ways of putting one particle in one half and<br />

seven in the other half, if we suppose the particles to be distinguishable,<br />

as is assumed in classical physics. Furthermore,<br />

equal distribution of the eight particles between the two halves<br />

can be carried out in 8!14!4! = 70 different ways (where<br />

n! = 1·2·3 ... (n-l)·n). Likewise, whatever the value of N, a<br />

number P of situations called complexions in physics may be<br />

defined, giving the number of ways of achieving any given distribution<br />

N1,N2• Its expression is P=N!IN1!N2!.<br />

For any given population, the larger the number of complexions<br />

the smaller the difference between N1 and N2• It is maximum<br />

when the population is equally distributed over the two<br />

halves. Moreover, the larger the value of N, the greater the<br />

difference between the number of complexions corresponding<br />

to the different ways of distribution. For values of N of the<br />

order of 1Q 2 3 values found in macroscopic systems, the overwhelming<br />

majority of possible distributions corresponds to<br />

the distribution N1 =N2 =NI2. For systems composed of a<br />

large number of particles, all states that differ from the state<br />

corresponding to an equal distribution are thus highly improbable.<br />

Boltzmann was the .first to realize that irreversible increase


ORDER OUT OF CHAOS 124<br />

in entropy could be considered as the expression of a growing<br />

molecular disorder, of the gradual f<strong>org</strong>etting of any initial dissymmetry,<br />

since dissymmetry decreases the number of complexions<br />

when compared to the state corresponding to the<br />

maximum of P. Boltzmann thus aimed to identify entropy S<br />

with the number of complexions: entropy characterizes each<br />

macroscopic state in terms of the number of ways of achieving<br />

this state. Boltzmann's famous equation S = k lg pt expresses<br />

this idea in quantitative form. The proportionality factor k in<br />

this formula is a universal constant, known as Boltzmann's<br />

constant.<br />

Boltzmann's results signify that irreversible thermodynamic<br />

change is a change toward states of increasing probability and<br />

that the attractor state is a macroscopic state corresponding to<br />

maximum probability. This takes us far beyond Newton. For<br />

the first time a physical concept has been explained in terms of<br />

probability. Its utility is immediately apparent. Probability can<br />

adequately explain a system's f<strong>org</strong>etting of all initial dissymmetry,<br />

of all special distributions (for example, the set of particles<br />

concentrated in a subregion of the system, or the<br />

distribution of velocities that is created when two gases of different<br />

temperatures are mixed). This f<strong>org</strong>etting is possible because,<br />

whatever the evolution peculiar to the system, it will<br />

ultimately lead to one of the microscopic states corresponding<br />

to the macroscopic state of disorder and maximum symmetry,<br />

since these macroscopic states correspond to the overwhelming<br />

majority of possible microscopic states. Once this state<br />

has been reached, the system will move only short distances<br />

from the state, and for short periods of time. In other words,<br />

the system will merely fluctuate around the attractor state.<br />

Boltzmann's order principle implies that the most probable<br />

state available to a system is the one in which the multitude of<br />

events taking place simultaneously in the system compensates<br />

for one another statistically. In the case of our first example,<br />

whatever the initial distribution, the system's evolution will<br />

ultimately lead it to the equal distribution N1 = N 2 • This state<br />

will put an end to the system's irreversible macroscopic evolutThe<br />

logarithmic expression indicates that entropy is an additive quantity<br />

(S 1 +2 = S 1 + S2), while the number of complexions is multiplicative<br />

(PI +2 =PI· P2).


125 ENERGY AND THE INDUSTRIAL AGE<br />

tion. Of course, the particles will go on moving from one half<br />

to the other, but on the average, at any given instant, as many<br />

will be going in one direction as in the other. As a result, their<br />

motion will cause only small, short-lived fluctuations around<br />

the equilibrium state N1 =N2• Boltzmann's probabilistic interpretation<br />

thus makes it possible to understand the specificity<br />

of t h e attractor studied by equilibrium thermodynamics.<br />

This is not the whole story, and we shall devote the third<br />

part of this book to a more detailed discussion. A few remarks<br />

suffice here. In classical mechanics (and, as we shall see, in<br />

quantum mechanics as well), everything is determined in<br />

terms of initial states and the laws of motion. How then does<br />

probability enter the description of nature? Here it is common<br />

to invoke our ignorance of the exact dynamic state of the system.<br />

This is the subjectivistic interpretation of entropy. Such<br />

an interpretation was acceptable when irreversible processes<br />

were considered to be mere nuisances corresponding to friction<br />

or, more generally, to losses in the functioning of thermal<br />

engines. But today the situation has changed. As we shall see,<br />

irreversible processes have an immense constructive importance:<br />

life would not be possible without them. The subjectivistic<br />

interpretation is therefore highly questionable. Are we<br />

ourselves merely the result of our ignorance, of the fact that<br />

we only observe macroscopic states.<br />

Moreover, both in thermodynamics as well as in its probabilistic<br />

interpretation, there appears a dissymmetry in time:<br />

entropy increases in the direction of the future, not of the past.<br />

This seems impossible when we consider dynamic equations that<br />

are invariant in respect to time inversion. As we shall see, the<br />

second law is a selection principle compatible with dynamics<br />

but not deducible from it. It limits the possible initial conditions<br />

available to a dynamic system. The second law therefore<br />

marks a radical departure from the mechanistic world of classical<br />

or quantum dynamics. Let us now return to Boltzmann's<br />

work.<br />

So far we have discussed isolated systems in which the number<br />

of particles as well as the total energy are fixed by the<br />

boundary conditions. However, it is possible to extend<br />

Boltzmann's explanation to open systems that interact with<br />

their environment. In a closed system, defined by boundary<br />

conditions such that its temperature Tis kept constant by heat


ORDER OUT OF CHAOS<br />

126<br />

exchange with the environment, equilibrium is not defined in<br />

terms of maximum entropy but in terms of the minimum of a<br />

similar function, free energy: F=E-TS , where E is the energy<br />

of the system and Tis the temperature (measured on the<br />

so-called Kelvin scale, where the freezing point of water is<br />

273°C and its boiling point is 373°C).<br />

This formula signifies that equilibrium is the result of competition<br />

between energy and entropy. Temperature is what<br />

determines the relative weight of the two factors. At low<br />

temperatures, energy prevails, and we have the formation of<br />

ordered (weak-entropy) and low-energy structures such as<br />

crystals. Inside these structures each molecule interacts with<br />

its neighbors, and the kinetic energy involved is small compared<br />

with the potential energy that results from the interactions<br />

of each molecule with its neighbors. We can imagine<br />

each particle as imprisoned by its interactions with its neighbors.<br />

At high temperatures, however, entropy is dominant and<br />

so is molecular disorder. The importance of relative motion<br />

increases, and the regularity of the crystal is disrupted; as the<br />

temperature increases, we first have the liquid state, then the<br />

gaseous state.<br />

The entropy S of an isolated system and the free energy F of<br />

a system at fixed temperature are examples of "thermodynamic<br />

potentials." The extremes of thermodynamic potentials<br />

such as S or F define the attractor states toward which systems<br />

whose boundary conditions correspond to the definition<br />

of these potentials tend spontaneously.<br />

Boltzmann's principle can also be used to study the coexistence<br />

of structures (such as the liquid phase and the solid<br />

phase) or the equilibrium between a crystallized product and<br />

the same product in solution. It is important to remember,<br />

however, that equilibrium structures are defined on the molecular<br />

level. It is the interaction between molecules acting<br />

over a range of the order of some to-s em, the same order of<br />

magnitude as the diameter of atoms in molecules, that makes a<br />

crystal structure stable and endows it with its macroscopic<br />

properties. Crystal size, on the other hand, is not an intrinsic<br />

property of structure. It depends on the quantity of matter in<br />

the crystalline phase at equilibrium.


127 ENERGY AND THE INDUSTRIAL AGE<br />

Carnot and Darwin<br />

Equilibrium thermodynamics provides a satisfactory explanation<br />

for a vast number of physicochemical phenomena. Yet it<br />

may be asked whether the concept of equilibrium structures<br />

encompasses the different structures we encounter in nature.<br />

Obviously the answer is no.<br />

Equilibrium structures can be seen as the results of statistical<br />

compensation for the activity of microscopic elements<br />

(molecules, atoms). By definition they are inert at the global<br />

level. For this reason they are also "immortal." Once they<br />

have been formed, they may be isolated and maintained indefinitely<br />

without further interaction with their environment.<br />

When we examine a biological cell or a city, however, the situation<br />

is quite different: not only are these systems open, but<br />

also they exist only because they are open. They feed on the<br />

flux of matter and energy coming to them from the outside<br />

world. We can isolate a crystal, but cities and cells die when<br />

cut off from their environment. They form an integral part of<br />

the world from which they draw sustenance, and they cannot<br />

be separated from the fluxes that they incessantly transform.<br />

However, it is not only living nature that is profoundly alien<br />

to the models of thermodynamic equilibrium. Hydrodynamics<br />

and chemical reactions usually involve exchanges of matter<br />

and energy with the outside world.<br />

It is difficult to see how Boltzmann's order principle can be<br />

applied to such situations. The fact that a system becomes<br />

more uniform in the course of time can be understood in terms<br />

of complexions; in a state of uniformity, when the "differences"<br />

created by the initial conditions have been f<strong>org</strong>otten,<br />

the number of complexions will be maximum. But it is impossible<br />

to understand spontaneous convection from this point of<br />

view. The convection current calls for coherence, for the cooperation<br />

of a vast number of molecules. It is the opposite of<br />

disorder, a privileged state to which only a comparatively<br />

small number of complexions may correspond. In Boltzmann's<br />

terms, it is an "improbable" state. If convection must<br />

be considered a "miracle,·· what then is there to say about life,


ORDER OUT OF CHAOS 128<br />

with its highly specific features present in the simplest <strong>org</strong>anisms?<br />

The question of the relevance of equilibrium models can be reversed.<br />

In order to produce equilibrium, a system must be<br />

"protected" from the fluxes that compose nature. It must be<br />

"canned," so to speak, or put in a bottle, like the homunculus<br />

in Goethe's Fa ust, who addresses to the alchemist who created<br />

him: "Come, press me tenderly to your breast, but not<br />

too hard, for fear the glass might break. This is the way things<br />

are: something natural, the whole world hardly suffices what<br />

is, but what is artificial demands a closed space." In the world<br />

that we are familiar with, equilibrium is a rare and precarious<br />

state. Even evolution toward equilibrium implies a world like<br />

ours, far enough away from the sun for the partial isolation of a<br />

system to be conceivable (no "canning" is possible at the temperature<br />

of the sun), but a world in which nonequilibrium remains<br />

the rule, a "lukewarm" world where equilibrium and<br />

nonequilibrium coexist.<br />

For a long time, however, physicists thought they could define<br />

the inert structure of crystals as the only physical order<br />

that is predictable and reproducible and approach equilibrium<br />

as the only evolution that could be deduced from the fundamental<br />

laws of physics. Thus any attempt at extrapolation<br />

from thermodynamic descriptions was to define as rare and<br />

unpredictable the kind of evolution described by biology and<br />

the social sciences. How, for example, could Darwinian evolution-the<br />

statistical selection of rare events-be reconciled<br />

with the statistical disappearance of all peculiarities, of all rare<br />

configurations, described by Boltzmann? As Roger Caillois21<br />

asks: "Can Carnot and Darwin both be right?"<br />

It is interesting to note how similar in essence the Darwinian<br />

approach is to the path explored by Boltzmann. This may be<br />

more than a coincidence. We know that Boltzmann had immense<br />

admiration for Darwin. Darwin's theory begins with an<br />

assumption of the spontaneous fluctuations of species; then<br />

selection leads to irreversible biological evolution. Therefore,<br />

as with Boltzmann, a randomness leads to irreversibility. Yet<br />

the result is very different. Boltzmann's interpretation implies<br />

the f<strong>org</strong>etting of initial conditions, the "destruction" of initial<br />

structures, while Darwinian evolution is associated with self<strong>org</strong>anization,<br />

ever-increasing complexity.


129 ENERGY AND THE INDUSTRIAL AGE<br />

To sum up our argument so far, equilibrium thermodynamics<br />

was the first response of physics to the problem of nature's<br />

complexity. This response was expressed in terms of the dissipation<br />

of energy, the f<strong>org</strong>etting of initial conditions, and evolution<br />

toward disorder. Classical dynamics, the science of eternal,<br />

reversible trajectories, was alien to the problems facing the<br />

nineteenth century, which was dominated by the concept of<br />

evolution. Equilibrium thermodynamics was in a position to<br />

oppose its view of time to that of other sciences: for thermodynamics,<br />

time implies degradation and death. As we have seen,<br />

Diderot had already asked the question: Where do we, <strong>org</strong>anized<br />

beings endowed with sensations, fit in an inert world<br />

subject to dynamics? There is another question, which has<br />

plagued us for more than a century: What significance does<br />

the evolution of a living being have in the world described by<br />

thermodynamics, a world of ever-increasing disorder? What is<br />

the relationship between thermodynamic time, a time headed<br />

toward equilibrium, and the time in which evolution toward<br />

increasing complexity is occurring?<br />

Was Bergson right? Is time the very medium of innovation,<br />

or is it nothing at all?


CHAPTERV<br />

THE THREE STAGES<br />

OF THERMODYNAMICS<br />

Flux and Force<br />

Let us return I to the description of the second law given in the<br />

previous chapter. The concept of entropy plays a central role<br />

in the description of evolution. As we have seen, its variation<br />

can be written as the sum of two terms-the term deS, linked<br />

to the exchanges between the system and the rest of the world,<br />

and a production term, diS, resulting from irreversible phenomena<br />

inside the system. This term is always positive except<br />

at thermodynamic equilibrium, when it becomes zero. For isolated<br />

systems (deS= 0), the equilibrium state corresponds to a<br />

state of maximum entropy.<br />

In order to appreciate the significance of the second law for<br />

physics, we need a more detailed description of the various<br />

irreversible phenomena involved in the entropy production diS<br />

or in the entropy production per unit time P= diS/dt.<br />

For us chemical reactions are of particular significance. Together<br />

with heat conduction, they form the prototype of irreversible<br />

processes. In addition to their intrinsic importance,<br />

chemical processes play a fundamental role in biology. The<br />

living cell presents an incessant metabolic activity. There<br />

thousands of chemical reactions take place simultaneously to<br />

transform the matter the cell feeds on, to synthesize the fundamental<br />

biomolecules, and to eliminate waste products. As<br />

regards both the different reaction rates and the reaction sites<br />

within the cell, this chemical activity is highly coordinated.<br />

The biological structure thus combines order and activity. In<br />

contrast, an equilibrium state remains inert even though it may<br />

be structured, as, for example, with a crystal. Can chemical<br />

131


ORDER OUT OF CHAOS 132<br />

processes provide us with the key to the difference between<br />

the behavior of a cry stal and that of a cell?<br />

We will have to consider chemical reactions from a dual<br />

point of view, both kinetic and thermodynamic.<br />

From the kinetic point of view, the fundamental quantity is the<br />

reaction rate. The classical theory of chemical kinetics is<br />

based on the assumption that the rate of a chemical reaction is<br />

proportional to the concentrations of the products taking part<br />

in it. Indeed, it is through collisions between molecules that a<br />

reaction takes place, and it is quite natural to assume that the<br />

number of collisions is proportional to the product of the concentrations<br />

of the reacting molecules.<br />

For the sake of example, let us take a simple reaction such<br />

as A + X B + Y. This "reaction equation" means that whenever<br />

a molecule of component A encounters a molecule of X,<br />

there is a certain probability that a reaction will take place and<br />

a molecule of B and a molecule of Y will be produced. A collision<br />

producing such a change in the molecules involved is a<br />

"reactive collision." Only a usually very small fraction (for<br />

example, 111 (6) of all collisions are of this kind. In most cases,<br />

the molecules retain their original nature and merely exchange<br />

energy.<br />

Chemical kinetics deals with changes in the concentration<br />

of the different products involved.in a reaction. This kinetics is<br />

described by differential equations, just as motion is described<br />

by the Newtonian equations. However, in this case, we are not<br />

calculating acceleration but the rates of change of concentration,<br />

and these rates are expressed as a function of the<br />

concentrations of the reactants. The rate of change of concentration<br />

of X, dXldt, is thus proportional to the product of<br />

the concentrations of A and X in the solution-that is,<br />

dXldt= -kA'X, where k is a proportionality factor that is<br />

linked to quantities such as temperature and pressure and that<br />

provides a measure for the fraction of reactive collisions taking<br />

place and leading to the reaction A + X Y + B. Since, in<br />

the example taken, whenever a molecule of X disappears, a<br />

molecule of A disappears too, and a molecule of Yand one of<br />

B are formed, the rates of change of concentration are related:<br />

dXldt=dAldt= -dYldt= -dBldt.<br />

But if the collision between a molecule of X and a molecule


133 THE THREE STAGES OF THERMODYNAMICS<br />

of A can set off a chemical reaction, the collision between molecules<br />

of Y and B can set off the opposite reaction. A second<br />

reaction Y + B-.X +A thus occurs within the system described,<br />

bringing about a supplementary variation in the concentration<br />

of X, dX/dt = k' YB. The total variation in<br />

concentration of a chemical compound is given by the balance<br />

between the forward and the reverse reaction. In our example,<br />

dX/dt (= -dY/dt= . . .)= -kAX+k'YB.<br />

If left to itself, a system in which chemical reactions occur<br />

tends toward a state of chemical equilibrium. Chemical equilibrium<br />

is therefore a typical example of an "attractor" state.<br />

Whatever its initial chemical composition, the system spontaneously<br />

reaches this final stage, where the forward and reverse<br />

reactions compensate one another statistically so that<br />

there is no longer any overall variation in the concentrations<br />

(dX/dt=O). This compensation implies that the ratio between<br />

equilibrium concentrations is given by AXIYB= k'lk= K. This<br />

result is known as the "law of mass action," or Guldberg and<br />

Waage's law, and K is the equilibrium constant. The ratio between<br />

concentrations determined by the law of mass action<br />

corresponds to chemical equilibrium in the same way that uniformity<br />

of temperature (in the case of an isolated system) corresponds<br />

to thermal equilibrium. The corresponding entropy<br />

production vanishes.<br />

Before we deal with the thermodynamic description of<br />

chemical reactions, let us briefly consider an additional aspect<br />

of the kinetic description. The rate of chemical reactions is<br />

affected not only by the concentrations of the reacting molecules<br />

and thermodynamic parameters (for example, pressure<br />

and temperature) but also may be affected by the presence in<br />

the system of chemical substances that modify the reaction<br />

rate without themselves being changed in the process. Substances<br />

of this kind are known as "catalysts." Catalysts can,<br />

for instance, modify the value of the kinetic constants k or k'<br />

or even allow the system to follow a new "reaction path. " In<br />

biology, this role is played by specific proteins, the "enzymes."<br />

These macromolecules have a spatial configuration<br />

that allows them to modify the rate of a given reaction. Often<br />

they are highly specific and affect only one reaction. A possible<br />

mechanism for the catalytic effect of enzymes is to present


ORDER OUT OF CHAOS<br />

134<br />

different "reaction sites" to which the different molecules involwd<br />

in the reaction tend to attach themselves, thus increasing<br />

the likelihood of their coming into contact and reacting.<br />

One very important type of catalysis, particularly in biology,<br />

is the one in which the presence of a product is required<br />

for its own synthesis. In other words, in order to produce the<br />

molecule X we must begin with a system already containing X.<br />

Very frequently, for instance, the molecule X activates an enzyme.<br />

By attaching itself to the enzyme it stabilizes that particular<br />

configuration in which the reaction site is available. To<br />

such an autocatalysis process correspond reaction schemes<br />

such as A+ 2X 3X; in the presence of molecules X, a molecule<br />

A is converted into a molecule X. Therefore we need X to<br />

produce more X. This reaction may be symbolized by the reaction<br />

"loop":<br />

A<br />

One important feature of systems involving such "reaction<br />

loops" is that the kinetic equations describing the changes occurring<br />

in them are nonlinear differential equations .<br />

. If we apply the same method as above, the kinetic equation<br />

obtained for the reaction A+ 2X 3X is dX/dt = kA)(2, where<br />

the rate of variation of the concentration of X is proportional<br />

to the square of its concentration.<br />

Another very important class of catalytic reactions in biology<br />

is that of crosscatalysis-for example, 2X + Y3X, B +X<br />

Y + D, which may be represented by the loop of Figure 3.<br />

This is a case of crosscatalysis, since X is produced from Y,<br />

and simultaneously Y from X. Catalysis does not necessarily<br />

increase the reaction rate; it may, on the contrary, lead to inhibition,<br />

which can also be represented by suitable feedback<br />

loops.<br />

The peculiar mathematical properties of the nonlinear differential<br />

equations describing chemical processes with catalytic<br />

steps are vitally important, as we shall see later, for the<br />

thermodynamics of far-from-equilibrium chemical processes.<br />

In addition, as we have already mentioned, molecular biology


135 THE THREE STAGES OF THERMODYNAMICS<br />

,.... -----£ a<br />

D<br />

or<br />

A -.x<br />

e.x -.v.o<br />

2.X•V -..Jx<br />

x-.E<br />

Figure 3. This graph represents the reaction paths for the "Brusselator"<br />

reactions, which are further described in the text.<br />

has established that these loops play an essential role in metabolic<br />

functions. For example, the relation between nucleic<br />

acids and proteins can be described in terms of a crosscatalytic<br />

effect: nucleic acids contain the information to produce<br />

proteins, which in turn produce nucleic acids.<br />

In addition to the rates of chemical reactions, we must also<br />

consider the rates of other irreversible processes, such as heat<br />

transfer and the diffusion of matter. The rates of irreversible<br />

processes are also called fluxes and are denoted by the symbol<br />

J. There is no general theory from which we can derive the<br />

form of the rates or fluxes. In chemical reactions the rate depends<br />

on the molecular mechanism, as can be verified by the<br />

examples already indicated. The thermodynamics of irreversible<br />

processes introduces a second type of quantity: in addition<br />

to the rates, or fluxes, J, it uses "generalized forces," X, that<br />

"cause" the fluxes. The simplest example is that of heat conduction.<br />

Fourier's law tells us that the heat flux J is proportional<br />

to the temperature gradient. This temperature gradient<br />

is the "force" causing the heat flux. By definition, flux and<br />

forces both vanish at thermal equilibrium. As we shall see, the<br />

production of entropy P= diS/dt can be calculated from the<br />

flux and the forces.<br />

Let us consider the definition of the generalized force corresponding<br />

to a chemical reaction. Recall the reaction A+ X


ORDER OUT OF CHAOS 136<br />

-+ Y + B. We have seen how, at equilibrium, the ratio between<br />

concentrations is given by the law of mass action. As Theophile<br />

De Donder has shown, a "chemical force" can be introduced,<br />

the "affinity" a that determines the direction of the<br />

chemical reaction rate just as the temperature gradient determines<br />

the direction in which heat will flow. In the case of the<br />

reaction we are considering, the affinity is proportional to log<br />

KB YIAX, where K is the equilibrium constant. It is immediately<br />

apparent that the affinity a vanishes at equilibrium<br />

where, following the law of mass action, we have AXIBY=K.<br />

The affinity increases (in absolute value) when we drive the<br />

system away from equilibrium. We can see this if we eliminate<br />

from the system a fraction of the molecules B once they are<br />

formed through the reaction A + X-+ Y + B. Affinity can be said<br />

to measure the distance between the actual state of the system<br />

and its equilibrium state. Moreover, as we have mentioned, its<br />

sign determines the direction of the chemical reaction. If a is<br />

positive, then there are "too many" molecules B and Y, and<br />

the net reaction proceeds in the direction B + Y -+A+ X. On the<br />

contrary, if a is negative there are "too few" B and Y, and the<br />

net reaction proceeds in the opposite direction.<br />

Affinity as we have defined it is a way of rendering more<br />

precise the ancient affinity described by the alchemists, who<br />

deciphered the elective· relationships between chemical<br />

bodies-that is, the "likes" and "dislikes" of molecules. The<br />

idea that chemical activity cannot be reduced to mechanical<br />

trajectories, to the calm domination of dynamic laws, has been<br />

emphasized from the beginning. We could cite Diderot at<br />

length. Later, Nietzsche, in a different context, asserted that it<br />

was ridiculous to speak of "chemical laws," as though chemical<br />

bodies were governed by laws similar to moral laws. In<br />

chemistry, he protested, there is no constraint, and each body<br />

does as it pleases. It is not a matter of "respect" but of a power<br />

struggle, of the ruthless domination of the weaker by the<br />

stronger.2 Chemical equilibrium, with vanishing affinity, corresponds<br />

to the resolution of this conflict. Seen from this point<br />

of view, the specificity of thermodynamic affinity thus rephrases<br />

an age-old problem in modern language,3 the problem<br />

of the distinction between the legal and indifferent world of<br />

dynamic law, and the world of spontaneous and productive<br />

activity to which chemical reactions belong.


137 THE THREE STAGES OF THERMODYNAMICS<br />

Let us emphasize the basic conceptual distinction between<br />

physics and chemistry. In classical physics we can at least<br />

conceive of reversible processes such as the motion of a frictionless<br />

pendulum. To neglect irreversible processes in dynamics<br />

always corresponds to an idealization, but, at least in<br />

some cases, it is a meaningful one. The situation in chemistry<br />

is quite different. Here the processes that define chemistrychemical<br />

transformations characterized by reaction ratesare<br />

irreversible. For this reason chemistry cannot be reduced<br />

to the idealization that lies at the basis of classical or quantum<br />

mechanics, in which past and future play equivalent roles.<br />

As could be expected, all possible irreversible processes appear<br />

in entropy production. Each of them enters through the<br />

product of its rate or flux J multiplied by the corresponding<br />

force X. The total entropy production per unit time, P= diS/dt,<br />

is the sum of these contributions. Each of them appears<br />

through the product JX.<br />

We can divide thermodynamics into three large fields, the<br />

study of which corresponds to three successive stages in its<br />

development. Entropy production, the fluxes, and the forces<br />

are all zero at equilibrium. In the close-to-equilibrium region,<br />

where thermodynamic forces are "weak," the rates Jk are linear<br />

functions of the forces. The third field is called the "nonlinear"<br />

region, since in it the rates are in general more<br />

complicated functions of the forces. Let us first emphasize<br />

some general features of linear thermodynamics that apply to<br />

close-to-equilibrium situations.<br />

Linear Thermodynamics<br />

In 1931, Lars Onsager discovered the first general relations<br />

in nonequilibrium thermodynamics for the linear, near-toequilibrium<br />

region. These are the famous "reciprocity relations."<br />

In qualitative terms, they state that if a force-say,<br />

"one" (corresponding, for example, to a temperature gradient)-may<br />

influence a flux "two" (for example, a diffusion<br />

process), then force "two" (a concentration gradient) will also<br />

influence the flux .. one .. (the heat flow). This has indeed been<br />

verified. For example, in each case where a thermal gradient


ORDER OUT OF CHAOS<br />

138<br />

induces a process of diffusion of matter, we find that a concentration<br />

gradient can set up a heat flux through the system.<br />

The general nature of Onsager's relations has to be emphasized.<br />

It is immaterial, for instance, whether the irreversible<br />

processes take place in a gaseous, liquid, or solid medium.<br />

The reciprocity expressions are valid independently of any microscopic<br />

assumptions.<br />

Reciprocity relations have been the first results in the thermodynamics<br />

of irreversible processes to indicate that this was<br />

not some ill-defined no-man 's-tand but a worthwhile subject of<br />

study whose fertility could be compared with that of equilibrium<br />

thermodynamics. Equilibrium thermodynamics was<br />

an achievement of the nineteenth century, nonequilibrium<br />

thermodynamics was developed in the twentieth century, and<br />

Onsager's relations mark a crucial point in the shift of interest<br />

away from equilibrium toward nonequilibrium.<br />

A second general result in this field of linear, nonequilibrium<br />

thermodynamics bears mention here. We have already<br />

spoken of thermodynamic potentials whose extrema correspond<br />

to the states of equilibrium toward which thermodynamic evolution<br />

tends irreversibly. Such are the entropy S for is


139 THE THREE STAGES OF THERMODYNAMICS<br />

of time. Therefore its time variation dS = 0 vanishes. But we<br />

have seen that the time variation of entropy is made up of two<br />

terms-the entropy flow deS and the positive entropy production<br />

diS. Therefore, dS=O implies that deS= -diS


ORDER OUT OF CHAOS 140<br />

any specificity. Carnot or Darwin? The paradox mentioned in<br />

Chapter IV remains. There is still no connection between the<br />

appearance of natural <strong>org</strong>anized forms on one side, and on the<br />

other the tendency toward "f<strong>org</strong>etting" of initial conditions,<br />

along with the resulting dis<strong>org</strong>anization.<br />

Far "from Equilibrium<br />

At the root of nonlinear thermodynamics lies something quite<br />

surprising, something that first appeared to be a failure: in<br />

spite of much effort, the generalization of the theorem of minimum<br />

entropy production for systems in which the fluxes are<br />

no longer linear functions of the forces appeared impossible.<br />

Far from equilibrium, the system may still evolve to some<br />

steady state, but in general this state can no longer be characterized<br />

in terms of some suitably chosen potential (such as<br />

entropy production for near-equilibrium states).<br />

The absence of any potential function raises a new question:<br />

What can we say about the stability of the states toward which<br />

the system evolves? Indeed, as long as the attractor state is<br />

defined by the minimum of a potential such as the entropy<br />

production, its stability is guaranteed. It is true that a fluctuation<br />

may shift the system away from this minimum. The second<br />

law of thermodynamics, however, imposes the return<br />

toward the attractor. The system is thus "immune" with respect<br />

to fluctuations. Thus whenever we define a potential, we<br />

are describing a "stable world" in which systems follow an<br />

evolution that leads them to a static situation that is established<br />

once and for all.<br />

When the thermodynamic forces acting on a system become<br />

such that the linear region is exceeded, however, the stability<br />

of the stationary state, or its independence from fluctuations,<br />

can no longer be taken for granted. Stability is no longer the<br />

consequence of the general laws of physics. We must examine<br />

the way a stationary state reacts to the different types of fluctuation<br />

produced by the system or its environment. In some<br />

cases, the analysis leads to the conclusion that a state is "unstable"-in<br />

such a state, certain fluctuations, instead of re-


141 THE THREE STAGES OF THERMODYNAMICS<br />

gressing, may be amplified and invade the entire system,<br />

compelling it to evolve toward a new regime that may be<br />

qualitatively quite different from the stationary states corresponding<br />

to minimum entropy production.<br />

Thermodynamics leads to an initial general conclusion concerning<br />

systems that are liable to escape the type of order governing<br />

equilibrium. These systems have to be "far from<br />

equilibrium." In cases where instability is possible, we have to<br />

ascertain the threshold, the distance from equilibrium, at<br />

which fluctuations may lead to new behavior, different from<br />

the "normal" stable behavior characteristic of equilibrium or<br />

near-equilibrium systems.<br />

Why is this conclusion so interesting?<br />

Phenomena of this kind are well known in the field of hydrodynamics<br />

and fluid flow. For instance, it has long been known<br />

that once a certain flow rate of flux has been reached, turbulence<br />

may occur in a fluid. Michel Serres has recently recalled4<br />

that the early atomists were so concerned about<br />

turbulent flow that it seems legitimate to consider turbulence<br />

as a basic source of inspiration of Lucretian physics. Sometimes,<br />

wrote Lucretius, at uncertain times and places, the<br />

eternal, universal fall of the atoms is disturbed by a very slight<br />

deviati0n-the "clinamen." The resulting vortex gives rise to<br />

the world, to all natural things. The clinamen, this spontaneous,<br />

unpredictable deviation, has often been criticized as<br />

one of the main weaknesses of Lucretian physics, as being<br />

something introduced ad hoc. In fact, the contrary is truethe<br />

clinamen attempts to explain events such as laminar flow<br />

ceasing to be stable and spontaneously turning into turbulent<br />

flow. Today hydrodynamic experts test the stability of fluid<br />

flow by introducing a perturbation that expresses the effect of<br />

molecular disorder added to the average flow. We are not so far<br />

from the clinamen of Lucretius!<br />

For a long time turbulence was identified with disorder or<br />

noise. Today we know that this is not the case. Indeed, while<br />

turbulent motion appears as irregular or chaotic on the macroscopic<br />

scale, it is, on the contrary, highly <strong>org</strong>anized on the<br />

microscopic scale. The multiple space and time scales involved<br />

in turbulence correspond to the coherent behavior of<br />

millions and millions of molecules. Viewed in this way, the<br />

transition from laminar flow to turbulence is a process of self-


ORDER OUT OF CHAOS 142<br />

<strong>org</strong>anization. Part of the energy of the system, which in laminar<br />

flow was in the thermal motion of the molecules, is being<br />

transferred to macroscopic <strong>org</strong>anized motion.<br />

The "Benard instability" is another striking example of the<br />

instability of a stationary state giving rise to a phenomenon of<br />

spontaneous self-<strong>org</strong>anization. The instability is due to a vertical<br />

temperature gradient set up in a horizontal liquid layer. The<br />

lower surface of the latter is heated to a given temperature,<br />

which is higher than that of the upper surface. As a result of<br />

these boundary conditions, a permanent heat flux is set up,<br />

moving from the bottom to the top. When the imposed gradient<br />

reaches a threshold value, the fluid's state of rest-the<br />

stationary state in which heat is conveyed by conduction<br />

alone, without convection-becomes unstable. A convection<br />

corresponding to the coherent motion of ensembles of molecules<br />

is produced, increasing the rate of heat transfer. Therefore,<br />

for given values of the constraints (the gradient of<br />

temperature), the entropy production of the system is increased;<br />

this contrasts with the theorem of minimum entropy<br />

production. The Benard instability is a spectacular phenomenon.<br />

The convection motion produced actually consists<br />

of the complex spatial <strong>org</strong>anization of the system. Millions of<br />

molecules move coherently, forming hexagonal convection<br />

cells of a characteristic size.<br />

In Chapter IV we introduced Boltzmann's order principle,<br />

which relates entropy to probability as expressed by the number<br />

of complexions P. Can we apply this relation here? To each<br />

distribution of the velocities of the molecules corresponds a<br />

number of complexions. This number measures the number of<br />

ways in which we can realize the velocity distribution by attributing<br />

some velocity to each molecule. The argument runs<br />

parallel to that in Chapter IV, where we expressed the number<br />

of complexions in terms of the distributions of molecules between<br />

two boxes. Here also the number of complexions is<br />

large when there is disorder-that is, a wide dispersion of velocities.<br />

In contrast, coherent motion means that many molecules<br />

travel with nearly the same speed (small dispersion of<br />

velocities). To such a distribution corresponds a number of<br />

complexions P so low that there seems almost no chance for<br />

the phenomenon of self-<strong>org</strong>anization to occur. Yet it occurs!<br />

We see, therefore, that calculating the number of complexions,


1-43<br />

THE THREE STAGES OF THERMODYNAMICS<br />

which entails the hypothesis of an equal a priori probability for<br />

each molecular state, is misleading. Its irrelevance is particularly<br />

obvious as far as the genesis of the new behavior is<br />

concerned. In the case of the Benard instability it is a fluctuation,<br />

a microscopic convection current, which would have<br />

been doomed to regression by the application of Boltzmann's<br />

order principle, but which on the contrary is amplified until it<br />

invades the whole system. Beyond the critical value of the imposed<br />

gradient, a new molecular order has thus been produced<br />

spontaneously. It corresponds to a giant fluctuation stabilized<br />

through energy exchanges with the outside world.<br />

In far-from-equilibrium conditions, the concept of probability<br />

that underlies Boltzmann's order principle is no longer<br />

valid in that the structures we observe do not correspond to a<br />

maximum of complexions. Neither can they be related to a<br />

minimum of the free energy F = E- TS. The tendency toward<br />

leveling out and f<strong>org</strong>etting initial conditions is no longer a general<br />

property. In this context, the age-old problem of the origin<br />

cf life appears in a different perspective. It is certainly true that<br />

life is incompatible with Boltzmann's order principle but not with<br />

the kind of behavior that can occur in far-from-equilibrium<br />

conditions.<br />

Classical thermodynamics leads to the concept of "equilibrium<br />

structures" such as crystals. Benard cells are structures<br />

too, but of a quite different nature. That is why we have<br />

introduced the notion of "dissipative structures," to emphasize<br />

the close association, at first paradoxical, in such situations<br />

between structure and order on the one side, and<br />

dissipation or waste on the other. We have seen in Chapter IV<br />

that heat transfer was considered a source of waste in classical<br />

thermodynamics. In the Benard cell it becomes a source of<br />

order.<br />

The interaction of a system with the outside world, its embedding<br />

in nonequilibrium conditions, may become in this way<br />

the starting point for the formation of new dynamic states of<br />

matter-dissipative structures. Dissipative structures actually<br />

correspond to a form of supramolecular <strong>org</strong>anization. Although<br />

the parameters describing crystal structures may be<br />

derived from the properties of the molecules of which they are<br />

composed, and in particular from the range of their forces of<br />

attraction and repulsion, Benard cells, like all dissipative


ORDER OUT OF CHAOS<br />

144<br />

structures, are essentially a reflection of the global situation of<br />

nonequilibrium producing them. The parameters describing<br />

them are macroscopic; they are not of the order of 10-8 cm,<br />

like the distance between the molecules of a crystal, but of the<br />

order of centimeters. Similarly, the time scales are differentthey<br />

correspond not to molecular times (such as periods of<br />

vibration of individual molecules, which may correspond to<br />

about 10-15 sec) but to macroscopic times: seconds, minutes,<br />

or hours.<br />

Let us return to the case of chemical reactions. There are<br />

some fundamental differences from the Benard problem. In<br />

the Benard cell the instability has a simple mechanical origin.<br />

When we heat the liquid layer from below, the lower part of the<br />

fluid becomes less dense, and the center of gravity rises. It is<br />

therefore not surprising that beyond a critical point the system<br />

tilts and convection sets in.<br />

But in chemical systems there are no mechanical features of<br />

this type. Can we expect any self-<strong>org</strong>anization? Our mental<br />

image of chemical reactions corresponds to molecules speeding<br />

through space, colliding at random in a chaotic way. Such<br />

an image leaves no place for self-<strong>org</strong>anization, and this may be<br />

one of the reasons why chemical instabilities have only recently<br />

become a subject of interest. There is also another difference.<br />

All flows become turbulent at a "sufficiently" large<br />

distance from equilibrium (the threshold is measured by dimensionless<br />

numbers such as Reynolds' number). This is not<br />

true for chemical reactions. Being far from equilibrium is a<br />

necessary requirement but not a sufficient one. For many<br />

chemical systems, whatever the constraints imposed and the<br />

rate of the chemical changes produced, the stationary state<br />

remains stable and arbitrary fluctuations are damped, as is the<br />

case in the close-to-equilibrium range. This is true in particular<br />

of systems in which we have a chain of transformations of<br />

the type A-+B-+C-+D . .. and that may be described by linear<br />

differential equations.<br />

The fate of the fluctuations perturbing a chemical system,<br />

as well as the kinds of new situations to which it may evolve,<br />

thus depend on the detailed mechanism of the chemical reactions.<br />

In contrast with close-to-equilibrium situations, the<br />

behavior of a far-from-equilibrium system becomes highly specific.<br />

There is no longer any universally valid law from which


145 THE THREE STAGES OF THERMODYNAMICS<br />

the overall behavior of the system can be deduced. Each system<br />

is a separate case; each set of chemical reactions must be<br />

investigated and may well produce a qualitatively different behavior.<br />

Nevertheless, one general result has been obtained, namely<br />

a necessary condition for chemical instability: in a chain of<br />

chemical reactions occurring in the system, the only reaction<br />

stages that, under certain conditions and circumstances, may<br />

jeopardize the stability of the stationary state are precisely the<br />

"catalytic loops"-stages in which the product of a chemical<br />

reaction is involved in its own synthesis. This is an interesting<br />

conclusion, since it brings us closer to some of the fundamental<br />

achievements of modern molecular biology (see Figure 4).<br />

Figure 4. Catalytic loops correspond to nonlinear terms. In the case of a<br />

one-independent-variable problem, this means the occurrence of at least<br />

one term where the independent variable appears with a power higher than<br />

1; in this simple case, it is easy to see the relation between such nonlinear<br />

terms and the potential instability of stationary states.<br />

Let us take for the independent variable X the time evolution dXIdt= f(X). It<br />

is always possible to decompose f(X) in two functions representing a gain<br />

and a loss f + (X) and f _ (X) , each of which is positive or 0, such that<br />

f(X) = f +(X)- f _(X). In this way, stationary states (dX!dt= 0) correspond to<br />

values where f +(X)= f _(X).<br />

Those states are graphically given by the intersections of the two graphs<br />

plotting f + and f _. If f + and f _ are linear, there can only be one intersection.<br />

In other cases, the type of the intersection permits us to infer the stability of<br />

the stationary state.<br />

Four cases are possible:<br />

Sl: stable with respect to negative fluctuations, unstable with respect to<br />

positive ones: If the system deviates slightly to the left of Sl , the positive<br />

difference between f + and f _ will reduce this deviation back to Sl; deviations<br />

to the right will be amplified.<br />

SS: stable with respect to positive and negative fluctuations.<br />

IS: stable only with respect to positive fluctuations.<br />

II: unstable with respect to positive and negative fluctuations.<br />

X


ORDER OUT 0 CHAOS 146<br />

Beyond the Threshold of Chemical Instability<br />

Today the study of chemical instabilities is common. Both theoretical<br />

and experimental work are being pursued in a large<br />

number of institutions and laboratories. Indeed, as will become<br />

clear, these investigations are of interest to a wide range<br />

of scientists-not only to mathematicians, physicists, chemists,<br />

and biologists, but also to economists and sociologists.<br />

In far-from-equilibrium conditions various new phenomena<br />

appear beyond the threshold of chemical instability. To describe<br />

them in a concrete fashion, it is useful to start with a<br />

simplified theoretical model, one that has been developed at<br />

Brussels during the past decade. American scientists have<br />

called this model the "Brusselator," and this name is used in<br />

scientific literature (Geographical associations seem to have<br />

become the rule in this field; in addition to the Brusselator,<br />

there is an "Oregonator," and most recently a "Paloaltonator"<br />

!). Let us briefly describe the Brusselator. The steps<br />

responsible for instability have already been noted (see Figure<br />

3). The product X, synthetized from A and broken down into<br />

the form of E, is linked by a relationship of crosscatalysis to<br />

produce Y. X is produced from Y during a trimolecular step<br />

but, conversely, Y is synthetized by a reaction between X and<br />

a product B.<br />

In this model, the concentrations of the products A, B, D, and<br />

E are given parameters (the "control substances"). The behavior<br />

of the system is explored for increasing values of B, with A<br />

remaining constant. The stationary state toward which such a<br />

system is likely to evolve-the state for which dX/dt = dY/dt<br />

= 0-corresponds to concentrations X0 =A and Y0 = BIA. This<br />

can be easily verified by writing the kinetic equations and<br />

looking for the stationary state. However, this stationary state<br />

ceases to be stable as soon as the concentration of B exceeds a<br />

critical threshold (everything else being kept equal). After the<br />

critical threshold has been reached, the stationary state becomes<br />

an unstable "focus" and the system leaves this focus to<br />

reach a "limit cycle."


147 THE THREE STAGES OF THERMODYNAMICS<br />

0<br />

2 3 4 y<br />

Figure 5. This scheme represents concentration of component X vs. concentration<br />

of component Y. The cycle's focus (point S) is the stationary state,<br />

which is unstable for 8>(1 + A2). All the trajectories (of which five are plotted),<br />

whatever their intitial state, lead to the same cycle.<br />

Instead of remaining stationary, the concentrations of X and Y<br />

begin to oscillate with a well-defined periodicity. The oscillation<br />

period depends both on the kinetic constants characterizing<br />

the reaction rates and the boundary conditions imposed on<br />

the system as a whole (temperature, concentration of A., B,<br />

etc.).<br />

Beyond the critical threshold the system spontaneously<br />

leaves the stationary state X0=A, Y0=BIA as the result of<br />

fluctuations. Whatever the initial conditions, it approaches the<br />

limit cycle, the periodic behavior of which is stable. We therefore<br />

have a periodic chemical process-a chemical clock. Let<br />

us pause a moment to emphasize how unexpected such a phenomenon<br />

is. Suppose we have two kinds of molecules, "red"<br />

and "blue." Because of the chaotic motion of the molecules,<br />

we would expect that at a given moment we would have more<br />

red molecules, say, in the left part of a vessel. Then a bit later<br />

more blue molecules would appear, and so on. The vessel<br />

would appear to us as "violet," with occasional irregular


ORDER OUT OF CHAOS 148<br />

flashes of red or blue. However, this is not what happens with<br />

a chemical clock; here the system is all blue, then it abruptly<br />

changes its col or to red, then again to blue. Because all these<br />

changes occur at regular time intervals, we have a coherent<br />

process.<br />

Such a degree of order stemming from the activity of biIlions<br />

of molecules seems incredible, and indeed, if chemical clocks<br />

had not been observed, no one would believe that such a process<br />

is possible. To change color all at once, molecules must<br />

have a way to "communicate." The system has to act as a<br />

whole. We will return repeatedly to this key word, communicate,<br />

which is of obvious importance in so many fields, from<br />

chemistry to neurophysiology. Dissipative structures introduce<br />

probably one of the simplest physical mechanisms for<br />

communication.<br />

There is an interesting difference between the simplest kind<br />

of mechanical oscillator, the spring, and a chemical clock. The<br />

chemical clock has a well-defined periodicity corresponding<br />

to the limit cycle its trajectory is following. On the contrary, a<br />

spring has a frequency that is amplitude-dependent. From this<br />

point of view a chemical clock is more reliable as a timekeeper<br />

than a spring.<br />

But chemical clocks are not the only type of self-<strong>org</strong>anization.<br />

Until now diffusion has been neglected. All substances were<br />

assumed to be evenly distributed over the reaction space. This<br />

is an idealization; small fluctuations will always lead to differences<br />

in concentrations and thus to diffusion. We therefore<br />

have to add diffusion to the chemical reaction equations. The<br />

diffusion-reaction equations of the Brusselator display an astonishing<br />

range of behaviors available to this system. Indeed,<br />

whereas at equilibrium and near-equilibrium the system remains<br />

spatially homogeneous, the diffusion of the chemical<br />

throughout the system induces, in the far-from-equilibrium region,<br />

the possibility of new types of instability, including the<br />

amplification of fluctuations breaking the initial spatial symmetry.<br />

Oscillations in time, chemical clocks, thus cease to be<br />

the only kind of dissipative structure available to the system.<br />

Far from it; for example, oscillations may appear that are now<br />

both time- and space-dependent. They correspond to chemical<br />

waves of X and Y concentrations that periodically pass through<br />

the system.


149<br />

THE THREE STAGES OF THERMODYNAMICS<br />

X hO<br />

hO.S8<br />

3<br />

o<br />

X h 1.10<br />

o<br />

h 1.88<br />

I--_ror.-____ 1 _ _____ _...,<br />

o<br />

X h2.04<br />

o<br />

h 3.435<br />

3<br />

21....-- - - -<br />

-- - --'"<br />

1 0<br />

o<br />

Figure 6. Chemical waves simulated on computer: successive steps of<br />

evolution of spatial profile of concentration of constituent X in the "Brusselator"<br />

trimolecular model. At time t= 3.435 we recover the same distribution as<br />

at time t= O. Concentration of A and B: 2, 5.45 (B>[1 + A2]). Diffusion coefficients<br />

for X and Y: 810-3,410-3•<br />

In addition, especially when the values of the diffusion constants<br />

of X and Y are quite different from each other, the system<br />

may display a stationary, time-independent behavior, and<br />

stable spatial structures may appear.


ORDER OUT OF CHAOS<br />

150<br />

Here we must pause once again, this time to emphasize how<br />

much the spontaneous formation of spatial structures contradicts<br />

the laws of equilibrium physics and Boltzmann's order<br />

principle. Again, the number of complexions corresponding to<br />

such structures would be extremely small in comparison with<br />

the number in a uniform distribution. Still , nonequilibrium<br />

processes may lead to situations that would appear impossible<br />

from the classical point of view.<br />

The number of different dissipative structures compatible<br />

with a given set of boundary conditions may be increased still<br />

further when the problem is studied in two or three dimensions<br />

instead of one. In a circular, two-dimensional space, for<br />

instance, the spatially structured stationary state may be<br />

characterized by the occurrence of a privileged axis.<br />

X<br />

--· --·<br />

Figure 7. Stationary state with privileged axis obtained by computer simulation.<br />

Concentration X is a function of geometrical coordinates p,a in the<br />

horizontal plane. The location of the perturbation applied to the uniform unstable<br />

solution (X Q<br />

, Y0) is indicated by an arrow.<br />

This corresponds to a new, extremely interesting symmetrybreaking<br />

process, especially when we recall that one of the<br />

first stages in morphogenesis of the embryo is the formation of<br />

a gradient in the system. We will return to these problems later<br />

in this chapter and again in Chapter VI.


151 THE THREE STAGES OF THERMODYNAMICS<br />

Up to now it has been assumed that the "control substances"<br />

(A, B, D, and E) are uniformly distributed throughout<br />

the reaction system. If this simplification is abandoned, additional<br />

phenomena can occur. For example, the system takes<br />

on a "natural size ," which is a function of the parameters describing<br />

it. In this way the system determines its own intrinsic<br />

size-that is, it determines the region that is spatially structured<br />

or crossed by periodic concentration waves.<br />

These results still give a very incomplete picture of the variety<br />

of phenomena that may occur far from equilibrium. Let us<br />

first mention the possibility of multiple states far from equilibrium.<br />

For given boundary conditions there may appear<br />

more than one stationary state-'-for instance one rich in the<br />

chemical X, the other poor. The shift from one state to another<br />

plays an important role in control mechanisms as they have<br />

been described in biological systems.<br />

Since the classical work of Lyapounov and Poincare,<br />

characteristic points such as focus or lines such as limit cycles<br />

Iii<br />

y<br />

Figure 8. (a) Bromide-ion concentration in the Belousov-Zhabotinsky reaction<br />

at times t1 and t1 + T (cf. R. H. Simoyi, A. Wolf, and H. L. Swinney,<br />

Physics Review Letters, Vol. 49 (1982), p. 245; see J. Hirsch, "Condensed<br />

Matter Physics," and on computers, Physics Today (May 1983), pp. 44-52).<br />

(b) Attractor lines calculated by Hao Bai-lin for a Brusselator with external<br />

periodic supply of component X (personal communication).


ORDER OUT OF CHAOS 152<br />

were known to mathematicians as the "attractors" of stable<br />

systems. What is new is their application to chemical systems.<br />

It is worth noting that the first paper dealing with instabilities<br />

in reaction-diffusion systems was published by Thring in 1952.<br />

In recent years new types of attractors have been identified.<br />

They appear only when the number of independent variables<br />

increases (there are two independent variables in the Brusselator,<br />

the variables X and Y). In particular, we can get "strange<br />

attractors" that do not correspond to periodic behavior.<br />

Figure 8, which summarizes some calculations by Hao Bailin,<br />

gives an idea of such very complicated attractor lines calculated<br />

for a model generalizing the Brusselator through the<br />

addition of an external periodic supply of X. What is remarkable<br />

is that most of the possibilities we have described have<br />

been observed in in<strong>org</strong>anic chemistry as well as in a number of<br />

biological situations.<br />

In in<strong>org</strong>anic chemistry the best-known example is the<br />

Belousov-Zhabotinsky reaction discovered in the early 1960s.<br />

The corresponding reaction scheme, the Oregonator, introduced<br />

by Noyes and his colleagues, is in essence similar to the<br />

Brusselator though more complex. The Belousov-Zhabotinsky<br />

reaction consists of the oxidation of an <strong>org</strong>anic acid (malonic<br />

acid) by a potassium bromate in the presence of a suitable catalyst,<br />

cerium, manganese, or ferroin.<br />

INFLOW<br />

MALONIC . ::: ...rP U ::- M :: P :


153 THE THREE STAGES OF THERMODYNAMICS<br />

Various experimental conditions may be set up giving different<br />

forms of auto<strong>org</strong>anization within the same system-a chemical<br />

clock, a stable spatial differentiation, or the formation of<br />

waves of chemical activity over macroscopic distances.5<br />

Let us now turn to a matter of the greatest interest: the relevance<br />

of these results for the understanding of living systems.<br />

The Encounter with Molecular Biology<br />

Earlier in this chapter we showed that in far-from-equilibrium<br />

conditions various types of self-<strong>org</strong>anization processes may<br />

occur. They may lead to the appearance of chemical oscillations<br />

or to spatial structures . We have seen that the basic condition<br />

for the appearance of such phenomena is the existence<br />

of catalytic effects.<br />

Although the effects of "nonlinear" reactions (the presence<br />

of the reaction product) have a feedback action on their<br />

"cause" and are comparatively rare in the in<strong>org</strong>anic world,<br />

molecular biology has discovered that they are virtually the<br />

rule as far as living systems are concerned. Autocatalysis (the<br />

presence of X accelerates its own synthesis), autoinhibition<br />

(the presence of X blocks a catalysis needed to synthesize it),<br />

and crosscatalysis (two products belonging to two different reaction<br />

chains activate each other's synthesis) provide the classical<br />

regulation mechanism guaranteeing the coherence of the<br />

metabolic function.<br />

Let us emphasize an interesting difference. In the examples<br />

known in in<strong>org</strong>anic chemistry, the molecules involved are<br />

simple but the reaction mechanisms are complex-in the<br />

Belousov-Zhabotinsky reaction, about thirty compounds have<br />

been identified. On the contrary, in the many biological examples<br />

we have, the reaction scheme is simple but the molecules<br />

(proteins, nucleic acids, etc.) are highly complex and specific.<br />

This can hardly be an accident. Here we encounter an initial<br />

element marking the difference between physics and biology.<br />

Biological systems have a past. Their constitutive molecules<br />

are the result of an evolution; they have been selected to take<br />

part in the autocatalytic mechanisms to generate very specific<br />

forms of <strong>org</strong>anization processes.<br />

A description of the network of metabolic activations and


ORDER OUT OF CHAOS<br />

154<br />

inhibitions is an essential step in understanding the functional<br />

logic of biological systems. This includes the triggering of syntheses<br />

the moment they are needed and the blocking of those<br />

chemical reactions whose unused products would accumulate<br />

in the cell.<br />

The basic mechanism through which molecular biology explains<br />

the transmission and exploitation of genetic information<br />

is itself a feedback loop, a "nonlinear" mechanism. Deoxyribonucleic<br />

acid (DNA), which contains in sequential form all<br />

the information required for the synthesis of the various basic<br />

proteins needed in cell building and functioning, participates<br />

in a sequence of reactions during which this information is translated<br />

into the form of different protein sequences. Among the<br />

proteins synthesized, some enzymes exert a feedback action<br />

that activates or controls not only the different transformation<br />

stages but also the autocatalytic mechanism of DNA replication,<br />

by which genetic information is copied at the same rate<br />

as the cells multiply.<br />

Here we have a remarkable case of the convergence of two<br />

sciences. The understanding attained here required complementary<br />

developments in physics and biology, one toward the<br />

complex and the other toward the elementary.<br />

Indeed, from the point of view of physics, we now investigate<br />

"complex" situations far removed from the ideal situations<br />

that can be described in terms of equilibrium thermodynamics.<br />

On the other hand, molecular biology succeeded in relating<br />

living structures to a relatively small number of basic<br />

biomolecules. Investigating the diversity of chemical mechanisms,<br />

it discovered the intricacy of the metabolic reaction<br />

chains, the subtle, complex logic of the control, inhibition, and<br />

activation of the catalytic function of the enzymes associated<br />

with the critical step of each of the metabolic chains. In this<br />

way molecular biology provides the microscopic basis for the<br />

instabilities that may occur in far-from-equilibrium conditions.<br />

In a sense, living systems appear as a well-<strong>org</strong>anized factory:<br />

on the one hand, they are the site of multiple chemical<br />

transformations; on the other, they present a remarkable "spacetime"<br />

<strong>org</strong>anization with highly nonuniform distribution of biochemical<br />

material. We can now link function and structure.<br />

Let us briefly consider two examples that have been studied<br />

extensively in the past few years.


155 THE THREE STAGES OF THERMODYNAMICS<br />

First we shall consider glycolysis, the chain of metabolic<br />

reactions during which glucose is broken down and an energyrich<br />

substance ATP (adenosine triphosphate) is synthetized,<br />

providing an essential source of energy common to all living<br />

cells. For each glucose molecule that is broken down , two<br />

molecules of ADP (adenosine disphosphate) are transformed<br />

into two molecules of ATP. Glycolysis provides a fine example<br />

of how complemetary the analytical approach of biology and<br />

the investigation of stability in far-from-equilibrium conditions<br />

are. 6<br />

Biochemical experiments have discovered the existence of<br />

temporal oscillations in concentrations related to the glycolytic<br />

cycle. 7 It has been shown that these oscillations are determined<br />

by a key step in the reaction sequence, a step activated<br />

by ADP and inhibited by ATP. This is a typical nonlinear phenomenon<br />

well suited to regulate metabolic functioning. Indeed,<br />

each time the cell draws on its energy reserves, it is<br />

exploiting the phosphate bonds, and AT P is converted into<br />

ADP. ADP accumulation inside the cell thus signifies intensive<br />

energy consumption and the need to replenish stocks. ATP<br />

accumulation, on the other hand, means that glucose can be<br />

broken down at a slower rate.<br />

Theoretical investigation of this process has shown that this<br />

mechanism is indeed liable to produce an oscillation phenomenon,<br />

a chemical clock. The theoretically calculated values<br />

of the chemical concentrations necessary to produce<br />

oscillation and the period of the cycle agree with the experimental<br />

data. Glycolytic oscillation produces a modulation of<br />

all the cell's energy processes which are dependent on ATP<br />

concentration and therefore indirectly on numerous other metabolic<br />

chains.<br />

We may go farther and show that in the glycolytic pathway<br />

the reactions controlled by some of the key enzymes are in farfrom-equilibrium<br />

conditions. Such calculations have been reported<br />

by Benno Hesss and have since been extended to other<br />

systems. Under usual conditions the glycolytic cycle corresponds<br />

to a chemical clock, but changing these conditions can<br />

induce spatial pattern formations in complete agreement with<br />

the predictions of existing theoretical models.<br />

A living system appears very complex from the thermodynamic<br />

point of view. Certain reactions are close to equi-


ORDER OUT OF CHAOS 156<br />

librium, and others are not. Not everything in a living system<br />

is "alive." The energy flow that crosses it somewhat resembles<br />

the flow of a river that generally moves smoothly but that<br />

from time to time tumbles down a waterfall; which liberates<br />

part of the energy it contains.<br />

Let us consider another biological process that also has<br />

been studied from the point of view of stability: the aggregation<br />

of slime molds, the Acrasiales amoebas (Dictyostelium<br />

discoideum). This process9A is an interesting case on the borderline<br />

between unicellular and pluricellular biology. When<br />

The aggregation of cellular slime molds furnishes a particularly re<br />

markable example of a self-<strong>org</strong>anization phenomenon in a biological<br />

sYltem in which a chemical clock plays an essential role. See Figure A.<br />

fruiting ?<br />

body <br />

spores 0<br />

00<br />

I<br />

(Y:J growth<br />

Q)<br />

\<br />

((t'': kx<br />

\ , .. , I<br />

, -<br />

.. - - '<br />

;'; aggregation<br />

\<br />


157 THE THREE STAGES OF THERMODYNAMICS<br />

In Dictyostelium discoideum, the aggregation proceeds in a periodic manner.<br />

Movies of aggregation process show the existence of concentric waves<br />

of amoebae moving toward the center with a periodicity of several minutes.<br />

The nature of the chemotactic factor is known: it is cyclic AMP (cAMP), a<br />

substance involved in numerous biochemical processes such as hormonal<br />

regulations. The aggregation centers release the signals of cAMP in a periodic<br />

fashion. The other cells respond by moving toward the centers and by<br />

relaying the signals to the periphery of the aggregation territory. The existence<br />

of a mechanism of relay of the chemotactic signals allows each center<br />

to control the aggregation of some 105 amoebae.<br />

The analysis of a model of the process of aggregation reveals the existence<br />

of two types of bifurcations. First the aggregation itself represents a<br />

breaking of spatial symmetry. The second bifurcation breaks the temporal<br />

symmetry.<br />

Initially the amoebae are homogeneously distributed. When some of them<br />

begin to secrete the chemotactic signals, there appear local fluctuations in<br />

the concentration of cAMP. For a critical value of some parameter of the<br />

system (diffusion coefficient of cAMp, motility of the amoebae, etc.), fluctuations<br />

are amplified: the homogeneous distribution becomes unstable and the<br />

amoebae evolve toward an inhomogeneous. distribution in space. This new<br />

distribution corresponds to the accumulation of amoebae around aggregation<br />

centers.<br />

To understand the origin of the periodicity in the aggregation of D. discoideum,<br />

it is necessary to study the mechanism of synthesis of the chemotactic<br />

signal. On the basis of experimental observations one can describe<br />

this mechanism by the scheme of Figure B.<br />

_----+:.-- cAM P <br />

Figure B<br />

ATP<br />

Y<br />

cAMP<br />

On the surface of the cell, receptors (R) bind the molecules of cAMP. The<br />

receptor faces the extracellular medium and is functionally linked to an enzyme,<br />

adenylate cyclase (C), which transforms intracellular ATP into cAMP.<br />

The cAMP thus synthesized is transported across the membrane into the<br />

extracellular medium, where it is degraded by phosphodiesterase, an enzyme<br />

that is secreted by the amoebae. The experiments show that binding of<br />

extracellular cAMP to the membrane receptor activates adenylate cyclase<br />

(positive feedback indicated by +).<br />

On the basis of this autocatalytic regulation, the analysiS of a model for


ORDER OUT OF CH AOS<br />

158<br />

cAMP synthesis has permitted unification of different types of behavior observed<br />

during aggregation.se<br />

Two key parameters of the model are the concentrations of adenylate<br />

cyclase (s) and of phosphodiesterase (k). Figure C (redrawn from A. GoLD·<br />

BETER and L. SEGEL, Differentiation, Vol. 17 [1980], pp. 127-35), shows the<br />

behavior of the modelized system in the space formed by s and k.<br />

cu<br />

-<br />

0<br />

:>.<br />

c<br />

cu<br />

"'Q<br />

<<br />

A<br />

Phosphodiesterase ,<br />

k<br />

Figure C<br />

Three regions can be distinguished for different values of k and s. Region<br />

A corresponds to a stable, nonexcitable stationary state; region B to a stationary<br />

state stable but excitable: the system is capable of amplifying small<br />

perturbations in the concentration of cAMP in a pulsatory manner (and thus<br />

of relaying cAMP signals); region C corresponds to a regime of sustained<br />

oscillations around an unstable stationary state.<br />

The arrow indicates a possible "developmental path" corresponding to a<br />

rise in phosphodiesterase (k) and adenylate cyclase (s), a rise that is observed<br />

to occur after the beginning of starvation. The crossing of regions A,<br />

B and C corresponds to the observed change of behavior: cells are at first<br />

incapable of responding to extracellular cAMP signals; thereafter they relay<br />

these signals and, finally, they become capable of synthetizing them periodically<br />

in an autonomous way. The aggregation centers would thus be the<br />

cells for which the parameters s and k have reached the more rapidly a point<br />

located inside region 0 after starvation has begun.


159 THE THREE STAGES OF THERMODYNAMICS<br />

the environment in which these amoebas live and multiply becomes<br />

poor in nutrients, they undergo a spectacular transformation.<br />

(See Figure A.) Starting as a population of isolated<br />

cells, they join to form a mass composed of several tens of<br />

thousands of cells. This "pseudoplasmodium" then undergoes<br />

differentiation, all the while changing shape. A "foot" forms,<br />

consisting of about one third of the cells and containing abundant<br />

cellulose. This foot supports a round mass of spores,<br />

which will detach themselves and spread, multiplying as soon<br />

as they come in contact with a suitable nutrient medium and<br />

thus forming a new colony of amoebas. This is a spectacular<br />

example of adaptation to the environment. The population<br />

lives in one region until it has exhausted the available resources.<br />

It then goes through a metamorphosis by means of<br />

which it acquires the mobility to invade other environments.<br />

An investigation of the first stage of the aggregation process<br />

reveals that it begins with the onset of displacement waves in<br />

the amoeba population, with a pulsating motion of convergence<br />

of the amoebaes toward a "center of attraction,"<br />

which appears to be produced spontaneously. Experimental<br />

investigation and modelization have shown that this migration<br />

is a response by the cells to the existence in the environment<br />

of a concentration gradient in a key substance, cyclic AMP,<br />

which is periodically produced by an amoeba which is the attractor<br />

center and later by other cells through a relay mechanism.<br />

Here we again see the remarkable role of chemical<br />

clocks. They provide, as we have already stressed, new means<br />

of communication. In the present case, the self-<strong>org</strong>anization<br />

mechanism leads to communication between cells.<br />

There is another aspect we wish to emphasize. Slime mold<br />

aggregation is a typical example of what may be termed "order<br />

through fluctuations" : the setting up of the attractor center<br />

giving off the AMP indicates that the metabolic regime corresponding<br />

to a normal nutritive environment has become unstable-that<br />

is, the nutritive environment has become<br />

exhausted. The fact that under such conditions of food shortage<br />

any given amoeba can be the first to emit cyclic AMP and<br />

thus become an attractor center corresponds to the random<br />

behavior of fluctuations. This fluctuation is then amplified and<br />

<strong>org</strong>anizes the medium.


ORDER OUT OF CHAOS<br />

160<br />

Bifurcations and Symmetry-Breaking<br />

Let us take a closer look at the emergence of self-<strong>org</strong>anization<br />

and the processes that occur when we go beyond this threshold.<br />

At equilibrium or near-equilibrium, there is only one<br />

steady state that will depend on the values of some control<br />

parameters. We shall call A the control parameter, which, for<br />

example, may be the concentration of substance B in the<br />

Brusselator described in section 4. We now follow the change<br />

in the state of the system as the value of B increases. In this<br />

way the system is pushed farther and farther away from equilibrium.<br />

At some point we reach the threshold of the stability<br />

of the "thermodynamic branch." Then we reach what is generally<br />

called a "bifurcation point." (These are the points whose<br />

role Maxwell emphasized in his thoughts on the relation between<br />

determinism and free choice [see Chapter II, section<br />

3].)<br />

X<br />

,.<br />

,<br />

,<br />

I<br />

I<br />

\<br />

\<br />

\<br />

\<br />

A B\ E<br />

t-------=--=-...·--<br />

Figure 10. Bifurcation diagram. The diagram plots the steady-state values<br />

of X as function of a bifurcation parameter . Continuous lines are stable<br />

stationary states; broken lines are unstable stationary states. The only way<br />

to get to branch D is to start with some concentration X0 higher than the<br />

value of X corresponding to branch E.


161 THE THREE STAGES OF THERMODYNAMICS<br />

Let us consider some typical bifurcation diagrams. At bifurcation<br />

point B, the thermodynamic branch becomes unstable<br />

in respect to fluctuations. For the value Ac of the control parameter<br />

A, the system may be in three different steady states:<br />

C, E, D. 1\vo of these states are stable, one unstable. It is very<br />

important to emphasize that the behavior of such systems depends<br />

on their history. Suppose we slowly increase the value<br />

of the control parameter A; we are likely to follow the path A,<br />

B, C in Figure 10. On the contrary, if we start with a large<br />

value of the concentration X and maintain the value of the control<br />

parameter constant, we are likely to come to point D. The<br />

state we reach depends on the previous history of the system.<br />

Until now history has been commonly used in the interpretation<br />

of biological and social phenomena, but that it may play<br />

an important role in simple chemical processes is quite unexpected.<br />

Consider the bifurcation diagram represented in Figure 11.<br />

This differs from the previous diagram in that at the bifurcation<br />

point two new stable solutions emerge. Thus a new question:<br />

Where will the system go when we reach the bifurcation<br />

point? We have here a "choice" between two possibilities;<br />

X<br />

Figure 11. Symmetrical bifurcation diagram. X is plotted as a function of A.<br />

For AAc there<br />

are two stable stationary states for each value of A (the formerly stable state<br />

becomes unstable).


ORDER OUT OF CHAOS 162<br />

they may represent either of the two nonuniform distributions<br />

of chemical X in space, as represented in Figures 12 and 13.<br />

X<br />

----r<br />

X<br />

---- r<br />

Figures 12 and 13. Two possible spatial distributions of the chemical component<br />

X, corresponding to each of the two branches in Figure 11. Figure 12<br />

corresponds to a "left" structure as component X has a higher concentration<br />

in the left part; similarly, Figure 13 corresponds to a "right" structure.<br />

The two structures are mirror images of one another. In Figure<br />

12 the concentration of X is larger at the left; in Figure 13 it is<br />

larger at the right. How will the system choose between left<br />

and right? There is an irreducible random element; the macroscopic<br />

equation cannot predict the path the system will<br />

take. Turning to a microscopic description will not help. There<br />

is also no distinction between left and right. We are faced with<br />

chance events very similar to the fall of dice.


163<br />

THE THREE STAGES OF THERMODYNAMICS<br />

We would expect that if we repeat the experiment many<br />

times and lead the system beyond the bifurcation point, half of<br />

the system will go into the left configuration, half into the<br />

right. Here another interesting question arises: In the world<br />

around us, some basic simple symmetries seem to be broken.to<br />

Everybody has obser ved that shells often have a preferential<br />

chirality. Pasteur went so far as to see in dissymmetry, in<br />

the breaking of symmetry, the very characteristic of life. We<br />

know today that DNA, the most basic nucleic acid, takes the<br />

form of a left-handed helix. How did this dissymmetry arise?<br />

One common answer is that it comes from a unique event that<br />

has by chance favored one of the two possible outcomes; then<br />

an autocatalytic process sets in, and the left-handed structure<br />

produces other left-handed structures. Others imagine a<br />

"war" between left- and right-handed structures in which one<br />

of them has annihilated the other. These are problems for<br />

which we have not yet found a satisfactory answer. To speak of<br />

unique events is not satisfactory; we need a more "systematic"<br />

explanation.<br />

We have recently discovered a striking example of the fundamental<br />

new properties that matter acquires in far-fromequilibrium<br />

conditions: external fields, such as the gravitational<br />

field, can be "perceived" by the system, creating the possibility<br />

of pattern selection.<br />

How would an external field-a gravitational field-change<br />

an equilibrium situation? The answer is provided by Boltzmann's<br />

order principle: the basic quantity involved is the ratio<br />

of potential energy/thermal energy. This is a small quantity for<br />

the gravitational field of earth; we would have to climb a mountain<br />

to achieve an appreciable change in pressure or in the<br />

composition of the atmosphere. But recall the Benard cell;<br />

from a mechanical perspective, its instability is the raising of<br />

its center of gravity as the result of thermal dilatation. In other<br />

words, gravitation plays an essential role here and leads to a<br />

new structure in spite of the fact that the Benard cell may have<br />

a thickness of only a few millimeters. The effect of gravitation<br />

on such a thin layer would be negligible at equilibrium, but<br />

because of the nonequilibrium induced by the difference in<br />

temperature, macroscopic effects due to gravitation become<br />

visible even in this thin layer. Nonequilibrium magnifies the<br />

effect of gravitation.tl


. . .. .. . . .<br />

ORDER OUT OF CHAOS<br />

164<br />

Gravitation obviously will modify the diffusion flow in a reaction<br />

diffusion equation. Detailed calculations show that this<br />

can be quite dramatic near a bifurcation point of an unperturbed<br />

system. In particular, we can conclude that very small<br />

gravitational fields can lead to pattern selection.<br />

Let us again consider a system with a bifurcation diagram<br />

such as represented in Figure 11. Suppose that for no gravitation,<br />

g=O, we have, as in Figures 12 and 13, an asymmetric<br />

"up/down" pattern as well as its mirror image, "down/up."<br />

Both are equally probable, but when g is taken into account,<br />

the bifurcation equations are modified because the diffusion<br />

flow contains a term proportional to g. As a result, we now<br />

obtain the bifurcation diagram represented in Figure 14. The<br />

original bifurcation has disappeared-this is true whatever the<br />

value of the field. One structure (a) now emerges continuously<br />

as the bifurcation parameter grows, while the other (b) can be<br />

attained only through a finite perturbation.<br />

x<br />

. .. ... .<br />

.<br />

.<br />

.<br />

.<br />

.<br />

<br />

.•. .<br />

.<br />

.<br />

...<br />

'.<br />

" ,<br />

.------<br />

• •••••••<br />

b)<br />

Figure 14. Phenomenon of assisted bifurcation in the presence of an external<br />

field. X is plotted as a function of parameter ". The symmetrical bifurcation<br />

that would occur in the absence of the field is indicated by the dotted<br />

line. The bifurcation value is "0; the stable branch (b) is at finite di'stance from<br />

branch (a).


165 THE THREE STAGES OF THERMODYNAMICS<br />

Therefore, if we follow the path (a), we expect the system to<br />

follow the continuous path. This expectation is correct as long<br />

as the distance s between the two branches remains large in respect<br />

to thermal fluctuations in the concentration of X. There<br />

occurs what we would like to call an "assisted" bifurcation.<br />

As before, at about the value Ac a self-<strong>org</strong>anization process<br />

may occur. But now one of the two possible patterns is preferred<br />

and will be selected.<br />

The important point is that, depending on the chemical process<br />

responsible for the bifurcation, this mechanism expresses<br />

an extraordinary sensitivity. Matter, as we mentioned earlier<br />

in this chapter, perceives differences that would be insignificant<br />

at equilibrium. Such possibilities lead us to think of the<br />

simplest <strong>org</strong>anisms, such as bacteria, which we know are able<br />

to react to electric or magnetic fields. More generally they<br />

show that far-from-equilibrium chemistry leads to possible<br />

"adaptation" of chemical processes to outside conditions. This<br />

contrasts strongly with equilibrium situations, in which large<br />

perturbations or modifications of the boundary conditions are<br />

necessary to determine a shift for one structure to another.<br />

The sensitivity of far-from-equilibrium states to external fluctuations<br />

is another example of a system's spontaneous "adaptative<br />

<strong>org</strong>anization" to its environment. Let us give an example12<br />

of self-<strong>org</strong>anization as a function of fluctuating external conditions.<br />

The simplest conceivable chemical reaction is the isomerization<br />

reaction where AB. In our model the product A can<br />

also enter into another reaction: A+ light-+ A *-+A+ heat. A<br />

absorbs light and gives it back as heat while leaving its excited<br />

state A*. Consider these two processes as taking place in a<br />

closed system: only light and heat can be exchanged with the<br />

outside. Nonlinearity exists in the system because the transformation<br />

from B to A absorbs heat: the higher the temperature,<br />

the faster the formation of A. But also the higher the concentration<br />

of A, the higher the absorption of light by A and its transformation<br />

into heat, and the higher the temperature. A<br />

catalyzes its own formation.<br />

We expect to find that the concentration of A corresponding<br />

to the stationary state increases with the light intensity. This is<br />

indeed the case. But starting from a critical point, there appears<br />

one of the standard far-from-equilibrium phenomena:<br />

the coexistence of multiple stationary states. For the same val-


ORDER OUT OF CHAOS<br />

166<br />

ues of light intensity and temperature, the system can be<br />

found in two different stable stationary states with different<br />

concentrations of A. A third, unstable state marks the threshold<br />

between the first two. Such a coexistence of stationary<br />

states gives birth to the well-known phenomenon of hysteresis.<br />

But this is not the whole story. If the light intensity, instead<br />

of being constant, is taken as randomly fluctuating, the<br />

situation is altered profoundly. The zone of coexistence between<br />

the two stationary states increases, and for certain values<br />

of the parameters coexistence among three stationary<br />

stable states becomes possible.<br />

In such a case, a random fluctuation in the external flux,<br />

often termed "noise," far from being a nuisance, produces<br />

new types of behavior, which would imply, under deterministic<br />

fluxes, much more complex reaction schemes. It is important<br />

to remember that random noise in the fluxes may be consid-<br />

X<br />

p<br />

1\ ..<br />

•<br />

'<br />

•<br />

I<br />

<br />

I ' ' I<br />

J ' 'I<br />

v<br />

•<br />

p':<br />

Q<br />

b1 b2 b<br />

Figure 15. This figure shows how a "hysteresis" phenomenon occurs if we<br />

have the value of the bifurcation parameter b first growing and then diminishing.<br />

If the system is initially in a stationary state belonging to the lower<br />

branch, it will stay there while b grows. But at b = b2, there will be a discontinuity:<br />

The system jumps from Q to Q', on the higher branch. Inversely,<br />

starting from a state on the higher branch, the system will remain there till<br />

b=b1, when it will jump down toP'. Such types of bistable behavior are<br />

observed in many fields, such as lasers, chemical reactions or biological<br />

membranes.


167 THE THREE STAGES OF THERMODYNAMICS<br />

ered as unavoidable in any "natural system." For example, in<br />

biological or ecological systems the parameters defining interaction<br />

with the environment cannot generally be considered as<br />

constants. Both the cell and the ecological niche draw their<br />

sustenance from their environment ; and humidity, pH, salt<br />

concentration, light, and nutrients form a permanently fluctuating<br />

environment. The sensitivity of nonequilibrium states,<br />

not only to fluctuations produced by their internal activity but<br />

also to those coming from their environment, suggests new<br />

perspectives for biological inquiry.<br />

Cascading Bifurcations<br />

and the Transitions to Chaos<br />

The preceding paragraph dealt only with the first bifurcation<br />

or, as mathematicians put it, the primary bifurcation, which<br />

occurs when we push a system beyond the threshold of stability.<br />

Far from exhausting the new solutions that may appear,<br />

this primary bifurcation introduces only a single characteristic<br />

time (the period of the limit cycle) or a single characteristic<br />

length. To generate the complex spatial temporal activity observed<br />

in chemical or biological systems, we have to follow the<br />

bifurcation diagram farther.<br />

We have already alluded to phenomena arising from the<br />

complex interplay of a multitude of frequences in hydrodynamical<br />

or chemical systems. Let us consider Benard structures,<br />

which appear at a critical distance from equilibrium.<br />

Farther away from thermal equilibrium the convection flow<br />

begins to oscillate in time; as the distance from equilibrium is<br />

increased still farther, more and more oscillation frequencies<br />

appear, and eventually the transition to equilibrium is complete.13<br />

The interplays among the frequencies produce possibilities<br />

of large fluctuations; the "region" in the bifurcation<br />

diagram defined by such values of the parameters is often<br />

called "chaotic." In cases such as the Benard instability, order<br />

or coherence is sandwiched between thermal chaos and nonequilibrium<br />

turbulent chaos. Indeed, if we continue to in­<br />

rease the gradient of temperature, the onvcction patterns<br />

become more complex; oscillations set in, and the ordered as-


ORDER OUT OF CHAOS ' 168<br />

TRACES OF Br- CONCENTRATION<br />

Homogeneous Steady State<br />

f\/l.J\NVVVV\ Sinusoidal Oscillations<br />

•<br />

•<br />

•<br />

Complex Periodic States<br />

(Subharmonic bifurcation)<br />

2:<br />

o<br />

0:<br />

u..<br />

&£J<br />

U<br />

Z<br />

<br />

!:!!<br />

c<br />

Chaos<br />

Mixed - Mode Oscillations<br />

Chaotic<br />

and<br />

Periodic<br />

Relaxation Oscillations<br />

TIME<br />

Figure 16. Temporal oscillations of the ion Br- in the Belousov­<br />

Zhabotinski reaction. The figure represents a succession of regions corresponding<br />

to qualitative differences. This is a schematic representation. The<br />

experimental data indicate the existence of much more complicated sequences.<br />

pect of the convection is largely destroyed. However, we<br />

should not confuse "equilibrium thermal chaos" and "nonequilibrium<br />

turbulent chaos." In thermal chaos as realized in<br />

equilibrium, all characteristic space and time scales are of molecular<br />

range, while in turbulent chaos we have such an abundance<br />

of macroscopic time and length scales that the system<br />

appears chaotic. In chemistry the relation between order and<br />

chaos appears highly complex: successive regimes of ordered<br />

(oscillatory) situations follow regimes of chaotic .behavior.<br />

This has, for instance, been observed as a function of the flow<br />

rate in the Belousov-Zhabotinsky reaction.


169 THE. THREE STAGES OF THERMODYNAMICS<br />

In many cases it is difficult to disentangle the meaning of<br />

words such as "order" and "chaos." Is a tropical forest an<br />

ordered or a chaotic system? The history of any particular<br />

animal species will appear very contingent, dependent on<br />

other species and on environmental accidents. Nevertheless,<br />

the feeling persists that, as such, the overall pattern of a tropical<br />

forest; as represented, for instance, by the diversity of species,<br />

corresponds to the very archetype of order. Whatever the<br />

precise meaning we will eventually give to this terminology, it<br />

is clear that in some cases the succession of bifurcations forms<br />

an irreversible evolution where the determinism of characteristic<br />

frequencies produces an increasing randomness stemming<br />

from the multiplicity of those frequencies.<br />

A remarkably simple road to "chaos" that has already attracted<br />

a lot of attention is the "Feigenbaum sequence." It<br />

concerns any system whose behavior is characterized by a<br />

very general feature-that is, for a determined range of parameter<br />

values the system's behavior is periodic, with a period T;<br />

beyond this range, the period becomes 2T, and beyond yet another<br />

critical threshold, the system needs 4 Tin order to repeat<br />

itself. The system is thus characterized by a succession of bifurcations,<br />

with successive period doubling. This constitutes a<br />

typical route going from simple periodic behavior to the complex<br />

aperiodic behavior occurring when the period has doubled<br />

ad infinitum. This route, as Feigenbaum discovered, is<br />

characterized by universal numerical features independent of<br />

the mechanism involved as long as the system possesses the<br />

qualitative property of period doubling. "In fact, most measurable<br />

properties of any such system in this aperiodic limit<br />

now can be determined in a way that essentially bypasses the<br />

details of the equations governing each specific system . . . . "14<br />

In other cases, such as those represented in Figure 16, both<br />

deterministic and stochastic elements characterize the history<br />

of the system.<br />

If we consider Figure 17 and a value of the control parameter<br />

of the order of X 6 , we see that the system already has a<br />

wealth of possible stable and unstable behaviors. The "historical"<br />

path along which the system evolves as the control parameter<br />

grows is characterized by a succession of stable regions,<br />

where deterministic laws dominate, and of instable ones, near<br />

the bifurcation points, where the system can "choose" be-


ORDER OUT OF CHAOS 170<br />

tween or among more than one possible future. Both the deterministic<br />

character of the kinetic equations whereby the set of<br />

possible states and their respective stability can be calculated,<br />

and the random fluctuations "choosing" between or among<br />

the states around bifurcation points are inextricably connected.<br />

This mixture of necessity and chance constitutes the<br />

history of the system.<br />

Solutions<br />

\ (c'),,<br />

' ,,<br />

' ,<br />

'<br />

I<br />

I<br />

,<br />

(/ ..<br />

, -<br />

-<br />

-<br />

t---


171 THE THREE STAGES OF THERMODYNAMICS<br />

From Euclid to Aristotle<br />

One of the most interesting aspects of dissipative structures is<br />

their coherence. The system behaves as a whole, as if it were<br />

the site of long-range forces. In spite of the fact that interactions<br />

among molecules do not exceed a range of some I0-8<br />

em, the system is structured as though each molecule were<br />

"informed" about the overall state of the system.<br />

It has often been said-and we have already repeated itthat<br />

modern science was born when Aristotelian space, for<br />

which one source of inspiration was the <strong>org</strong>anization and solidarity<br />

of biological functions, was replaced by the homogeneous<br />

and isotropic space of Euclid. However, the theory of<br />

dissipative structures moves us closer to Aristotle's conception.<br />

Whether we are dealing with a chemical clock, concentration<br />

waves, or the inhomogeneous distribution of chemical<br />

products, instability has the effect of breaking symmetry, both<br />

temporal and spatial. In a limit cycle, no two instants are<br />

equivalent; the chemical reaction acquires a phase similar to<br />

that characterizing a light wave, for example. Again, when a<br />

favored direction results from an instability, space ceases to be<br />

isotropic. We move from Euclidian to Aristotelian space!<br />

It is tempting to speculate that the breaking of space and<br />

time symmetry plays an important part in the fascinating phenomena<br />

of morphogenesis. These phenomena have often led<br />

to the conviction that some internal purpose must be involved,<br />

a plan realized by the embryo when its growth is complete. At<br />

the beginning of this century, German embryologist Hans<br />

Driesch believed that some immaterial "entelechy" was responsible<br />

for the embryo's development. He had discovered<br />

that the embryo at an early stage was capable of withstanding<br />

the severest perturbations and, in spite of them, of developing<br />

into a normal, functional <strong>org</strong>anism. On the other hand, when<br />

we observe embryological development on film, we "see"<br />

jumps corresponding to radical re<strong>org</strong>anizations followed by<br />

periods of more "pacific" quantitative growth. There are, fortunately,<br />

few mistakes. The jumps are performed in a reproducible<br />

fashion. We might speculate that the basic<br />

mechanism of evolution is based on the play between bifurca-


ORDER OUT OF CHAOS 172<br />

tions as mechanisms of exploration and the selection of chemical<br />

interactions stabilizing a particular trajectory. Some forty<br />

years ago, the biologist Waddington introduced such an idea.<br />

The concept of "chreod" that he introduced to describe the<br />

stabilized paths of development would correspond to possible<br />

lines of development produced as a result of the double imperative<br />

of flexibility and security. 15 Obviously the problem is<br />

very complex and can be dealt with only briefly here.<br />

Many years ago embryologists introduced the concept of a<br />

morphogenetic field and put forward the hypothesis that the<br />

differentiation of a cell depends on its position in that field.<br />

But how does a cell "recognize" its position? One idea that is<br />

often debated is that of a "gradient" of a characteristic substance,<br />

of one or more "morphogens." Such gradients could<br />

actually be produced by symmetry-breaking instabilities in<br />

far-from-equilibrium conditions. Once it has been produced, a<br />

chemical gradient can provide each cell with a different chemical<br />

environment and thus induce each of them to synthesize a<br />

specific set of proteins. This model, which is now widely used,<br />

seems to be in agreement with experimental evidence. In particular,<br />

we may refer to Kauffman's work16 on drosophila. A<br />

reaction-diffusion system is taken as responsible for the commitment<br />

to alternative development programs that appear to<br />

occur in different groups of cells in the early embryo. Each<br />

3 1 3<br />

4 r--------r------------ 4<br />

Figure 18. Schematic representation of the structure of the drosophila embryo<br />

as it results from successive binary choices. See text for more detail.


173 THE THREE STAGES OF THERMODYNAMICS<br />

compartment would be specified by a unique combination of<br />

binary choices, each of these choices being the result of a spatial<br />

symmetry-breaking bifurcation. The model leads to successful<br />

predictions about the result of transplantations as a<br />

function of the "distance" between the original and final regions-that<br />

is, of the number of differences among the states<br />

of the binary choices or "switches" that specify each of them.<br />

Such ideas and models are especially important in biological<br />

systems where the embryo begins to develop in an apparently<br />

symmetrical state (for example, Fucus, Acetabularia).<br />

We may ask if the embryo is really homogeneous at the beginning.<br />

And even if small inhomogeneities are present in the initial<br />

environment, do they cause or channel evolution toward a<br />

given structure? Precise answers to such questions are not<br />

available at present. However, one thing seems established:<br />

the instability connected with chemical reactions and transport<br />

appears as the only general mechanism capable of breaking<br />

the symmetry of an initially homogeneous situation.<br />

The very possibility of such a solution takes us far beyond<br />

the age-old conflict between reductionists and antireductionists.<br />

Ever since Aristotle (and we have cited Stahl, Hegel,<br />

Bergson, and other antireductionists), the same conviction<br />

has been expressed: a concept of complex <strong>org</strong>anization is required<br />

to connect the various levels of description and account<br />

for the relationship between the whole and the behavior of the<br />

parts. In answer to the reductionists, for whom the sole<br />

"cause" or <strong>org</strong>anization can lie only in the part, Aristotle with<br />

his formal cause, Hegel with his emergence of Spirit in Nature,<br />

and Bergson with his simple, irrepressible, <strong>org</strong>anizationcreating<br />

act, assert that the whole is predominant. To cite<br />

Bergson,<br />

In general, when the same object appears in one aspect<br />

as simple and in another as infinitely complex, the two<br />

aspects have by no means the same importance, or rather<br />

the same degree of reality. In such cases, the simplicity<br />

belongs to the object itself, and the infinite complexity to<br />

the views we take in turning around it, to the symbols by<br />

which our senses or intellect represent it to us, or, more<br />

gc::nc::rally, to c::lc::mc::nts of a different order, with which we:<br />

try to imitate it artificially, but with which it remains in-


ORDER OUT OF CHAOS 174<br />

commensurable, being of a different nature. An artist of<br />

genius has painted a figure on his canvas. We can imitate<br />

his picture with many-coloured squares of mosaic. And<br />

we shall reproduce the curves and shades of the model so<br />

much the better as our squares are smaller, more numerous<br />

and more varied in tone. But an infinity of elements<br />

infinitely small, presenting an infinity of shades, would<br />

be necessary to obtain the exact equivalent of the figure<br />

that the artist has conceived as a simple thing, which he<br />

has wished to transport as a whole to the canvas, and<br />

which is the more complete the more it strikes us as the<br />

projection of an indivisible intuition. 17<br />

In biology, the conflict between reductionists and antireductionists<br />

has often appeared as a conflict between the assertion<br />

of an external and an internal purpose. The idea of an immanent<br />

<strong>org</strong>anizing intelligence is thus often opposed by an <strong>org</strong>anizational<br />

model borrowed from the technology of the time<br />

(mechanical, heat, cybernetic machines), which immediately<br />

elicits the retort: "Who" built the machine, the automaton<br />

that obeys external purpose?<br />

As Bergson emphasized at the beginning of this century,<br />

both the technological model and the vitalist idea of an internal<br />

<strong>org</strong>anizing power are expressions of an inability to conceive<br />

evolutive <strong>org</strong>anization without immediately referring it<br />

to some preexisting goal. Today, in spite of the spectacular success<br />

of molecular biology, the conceptual situation remains<br />

about the same: Bergson's argument could be applied to contemporary<br />

metaphors such as "<strong>org</strong>anizer," "regulator," and<br />

"genetic program." Unorthodox biologists such as Paul Weiss<br />

and Conrad Waddington 18 have rightly criticized the way this<br />

kind of qualification attributes to individual molecules the<br />

power to produce the global order biology aims to understand,<br />

and, by so doing, mistakes the formulation of the problem for<br />

its solution.<br />

It must be recognized that technological analogies in biology<br />

are not without interest. However, the general validity of<br />

such analogies would imply that, as in an electronic circuit, for<br />

example, there is a basic homogeneity between the description<br />

of molecular interaction and that of global behavior: The functioning<br />

of a circuit may be deduced from the nature and posi-


175 THE THREE STAGES OF THERMODYNAMICS<br />

tion of its relays; both refer to the same scale, since the relays<br />

were designed and installed by the same engineer who built<br />

the whole machine. This cannot be the rule in biology.<br />

It is true that when we come to a biological system such as<br />

the bacterial chemotaxis, it is hard not to speak of a molecular<br />

machine consisting of receptors, sensory and regulatory processing<br />

systems, and motor response. We know of approximately<br />

twenty or thirty receptors that can detect highly<br />

specific classes of compounds and make a bacterium swim up<br />

spatial gradients of attractants or down gradients of repellents.<br />

This "behavior" is determined by the output of the processing<br />

system-that is, the switching on or off of a tumble that generates<br />

a change in the bacterium's direction.l9<br />

But such cases, fascinating as they are, do not tell the whole<br />

story. In fact it is tempting to see them as limiting cases, as the<br />

end products of a specific kind of selective evolution, emphasizing<br />

stability and reproducible behavior against openness<br />

and adaptability. In such a perspective, the relevance of the<br />

technological metaphor is not a matter of principle but of opportunity.<br />

The problem of biological order involves the transition from<br />

the molecular activity to the supermolecular order of the cell.<br />

This problem is far from being solved.<br />

Often biological order is simply presented as an improbable<br />

physical state created and maintained by enzymes resembling<br />

Maxwell's demon, enzymes that maintain chemical differences<br />

in the system in the same way as the demon maintains<br />

temperature and pressure differences. If we accept this, biology<br />

would be in the position described by Stahl. The laws of<br />

nature allow only death. Stahl's notion of the <strong>org</strong>anizing action<br />

of the soul is replaced by the genetic information contained in<br />

the nucleic acids and expressed in the formation of enzymes<br />

that permit life to be perpetuated. Enzymes postpone death<br />

and the disappearance of life.<br />

In the context of the physics of irreversible processes, the<br />

results of biology obviously have a different meaning and different<br />

implications. We know today that both the biosphere as<br />

a whole as well as its components, living or dead, exist in farfrom-equilibrium<br />

conditions. In this context life, far from<br />

being outside the natural order, appears as the supreme expression<br />

of the self-<strong>org</strong>anizing processes that occur.


ORDER OUT OF CHAOS<br />

176<br />

We are tempted to go so far as to say that once the condi·<br />

tions for self-<strong>org</strong>anization are satisfied, life becomes as predictable<br />

as the Benard instability or a falling stone. It is a<br />

. remarkable fact that recently discovered fossil forms of life<br />

appear nearly simultaneously with the first rock formations<br />

(the oldest microfossils known today date back 3.8 . 109 years,<br />

while the age of the earth is supposed to be 4. 6,109 years; the<br />

formation of the first rocks is also dated back to 3.8 . 109 years).<br />

The early appearance of life is certainly an argument in favor of<br />

the idea that life is the result of spontaneous seif-<strong>org</strong>anization<br />

that occurs whenever conditions for it permit. However, we<br />

must admit that we remain far from any quantitative theory.<br />

To return to our understanding of life and evolution, we are<br />

now in a better position to avoid the risks implied by any de·<br />

nunciation of reductionism. A system far from equilibrium may<br />

be described as <strong>org</strong>anized not because it realizes a plan alien<br />

to elementary activities, or transcending them, but, on the<br />

contrary, because the amplification of a microscopic fluctuation<br />

occurring at the "right moment" resulted in favoring one reaction<br />

path over a number of other equally possible paths. Under cer·<br />

tain circumstances, therefore, the role played by individual behavior<br />

can be decisive. More generally, the "overall" behavior<br />

cannot in general be taken as dominating in any way the elementary<br />

processes constituting it. Self-<strong>org</strong>anization processes<br />

in far-from-equilibrium conditions correspond to a delicate interplay<br />

between chance and necessity, between fluctuations<br />

and deterministic laws. We expect that near a bifurcation, fluctuations<br />

or random elements would play an important role,<br />

while between bifurcations the deterministic aspects would<br />

become dominant. These are the questions we now need to<br />

investigate in more detail.


CHAPTER VI<br />

ORDER THROUGH<br />

FWCTUATIONS<br />

Fluctuations and Chemistry<br />

In our Introduction we noted that a reconceptualization of the<br />

physical sciences is occurring today. They are moving from<br />

deterministic, reversible processes to stochastic and irreversible<br />

ones. This change of perspective affects chemistry in a<br />

striking way. As we have seen in Chapter V, chemical processes,<br />

in contrast to the trajectories of classical dynamics,<br />

correspond to irreversible processes. Chemical reactions lead<br />

to entropy production. On the other hand, classical chemistry<br />

continues to rely on a deterministic description of chemical<br />

evolution. As we have seen in Chapter V, it is necessary to<br />

produce differential equations involving the concentration of<br />

the various chemical components. Once we know these concentrations<br />

at some initial time (as well as at appropriate<br />

boundary conditions when space-dependent phenomena such as<br />

diffusion are involved), we may calculate what the concenK-ation<br />

will be at a later time. It is interesting to note that the deterministic<br />

view of chemistry fails when far-from-equilibrium<br />

processes are involved.<br />

We have repeatedly emphasized the role of fluctuations. Let<br />

us summarize here some of the more striking features. Whenever<br />

we reach a bifurcation point, deterministic description<br />

breaks down. The type of fluctuation present in the system<br />

will lead to the choice of the branch it will follow. Crossing a<br />

bifurcation is a stochastic process, such as the tossing of a<br />

coin. Chemical chaos provides another example (see Chapter<br />

V). Here we can no longer follow an individual chemical trajectory.<br />

We cannot predict the details of temporal evolution.<br />

177


ORDER OUT OF CHAOS 178<br />

Once again, only a statistical description is possible. The existence<br />

of an instability may be viewed as the result of a fluctuation<br />

that is first localized in a small part of the system and then<br />

spreads and leads to a new macroscopic state.<br />

This situation alters the traditional view of the relation between<br />

the microscopic level as described by molecules or<br />

atoms and the macroscopic level described in ter ms of global<br />

variables such as concentration. In many situations fluctuations<br />

correspond only to small corrections. As an example, let<br />

us take a gas composed of N molecules enclosed in a vessel of<br />

volume V. Let us divide this volume into two equal parts.<br />

What is the number of particles X in one of these two parts?<br />

Here the variable X is a "random" variable, and we would<br />

expect it to have a value in the neighborhood of N/2.<br />

A basic theorem in probability theory, the law of large numbers,<br />

provides an estimate of the "error" due to fluctuations.<br />

In essence, it states that if we measure X we have to expect a<br />

value of the order N/2±v'Nii. If N is large, the difference<br />

introduced by fluctuations v'Nfi may also be large (if<br />

N= 1Q24, VN= 1012); however, the relative error introduced<br />

by fluctuations is of the order of (v'N!i)/(N/2) or llYN and<br />

thus tends toward zero for a sufficiently large value of N. As<br />

soon as the system becomes large enough, the law of large<br />

numbers enables us to make a clear distinction between mean<br />

values and fluctuations, and the latter may be neglected.<br />

Hewever, in nonequilibrium processes we may find just the<br />

opposite situation. Fluctuations determine the global outcome.<br />

We could say that instead of being corrections in the<br />

average values, fluctuations now modify those averages. This<br />

is a new situation. For this reason we would like to introduce a<br />

neologism and call situations resulting from fluctuation "order<br />

through fluctuation." Before giving examples, let us make<br />

some general remarks to illustl·ate the conceptual novelty of<br />

this situation.<br />

Readers may be familiar with the Heisenberg uncertainty<br />

relations, which express in a striking way the probabilistic aspects<br />

of quantum theory. Since we can no longer simultaneously<br />

measure position and coordinates in quantum theory,<br />

classical determinism is breaking down. This was believed to<br />

be of no importance for the description of macroscopic objects


179 ORDER THROUGH FLUCTUATIONS<br />

such as living systems. But the role of fluctuations in nonequilibrium<br />

systems shows that this is not the case. Randomness<br />

remains essential on the macroscopic level as well. It is interesting<br />

to note another analogy with quantum theory, which<br />

assigns a wave behavior to all elementary particles. As we<br />

have seen, chemical systems far from equilibrium may also<br />

lead to coherent wave behavior: these are the chemical clocks<br />

discussed in Chapter V. Once again, some of the properties<br />

quantum mechanics discovered on the microscopic level now<br />

appear on the macroscopic level.<br />

Chemistry is actively involved in the reconceptualization of<br />

science. I We are probably only at the beginning of new directions<br />

of research. It may well be, as some recent calculations<br />

suggest, that the idea of reaction rate has to be replaced in<br />

some cases by a statistical theory involving a distribution of<br />

reaction probabilities.2<br />

Fluctuations and Correlations<br />

Let us go back to the types of chemical reaction discussed in<br />

Chapter V. To take a specific example, consider a chain of<br />

reactions such as APXPF. The kinetic equations in Chapter V<br />

refer to the average concentrations. To emphasize this we shall<br />

now write instead of X. We can then ask what is the<br />

probability at a given time of finding a number X for the concentration<br />

of this component. Obviously this probability will<br />

fluctuate, as do the number of collisions among the various<br />

molecules involved. It is easy to write an equation that describes<br />

the change in this probability distribution P(X,t) as a<br />

result of processes that produce molecule X and of processes<br />

that destroy that molecule. We may perform the calculation for<br />

equilibrium systems or for steady-state systems. Let us first<br />

mention the results obtained for equilibrium systems.<br />

At equilibrium we virtually recover a classical probabilistic<br />

distribution, the Poisson distribution, which is described in<br />

every textbook on probabilities, since it is valid in a variety of<br />

situations, such as the distribution of telephone calls, waiting<br />

times in restaurants, or the fluctuation of the concentration of


ORDER OUT OF CHAOS 180<br />

particles in a gas or a liquid. The mathematical form of this<br />

distribution is of no importance here. We merely want to emphasize<br />

two of its aspects. First, it leads to the law of large<br />

numbers as formulated in the first section of this chapter. Thus<br />

fluctuations indeed become negligible in a large system. Moreover,<br />

this law enables us to calculate the correlation between<br />

the number of particles X at two different points in space separated<br />

by some distance r. The calculation demonstrates that at<br />

equilibrium there is no such correlation. The probability of<br />

finding two molecules X and X' at two different points rand r'<br />

is the product of finding X at rand X' at r' (we cons.ider distances<br />

that are large in respect to the range of intermolecular<br />

forces).<br />

One of the most unexpected results of recent research is that<br />

this situation changes drastically when we move to nonequilibrium<br />

situations. First, when we come close to bifurcation<br />

points the fluctuations become abnormally high and the law of<br />

large numbers is violated. This is to be expected, since the<br />

system may then "choose" among various regimes. Fluctuations<br />

can even reach the same order of magnitude as the mean<br />

macroscopic values. Then the distinction between fluctuations<br />

and mean values breaks down. Moreover, in the case of a nonlinear<br />

type of chemical reaction discussed in Chapter V, longrange<br />

correlations appear. Particles separated by macroscopic<br />

distances become linked. Local events have repercussions<br />

throughout the whole system. It is interesting to note3 that<br />

such long-range correlations appear at the precise point of<br />

transition from equilibrium to nonequilibrium. From this point<br />

of view the transition resembles a phase transition. However,<br />

the amplitudes of these long-range correlations are at first<br />

small but increase with distance from equilibrium and may become<br />

infinite at the bifurcation points.<br />

We believe that this type of behavior is quite interesting,<br />

since it gives a molecular basis to the problem of communication<br />

mentioned in our discussion of the chemical clock. Even<br />

before the macroscopic bifurcation, the system is <strong>org</strong>anized<br />

through these long-range correlations. We come back to one of<br />

the main ideas of this book: nonequilibrium as a source of order.<br />

Here the situation is especially clear. At equilibrium molecules<br />

behave as essentially independent entities; they· ignore<br />

one another. We would like to call them "hypnons," "sleep-


181 ORDER THROUGH FLUCTUATIONS<br />

walkers." Though each of them may be as complex as we like,<br />

they ignore one another. However, nonequilibrium wakes them<br />

up and introduces a coherence quite foreign to equilibrium.<br />

The microscopic theory of irreversible processes that we shall<br />

develop in Chapter IX will present a similar picture of matter.<br />

Matter's activity is related to the nonequilibrium conditions<br />

that it itself may generate. Just as in macroscopic behavior, the<br />

laws of fluctuations and correlations are universal at equilibrium<br />

(when we find the Poisson type of distribution); they<br />

become highly specific depending on the type of nonlinearity<br />

involved when we cross the boundary between equilibrium<br />

and nonequilibrium.<br />

The Amplification of Fluctuations<br />

Let us first take two examples wherein the growth of a fluctuation<br />

preceding the formation of a new structure can be followed<br />

in detail. The first is the aggregation of slime molds,<br />

which when threatened with starvation coalesce into a single<br />

supracellular mass. We have already mentioned this in Chapter<br />

V. Another illustration of the role of fluctuations is the first<br />

stage in the construction of a termites' nest. This was first<br />

described by Grasse, and Deneubourg has studied it from the<br />

standpoint that interests us here."'<br />

Self-Aggregation Process in an Insect Population<br />

Larvae of a coleoptera (Dendroctonus micans [Scot.]). are initially distributed<br />

at random between two horizontal sheets of glass, 2 mm apart. The<br />

borders are open and the surface is equal to 400 cm2.<br />

The aggregation process appears to result from the competition between<br />

two factors: the random moves of the larvae, and their reaction to a chemical<br />

product, a "pheromon" they synthetize from terpanes contained in the tree<br />

on which they feed and that each of them emits at a rate depending on its<br />

nutrition state. The pheromon diffuses in space, and the larvae move in the<br />

direction of its concentration gradient. Such a reaction provides an autocatalytic<br />

mechanism since, as they gather in a cluster, the larvae contribute<br />

to enhance the attractiveness of the corresponding region. The higher the<br />

local density of larvae in this region, the stronger the gradient and the more<br />

intense the tendency to move toward the crowded point.<br />

The experiment shows that the density of the larvae population determines<br />

not only the rate of the aggregation process but its effectiveness as<br />

well-that is, the number of larvae that will finally be part of the cluster. At


ORDER OUT OF CHAOS<br />

182<br />

high density (Figure A) a cluster appears and rapidly grows at the center of<br />

the experimental setup. At very low density (Figure B), no stable cluster appears.<br />

Moreover, other experiments have explored the possibility for a cluster to<br />

develop starting from a "nucleus" artificially created in a peripheral region of<br />

the system. Different solutions appear depending on the number of larvae in<br />

this initial nucleus.<br />

,.<br />

J<br />

(<br />

I'<br />

\<br />

t<br />

'<br />

I<br />

Figure A. Self-aggregation at high density. The times are 0 minutes and 21<br />

minutes.


..<br />

183 ORDER THROUGH FLUCTUATIONS<br />

If this number is small compared with the total number of larvae, the cluster<br />

fails to develop (Figure D). If it is large, the cluster grows (Figure E). For<br />

intermediate values of the initial nucleus, new types of structure may develop:<br />

Two, three or four other clusters appear and coexist, with a time of life<br />

at least greater than the time of observation (Figures F and G).<br />

No such multicluster structure was ever observed in experiments with homogeneous<br />

initial conditions. It would seem they correspond, in a bifurcation<br />

4)<br />

"'<br />

\1<br />

1<br />

"-<br />

.,..<br />

'<br />

I .,.<br />

,.<br />

"'<br />

<br />

"<br />

\<br />

..<br />

A<br />

...<br />

-<br />

(,l<br />

l<br />

•<br />

Figure B. Self-aggregation at low density. The times are 0 minutes and 22<br />

minutes.


ORDER OUT OF CH AOS<br />

184<br />

diagram, to stable states compatible with the value of the parameters<br />

characterizing the system but that cannot be attained by this system starting<br />

from homogeneous conditions. The nucleus would play the part of a finite<br />

perturbation necessary to excite the system and deport it in a region of the<br />

bifurcation diagram corresponding to such families of multicluster solutions .<br />

•<br />

Q<br />

100<br />

•<br />

80<br />

60<br />

40<br />

20<br />

10<br />

--f--- -- -<br />

,--..--- -<br />

-<br />

.. .... ..<br />

. . .<br />

-<br />

...... .. ....<br />

. ..... ..<br />

f<br />

--- r<br />

····]<br />

20 30 40<br />

. -- ..<br />

,-<br />

,-f<br />

.. · ·······t<br />

Figure C. Percent of the total number of larvae in the central cluster in<br />

function of time at three different densities.<br />

N<br />

CRITICAl NUCLEUS dKnc of e10 lot- nt ... nud


185 ORDER T;-iROUGH FWCTUATIONS<br />

N<br />

CR.tTiCAL NUCLEUS !l'owlh of• :or.-- inilinlnutk .. 1•··•1<br />

growth of a 30 la


ORDER OUT OF CHAOS<br />

186<br />

• > <br />

, , ..<br />

"<br />

.,<br />

')<br />

., '<br />

,. ,<br />

., .I I<br />

I<br />

..<br />

"<br />

... ...<br />

., "<br />

., \<br />

I<br />

...<br />

I<br />

"'<br />

.,<br />

'<br />

"<br />

, •<br />

.,<br />

;)<br />

")<br />

1<br />

\.<br />

'\<br />

., _,<br />

)<br />

...<br />

...<br />

'\<br />

I<br />

...<br />

<br />

I<br />

<br />

,'<br />

'' I •<br />

,J<br />

'<br />

...<br />

Figure G. Growth of a cluster (I) introduced peripherally, which induce the<br />

formation of a second little cluster (II).<br />

The construction of a termites' nest is one of those coherent<br />

activities that have led some scientists to speculate about a<br />

.. collective mind" in insect communities. But curiously, it appears<br />

that in fact the termites need very little information to<br />

participate in the construction of such a huge and complex<br />

edifice as the nest. The first stage in this activity, the construction<br />

of the base, has been shown by Grasse to be the<br />

result of what appears to be disordered behavior among termites.<br />

At this stage, they transport and drop lumps of earth in<br />

a random fashion, but in doing so they impregnate the lumps<br />

with a hormone that attracts other termites. The situation<br />

could thus be represented as follows: the initial "fluctuation"<br />

would be the slightly larger concentration of lumps of earth,<br />

which inevitably occurs at one time or another at some point<br />

in the area. The amplification of this event is produced by the


187 ORDER THROUGH FLUCTUATIONS<br />

increased density of termites in the region, attracted by the<br />

· slightly higher hormone concentration. As termites become<br />

more numerous in a region, the probability of their dropping<br />

lumps of earth there increases, leading in turn to a still higher<br />

concentration of the hormone. In this way "pillars" are<br />

formed, separated by a distance related to the range over<br />

which the hormone spreads. Similar examples have recently<br />

been described.<br />

Although Boltzmann's order principle enables us to describe<br />

chemical or biological processes in which differences<br />

are leveled out and initial conditions f<strong>org</strong>otten, it cannot explain<br />

situations such as these, where a few "decisions" in an<br />

unstable situation may channel a system formed by a large<br />

number of interactive entities toward a global structure.<br />

When a new structure results from a finite perturbation, the<br />

fluctuation that leads from one regime to the other cannot possibly<br />

overrun the initial state in a single move. It must first estab<br />

lish itself in a limited region and then invade the whole space:<br />

there is a nucleation mechanism. Depending on whether the<br />

size of the initial fluctuating region lies below or above some<br />

critical value (in the case of chemical dissipative structures,<br />

this threshold depends in particular on the kinetic constants<br />

and diffusion coefficients), the fluctuation either regresses or<br />

else spreads to the whole system. We are familiar with nucleation<br />

phenomena in the classical theory of phase change: in a<br />

gas, for example, condensation droplets incessantly form and<br />

evaporate. That temperature and pressure reach a point where<br />

the liquid state becomes stable means that a critical droplet<br />

size can be defined (which is smaller the lower the temperature<br />

and the higher the pressure). If the size of a droplet exceeds<br />

this "nucleation threshold," the gas almost instantaneously<br />

transforms into a liquid.<br />

Moreover, theoretical studies and numerical simulations<br />

show that the critical nucleus size increases with the efficacy<br />

of the diffusion mechanisms that link all the regions of systems.<br />

In other words, the faster communication takes place<br />

within a system, the greater the percentage of unsuccessful<br />

fluctuations and thus the more stable the system. This aspect<br />

of the critical-size problem means that in such situations the<br />

"outside world,·· the environment of the fluctuating region,<br />

always tends to damp fluctuations. These will be destroyed or


ORDER OUT OF CHAOS<br />

188<br />

(a)<br />

(b)<br />

Figure 19. Nucleation of a liquid droplet in a supersaturated vapor. (a)<br />

droplet smaller than the critical size; (b) droplet larger than the critical size.<br />

The existence of the threshold has been experimentally verified for dissipative<br />

structures.<br />

amplified according to the effectiveness of the communication<br />

between the fluctuating region and the outside world. The critical<br />

size is thus determined by the competition between the<br />

system's "integrative power" and the chemical mechanisms<br />

amplifying the fluctuation.<br />

This model applies to the results obtained recently in in<br />

vitro experimental studies of the onset of cancer tumors.s An<br />

individual tumor cell is seen as a "fluctuation," uncontrollably<br />

and permanently able to appear and to develop through replication.<br />

It is then confronted with the population of cytotoxic<br />

cells that either succeeds in destroying it or fails. Following<br />

the values of the different parameters characteristic of the replication<br />

and destruction processes, we can predict a regression<br />

or an amplification of the tumor. This kind of kinetic study has<br />

led to the recognition of unexpected features in the interaction<br />

between cytotoxic cells and the tumor. It seems that cytotoxic<br />

cells can confuse dead tumor cells with living ones. As a result,<br />

the destruction of the cancer cells becomes increasingly<br />

difficult.<br />

The question of the limits of complexity has often been<br />

raised. Indeed, the more complex a system is, the more numerous<br />

are the types of fluctuations that threaten its stability.<br />

How then, it has been asked, can systems as complex as eco<br />

logical or human <strong>org</strong>anizations possibly exist? How do they


189 ORDER THROUGH FLUCTUATIONS<br />

manage to avoid permanent chaos? The stabilizing effect of<br />

communication, of diffusion processes, could be a partial an·<br />

swer to these questions. In complex systems, where species<br />

and individuals interact in many different ways, diffusion and<br />

communication among various parts of the system are likely<br />

to be efficient. There is competition between stabilization<br />

through communication and instability through fluctuations.<br />

The outcome of that competition determines the threshold of<br />

stability.<br />

Structural Stability<br />

When can we begin to speak about "evolution" in its proper<br />

sense? As we have seen, dissipative structures require far-fromequilibrium<br />

conditions. Yet the reaction diffusion equations contain<br />

parameters that can be shifted back to near-equilibrium<br />

conditions. The system can explore the bifurcation diagram in<br />

both directions. Similarly, a liquid can shift from laminar flow<br />

to turbulence and back. There is no definite evolutionary pat·<br />

tern involved.<br />

The situation for models involving the size of the system as<br />

a bifurcation parameter is quite different. Here, growth occurring<br />

irreversibly in time produces an irreversible evolution.<br />

But this remains a special case, even if it can be relevant for<br />

morphogenetic development.<br />

Be it in biological, ecological, or social evolution, we cannot<br />

take as given either a definite set of interacting units, or a definite<br />

set of transformations of these units. The definition of the<br />

system is thus liable to be modified by its evolution. The simplest<br />

example of this kind of evolution is associated with the<br />

concept of structural stability. It concerns the reaction of a<br />

given system to the introduction of new units able to multiply<br />

by taking part in the system's processes.<br />

The problem of the stability of a system vis-a-vis this kind of<br />

change may be formulated as follows: the new constituents, introduced<br />

in small quantities, lead to a new set of reactions among<br />

the system's components. This new set of reactions then enters<br />

into competition with the system·s previous mode of functioning.<br />

If the system is "structurally stable" as far as this


ORDER OUT OF CHAOS<br />

190<br />

intrusion is concerned, the new mode of functioning will be<br />

unable to establish itself and the "innovators" will not survive.<br />

If, however, the structural fluctuation successfully imposes itself-if,<br />

for example, the kinetics whereby the "innovators"<br />

multiply is fast enough for the latter to invade the system instead<br />

of being destroyed-the whole system will adopt a new<br />

mode of functioning: its activity will be governed by a new<br />

"syntax. "6<br />

The simplest example of this situation is a population of<br />

macromolecules reproduced by polymerization inside a system<br />

being fed with the monomers A and B. Let us assume the<br />

polymerization process to be autocatalytic-that is, an already<br />

synthesized polymer is used as a model to form a chain having<br />

the same sequence. This kind of synthesis is much faster than<br />

a synthesis in which there is no model to copy. Each type of<br />

polymer, characterized by a particular sequence of A and B,<br />

can be described by a set of parameters measuring the speed<br />

of the synthesis of the copy it catalyzes, the accuracy of the<br />

copying process, and the mean life of the macromolecule itself.<br />

It may be shown that, under certain conditions, a single<br />

type of polymer having a sequence, shall we say, ABABABA . ..<br />

dominates the population, the other polymers being reduced<br />

to mere "fluctuations" with respect to the first. The . problem<br />

of structural stability arises each time that, as a result of a<br />

copying "error," a new type of polymer characterized by a<br />

hitherto unknown sequence and by a new set of parameters<br />

appears in the system and begins to multiply, competing with<br />

the dominant species for the available A and B monomers.<br />

Here we encounter an elementary case of the classic Darwinian<br />

idea of the "survival of the fittest."<br />

Such ideas form the basis for the model of prebiotic evolution<br />

developed by Eigen and his coworkers. The details of<br />

Eigen's argument are easily accessible elsewhere.? Let us<br />

briefly state that it seems to show that there is only one type of<br />

system that can resist the "errors" that autocatalytic populations<br />

continually make-a polymer system structurally stable<br />

for any possible "mutant polymer." This system is composed<br />

of two sets of polymer molecules. The molecules of the first<br />

set are of the "nucleic acid" type; each molecule is capable of<br />

reproducing itself and act s as a catalyst in the synthesis of


191 ORDER THROUGH FLUCTUATIONS<br />

a molecule of the second set, which is of the proteic type: each<br />

molecule of this second set catalyzes the self-reproduction of a<br />

molecule of the first set. This transcatalytic association between<br />

molecules of the two sets may turn into a cycle (each<br />

"nucleic acid" reproduces itself with the help of a "protein").<br />

It is then capable of stable survival, sheltered from the continual<br />

emergence of new polymers with higher reproductive<br />

efficiency: indeed, nothing can intrude into the self-replicating<br />

cycle formed by "proteins" and "nucleic acids." A new kind<br />

of evolution may thus begin to grow on this stable foundation,<br />

heralding the genetic code.<br />

Eigen 's approach is certainly of great interest. Darwinian<br />

selection for faithful self-reproduction is certainly important<br />

in an environment with a limited capacity. But we tend to believe<br />

that this is not the only aspect involved in pre biotic evolution.<br />

The "far-from-equilibrium" conditions related to critical<br />

amounts of flow of energy and matter are also important. It<br />

seems reasonable to assume that some of the first stages moving<br />

toward life were associated with the formation of mechanisms<br />

capable of absorbing and transforming chemical energy,<br />

so as to push the system into "far-from-equilibrium" conditions.<br />

At this stage life, or "prelife," probably was so diluted<br />

that Darwinian selection did not play the essential role it did in<br />

later stages.<br />

Much of this book has centered around the relation between<br />

the microscopic and the macroscopic. One of the most important<br />

problems in evolutionary theory is the eventual feedback<br />

between macroscopic structures and microscopic events: macroscopic<br />

structures emerging from microscopic events would<br />

in turn lead to a modification of the microscopic mechanisms.<br />

Curiously, at present, the better understood cases concern social<br />

situations. When we build a road or a bridge, we can predict<br />

how this will affect the behavior of the population, and<br />

this will in turn determine other modifications of the modes of<br />

communication in the region. Such interrelated processes generate<br />

very complex situations, the understanding of which is<br />

needed before any kind of modelization. This is why what we<br />

will now describe are only very simple cases.


ORDER OUT OF CHAOS 192<br />

Logistic Evolution<br />

In social cases, the problem of structural stability has a large<br />

number of applications. But it must be emphasized that such<br />

applications imply a drastic simplification of a situation defined<br />

simply in terms of competition between self-replicating<br />

processes in an environment where only a limited amount of<br />

the needed resources exists.<br />

In ecology the classic equation for such a problem is called<br />

the "logistic equation." This equation describes the evolution<br />

of a population containing N individuals, taking into account<br />

the birthrate, the death rate, and the amount of resources available<br />

to the population. The logistic equation can be written<br />

dN/dt = rN(K- N) - mN, where r and m are characteristic<br />

birth and death constants and K the "carrying capacity" of the<br />

environment. Whatever the initial value of N, as time goes on<br />

it will reach the steady-state value N = K-mlr determined by<br />

the differences of the carrying capacity and the ratio of death<br />

and birth constants. When this value is reached, the environ-<br />

K - ..!Il<br />

r<br />

N<br />

Figure 20. Evolution of a population N as a function of time t according to<br />

the logistic curve. The stationary state N=O is unstable while the stationary<br />

state N = K- mlr is stable with respect to fluctuations of N.<br />

t


193 ORDER THROUGH FLUCTUATIONS<br />

ment is saturated, and at each instant as many individuals die<br />

as are born.<br />

The apparent simplicity of the logistic equation conceals to<br />

some extent the complexity of the mechanisms involved. We<br />

have already mentioned the effect of external noise, for example.<br />

Here it has an especially simple meaning. Obviously, if<br />

only because of climatic fluctuations, the coefficients K, m,<br />

and r cannot be taken as constant. We know that such fluctuations<br />

can completely upset the ecological equilibrium and<br />

even drive the population to extinction. Of course, as a result,<br />

new processes, such as the storage of food and the formation<br />

of new colonies, will begin and eventually evolve so that some<br />

effects of external fluctuation may be avoided.<br />

But there is more. Instead of writing the logistic equation as<br />

continuous in time, let us compare the population at fixed time<br />

intervals (for example, separated by a year). This "discrete"<br />

logistic equation can be written in the form Nt + 1 = Nt(l + r<br />

[1-N/KJ), wliere Nt and Nt+ 1 are the populations separated<br />

by a one-year interval (we neglect here the death term). The<br />

remarkable feature, noted by R. May,8 is that such equations,<br />

in spite of their simplicity, admit a bewildering number of solu ·<br />

tions. For values of the parameter Or2, we have, as in the<br />

continuous case, a uniform approach to equilibrium. For values<br />

of r lower than 2.444, a limit cycle sets in: we now have a<br />

periodic behavior with a two-year period. This is followed by<br />

four-, eight-, etc., year cycles, until the behavior can only be<br />

described as chaotic (if r is larger than 2.57). Here we have a<br />

transition to chaos as described in Chapter V. Does this chaos<br />

arise in nature? Recent studies9 seem to indicate that the parameters<br />

characterizing natural populations keep them from<br />

the chaotic region. Why is this so? Here we have one of the<br />

very interesting problems created by the confluence of evolutionary<br />

problems with the mathematics produced by computer<br />

simulation.<br />

Up to now we have taken a static point of view. Let us now<br />

move to mechanisms, whereby the parameters K, r, and m<br />

may vary during biological or ecological evolution.<br />

We have to expect that during evolution the values of the<br />

ecological parameters K, r, and m will vary (as well as many<br />

other parameters and variables, whether they are quantifiable<br />

or not). Living societies continually introduce new ways of ex-


ORDER OUT OF CHAOS 194<br />

ploiting exi.;ting resources or of discovering new ones (that is,<br />

K increases) and continually discover new ways of extending<br />

their lives or of multiplying more quickly. Each ecological<br />

equilibrium defined by the logistic equation is thus only temporary,<br />

and a logistically defined niche will be occupied successively<br />

by a series of species, each capable of ousting the<br />

preceding one when its "aptitude" for exploiting the niche, as<br />

measured by the quantity K-mlr, becomes greater. (See Figure<br />

21.) Thus the logistic equation leads to the definition of a<br />

very simple situation where we can give a quantitative formulation<br />

of the Darwinian idea of the "survival of the fittest."<br />

The "fittest" is the species for which at a given time the quantity<br />

K-mlr is the largest.<br />

As restricted as the problem described by the logistic equation<br />

is, it nonetheless leads to some marvelous examples of<br />

nature's inventiveness .<br />

Take the example of caterpillars, who must remain undetected,<br />

since the slowness of their movement


195 ORDER THROUGH FLUCTUATIONS<br />

of these strategies is effective against all predators at all times,<br />

particularly if a predator is hungry enough. The ideal strategy<br />

is to remain totally undetected. Some caterpillars approach<br />

this ideal, and the variety and sophistication of the strategies<br />

used by the hundreds of lepidopteran species to remain undetected<br />

bring to mind the words of distinguished nineteenthcentury<br />

naturalist Louis Agassiz: "The possibilities of existence<br />

run so deeply into the extravagant that there is scarcely<br />

any conception too extraordinary for Nature to realize. " I O<br />

We cannot resist giving an example reported by Milton<br />

Love. '' The sheep liver trematode has to pass from an ant to a<br />

sheep, where it will finally reproduce itself. The chances of<br />

sheep swallowing an infected ant are very small, but the ant<br />

behaves in a remarkable way: it starts to maximize the probability<br />

of its encounter with a sheep. The trematode has truly<br />

"body snatched" its host. It has burrowed into the ant's brain,<br />

compelling its victim to behave in a suicidal way: the possessed<br />

ant, instead of staying on the ground, climbs to the tip<br />

of a blade of grass and there, immobile, waits for a sheep. This<br />

is indeed an incredibly "clever" solution to the parasites problem.<br />

How it was selected remains a puzzle.<br />

Other situations in biological evolution may be investigated<br />

using models similar to the logistic equation. For instance, it is<br />

possible to calculate the conditions of interspecies competition<br />

under which it may be advantageous for a fraction of the<br />

population to specialize in warlike and nonproductive activity<br />

(for example, the "soldiers" among the social insects). We can<br />

also determine the kind of environment in which a species that<br />

has become specialized, that has restricted the range of its<br />

food resources, will survive more easily than a nonspecialized<br />

species that consumes a wider range of resources. t2 But here<br />

we are approaching some very different problems, which concern<br />

the <strong>org</strong>anization of internally differentiated populations.<br />

Clear distinctions are absolutely necessary if we are to avoid<br />

confusion. In populations where individuals are not interchangeable<br />

and where each, with its own memory, character,<br />

and experience, is called upon to play a singular role, the relevance<br />

of the logistic equation and, more generally, of any simple<br />

Darwinian reasoning becomes quite relative. We shall<br />

return to this problem.<br />

It is interesting to note that the type of curve represented in


ORDER OUT OF CHAOS<br />

196<br />

Figure 21 showing the succession of growths and peaks defined<br />

by a given logistic equation's family with increasing<br />

K - mlr has also been used to describe the multiplication of<br />

certain technical procedures or products. Here too, the discovery<br />

or introduction of a new technique or product breaks<br />

some kind of social, technological, or economic equilibrium.<br />

This equilibrium would correspond to the maximum reached<br />

by the growth curve of the techniques or products with which<br />

the innovation is going to have to compete and that play a similar<br />

role in the situation described by the equation. 13 Thus, to<br />

choose but one example, not only did the spread of the steamship<br />

lead to the disappearance of most sailing ships, but, by<br />

reducing the cost of transportation and increasing its speed, it<br />

caused an increase in the demand for sea transport ("K") and<br />

consequently an increase in the population of ships. We are<br />

obviously representing here an extremely simple situation,<br />

supposedly governed by purely economic logic. Indeed, in<br />

this case innovation seems merely to satisfy, albeit in a different<br />

way, a preexisting need that remains unchanged. However,<br />

in ecology as in human societies, many innovations are<br />

successful without such a preexisting "niche." Such innovations<br />

transform the environment in which they appear, and as<br />

they spread, they create the conditions necessary for their<br />

own multiplication. their "niche." In social situations. in particular,<br />

the creation of a "demand," and even of a "need" for<br />

this demand to fulfill, often appears as correlated with the production<br />

of the goods or techniques that satisfy the demand.<br />

Evolutionary Feedback<br />

A first step toward accounting for this dimension of the evolutionary<br />

process can be achieved by making the "carrying capacity"<br />

of a system a function of the way it is exploited instead<br />

of taking it as given.<br />

In this way some supplementary dimensions of economic<br />

activities, and more particularly the "multiplying effects," can<br />

be represented. Thus we can describe the self-accelerating<br />

properties of systems and the spatial differentiation between<br />

different levels of activity.


197 ORDER THROUGH FLUCTUATIONS<br />

Geographers have already constructed a model correlating<br />

these processes, the Christaller model, defining the optimal<br />

spatial distribution of centers of economic activity. Important<br />

centers would be at the intersection of an hexagonal network,<br />

each being surrounded by a ring of towns of the next smallest<br />

size, each being, etc . . . . Obviously, in actual cases, such a<br />

regular hierarchical distribution is very infrequent: historical,<br />

political, and geographical factors abound, disrupting the spatial<br />

symmetry. But there is more. Even if all the important<br />

sources of asymmetrical development were excluded and we<br />

started from a homogeneous economic and geographical space,<br />

the modeling of the genesis of a distribution such as defined by<br />

Christaller establishes that the kind of static optimalization he<br />

describes constitutes a possible but quite unlikely result of the<br />

process.<br />

The model in question14 stages only the minimal set of variables<br />

implied by a calculation such as Christaller. A set of<br />

equations extending the logistic equations is constructed,<br />

starting from the basic supposition that populations tend to<br />

migrate as a function of local levels of economic activity,<br />

which thus define a kind of local "carrying capacity," here<br />

reduced to an "employment" capacity. But the local population<br />

is .also a potential consumer for locally produced goods.<br />

We have, in fact, a double positive feedback, called the "urban<br />

multiplier," for a local development: both the local population<br />

and the economic infrastructute produced by the already attained<br />

level of activity accelerate the increase of this activity.<br />

But each local level of activity is also determined by competition<br />

with similar centers of activity located elsewhere. The<br />

sale of produced goods or services depends on the cost of<br />

transporting them to consumers and on the size of the "enterprise."<br />

The expansion of each such enterprise depends on a<br />

demand that this expansion itself helps to create and for which<br />

it competes. Thus the respective growth of population and<br />

manufacturing or service activities is linked by strong feedback<br />

and nonlinearities.<br />

The model starts with a hypothetical initial condition, where<br />

"level 1" activity (rural) exists at the different points; it then<br />

permits us to follow successive launchings of activities corresponding<br />

to "superior" levels in Christaller's hierarchy-that<br />

is, implying exportation oo a greater range. Even if the initial


ORDER OUT OF CHAOS<br />

198<br />

65<br />

•<br />

66<br />

•<br />

69<br />

•<br />

64<br />

•<br />

62<br />

•<br />

67<br />

•<br />

65<br />

•<br />

65<br />

•<br />

67<br />

62<br />

• •<br />

66<br />

•<br />

63<br />

•<br />

62<br />

•<br />

66<br />

•<br />

66<br />

•<br />

66<br />

•<br />

65<br />

•<br />

63<br />

•<br />

63<br />

•<br />

62<br />

•<br />

67 67<br />

• •<br />

67<br />

•<br />

65 66<br />

• •<br />

61<br />

•<br />

60<br />

•<br />

69<br />

•<br />

&9<br />

•<br />

68 68<br />

• •<br />

Figure 22. A possible history of "urbanization." • have only function 1;<br />

• have functions 1 and 2; A have functions 1, 2 and 3. are the largest<br />

centers, with functions 1, 2, 3, and 4. At t=O (not represented), all points


199<br />

ORDER THROUGH FLUCTUATIONS<br />

61<br />

•<br />

66<br />

•<br />

67<br />

•<br />

64<br />

•<br />

62<br />

•<br />

64<br />

•<br />

62<br />

•<br />

67<br />

•<br />

66<br />

•<br />

63<br />

•<br />

67<br />

•<br />

64 69<br />

• •<br />

have a "population" of 67 units. At C, the largest center is gOing through a<br />

maximum (152 population units); this is followed by an "urban sprawl," with<br />

creation of satellite cities; this also occurs around the second min center.


ORDER OUT OF CHAOS<br />

200<br />

62<br />

•<br />

65<br />

•<br />

65<br />

•<br />

62<br />

•<br />

64<br />

•<br />

61<br />

•<br />

67<br />

•<br />

65<br />

•<br />

66<br />

•<br />

66<br />

•<br />

64<br />

•<br />

66 66<br />

• •


201 ORDER THROUGH FLUCTUATIONS


ORDER OUT OF CHAOS<br />

202<br />

i7<br />

•<br />

71<br />


203<br />

ORDER THROUGH FLUCTUATIONS<br />

state is quite homogeneous, the model shows that the mere<br />

play of chance factors-factors uncontrolled by the model,<br />

such as the place and time where the different enterprises<br />

start-is sufficient to produce symmetry breakings: the appearance<br />

of highly concentrated zones of activity while others<br />

suffer a reduction in economic activity and are depopulated.<br />

The different computer simulations show growth and decay,<br />

capture and domination, periods of opportunity for alternative<br />

developments followed by solidification of the existing domination<br />

structures.<br />

Whereas Christaller's symmetrical distribution ignores history,<br />

this scenario takes it into account, at least in a very minimal<br />

sense, as an interplay between .. laws," in this case of a<br />

purely economic nature, and the .. chance" governing the sequence<br />

of launchings.<br />

Modelizations of Corrplexity<br />

In spite of its simplicity, our model succeeds in showing some<br />

properties of the evolution of complex systems, and in particular,<br />

the difficulty of .. governing" a development determined by<br />

multiple interacting elements. Each individual action or each<br />

local intervention has a collective aspect that can result in<br />

quite unanticipated global changes. As Waddington emphasized,<br />

at present we have very little understanding of how a<br />

complex system is likely to respond to a given change. Often<br />

this response runs counter to our intuition. The term .. counterintuitive"<br />

was introduced at MIT to express our frustration:<br />

.. The damn thing just does not do what it should do!" To take<br />

the classic example cited by Waddington, a program of slum<br />

clearance results in a situation worse than before. New buildings<br />

attract a larger number of people into the area, but if there<br />

are not enough jobs for them, they remain poor, and their<br />

dwellings become even more overcrowded.t5 We are trained to<br />

think in terms of linear causality, but we need new "tools of<br />

thought": one of the greatest benefits of models is precisely to<br />

help us discover these tools and learn how to use them.<br />

As we have already emphasized, logistic equations are most<br />

relevant when the crucial dimension is the growth of a popula-


ORDER OUT OF CHAOS<br />

204<br />

tion, be it of animals, activities, or habits. What is presupposed<br />

is that each member of a given population can be taken<br />

as the equivalent of any of the others. But this general equivalence<br />

can itself be seen not as a simple general fact but as an<br />

approximation, the validity of which depends on the constraints<br />

and pressures to which this population was submitted<br />

and on the strategy it used to cope with them.<br />

Take, for example, the distinction ecologists have proposed<br />

between K and r strategies. K and r refer to the parameters in<br />

logistic equations. Though this distinction is only relative, it is<br />

especially clear when it characterizes the divergence resulting<br />

from a systematic interaction between two populations, particularly<br />

the prey-predator interaction. In this view, the typical<br />

evolution for a prey population will be the increase in the reproduction<br />

rate r. The predator will evolve toward more effective<br />

ways of capturing its prey-that is, toward an amelioration<br />

of K. But this amelioration, defined in a logistic frame, is liable<br />

to have consequences that go beyond the situations defined by<br />

logistic equations.<br />

As Stephen 1. Gould remarked, 16 a K strategy implies individuals<br />

becoming more and more able to learn from experience<br />

and to store memories-that is, individuals more<br />

complex with a longer period of maturation and apprenticeship.<br />

This in turn means individuals both more "valuable"­<br />

representing a larger biological investment-and characterized<br />

by a longer period of vulnerability. The development of<br />

"social" and "family" ties thus appears as a logical counterpart<br />

of the K strategy. From that point on, other factors, besides<br />

the mere number of individuals in the population,<br />

become more and more relevant and the logistic equation measuring<br />

the success by the number of individuals becomes misleading.<br />

We have here a particular example of what makes<br />

modelization so risky. In complex systems, both the definition<br />

of entities and of the interactions among them can be modified<br />

by evolution. Not only each state of a system but also the very<br />

definition of the system as modelized is generally unstable, or<br />

at least metastable.<br />

We come to problems where methodology cannot be separated<br />

from the question of the nature of the object investigated.<br />

We cannot ask the same questions about a population of<br />

flies that reproduce and die by millions without apparently


205 ORDER THROUGH FLUCTUATIONS<br />

learning from or enlarging their experience and about a population<br />

of primates where each individual is an entanglement of<br />

its own experiences and the traditions of the populations in<br />

which he lives.<br />

We also find that, within anthropology itself, basic choices<br />

must be made between various approaches to collective phenomena.<br />

It is well known, for example, that structural anthropology<br />

privileges those aspects of society where the tools<br />

of logic and finite mathematics can be used, aspects such as<br />

the elementary structures of kinship or the analysis of myths,<br />

whose transformations are often compared to crystalline<br />

growth. Discrete elements are counted and combined. This<br />

contrasts with approaches that analyze evolution in terms of<br />

processes involving large, partially chaotic populations. We<br />

are dealing with two different outlooks and two types of models:<br />

Levi-Strauss defines them respectively as "mechanical"<br />

and "statistical." In the mechanical model "the elements are<br />

of the same scale as the phenomena" and individual behavior<br />

is based on prescriptions referring to the structural <strong>org</strong>anization<br />

of society. The anthropologist makes the logic of this behavior<br />

explicit. The sociologist, on the other hand, works with<br />

statistical models for large populations and defines averages<br />

and thresholds.I7<br />

A society defined entirely in terms of a functional model<br />

would correspond to the Aristotelian idea of natural hierarchy<br />

and order. Each official would perform the duties for which he<br />

has been appointed. These duties would translate at each level<br />

the different aspects of the <strong>org</strong>anization of the society as a<br />

whole. The king gives orders to the architect, the architect to<br />

the contractor, the contractor to the workers. Everywhere a<br />

mastermind is at work. On the contrary, termites and other<br />

social insects seem to approach the "statistical" model. As we<br />

have seen, there seems to be no mastermind behind the construction<br />

of the termites' nest, when interactions among individuals<br />

produce certain types of collective behavior in some<br />

circumstances, but none of these interactions refer to any<br />

global task, being all purely local. Such a description necessarily<br />

implies averages and reintroduces the question of stability<br />

and bifurcations.<br />

Which events will regress, and which arc likely to affect the<br />

whole system? What are the situations of choice, and what are


ORDER OUT OF CHAOS<br />

206<br />

the regimes of stability? Since size or the system's density<br />

may play the role of a bifurcation parameter, how may purely<br />

quantitative growth lead to qualitatively new choices? Questions<br />

such as these call for an ambitious program indeed. As<br />

with the rand K strategies, they lead us to connect the choice<br />

of a "good" model for social behavior and history. How does<br />

the evolution of a population lead it to become more "mechanical"?<br />

This question seems parallel to questions we have already<br />

met in biology. How, for example, does the selection of<br />

the genetic information governing the rates and regulations of<br />

metabolic reactions favor certain paths to such an extent that<br />

development seems to be purposive or appear as the translation<br />

of a "message"?<br />

We believe that models inspired by the concept of "order<br />

through fluctuations" will help us with these questions and<br />

even permit us in some circumstances to give a more precise<br />

formulation to the complex interplay between individual and<br />

collective aspects of behavior. From the physicist's point of<br />

view, this involves a distinction between states of the system<br />

in which all individual initiative is doomed to insignificance on<br />

the one hand, and on the other, bifurcation regions in which an<br />

individual, an idea, or a new behavior can upset the global<br />

state. Even in those regions, amplification obviously does not<br />

occur with just any individual, idea, or behavior, but only with<br />

those that are "dangerous"-that is, those that can exploit to<br />

their advantage the nonlinear relations guaranteeing the stability<br />

of the preceding regime. Thus we are led to conclude<br />

that the same nonlinearities may produce an order out of the<br />

chaos of elementary processes and still, under different circumstances,<br />

be responsible for the destruction of this same<br />

order, eventually producing a new coherence beyond another<br />

bifurcation.<br />

"Order through fluctuations" models introduce an unstable<br />

world where small causes can have large effects, but this world<br />

is not arbitrary. On the contrary, the reasons for the amplification<br />

of a small event are a legitimate matter for rational inquiry.<br />

Fluctuations do not cause the transformation of a systefll 's activity.<br />

Obviously, to use an image inspired by Maxwell, the<br />

match is responsible for the forest fire, but reference to a<br />

match does not suffice to understand the fire. Moreover, the<br />

fact that a fluctuation evades control does not mean that we


207 ORDER THROUGH FLUCTUATIONS<br />

cannot locate the reasons for the instability its amplification<br />

causes.<br />

An Open World<br />

In view of the complexity of the questions raised here, we can<br />

hardly avoid stating that the way in which biological and social<br />

evolution has traditionally been interpreted represents a particularly<br />

unfortunate use of the concepts and methods borrowed<br />

from physics18-unfortunate because the area of<br />

physics where these concepts and methods are valid was very<br />

restricted, and thus the analogies between them and social or<br />

economic phenomena are completely unjustified.<br />

The foremost example of this is the paradigm of optimization.<br />

It is obvious that the management of human society as<br />

well as the action of selective pressures tends to optimize<br />

some aspects of behaviors or modes of connection, but to consider<br />

optimization as the key to understanding how populations<br />

and individuals survive is to risk confusing causes with<br />

effects.<br />

Optimization models thus ignore both the possibility of radical<br />

transformations-that is, transformations that change the<br />

definition of a problem and thus the kind of solution soughtand<br />

the inertial constraints that may eventually force a system<br />

into a disastrous way of functioning. Like doctrines such as<br />

Adam Smith's invisible hand or other definitions of progress in<br />

terms of maximization or minimization criteria, this gives a<br />

reassuring representation of nature as an all-powerful and rational<br />

calculator, and of a coherent history characterized by<br />

global progress. To restore both inertia and the possibility of<br />

unanticipated events-that is, restore the open character of<br />

history-we must accept its fundamental uncertainty. Here we<br />

could use as a symbol the apparently accidental character of<br />

the great cretaceous extinction that cleared the path for the<br />

development of mammals, a small group of ratlike creatures.'<br />

This has been a general presentation, a kind of "bird's-eye<br />

view," and thus has omitted many topics of great interest:<br />

flames, plasmas, and lasers, for example, present nonequilibrium<br />

instabilities of great theoretical and practical interest.


ORDER OUT OF CHAOS<br />

208<br />

Everywhere we look, we find a nature that is rich in diversity<br />

and innovations. The conceptual evolution we have described<br />

is itself embedded in a wider history, that of the progressive<br />

rediscovery of time.<br />

We have seen new aspects of time being progressively incorporated<br />

into physics, while the ambitions to omniscience inherent<br />

in classical science were progressively rejected. In this<br />

chapter we have moved from physics through biology and<br />

ecology to human society, but we could have proceeded in the<br />

inverse order. Indeed, history began by concentrating mainly<br />

on human societies, after which attention was given to the<br />

temporal dimensions of life and of geology. The incorporation<br />

of time into physics thus appears as the last stage of a progressive<br />

reinsertion of history into the natural and social sciences.<br />

Curiously, at every stage of the process, a decisive feature of<br />

this "historicization" has been the discovery of some temporal<br />

heterogeneity. Since the Renaissance, We stern society has<br />

come into contact with different populations that were seen as<br />

corresponding to different stages of development; nineteenthcentury<br />

biology and geology learned to discover and classify<br />

fossils and to recognize in landscapes the memories of a past<br />

with which we coexist; finally, twentieth-century physics has<br />

also discovered a kind of fossil, residual black-body radiation,<br />

which tells us about the beginnings of the universe. Today we<br />

know that we live in a world where different interlocked times<br />

and the fossils of many pasts coexist.<br />

We must now proceed to another question. We have said<br />

that life is starting to seem as "natural as a falling body." What<br />

has the natural process of self-<strong>org</strong>anization to do with a falling<br />

body? What possible link can there be between dynamics, the<br />

science of force and trajectories, and the science of complexity<br />

and becoming, the science of living processes and of the<br />

natural evolution of which they are part? At the end of the<br />

nineteenth century, irreversibility was associated with the phenomena<br />

of friction, viscosity, and heating. Irreversibility lay at<br />

the origin of energy losses and waste. At that time it was still<br />

possible to subscribe to the fiction that irreversibility was only<br />

a result of our ineptitude, of our unsophisticated machines,<br />

and that nature remained fundamentally reversible. Now it is


209 ORDER THROUGH FLUCTUATIONS<br />

no longer possible: today even physics tells us that irreversible<br />

processes play a constructive and indispensable role.<br />

So we come to a question that can be avoided no longer.<br />

What is the relation between this new science of complexity<br />

and the science of simple, elementary behavior? What is the<br />

relation between these two opposing views of nature? Are<br />

there two sciences, two truths for a single world? How is that<br />

possible?<br />

In a certain sense, we have come back to the beginning of<br />

modern science. Now, as at Newton's time, two sciences<br />

come face to face-the science of gravitation, which describes<br />

an atemporal nature subject to laws, and the science of fire,<br />

chemistry. We now understand why it was impossible for the<br />

first synthesis produced by science, the Newtonian synthesis,<br />

to be complete; the forces of interaction described by dynamics<br />

cannot explain the complex and irreversible behavior of<br />

matter. Ignis mutat res. According to this ancient saying,<br />

chemical structures are the creatures of fire, the results of irreversible<br />

processes. How can we bridge the gap between being<br />

and becoming-two concepts in conflict, yet both necessary<br />

to reach a coherent description of this strange world in which<br />

we live?


BOOK THREE<br />

FROM BEING TO<br />

BECOMING


I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I


CHAPTER VII<br />

REDISCOVERING TIME<br />

A Change of Emphasis<br />

Whitehead wrote that a "clash of doctrines is not a disaster, it<br />

is an opportunity. " t If this statement is true, few opportunities<br />

in the history of science have been so promising: two worlds<br />

have come face to face, the world of dynamics and the world of<br />

thermodynamics.<br />

Newtonian science was the outcome, the crowning synthesis<br />

of centuries of experimentation as well as of converging<br />

lines of theoretical research. The same is true for thermodynamics.<br />

The growth of science is quite different from the<br />

uniform unfolding of scientific disciplines, each in turn divided<br />

into an increasing number of watertight compartments. Quite<br />

the contrary, the convergence of different problems and points<br />

of view may break open the compartments and stir up scientific<br />

culture. These turning points have consequences that go<br />

beyond their scientific context and influence the intellectual<br />

scene as a whole. Inversely, global problems often have been<br />

sources of inspiration to science.<br />

The clash of doctrines, the conflict between being and becoming,<br />

indicates that a new turning point has been reached,<br />

that a new synthesis is needed. Such a synthesis is taking<br />

shape today, every bit as unexpected as the preceding ones.<br />

We again find a remarkable convergence of research, all of<br />

which contributes to identifying the difficulties inherent in the<br />

Newtonian concept of a scientific theory.<br />

The ambition of Newtonian science was to present a vision<br />

of nature that would be universal, deterministic, and objective<br />

inasmuch as it contains no reference to the observer, complete<br />

inasmuch as it attains a level of description that escapes the<br />

clutches of time.<br />

213


ORDER OUT OF CHAOS 214<br />

We have reached the core of the problem. "What is time?''<br />

Must we accept the opposition, traditional since Kant, between<br />

the static time of classical physics and the existential<br />

time we experience in our lives? According to Carnap:<br />

Once Einstein said that the problem of the Now worried<br />

him seriously. He explained that the experience of<br />

the Now means something special for man, something<br />

essentially different from the past and the future, but<br />

that this important difference does not and cannot occur<br />

within physics. That this experience cannot be grasped<br />

by science seemed to him a matter of painful but inevitable<br />

resignation. I remarked that all that occurs objectively<br />

can be described in science; on the one hand the<br />

temporal sequence of events is described in physics; and,<br />

on the other hand, the peculiarities of man's experiences<br />

with respect to time, including his different attitude towards<br />

past, present and future, can be described and (in<br />

principle) explained in psychology. But Einstein thought<br />

that these scientific descriptions cannot possibly satisfy<br />

our human needs; that there is something essential about<br />

the Now which is just outside of the realm of science.2<br />

It is interesting to note that Bergson, in a sense following an<br />

opposite road, also reached a dualistic conclusion (see Chapter<br />

III). Like Einstein, Bergson started with a subjective time<br />

and then moved to time in nature, time as objectified by physics.<br />

However, for him this objectivization led to a debasement<br />

of time. Internal existential time has qualitative features that<br />

are lost in the process. It is for this reason that Bergson introduced<br />

the distinction between physical time and duration, a<br />

concept referring to existential time.<br />

But we cannot stop here. As J. T. Fraser says, "The resulting<br />

dichotomy between time felt and time understood is a hallmark<br />

of scientific-industrial civilization, a sort of collective<br />

schizophrenia. "3 As we have already emphasized, where classical<br />

science used to emphasize permanence, we now find<br />

change and evolution; we no longer see in the skies the trajectories<br />

that filled Kant's heart with the same admiration as the<br />

moral law residing in him. We now see strange objects: quasars,·<br />

pulsars, galaxies exploding and being torn apart, stars


215 REDISCOVERING TIME<br />

that, we are told, collapse into "black holes" irreversibly devouring<br />

everything they manage to ensnare.<br />

Time has penetrated not only biology, geology, and the social<br />

sciences but also the two levels from which it has been<br />

traditionally excluded, the microscopic and the cosmic. Not<br />

only life, but also the universe as a whole has a history; this<br />

has profound implications.<br />

The first theoretical paper dealing with a cosmological<br />

model from the point of view of general relativity was published<br />

by Einstein in 1917. It presented a static, timeless view<br />

of the universe, Spinoza's vision translated into physics. But<br />

then comes the unexpected. It became immediately evident<br />

that there were other, time-dependent solutions to Einstein's cosmological<br />

equations. We owe this discovery to the Russian astrophysicist<br />

A. Friedmann and the Belgian G. Lemaitre. At the<br />

same time Hubble and his coworkers were studying the motions<br />

of galaxies, and they demonstrated that the velocity of<br />

distant galaxies is proportional to their distance from earth.<br />

The relation with the expanding universe discovered by Friedmann<br />

and Lemaitre was obvious. Yet for many years physicists<br />

remained reluctant to accept such an "historical"<br />

description of cosmic evolution. Einstein himself was wary of<br />

it. Lemaitre often said that when he tried to discuss with Einstein<br />

the possibility of making the initial state of the universe<br />

more precise and perhaps finding there the explanation of cosmic<br />

rays, Einstein showed no interest.<br />

Today there is new evidence, the famous residual blackbody<br />

radiation, the light that illuminated the explosion of the<br />

hyperdense fireball with which our universe began. The whole<br />

story appears as another irony of history. In a sense, Einstein<br />

has, against his will, become the Darwin of physics. Darwin<br />

taught us that man is embedded in biological evolution; Einstein<br />

has taught us that we are embedded in an evolving universe.<br />

Einstein's ideas led him to a new continent, as unexpected to<br />

him as was America to Columbus. Einstein, like many physicists<br />

of his generation, was guided by a deep conviction that<br />

there was a fundamental, simple level in nature. Yet today this<br />

level is becoming less and less accessible to experiment. The<br />

only objects whose behavior is truly "simple" exist in our own<br />

world, at the macroscopic level. Classical science carefully<br />

chose its objects from this intermediate range. The first ob-


ORDER OUT OF CHAOS 216<br />

jects singled out by Newton-falling bodies, the pendulum,<br />

planetary motion-were simple. We know now, however, that<br />

this simplicity is not the hallmark of the fundamental: it cannot<br />

be attributed to the rest of the world.<br />

Does this suffice? We now know that stability and simplicity<br />

are exceptions. Should we merely disregard the totalizing totalitarian<br />

claims of a conceptualization that, in fact, applies<br />

only to simple and stable objects? Why worry about the incompatibility<br />

between dynamics and thermodynamics?<br />

We must not f<strong>org</strong>et the words of Whitehead, words constantly<br />

confirmed by the history of science: a clash of doctrines<br />

is an opportunity, not a disaster. It has often been<br />

suggested that we simply ignore certain issues for practical<br />

reasons on the grounds that they are based on idealizations<br />

that are difficult to implement. At the beginning of this century,<br />

several physicists suggested abandoning determinism on<br />

the grounds that it was inaccessible in real experience. 4 Indeed,<br />

as we have already emphasized, we never know the exact<br />

positions and velocities of the molecules in a large system;<br />

thus an exact prediction of the system's future evolution is impossible.<br />

More recently, Brillouin hoped to destroy determinism<br />

by appealing to the commonsense truth that accurate<br />

prediction requires an accurate knowledge of the initial conditions<br />

and that this knowledge must be paid for; the exact prediction<br />

necessary to make determinism work requires that an<br />

"infinite" price be paid.<br />

These objections, while reasonable, do not affect the conceptual<br />

world of dynamics. They shed no new light on reality.<br />

Moreover, the improvements in technology could bring us<br />

closer and closer to the idealization implied by classical dynamics.<br />

In contrast, demonstrations of "impossibility" have a fundamental<br />

importance. They imply the discovery of an unexpected<br />

intrinsic structure of reality that dooms an intellectual<br />

enterprise to failure. Such discoveries will exclude the possibility<br />

of an operation that previously could have been imagined<br />

as feasible, at least in principle. "No engine can have an<br />

efficiency greater than one," "no heat engine can produce<br />

useful work unless it is in contact with two sources" are examples<br />

of statements of impossibility which have led to profound<br />

conceptual innovations.


217 .REDISCOVERING TIME<br />

Thermodynamics, relativity, and quantum mechanics are all<br />

rooted in the discovery of impossibilities, of limits to the ambitions<br />

of classical physics. Thus they marked the end of an exploration<br />

that had reached its limits. But we can now see these<br />

scientific innovations in a different light, not as an end but a<br />

beginning, as the opening up of new opportunities. We shall<br />

see in Chapter IX that the second law of thermodynamics expresses<br />

an "impossibility," even on the microscopic level, but<br />

even there the newly discovered impossibility becomes a start·<br />

ing point for the emergence of new concepts.<br />

The End of Universality<br />

Scientific description must be consistent with the resources<br />

available to an observer who belongs to the world he describes<br />

and cannot refer to some being who contemplates the physical<br />

world "from the outside." This is one of the fundamental requirements<br />

of relativity theory. In connection with the propagation<br />

of signals a limit appears that cannot be transgressed<br />

by any observer. Indeed, c, the velocity of light in vacuum<br />

(c=300,000 km/sec), is the limiting velocity for the propagation<br />

of all signals. Thus this limiting velocity plays a fundamental<br />

role. It limits the region in space that may influence the<br />

point where an observer is located.<br />

There is no universal constant in Newtonian physics. This is<br />

the reason for its claim to universality, why it can be applied in<br />

the same way whatever the scale of the objects: the motion of<br />

atoms, planets, and stars are governed by a single law.<br />

The discovery of universal constants signified a radical<br />

change. Using the velocity of light as the comparison standard,<br />

physics has established a distinction between low and<br />

high velocities, those approaching the speed of light.<br />

Likewise, Planck's constant, h; sets up a natural scale according<br />

to the object's mass. The atom can no longer be regarded<br />

as a tiny planetary system. Electrons belong to a<br />

different scale than planets and all other heavy, slow-moving,<br />

macroscopic objects, including ourselves.<br />

Universal constants not only destroy the homogeneity of the<br />

universe by introducing physical scales in terms of which vari-


ORDER OUT OF CHAOS 218<br />

ous behaviors become qualitatively different, they also lead to<br />

a new conception of objectivity. No observer can transmit signals<br />

at a velocity higher than that of light in a vacuum. Hence<br />

Einstein's remarkable conclusion: we can no longer define the<br />

absolute simultaneity of two distant events; simultaneity can<br />

be defined only in terms of a given reference frame. The scope<br />

of this book does not permit an extensive account of relativity<br />

theory. Let us merely point out that Newton's laws did not<br />

assume that the observer was a "physical being." Objective<br />

description was defined precisely as the absence of any reference<br />

to its author. For "nonphysical" intelligent beings capable<br />

of communicating at an infinite velocity, the laws of<br />

relativity would be irrelevant. The fact that relativity is based<br />

on a constraint that applies only to physically localized observers,<br />

to beings who can be in only one place at a time and<br />

not everywhere at once, gives this physics a "human" quality.<br />

This does not mean, however, that it is a "subjective" physics,<br />

the result of our preferences and convictions; it remains subject<br />

to intrinsic constraints that identify us as part of the physical<br />

world we are describing. It is a physics that presupposes an<br />

observer situated within the observed world . .Our dialogue<br />

with nature will be successful only if it is carried on from<br />

within nature.<br />

The Rise of Quantum Mechanics<br />

Relativity altered the classical concept of objectivity. However,<br />

it left unchanged another fundamental characteristic of<br />

classical physics, namely, the ambition to achieve a "complete"<br />

description of nature. After relativity, physicists could<br />

no longer appeal to a demon who observed the entire universe<br />

from outside, but they could still conceive of a supreme mathematician<br />

who, as Einstein claimed, neither cheats nor plays<br />

dice. This mathematician would possess the formula of the<br />

universe, which would include a complete description of nature.<br />

In this sense, relativity remains a continuation of classical<br />

physics.<br />

Quantum mechanics, on the other hand, is the first physical<br />

theory truly to have broken with the past. Quantum mechanics<br />

not only situates us in nature, it also labels us as "heavy"


219 REDISCOVERING TIME<br />

beings composed of a macroscopic number of atoms. In order<br />

to visualize more clearly the consequences of the velocity of<br />

light as a universal constant, Einstein imagined himself riding<br />

a photon. But quantum mechanics discovered that we are too<br />

heavy to ride photons or electrons. We cannot possibly replace<br />

such airy beings, identify ourselves with them, and describe<br />

what they would think, if they were able to think, and what<br />

they would experience, if they were able to feel anything.<br />

The history of quantum mechanics, like that of all conceptual<br />

innovations, is complex, full of unexpected events; it is<br />

the history of a logic whose implications were discovered long<br />

after it was conceived in the urgency of experiment and in a<br />

difficult political and cultural environment.5 This history cannot<br />

be related here; we only wish to emphasize its role in the<br />

construction of the bridge from being to becoming, which is<br />

our main subject.<br />

The birth of quantum mechanics was in itself part of the<br />

quest for this bridge. Planck was interested in the interaction<br />

between matter and radiation. Underlying his work was the<br />

ambition to accomplish for the matter-light interaction what<br />

Boltzmann had achieved for the matter-matter interaction,<br />

namely, to discover a kinetic model for irreversible processes<br />

leading to equilibrium. 6 To his surprise, he was forced, in order<br />

to reach experimental results valid at thermal equilibrium,<br />

to assume that an exchange of energy between matter and radiation<br />

occurred only in discrete steps involving a new universal<br />

constant. This universal constant "h" measures the "size"<br />

of each step.<br />

In this case, as in many others, the challenge of irreversibility<br />

led to decisive progress in physics.<br />

This discovery remained isolated until Einstein presented<br />

the first general interpretation of Planck's constant. He understood<br />

that it had far-reaching implications for the nature of<br />

light. He introduced a revolutionary concept: the waveparticle<br />

duality of light.<br />

Since the beginning of the nineteenth century, light had<br />

been associated with wave properties manifest in phenomena<br />

such as diffraction or interference. However, at the end of the<br />

nineteenth century, new phenomena were discovered, notably<br />

the photoelectric effect-that is, the expulsion of electrons as<br />

the result of the absorption of light. These new experimental


ORDER OUT OF CHAOS 220<br />

results were difficult to explain in terms of the traditional wave<br />

properties of light. Einstein solved the riddle by assuming that<br />

light may be both wave and particle and that these two aspects<br />

are related through Planck's constant. More precisely, a light<br />

wave is characterized by its frequency u and its wavelength X;<br />

h permits us to go from frequency and wavelength to mechanical<br />

quantities such as energy e and momentum p. The relations<br />

between u and A on the one side and e and p on the other are<br />

Very simple: B = hu, p = h/X, and both involve h. 1\.venty years<br />

later, Louis de Broglie extended this wave-particle duality<br />

from light to matter; thus the starting point for the modern<br />

formulation of quantum mechanics.<br />

In 1913 Niels Bohr had linked the new quantum physics to<br />

the structure of atoms (and later of molecules). As a result of<br />

the wave-particle duality, he showed that there exist discrete<br />

sequences of electron orbits. When an atom is excited, the<br />

electron jumps from one orbit to another. At this very instant<br />

the atom emits or absorbs a photon the frequency of which<br />

corresponds to the difference between the energies characterizing<br />

the electron's motion in each of the two orbits. This<br />

difference is calculated in terms of Einstein's formula relating<br />

energy to frequency.<br />

Thus we reach the decisive years 1925-27, a "golden age" of<br />

physics.7 During this short period, Heisenberg, Born, Jordan,<br />

Schrodinger, and Dirac made quantum physics into a consistent<br />

new theory. This theory incorporates Einstein's and de<br />

Broglie's wave-particle duality in the framework of a new generalized<br />

form of dynamics: quantum mechanics. For our purposes<br />

here, the conceptual novelty of quantum mechanics is<br />

essential.<br />

First and foremost, a new formulation, unknown in classical<br />

physics, had to be introduced to allow "quantitization" to be<br />

incorporated into the theoretical language. The essential fact<br />

is that an atom can be found only in discrete energy levels<br />

corresponding to the various electron orbits. In particular, this<br />

means that energy (or the Hamiltonian) can no longer be<br />

merely a function of the position and the moment, as it is in<br />

classical mechanics. Otherwise, by giving the positions and<br />

moments slightly different values, energy could be made to<br />

vary continuously. But as observation reveals, only discrete<br />

levels exist.


22'1<br />

REDISCOVERING TIME<br />

We therefore have to replace the conventional idea that the<br />

Hamiltonian is a function of position and momenta with something<br />

new; the basic idea of quantum mechanics is that the<br />

Hamiltonian as well as the other quantities of classical mechanics,<br />

such as coordinates q or momenta p, now become<br />

operators. This is one of the boldest ideas ever introduced in<br />

science, and we would like to discuss it in detail.<br />

It is a simple idea, even if at first it seems somewhat abstract.<br />

We have to distinguish the operator-a mathematical<br />

operation-and the object on which it operates-a function.<br />

As an example, take as the mathematical "operator" the derivative<br />

represented by d/dx and suppose it acts on a function-say,<br />

x2; the result of this operation is a new function, this<br />

time "2x." However, certain functions behave in a peculiar<br />

way with respect to derivation. For example, the derivative of<br />

"e3x" is "3e3x": here we return to the original function simply<br />

multiplied by some number-here, 3. Functions that are<br />

merely recovered by a given operator to them are known as the<br />

"eigenfunctions" of this operator, and the numbers by which<br />

the eigenfunction is multiplied after the application of the operator<br />

are the "eigenvalues" of the operator.<br />

To each operator there thus corresponds an ensemble, a "reservoir"<br />

of numerical values; this ensemble forms its "spectrum."<br />

This spectrum is "discrete" when the eigenvalues form<br />

a discrete series. There exists, for instance, an operator with<br />

all the integers 0, 1, 2 . . . as eigenvalues. A spectrum may<br />

also be continuous-for example, when it consists of all the<br />

numbers between 0 and 1.<br />

The basic concept of quantum mechanics may thus be expressed<br />

as follows: to all physical quantities in classical mechanics<br />

there corresponds in quantum mechanics an operator,<br />

and the numerical values that may be taken by this physical<br />

quantity are the eigenvalues of this operator. The essential<br />

point is that the concept of physical quantity (represented by<br />

an operator) is now distinct from that of its numerical values<br />

(represented by the eigenvalues of the operator). In particular,<br />

energy will now be represented by the Hamiltonian operator,<br />

and the energy levels-the observed values of the energywill<br />

be identified with the eigenvalues corresponding to this<br />

operator.<br />

The introduction of operators opened up to physics a micro-


ORDER OUT OF CHAOS 222<br />

scopic world of unsuspected richness, and we regret that we<br />

cannot devote more space to this fascinating subject, in which<br />

creative imagination and experimental observation are so successfully<br />

combined. Here we wish merely to stress that the<br />

microscopic world is governed by laws having a new structure,<br />

thereby putting an end once and for all to the hope of discovering<br />

a single conceptual scheme common to all levels of description.<br />

A new mathematical language invented to deal with a certain<br />

situation may actually open up fields of inquiry that are<br />

full of surprises, going far beyond the expectations of its originators.<br />

This was true for differential calculus, which lies at<br />

the root of the formulation of classical dynamics. It is true as<br />

well for operator calculus. Quantum theory, initiated as demanded<br />

by the result of unexpected experimental discoveries,<br />

was quick to reveal itself as pregnant with new content.<br />

Today, more than fifty years after the introduction of operators<br />

into quantum mechanics, their significance remains a subject<br />

of lively discussion. From the historical point of view, the<br />

introduction of operators is linked to the existence of energy<br />

levels, but today operators have applications even in classical<br />

physics. This implies that their significance has been extended<br />

beyond the expectations of the founders of quantum mechanics.<br />

Operators now come into play as soon as, for one reason<br />

or another, the notion of a dynamic trajectory has to be discarded,<br />

and with it, the deterministic description a trajectory<br />

implies.<br />

Heisenbergs Uncertainty Relation<br />

We have seen that in quantum mechanics to each physical<br />

quantity corresponds an operator that acts on functions. Of<br />

special importance are the eigenfunctions and the eigenvalues<br />

corresponding to the operator under consideration. The eigenvalues<br />

correspond precisely to the numerical values the physical<br />

quantity can now take. Let us take a closer look at the<br />

operators quantum mechanics associates with coordinates q<br />

and momenta p; their coordinates are, as we have seen in<br />

Chapter II, the canonical variables.<br />

In classical mechanics coordinates and momenta are inde-


223 REDISCOVERING TIME<br />

pendent in the sense that we can ascribe to a coordinate a<br />

numerical value quite independent of the value we have ascribed<br />

to the momentum. However, the existence of Planck's<br />

constant h implies the reduction in the number of independent<br />

variables. We could have guessed this right away from the<br />

Einstein-de Broglie relation A= hlp, which, as we have seen,<br />

connects wavelength to momentum. Planck's constant h expresses<br />

a relation between lengths (closely related to the concept<br />

of coordinates) and momenta. Therefore, positions and<br />

momenta can no longer be independent variables, as in classical<br />

mechanics. The operators corresponding to positions and<br />

momenta can be expressed in terms of the coordinate alone or<br />

in terms of the momentum, something explained in all textbooks<br />

dealing with quantum mechanics.<br />

The important point is that in all cases, only one type of<br />

quantity appears (either coordinate or momentum), but not<br />

both. In this sense we may say that the quantum mechanics<br />

divides the number of classical mechanical variables by a factor<br />

of two.<br />

One fundamental property results from the relation between<br />

operators in quantum mechanics: the two operators q0 P and<br />

Pop do not commute-that is, the results of q0pP0p and of<br />

Pop%p applied to the same function are different. This has profound<br />

implications, since only commuting operators admit<br />

common eigenfunctions. Thus we cannot identify a function<br />

that would be an eigenfunction of both coordinate and momentum.<br />

As a consequence of the definition of the coordinate and<br />

momentum operators in quantum mechanics, there can be no<br />

state in which the physical quantities, coordinate q and momentum<br />

p, both have a well-defined value. This situation, unknown<br />

in classical mechanics, is expressed by Heisenberg's<br />

famous uncertainty relations. We can measure a coordinate<br />

and a momentum, but the dispersions of the respective possible<br />

predictions as expressed by f::j,q,f::j,p are related by the<br />

Heisenberg inequality f::j,qf::j,p;;::.h. We can make f::j,q as small as<br />

we want, but then f::j,p goes to infinity, and vice versa.<br />

Much has been written about Heisenberg's uncertainty relations,<br />

and our discussion is admittedly oversimplified. But we<br />

wish to give our readers some understanding of the new problem<br />

that re:sult:s from the u:se of operators; Heisenberg's uncertainty<br />

relation necessarily leads to a revision of the concept of


ORDER OUT OF CHAOS 22 4<br />

causality. It is possible to determine the coordinate precisely.<br />

But the moment we do so, the momentum will acquire an arbitrary<br />

value, positive or negative. In other words, in an instant<br />

the position of the object will become arbitrarily distant.<br />

The meaning of localization becomes blurred: the concepts<br />

that form the basis of classical mechanics are profoundly altered.<br />

These consequences of quantum mechanics were unacceptable<br />

to many physicists, including Einstein; and many experiments<br />

were devised to demonstrate their absurdity. An<br />

attempt was also made to minimize the conceptual change involved.<br />

In particular, it was suggested that the foundation of<br />

quantum mechanics is in some way related to perturbations<br />

resulting from the process of observation. A system was<br />

thought to possess intrinsically well-defined mechanical parameters<br />

such as coordinates and momenta; but some of them<br />

would be made fuzzy by measurement, and Heisenberg's uncertainty<br />

relation would only express the perturbation created<br />

by the measurement process. Classical realism thus would remain<br />

intact on the fundamental level, and we would simply<br />

have to add a positivistic qualification. This interpretation<br />

seems too narrow. It is not the quantum measurement process<br />

that disturbs the results. Far from it: Planck's constant forces<br />

us to revise our concepts of coordinates and momenta. This<br />

conclusion has been confirmed by recent experiments designed<br />

to test the assumption of local hidden variables that<br />

were introduced to restore classical determinism. s The results<br />

of those experiments confirm the striking consequences of<br />

quantum mechanics.<br />

That quantum mechanics obliges us to speak less absolutely<br />

about the localization of an object implies, as Niels Bohr often<br />

emphasized, that we must give up the realism of classical<br />

physics. For Bohr, Planck's constant defines the interaction<br />

between a quantum system and the measurement device as<br />

nondecomposable. It is only to the quantum phenomenon as a<br />

whole, including the measurement interaction, that we can ascribe<br />

numerical values. All description thus implies a choice<br />

of the measurement device, a choice of the question asked. In<br />

this sense, the answer, the result of the measurement, does not<br />

give us access to a given reality. We have to decide which measurement<br />

we are going to perform and which question our ex-


225 REDISCOVERING TIME<br />

periments will ask the system. Thus there is an irreducible<br />

multiplicity of representations for a system, each connected<br />

with a determined set of operators.<br />

This implies a departure from the classical notion of objectivity,<br />

since in the classical view the only "objective" description<br />

is the complete description of the system as it is,<br />

independent of the choice of how it is observed.<br />

Bohr always emphasized the novelty of the positive choice<br />

introduced through measurement. The physicist has to choose<br />

his language, to choose the macroscopic experimental device.<br />

Bohr expressed this idea through the principle of complementarity,9<br />

which may be considered as an extension of Heisenberg's<br />

uncertainty relations. We can measure coordinates or<br />

momenta, but not both. No single theoretical language articulating<br />

the variables to which a well-defined value can be attributed<br />

can exhaust the physical conent of a system. Various<br />

possible languages and points of view about the system may<br />

be complementary. They all deal with the same reality, but it is<br />

impossible to reduce them to one single description. The irreducible<br />

plurality of perspectives on the same reality expresses<br />

the impossibility of a divine point of view from which the<br />

whole of reality is visible. However, the lesson of the principle<br />

of complementarity is not a lesson in resignation. Bohr used to<br />

say that the significance of quantum mechanics always made<br />

him dizzy, and we do indeed feel dizzy when we are torn from<br />

the comfortable routine of common sense.<br />

The real lesson to be learned from the principle of complementarity,<br />

a lesson that can perhaps be transferred to other<br />

fields of knowledge, consists in emphasizing the wealth of reality,<br />

which overflows any single language, any single logical<br />

structure. Each language can express only part of reality. Music,<br />

for example, has not been exhausted by any of its realizations,<br />

by any style of composition, from Bach to Schonberg.<br />

We have emphasized the importance of operators because<br />

they demonstrate that the reality studied by physics is also a<br />

mental construct; it is not merely given. We must distinguish<br />

between the abstract notion of a coordinate or of momentum,<br />

represented mathematically by operators, and their numerical<br />

realization, which can be reached through experiments. One<br />

of the reasons for the opposition between the ··two cultures"<br />

may have been the belief that literature corresponds to a con-


ORDER OUT OF CHAOS 226<br />

ceptualization of reality, to "fiction," while science seems to<br />

express objective "reality. " Quantum mechanics teaches us<br />

that the situation is not so simple. On all levels reality implies<br />

an essential element of conceptualization.<br />

The Temporal Evolution of Quantum Systems<br />

We shall now move on to discuss the temporal evolution of<br />

quantum systems. As in classical mechanics, the Hamiltonian<br />

plays a fundamental role. As we have seen, in quantum mechanics<br />

it is replaced by the Hamiltonian operator Hop· This<br />

energy operator plays a central role: on the one hand, its<br />

eigenvalues correspond to the energy levels; on the other<br />

hand, as in classical mechanics, the Hamiltonian operator determines<br />

the temporal evolution of the system. In quantum<br />

mechanics the role played by the canonical equation of classical<br />

mechanics is taken by the Schrodinger equation, which<br />

expresses the time evolution of the fu nction characterizing the<br />

quantum state as the result of the application of the operator<br />

Hop on the wave function \jJ (there are, of course, other formulations,<br />

which we cannot describe here). The term "wave<br />

function" has been chosen to emphasize once again the waveparticle<br />

duality so fu ndamental in all of quantum physics. \jJ is<br />

a wave amplitude that evolves according to a particle type of<br />

equation determined by the Hamiltonian. Schrodinger's equation,<br />

like the canonical equation of classical physics, expresses<br />

a reversible and deterministic evolution. The<br />

reversible change of wave function corresponds to a reversible<br />

motion along a trajectory. If the wave fu nction at a given instant<br />

is known, Schrodinger's equation allows it to be calculated<br />

for any previous or subsequent instant. From this viewpoint,<br />

the situation is strictly similar to that in classical mechanics.<br />

This is because the uncertainty relations of quantum mechanics<br />

do not include time. Time remains a number, not an operator,<br />

and only operators can appear in Heisenberg's uncertainty<br />

re lations.<br />

Quantum mechanics deals with only half of the variables of<br />

dassical mechanics. As a result, classical determinism becomes<br />

inapplicable, and in quantum physics statistical consid-


227 REDISCOVERING TIME<br />

erations play a central role. It is through the wave intensity ttJ2<br />

(the square of the amplitude) that we make contact with statistical<br />

considerations.<br />

The standard statistical interpretation of quantum mechanics<br />

runs as follows: consider the eigenfunctions of some operator-say,<br />

the energy operator H0 P -and the corresponding<br />

eigenvalues. In general the wave function tts will not be the<br />

eigenfunction of the energy operator, but it can be expressed<br />

as the superposition of these eigenfunctions. The respective<br />

importance of each eigenfunction in this superposition allows<br />

us to calculate the probability for the appearance of the various<br />

possible corresponding eigenvalues.<br />

Here again we notice a fundamental departure from classical<br />

theory. Only probabilities can be predicted, not single<br />

events. This was the second time in the history of science that<br />

probabilities were used to explain some basic features of nature.<br />

The first time was in Boltzmann's interpretation of entropy.<br />

There , however, a subjective point of view remained<br />

possible; in this view, "only" our ignorance in the face of the<br />

complexity of the systems considered prevented us from achieving<br />

a complete description. (We shall see that today it is possible<br />

to overcome this attitude.) Here, as before, the use of probabilities<br />

was unacceptable to many physicists-including Einstein-who<br />

wished to achieve a "complete" deterministic<br />

description. Just as with irreversibility, an appeal to our ignorance<br />

seemed to offer a way out: our inaptitude would make us<br />

responsible for statistical behavior in the quantum world, just<br />

as it makes us responsible for irreversibility.<br />

Once again we come to the problem of hidden variables.<br />

However, as we have said, there has been no experimental evidence<br />

to justify the introduction of such variables, and the rol<br />

of probabilities seems irreducible.<br />

There is only one case in which the Schrodinger equation<br />

leads to a deterministic prediction: that is when tlJ, instead of<br />

being a superposition of eigenfunctions, is reduced to a single<br />

one. In particular, in an ideal measurement process, a system<br />

may be prepared in such a way that the result of a given measurement<br />

may be predicted. We then know that the system is<br />

described by the corresponding eigenfunction. From then on,<br />

the system may be described with certainty as being in the<br />

eigenstate indicated by the measurement result.


ORDER OUT OF CHAOS 228<br />

The measurement process in quantum mechanics has a spe·<br />

cial significance that is attracting considerable interest today.<br />

Suppose we start with a wave function, which is indeed a superposition<br />

of eigenfunctions. As a result of the measurement<br />

process, this single collection of systems all represented by<br />

the same wave function is replaced by a collection of wave<br />

functions corresponding to the various eigenvalues that may<br />

be measured. Stated technically, a measurement leads from a<br />

single wave function (a "pure" state) to a mixture .<br />

As Bohr and Rosenfeld IO repeatedly pointed out, every<br />

measurement contains an element of irreversibility, an appeal<br />

made to irreversible phenomena, such as chemical processes<br />

corresponding to the recording of the "data." Recording is accompanied<br />

by an amplification whereby a microscopic event<br />

produces an effect on a macroscopic level-that is, a level at<br />

which we can read the measuring instruments. The measurement<br />

thus presupposes irreversibility.<br />

This was in a sense already true in classical physics. How·<br />

ever, the problem of the irreversible character of measurement<br />

is more urgent in quantum mechanics because it raises questions<br />

at the level of its formulation.<br />

The usual approach to this problem states that quantum mechanics<br />

has no choice but to postulate the coexistence of two<br />

mutually irreducible processes. the reversible and continuous<br />

evolution described by Schrodinger's equation and the irreversible<br />

and discontinuous reduction of the wave function to<br />

one of its eigenfunctions at the time of measurement. Thus the<br />

paradox: the reversible Schrodinger equation can be tested<br />

only by irreversible measurements that the equation is by definition<br />

unable to describe. It is thus impossible for quantum<br />

mechanics to set up a closed structure.<br />

In the face of these difficulties, some physicists have once<br />

more taken refuge in subjectivism, stating that we-our measurement<br />

and even, for some, our mind-determine the evolution<br />

of the system that breaks the law of natural, "objective"<br />

reversibility. 11 Others have concluded that Schrodinger's equation<br />

was not "complete" and that new terms must be added to<br />

account for the irreversibility of the measurement. Other more<br />

improbable "solutions" have also been proposed, such as<br />

Everett's many-world hypothesis (see d'Espagnat, ref. 8). For<br />

us, however, the coexistence in quantum mechanics of revers-


229 REDISCOVERING TIME<br />

ibility and irreversibility shows that the classical idealization<br />

that describes the dynamic world as self-contained is impossible<br />

at the microscopic level. This is what Bohr meant when he<br />

noted that the language we use to describe a quantum system<br />

cannot be separated from the macroscopic concepts that describe<br />

the functioning of our measurement instruments. Schrodinger's<br />

equation does not describe a separate level of reality;<br />

rather it presupposes the macroscopic world to which we belong.<br />

·<br />

The problem of measurement in quantum mechanics is thus<br />

an aspect of one of the problems to which this book is devoted-the<br />

connection between the simple world described by<br />

Hamiltonian trajectories and Schrodinger's equation, and the<br />

complex macroscopic world of irreversible processes.<br />

In Chapter IX, we shall see that irreversibility enters classical<br />

physics when the idealization involved in the concept of a<br />

trajectory becomes inadequate. The measurement problem in<br />

quantum mechanics is susceptible to the same type of solution.12<br />

Indeed, the wave function represents the maximum<br />

knowledge of a quantum system. As in classical physics, the<br />

object of this maximum knowledge satisfies a reversible evolution<br />

equation. In both cases, irreversibility enters when the<br />

ideal object corresponding to maximum knowledge has to be<br />

replaced by less idealized concepts. But when does this happen?<br />

This is the question of the physical mechanisms of irreversibility<br />

to which we shall turn in Chapter IX. But let us first<br />

summarize some other features of the renewal of contemporary<br />

science.<br />

A Nonequilibrium Universe<br />

The two scientific revolutions described in this chapter started<br />

as attempts to incorporate universal constants, c and h, into<br />

the framework of classical mechanics. This led to far-reaching<br />

consequences, some of which we have described here. From<br />

other perspectives, relativity and quantum mechanics seemed<br />

to adhere to the basic world view expressed in Newtonian mechanics.<br />

This is especially true regarding the role and meaning<br />

of time. In quantum mechanics, once the wave function at time


ORDER OUT OF CHAOS<br />

230<br />

zero is known, its value ljJ(t) both for future and past is deter·<br />

mined. Likewise, in relativity theory the static geometric charac·<br />

ter of time is often emphasized by the use of four-dimensional<br />

notation (three dimensions for space and one for time). As expressed<br />

concisely by Minkowski in 1908, "space by itself and<br />

time by itself are doomed to fade away into mere shadows, and<br />

only a kind of union of the two will preserve an independent<br />

reality . .. only a world in itself will subsist." 13<br />

But over the past five decades this situation has radically<br />

changed. Quantum mechanics has become the main tool for<br />

dealing with elementary particles and their transformations. It<br />

is outside the scope of this book to describe the bewildering<br />

variety of elementary particles that have appeared during the<br />

past few years.<br />

We want only to recall that, using both quantum mechanics<br />

and relativity, Dirac demonstrated that we have to associate to<br />

each particle of mass m and charge e an antiparticle of the<br />

same mass but of opposite charge. Positrons, the antiparticles<br />

of electrons, as well as antiprotons, are currently being produced<br />

in high-energy accelerators. Antimatter has become a<br />

common subject of study in particle physics. Particles and<br />

their corresponding antiparticles annihilate each other when they<br />

collide, producing photons, massless particles corresponding<br />

to light. The equations of quantum theory are symmetric in<br />

respect to the exchange particle-antiparticle, or more precisely,<br />

they are symmetric in respect to a weaker requirement<br />

known as the CPT symmetry. In spite of this symmetry, there<br />

exists a remarkable dissymmetry between particles and antiparticles<br />

in the world around us. We are made of particles<br />

(electrons, protons), while antiparticles remain rare laboratory<br />

products. If particles and antiparticles coexisted in equal<br />

amount, all matter would be annihilated. There is strong evidence<br />

that antimatter does not exist in our galaxy, but the possibility<br />

that it exists in distant galaxies cannot be excluded. We<br />

can imagine a mechanism in the universe that separates particles<br />

and antiparticles, hides antiparticles somewhere. However,<br />

it seems more likely that we live in a "nonsymmetrical"<br />

universe where matter completely dominates antimatter.<br />

How is this possible? A model explaining the situation was<br />

presented by Sakharov in 1966, and today much work is being<br />

done along these lines.14 One essential element of the model is


231 REDISCOVERING TIME<br />

that, at the time of the formation of matter, the universe had to<br />

be in n{)nequilibrium conditions, for at equilibrium the law of<br />

mass action discussed in Chapter V would have required equal<br />

amounts of matter and antimatter.<br />

What we want to emphasize here is that nonequilibrium has<br />

now acquired a new, cosmological dimension. Without nonequilibrium<br />

and without the irreversible processes linked to it,<br />

the universe would have a completely different structure.<br />

There would be no appreciable amount of matter, only some<br />

fluctuating local excesses of matter over antimatter, or vice<br />

versa.<br />

From a mechanistic theory that was modified to account for<br />

the existence of the universal constant h, quantum theory has<br />

evolved into a theory of mutual transformations of elementary<br />

particles. In recent attempts to formulate a "unified theory of<br />

elementary particles" it has even been suggested that all particles<br />

of matter, including the proton, are unstable (however, the<br />

lifetime of the proton would be enormous, of the order of 1 Q 3 0<br />

years). Mechanics, the science of motion, instead of corresponding<br />

to the fundamental level of description, becomes a<br />

mere approximation, useful only because of the long lifetime<br />

of elementary particles such as protons.<br />

Relativity theory has gone through the same transformations.<br />

As we mentioned, it started as a geometric theory that<br />

strongly emphasized timeless features. Today it is the main<br />

tool for investigating the thermal history of the universe, for<br />

providing clues to the mechanisms that led to the present<br />

structure of the universe. The problem of time, of irreversibility,<br />

has therefore acquired a new urgency. From the field of<br />

engineering, of applied chemistry, where it was first formulated,<br />

it has spread to the whole of physics, from elementary<br />

particles to cosmology.<br />

From the perspective of this book, the importance of quantum<br />

mechanics lies in its introduction of probability into microscopic<br />

physics. This should not be confused with the stochastic<br />

processes that describe chemical reactions as discussed in<br />

Chapter V. In quantum mechanics, the wave function evolves<br />

in a deterministic fashion, except in the measurement process.<br />

We have seen that in the fifty years since the formulation of<br />

quantum mechanics the study of nonequilibrium processes<br />

has revealed that fluctuations, stochastic elements, are impor-


ORDER OUT OF CHAOS 232<br />

tant even on the microscopic scale. We have repeatedly stated<br />

in this book that the reconceptualization of physics going on<br />

today leads from deterministic, reversible processes to stochastic<br />

and irreversible ones. We believe that quantum mechanics<br />

occupies a kind of intermediate position in this process. There<br />

probability appears, but not irreversibility. We expect, and we<br />

shall give some reasons for this in Chapter IX, that the next<br />

step will be the introduction of fundamental irreversibility on<br />

the microscopic level. In contrast with the attempts to restore<br />

classical orthodoxy through hidden variables or other means,<br />

we shall argue that it is necessary to move even farther away<br />

from deterministic descriptions of nature and adopt a statistical,<br />

stochastic description.


CHAPTER VIII<br />

THE CLASH OF<br />

DOCTRINES<br />

Probability and Irreversibility<br />

We shall see that nearly everywhere the physicist has<br />

purged from his science the use of one-way time, as<br />

though aware that this idea introduces an antrlropomorphlc<br />

element alien to the ideals of physics. Nevertheless,<br />

in several important cases unidirectional time<br />

and unidirectional causality have been invoked, but always,<br />

as we shall proceed to show. in support of some<br />

false doctrine.<br />

G. N. LEWIS'<br />

The law that entropy always increases-the second<br />

law of thermodynamics-holds, I think, the supreme<br />

position among the laws of Nature. If someone points<br />

out to you that your pet theory of the universe is in<br />

disagreement with Maxwell's equations-then so<br />

much the worse for Maxwells equations. If it is found to<br />

be contradicted by observation-well, these experimentalists<br />

do bungle things sometimes. But if your<br />

theory is found to be against the second law of thermodynamics<br />

I can give you no hope; there is nothing<br />

for it but to collapse in deepest humiliation.<br />

A S. EDDINGTON2<br />

With Clausius' formulation of the second law of thermodynam·<br />

ics, the conflict between thermodynamics and dynamics be·<br />

came obvious. There is hardly a single question in physis that<br />

has been more often and more actively discussed than the rela·<br />

233


ORDER OUT OF CHAOS<br />

234<br />

tion between thermodynamics and dynamics. Even now, a<br />

hundred and fifty years after Clausius, the question still<br />

arouses strong feelings. No one can remain neutral in this conflict,<br />

which involves the meaning of reality and time. Must<br />

dynamics, the mother of modern science, be abandoned in<br />

favor of some form of thermodynamics? That was the view of<br />

the "energeticists," who exerted great influence during the<br />

nineteenth century. Is there a way to "save" dynamics, to recoup<br />

the second law without giving up the formidable structure<br />

built by Newton and his successors? What role can<br />

entropy play in a world described by dynamics?<br />

We have already mentioned the answer proposed by Boltzmann.<br />

Boltzmann's famous equation S = k log P relates entropy<br />

and probability: entropy grows because probability grows.<br />

Let us immediately emphasize that in this perspective the second<br />

law would have great practical importance but would be of<br />

no fundamental significance. In his excellent book The Ambidextrous<br />

Universe, Martin Gardner writes: "Certain events go<br />

only one way not because they can't go the other way but because<br />

it is extremely unlikely that they go backward. "3 By improving<br />

our abilities to measure less and less unlikely events,<br />

we could reach a situation in which the second law would play<br />

as small a role as we want. This is the point of view that is<br />

often taken today. However, this was not Planck's point of<br />

view:<br />

It would be absurd to assume that the validity of the second<br />

law depends in any way on the skill of the physicist<br />

or chemist in observing or experimenting. The gist of the<br />

second law has nothing to do with experiment; the law<br />

asserts briefly that there exists in nature a quantity which<br />

changes always in the same sense in all natural processes.<br />

The proposition stated in this general form may<br />

be correct or incorrect; but whichever it may be, it will<br />

remain so, irrespective of whether thinking and measuring<br />

beings exist on the earth or not, and whether or not,<br />

assuming they do exist, they are able to measure the details<br />

of physical or chemical processes more accurately<br />

by one, two, or a hundred decimal places than we can.<br />

The limitation to the law, if any, must lie in the same<br />

province as its essential idea, in the observed Nature, and


235 THE CLASH OF DOCTRINES<br />

not in the Observer. That man's experience is called upon<br />

in the deduction of the law is of no consequence; for that<br />

is, in fact, our only way of arriving at a knowledge of<br />

natural law. 4<br />

However, Planck's views remained isolated . As we noted,<br />

most scientists considered the second law the result of approximations,<br />

the intrusion of subjective views into the exact world<br />

of physics. For example, in a celebrated sentence Born stated,<br />

"Irreversibility is the effect of the introduction of ignorance<br />

into the basic laws of physics. "5<br />

In the present chapter we wish to describe some of the basic<br />

steps in the development of the interpretation of the second<br />

law. We must first understand why this problem appeared to<br />

be so difficult. In Chapter IX we shall go on to present a new<br />

approach that, we hope, will clearly express both the radical<br />

originality and the objective meaning of the second law. Our<br />

conclusion will agree with Planck's view. We shall show that,<br />

far from destroying the formidable structure of dynamics, the<br />

second law adds an essential new element to it.<br />

First we wish to clarify Boltzmann's association of probability<br />

and entropy. We shall begin by describing the "urn<br />

model" proposed by P. and T. Ehrenfest. 6 Consider N objects<br />

(for example, balls) distributed between two containers A and<br />

B. At regular time intervals (for example, every second) a ball<br />

tioe n<br />

time n+1 1<br />

D EJ<br />

A<br />

A<br />

!-tottery<br />

B<br />

or N-k+1 N-k-1<br />

Figure 23. Ehrenfest's urn model. N balls are distributed between two containers<br />

A and B. At time n there are k balls in A and N- k balls in B. At regular<br />

time intervals a ball is taken at random from A and put in B.<br />

B


ORDER OUT OF CHAOS 236<br />

is chosen at random and moved from one container to the<br />

other. Suppose that at time n there are k balls in A and N- k<br />

balls in B. Then at time n + I there can be in A either k- I or<br />

k+ I balls. We have the transition probabilities kiN for k-+k - 1<br />

and 1-k/N for k-+k + 1. Suppose we continue the game. We<br />

expect that as a result of the exchanges of balls the most probable<br />

distribution in Boltzmann's sense will be reached. When the<br />

number N of balls is large, this distribution corresponds to an<br />

equal number N/2 of balls in each urn. This can be verified by<br />

elementary calculations or by performing the experiment.<br />

N<br />

k - -<br />

2<br />

t<br />

Figure 24. Approach to equilibrium (k = Nt2) in Ehrenfest's urn model<br />

(schematic representation).<br />

The Ehrenfest model is a simple example of a "Markov process"<br />

(or Markov "chain"), named after the great Russian<br />

mathematician Markov, who was one of the first to describe<br />

such processes (Poincare was another). In brief, their characteristic<br />

feature is the existence of well-defined transition probabilities<br />

independent of the previous history of the system.<br />

Markov chains have a remarkable property: they can be described<br />

in terms of entropy. Let us call P(k) the probability of<br />

finding k balls in A. We may then associate to it an "J-{ quantity,"<br />

which has the precise properties of entropy that we discussed<br />

in Chapter IV. Figure 25 gives an example of its<br />

evolution. The Jf quantity varies uniformly with time, as does<br />

the entropy of an isolated system. It is true that J-{ decreases<br />

with time, while the entropy S increases, but that is a matter of<br />

·<br />

definition: J-{ plays the role of -s.


237 THE CLASH OF DOCTRINES<br />

Figure 25. Time evolution of the J{ quantity (defined in the text) corresponding<br />

to the Ehrenfest model. This quantity decreases monotonously<br />

and vanishes for long times.<br />

The mathematical meaning of this "J-l quantity" is worth<br />

considering in more detail: it measures the difference between<br />

the probabilities at a given time and those that exist at the<br />

equilibrium state (where the number of balls in each urn is<br />

N/2). The argument used in the Ehrenfest urn model can be<br />

generalized. Let us consider the partition of a square-that is,<br />

we subdivide the square into a number of disjointed regions<br />

(see Figure 26). Then we consider the distribution of particles<br />

in the square and call P(k,t) the probability of finding a particle<br />

in the region k. Similarly, we call Peqm(k) this quantity when<br />

uniformity is reached. We assume that, as in the urn model,<br />

there exist well-defined transition probabilities. The definition<br />

of the J-l. quantity is<br />

t<br />

J{ = P(k,t) log J


ORDER OUT OF CHAOS 238<br />

sponding values of P(k,t) would be P( l ,t) = 1, all others zero. As<br />

the result we find .Jl= log (II[ 1/8]) =log 8. As time goes by, the<br />

particles become equally distributed and P(k,t) = Peqm(k) = 1/8.<br />

As the result the :I{ quantity vanishes. It can be shown that, in<br />

accordance with Figure 25, the decrease in the value of :H proceeds<br />

in a uniform fashion. (The demonstration is given in all<br />

textbooks dealing with the theory of stochastic processes.)<br />

This is why :H plays the role of -S, entropy. The uniform decrease<br />

of .H has a very simple meaning: it measures the progressive<br />

uniformization of the system. The initial information<br />

is lost, and the system evolves from "order" to "disorder."<br />

Note that a Markov process implies fluctuations, as clearly<br />

indicated in Figure 24. If we would wait long enough we would<br />

recover the initial state. However, we are dealing with averages.<br />

The :JiM quantity that decreases uniformly is expressed<br />

in terms of probability distributions and not in terms of individual<br />

events. It is the probability distribution that evolves irreversibly<br />

(in the Ehrenfest model, the distribution function<br />

tends uniformly to a binomial distribution). Therefore, on the<br />

level of distribution functions, Markov chains lead to a onewaynss<br />

in time.<br />

This arrow of time marks the difference between Markov<br />

chains and temporal evolution in quantum mechanics, where<br />

the wave function, though related to probabilities, evolves reversibly.<br />

It also illustrates the close relation between stochastic<br />

processes, such as Markov chains, and irreversibility.<br />

However, the increasing of entropy (or decreasing of Jf) is not<br />

based on an arrow of time present in the laws of nature but on<br />

our decision to use present knowledge to predict future (and<br />

not past) behavior. Gibbs states it in his usual lapidary manner:<br />

But while the distinction of prior and subsequent events<br />

may be immaterial with respect to mathematical fictions,<br />

it is quite otherwise with respect to the events of the real<br />

world. It should not be f<strong>org</strong>otten, when our ensembles<br />

are chosen to illustrate the probabilities of events in the<br />

real world, that while the probabilities of subsequent<br />

events may often be determined from the probabilities of<br />

prior events, it is rarely the case that probabilities of prior


239<br />

THE CLASH OF DOCTRINES<br />

events can be determined from those of subsequent<br />

events, for we are rarely justified in excluding the consideration<br />

of the antecedent probability of the prior events. 7<br />

It is an important point, which has led to a great deal of discussion.<br />

8 Probability calculus is indeed time-oriented. The prediction<br />

of the future is different from retrodiction. If this was<br />

the whole story, we would have to conclude that we are forced<br />

to accept a subjective interpretation of irreversibility, since the<br />

distinction between future and past would depend only on us.<br />

In other words, in the subjective interpretation of iri;"eversibility<br />

(further reinforced by the ambiguous analogy with information<br />

theory), the observer is responsible for the temporal<br />

asymmetry characterizing the system's development. Since<br />

the observer cannot in a single glance determine the positions<br />

and velocities of all the particles composing a complex system,<br />

he cannot know the instantaneous state that simultaneously<br />

contains its past and its future, nor can he grasp the<br />

reversible law that would allow him to predict its developments<br />

from one moment to the next. Neither can he manipulate the<br />

system like the demon invented by Maxwell, who can separate<br />

fast- and slow-moving particles and impose on a system an<br />

antithermodynamic evolution toward an increasingly less uniform<br />

temperature distribution.9<br />

Thermodynamics remains the science of complex systems;<br />

but, from this perspective, the only specific feature of complex<br />

systems is that our knowledge of them is limited and that our<br />

uncertainty increases with time. Instead of recognizing in irreversiblity<br />

something that links nature to the observer, the scientist<br />

is compelled to admit that nature merely mirrors his<br />

ignorance. Nature is silent; irreversibility, far from rooting us<br />

in the physical world, is merely the echo of human endeavor<br />

and of its limits.<br />

However, one immediate objection can be raised. According<br />

to such interpretations, thermodynamics ought to be as universal<br />

as our ignorance. There should exist only irreversible<br />

processes. This is the stumbling block for all universal interpretations<br />

of entropy that concentrate on our ignorance of initial<br />

(or boundary) conditions. Irreversibility is not a universal<br />

propercy. In order to link dynamics and thermodynamics, a


ORDER OUT OF CHAOS 240<br />

physical criterion is required to distinguish between reversible<br />

and irreversible processes.<br />

We shall take up this question in Chapter IX. Here let us<br />

return to the history of science and Boltzmann's pioneering<br />

work.<br />

Boltzn1anns Breakthrough<br />

Boltzmann's fundamental contribution dates from 1872, about<br />

thirty years before the discovery of Markov chains. His ambition<br />

was to derive a "mechanical" interpretation of entropy. In<br />

other words, while in Markov chains the transition probabilities<br />

are given from outside as, for example, in the Ehrenfest<br />

model, we now have to relate them to the dynamic behavior of<br />

the system. Boltzmann was so fascinated by this problem that<br />

he devoted most of his scientific life to it. In his Populiire<br />

Schriften10 he wrote: "If someone asked me what name we<br />

should give to this century, I would answer without hesitation<br />

that this is the century of Darwin." Boltzmann was deeply attracted<br />

by the idea of evolution, and his ambition was to become<br />

the "Darwin" of the evolution of matter.<br />

The first step toward the mechanistic interpretation of entropy<br />

was to reintroduce the concept of "collisions" of molecules<br />

or atoms into the physical description, and along with it<br />

the possibility of a statistical description. This step had been<br />

taken by Clausius and Maxwell. Since collisions are discrete<br />

events, we may count them and estimate their average frequency.<br />

We may also classify collisions-for example, distinguish<br />

between collisions producing a particle with a given<br />

velocity v and collisions destroying a particle with a velocity v,<br />

producing molecules with a different velocity (the "direct"<br />

and "inverse" collisions); l l<br />

The question Maxwell asked was whether it was possible to<br />

define a state of a gas such that the collisions that incessantly<br />

modify the velocities of the molecules no longer determine<br />

any evolution in the distribution of these velocities-that is, in<br />

the mean number of particles for each velocity value. What is<br />

the velocity distribution such that the effects of the different<br />

collisions compensate each other on the population scale?


241 THE CLASH OF DOCTRINES<br />

Maxwell demonstrated that this particular state, which is<br />

the thermodynamic equilibrium state, occurs when the velocity<br />

distribution becomes the well-known "bell-shaped curve,"<br />

the "gaussian," which Quetelet, the founder of "social physics,"<br />

had considered to be the very expression of randomness.<br />

Maxwell's theory permits us to give a simple interpretation of<br />

some of the basic laws describing the behavior of gases. An<br />

increase in temperature corresponds to an increase in the<br />

mean velocity of the molecules and thus of the energy associated<br />

with their motion. Experiments have verified Maxwell's<br />

law with great accuracy, and it still provides a basis for the<br />

solution of numerous problems in physical chemistry (for example,<br />

the calculation of the number of collisions in a reactive<br />

mixture).<br />

Boltzmann, however, wanted to go farther. He wanted to<br />

describe not only the state of equilibrium but also evolution<br />

toward equilibrium-that is, evolution toward the Maxwellian<br />

distribution. He wanted to discover the molecular mechanism<br />

that corresponds to the increase of entropy, the mechanism<br />

that drives a system from an arbitrary distribution of velocities<br />

toward equilibrium.<br />

Characteristically, Boltzmann approached the question of<br />

physical evolution not at the level of individual trajectories but<br />

at the level of a population of molecules. This, Boltzmann felt,<br />

was virtually tantamount to accomplishing Darwin's feat, but<br />

this time in physics: the driving force behind biological evolution-natural<br />

selection-cannot be defined for one individual<br />

but only for a large population. It is therefore a statistical concept.<br />

Boltzmann's result may be described in relatively simple<br />

terms. The evolution of the distribution function f ( v, t) of the<br />

velocities v in some region of space and at time t appears as<br />

the sum of two effects; the number of particles at any given<br />

time t having a velocity v varies both as the result of the free<br />

motion of the particles and as the result of collisions between<br />

particles. The first result can be easily calculated in the terms<br />

of classical dynamics. It is in the investigation of the second<br />

result, due to collisions, that the originality of Boltzmann's<br />

method lies. In the face of the difficulties involved in following<br />

the trajectories (including the interactions), Boltzmann came<br />

to use concepts similar to those outlined in Chapter V (in con-


ORDER OUT OF CHAOS 242<br />

nection with chemical reactions) and to calculate the average<br />

number of collisions creating or destroying a molecule corresponding<br />

to a velocity v.<br />

Here once again there are two processes with opposite<br />

effects-"direct" collisions, those producing a molecule with<br />

velocity v starting from two molecules with velocities v ' and<br />

v " , and "inverse" collisions, in which a molecule with velocity<br />

v is destroyed by collision with a molecule with velocity v "' .<br />

As with chemical reactions (see Chapter V, section 1), the frequency<br />

of such events is evaluated as being proportional to the<br />

product of the number of molecules taking part in these processes.<br />

(Of course, historically speaking, Boltzmann's method<br />

[1872] preceded that of chemical kinetics.)<br />

The results obtained by Boltzmann are quite similar to those<br />

obtained in Markov chains. Again we shall introduce an J{<br />

quantity, this time referring to the velocity distribution f. It<br />

may be written J{= f flog f dv. Once again, this quantity can<br />

only decrease in time until equilibrium is reached and the velocity<br />

distribution becomes the equilibrium Maxwellian distribution.<br />

In recent years there have been numerous numerical verifications<br />

of the uniform decrease of J{ with time. All of them<br />

confirm Boltzmann's prediction. Even today, his kinetic equation<br />

plays an important role in the physics of gases: transport<br />

coefficients such as those characterizing heat conductivity or<br />

diffusion can be calculated in good agreement with experimental<br />

data.<br />

However, it is from the conceptual standpoint that Boltzmann's<br />

achievement is greatest: the distinction between reversible<br />

and irreversible phenomena, which, as we have seen,<br />

underlies the second law, is now transposed onto the microscopic<br />

level. The change of the velocity distribution due to<br />

free motion corresponds to the reversible part, while the contribution<br />

due to collisions corresponds to the irreversible part.<br />

For Boltzmann this was the key to the microscopic interpretation<br />

of entropy. A principle of molecular evolution had been<br />

produced! It is easy to understand the fascination this discovery<br />

exerted on the physicists who followed Boltzmann, including<br />

Planck, Einstein, and Schrodinger. t2<br />

Boltzmann's breakthrough was a decisive step in the direc-


243 THE CLASH OF DOCTRINES<br />

tion of the physics of processes. What determines temporal<br />

evolution in Boltzmann's equation is no longer the Hamiltonian,<br />

depending on the type of forces; now, on the contrary,<br />

functions associated with the processes-for example, the<br />

cross section of scattering-will generate motion. Can we conclude<br />

that the problem of irreversibility has been solved, that<br />

Boltzmann's theory has reduced entropy to dynamics? The<br />

answer is clear: No, it has not. Let us have a closer look at this<br />

question.<br />

Questioning Boltzmanns Interpretation<br />

As soon as Boltzmann's fundamental paper appeared in 1872,<br />

objections were raised. Had Boltzmann really "deduced" irreversibility<br />

from dynamics? How could the reversible laws of<br />

trajectories lead to irreversible evolution? Is Boltzmann's kinetic<br />

equation in any way compatible with dynamics? It is<br />

easy to see that the symmetry present in Boltzmann's equation<br />

is in contradiction with the symmetry of classical mechanics.<br />

We have already seen that velocity inversion (v -v) produces<br />

in classical dynamics the same effect as time inversion<br />

(t-t). This is a basic symmetry of classical dynamics, and<br />

we would expect that Boltzmann's kinetic equation, which describes<br />

the time change of the distribution function, would<br />

share this symmetry. But this is not so. The collision term<br />

calculated by Boltzmann remains invariant with respect to velocity<br />

inversion. There is a simple physical reason for this.<br />

Nothing in Boltzmann's picture distinguishes a collision that<br />

proceeds toward the future from a collision proceeding toward<br />

the past. This is the basis of Poincare's objection to Boltzmann's<br />

derivation. A correct calculation can never lead to<br />

conclusions that contradict its premises.B· 14 As we have seen,<br />

the symmetry properties of the kinetic equation obtained by<br />

Boltzmann for the distribution function contradict those of dynamics.<br />

Boltzmann cannot, therefore, have "deduced" entropy<br />

from dynamics. He must have introduced something<br />

new, something foreign to dynamics. Thus his results ;an rep-


ORDER OUT OF CHAOS<br />

244<br />

resent at best only a phenomenological model that, however<br />

useful, has no direct relation with dynamics. This was also the<br />

objection that Zermelo (1896) brought against Boltzmann.<br />

Loschmidt's objection, on the other hand, makes it possible<br />

to determine the limits of validity of Boltzmann's kinetic<br />

model. In fact, Loschmidt observed (1876) that this model can<br />

no longer be valid after a reversal of the velocities corresponding<br />

to the transformation v-+-v.<br />

Let us explain this by means of a thought experiment. We<br />

start with a gas in a nonequilibrium condition and let it evolve<br />

till t0• We then invert the velocities. The system reverts to its<br />

past state. As a consequence, Boltzmann's entropy is the<br />

same at t=O and at t=2t0•<br />

We may multiply such thought experiments. Start with a<br />

mixture of hydrogen and oxygen; after some time water will<br />

appear. If we invert the velocities, we should go back to an<br />

initial state with hydrogen and oxygen and no water.<br />

It is interesting that in laboratory or computer experiments,<br />

we actually can perform a velocity inversion. For example, in<br />

Figures 26 and 27, Boltzmann's J{ quantity has been calculated<br />

for two-dimensional hard spheres (hard disks), starting<br />

first collision<br />

ae<br />

•<br />

•<br />

0 20 40 TIME 60<br />

Figure 26. Evolution of .Jf with time for N "hard spheres" by computer<br />

simulation; (a) corresponds to N=100, (b) to N=484, (c) to N=1225.


..<br />

245 THE CLASH OF DOCTRINES<br />

with disks on lattice sites with an isotropic velocity distribution.<br />

The results follow Boltzmann's predictions.<br />

If, after fifty or a hundred collisions, corresponding to about<br />

I0-6 sec in a dilute gas, the velocities are inverted, a new ensemble<br />

is obtained.15 Now, after the velocity inversion, Boltzmann's<br />

J{ quantity increases instead of decreasing.<br />

I<br />

•<br />

•<br />

•<br />

-<br />

•<br />

. . •<br />

-<br />

. -<br />

•<br />

N:IOO<br />

•<br />

•<br />

•<br />

...<br />

-<br />

• u<br />

•<br />

•<br />

•• •<br />

• 0<br />

• .<br />

'\<br />

on<br />

•<br />

• •<br />

•<br />

y<br />

<br />

0<br />

-. . . ....<br />

.. . - .<br />

.<br />

·<br />

·: : :" ... ..<br />

. ... . . .. . 2 '"·<br />

_,. •<br />

.<br />

••• t •<br />

, .. .<br />

... •<br />

. ...<br />

.<br />

..<br />

-<br />

t<br />

' -<br />

.<br />

-'"'<br />

.... , ·· ... .<br />

.... .,<br />

·11-------::-----------------<br />

equit .<br />

...<br />

:<br />

0 20 TIME 60<br />

Figure 27. Evolution of :H when velocities are inverted after 50 or 100<br />

collisions. Simulation with 100 "hard spheres."<br />

A similar situation can be produced in spin echo experiments<br />

or plasma echo experiments. There also, over limited<br />

periods of time, an "antithermodynamic" behavior in Boltzmann's<br />

sense may be observed.<br />

But it is important to note that the velocity inversion experiment<br />

becomes increasingly more difficult when the time interval<br />

t0 after which the inversion occurs is increased.<br />

To be able to retrace its past, the gas must remember everything<br />

that happened to it during the time interval from 0 to t0•<br />

There must be "storage" of information. We can express this<br />

storage in terms of correlations between particles. We shall<br />

come back to the question of correlations in Chapter IX. Let<br />

us only mention here that it is precisely this relation between<br />

correlations and collisions that is the element missing from


ORDER OUT OF CHAOS 246<br />

Boltzmann's considerations. When Loschmidt confronted him<br />

with this, Boltzmann had to accept that there was no way out:<br />

the collisions occurring in the opposite direction "undo" what<br />

was done previously, and the system has to revert to its initial<br />

state. Therefore, the function J{ must also increase until it<br />

again reaches its initial value. Velocity inversion thus calls for<br />

a distinction between the situations to which Boltzmann's reasoning<br />

applies and those to which it does not.<br />

Once the problem was stated (1894), it was easy to identify<br />

the nature of this limitation.l6, 17 The validity of Boltzmann's<br />

statistical procedure depends on the assumption that before<br />

they collide, the molecules behave independently of one another.<br />

This constitutes an assumption about the initial conditions,<br />

called the "molecular chaos" assumption. The initial<br />

conditions created by a velocity inversion do not conform to<br />

this assumption. If the system is made "to go backward in<br />

time," a new "anomalous" situation is created in the sense<br />

that certain molecules are then "destined" to meet at a predeterminable<br />

instant and to undergo a predetermined change<br />

of velocity at this time, however far apart they may be at the<br />

instant of velocity inversion.<br />

· Velocity inversion thus creates a highly <strong>org</strong>anized system,<br />

and thus the molecular chaos assumption fails. The various<br />

collisions produce, as if by a preestablished harmony, an apparently<br />

purposeful behavior.<br />

But there is more. What does the transition from order to<br />

disorder signify? In the Ehrenfest urn experiment, it is clearthe<br />

system will evolve till uniformity is reached. But other situations<br />

are not so clear; we may do computer experiments in<br />

which interacting particles are initially distributed at random.<br />

In time a lattice is formed. Do we still move from order to<br />

disorder? The answer is not obvious. To understand order and<br />

disorder we first have to define the objects in terms of which<br />

these concepts are used. Moving from dynamic to thermodynamic<br />

objects is easy in the case of dilute gases-as shown by<br />

the work of Boltzmann. However, it is not so easy in the case<br />

of dense systems whose molecules interact.<br />

Because of such difficulties, Boltzmann's creative and pioneering<br />

work remained incomplete.


247 THE CLASH OF DOCTRINES<br />

Dynarnics and Thermodynamics:<br />

Two Separate Worlds<br />

We already noted that trajectories are incompatible with the<br />

idea of irreversibility. However, the study of trajectories is not<br />

the only way in which we can give a formulation of dynamics.<br />

There is also the theory of ensembles introduced by Gibbs and<br />

Einstein,6. 18 which is of special interest in the case of systems<br />

formed by a large number of molecules. The essential new<br />

element in the Gibbs-Einstein ensemble theory is that we can<br />

formulate the dynamic theory independently of any precise<br />

specification of initial conditions.<br />

The theory of ensembles represents dynamic systems in<br />

"phase space." The dynamic state of a point particle is specified<br />

by position (a vector with three components) and by momentum<br />

(also a vector with three components). We may<br />

represent this state by two points, each in a three-dimensional<br />

space, or by a single point in the six-dimensional space formed<br />

by the coordinates and momenta. This is the phase space.<br />

This geometric representation can be extended to an arbitrary<br />

system formed by n particles. We then need n x 6 numbers to<br />

specify the state of the system, or alternatively we may specify<br />

this system by a single point in the 6n-dimensional phase<br />

space. The evolution in time of such a system will then be<br />

described by a trajectory in the phase space.<br />

It has already been stated that the exact initial conditions of<br />

a macroscopic system are never known. Nevertheless, nothing<br />

prevents us from representing this system by an "ensemble"<br />

of points-namely, the points corresponding to the various dynamic<br />

states ompatible with the information we have concerning<br />

the system. Each region of phase space may contain an<br />

infinite number of representative points, the density of which<br />

measures the probability of actually finding the system in this<br />

region. Instead of introducing an infinity of discrete points, it<br />

is more convenient to introduce a continuous density of representative<br />

points in the phase space. We shall call p (q1<br />

• • • q 3 0,<br />

p1 • • • p30) this density in phase space where q1 ,q2 • • • q3n are<br />

the coordinates of the n points; similarly, p1 ,p2 • • • p30 are the<br />

momenta (each point has three coordinates and three mo-


ORDER OUT OF CHAOS 248<br />

menta). This density measures the probability of finding a dynamic<br />

system around the point ql ... q3n,PI • • • P3n in phase<br />

space.<br />

Presented in such a way, the density function p may appear<br />

as an idealization, an artificial construct, whereas the trajectory<br />

of a point in phase space would correspond "directly" to<br />

the description of "natural" behavior. But in fact it is the<br />

point, not the density, that corresponds to an idealization. Indeed,<br />

we never know an initial state with the infinite degree of<br />

precision that would reduce a region in phase space to a single<br />

point; we can only determine an ensemble of trajectories starting<br />

from the ensemble of representative points corresponding<br />

to what we know about the initial state of the system. The<br />

density function p represents knowledge about a system, and<br />

the more accurate this knowledge, the smaller the region in the<br />

phase space where the density function is different from zero<br />

and where the system may be found. Should the density function<br />

everywhere have a uniform value, we would know nothing<br />

about the state of the system. It might be in any of the possible<br />

states compatible with its dynamic structure.<br />

From this perspective, a point thus represents the maximum<br />

knowledge we can have about a system. It is the result of a<br />

limiting process, the result of the ever-growing precision of our<br />

knowledge. As we shall see in Chapter IX, a fundamental<br />

problem will be to determine when such a limiting process is<br />

really possible. Through increased precision, this process<br />

means we go from a region where the density function p is<br />

different from zero to another, smaller region inside the first.<br />

We can continue this until the region containing the system<br />

becomes arbitrarily small. But as we shall see, we must be<br />

cautious: arbitrarily small does not mean zero, and it is not<br />

certain a priori that this limiting process will lead to the possibility<br />

of consistently predicting a single well-defined trajectory.<br />

The introduction of the theory of ensembles by Gibbs and<br />

Einstein was a natural continuation of Boltzmann's work. In<br />

this perspective the density function p in phase space replaces<br />

the velocity distribution function f used by Boltzmann. However,<br />

the physical content of p exceeds that off. Just like/, the<br />

density function p determines the velocity distribution, but it<br />

also contains other information, such as the probability of


29<br />

THE CLASH OF DOCTRINES<br />

meeting two particles a certain distance apart. In particular,<br />

correlations between particles, which we discussed in the preceding<br />

section, are now included in the density function p. In<br />

fact, this function contains the complete information about all<br />

statistical features of the n-b9dy system.<br />

We must now describe the evolution of the density function<br />

in phase space. At first sight, this appears to be an even more<br />

ambitious task than the one Boltzmann set himself for the velocity<br />

distribution function. But this is not the case. The Hamiltonian<br />

equations discussed in Chapter 11 allow us to obtain<br />

an exact evolution equation for p without any further approximations.<br />

This is the so-called Liouville equation, to which we<br />

shall return in Chapter IX. Here we wish merely to point out<br />

, that the properties of Hamiltonian dynamics imply that the<br />

evolution of the density function p in phase space is that of an<br />

incompressible fluid. Once the representative points occupy a<br />

region of volume V in phase space, this volume remains constant<br />

in time. The shape of the region may be deformed in an<br />

arbitrary way, but the value of the volume remains the same.<br />

Gibbs' theory of ensembles thus permits a rigorous combination<br />

of the statistical point of view (the study of the "population"<br />

described by p) and the laws of dynamics. It also permits<br />

a more accurate representation of the thermodynamic equilibrium<br />

state. Thus, in the case of an isolated system, the ensemp<br />

Figure 28. Time evolution in the phase space of a "volume" containing the<br />

representative points of a system: the volume is conserved while the shape<br />

is modified. The position in phase space is specified by coordinate& q and<br />

momentum p.<br />

q


ORDER OUT OF CHAOS<br />

250<br />

ble ci representative points corresponds to systems that all have<br />

the same energy E. The density p will differ from zero only on<br />

the "microcanonical surface" corresponding to the specified<br />

value of the energy in phase space. Initially, the density p may<br />

be distributed arbitrarily over this surface. At equilibrium, p<br />

must no longer vary with time and has to be independent of<br />

the specific initial state. Thus the approach to equilibrium has<br />

a simple meaning in terms of the evolution of p. The distribution<br />

function p becomes uniform over the microcanonical surface.<br />

Each ci the points on this surface has the same probability<br />

of actually representing the system. This corresponds to the<br />

"microcanonical ensemble."<br />

Does the theory of ensembles bring us any closer to the<br />

solution of the problem of irreversibility? Boltzmann's theory<br />

describes thermodynamic entropy in terms of the velocity distribution<br />

function f. He achieved this result through the introduction<br />

of his J{ quantity. As we have seen, the system evolves<br />

in time until the Maxwellian distribution is reached, while,<br />

during this evolution, the quantity J-{ decreases uniformly. Can<br />

we now, in a more general fashion, take the evolution of the<br />

distribution p in phase space toward the microcanonical ensemble<br />

as the basis for entropy increase? Would it be enough<br />

to replace Boltzmann's quantity J{ expressed in terms ofjby a<br />

"Gibbsian" quantity 3£0 defined in exactly the same way, but<br />

this time in terms of p? Unfortunately, the answer to both<br />

questions is "No." If we use the Liouville equation, which<br />

describes the evolution of the density phase space P. and take<br />

into account the conservation of volume in phase space we<br />

have mentioned, the conclusion is immediate: :Jf0 is a constant<br />

and thus cannot represent entropy. With respect to Boltzmann,<br />

this appears as a step backward rather than forward!<br />

Though it is negative, Gibbs' conclusion remains very important.<br />

We have already discussed the ambiguity of the ideas<br />

of order and disorder. What the constancy of 3£0 tells us is<br />

that there is no change of order whatsoever in the frame of<br />

dynamic theory! The "information" expressed by 3£0 remains<br />

constant. This can be understood as follows: we have seen that<br />

collisions introduce correlations. From the perspective of velocities,<br />

the result of collisions is randomization; therefore we<br />

can describe this process as a transition from order to disor-


251 THE CLASH OF DOCTRINES<br />

der, but the appearance of correlations as the result of collision<br />

points in the opposite direction, toward a transition from disorder<br />

to order! Gibbs' result shows that the two effects exactly<br />

cancel each other.<br />

We come, therefore, to an important conclusion. Whatever<br />

representation we use, be it the idea of trajectories or the<br />

Gibbs-Einstein ensemble theory, we will never be able to deduce<br />

a theory of irreversible processes that will be valid for<br />

every system that satisfies the laws of classical (or quantum)<br />

dynamics. There isn't even a way to speak of a transition from<br />

order to disorder! How should we understand these negative results?<br />

Is any theory of irreversible processes in absolute conflict<br />

with dynamics (classical or quantum)? It has often been<br />

proposed that we include some cosmological terms that would<br />

express the influence of the expanding universe on the equations<br />

of motion. Cosmological terms would ultimately provide<br />

the arrow of time. However, this is difficult to accept. On the<br />

one hand, it is not clear how we should add these cosmological<br />

terms; on the other, precise dynamic experiments seem to rule<br />

out the existence of such terms, at least on the terrestrial scale<br />

with which we are concerned here (think, for example, about<br />

the precision of space trip experiments, which confirm Newton's<br />

equations to a high degree). On the other hand, as we<br />

have already stated, we live in a pluralistic universe in which<br />

reversible and irreversible processes coexist, all embedded in<br />

the expanding universe.<br />

An even more radical conclusion is to affirm with Einstein<br />

that time as irreversibility is an illusion that will never find a<br />

place in the objective world of physics. Fortunately there is<br />

another way out, which we shall explore in Chapter IX. Irreversibility,<br />

as has been repeatedly stated, is not a universal<br />

property. Therefore , no general derivation of irreversibility from<br />

dynamics is to be expected.<br />

Gibbs' theory of ensembles introduces only one additional<br />

element with respect to trajectory dynamics, but a very important<br />

one-our ignorance of the precise initial conditions. It is<br />

unlikely that this ignorance alone leads to irreversibility.<br />

We should therefore not be astonished at our failure. We<br />

have not yet formulated the specific features that a dynamic<br />

system has to possess to lead to irreversible processes.


ORDER OUT OF CHAOS 252<br />

Why have so many scientists accepted so readily the subjective<br />

interpretation of irreversibility? Perhaps part of its attraction<br />

lies in the fact that, as we have seen, the irreversible<br />

increase of entropy was at first associated with imperfect manipulation,<br />

with our lack of control over operations that are<br />

ideally reversible.<br />

But this interpretation becomes absurd as soon as the irrelevant<br />

associations with technological problems are set aside. We<br />

must remember the context that gave the second law its significance<br />

as nature's arrow of time. According to the subjective<br />

interpretation, chemical affinity, heat conduction, viscosity,<br />

all the properties connected with irreversible entropy production<br />

would depend on the observer. Moreover, the extent to<br />

which phenomena of <strong>org</strong>anization originating in irreversibility<br />

play a role in biology makes it impossible to consider them as<br />

simple illusions due to our ignorance. Are we ourselves-living<br />

creatures capable of observing and manipulating-mere<br />

fictions produced by our imperfect senses? Is the distinction<br />

between life and death an illusion?<br />

Thus recent developments in thermodynamic theory have<br />

increased the violence of the conflict between dynamics and<br />

thermodynamics. Attempts to reduce the results of thermodynamics<br />

to mere approximations due to our imperfect knowledge<br />

seem wrong headed when the constructive role of entropy<br />

is understood and the possibility of an amplification of fluctuations<br />

is discovered. Conversely, it is difficult to reject dynamics<br />

in the name of irreversibility: there is no irreversibility in<br />

the motion of an ideal pendulum. Apparently there are two<br />

conflicting worlds, a world of trajectories and a world of processes,<br />

and there is no way of denying one by asserting the<br />

other.<br />

To a certain extent, there is an analogy between this conflict<br />

and the one that gave rise to dialectical materialism. We have<br />

described in Chapters V and VI a nature that might be called<br />

"historical"-that is, capable of development and innovation.<br />

The idea of a history of nature as an integral part of materialism<br />

was asserted by Mar x and, in greater detail, by Engels.<br />

Contemporary developments in physics, the discovery of the<br />

constructive role played by irreversibility, have thus raised<br />

within the natural sciences a question that has long been asked<br />

by materialists. For them, understanding nature meant under-


253<br />

THE CLASH OF DOCTRINES<br />

standing it as being capable of producing man and his societies.<br />

Moreover, at the time Engels wrote his Dialectics of Nature,<br />

the physical sciences seemed to have rejected the mechanistic<br />

world view and drawn closer to the idea of an historical<br />

development of nature. Engels mentions three fundamental<br />

discoveries: energy and the laws governing its qualitative<br />

transformations, the cell as the basic constituent of life, and<br />

Darwin's discovery of the evolution of species. In view of<br />

these great discoveries, Engels came to the conclusion that the<br />

mechanistic world view was dead.<br />

But mechanicism remained a basic difficulty facing dialectical<br />

materialism. What are the relations between the general<br />

laws of dialectics and the equally universal laws of mechanical<br />

motion? Do the latter "cease" to apply after a certain stage has<br />

been reached, or are they simply false or incomplete? To come<br />

back to our previous question, how can the world of processes<br />

and the world of trajectories ever be linked together?I9<br />

However, while it is easy to criticize the subjectivistic interpretation<br />

of irreversibility and to point out its weakness, it is<br />

not so easy to go beyond it and formulate an "objective" theory<br />

of irreversible processes. The history of this subject has<br />

some dramatic overtones. Many people believe that it is the<br />

recognition of the basic difficulties involved that may have led<br />

to Boltzmann's suicide in 1906.<br />

Boltzmann and the Arrow of Time<br />

As we have noted, Boltzmann at first thought that he could<br />

prove that the arrow of time was determined by the evolution<br />

of dynamic systems toward states of higher probability or a<br />

higher number of complexions: there would be a one-way increase<br />

of the number of complexions with time. We have also<br />

discussed the objections of Poincare and Zermelo. Poincare<br />

proved that every closed dynamic system reverts in time toward<br />

its previous state. Thus, all states are forever recurring.<br />

How could such a thing as an "arrow of time" be associated<br />

with entropy increase? This led to a dramatic change in Boltzmann's<br />

attitude. He abandoned his attempt to prove that an objective<br />

arrow of time exists and introduced instead an idea that,


ORDER OUT OF CHAOS<br />

254<br />

in a sense, reduced the law of entropy increase to a tautology.<br />

Now he argued that the arrow of time is only a convention that<br />

we (or perhaps all living beings) introduce into a world in<br />

which there is no objective distinction between past and future.<br />

Let us cite Boltzmann's reply to Zermelo:<br />

We have the choice of two kinds of picture. Either we<br />

assume that the whole universe is at the present moment<br />

in a very improbable state. Or else we assume that the<br />

aeons during which this improbable state lasts, and the<br />

distance from here to Sirius, are minute if compared with<br />

the age and size of the whole universe. In such a universe,<br />

which is in thermal equilibrium as a whole and<br />

therefore dead, relatively small regions of the size of our<br />

galaxy will be found here and there; regions (which we<br />

may call "worlds") which deviate significantly from thermal<br />

equilibrium for relatively short stretches of those<br />

"aeons" of time. Among these worlds the probabilities of<br />

their state (i.e. the entropy) will increase as often as they<br />

decrease. In the universe as a whole the two directions of<br />

time are indistinguishable, just as in space there is no up<br />

or down. However, just as at a certain place on the earth's<br />

surface we can call "down" the direction towards the<br />

centre of the earth, so a living <strong>org</strong>anism that finds itself in<br />

such a world at a certain period of time can define the<br />

"direction" of time as going from the less probable state<br />

to the more probable one (the former will be the "past"<br />

and the latter the "future"), and by virtue of this definition<br />

he will find that his own small region, isolated from<br />

the rest of the universe, is "initially" always in an improbable<br />

state. It seems to me that this way of looking at<br />

things is the only one which allows us to understand the<br />

validity of the second law, and the heat death of each individual<br />

world, without invoking a unidirectional change of<br />

the entire universe from a definite initial state to a final<br />

state.20<br />

Boltzmann's idea can be made clearer by referring to a diagram<br />

proposed by Karl Popper (Figure 29). The arrow of time<br />

would be as arbitrary as the vertical direction determined by<br />

the gravitational field.


255 THE CLASH OF DOCTRINES<br />

Arruw of tiM 1Jf c.hia<br />

•trett"h o£ tiM only<br />

ArrllW of ti• or this<br />

atrf'tch ,,, ti•• onlr<br />

t.qui liltri ... level<br />

tnt ropy c-urv• clettr•i ni.aa<br />

the dir•rtion of ti•<br />

Figure 29. Popper's schematic representation of Boltzmann's cosmological<br />

interpretation of the arrow of time (see text).<br />

Commenting on Boltzmann's text, Popper has written:<br />

I think that Boltzmann's idea is staggering in its boldness<br />

and beauty. But I also think that it is quite untenable, at<br />

least for a realist. It brands unidirectional change as an<br />

illusion. This makes the catastrophe of Hiroshima an illusion.<br />

Thus it makes our world an illusion, and with it all<br />

our attempts to find out more about our world. It is therefore<br />

self-defeating (like every idealism). Boltzmann's idealistic<br />

ad hoc hypothesis clashes with his own realistic<br />

and almost passionately maintained anti-idealistic philosophy,<br />

and with his passionate wish to know.21<br />

We fully agree with Popper's comments, and we believe that<br />

it is time to take up Boltzmann's task once again. As we have<br />

said, the twentieth century has seen a great conceptual revolution<br />

in theoretical physics, and this has produced new hopes<br />

for the unification of dynamics and thermodynamics. We are<br />

now entering a new era in the history of time, an era in which<br />

both being and becoming can be incorporated into a single<br />

noncontradictory vision.


CHAPTER IX<br />

IRREVERSIBILITY­<br />

THEENTROPY<br />

BARRIER<br />

Entropy and the Arrow of Time<br />

In the preceding chapter we described some difficulties in the<br />

microscopic theory of irreversible processes. Its relation with<br />

dynamics, either classical or quantum, cannot be simple, in<br />

the sense that irreversibility and its concomitant increase of<br />

entropy cannot be a general consequence of dynamics. A microscopic<br />

theory of irreversible processes will require additional,<br />

more specific conditions. We must accept a pluralistic<br />

world in which reversible and irreversible processes coexist.<br />

Yet such a pluralistic world is not easy to accept.<br />

In his Dictionnaire Philosophique Voltaire wrote the following<br />

about destiny:<br />

. . . everything is governed by immutable laws • . . everything<br />

is prearranged ... everything is a necessary<br />

effect . . .. There are some people who, frightened by<br />

this truth, allow half of it, like debtors who offer their<br />

creditors half their debt, asking for more time to pay the<br />

remainder. There are, they say, events which are necessary<br />

and others which are not. It would be strange if a<br />

part of what happens had to happen and another part did<br />

not . • . . I necessarily must have the passion to write this,<br />

and you must have the passion to condemn me; we are<br />

both equally foolish, both toys in the hands of destiny.<br />

Your nature is to do ill, mine is to love truth, and to publish<br />

it in spite of you.1<br />

257


ORDER OUT OF CHAOS 258<br />

However convincing they may sound, such a priori arguments<br />

can lead us astray. Voltaire's reasoning is Newtonian:<br />

nature always conforms to itself. But, curiously, today we find<br />

ourselves in the strange world mocked by Voltaire; we are astonished<br />

to discover the qualitative diversity presented by nature.<br />

It is not surprising that people have vacillated between the<br />

two extremes; between the elimination of irreversibility from<br />

physics, advocated by Einstein, as we have mentioned,2 and,<br />

on the contrary, the emphasis on the importance of irreversibility,<br />

as in Whitehead's concept of process. There can be no doubt<br />

that irreversibility exists on the macroscopic level and has an<br />

important constructive role, as we have shown in Chapters V<br />

and VI . Therefore there must be something in the microscopic<br />

world of which macroscopic irreversibility is the manifestation.<br />

The microscopic theory has to account for two closely<br />

related elements. First of all, we must follow Boltzmann in<br />

attempting to construct a microscopic model for entropy<br />

(Boltzmann's .J{ fu nction) that changes uniformly in time. This<br />

change has to define our arrow of time. The increase of entropy<br />

for isolated systems has to express the aging of the system.<br />

Often we may have an arrow of time without being able to<br />

associate entropy with the type of processes considered. Popper<br />

gives a simple example of a system presenting a unidirectional<br />

process and therefore an arrow of time.<br />

Suppose a film is taken of a large surface of water initially<br />

at rest into which a stone is dropped. The reversed<br />

film will show contracting, circular waves of increasing<br />

amplitude. Moreover, immediately behind the highest<br />

wave crest, a circular region of undisturbed water will<br />

close in towards the centre. This cannot be regarded as a<br />

possible classical process. It would demand a vast number<br />

of distant coherent generators of waves the coordination<br />

of which, to be explicable, would have to be shown,<br />

in the film, as originating from one centre. This, however,<br />

raises precisely the same difficulty again, if we try to reverse<br />

the amended film. 3


259<br />

IRREVERSIBILITY-THE ENTROPY BARRIER<br />

Indeed, whatever the technical means, there will always be<br />

a distance from the center beyond which we are unable to generate<br />

a contracting wave. There are unidirectional processes.<br />

Many other processes of the type presented by Popper can be<br />

imagined: we never see energy coming from all directions converge<br />

on a star, together with the backward-running nuclear<br />

reactions that would absorb that energy.<br />

In addition, there exist other arrows of time-for example,<br />

the cosmological arrow (see the excellent account by M.<br />

Gardner'). If we assume that the universe started with a Big<br />

Bang, this obviously implies a temporal order on the cosmological<br />

level. The size of the universe continues to increase,<br />

but we cannot identify the radius of the universe with entropy.<br />

Indeed, as we already mentioned, inside the expanding universe<br />

we find both reversible and irreversible processes. Similarly,<br />

in elementary-particle physics there exist processes that<br />

present the so-called T-violation. The T-violation implies that<br />

the equations describing the evolution of the system for +t are<br />

different from those describing the evolution for -t. However,<br />

this T-violation does not prevent us from including it in the<br />

usual (Hamiltonian) formulation of dynamics. No entropy<br />

fu nction can be defined as a result of the T-violation.<br />

We are reminded of the celebrated discussion between Einstein<br />

and Ritz published in 1909.5 This is a quite unusual paper,<br />

a very short one, less than a printed page long. It simply is<br />

a statement of disagreement. Einstein argued that irreversibility<br />

is a consequence of the probability concept introduced by<br />

Boltzmann. On the contrary, Ritz argued that the distinction<br />

between "retarded" and "advanced" waves plays an essential<br />

role. This distinction reminds us of Popper's argument. The<br />

waves we observe in the pond are retarded waves; they follow<br />

the dropping of the stone.<br />

Both Einstein and Ritz introduced essential elements into<br />

the discussion of irreversibility, but each of them emphasized<br />

only part of the story. We have already mentioned in Chapter<br />

VIII that probability presupposes a direction of time and<br />

therefore cannot be used to derive the arrow of time. We have<br />

also mentioned that the exclusion of processes such as advanced<br />

waves does not necessarily lead to a formulation of the<br />

second law. We need both types of arguments.


ORDER OUT OF CHAOS<br />

260<br />

Irreversibility as a Syrr1metry-Breaking Process<br />

Before discussing the problem of irreversibility, it is useful to<br />

remember how another type of symmetry-breaking, spatial<br />

symmetry-breaking, can be derived. In the equations describing<br />

reaction-diffusion systems, left and right play the same<br />

role (the diffusion equations remain invariant when we perform<br />

the space inversion r-+-r). Still, as we have seen, bifurca­<br />

.tions may lead to solutions in which this symmetry is broken<br />

(see Chapter V) . For example, the concentration of some of<br />

the components may become higher on the left than on the<br />

right. The symmetry of the equations only requires that the symmetry-breaking<br />

solutions appear in pairs.<br />

There are, of course, many reaction-diffusion equations that<br />

do not present bifurcations and that therefore cannot break<br />

spatial symmetry. The breaking of spatial symmetry requires<br />

other highly specific conditions. This is valuable for understanding<br />

temporal symmetry-breaking, in which we are primarily<br />

interested here. We have to find systems in which the<br />

equations of motion may have realizations of lower symmetry.<br />

The equations are indeed invariant in respect to time inversion<br />

t-+-t. However, the realization of these equations may<br />

correspond to evolutions that lose this symmetry. The only<br />

condition imposed by the symmetry of equations is that such<br />

realizations appear in pairs. If, for example, we find one solution<br />

going to equilibrium in the far distant future (and not in<br />

the far distant past), we should also find a solution that goes to<br />

equilibrium in the far distant past (and not in the far distant<br />

future). Symmetry-broken solutions appear in pairs.<br />

Once we find such a situation we can express the intrinsic<br />

meaning of the second law. It becomes a selection principle<br />

stating that only one of the two types of solutions can be realized<br />

or may be observed in nature. Whenever applicable, the<br />

second law expresses an intrinsic polarization of nature. It can<br />

never be the outcome of dynamics itself. It has to appear as a<br />

supplementary selection principle that when realized is propagated<br />

by dynamics. Only a few years ago it seemed impossible<br />

to attempt such a program. However, over the past few de-


261 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

cades dynamics has made remarkable progress, and we can<br />

now understand in detail how these symmetry-breaking solutions<br />

emerge in dynamic systems "of sufficient complexity"<br />

and what the selection rule expressed by the second law of<br />

thermodynamics means on the microscopic level. This is what<br />

we want to show in the next part of this chapter.<br />

The Limits of Classical Concepts<br />

Let us start with classical mechanics. As we have already<br />

mentioned, if trajectory is to be the basic irreducible element,<br />

the world would be as reversible as the trajectories out of<br />

which it is formed. In this description there would be no entropy<br />

and no arrow of time; but, as a result of unexpected re·<br />

cent developments, the validity of the trajectory concept<br />

appears far more limited than we might have expected. Let us<br />

return to Gibbs' and Einstein's theory of ensembles, introduced<br />

in Chapter VIII. We have seen that Gibbs and Einstein<br />

introduced phase space into physics to account for the fact<br />

that we do not "know" the initial state of systems formed by a<br />

large number of particles. For them, the distribution function<br />

in phase space was only an auxiliary construction expressing<br />

our de fa cto ignorance of a situation that was well determined<br />

de jure. However, the entire problem takes on new dimensions<br />

once it can be shown that for certain types of systems an infi·<br />

nitely precise determination of initial conditions leads to a<br />

self-contradictory procedure. Once this is so, the fact that we<br />

never know a single trajectory but a group, an ensemble of<br />

trajectories in phase space, is not a mere expression of the<br />

limits of our knowledge. It becomes a starting point of a new<br />

way of investigating dynamics.<br />

It is true that in simple cases there is no problem. Let us<br />

take the example of a pendulum. It may oscillate or else rotate<br />

about its axis according to the initial conditions. For it to rotate,<br />

its kinetic energy must be large enough for it not to "fall<br />

back" before reaching a vertical position. These two types of<br />

motion define two disjointed regions of phase space. The reason<br />

for this is very simple: rotation requires more energy than<br />

oscillation (see Figure 30).


ORDER OUT OF CHAOS 262<br />

y<br />

0)<br />

a<br />

v<br />

b)<br />

Figure 30. Representation of a pendulum's motion in a space where Vis<br />

the velocity and e the angle of deflection. (a) typical trajectories in (V, e)<br />

space; (b) the shaded regions correspond to oscillations; the region outside<br />

corresponds to rotations.<br />

If our measurements allow us to establish that the system is<br />

initially in a given region, we may safely predict the type of<br />

motion displayed by the pendulum. We can increase the accu-


263<br />

IRREVERSIBILITY-THE ENTROPY BARRIER<br />

racy of our measurements and localize the initial state of the<br />

pendulum in a smaller region circumscribed by the first. In<br />

any case, we know the system's behavior for all time; nothing<br />

new or unexpected is likely to occur.<br />

One of the most surprising results achieved in the twentieth<br />

century is that such a description is not valid in general. On<br />

the contrary, "most" dynamic systems behave in a quite unstable<br />

way. 6 Let us indicate one kind of trajectory (for example,<br />

that of oscillation) by + and another kind (for example,<br />

that corresponding to rotation) by *· Instead of Figure 30,<br />

where the two regions were separated, we find, in general, a<br />

mixture of states that makes the transition to a single point<br />

ambiguous (see Figure 31). If we know only that the initial<br />

state of our system is in region A, we cannot deduce that its<br />

trajectory is of type +; it may equally well be of type *· We<br />

achieve nothing by increasing the accuracy by going from region<br />

A to a smaller region within it, for the uncertainty remains.<br />

In all regions, however small, there are always states<br />

belonging to each of the two types of trajectories.1<br />

v<br />

Figure 31. Schematic representation of any region, arbitrarily small, of the<br />

phase space V tor a system presenting dynamic instability. As in the case of<br />

the pendulum, there are two types of trajectories (represented here by +<br />

and •); however, in contrast with the pendulum, both motions appear in every<br />

region arbitrary small.


265 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

that at the end of the nineteenth century Bruns and Poincare<br />

demonstrated that most dynamic systems, starting with the<br />

famous "three body" problem, were not integrable.<br />

On the other hand, the very idea of approaching equilibrium<br />

in terms of the theory of ensembles requires that we go beyond<br />

the idealization of integrable systems. As we saw in Chapter<br />

VII I, according to the theory of ensembles, an isolated system<br />

is in equilibrium when it is represented by a "microcanonical<br />

ensemble," when all points on the surface of given energy<br />

have the same probability. This means that for a system to<br />

evolve to equilibrium, energy must be the only quantity conserved<br />

during its evolution. It must be the only "invariant."<br />

Whatever the initial conditions, the evolution of the system<br />

must allow it to reach all points on the surface of given energy.<br />

But for an integrable system, energy is far from being the only<br />

invariant. In fact, there are as many invariants as degrees of<br />

freedom, since each generalized momentum remains constant.<br />

Therefore we have to expect that such a system is "imprisoned"<br />

in a very small "fraction" of the constant-energy (see<br />

Figure 32) surface formed by the intersection of all these invar­<br />

-iant surfaces.<br />

p<br />

Figure 32. Temporal evolution of a cell in phase space p, q. The "volume"<br />

of the cell and its form are maintained in time; moreover, most of the phase<br />

space is inaccessible to the system.<br />

q


265 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

that at the end of the nineteenth century Bmns and Poincare<br />

demonstrated that most dynamic systems, starting with the<br />

famous "three body" problem, were not integrable.<br />

On the other hand, the very idea of approaching equilibrium<br />

in terms of the theory of ensembles requires that we go beyond<br />

the idealization of integrable systems. As we saw in Chapter<br />

VIII, according to the theory of ensembles, an isolated system<br />

is in equilibrium when it is represented by a "microcanonical<br />

ensemble," when all points on the surface of given energy<br />

have the same probability. This means that for a system to<br />

evolve to equilibrium, energy must be the only quantity conserved<br />

during its evolution. It must be the only "invariant."<br />

Whatever the initial conditions, the evolution of the system<br />

must allow it to reach all points on the surface of given energy.<br />

But for an integrable system, energy is far from being the only<br />

invariant. In fact, there are as many invariants as degrees of<br />

freedom, since each generalized momentum remains constant.<br />

Therefore we have to expect that such a system is "imprisoned"<br />

in a very small "fraction" of the constant-energy (see<br />

Figure 32) surface formed by the intersection of all these invariant<br />

.<br />

surfaces.<br />

p<br />

..--- ...<br />

- ...<br />

1),-/<br />

. , ,<br />

<br />

," , ---""'04\<br />

"<br />

, I \ \<br />

I ,<br />

I ,<br />

, \<br />

I<br />

\<br />

\ , 'I<br />

, rs-'" ,<br />

I<br />

\ \ , "<br />

/<br />

, /<br />

,<br />

,/<br />

"<br />

'"<br />

..... _..-.,-.<br />

Figure 32. Temporal evolution of a cell in phase space p, q. The "volume"<br />

of the cell and its form are maintail1ed in time; moreover, most of the phase<br />

space is inaccessible to the system.<br />

q


ORDER OUT OF CHAOS 266<br />

To avoid these difficulties, Maxwell and Boltzmann introduced<br />

a new, quite different type of dynamic system. For these<br />

systems energy would be the only invariant. Such systems are<br />

called "ergodic" systems (see Figure 33).<br />

Great contributions to the theory of ergodic systems have<br />

been made by Birchoff, von Neumann, Hopf, Kolmogoroff,<br />

and Sinai, to mention only a few. s. 9. tO<br />

Today we know that<br />

there are large classes of dynamic (though non-Hamiltonian)<br />

systems that are ergodic. We also know that even relatively<br />

simple systems may have properties stronger than ergodicity.<br />

For these systems, motion in phase space becomes highly chaotic<br />

(while always preserving a volume that agrees with the<br />

Liouville equation mentioned in Chapter VII).<br />

p<br />

q<br />

Figure 33. Typical evolution in phase space of a cell corresponding to an<br />

ergodic system. Time going on, the "volume" and the form are conserved<br />

but the cell now spirals through the whole phase space.


267<br />

IRREVERSIBILITY-THE ENTROPY BARRIER<br />

Suppose that our knowledge of initial conditions permits us<br />

to localize a system in a small cell of the phase space. During<br />

its evolution, we shall see this initial cell twist and turn and,<br />

like an amoeba, send out "pseudopods" in all directions,<br />

spreading out in increasingly thinner and ever more twisted<br />

filaments until it finally invades the whole space. No sketch<br />

can do justice to the complexity of the actual situation. Indeed,<br />

during the dynamic evolution of a mixing system, two<br />

· points as close together in phase space as we might wish may<br />

head in different directions. Even if we possess a lot of information<br />

about the system so that the initial cell formed by its<br />

representative points is very small, dynamic evolution turns<br />

this cell into a true geometric "monster" stretching its network<br />

of filaments through phase space.<br />

p<br />

Figure 34. Typical evolution in phase space of a cell corresponding to a<br />

"mixing" system. The volume is still conserved but no more its form: the cell<br />

progressively spreads through the whole phase space.<br />

q


ORDER OUT OF CHAOS<br />

268<br />

We would like to illustrate the distinction between stable<br />

and unstable systems with a few simple examples. Consider a<br />

phase space with two dimensions. At regular time intervals,<br />

we shall replace these coordinates by new ones. The new point<br />

on the horizontal axis is p-q, the new ordinate p. Figure 35<br />

shows what happens when we apply this operation to a square.<br />

(0,-1) p<br />

Figure 35. Transformation of a volume in phase space generated by a<br />

discrete transformation: the abscissa p becomes p-q, the ordinate q becomes<br />

p. The transformation is cyclic: after six times the initial cell is recovered.<br />

The square is deformed, but after six transformations we return<br />

to the original square. The system is stable: neighboring<br />

points are transformed into neighboring points. Moreover, it<br />

corresponds to a cyclic transformation (after six operations<br />

the original square reappears).<br />

Let us now consider two examples of highly unstabl systems-the<br />

first mathematical, the second of obvious physical


11<br />

269 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

relevance. The first system consists of a transformation that,<br />

for obvious reasons, mathematicians call the "baker transformation.<br />

"9, to We take a square and flatten it into a rectangle,<br />

then we fold half of the rectangle over the other half to form a<br />

square again. This set of operations is shown in Figure 36 and<br />

may be repeated as many times as one likes.<br />

q:1<br />

n<br />

<br />

1<br />

q=1<br />

2 p:1<br />

B<br />

q=1<br />

)<br />

q:11J<br />

p:1<br />

112 p:1<br />

e-1<br />

Figure 36. Realization of the baker transformation (B) and of its inverse<br />

(B-1). The path of the two spots gives an idea of the transformations.<br />

Each time the surface of the square is broken up and redistributed.<br />

The square corresponds here to the phase space.<br />

The baker transformation transforms each point into a welldefined<br />

new point. Although the series of points obtained in<br />

this way is "deterministic," the system displays in addition<br />

irreducibly statistical aspects. Let us take , for instance, a system<br />

described by an initial condition such that a region A of<br />

the square is initially filled in a uniform way with representative<br />

points. It may be shown that after a sufficient number of<br />

repetitions of the transformation, this cell, whatever its size<br />

and localization, will be broken up into pieces (see Figure 37).<br />

The essential point is that any region, whatever its size, thus


ORDER OUT OF CHAOS 270<br />

p<br />

Figure 37. Time evolution of an unstable system. Time going on, region A<br />

splits into regions A' and A", which in turn will be divided.<br />

q<br />

always contains different trajectories diverging at each fragmentation.<br />

Although the evolution of a point is reversible and<br />

deterministic, the description of a region, however small, is<br />

basically statistical.<br />

A similar example involves the scattering of hard spheres.<br />

We may consider a small sphere rebounding on a collection of<br />

large, randomly distributed spheres. The latter are supposed<br />

to be fixed. This is the model physicists call the "Lorentz<br />

model," after the name of a great Dutch physicist, Hendrik<br />

Antoon Lorentz.<br />

The trajectory of the small mobile sphere is well defined.<br />

However, whenever we introduce the smallest uncertainty in<br />

the initial conditions, this uncertainty is amplified through<br />

successive collisions. As time passes, the probability of finding<br />

the small sphere in a given volume becomes uniform.<br />

Whatever the number of transformations, we never return to<br />

the original state.<br />

In the last two examples we have strongly unstable dynamic<br />

systems. The situation is reminiscent of instabilities as they<br />

appear in thermodynamic systems (see Chapter V). Arbitrarily<br />

small differences in initial conditions are amplified. As a<br />

result we can no longer perform the transition from ensembles


271 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

<br />

Q<br />

I<br />

I<br />

I<br />

I<br />

I<br />

'o<br />

I I<br />

I I<br />

I 1<br />

1 I<br />

1 I<br />

I<br />

' ----- <br />

- -o---- --:..<br />

...<br />

.... .. ,"<br />

... ..<br />

... ..<br />

... ...<br />

... "<br />

...<br />

"<br />

"<br />

,<br />

"<br />

/<br />

"<br />

"<br />

•<br />

<br />

Figure 38. Schematic representation of the instability of the trajectory of a<br />

small sphere rebounding on large ones. The least imprecision about the<br />

position of the small sphere makes it impossible to predict which large<br />

sphere it will hit after the first collision.<br />

in phase space to individual trajectories. The description in<br />

terms of ensembles has to be taken as the starting point. Statistical<br />

concepts are no longer merely an approximation with<br />

respect to some "objective truth." When faced with these unstable<br />

systems, Laplace's demon is just as powerless as we.<br />

Einstein's saying, "God does not play dice," is well known.<br />

In the same spirit Poincare stated that for a supreme mathematician<br />

there is no place for probabilities. However, Poincare<br />

himself mapped the path leading to the answer to this problem.11<br />

He noticed that when we throw dice and use probability<br />

calculus, it does not mean that we suppose dynamics to be<br />

wrong. It means something quite different. We apply the probability<br />

concept because in each interval of initial conditions,<br />

however small, there are as "many" trajectories that lead to<br />

each of the faces of the dice. This is precisely what happens<br />

with unstable dynamic systems. God could, if he wished to,


ORDER OUT OF CHAOS 272<br />

calculate the trajectories in an unstable dynamic world. He<br />

would obtain the same result as probability calculus permits<br />

us to reach. Of course, if he made use of his absolute knowledge,<br />

then he could get rid of all randomness.<br />

In conclusion, there is a close relationship between instability<br />

and probability. This is an important point, and we want to<br />

discuss it now.<br />

From Randomness to Irreversibility<br />

Consider a succession of squares to which we apply the baker<br />

transformation. This succession is represented in Figure 39. A<br />

shaded region may be imagined to be filled with ink, an unshaded<br />

region by water. At time zero we have what is called a<br />

generating partition. Out of this partition we form a series of<br />

either horizontal partitions when we go into the future or vertical<br />

partitions going into the past. These are the basic partitions.<br />

An arbitrary distribution of ink in the square can be<br />

written formally as a superposition of the basic partitions. To<br />

each basic partition we may associate an "internal" time that<br />

is simply the number of baker transformations we have to perform<br />

to go from the generating partition to the one under consideration.I2<br />

We therefore see that this type of system admits a<br />

kind of internal age.*<br />

The internal time Tis quite different from the usual mechanical<br />

time, since it depends on the global topology of the system.<br />

We may even speak of the "timing of space," thus coming<br />

close to ideas recently put forward by geographers, who have introduced<br />

the concept of "chronogeography." 13 When we look at<br />

the structure of a town , or of a landscape, we see temporal<br />

elements interacting and coexisting. Brasilia or Pompeii would<br />

correspond to a well-defined internal age, somewhat like one<br />

of the basic partitions in the baker transformation. On the contrary,<br />

modern Rome, whose buildings originated in quite dif-<br />

*It may be noticed that this internal time, which we shall denote by T. is in<br />

fact an operator like those introduced in quantum mechanics (see Chapter<br />

VII). Indeed, an arbitrary partition of the square does not have a welldefined<br />

time but only an "average" time corresponding to the superposition<br />

of the basic partitions out of which it is formed.


273 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

-m-<br />

-<br />

-=-<br />

past<br />

•1 0 z<br />

t<br />

genart1ting partition<br />

future<br />

Figure 39. Starting with the "generating partition" (see text) at time 0, we<br />

repeatedly apply the baker transformation. We generate horizontal stripes in<br />

this way. Similarly going into the past we obtain vertical stripes.<br />

ferent periods, would correspond to an average time exactly as<br />

an arbitrary partition may be decomposed into elements corresponding<br />

to different internal times.<br />

Let us again look at Figure 39. What happens if we go into<br />

the far distant future? The horizontal bands of ink will get<br />

closer and closer. Whatever the precision of our measurements,<br />

after some time we shall conclude that the ink is uniformly<br />

distributed over the volume. It is therefore not surprising that<br />

this kind of approach to "equilibrium" may be mapped into a<br />

stochastic process, such as the Markov chain we described in<br />

Chapter VIII. Recently this has been shown with full mathematical<br />

rigor, 14 but the result seems to us quite natural. As<br />

time passes, the distribution of ink reaches equilibrium, exactly<br />

like the distribution of balls in the urn experiment discussed<br />

in Chapter VIII. However, when we look into the past,<br />

again beginning from the generating partition at time zero, we<br />

see the same phenomenon. Now ink is distributed along shrinking<br />

vertical sections and, again, sufficiently far in the past we<br />

shall find a uniform distribution of ink. We may therefore conclude<br />

that we can also model this process in terms of a Markov<br />

chain, now, however, oriented toward the past. We see that out<br />

of the unstable dynamic processes we obtain two Markov<br />

chains, one reaching equilibrium in the future, one in the past.<br />

We believe that it is a very interesting result and we would<br />

like to commen t it. Internal time provides us with a new 'nonlocal'<br />

description.<br />

When we know the 'age' of the system, (that is, the corresponding<br />

partition), we can still not associate to it a well-defined<br />

local trajectory.


ORDER OUT OF CHAOS 274<br />

We know only that the system is in a shaded region (Figure<br />

39). Similarly, if we know some precise initial conditions corresponding<br />

to a point in the system, we don't know the partition<br />

to which it belongs, nor the age of the system. For such<br />

systems we know therefore two complementary descriptions,<br />

and the situation becomes somewhat reminiscent of the one<br />

we described in Chapter VII, when we discussed quantum mechanics.<br />

It is because of the existence of this new alternative, nonlocal<br />

description, that we can make the transition from dynamics<br />

to probabilities. We call the systems for which this is<br />

possible "intrinsically random systems".<br />

In classical deterministic systems, we may use transition<br />

probabilities to go from one point to another on a quite degenerate<br />

sense. This transition probability will be equal to one if<br />

the two points lie on the same dynamic trajectory, or zero if<br />

they are not.<br />

In contrast, in genuine probability theory, we need transition<br />

probabilities which are positive numbers between zero<br />

and one. How is this possible? Here we see in full light the<br />

conflict between subjectivistic views and objective interpretations<br />

of probability. The subjective interpretation corresponds<br />

to the situation where individual trajectories are not known.<br />

Probability (and, eventually, irreversibility, closely related to<br />

it) would originate from our ignorance. But fortunately, there<br />

is another objective interpretation: probability arises as a result<br />

of an alternative description of dynamics, a non-local description<br />

which arises in strongly unstable dynamical systems.<br />

Here, probability becomes an objective property generated<br />

from the inside of dynamics, so to speak, and which expresses<br />

a basic structure of the dynamical system. We have stressed<br />

the importance of Boltzmann's basic discovery: the connection<br />

between entropy and probability. For intrinsic random<br />

systems, the concept of probability acquires a dynamical<br />

meaning. We have now to make the transition from intrinsic<br />

random systems to irreversible systems. We have seen that out<br />

of unstable dynamical processes, we obtain two Markov<br />

chains.<br />

We may see this duality in a different way. Take a distribution<br />

concentrated on a line (instead of being distributed on a<br />

surface). This line may be vertical or horizontal. Let us look at


275 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

what will happen to this line when we apply the baker transformation<br />

going to the future. The result is represented in Figure<br />

40. The vertical line is cut successively into pieces and will<br />

reduce to a point in the far distant future. The horizontal line,<br />

on the contrary, is duplicated and will uniformly "cover" the<br />

surface in the far distant future. Obviously, just the opposite<br />

happens if we go to the past. For reasons that are easy to understand,<br />

the vertical line is called a contracting fiber, the<br />

horizontal line a dilating fiber.<br />

We see now the complete analogy with bifurcation theory. A<br />

contracting fiber and a dilating fiber correspond to two realizations<br />

of dynamics, each involving symmetry-breaking and appearing<br />

in pairs. The contracting tiber corresponds to equilibrium in<br />

the far distant past, the dilating fiber to the future. We therefore<br />

have two Markov chains oriented in opposite time directions.<br />

Now we have to make the transistion from intrinsically random<br />

to intrinsically irreversible systems. To do so we must<br />

understand more precisely the difference between contracting<br />

and dilating fibers. We have seen that another system as unsta-<br />

I<br />

Figure 40. Contracting and dilating fibers in the baker transformation; time<br />

going on, the contracting fiber A1 is shortened (sequence A1, 81, C1), while<br />

the dilating fibers are duplicated (sequences A2, 82, C2).<br />

ble as the baker transformation can describe the scattering of<br />

hard spheres. Here contracting and dilating fibers have a sim-


ORDER OUT OF CHAOS 276<br />

pie physical interpretation. A contracting fiber corresponds to<br />

a collection of hard spheres whose velocities are randomly distributed<br />

in the far distant past, and all become parallel in the<br />

far distant future. A dilating fiber corresponds to the inverse<br />

situation, in which we start with parallel velocities and go to a<br />

random distribution of velocities. Therefore the difference is<br />

very similar to the one between incoming waves and outgoing<br />

waves in the example given by Popper. The exclusion of the<br />

contracting fibers corresponds to the experimental fact that<br />

whatever the ingenuity of the experimenter, he will never be<br />

able to control the system to produce parallel velocities after<br />

an arbitrary number of collisions. Once we exclude con·<br />

tracting fibers we are left with only one of the two possible<br />

Markov chains we have introduced. In other words, the second<br />

law becomes a selection principle of initial conditions.<br />

Only initial conditions that go to equilibrium in the fu ture are<br />

retained.<br />

Obviously the validity of this selection principle is maintained<br />

by dynamics. It can easily be seen in the example of the<br />

baker transformation that the contracting fiber remains a contracting<br />

fiber for all times, and likewise for a dilating fiber. By<br />

suppressing one of the two Markov chains we go from an<br />

intrinsically random system to an intrinsically irreversible system.<br />

We find three basic elements in the description of irreversibility:<br />

instability<br />

f<br />

intrinsic randomness<br />

f<br />

·intrinsic irreversibility<br />

Intrinsic irreversibility is the strongest property: it implies<br />

randomness and instability.t4, ts<br />

How is this conclusion compatible with dynamics? As we<br />

have seen, in dynamics "information" is conserved, while in<br />

Markov chains information is lost (and entropy therefore increases;<br />

see Chapter VIII). There is, however , no contradiction;<br />

when we go from the dynamic description of the baker<br />

transformation to the thermodynamic description, we have to


277 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

modify our distribution function; the "objects" in terms of<br />

which entropy increases are different fro m the ones considered<br />

in dynamics. The new distribution function, p, corresponds<br />

to an intrinsically time-oriented description of the<br />

dynamic system. In this book we cannot dwell on the mathematical<br />

aspects of this transformation. Let us only emphasize<br />

that it has to be noncanonical (see Chapter II). We must abandon<br />

the usual formulations of dynamics to reach the thermodynamic<br />

description.<br />

It is quite remarkable that such a transformation exists and<br />

that as a result we can now unify dynamics and thermodynamics,<br />

the physics of being and the physics of becoming. We shall<br />

come back to these new thermodynamic objects later in this<br />

chapter as well as in our concluding chapter. Let us emphasize<br />

only that at equilibrium, whenever entropy reaches its maximum,<br />

these objects must behave randomly.<br />

It also seems quite remarkable that irreversibility emerges, so<br />

to speak, from instability, which introduces irreducible statistical<br />

features into our description. Indeed, what could an arrow<br />

of time mean in a deterministic world in which both future and<br />

past are contained in the present? It is because the future is<br />

not contained in the present and that we go from the present to<br />

the future that the ar row of time is associated with the transition<br />

from present to future. This construction of irreversibility out of<br />

randomness has, we believe, many consequences that go<br />

beyond science proper and to which we shall come back in our<br />

concluding chapter. Let us clarify the difference between the<br />

states permitted by the second law and those it prohibits.<br />

The Entropy Barrier<br />

Time flows in a single direction, from past to future. We cannot<br />

manipulate time, we cannot travel back to the past. Travel<br />

through time has preoccupied writers, from The Thousand<br />

and One Nights to H. G. Wells' The Time Machine. In our<br />

time, Nabokov's short novel, Look at the Harlequins!, 1' describes<br />

the torment of a narrator who finds himself as unable<br />

to switch from one spatial direction to the other as we are to


ORDER OUT OF CHAOS 278<br />

"twirl time." In the fifth volume of Science and Civilization in<br />

China, Needham describes the dream of the Chinese alchemists:<br />

their supreme aim was not to achieve the transmutation<br />

of metals into gold but to manipulate time, to reach immortality<br />

through a radical slowdown of natural decaying processes.<br />

I? We are now better able to understand why we cannot<br />

"twirl time," to use Nabokov's expression.<br />

An infinite entropy barrier separates possible initial conditions<br />

fr om prohibited ones. Because this barrier is infinite,<br />

technological progress never will be able to overcome it. We<br />

have to abandon the hope that one day we will be able to travel<br />

back into our past. The situation is somewhat analogous to the<br />

barrier presented by the velocity of light. Technological progress<br />

can bring us closer to the velocity of light, but in the present<br />

views of physics we will never cross it.<br />

To understand the origin of this barrier, let us return to the<br />

expression of the :H quantity as it appears in the theory of<br />

Markov chains (see Chapter VIII). To each distribution we can<br />

associate a number-the corresponding value of J-{. We can<br />

say that to each distribution corresponds a well-defined information<br />

content. The higher the information content, the more<br />

difficult it will be to realize the corresponding state. What we<br />

wish to show here is that the initial distribution prohibited by<br />

the second law would have an infinite information content.<br />

That is the reason why we can neither realize them nor find<br />

them in nature.<br />

Let us first come back to the meaning of :Has presented in<br />

Chapter VIII. We have to subdivide the relevant phase space<br />

into sectors or boxes. With each box k, we associate an probability<br />

Peqm(k) at equilibrium as well as a non-equilibrium probability<br />

P(k,t).<br />

The :H is a measure of the difference between P(k,t) and<br />

Peqm(k), and vanishes at equilibrium when this difference disappears.<br />

Therefore, to compare the Baker transformation with<br />

Markov chains, we have to make more precise the corresponding<br />

choice of boxes. Suppose we consider a system at time 2<br />

(see Figure 39), and suppose that this system originated at<br />

time ti. Then, a result of our dynamical theory is that the<br />

boxes correspond to all possible intersections among the partitions<br />

between timet; and t=2. If we consider now Figure 39.<br />

we see that when ti is receding towards the past, the boxes will


279 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

become steadily thinner as we have to introduce more and<br />

more vertical subdivisions. This is expressed in Figure 41, sequence<br />

B, where, going from top to bottom, we have ti = 1, 0,<br />

- 1, and finally ti = - 2. We see indeed that the number of<br />

boxes increa ses in this way from 4 to 32.<br />

Once we have the boxes, we can compare the non-equilibrium<br />

distribution with the equilibrium distribution for each<br />

box. In the present case, the non-equilibrium distribution is<br />

either a dilating fiber (sequence A) or a contracting fiber (sequence<br />

C). The important point to notice is that when ti is<br />

receding to the past, the dilating fiber occupies an increasing<br />

large number of boxes: for ti = -2 it occupies 4 boxes, for<br />

ti = - 2 it occupies 8 boxes, and so on.<br />

As a result, when we apply the fo rmula given in Chapter<br />

VIII, we obtain a finite result, even if the number of boxes<br />

goes to infinity for tr-+ -oo.<br />

In contrast, the contracting fiber remains always localized<br />

in 4 boxes whatever ti. As a result, .Jl, when applied to a con-<br />

A 8 c<br />

Figure 41. Dilating (sequence A) and contracting (sequence C) fibers<br />

cross various numbers of the boxes which subdivide a Baker transformation<br />

phase space. All "squares" on a given sequence refer to the same time, t=2,<br />

but the number of the boxes subdividing each square depends on the initial<br />

time of the system ti.


ORDER OUT OF CHAOS<br />

280<br />

tracting fiber, diverges to infinity when ti recedes to the past.<br />

In summary, the difference between a dynamical system and<br />

the Markov chain is that the number of boxes to be considered<br />

in a dynamical system is infinite. It is this fact that leads to a<br />

selection principle. Only measures or probabilities, which in<br />

the limit of infinite number of boxes give a finite information<br />

or a finite J{ quantity, can be prepared or observed. This excludes<br />

contracting fibers.ts For the same reason we must also<br />

exclude distributions concentrated on a single point. Initial<br />

conditions corresponding to a single point in unstable systems<br />

would again correspond to infinite information and are therefore<br />

impossible to realize or observe. Again we see that the<br />

second law appears as a selection principle.<br />

In the classical scheme, initial conditions were arbitrary.<br />

This is no longer so for unstable systems. Here we can associate<br />

an information content to each initial condition, and this<br />

information content itself depends on the dynamics of the system<br />

(as in the baker transformation we used the successive<br />

fragmentation of the cells to calculate the information content).<br />

Initial conditions and dynamics are no longer independent.<br />

The second law as a selection rule seems to us so<br />

important that we would like to give another illustration based<br />

on the dynamics of correlations.<br />

The Dynamics of Correlations<br />

In Chapter VIII we discussed briefly the velocity inversion<br />

experiment. We may consider a dilute gas and follow its evolution<br />

in time. At time t0 we proceed to a velocity inversion of<br />

each molecule. The gas then returns to its initial state. We<br />

have already noted that for the gas to retrace its past there<br />

must be some storage of information. This storage can be described<br />

in terms of "correlations" between particles.t9<br />

Consider first a cloud of particles directed toward a target (a<br />

heavy, motionless particle). This situation is described in Figure<br />

42. In the far distant past, there were no correlations between<br />

particles. Now, scattering has two effects, already<br />

mentioned in Chapter VIII. It disperses the particles (it makes<br />

the velocity distribution more symmetrical) and, in addition, it


281 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

produces correlations between the scattered particles and the<br />

scatterer. The correlations can be made explicit by performing<br />

a velocity inversion (that is, by introducing a spherical mirror).<br />

Figure 43 represents this situation (the wavy lines represent<br />

the correlations). Therefore, the role of scattering is as follows:<br />

In the direct process, it makes the velocity distribution<br />

more symmetrical and creates correlations; in the inverse process,<br />

the velocity distribution becomes less symmetrical and<br />

correlations disappear. We see that the consideration of cor-<br />

•<br />

•<br />

.<br />

.. 0<br />

- .. ...<br />

• •<br />

Figure 42. Scattering of particles. Initially all particles have the same velocity.<br />

After the collision, the velocities are no more identical, and the scattered<br />

particles are correlated with the scatterer (correlations are always<br />

represented by wavy lines).<br />

relations introduces a basic distinction between the direct and<br />

the inverse processes.<br />

We can apply our conclusions to many-body systems. Here<br />

also we may consider two types of situations: in one, uncorrelated<br />

particles enter, are scattered, and correlated particles are<br />

produced (see Figure 44). In the opposite situation, correlated<br />

• •<br />

• ..<br />

<br />

0 • ..<br />

• ..<br />

Figure 43. Effect of a velocity inversion after a collision: after the new<br />

"inverted" collision, the correlations are suppressed and all particles have<br />

the same velocity.


ORDER OUT OF CHAOS 282<br />

particles enter, the correlations are destroyed through collisions,<br />

and uncorrelated particles re sult (see Figure 45).<br />

The two situations diffe r in the temporal order of collisions<br />

and correlations. In the first case , we have "postcollisional"<br />

correlations. With this distinction between pre- and postcollisional<br />

correlations in mind, let us re turn to the ve locity inversion<br />

experiment. We start at t = 0, with an initial state<br />

corresponding to no correlations between particles. During<br />

the time ot0 we have a "normal" evolution. Collisions bring<br />

the ve locity distribution closer to the Maxwellian equilibrium<br />

distribution. They also create postcollisional correlations be-<br />

0 0<br />

0 0<br />

before<br />

after<br />

Figure 44. Creation of postcollisional correlations represented by wavy<br />

lines; for details see text.<br />

tween the particles. At t0 after the velocity inversion, a completely<br />

new situation arises. Postcollisional correlations are<br />

now transformed into precollisional correlations. In the time<br />

interval between t0 and 2t0, these precollisional correlations<br />

disappear, the velocity distribution becomes less symmet-rical,<br />

and at time 2t0 we are back in the noncorrelational state. The<br />

history of this system therefore has two stage s. During the<br />

0 0<br />

0<br />

<br />

0<br />

I before<br />

after<br />

Figure 45. Destruction of precollisional correlations (wavy lines) through<br />

collisions.


283<br />

IRREVERSIBILITY-THE ENTROPY BARRIER<br />

first, collisions are transformed into correlations; in the second,<br />

correlations turn back into collisions. Both types of processes<br />

are compatible with the laws of dynamics. Moreover, as<br />

we have mentioned in Chapter VIII, the total "information"<br />

described by dynamics remains constant. We have also seen<br />

that in Boltzmann's description the evolution from time 0 till t0<br />

corresponds to the usual decrease of J{, while from t0 to 2t0 we<br />

have an abnormal situation: J{ would increase and entropy decrease.<br />

We would then be able to devise experiments in the<br />

laboratory or on computers in which the second Jaw would be<br />

violated! The irreversibility during time 0-t0 would be "compensated"<br />

by "anti-irreversibility" during time t0 - 2t0•<br />

This is quite unsatisfactory. All these difficulties disappear<br />

if we go, as in the baker transformation, to the new "thermodynamic<br />

representation" in terms of which dynamics becomes a<br />

probabilistic process like a Markov chain. We must also take<br />

into account that velocity inversion is not a "natural" process;<br />

it requires that "information" be given to molecules from the<br />

outside for them to invert their velocity. We need a kind of<br />

Maxwellian demon to perform the velocity inversions, and<br />

Maxwell's demon has a price. Let us represent the J{ quantity<br />

(for the probabilistic process) as a function of time. This is<br />

done in Figure 46. In this approach, in contrast with Boltzmann's,<br />

the effect of correlations is retained in the new definition<br />

of :H. Therefore at the velocity inversion point t0 the J{<br />

quantity will jump, since we abruptly create abnormal precollisional<br />

correlations that will have to be destroyed later. This<br />

jump corresponds to the entropy or information price we have<br />

to pay.<br />

Now we have a faithful representation of the second law: at<br />

every moment the J{ quantity decreases (or the entropy increases).<br />

There is one exception at time t0: J{ jumps upward,<br />

but that corresponds to the very moment at which the system<br />

is open. We can invert the velocities only by acting from the<br />

outside.<br />

There is another essential point: at time t0 the new J{ quantity<br />

has two different values, one for the system before velocity<br />

inversion and the other after a velocity inversion. These<br />

two situations have different entropies. This resembles what<br />

occurs in the baker transformation when the contracting and<br />

dilating fibers are velocity inversions of each other.


ORDER OUT OF CHAOS<br />

284<br />

Suppose we wait a sufficient time before making the velocity<br />

in-version. The postcollisional correlations wmdd have<br />

an arbitrary range, and the entropy price for velocity inversion<br />

would become too high. The velocity inversion would then<br />

require too high an entropy price and thus would be excluded.<br />

In physical terms this means that the second law excludes persistent<br />

long-range precollisional correlations.<br />

The analogy with the macroscopic description of the second<br />

law is striking. From the point of view of energy conservation<br />

(see Chapters IV and V), heat and work play the same role,<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

I<br />

----·-------+ t<br />

t 0 2t 0<br />

Figure 46. Time variation of the .J l -function in the vlocity inversion experiment:<br />

at time t0, the velocities are inversed and J-l presents a discontinuity.<br />

At time 2t0 the system is in the same state as at ti_me 0, and .J l<br />

recovers the value it had initially. At all times (except at t0), J-l is decreasing.<br />

The important fact is that at time t0 the J-l-quantity takes two different values<br />

(see text).<br />

but no longer from the point of view of the second law. Briefly<br />

speaking, work is a more coherent form of energy and always<br />

can be converted into heat, but the inverse is not true. There is<br />

on the microscopic level a similar distinction between collisions<br />

and correlations. From the point of view of dynamics, collisions<br />

and correlations play equivalent roles. Collisions give<br />

rise to correlations, and correlations may destroy the effect of<br />

I


285 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

collisions. But there is an essential difference. We can control<br />

collisions and produce correlations, but we cannot control correlations<br />

in a way that will destroy the effects collisions have<br />

brought into the system. It is this essential difference that is<br />

missing in dynamics but that can be incorporated into thermodynamics.<br />

Note that thermodynamics does not enter into conflict<br />

with dynamics at any point. It adds an additional, essential<br />

element to our understanding of the physical world.<br />

Entropy as a Selection Principle<br />

It is amazing how closely the microscopic theory of irreversible<br />

processes resembles traditional macroscopic theory. In both<br />

cases entropy initially has a negative meaning. In its macroscopic<br />

aspect it prohibits some processes, such as heat flowing from<br />

cold to hot. In its microscopic aspect it prohibits certain<br />

classes of initial conditions. The distinction between what is<br />

permitted and what is prohibited is maintained in time by the<br />

laws of dynamics. It is from the negative aspect that the positive<br />

aspect emerges: the existence of entropy together with its<br />

probability interpretation. Irreversibility no longer emerges as<br />

if by a miracle at some macroscopic level. Macroscopic irreversibility<br />

only makes apparent the time-oriented polarized<br />

nature of the universe in which we live.<br />

As we have emphasized repeatedly, there exist in nature systems<br />

that behave reversibly and that may be fully described by<br />

the laws of classical or quantum mechanics. But most systems<br />

of interest to us, including all chemical systems and therefore<br />

all biological systems, are time-oriented on the macroscopic<br />

level. Far from being an "illusion," this expresses a broken<br />

time-symmetry on the microscopic level. Irreversibility is either<br />

true on all levels or on none. It cannot emerge as if by a<br />

miracle, by going from one level to another.<br />

Also we have already noticed, irreversibility is the starting<br />

point of other symmetry breakings. For example, it is generally<br />

accepted that the difference between particles and antiparticles<br />

could arise only in a nonequilibrium world. This may<br />

be extended to many other situations. It is likely that irreversibility<br />

also played a role in the appearance of chiral symmetry


ORDER OUT OF CHAOS 286<br />

through the selection of the appropriate bifurcation. One of<br />

the most active subjects of research now is the way in which<br />

irreversibility can be "inscribed" into the structure of matter.<br />

The reader may have noticed that in the derivation of microscopic<br />

irreversibility we have been concentrating on classical<br />

dynamics. However, the ideas of correlations and the distinction<br />

between pre- and postcollisional correlations apply to<br />

quantum systems as well. The study of quantum systems is<br />

more complicated than the study of classical systems. There is<br />

a reason for this: the difference between classical and quantum<br />

mechanics. Even small classical systems, such as those<br />

formed by a few hard spheres, may present intrinsic irreversibility.<br />

However, to reach irreversibility in quantum systems we<br />

need large systems, such as those realized in liquids, gases, or<br />

in field theory. The study of large systems is obviously more<br />

difficult mathematically, and that is why we will not go into the<br />

matter here. However, the situation remains essentially the<br />

same in quantum theory. There also irreversibility begins as<br />

the result of the limitation of the concept of wave function due<br />

to a form of quantum instability.<br />

Moreover, the idea of collisions and correlations may also be<br />

used in quantum theory. Therefore, as in classical theory, the<br />

second law prohibits long-range precollisional correlations.<br />

The transition to a probability process introduces new entities,<br />

and it is in terms of these new entities that the second<br />

law can be understood as an evolution from order to disorder.<br />

This is an important conclusion. The second law leads to a<br />

new concept of matter. We would like to describe this concept<br />

now.<br />

Active Matter<br />

Once we associate entropy with a dynamic system, we come<br />

back to Boltzmann's conception: the probability will be maximum<br />

at equilibrium. The units we use to describe thermo<br />

dynamic evolution will therefore behave in a chaotic way at<br />

equilibrium. In contrast , in near-equilibrium conditions correlations<br />

and coherence will appear.<br />

We come to one of our main conclusions; At all levels, be it


287 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

the level of macroscopic physics, the level of fluctuations, or<br />

the microscopic level, nonequilibrium is the source of order.<br />

Nonequilibrium brings "order out of chaos." But as we already<br />

mentioned, the concept of order (or disorder) is more<br />

complex than was thought. It is only in some limiting situations,<br />

such as with dilute gases, that it acquires a simple meaning<br />

in agreement with Boltzmann's pioneering work.<br />

Let us once more contrast the dynamic description of the<br />

physical world in terms of forces and fields with the thermo<br />

dynamic description. As we mentioned, we can construct<br />

computer experiments in which interacting particles initially<br />

distributed at random form a lattice. The dynamic interpretation<br />

would be the appearance of order through interparticle<br />

forces. The thermodynamic interpretation is, on the contrary,<br />

the approach to disorder (when the system is isolated), but to<br />

disorder expressed in quite different units, which are in this<br />

case collective modes involving a large number of particles. It<br />

seems to us worthwhile to reintroduce the neologism we used<br />

in Chapter VI to define the new units in terms of which the<br />

system is incoherent at equilibrium: we call them "hypnons,"<br />

sleepwalkers, since they ignore each other at equilibrium.<br />

Each of them may be as complex as you wish (think about<br />

molecules of the complexity of an enzyme), but at equilibrium<br />

their complexity is turned "inward." Again, inside a molecule<br />

there is an intensive electric field, but this field in a dilute gas<br />

is negligible as far as other molecules are concerned.<br />

One of the min subjects in present-day physics is the problem<br />

of elementary particles. However, we know that elementary<br />

particles are far from elementary. New layers of structure<br />

are disclosed at higher and higher energies. But what, after all,<br />

is an elementary particle? Is the planet earth an elementary<br />

particle? Certainly not, because part of this energy is in its<br />

interaction with the sun, the moon, and the other planets. The<br />

concept of elementary particles requires an "autonomy" that<br />

is very difficult to describe in terms of the usual concepts.<br />

Thke the case of electrons and photons. We are faced with a<br />

dilemma: either there are no well-defined particles (because<br />

the energy is partly between the electrons and protons), or<br />

there are noninteracting particles if we can eliminate the interaction.<br />

Even if we knew how to do that, it seems too radical a


ORDER OUT OF CHAOS 288<br />

procedure. Electrons absorb photons or emit photons. A way<br />

out may be to go to the physics of processes. The units, the<br />

elementary particles, would then be defined as hypnons, as<br />

the entities that evolve independently at equilibrium. We hope<br />

that there soon will be experiments available to test this hypothesis;<br />

it would be quite appealing if atoms interacting with<br />

photons (or unstable elementary particles) already carried the<br />

arrow of time that expresses the global evolution of nature.<br />

A subject widely discussed today is the problem of cosmic<br />

evolution. How could the world near the moment of the Big<br />

Bang be so "ordered"? Yet this order is necessary if we wish<br />

to understand cosmic evolution as the gradual movement from<br />

order to disorder.<br />

To give a satisfactory answer we need to know what "hypnons"<br />

could have been appropriate to the extreme conditions<br />

of temperature and density that characterized the early universe.<br />

Thermodynamics alone will not, of course, solve these<br />

problems; neither will dynamics, even in its most refined form<br />

field theory. That is why the unification of dynamics and thermodynamics<br />

opens new perspectives.<br />

In any case, it is striking how the situation has changed since<br />

the formulation of the second law of thermodynamics one hundred<br />

fifty years ago. At first it seemed that the atomistic view<br />

contradicted the concept of entropy. Boltzmann attempted to<br />

save the mechanistic world view at the cost of reducing the<br />

second law to a statement of probability with great practical<br />

importance but no fundamental significance. We do not know<br />

what the definitive solution will be; but today the situation is<br />

radically different. Matter is not given. In the present-day view<br />

it has to be constructed out of a more fundamental concept in<br />

terms of quantum fields. In this construction of matter, thermodynamic<br />

concepts (irreversibility, entropy) have a role to<br />

play.<br />

Let us summarize what has been achieved here. The central<br />

role of the second law (and of the correlative concept of irreversibility)<br />

at the level of macroscopic systems has already<br />

been emphasized in Books One and 1\vo.<br />

What we have tried to show in Book Three is that we now<br />

can go beyond the macroscopic level, and discover the microscopic<br />

meaning of irreversibility.<br />

However, this requires basic changes in the way in which we


289 IRREVERSIBILITY-THE ENTROPY BARRIER<br />

conceive the fundamental laws of physics. It is only when the<br />

classical point of view is lost-as it is in the case of sufficiently<br />

unstable systems-that we can speak of 'intrinsic ra ndomness'<br />

and 'intrinsic irreversibility. '<br />

It is for such systems that we may introduce a new enlarged<br />

description of time in terms of the operator time T. As we have<br />

shown in the example of the Baker transformation (Chapter IX<br />

"From randomness to irreversibility"), this operator has as<br />

eigenfunctions partitions of the phase space (see Figure 39).<br />

We come therefore to a situation quite reminiscent of that of<br />

quantum mechanics. We have indeed two possible descriptions.<br />

Either we give ourselves a point in phase space, and<br />

then we don't know to which partition it belongs, and therefore<br />

we don't know its internal age ; or we know its internal<br />

age, but then we know only the partition, but not the exact<br />

localization of the point.<br />

Once we have introduced the internal time T, we can use<br />

entropy as a selection principle to go from the initial description<br />

in terms of the distribution function p to a new one, p<br />

where the distribution p has an intrinsic arrow of time, satisfying<br />

the second law of thermodynamics. The basic difference<br />

between p and p appears when these functions are expanded in<br />

terms of the eigenfunction of the operator time T (see Chapter<br />

VII, "The rise of quantum mechanics"). In p, all internal ages,<br />

be they from past or future, appear symmetrically. In contrast,<br />

in p, past and future play different roles. The past is included,<br />

but the future remains uncertain. That is the meaning of the<br />

arrow of time. The fascinating aspect is that there appears now<br />

a relation betwen initial conditions and the laws of change. A<br />

state with an arrow of time emerges from a law, which has also<br />

an arrow of time, and which transforms the state, however<br />

keeping this arrow of time.<br />

We have concentrated mostly on the classical situation.2o<br />

However, our analysis applies as well to quantum mechanics,<br />

where the situation is more complicated, as the existence of<br />

Planck's constant destroys already the concept of a trajectory,<br />

and leads therefore also to a kind of delocalization in phase<br />

space. In quantum mechanics we have therefore to superpose<br />

the quantum delocalization with de localization due to irreversibility.<br />

As emphasized in Chapter VII, the two great revolutions in


ORDER OUT OF CHAOS<br />

290<br />

the physics of our century correspond to the incorporation, in<br />

the fundamental structure of physics, of impossibilities foreign<br />

to classical mechanics: the impossibility of signals propagating<br />

with a velocity larger than the velocity of light , and the impossibility<br />

of measuring simultaneously coordinates and momentum.<br />

It is not astonishing that the second principle, which as well<br />

limits our ability to manipulate matter, also leads to deep<br />

changes in the structure of the basic laws of physics.<br />

Let us conclude this part of our monograph wit h a word of<br />

caution. The phenomenological theory of irreversible processes<br />

is at present well established. In contrast , the basic microscopic<br />

theory of irreversible processes is quite new. At the<br />

time of correcting the proofs of this book, experiments are in<br />

preparation to test these views. As long as they have not been<br />

performed, a speculative element is unavoidable.


CONCLUSION<br />

FROM EARTH TO HEAVEN­<br />

THE REENCHANTMENT<br />

OF NATURE<br />

In any attempt to bridge the domains of experience belonging<br />

to the spiritual and physical s1des of our nature, time occupies<br />

the key position.<br />

A. S. EDDINGTON1<br />

An Open Science<br />

Science certainly involves manipulating nature, but it is also<br />

an attempt to understand it, to dig deeper into questions that<br />

have been asked generation after generation. One of these questions<br />

runs like a leitmotiv, almost as an obsession, through this<br />

book, as it does through the history of science and philosophy.<br />

This is the question of the relation between being and becoming,<br />

between permanence and change.<br />

We have mentioned pre-Socratic speculations: Is change,<br />

whereby things are born and die, imposed from the outside on<br />

some kind of inert matter? Or is it the result of the intrinsic and<br />

independent activity of matter? Is an external driving force<br />

necessary, or is becoming inherent in matter? Seventeenthcentury<br />

science arose in opposition to the biological model of<br />

a spontaneous and autonomous <strong>org</strong>anization of natural beings.<br />

But it was confronted with another fundamental alternative. Is<br />

nature intrinsically random? Is ordered behavior merely the<br />

transient result of the chance collisions of atoms and of their<br />

unstable associations?<br />

One of the main sources of fascination in modern science<br />

was precisely the feeling that it had discovered eternal laws at<br />

291


ORDER OUT OF CHAOS 292<br />

the core of nature's transformations and thus had exorcised<br />

time and becoming. This discovery of an order in nature produced<br />

the fe eling of intellectual security described by French<br />

sociologist Levy-Bruhl:<br />

Our feeling of intellectual security is so deeply anchored<br />

in us that we even do not see how it could be shaken.<br />

Even if we suppose that we could observe some phenomenon<br />

seemingly quite mysterious, we still would remain<br />

persuaded that our ignorance is only provisional, that this<br />

phenomenon must satisfy the general laws of causality,<br />

and that the reasons for which it has appeared will be<br />

determined sooner or later. Nature around us is order<br />

and reason, exactly as is the human mind. Our everyday<br />

activity implies a perfect confidence in the universality of<br />

the laws of nature.2<br />

This feeling of confidence in the "reason" of nature has<br />

been shattered, partly as the result of the tumultuous growth<br />

of science in our time. As we stated in the Preface, our vision<br />

of nature is undergoing a radical change toward the multiple,<br />

the temporal, and the complex. Some of these changes have<br />

been described in this book.<br />

We were seeking general, all-embracing schemes that could<br />

be expressed in terms of eternal laws, but we have found time,<br />

events, evolving particles. We were also searching for symmetry,<br />

and here also we were surprised, since we discovered<br />

symmetry-breaking processes on all levels, from elementary<br />

particles up to biology and ecology. We have described in this<br />

book the clash between dynamics, with the temporal symmetry<br />

it implies, and the second law of thermodynamics, with its<br />

directed time.<br />

A new unity is emerging: irreversibility is a source of order<br />

at all levels. Irreversibility is the mechanism that brings order<br />

out of chaos. How could such a radical transformation of our<br />

views on nature occur in the relatively short time span of the<br />

past few decades? We believe that it shows the important role<br />

intellectual construction plays in our concept of reality. This<br />

was very well expressed by Bohr, when he said to Werner Heisenberg<br />

on the occasion of a visit at Kronberg Castle:


293<br />

. FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

Isn't it strange how this castle changes as soon as one<br />

imagines that Hamlet lived here? As scientists we believe<br />

that a castle consists only of stones, and admire the way<br />

the architect put them together. The stones, the green<br />

roof with its patina, the wood carvings in the church, constitute<br />

the whole castle. None of this should be changed by<br />

the fact that Hamlet lived here, and yet it is changed completely.<br />

Suddenly the walls and the ramparts speak a different<br />

language . ... Yet all we really know about Hamlet<br />

is that his name appears in a thirteenth-century chronicle<br />

. ... But everyone knows the questions Shakespeare<br />

had him ask, the human depths he was made to reveal,<br />

and so he too had to have a place on earth, here in Kronberg}<br />

The question of the meaning of reality was the central subject<br />

of a fascinating dialogue between Einstein and Tagore."'<br />

Einstein emphasized that science had to be independent of the<br />

existence of any observer. This led him to deny the reality of<br />

time as irreversibility, as evolution. On the contrary, Thgore<br />

maintained that even if absolute truth could exist, it would be<br />

inaccessible to the human mind. Curiously enough, the present<br />

evolution of science is running in the direction stated by<br />

the great Indian poet. Whatever we call reality, it is revealed to<br />

us only through the active construction in which we participate.<br />

As it is concisely expressed by D. S. Kothari, "The simple<br />

fact is that no measurement, no experiment or observation<br />

is possible without a relevant theoretical framework. "5<br />

Time and Times<br />

The statement that time is basically a geometric parameter<br />

that makes it possible to follow the unfolding of the succession<br />

of dynamical states has been asserted in physics for more than<br />

three centuries. Emile Meyerson6 tried to describe the history<br />

of modern science as the gradual implementation of what he<br />

regarded as a basic category of human reason: the different<br />

and the changing had to be reduced to the identical and the<br />

permanent. Time had to be eliminated.


ORDER OUT OF CHAOS<br />

294<br />

Nearer to our own time, Einstein appears as the incarnation<br />

of this drive toward a formulation of physics in which no reference<br />

to irreversibility would be made on the fundamental level.<br />

An historic scene took place at the Societe de Philosophie in<br />

Paris on April 6, 1922,1 when Henri Bergson attempted to defend<br />

the cause of the multiplicity of coexisting "lived" times<br />

against Einstein. Einstein's reply was absolute: he categorically<br />

rejected "philosophers' time." Lived experience cannot<br />

save what has been denied by science.<br />

Einstein's reaction was somewhat justified. Bergson had obviously<br />

misunderstood Einstein's theory of relativity. However,<br />

there also was some prejudice on Einstein's part: duree,<br />

Bergson's "lived time," refers to the basic dimensions of becoming,<br />

the irreversibility that Einstein was willing to admit<br />

only at the phenomenological level. We have already referred<br />

to Einstein's conversations with Carnap (see Chapter VII).<br />

For him distinctions among past, present, and future were outside<br />

the scope of physics.<br />

It is fascinating to follow the correspondence between Einstein<br />

and the closest friend of his young days in Zurich, Michele<br />

Besso. 8 Although he was an engineer and scientist, toward the<br />

end of his life Besso became increasingly concerned with philosophy,<br />

literature, and the problems that surround the core of<br />

human existence. Untiringly he kept asking the same questions:<br />

What is irreversibility? What is its relationship with the<br />

laws of physics? And untiringly Einstein would answer with a<br />

patience he showed only to this closest friend: irreversibility is<br />

merely an illusion produced by "improbable" initial conditions.<br />

This dialogue continued over many years until Besso, older<br />

than Einstein by eight years, passed away, only a few months<br />

before Einstein's death. In a last letter to Besso's sister and<br />

son, Einstein wrote: "Michele has left this strange world just<br />

before me. This is of no importance. For us convinced physicists<br />

the distinction between past, present and future is an illusion,<br />

although a persistent one." In Einstein's drive to perceive<br />

the basic laws of physics, the intelligible was identified with<br />

the immutable.<br />

Why was Einstein so strongly opposed to the introduction<br />

of irreversibility into physics? We can only guess. Einstein<br />

was a rather lonely man; he had few friends, few coworkers,


295<br />

FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

few students. He lived in a sad time: the two World Wars , the<br />

rise of anti-Semitism. It is not surprising that for Einstein science<br />

was the road that led to victory over the turmoil of time.<br />

What a contrast, however, with his scientific work. His world<br />

was full of observers, of scientists situated in various coordinate<br />

systems in motion with one another, situated on various<br />

stars differing by their gravitational fields. All these observers<br />

were exchanging information through signals all over the universe.<br />

What Einstein wanted to preserve above all was the objective<br />

meaning of this communication. However, we can<br />

perhaps state that Einstein stopped short of accepting that<br />

communication and irreversibility are closely related. Communication<br />

is at the base of what probably is the· most irreversible<br />

process accessible to the human mind, the progressive<br />

increase of knowledge.<br />

The Entropy Barrier<br />

In Chapter IX we described the second law as a selection principle:<br />

to each initial condition there corresponds an "information."<br />

All initial conditions for which this information is finite<br />

are permitted. However, to reverse the direction of time we<br />

would need infinite information; we cannot produce situations<br />

that would evolve into our past ! This is the entropy barrier we<br />

have introduced.<br />

There is an interesting analogy with the concept of the velocity<br />

of light as the maximum velocity of transmission of signals.<br />

As we have seen in Chapter VII, this is one of the basic<br />

postulates of Einstein's relativity theory. The existence of the<br />

velocity of light barrier is necessary to give meaning to causality.<br />

Suppose we could, in a science-fiction ship, leave the earth at<br />

a velocity greater than the velocity of light. We could overtake<br />

light signals and in this way precede our own past. Likewise,<br />

the entropy barrier is necessary to give meaning to communication.<br />

We have already mentioned that irreversibility and<br />

communication are closely related. Norbert Wiener has argued<br />

that the existence of two directions of time would have<br />

disastrous consequences. It is worthwhile to cite a passage<br />

from his well-known book Cybernetics:


ORDER OUT OF CHAOS 296<br />

Indeed, it is a very interesting intellectual experiment<br />

to make the fantasy of an intelligent being whose time<br />

should run the other way to our own. To such a being, all<br />

communication with us would be impossible. Any signal<br />

he might send would reach us with a logical stream of<br />

consequents from his point of view, antecedents from<br />

ours. These antecedents would already be in our experience,<br />

and would have served to us as the natural explanation<br />

of his signal, without presupposing an intelligent<br />

being to have sent it. If he drew us a square, we should<br />

see the remains of his figure as its precursors, and it would<br />

seem to be the curious crystallization-always perfectly<br />

explainable-of these remains. Its meaning would seem<br />

to be as fortuitous as the faces we read into mountains<br />

and cliffs. The drawing of the square would appear to us<br />

as a catastrophe-sudden indeed, but explainable by natural<br />

laws-by which that square wo_Id cease to exist.<br />

Our counterpart would have exactly similar ideas concerning<br />

us. Within any world with which we can communicate,<br />

the direction of time is uniform.9<br />

It is precisely the infinite entropy barrier that guarantees the<br />

uniqueness of the direction of time, the impossibility of switching<br />

from one direction of time to the opposite one.<br />

In the course of this book, we have stressed the importance<br />

of demonstrations of impossibility. In fact, Einstein was the<br />

first to grasp that importance when he based his concept of relative<br />

simultaneity on the impossibility of transmitting information<br />

at a velocity greater than that of light. The whole theory of<br />

relativity is built around the exclusion of "unobservable" simultaneities.<br />

Einstein considered this step as somewhat similar to<br />

the step taken in thermodynamics when perpetual motion was<br />

excluded. But some of his contemporaries-Heisenberg, for<br />

example-emphasized an important difference between these<br />

two impossibilities. In the case of thermodynamics, a certain<br />

situation is defined as being absent from nature; in the case of<br />

relativity, it is an observation that is defined as impossiblethat<br />

is, a type of dialogue, of communication between nature<br />

and the person who describes it. Thus Heisenberg saw himself<br />

as following Einstein's example, in spite of Einstein's skepticism,<br />

when he grounded quantum mechanics on the ex.clusion


297<br />

FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

of what the quantum uncertainty principle defines as unobservable.<br />

As long as the second law was considered to express only a<br />

practical improbability, it had little theoretical interest. You<br />

could always hope to overcome it by sufficient technical skill.<br />

But we have seen that this is not so. At its roots there is a<br />

selection of possible initial states. It is only after these states<br />

have been selected that the probability interpretation becomes<br />

possible. Indeed, as Boltzmann stated for the first time, the<br />

increase of entropy expresses the increase of probability, of<br />

disorder. However, his interpretation results from the conclusion<br />

that entropy is a selection principle breaking the time<br />

symmetry. It is only after this symmetry-breaking that any<br />

probabilistic interpretation becomes possible.<br />

In spite of the fact that we have recouped much of Boltzmann's<br />

interpretation of entropy, the basis of our interpretation<br />

of his second law is radically different, since we have in<br />

succession<br />

the second law as a symmetry-breaking selection principle<br />

<br />

probabilistic interpretation<br />

<br />

irreversibility as increase of disorder<br />

It is only the unification of dynamics and thermodynamics<br />

through the introduction of a new selection principle that gives<br />

the second law its fundamental importance as the evolutionary<br />

paradigm of the sciences. This point is of such importance that<br />

we shall dwell on it in more detail.<br />

The Evolutionary Paradigm<br />

f<br />

The world of dynamics, be it classical or quantum, is a reversible<br />

world. As we have emphasized in Chapter VIII, no<br />

evolution can be ascribed to this world; the "information" expressed<br />

in terms of dynamical units remains constant. It is<br />

therefore of great importance that the existence of an evolutionary<br />

paradigm can now be established in physics-not only


ORDER OUT OF CHAOS 298<br />

on the level of macroscopic description but also on all levels.<br />

Of course, there are conditions: as we have seen, a minimum<br />

complexity is necessary. But the immense importance of irreversible<br />

processes shows that this requirement is satisfied for<br />

most systems of interest. Remarkably, the perception of oriented<br />

time increases as the level of biological <strong>org</strong>anization increases<br />

and probably reaches its culminating point in human<br />

consciousness.<br />

How general is this evolutionary paradigm? It includes isolated<br />

systems that evolve to disorder and open systems that<br />

evolve to higher and higher forms of complexity. It is not surprising<br />

that the entropy metaphor has tempted a number of<br />

writers dealing with social or economic problems. Obviously<br />

here we have to be careful; human beings are not dynamic<br />

objects, and the transition to thermodynamics cannot be formulated<br />

as a selection principle maintained by dynamics. On<br />

the human level irreversibility is a more fundamental concept,<br />

which is for us inseparable from the meaning of our existence.<br />

Still it is essential that in this perspective we no longer see the<br />

internal feeling of irreversibility as a subjective impression that<br />

alienates us from the outside world, but as marking our participation<br />

in a world dominated by an evolutionary paradigm.<br />

Cosmological problems are notoriously difficult. We still do<br />

not know what the role of gravitation was in the early universe.<br />

Can gravitation be included in some form of the second law, or<br />

is there a kind of dialectical balance between thermodynamics<br />

and gravitation? Certainly irreversibility could not have appeared<br />

abruptly in a time-reversible world. The origin of irreversibility<br />

is a cosmological problem and would require an<br />

analysis of the universe in its first stages. Here our aim is more<br />

modest. What does irreversibility mean today? How does it<br />

relate to our position in the world we describe?<br />

Actors and Spectators<br />

The denial of becoming by physics created deep rifts within<br />

science and estranged science from philosophy. What had<br />

originally been a daring wager with the dominant Aristotelian


299<br />

FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

tradition gradually became a dogmatic assertion directed<br />

against all those (chemists, biologists, physicians) for whom a<br />

qualitative diversity existed in nature. At the end of the nineteenth<br />

century this conflict had shifted from inside science to<br />

the relation between "science" and the rest of culture, especially<br />

philosophy. We have described in Chapter III this aspect<br />

of the history of Western thought, with its continual<br />

struggle to achieve a new unity of knowledge. The "lived<br />

time" of the phenomenologists, the Lebenswelt opposed to<br />

the objective world of science, may be related to the need to<br />

erect bulwarks against the invasion of science.<br />

Today we believe that the epoch of certainties and absolute<br />

oppositions is over. Physicists have no privilege whatsoever to<br />

any kind of extraterritoriality. As scientists they belong to<br />

their culture, to which, in their turn, they make an essential<br />

contribution. We have reached a situation close to the one recognized<br />

long ago in sociology: Merleau-Ponty had already<br />

stressed the need to keep in mind what he termed a "truth<br />

within situations. "<br />

So long as I keep before me the ideal of an absolute observer,<br />

of knowledge in the absence of any viewpoint, I<br />

can only see my situation as being a source of error. But<br />

once I have acknowledged that through it I am geared to<br />

all actions and all knowledge that are meaningful to me,<br />

and that it is gradually filled with everything that may be<br />

for me, then my contact with the social in the finitude of<br />

my situation is revealed to me as the starting point of all<br />

truth, including that of science and, since we have some<br />

idea of the truth, since we are inside truth and cannot get<br />

outside it, all that I can do is define a truth within the<br />

situation. 10<br />

It is this conception of knowledge as both objective and participatory<br />

that we have explored through this book.<br />

In his Themesll Merleau-Ponty also asserted that the "philosophic"<br />

discoveries of science, its basic conceptual transformations,<br />

are often the result of negative discoveries. which<br />

provide the occasion and the starting point for a reversal of<br />

point of view. Demonstrations of impossibility, whether in rel-


ORDER OUT OF CHAOS<br />

300<br />

ativity, quantum mechanics, or thermodynamics, have shown<br />

us that nature cannot be described "from the outside," as if by<br />

a spectator. Description is dialogue, communication, and this<br />

communication is subject to constraints that demonstrate that<br />

we are macroscopic beings embedded in the physical world.<br />

We may summarize the situation as we see it today in the<br />

following diagram:<br />

observer ---+<br />

t<br />

dissipative structures<br />

t<br />

dynamics<br />

irreversibility +- randomness +- unstable dynamic systems<br />

We start from the observer, who measures coordinates and<br />

momenta and studies their change in time. This leads him to<br />

the discovery of unstable dynamic systems and other concepts<br />

of intrinsic randomness and intrinsic irreversibility, as we discussed<br />

them in Chapter IX. Once we have intrinsic irreversibility<br />

and entropy, we come in far-from-equilibrium systems<br />

to dissipative structures, and we can understand the timeoriented<br />

activity of the observer.<br />

There is no scientific activity that is not time-oriented. The<br />

preparation of an experiment calls for a distinction between<br />

"before.. and "after... It is only because we are aware of irreversibility<br />

that we can recognize reversible motion. Our diagram<br />

shows that we have now come full circle, that we can see<br />

ourselves as part of the universe we describe.<br />

The scheme we have presented is not an a priori schemededucible<br />

from some logical structure. There is, indeed, no<br />

logical necessity for dissipative structures actually to exist in<br />

nature; the "cosmological fact" of a universe far from equilib<br />

rium is needed for the macroscopic world to be a world inhabited<br />

by "observers"-that is, to be a living world. Our scheme thus<br />

does not correspond to a logical or epistemological truth but<br />

refers to our condition as macroscopic beings in a world far<br />

from equilibrium. Moreover, an essential characteristic of our<br />

scheme is that it does not suppose any fundamental mode of<br />

description; each level of description is implied by another and<br />

implies the other. We need a multiplicity of levels that are all<br />

connected, none of which may have a claim to preeminence.<br />

I


301 FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

We have already emphasized that irreversibility is not a universal<br />

phenomenon. We can perform experiments corresponding<br />

to thermodynamic equilibrium in limited portions of space.<br />

Moreover, the importance of time scales varies. A stone<br />

evolves according to the time scale of geological evolution;<br />

human societies, especially in our time, obviously have a<br />

much shorter time scale. We have already mentioned that irreversibility<br />

starts with a minimum complexity of the dynamic<br />

systems involved. It is interesting that with an increase of<br />

complexity, going from the stone to human societies, the role<br />

of the arrow of time, of evolutionary rhythms, increases. Molecular<br />

biology shows that everything in a cell is not alive in<br />

the same way. Some processes reach equilibrium, others are<br />

dominated by regulatory enzymes far from equilibrium. Similarly,<br />

the arrow of time plays quite different roles in the universe<br />

around us. From this point of view, in the sense of this<br />

time-oriented activity, the human condition seems unique. It<br />

seems to us, as we said in Chapter IX, quite important that<br />

irreversibility, the arrow of time, implies randomness. "Time<br />

is construction." This conclusion, one that Valery 12 reached<br />

quite independently, carries a message that goes beyond science<br />

proper.<br />

A Whirlwind in a Turbulent Nature<br />

In our society, with its wide spectrum of cognitive techniques,<br />

science occupies a peculiar position, that of a poetical interrogation<br />

of nature, in the etymological sense that the poet is a<br />

"maker"-active, manipulating, and exploring. Moreover, science<br />

is now capable of respecting the nature it investigates.<br />

Out of the dialogue with nature initiated by classical science,<br />

with its view of nature as an automaton, has grown a quite<br />

different view in which the activity of questioning nature is<br />

part of its intrinsic activity.<br />

As we have written at the start of this chapter, our feeling of<br />

intellectual security has been shattered. We can now appreciate<br />

in a nonpolemical fashion the relation between science and<br />

philosophy. We have already mentioned the Einstein-Bergson<br />

conflict. Bergson was certainly "wrong" on some technical<br />

points, but his task as a philosopher was to attempt to make


ORDER OUT OF CHAOS 302<br />

explicit inside physics the aspects of time he thought science<br />

was neglecting.<br />

Exploring the implications and the coherence of those fundamental<br />

concepts, which appear both scientific and philosophical,<br />

may be risky, but it can be very fruitful in the dialogue<br />

between science and philosophy. Let us illustrate this with<br />

some brief references to Leibniz, Peirce, Whitehead, and Lucretius.<br />

Leibniz introduced the strange concept of monads, noncommunicating<br />

physical entities that have "no windows through<br />

which something can get in or out. " His views have often been<br />

dismissed as mad, and still, as we have seen in Chapter 11, it is<br />

an essential property of all integrable systems that there exist<br />

a transformation that may be described in terms of noninteracting<br />

entities. These entities translate their own initial<br />

state throughout their motion, but at the same time, like monads,<br />

they coexist with all the others in a "preestablished" harmony:<br />

in this representation, the state of each entity, although<br />

perfectly self-determined, reflects the state of the whole system<br />

down to the smallest detail.<br />

All integrable systems thus can be viewed as "monadic" systems.<br />

Conversely, Leibnizian monadology can be translated into<br />

dynamic language: the universe is an integrable system.13<br />

Monadology thus becomes the most consequential formulation<br />

of a universe from which all becoming is eliminated. By<br />

considering Leibniz's efforts to understand the activity of matter,<br />

we can measure the gap that separates the seventeenth<br />

century from our time. The tools were not yet ready; it was<br />

impossible, on the basis of a purely mechanical universe, for<br />

Leibniz to give an account of the activity of matter. Still some<br />

of his ideas, that substance is activity, that the universe is an<br />

interrelated unit, remain with us and are today taking on a new<br />

form.<br />

We regret that we cannot devote sufficient space to the work<br />

of Charles S. Peirce. At least let us cite one remarkable passage:<br />

You have all heard of the dissipation of energy. It is found<br />

that in all transformations of energy a part is converted<br />

into heat and heat is always tending to equalize its tem-


303 FROM EARTH TO HEAVEN- THE REENCHANTMENT OF NATURE<br />

perature. The consequence is that the energy of the universe<br />

is tending by virtue of its necessary laws toward a<br />

death of the universe in which there shall be no force but<br />

heat and the temperature everywhere the same . . . .<br />

But although no force can counteract this tendency,<br />

chance may and will have the opposite influence. Force is<br />

in the long run dissipative; chance is in the long run concentrative.<br />

The dissipation of energy by the regular laws<br />

of nature is by these very laws accompanied by circumstances<br />

more and more favorable to its reconcentration<br />

by chance. There must therefore be a point at which the<br />

two tendencies are balanced and that is no doubt the actual<br />

condition of the whole universe at the present time. J4<br />

Once again, Peirce's metaphysics was considered as one more<br />

example of a philosophy alienated from reality. But, in fact,<br />

today Peirce's work appears to be a pioneering step toward the<br />

understanding of the pluralism involved in physical laws.<br />

Whitehead's philosophy takes us to the other end of the spectrum.<br />

For him, being is inseparable from becoming. Whitehead<br />

wrote: "The elucidation of the meaning of the sentence 'everything<br />

flows' is one of metaphysics' main tasks." 15 Physics and<br />

metaphysics are indeed coming together today in a conception<br />

of the world in which process, becoming, is taken as a primary<br />

constituent of physical existence and where, unlike Leibniz'<br />

monads, existing entities can interact and therefore also be<br />

born and die.<br />

The ordered world of classical physics, or a monadic theory<br />

of parallel changes, resembles the equally parallel, ordered,<br />

and eternal fall of Lucretius' atoms through infinite space. We<br />

have already mentioned the clinamen and the instability of<br />

laminar flows. But we can go farther. As Serresi6 points out,<br />

the infinite fall provides a model on which to base our conception<br />

of the natural genesis of the disturbance that causes<br />

things to be born. If the vertical fall were not disturbed "without<br />

reason" by the clinamen, which leads to encounters and<br />

associations between uniformly falling atoms, no nature could<br />

be created; all that would be reproduced would be the repetitive<br />

connection between equivalent causes and effects governed<br />

by the laws of fate (foedera fati).


ORDER OUT OF CHAOS 304<br />

Denique si semper motus conectitur omnls<br />

et uetere exoritur [semper] novus ordine certo<br />

nee declinando faciunt primordia motus<br />

principium quoddam quod fatl foedera rumpat,<br />

ex lnfinito ne causa m causa sequatur,<br />

libera per terras unde haec animantibus exstat . . . ?17<br />

Lucretius may be said to have invented the clinamen in the<br />

same way that archaeological remains are "invented": one<br />

"guesses" they are there before one begins to dig. If only uniformly<br />

reversible trajectories existed, where would the irreversible<br />

processes we produce and experience come from?<br />

The point where the trajectories cease to be determined,<br />

where thefoederafati governing the ordered and monotonous<br />

world of deterministic change break down, marks the beginning<br />

of nature. It also marks the beginning of a new science<br />

that describes the birth, proliferation, and death of natural<br />

beings. "The physics of falling, of repetition, of rigorous concatenation<br />

is replaced by the creative science of change and<br />

circumstances. "18 The foedera fati are replaced by the<br />

foedera naturae, which, as Serres emphasizes, denote both<br />

"laws" of nature-local, singular, historical relations-and an<br />

"alliance," a form of contract with nature.<br />

In Lucretian physics we thus again find the link we have<br />

discovered in modern knowledge between the choices underlying<br />

a physical description and a philosophic, ethical, or religious<br />

conception relating to man's situation in nature. The<br />

physics of universal connections is set against another science<br />

that in the name of law and domination no longer struggles<br />

with disturbance or randomness. Classical science from Archime<br />

des to Clausius was opposed to the science of turbulence<br />

and of bifurcating changes.<br />

It is here that Greek wisdom reaches one of its pinnacles.<br />

Where man is in the world, of the world, in matter, of<br />

matter, he is not a stranger, but a friend, a member of the<br />

family, and an equal. He has made a pact with things.<br />

Conversely, many other systems and many other sciences<br />

are based on breaking this pact. Man is a stranger<br />

to the world, to the dawn, to the sky, to things. He hates


305<br />

FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

them, and fights them. His environment is a dangerous<br />

enemy to be fought, to be kept enslaved . . • . Epicurus<br />

and Lucretius live in a reconciled universe. Where the<br />

science of things and the science of man coincide. I am a<br />

disturbance, a whirlwind in turbulent nature. 1<br />

Beyond Tautology<br />

The world of classical science was a world in which the only<br />

events that could occur were those deducible from the instantaneous<br />

state of the system. Curiously, this conception,<br />

which we have traced back to Galileo and Newton, was not<br />

new at that time. Indeed, it can be identified with Aristotle's<br />

conception of a divine and immut a ble heaven. In Aristotle's<br />

opinion, it was only the heavenly world to which we could<br />

hope to apply an exact mathematical description. In the Introduction,<br />

we echoed the complaint that science has "disenchanted"<br />

the world. But this disenchantment is paradoxically<br />

due to the glorification of the earthly world, henceforth<br />

worthy of the kind of intellectual pursuit Aristotle reserved for<br />

heaven. Classical science denied becoming, natural diversity,<br />

both considered by Aristotle as attributes of the sublunar, inferior<br />

world. In this sense, classical science brought heaven to<br />

earth. However, this apparently was not the intention of the<br />

fathers of modern science. In challenging Aristotle's claim that<br />

mathematics ends where nature begins, they did not seek to<br />

discover the immutable concealed behind the changing, but<br />

rather to extend changing, corruptible nature to the boundaries<br />

of the universe. In his Dialogue Concerning the Two<br />

Chief World Systems, Galileo is amazed at the notion that the<br />

world would be a nobler place if the great flood had left only a<br />

sea of ice behind, or if the earth had the incorruptible hardness<br />

of jasper; let those who think the earth would be more<br />

beautiful after being changed into a crystal ball be transformed<br />

by Medusa's stare into a diamond statue!<br />

But the objects chosen by the first physicists to explore the<br />

validity of a quantitative description-that is, the ideal pendulum<br />

with its conservative motion, simple machines, planetary<br />

orbits, etc.-were found to correspond to a unique mathemati-


ORDER OUT OF CHAOS<br />

306<br />

cal description that actually reproduced the divine ideality of<br />

Aristotle's heavenly bodies.<br />

Like Aristotle's gods, the objects of classical dynamics are<br />

concerned only with themselves. They can learn nothing from<br />

the outside. At any instant, each point in the system knows all<br />

it will ever need to know-that is, the distribution of masses in<br />

space and their velocities. Each state contains the whole truth<br />

concerning all possible other states, and each can be used to<br />

predict the others, whatever their respective positions on the<br />

time axis. In this sense, this description leads to a tautology,<br />

since both future and past are contained in the present.<br />

The radical change in the outlook of modern science, the<br />

transition toward the temporal, the multiple, may be viewed as<br />

the reversal of the movement that brought Aristotle's heaven to<br />

earth. Now we are bringing earth to heaven. We are discovering<br />

the primacy of time and change, from the level of elementary<br />

particles to cosmological models.<br />

Both at the macroscopic and microscopic levels, the natural<br />

sciences have thus rid themselves of a conception of objective<br />

reality that implied that novelty and diversity had to be denied<br />

in the name of immutable universal laws. They have rid themselves<br />

of a fascination with a rationality taken as closed and a<br />

knowledge seen as nearly achieved. They are now open to the<br />

unexpected, which they no longer define as the result of imperfect<br />

knowledge or insufficient<br />

·<br />

control.<br />

This opening up of science has been well defined by Serge<br />

Moscovici as the "Keplerian revolution," to distinguish it<br />

from the "Copernican revolution" in which the idea of an absolute<br />

point of view was maintained. In many of the passages<br />

cited in the Introduction to this book, science was likened to a<br />

"disenchantment" of the world. Let us quote Moscovici's description<br />

of the changes going on in the sciences today:<br />

Science has become involved in this adventure, our adventure,<br />

in order to renew everything it touches and<br />

warm all that it penetrates-the earth on which we live<br />

and the truths which enable us to live. At each turn it is<br />

not the echo of a demise, a bell tolling for a passing away<br />

that is heard, but the voice of rebirth and beginning, ever<br />

afresh, of mankind and materiality, fixed for an instant in<br />

their ephemeral permanence. That is why the great dis-


307<br />

FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

coveries are not revealed on a deathbed like that of<br />

Copernicus, but offered like Kepler's on the road of<br />

dreams and passion.2o<br />

The Creative Course of Time<br />

It is often said that without Bach we would not have had the<br />

"St. Matthew Passion" but that relativity would have been discovered<br />

without Einstein. Science is supposed to take a deterministic<br />

course, in contrast with the unpredictability involved<br />

in the history of the arts. When we look back on the strange<br />

history of science, three centuries of which we have tried to<br />

outline, we may doubt the validity of such assertions. There<br />

are striking examples of facts that have been ignored because<br />

the cultural climate was not ready to incorporate them into a<br />

consistent scheme. The discovery of chemical clocks probably<br />

goes back to the nineteenth century, but their result seemed to<br />

contradict the idea of uniform decay to equilibrium. Meteorites<br />

were thrown out of the Vienna museum because there<br />

was no place for them in the description of the solar system.<br />

Our cultural environment plays an active role in the questions<br />

we ask, but beyond matters of style and social acceptance, we<br />

can identify a number of questions to which each generation<br />

returns.<br />

The question of time is certainly one of those questions.<br />

Here we disagree somewhat with Thomas Kuhn's analysis of<br />

the formation of "normal" science.21 Scientific activity best<br />

corresponds to Kuhn's view when it is considered in the context<br />

of the contemporary university, in which research and the<br />

training of future researchers is combined. Kuhn's analysis, if<br />

it is taken as a description of science in general, leading to<br />

conclusions about what knowledge must be, can be reduced to<br />

a new psychosocial version of the positivist conception of scientific<br />

development, namely, the tendency to increasing specialization<br />

and compartmentalization; the identification of<br />

"normal" scientific behavior with.that of the "serious," "silent"<br />

researcher who wastes no time on "general" questions<br />

about the overall significance of his research but sticks to specialized<br />

problems; and the essential independence of scientific<br />

development from cultural, economic, and social problems.


ORDER OUT OF CHAOS<br />

306<br />

The academic structure in which the "normal science" de·<br />

scribed by Kuhn came into being took shape in the nineteenth<br />

century. Kuhn emphasizes that it is by repeating in the form of<br />

exercises solutions to the paradigmatic problems of previous<br />

generations that students learn the concepts upon which research<br />

is based. It is in this way that they are given the criteria<br />

that define a problem as interesting and a solution as acceptable.<br />

The transition from student to researcher takes place<br />

gradually; the scientist continues to solve problems using similar<br />

techniques.<br />

Even in our time, for which Kuhn's description has the<br />

greatest relevance, it refers to only one specific aspect of scientific<br />

activity. The importance of this aspect varies according<br />

to the individual researchers and the institutional context.<br />

In Kuhn's view the transformation of a paradigm appears as<br />

a crisis: instead of remaining a silent, almost invisible rule,<br />

instead of remaining unspoken, the paradigm is actually questioned.<br />

Instead of working in unison, the members of the community<br />

begin to ask "basic" questions and challenge the<br />

legitimacy of their methods. The group, which by training was<br />

homogeneous, now diversifies. Different points of view, cultural<br />

exper.iences, and philosophic convictions are now expressed<br />

and often play a decisive role in the discovery of a new<br />

paradigm. The emergence of the new paradigm further increases<br />

the vehemence of the debate. The rival paradigms are<br />

put to the test until the academic world determines the victor.<br />

With the appearance of a new generation of scientists, silence<br />

and unanimity take over again. New textbooks are written,<br />

and once again things "go without saying."<br />

In this view the driving force behind scientific innovation is<br />

the intensely conservative behavior of scientific communities,<br />

which stubbornly apply to nature the same techniques, the<br />

same concepts, and always end up by encountering an equally<br />

stubborn resistance from nature. When nature is eventually<br />

seen as refusing to express itself in the accepted language, the<br />

crisis explodes with the kind of violence that results from a<br />

breach of confidence. At this stage, all intellectual resources<br />

are concentrated on the search for a new language. Thus scientists<br />

have to deal with crises imposed upon them against<br />

their will.<br />

The questions we have investigated have led us to emphasize


309<br />

FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

aspects that differ considerably from those to which Kuhn's<br />

description applies. We have dwelled on continuities, not the<br />

"obvious" continuities but the hidden ones, those involving<br />

difficult questions rejected by many as illegitimate or false but<br />

that keep coming back generation after generation-questions<br />

such as the dynamics of complex systems, the relation of the<br />

irreversible world of chemistry and biology with the reversible<br />

description provided by classical physics. In fact, the interest<br />

of such questions is hardly surprising. To us, the problem is<br />

rather to understand how they could ever have been neglected<br />

after the work of Diderot, Stahl, Venel, and others.<br />

The past one hundred years have been marked by several<br />

crises that correspond closely to the description given by<br />

Kuhn-none of which were sought by scientists. Examples<br />

are the discovery of the instability of elementary particles, or<br />

of the evolving universe. However, the recent history of science<br />

is also characterized by a series of problems that are the<br />

consequences of deliberate and lucid questions asked by scientists<br />

who knew that the questions had both scientific and<br />

philosophical aspects. Thus scientists are not doomed to behave<br />

like "hypnons"!<br />

It is important to point out that the new scientific development<br />

we have described, the incorporation of irreversibility<br />

into physics, is not to be seen as some kind of "revelation,"<br />

the possession of which would set its possessor apart from the<br />

cultural world he lives in. On the contrary, this development<br />

clearly reflects both the internal logic of science and the<br />

cultural and social context of our time.<br />

In particular, how can we consider as accidental that the<br />

rediscovery of time in physics is occurring at a time of extreme<br />

acceleration in human history? Cultural context cannot be the<br />

complete answer, but it cannot be denied either. We have to<br />

incorporate the complex relations between "internal" and<br />

"external" determinations of the production of scientific concepts.<br />

In the preface of this book, we have emphasized that its<br />

French title (La nouvelle alliance) expresses the coming together<br />

of the "two cultures". Perhaps the confluence is nowhere<br />

as clear as in the problem of the microscopic<br />

foundations of irreversibility we have studied in Book Three.<br />

As mentioned repeatedly, both classical and quantum me-


ORDER OUT OF CHAOS 310<br />

chanics are based on arbitrary initial conditions and deterministic<br />

laws (for trajectories or wave functions). In a sense,<br />

laws made simply explicit what was already present in the initial<br />

conditions. This is no longer the case when irreversibilty is<br />

taken into account. In this perspective, initial conditions arise<br />

from previous evolution and are transformed into states of the<br />

same class through subsequent evolution.<br />

We come therefore close to the central problem of Western<br />

ontology: the relation between Being and Becoming. We have<br />

given a brief account of the problem in Chapter III. It is<br />

remarkable that two of the most influential works of the century<br />

were precisely devoted to this problem. We have in mind<br />

Whitehead's Process and Reality and Heidegger's Sein und<br />

Zeit. In both cases, the aim is to go beyond the identification<br />

of Being with timelessness, following the Voie Royale of western<br />

philosophy since Plato and Aristotle.22<br />

But obviously, we cannot reduce Being to Time, and we<br />

cannot deal with a Being devoid of any temporal connotation.<br />

The direction which the microscopic theory of irreversibility<br />

takes gives a new content to the speculations of Whitehead<br />

and Heidegger.<br />

It would go beyond the aim of this book to develop this problem<br />

in greater detail; we hope to do it elsewhere. Let us notice<br />

that initial conditions, as summarized in a state of the system,<br />

are associated with Being; in contrast, the laws involving temporal<br />

changes are associated with Becoming.<br />

In our view, Being and Becoming are not to be opposed one<br />

to the other: they express two related aspects of reality.<br />

A state with broken time symmetry arises from a law with<br />

broken time symmetry, which propagates it into a state belonging<br />

to the same category.<br />

In a recent monograph (From Being to Becoming), one of<br />

the authors concluded in the following terms: "For most of the<br />

founders of classical science-even for Einstein-science was<br />

an attempt to go beyond the world of appearances, to reach a<br />

timeless world of supreme rationality-the world of Spinoza.<br />

But perhaps there is a more subtle form of reality that involves<br />

both laws and games, time and eternity. "<br />

This is precisely the direction which the microscopic theory<br />

of irreversible processes is taking.


311 FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

The Human Condition<br />

We agree completely with Herman Weyl:<br />

Scientists would be wrong to ignore the fact that theo<br />

retical construction is not the only approach to the phenomena<br />

of life; another way, that of understanding from<br />

within (interpretation), is open to us . . . . Of myself, of<br />

my own acts of perception, thought, volition, feeling and<br />

doing, I have a direct knowledge entirely different from<br />

the theoretical knowledge that represents the "parallel"<br />

cerebral processes. in symbols. This inner awareness of<br />

myself is the basis for the understanding of my fellowmen<br />

whom I meet and acknowledge as beings of my own<br />

kind, with whom I communicate sometimes so intimately<br />

as to share joy and sorrow with them. 23<br />

Until recently, however, there was a striking contrast. The external<br />

universe appeared to be an automaton following deterministic<br />

causal laws, in contrast with the spontaneous activity<br />

and irreversibility we experience. The two worlds are now<br />

drawing closer together. Is this a loss for the natural sciences?<br />

Classical science aimed at a "transparent" view of the physical<br />

universe. In each case you would be able to identify a<br />

cause and an effect. Whenever a stochastic description becomes<br />

necessary, this is no longer so. We can no longer speak<br />

of causality in each individual experiment; we can only speak<br />

about statistical causality. This has, in fact, been the case ever<br />

since the advent of quantum mechanics, but it has been greatly<br />

amplified by recent developments in which randomness and<br />

probability play an essential role, even in classical dynamics<br />

or chemistry. Therefore, the modern trend as compared to the<br />

classical one leads to a kind of "opacity" as compared to the<br />

transparency of classical thought.<br />

Is this a defeat for the human mind? This is a difficult question.<br />

As scientists, we have no choice; we cannot describe for<br />

you the world as we would like to see it, but only as we are<br />

able to see it through the ombined impat of experimental


ORDER OUT OF CHAOS 312<br />

results and new theoretical concepts. Also, we believe that<br />

this new situation reflects the situation we seem to find in our<br />

own mental activity. Classical psychology centered around<br />

conscious, transparent activity; modern psychology attaches<br />

much weight to the opaque functioning of the unconscious.<br />

Perhaps this is an image of the basic features of human existence.<br />

Remember Oedipus, the lucidity of his mind in front of<br />

the sphinx and its opacity and darkness when confronted with<br />

his own origins. Perhaps the coming together of our in sights<br />

about the world around us and the world inside us is a satisfying<br />

feature of the recent evolution in science that we have tried<br />

to describe.<br />

It is hard to avoid the impression that the distinction between<br />

what exists in time, what is irreversible, and, on the<br />

other hand, what is outside of time, what is eternal, is at the<br />

origin of human symbolic activity. Perhaps this is especially so<br />

in artistic activity. Indeed, one aspect of the transformation of<br />

a natural object, a stone, to an object of art is closely related to<br />

our impact on matter. Artistic activity breaks the temporal<br />

symmetry of the object. It leaves a mark that translates our<br />

temporal dissymmetry into the temporal dissymmetry of the<br />

object. Out of the reversible, nearly cyclic noise level in which<br />

we live arises music that is both stochastic and time-oriented.<br />

The Renewal of Nature<br />

It is quite remarkable that we are at a moment both of profound<br />

change in the scientific concept of nature and of the<br />

structure of human society as a result of the demographic explosion.<br />

As a result, there is a need for new relations between<br />

man and nature and between man and man. We can no longer<br />

accept the old a priori distinction between scientific and ethical<br />

values. This was possible at a time when the external world<br />

and our internal world appeared to conflict, to be nearly<br />

orthogonal. Today we know that time is a construction and<br />

therefore carries an ethical responsibility.<br />

The ideas to which we have devoted much space in this<br />

book-the ideas of instability, of fluctuation-diffuse into the<br />

social sciences. We know now that societies are immensely


313 FROM EARTH TO HEAVEN-THE REENCHANTMENT OF NATURE<br />

complex systems involving a potentially enormous number of<br />

bifurcations exemplified by the variety of cultures that have<br />

evolved in the relatively short span of human history. We know<br />

that such systems are highly sensitive to fluctuations. This<br />

leads both to hope and a threat: hope, since even small fluctuations<br />

may grow and change the overall structure. As a result,<br />

individual activity is not doomed to insignificance. On<br />

the other hand, this is also a threat, since in our universe the<br />

security of stable, permanent rules seems gone forever. We are<br />

living in a dangerous and uncertain world that inspires no<br />

blind confidence, but perhaps only the same feeling of<br />

qualified hope that some Talmudic texts appear to have attributed<br />

to the God of Genesis:<br />

1\venty-six attempts preceded the present genesis, all of<br />

which were destined to fail. The world of man has arisen<br />

out of the chaotic heart of the preceding debris; he too is<br />

exposed to the risk of failure, and the return to nothing.<br />

"Let's hope it works" [Halway Sheyaamod] exclaimed<br />

God as he created the World, and this hope, which has<br />

accompanied all the subsequent history of the world and<br />

mankind, has emphasized right from the outset that this<br />

history is branded with the mark of radical uncertainty.24


NOTES<br />

Introduction<br />

I. I. BERLIN, Against the Current, selected writings ed. H. Hardy<br />

(New York: The Viking Press, 1980), p. xxvi.<br />

2. See TITUS LucRETIUS CARUS, De Natura Rerum, Book I, v.<br />

267-70. ed . and comm. C. Bailey (Oxford: Oxford University<br />

Press 1947, 3 vols.)<br />

3. R. LENOBLE, Histoire de /'idee de nature (Paris: Albin Michel,<br />

1969).<br />

4. B. PASCAL, "Pensees," frag. 792, in Oeuvres Completes (Paris:<br />

Brunschwig-Boutroux-Gazier, 1904-14).<br />

5. J. MoNOD, Chance and Necessity (New York: Vintage Books,<br />

1972), pp. 172-73.<br />

6. G. V1co, The New Science, trans. T. G. Bergin and M. H. Fisch<br />

(New York: 1968), par. 331.<br />

7. J. P. VERNANT et al., Divination et rationalite, esp. J. BOTTERO,<br />

"Symptomes, signes, ecritures" (Paris: Seuil, 1974).<br />

8. A. KoYRE , Galileo Studies (Hassocks, Eng.: The Harvester<br />

Press, 1978).<br />

9. K. PoPPER, Objective Knowledge (Oxford: Clarendon Press,<br />

1972).<br />

10. P. FoRMAN, "Weimar Culture, Causality and Quantum Theory,<br />

1918-1927; Adaptation by German Physicists and Mathematicians<br />

to an Hostile Intellectual Environment," Historical Studies<br />

in Physical Sciences, Vol. 3 (1971), pp. 1-1 15.<br />

11. J. NEEDHAM and C. A. RoNAN, A Shorter Science and Civilization<br />

in China, Vol. I (Cambridge: Cambridge University Press,<br />

1978), p. 170.<br />

12. A. EDDINGTON, The Nature of the Physical World (Ann Arbor:<br />

University of Michigan Press, 1958), pp. 68-80.<br />

13. Ibid., p. 103.<br />

14. BERLIN, op. cit., p. 109.<br />

15. K. POPPER, Unended Quest (La Salle, Ill.: Open Court Publishing<br />

Company, 1976), pp. 161-62.<br />

16. G. BRUNO, 5th dialogue, "De Ia causa," Opere ltaliane, I (Bari:<br />

1907); cf. I. LECLERC, The Nature of Physical Existence<br />

(London: Ge<strong>org</strong>e Allen & Unwin, 1972).<br />

315


ORDER OUT OF CHAOS 316<br />

17. P. VA LRY, Cahiers, (2 vols.) ed. Mrs. Robinson-Valery, (Paris:<br />

Gallimard, 1973-74).<br />

18. E. ScHRODINGER, "A re there Quantum Jumps?" The British<br />

Journal fo r the Philosophy of Science, Vol. III (1952), pp. 109-<br />

10; this text has been quoted with indignation by P. W. Bridgmann<br />

in his contribution to Determinism and Freedom in the<br />

Age of Modern Science, ed. S. Hook (New York: New York<br />

University Press, 1958).<br />

19 .. A. EINSTEIN, "Prinzipien der Forschung, Rede zur 60.<br />

Geburstag van Max Planck" (1918) in Mein Weltbild, Ullstein<br />

Verlag 1977, pp. 107-10, trans. Ideas and Opinions (New York:<br />

Crown, 1954), pp. 224-27.<br />

20. R DDRRENMATT, The Physicists. (New York: Grove, 1964).<br />

21. S. MoscoviCI, Essai sur /'histoire humaine de Ia nature, Collection<br />

Champs (Paris: Flammarion, 1977).<br />

22. Quoted in Ronan, op. cit., p. 87.<br />

23. MoNon , op. cit., p. 180.<br />

Chapter 1<br />

1. J. T. DESAGULIERS, "The Newtonian System of the World, The<br />

Best Model of Government: an Allegorical Poem," 1728, quoted<br />

in H. N. FA IRCHILD, Religious Trends in English Poetry, Vol. I<br />

(New York: Columbia University Press, 1939), p. 357.<br />

2. Ibid., p. 358.<br />

3. Gerd Buchdahl emphasized and illustrated this ambiguity of the<br />

cultural influence of the Newtonian model in its dimensions both<br />

empirical (Opticks) and systematic (Principia) in The Image of<br />

Newton and Locke in the Age of Reason, Newman History and<br />

Philosophy of Science Series (London: Sheed & Ward, 1961).<br />

4. La Science et Ia diversite des cultures, (Paris: UNESCO, PUF,<br />

1974), pp. 15-16.<br />

5. C. C. GILLISPIE, The, Edge of Objectivity (Princeton, N.J.:<br />

Princeton University Press, 1970), pp. 199-200.<br />

6. M. HEIDEGGER, The Question Concerning Te chnology (New<br />

York: Harper & Row, 1977), p. 20.<br />

7. Ibid., p. 21.<br />

8. Ibid., p. 16.<br />

9. "The Coming of the Golden Age," Paradoxes of Progress (San<br />

Francisco: Freeman & Company, 1978).<br />

10. See, for instance. P. DAVIES, Other Worlds (Toronto: J. M. Dent<br />

& Sons, 1980).


317 NOTES<br />

11. A. K oESTLER, The Roots of Coincidence (London: Hutchinson,<br />

1972), pp. 138-39.<br />

12. A. KoYRE, Newtonian Studies (Chicago: University of Chicago<br />

Press, 1968), pp. 23-24.<br />

13. In "Race and History" (Structural Anthropology II, New York:<br />

Basic Books, 1976), Claude Levi-Strauss discusses the conditions<br />

that lead to the Neolithic and Industrial revolutions. The<br />

model he introduces, involving chain reactions and catalysis (a<br />

process with kinetics characterized by threshold and amplification<br />

phenomena) attests to an affinity between the problems of<br />

stability and fluctuation we discuss in Chapter VI as well as certain<br />

themes of the "structural approach" in anthropology.<br />

14. "Inside each society, the order of myth excludes dialogue: the<br />

group's myths are not discussed, they are transformed when<br />

they are thought to be repeated." C. LEvi-STRAUSS, L'Homme<br />

Nu (Paris: Pion, 1971), p. 585. Thus mythical discourse is to be<br />

distinguished from critical (scientific and philosophic) dialogue<br />

more because of the practical conditions of its reproduction than<br />

because of an intrinsic inability of such or such emitter to think<br />

in a rational way. The practice of critical dialogue has given to<br />

the cosmological discourse claiming truthfulness its spectacular<br />

evolutive acceleration.<br />

15. This is, of course, one of the main themes of Alexandre Koyre.<br />

16. The definition of such an "absurdity" opposes the age-long idea<br />

that a sufficiently tricky device would permit one to cheat nature.<br />

See the efforts devoted by engineers till the twentieth century<br />

to the construction of perpetual-motion machines in A. Ord<br />

Hume, Perpetual Motion: The History of an Obsession (New<br />

York: St. Martin's Press, 1977).<br />

17. Popper translated into a norm this excitement born out of the<br />

risks involved in the experimental games. He affirms, in The<br />

Logic of Scientific Discovery, that the scientific must look for<br />

the most "improbable" hypothesis-that is, the most risky<br />

one-to try to refute it as well as the corresponding theories.<br />

18. R. FEYNMAN, The Character of Physical Law (Cambridge,<br />

Mass.: M.I.T. Press, 1967), second chapter.<br />

19. J. NEEDHAM, "Science and Society in East and West," The<br />

Grand Titration (London: Allen & Unwin, 1969).<br />

20. A. N. WHITEHEAD, Science and the Modern World (New York:<br />

The Free Press, 1967), p. 12.<br />

21. NEEDHAM, op. cit., p. 308.<br />

22. NEEDHAM, op. cit., p. 330.<br />

23. R. HooYKAAS emphasized this "dedivinization" of the world by<br />

the Christian metaphor of the world machine in Religion and the


ORDER OUT OF CHAOS 318<br />

Rise of Modern Science (Edinburgh and London: Scottish Academic<br />

Press, 1972), esp. pp. 14-16.<br />

24. WHITEHEAD, Op. cit., p. 54.<br />

25. The famous text about nature being written in mathematical<br />

signs is to be fo und in 11 Saggiatore. See also The Dialogue Concerning<br />

the Two Chief World Systems, 2nd rev. ed. (Berkeley:<br />

University of California Press, 1967).<br />

26. At least it was triumphant in the academies created in France,<br />

Prussia, and Russia by absolute sovereigns. In The Scientist's<br />

Role in Society (Englewood Cliffs, N.J.: Foundations of Modern<br />

Sociology Series, Prentice-Hall, 1971), Ben David emphasized<br />

the distinction between physicists of these countries, dedicated<br />

to physics as a glamorous and purely theoretical science, and<br />

the English physicists immerged in a wealth of empirical and<br />

technical problems. Ben David proposed a connection between<br />

the fascination for a theoretical science and the relegation far<br />

from political power of the social class supporting the "scientific<br />

movement. "<br />

27. In his biography of dlembert-Jean d'Alembert, Science and<br />

Enlightenment (Oxford: Clarendon Press, 1970)-Thomas<br />

Hankins emphasized how closed and small was the first true<br />

scientific community, in the modern sense of the term, namely,<br />

that of the eighteenth-century physicists and mathematicians,<br />

and how intimate were their relations with. the absolute sovereigns.<br />

28. EINSTEIN, Op. cit., pp. 225-26.<br />

29. E. MACH, "The Economical Nature of Physical Inquiry, " Popular<br />

Scientific Lectures (Chicago: Open Court Publishing Company,<br />

1895), pp. 197-98.<br />

30. J. DONNE, An Anatomy of the World wherein . .. the frailty and<br />

the decay of the whole world is represented (London, catalog of<br />

the British Museum, 161 1).<br />

Chapter 2<br />

I. On this point, see T. HANKINS, "The Reception of Newton's<br />

Second Law of Motion in the Eighteenth Century, " Archives Internationales<br />

d'Histoire des Sciences, Vol. XX (1967), pp. 42-<br />

65, and I. B. CoHEN, "Newton's Second Law and the Concept<br />

of Force in the Principia," The Annus Mirabilis of Sir Isaac<br />

Newton, Tricentennial Celebration, The Texas Quarterly,<br />

Vol. X, No. 3 (1967), pp. 25-157. The four following paragraphs<br />

rest, for what concerns atomism and the conservation theories,


319 NOTES<br />

on W. Scarr, The Conflict Between Atomism and Conservation<br />

Theory (London: Macdonald, 1970).<br />

2. A. KovRE: , Galileo Studies (Hassocks, Eng.: The Harvester<br />

Press, 1978), pp. 89-94.<br />

3. In his history of mechanics-The Science of Mechanics: A Critical<br />

and Historical Account of Its Development (La Salle, Ill.:<br />

Open Court Publishing Company, 1960)-Ernst Mach laid stress<br />

on this dual filiation of modern dynamics of both the trajectories<br />

science and the engineer's computations.<br />

4. This at least is the conclusion of today's historians who began<br />

the study of the impressive mass of Newton's ·tchemical Papers,"<br />

which till now were ignored or disdained as "nonscientific."<br />

See B. J. DoBBS, The Foundations of Newton's Alchemy<br />

(Cambridge: Cambridge University Press, 1975); R. WESTFALL,<br />

"Newton and the Hermetic Tradition" in Science, Medicine and<br />

Society, ed. A. G. DEBUS (London: Heinemann, 1972); and R.<br />

WESTFALL, "The Role of Alchemy in Newton's Career," Reason,<br />

Experiment and Mysticism, ed. M. L. RIGHINI BONELLI<br />

and W. R. SHEA (London: Macmillan, 1975). As Lord Keynes,<br />

who played a crucial part in the collection of these papers, summarized<br />

(quoted in DoBBS, op. cit., p. 13), "Newton was not the<br />

first of the age of reason. He was the last of the Babylonians and<br />

Sumerians, the last great mind which looked out on the visible<br />

and intellectual world with the same eyes as those who began to<br />

build our intellectual inheritance rather less than 10,000 years<br />

ago."<br />

5. DoBBS, op. cit., also examined the role of the "mediator" by<br />

which two substances are made "sociable." We may recall here<br />

the importance of the mediator in Goethe's Elective Affinities<br />

(Engl. trans. Greenwood 1976). For what concerns chemistry,<br />

Goethe was not far from Newton.<br />

6. The story of Newton's "mistake" is told in HANKINs's, Jean<br />

d'Alembert, pp. 29-35.<br />

7. G. L. BuFFON, "Reflexions sur Ia loi d'attraction," appendix to<br />

Introduction a I' histoire des minhaux (1774), To me IX of<br />

Oeuvres Completes (Paris: Garnier Frere s), pp. 75, 77.<br />

8. G. L. BuFFON, Histoire naturelle. De Ia Nature, Seconde Vue<br />

(1765), quoted in H. METZGER, Newton, Stahl, Boerhaave et Ia<br />

doctrine chimique (Paris: Blanchard, 1974), pp. 57-58.<br />

9. A. THACKRAY describes the way French chemistry became Buffonian<br />

in Atom and Power: An Essay on Newtonian Matter Th e­<br />

ory and the Development of Chemistry (Cambridge, Mass.:<br />

Harvard University Press, 1 970). pp. 199-233. Berthollet's<br />

Statique chimique accomplished Buffon's program and also


ORDER OUT OF CHAOS 320<br />

closed it, since his disciples gave up the attempt to understand<br />

chemical reactions in terms compatible with Newtonian con·<br />

cepts.<br />

10. We do not wish to try to explain here the reasons of Newton's<br />

triumph in France, nor of its fall, but to emphasize the at least<br />

chronological connection between these events and the stages of<br />

the process of professionalization of science. See M. CROS·<br />

LAND, The Society of Arcueil: A View of French Science at the<br />

Time of Napoleon (London: Heinemann, 1960), as well as his<br />

Gay Lussac (Cambridge: Cambridge University Press, 1978).<br />

11. Thomas Kuhn made of this role of scientific institutions, taking<br />

over the formation of the future scientists-that is assuring their<br />

own reproduction, the main characteristic of scientific activity<br />

as we know it today. This problem has also been approached by<br />

M. Crosland, R. Hahn, and W. Farrar in The Emergence of Science<br />

in Western Europe, ed. M. CROSLAND (London: Mac·<br />

millan, 1975).<br />

12. The role of "mundane" interest so despised by philosophers<br />

such as Gaston Bachelard in France should be taken as the sign<br />

of the open character of eighteenth-century science. In a way,<br />

we can truly speak about a regression during the nineteenth century,<br />

at least for what concerns the scientific culture. And we<br />

could learn today from the multiplicity of local academies and<br />

circles where scientific matters were discussed by nonprofessionals.<br />

13. Quoted in J. ScHLANGER, Les metaphores de I' <strong>org</strong>anisme<br />

(Paris: Vrin, 1971), p. 108.<br />

14. J. C. MAXWELL, Science and Free Will, in CAMPBELL and GAR·<br />

NETT , op. cit., p. 443. L. CAMPBELL & W. GARNETT, The Life of<br />

James Clerk Maxwell (London, Macmillan, 1882).<br />

15. This problem is one of the main themes of French philosopher<br />

Michel Serres. See, for instance, "Conditions" in his La naissance<br />

de Ia physique dans le texte de Lucrece (Paris: Minuit,<br />

1977). Some texts by M. Serres are now available in English<br />

translation, thanks to the pious zeal of the French Studies Department<br />

of Johns Hopkins University. See M. SERRES,<br />

Hermes: Literature, Science, Philosophy. (Baltimore: The<br />

Johns Hopkins University Press, 1982.)<br />

16. See, about the fate of Laplace's demon, E. CASSIRER, Determinism<br />

and Interdeterminism in Modern Physics (New Haven,<br />

Conn.: Yale University Press, 1956), pp. 3-25.


321 NOTES<br />

Chapter 3<br />

1. R. NISBET, History of the Idea of Progress (New York: Basic<br />

Books, 1980), p. 4.<br />

2. D. DIDEROT, d'Alembert's Dream (Harmondsworth:, Eng.: Penguin<br />

Books, 1976), pp. 166-67.<br />

3. D. DIDEROT, "Conversation Between d'Alembert and Diderot,"<br />

d'Alembert's Dream, pp. 158-59.<br />

4. D. DIDEROT, Pensees sur /'Interpretation de Ia Nature (1754),<br />

Oeuvres Completes, Tome II (Paris: Garnier Freres, 1875), p. II.<br />

5. Diderot ascribes this opinion to the physician Bordeu in the<br />

Dream.<br />

6. See, for instance, A. LovEJOY, The Great Chain of Beings<br />

(Cambridge, Mass.: Harvard University Press, 1973).<br />

7. The historian Gillispie proposed a relation between the protest<br />

against mathematical physics, as popularized by Diderot in the<br />

Encyclopedie, and the revolutionaries' hostility against this official<br />

science, as manifested by the closure of the Academy and<br />

Lavoisier's death. This is a very controversial point, but what is<br />

sure is that the Newtonian triumph in France coincides with the<br />

Napoleonic institutions, spelling the final victory of state academy<br />

over craftsmen (see C. C. GILLISPIE, "The Encyclopedia<br />

and the Jacobin Philosophy of Science: A Study in Ideas and<br />

Consequences," Critical Problems in the History of Science, ed.<br />

M. CLAGETT (Madison, Wis.: University of Wisconsin Press,<br />

1959), pp. 255-89.<br />

8. G. E. STAHL, "Veritable Distinction a etablir entre le mixte et le<br />

vivant du corps humain," Oeuvres medicophilosophiques et<br />

pratiques, To me II (Montpellier: Pitrat et Fils, 1861), esp.<br />

pp. 279-82.<br />

9. See J. SCHLANGER, Les metaphores de /'<strong>org</strong>anisme, for a description<br />

of the transformation of the meaning of "<strong>org</strong>anization"<br />

between Stahl and the Romanticists.<br />

10. Philosophy of Nature, §261.<br />

11. This is Knight's conclusion in "The German Science in the Romantic<br />

Period," The Emergence of Science in Western Europe.<br />

12. H. BERGSON, La pensee et le mouvant in Oeuvres (Paris: E ditions<br />

du Centenaire, PUP, 1970), p. 1285 ; trans. The Creative<br />

Mind (Totowa, N.J.: Littlefield, Adams, 1975), p. 42.<br />

13. Ibid., p. 1287; trans., p. 44.<br />

14. Ibid., p. 1286; trans. , p. 44.


ORDER OUT OF CHAOS 322<br />

15. H. BERGSON, L'evolution creatrice in Oeuvres, p. 784; trans.<br />

Creative Evolution (London: Macmillan, 191 1), p. 361.<br />

16. Ibid., p. 538; trans., p. 54.<br />

17. Ibid., p. 784; trans., p. 361.<br />

18. BERGSON, La pensee et /e mouvant, p. 1273; trans., p. 32.<br />

19. Ibid:, p. 1274; trans., p. 33.<br />

20. A. N. WHITEHEAD, Science and the Modern World, p. 55.<br />

21. A. N. WHITEHEAD, Process and Reality: An Essay in Cosmology<br />

(New York: The Free Press, 1969), p. 20.<br />

22. Ibid., p. 26.<br />

23. Joseph Needham and C. H. Waddington both acknowledged the<br />

importance of Whitehead's influence for what concerns their endeavor<br />

to describe in a positive way the <strong>org</strong>anism as a whole.<br />

24. H. HELMHOLTZ, Uber die Erhaltung der Kraft (1847), trans. in<br />

S. BRUSH, Kinetic Theory, Vol. I, The Nature of Gases and<br />

Heat (Oxford: Pergamon Press, 1965), p. 92. See also Y.<br />

ELKANA, Th e Discovery of the Conservation of Energy<br />

(London: Hutchinson Educational, 1974) and P. M. HEIMANN,<br />

"Helmholtz and Kant: The Metaphysical Foundations of Uber<br />

die Erhaltung der Kraft," Studies in the History and Philosophy<br />

of Sciences, Vol. 5 (1974), pp. 205-38.<br />

25. H. REICHENBACH, The Direction of Time (Berkeley: University<br />

of California Press, 1956), pp. 16-17.<br />

Chapter 4<br />

I. W. ScoTT, About the novelty of these problems, see The Conflict<br />

Between Atomism and Conservation Theory, Book II, and about<br />

the industrial context where these concepts were created, D.<br />

CARDWELL, From Wa tt to Clausius (London, Heinemann,<br />

1971). Particularly interesting in this respect is the convergence<br />

between on one hand the need determined by industrial problems<br />

and on the other the positivist simplifications by operational<br />

definitions.<br />

2. J. HERIVEL, Joseph Fourier: The Man and the Physicist (Oxford:<br />

Clarendon Press, 1975). In this biography we learn the following<br />

curious information: Fourier would have brought back<br />

from his trip with Bonaparte to Egypt a sickness causing permanent<br />

deperditions of heat.<br />

3. See, more particularly, the introduction to Comte's Philosophie<br />

Premiere (Paris: Herman, 1975), ·uguste Comte auto-traduit<br />

dans l'encyclopedie" in La Traduction (Paris: Minuit, 1974) and<br />

"Nuage," La Distribution (Paris: Minuit, 1977).


323 NOTES<br />

4. C. SMITH, "Natural Philosophy and Thermodynamics: William<br />

Thomson and the Dynamical Theory of Heat," The British Journal<br />

for the Philosophy of Science, Vol . 9 (1976), pp. 293-3 19 and<br />

M. CROSLAND and C. SMITH, "The Transmission of Physics<br />

from France to Britain, 1800-1840," Historical Studies in the<br />

Physical Sciences, Vol . 9 (1978), pp. 1-61.<br />

5. For what follows, see Y. ELKANA, The Discovery of the Conservation<br />

of Energy Principle, as well as the famous paper by<br />

Thomas Kuhn, "Energy conservation as an Example of Simultaneous<br />

Discovery," originally published in Critical Problems in<br />

the History of Science and recently in T. KuHN, The Essential<br />

Tension (Chicago: University of Chicago Press, 1977).<br />

6. ELKANA followed the slow crystallization of the concept of energy;<br />

see his book and "Helmholtz's Kraft: An Illustration of<br />

Concepts in Flux," Historical Studies in the Physical Sciences,<br />

Vol . 2 (1970), pp. 263-98.<br />

7. J. JouLE, "Matter, Living Force and Heat," The Scientific Papers<br />

of James Prescott Joule, Vol. 1 (London: Taylor & Francis,<br />

1884), pp. 265-76 (quotation, p. 273).<br />

8. The English translations of Mayer's two great papers, "On the<br />

Forces of In<strong>org</strong>anic Nature" and "The Motions of Organisms<br />

and Their Relation to Metabolism," are in Energy: Historical<br />

Development of a Concept, ed. R. B. LINDSAY (Stroudsburg,<br />

Pa.: Benchmarks Papers on Energy 1, Dowden, Hutchinson &<br />

Ross, 1975).<br />

9. E. BENTON, "Vitalism in the Nineteenth Century Scientific<br />

Thought: A 'JYpology and Reassessment," Studies in History<br />

and Philosophy of Science, Vol. 5 (1974), pp. 17-48.<br />

10. H.H ELMHOLTZ, "Uber die Erhaltung der Kraft," op. cit.,<br />

pp. 90-91.<br />

11. G. DELEUZE, Nietzsche et Ia phi/osophie (Paris: PUF, 1973),<br />

pp. 48-55.<br />

12. In this study of Zola's "Docteur Pascal ," Feux et signaux de<br />

brume Paris: Grassel (1975), p. 109, Michel Serres wrote: "The<br />

century that was practically drawing to a close when the novel<br />

appeared had opened with the majestic stability of the solar system,<br />

and was now filled with dismay at the relentless degradations<br />

of fire. Hence the fierce, positive dilemma: perfect cycle<br />

without residue, eternal and positively valued, i.e., the cosmology<br />

of the sun; or else a missed cycle, losing its difference, irreversible,<br />

historical and despised-a cosmology, a thermogony of<br />

fire which must either be extinguished or destroyed, without alternative.<br />

One dreams of Laplace, whilst Carnot and the others<br />

have forever smashed the cubby-hole, the niche, where one


ORDER OUT OF CHAOS 324<br />

could sleep in peace; one is dreaming, that is certain: then<br />

cultural archaisms having returned through another door,<br />

through another opening of the same door, are powerfully reawakened:<br />

immortal flame, purifying blaze or evil fire?"<br />

13. The continuity between Carnot father and son has been emphasized<br />

by Cardwell (From Wa tt to Clausius) and Scott (The Conflict<br />

Between Atomism and Conservation Theory).<br />

14. P. DAVIES, The Runaway Universe (New York: Penguin Books,<br />

1980), p. 197.<br />

15. R DYSON, "Energy in the Universe," Scientific American, Vol.<br />

225 (197 1), pp. 50-59.<br />

1 6. What was particularly important was to grasp that, unlike what<br />

happens in mechanics, it is not just any situation of a thermodynamic<br />

system that can be characterized as a "state"; quite<br />

the contrary. See E. DAUB, "Entropy and Dissipation," Historical<br />

Studies in the Physical Sciences, Vol. 2 (1970), pp. 321-54.<br />

17. In his autobiography, Scientific Autobiography (London:<br />

Williams & N<strong>org</strong>ate, 1950), Max Planck recalled how isolated he<br />

had been when he first emphasized the peculiarity of heat and<br />

pointed out that it is the conversion of heat into another form of<br />

energy that raises the irreversibility problem. Energeticists such<br />

as Ostwald wanted all forms of energy to be given the same<br />

status. For them, the fall of a body between two altitude levels<br />

takes place by virtue of the same kind of productive difference<br />

as the passage of heat between two bodies at different temperatures.<br />

Thus, Ostwald's comparison did away with the crucial<br />

distinction between an ideally reversible process, such as the<br />

mechanical motion, and an intrinsically irreversible one, such as<br />

heat diffusion. By doing so, he was actually taking up a position<br />

similar to what we have attributed to Lagrange: where Lagrange<br />

considered conservation of energy as a property belonging only<br />

to ideal cases but also the only one capable of being treated<br />

rigorously, Ostwald held conservation of energy as the property<br />

of any natural transformation, but defined conservation of energy<br />

differences (required by all transformation since only a difference<br />

can produce another difference) as an abstract ideal, but<br />

the sole object for a rational science.<br />

18. The splitting of the entropy variation into two different terms<br />

was introduced in l. Prigogine, Etude thermodynamique des<br />

Phenomenes irreversibles, These d'agn!gation presentee a Ia<br />

faculte des sciences de l'Universite Libre de Bruxelles 1945<br />

(Paris: Dunod, 1947).<br />

19. R. CLAUSIUS, Ann. Phys., Vol. 125 (1865), p. 353.<br />

20. M. PLANCK, "The Unity of the Physical Universe," A Survey of


325 NOTES<br />

Physics, Collection of Lectures and Essays (New York: E. P.<br />

Dutton, 1925), p. 16.<br />

21. R. CAILLOIS, "La dissymetrie," Coherences aventureuses, Collection<br />

Idees (Paris: Gallimard, 1973), p. 198.<br />

Chapter 5<br />

1. For what concerns the content of this and the fol lowing chapter,<br />

see P. GLANSDORFF and I. PRIGOGINE, Thermodynamic Theory<br />

of Structure, Stability and Fluctuations (New York: John Wiley<br />

& Sons, 1971) and G. NICOLls and I. PRIGOGINE, Self-Organization<br />

in Non-Equilibrium Systems (New York: John Wiley &<br />

Sons, 1977), where further references may be found.<br />

2. R NIETZSCHE, Der Wille zur Macht, Siimtliche Werke (Stuttgart:<br />

Kroner, 1964), aphorism 630.<br />

3. Which precise content can be given to the general law of entropy<br />

growth? For a theoretician physicist such as de Donder, chemical<br />

activity, which appeared obscure and inaccessible to the rational<br />

approach of mechanics, was mysterious enough to<br />

become the synonym of the irreversible process. Thus chemistry,<br />

whose question physicists had never truly answered, and the<br />

new enigma of irreversibility came to join in a challenge not to be<br />

ignored anymore. See Th. De Donder, L' Affinite (Paris:<br />

Gauthier-Villars, 1962) and L. Onsag'er Phys. Rev. 37, 405 (193 1).<br />

4. M. SERRES, La naissance de Ia physique dans le texte de Lucrece,<br />

op. cit.<br />

5. For more details concerning chemical oscillations, see A.<br />

WINFREE, "Rotating Chemical Reactions," Scientific Amer·<br />

ican, Vol. 230 (1974), pp. 82-95.<br />

6. A. GoLDBETER and G. Nicous, ''An Allosteric Model with<br />

Positive Feedback Applied to Glycolytic Oscillations," Progress<br />

in Theoretical Biology, Vol. 4 (1976), pp. 65-160; A. GOLDBETER<br />

and S. R. CAPLAN, "Oscillatory Enzymes," Annual Review of<br />

Biophysics and Bioengineering, Vol. 5 (1976), pp. 449-73 ..<br />

7. B. HESS, A. BOITEUX, and J. KROG ER, "Cooperation of<br />

Glycolytic Enzymes," Advances in·Enzyme Regulation, Vol. 7<br />

(1969), pp. 149-67; see also B. HESS, A. GOLDBETER, and R.<br />

LEFEVER, "Temporal, Spatial and Functional Order in Regulated<br />

Biochemical Cellular Systems," Advances in Chemical<br />

Physics, Vol. XXXVIII (1978), pp. 363-413.<br />

8_ R HEss, Ciba Foundation Symposium. Vol. 31 (1975), p. 369.<br />

9A.G. GERESCH, "Cell Aggregation and Differentiation in Die-


ORDER OUT OF CHAOS 326<br />

tyostelium Discoideum," in Developmental Biology, Vol. 3<br />

(1968), pp. 157-197.<br />

98. A. GoLDBETER and L. A. SEGEL, "Unified Mechanism for Relay<br />

and Oscillation of Cyclic AMP in Dictyostelium Discoideum,"<br />

Proceedings of the National Academy of Sciences,<br />

Vol. 74 (1977), pp. 1543-47.<br />

10. See M. GARDNER, The Ambidextrous Universe (New Yo rk:<br />

Charles Scribner's Sons, 1979).<br />

II. O. K. KONDEPUDI and I. PRIGOGINE, Physica, Vol. l07A (1981),<br />

pp. 1-24; D. K. KoNDEPUDI , Physica, Vol. 115A (1982),<br />

pp. 552-66. It could even be that chemistry may bring to the<br />

macroscopic scale the violation of parity in weak fo rces; D. K.<br />

KoNDEPUDI and G. W. NELSON, Physical Review Letters, Vol.<br />

50, No. 14 (1983), pp. 1023-26.<br />

12. R. LEFEVER and W. HoRSTHEMKE, "Multiple Transitions Induced<br />

by Light Intensity Fluctuations in Illuminated Chemical<br />

Systems," Proceedings of the National Academy of Sciences,<br />

Vol. 76 (1979), pp. 2490-94. See also W. HoRSTHEMKE and M.<br />

MALEK MANSOUR, "Influence of External Noise on Nonequilibrium<br />

Phase Transitions," Zeitschrift fu r Physik B, Vo l. 24<br />

(1976), pp. 307-1 3; L. ARNOLD, W. HORSTHEMKE, and R.<br />

LEFEVER, "White and Coloured External Noise and Transition<br />

Phenomena in Nonlinear Systems," Zeitschrift fii r Physik B,<br />

Vol. 29 (1978), pp. 367-73; W. HORSTHEMKE, "Nonequilibrium<br />

Transitions Induced by External White and Coloured Noise,"<br />

Dynamics of Synergetic Systems, ed. H. HAKEN (Berlin:<br />

Springer Ve rlag, 1980); for an application to a biological problem,<br />

R. LEFEVER and W. HORSTHEMKE, "Bistability in Fluctuating<br />

Environments: Implication in Tu mor Immunology,"<br />

Bulletin of Mathematic Biology, Vol. 41 (1979).<br />

13. H. L. SwiNNEY and J. P. GoLLUB, "The Transition to Thrbulence,"<br />

Physics Today, Vol. 31 , No.8 (1978), pp. 41-49.<br />

14. M. J. FEIGENBAUM, "Universal Behavior in Nonlinear Systems,"<br />

Los Alamos Science, No I (Summer 1980), pp. 4-27.<br />

15. The concept of chreod is part of the qualitative description of<br />

embryological development Waddington proposed more than<br />

twenty years ago. It is truly a bifurcating evolution: a progressive<br />

exploration along which the embryo grows in an "epigenetic<br />

landscape" where coexist stable segments and segments<br />

where a choice among several developmental paths is possible.<br />

See C. H. WA DDINGTON, The Strategy of the Genes (London:<br />

Allen & Unwin, 1957). C. H. Waddington's chreods are also a<br />

central reference in Rene Thorn's biological thought. They could<br />

thus become a meeting point for two approaches: the one we are<br />

presenting, starting from local mechanisms and exploring the


327 NOTES<br />

spectrum of collective behaviors they can generate; and Thorn's,<br />

starting from global mathematical entities and connecting the<br />

qualitatively distinct forms and transformations they imply with<br />

the phenomenological description of morphogenesis.<br />

16. S. A. KAUFFMAN, R. M. SHYMKO, and K. TRABERT, "Control of<br />

Sequential Compartment Formation in Drosophila," Science,<br />

Vol. 199 (1978), pp. 259-69.<br />

17. H. BERGSON, Creative Evolution, pp. 94-95.<br />

18. See C. H. WADDINGTON, The Evolution of an Evolutionist<br />

(Edinburgh: Edinburgh University Press, 1975) and P. WEISS,<br />

"The Living System: Determinism Stratified," Beyond Reductionism,<br />

ed. A. KOESTLER and J. R. SMYTHIES (London:<br />

Hutchinson, 1969).<br />

19. D. E. KosHLAND, · Model Regulatory System: Bacterial<br />

Chemotaxis," Physiological Review, Vol. 59, No. 4, pp. 811-62.<br />

Chapter 6<br />

J. G. NICOLlS and I. PRIGOGINE, Self-Organization in Nonequilibrium<br />

Systems (New York: John Wiley & Sons, 1977).<br />

2. R BARAS, G. Nicous, and M. MALEK MANSOUR, "Stochastic<br />

Theory of Adiabatic Explosion," Journal of Statistical Physics,<br />

Vol. 32, No. 1 (1983), pp. I.<br />

3. See, for example, M. MALEK MANSOUR, C. VAN DEN BROECK,<br />

G. Nicous, and J. W. TURNER, Annals of Physics, Vol. 131, No.<br />

2 (1981), p. 283.<br />

4. J. L. DENEUBOURG, 'pplication de l'ordre par fluctuation a Ia<br />

description de certaines etapes de Ia construction du nid chez<br />

les termites," Insectes Sociaux, Journal International pour<br />

/'etude des Arthropodes sociaux, To me 24, No. 2 (1977),<br />

pp. 117-30. This first model is presently being extended in connection<br />

with new experimental studies; 0. H. BRUINSMA, ·n<br />

Analysis of Building Behaviour of the Termite macrotermes subhyaiinus,"<br />

Proceedings of the VIII Congress IUSSI (Waegeningen,<br />

1977).<br />

5. R. P. GARAY and R. LEFEVER, · Kinetic Approach to the Immunology<br />

of Cancer: Stationary States Properties of Effector­<br />

Target Cell Reactions," Journal of Theoretical Biology, Vol. 73<br />

(1978), pp. 417-38, and private communication.<br />

6. P. M. ALLEN, "Darwinian Evolution and a Predator-Prey Ecology,<br />

" Bulletin of Mathematical Biology, Vol. 37 (1975),<br />

PP- 389-405; "Evolution, Population and Stability, " Proceedings<br />

of the National Academy of Sciences, Vol. 73, No. 3 (1976),<br />

pp. 665-68. See also R. CzAPLEWSKI, ']\_ Methodology for Eval-


ORDER OUT OF CHAOS 328<br />

uation of Parent-Mutant Competition," Journal for Theoretical<br />

Biology, Vol. 40 (1973), pp. 429-39.<br />

7. See, for the present state of this work, M . . EIGEN and P. ScHus­<br />

TER, The Hypercycle (Berlin: Springer Verlag, 1979).<br />

8. R. MAY in Science, Vol. 186 (1974), pp. 645-47; see also R. MAY,<br />

"Simple Mathematical Models with very Complicated Dynamics,"<br />

Nature, Vol. 261 (1976), pp. 459-67.<br />

9. M. P. HASSELL, The Dynamics in Arthropod Predator-Prey Systems<br />

(Princeton, N.J.: Princeton University Press, 1978).<br />

10. B. HEINRICH, ·rtful Diners," Natural History, Vol. 89, No. 6<br />

(1980), pp. 42-5 1, esp. quote, p. 42.<br />

11. M. LovE, "The Alien Strategy," Natural History, Vol. 89, No. 5<br />

(1980), pp. 30-32.<br />

12. J. L. DENEUBOURG and P. M. ALLEN, "Modeles theoriques de<br />

Ia division du travail des les societes d'insectes," Academie<br />

Royale de Belgique, Bulletin de Ia Classe des Sciences, Tome<br />

LXII (1976), pp. 416-29; P. M. ALLEN, "Evolution in an Ecosystem<br />

with Limited Resources," op. cit., pp. 408-15.<br />

13. E. W. MoNTROLL, "Social Dynamics and the Quantifying of Social<br />

Forces," Proceedings of the National Academy of Sciences,<br />

Vol. 75, No. 10 (1978), pp. 4633-37.<br />

14. P. M. ALLEN and M. SANGLIER, "Dynamic Model of Urban<br />

Growth," Journal for Social and Biological Structures, Vol. 1<br />

(1978), pp. 265-80, and "Urban Evolution, Self-Organization<br />

and Decision-making," Environment and Planning A, Vol. 13<br />

(1981), pp. 167-83.<br />

15. C. H. WADDINGTON, To ols for Thought, (St. Albans, Eng.: Pal·<br />

adin, 1976), p. 228.<br />

16. S. J. GouLD, Ontogeny and Phylogeny, op. cit. Belknap Press<br />

Harvard University Press, 1977.<br />

17. C. L:Evi-STRAUSS, "Methodes et enseignement," Anthropologie<br />

structurale (Paris: Pion), pp. 311-17.<br />

18. See, for instance, C. E. RusSET, The Concept of Equilibrium in<br />

American Social Thought (New Haven, Conn.: Yale University<br />

Press, 1966).<br />

19. S. J. GouLD, "The Belt of an Asteroid," Natural History, Vol.<br />

89, No. 1 (1980), pp. 26-33.<br />

Chapter 7<br />

1. A.N. WHITEHEAD, Science and the Modern World, p. 186.<br />

2. The Philosophy of Rudolf Carnap, ed. P.A. ScHILP.P (Cambridge:<br />

Cambridge University Press, 1963).


329 NOTES<br />

3. J. FRASER, "The Principle of Temporal Levels: A Framework for<br />

the Dialogue?" communication at the conference Scientific<br />

Concepts of Time in Humanistic and Social Perspectives (Bellagio,<br />

July 1981).<br />

4. See on this point S. BRUSH, Statistical Physics and Irreversible<br />

Processes, esp. pp. 616-25.<br />

5. Feuer has rather convincingly shown how the cultural context of<br />

Bohr's youth could have helped his decision to look for a nonmechanistic<br />

model of the atom; Einstein and the Generation of<br />

Science (New York: Basic Books, 1974). See also W. HEISEN­<br />

BERG, Physics and Beyond (New York: Harper & Row, 1971)<br />

and D. SERWER, "Unmechanischer Zwang: Pauli, Heisenberg<br />

and the Rejection of the Mechanical Atom 1923-1925," Historical<br />

Studies in the Physical Sciences, Vol. 8 (1977), pp. 189-256.<br />

6. In Black-Body Theory and the Quantum Discontinuity, 1894-<br />

1912 (Oxford: Clarendon Press and New York: Oxford University<br />

Press, 1978), Thomas Kuhn has beautifully shown how<br />

closely Planck followed Boltzmann's statistical treatment of irreversibility<br />

in his own work.<br />

7. J. MEHRA and H. RECHENBERG, The Historical Development of<br />

Quantum Theory, Vols. 1-4 (New York: Springer Verlag, 1982).<br />

8. See, about the conceptual framework of the experimental tests<br />

recently conceived for hidden variables in quantum mechanics,<br />

B. o'EsPAGNAT, Conceptual Foundations of Quantum Mechanics,<br />

2nd aug. ed. (Reading, Mass.: Benjamin, 1976). See also B.<br />

o'EsPAGNAT, "The Quantum Theory and Reality," Scientific<br />

American, Vol. 241 (1979), pp. 128-40.<br />

9. See, for the complementarity principle, B. o'EsPAGNAT, op. cit.;<br />

M. JAMMER, The Philosophy of Quantum Mechanics (New<br />

York: John Wiley & Sons, 1974); and A. PETERSEN, Quantum<br />

Mechanics and the Philosophical Tradition (Cambridge, Mass.:<br />

MIT Press, 1968). C. GEORGE and I. PRIGOGINE, "Coherence<br />

and Randomness in Quantum Theory, " Physica, Vol. 99A<br />

(1979), pp. 369-82.<br />

10. L. RosENFELD, "The Measuring Process in Quantum Mechanics,"<br />

Supplement of the Progress of Theoretical Physics .(1965),<br />

p. 222.<br />

11. About the quantum mechanics paradoxes, which can truly be<br />

said to be the nightmares of the classical mind, since they all<br />

(SchrOdinger's cat, Wigner's friend, multiple universes) call to<br />

life again the phoenix idea of a closed objective theory (this time<br />

in the guise of SchrOdinger's equation), see the books by d'Espagnat<br />

and Jammer.<br />

12. B. MISRA, I. PRIGOGINE, and M. COURBAGE, "Lyapounov Vari-


ORDER OUT OF CHAOS<br />

330<br />

able; Entropy and Measurement in Quantum Mechanics," Pro<br />

ceedings of the National Academy of Sciences, Vol. 76 (1979),<br />

pp. 4768-4772. I. PRIGOGINE and C. GEORGE, "The Second<br />

Law as a Selection Principle: The Microscopic Theory of Dissipative<br />

Processes in Quantum Systems," to appear in Proceedings<br />

of the National Academy of Sciences. Vol 80 (1983)<br />

4590-94.<br />

13. H. MINKOWSKI, "Space and Time," The Principles of Relativity<br />

(New York: Dover Publications, 1923).<br />

14. A. D. SAKHAROV, Pisma Zh. Eksp. Teor. Fiz., Vol. 5, No. 23<br />

(1%7).<br />

Chapter 8<br />

I. G. N. LEWIS, "The Symmetry of Time in Physics," Science,<br />

Vol. 71 (1930), p. 570.<br />

2. A. S. EDDINGTON, The Nature of the Physical World (New<br />

York: Macmillan, 1948), p. 74.<br />

3. M. GARDNER, The Ambidextrous Universe: Mirror Asymmetry<br />

and Time-Reversed Worlds (New York: Charles Scribner's Sons,<br />

1979), p. 243.<br />

4. M. PLANCK, Treatise on Thermodynamics (New York: Dover<br />

Publications, 1945), p. 106.<br />

5. Quote by K. DENBIGH, "How Subjective Is Entropy?" Chemistry<br />

in Britain, Vol. 17 (198 1), pp. 168-85.<br />

6. See, for instance, M. KAC, Probability and Related Topics in<br />

Physical Sciences (London: Interscience Publications, 1959).<br />

7. J. W. GIBBS, Elementary Principles in Statistical Mechanics<br />

(New York: Dover Publications, 1960), Chap. XII.<br />

8. For instance, S. Watanabe introduces a strong distinction between<br />

the world to be contemplated and the world upon which<br />

we, as active agents, work; he states there is no consistent way<br />

of speaking about entropy increase if it is not in connection with<br />

our actions on the world. However, all our physics is in fact<br />

about the world to be acted on, and Watanabe's distinction thus<br />

does not help to clarify the relation between "microscopic deterministic<br />

symmetry" and "macroscopic probabilistic asymmetry."<br />

The question is left without an answer. How can we<br />

meaningfully say that the sun is irreversibly burning? See S.<br />

WATANABE, "Time and the Probabilistic View of the World,"<br />

The Voices of Time, ed. J. FRASER (New York: Braziller, 1966).<br />

9. Maxwell's demon appears in J. C. MAXWELL. TheorY of Heat<br />

(London: Longmans, 1871), Chap. XXII; see also E. DAUB,


331<br />

NOTES<br />

"Maxwell's Demon" and P. HEIMANN, "Molecular Forces, Statistical<br />

Representation and Maxwell's Demon," both in Studies<br />

in History and Philosophy of Science, Vol. I (1970); this volume<br />

is entirely devoted to Maxwell.<br />

10. L. BoLTZMANN, Populiire Schriften, new ed. (Braunschweig­<br />

Wiesbaden: Vieweg, 1979). As Elkana emphasizes in "Boltzmann's<br />

Scientific Research Program and Its Alternatives," Interaction<br />

Between Science and Philosophy (Atlantic,<br />

Highlands, N.J.: Humanities Press, 1974), the Darwinian idea of<br />

evolution is explicitly expressed mostly in Boltzmann's view<br />

about scientific knowledge-that is, in his defense of mechanistic<br />

models against energeticists. See, for instance, his 1886 lecture<br />

"The Second Law of Thermodynamics," Theoretical<br />

Physics and Philosophical Problems, ed. B. McGuiNNESS (Dordrecht,<br />

Netherlands: D. Reidel , 1974).<br />

11. For a recent account see I. PRIGOGINE, From Being to Becoming-Time<br />

and Complexity in the Physical Sciences (San Francisco:<br />

W. H. Freeman & Company, 1980).<br />

12. In his Scientific Autobiography, Planck describes his changing<br />

relationship with Boltzmann (who was first hostile to the phenomenological<br />

distinction introduced by Planck between reversible<br />

and irreversible processes). See also on this point Y.<br />

ELKANA, op. cit., and S. BRUSH, Statistical Physics and Irreversible<br />

Processes, pp. 640-5 1; for Einstein, op. cit., pp. 672-74;<br />

for Schrodinger, E. SCHR6DINGER, Science, Theory and Man<br />

(New York: Dover Publications, 1957).<br />

13. H. POINCARE, "La mecanique et I' experience," Revue de Metaphysique<br />

et de Morale, Vol. 1 (1893), pp. 534-37. H. POINCARE,<br />

Lefons de Thermodynamique, ed. J. Blondin (1892; Paris: Hermann<br />

1923).<br />

14. See for a study of the controversies around Boltzmann's entropy,<br />

see on this point S. BRUSH, The Kind of Motion We Call Heat,<br />

op. cit., and Planck's remarks in his biography (Loschmidt was<br />

Planck's student).<br />

15. I. PRIGOGINE, C. GEORGE, R HENIN, and L. ROSENFELD, ·<br />

Unified Formulation of Dynamics and Thermodynamics,"<br />

Chemica Scripta, Vol. 4 (1973), pp. 5-32.<br />

16. D. PARK, The Image of Eternity: Roots of Time in the Physical<br />

World (Amherst, Mass.: University of Massachusetts Press,<br />

1980).<br />

17. See also on this point S. BRUSH, The Kind of Motion We Call<br />

Heat-Book I, Physics and the Atomists; Book II, Statistical<br />

Physics and Irre versible Processes (Amsterdam: North Holland<br />

Publishing Company, 1976), as well as his commented anthology,


ORDER OUT OF CHAOS<br />

332<br />

Kinetic Theory: Vol. I, The Nature of Gases and Heat: Vol. II.<br />

Irreversible Processes (Oxford: Pergamon Press, 1965 and 1966).<br />

18. J. W. GIBBS, Elementary Principles in Statistical Mechanics<br />

(New Yo rk: Dover Publications, 1%0), Chap XII. For an historical<br />

account, see J. MEHRA, "Einstein and the Foundation of<br />

Statistical Mechanics, Physica, Vol. 79A, No. 5 (1974), p. 17.<br />

19. Many Marxist nature philosophers seem to take inspiration from<br />

Engels (quoted by Lenin in his Philosophic Notebooks) when he<br />

wrote in Anti-Diihring (Moscow: Foreign Languages Publishing<br />

House, 1954), p. 167, "Motion is a contradiction: even simple<br />

mechanical change of a position can only come about through a<br />

body being at one and the same moment of time both in one<br />

place and in another place, being in one and the same place and<br />

also not in it. And the continuous and simultaneous solution of<br />

this contradiction is precisely what motion is."<br />

20. L. BoLTZMANN, Lectures on Gas Theory (Berkeley: University<br />

of California Press, 1964), p. 446f, quoted in K. POPPER, Unended<br />

Quest (La Salle, Ill.: Open Court Publishing Company,<br />

1976), p. 160.<br />

21. POPPER, op. cit., p. 160.<br />

Chapter 9<br />

1. VoLTAIRE, Dictionnaire Philosophique. (Paris: Garnier, 1954.)<br />

2. See note 2, Chapter VII.<br />

3. K. PoPPER, "The Arrow of Time," Nature, Vol. 177 (1956),<br />

p. 538.<br />

4. See M. GARDNER, The Ambidextrous Universe, pp. 271-72.<br />

5. A. EINSTEIN and W RITZ, Phys. Zsch., Vol. lO (1909), p. 323.<br />

6. H. POINCARE, Les methodes nouvelles de Ia mecanique celeste<br />

(New York: Dover Publications, 1957); E. T. WHITTAKER, A<br />

Treatise on the Analytical Dynamics of Pa rticles and Rigid<br />

Bodies (Cambridge: Cambridge University Press, 1965).<br />

7. J. MosER, Stable and Random Motions in Dynamical Systems<br />

(Princeton, N.J. : Princeton University Press, 1974).<br />

8. For a general review, see J. LEBOWITZ and 0. PENROSE, "Modern<br />

Ergodic Theory," Physics To day (Feb. 1973), pp. 23-29.<br />

9. For a more detailed study, see R. BALESCU, Equilibrium and<br />

Non-Equilibrium Statistical Mechanics (New York: John Wiley<br />

& Sons, 1975).<br />

10. V. ARNOLD and A. Av Ez, Ergodic Problems of Classical Mechanics<br />

(New York: Benjamin, 1968).


333 NOTES<br />

11. H. PoiNCARE, "Le Hasard," Science et Methode (Paris: Flammarion,<br />

1914), p. 65.<br />

12. B. MISRA, I. PRIGOGINE and M. CouRBAGE, "From Deterministic<br />

Dynamics to Probabilistic Descriptions," Physica, Vol. 98A<br />

(1979), pp. 1-26.<br />

13. D. N. PA RKS and N. J. THRIFf, Times, Spaces and Places: A<br />

Chronogeographic Perspective (New York: John Wiley & Sons,<br />

1980).<br />

14. M. COURBAGE and I. PRIGOGINE, "Intrinsic Randomness and<br />

Intrinsic Irreversibility in Classical Dynamical Systems," Proceedings<br />

of the National Academy of Sciences, 80 (April 1983).<br />

15. I. PRIGOGINE and C. GEORGE, "The Second Law as a Selection<br />

Principle: The Microscopic Theory of Dissipative Processes in<br />

Quantum Systems," Proceedings of the National Academy of<br />

Sciences, Vol. 80 (1983), pp. 4590-4594.<br />

16. V. NABOKOV, Look at the Harlequins! (McGraw-Hill 1974).<br />

17. J. NEEDHAM, "Science and Society in East and West," The<br />

Grand Titration (London: Allen & Unwin, 1%9).<br />

18. See for more details B. MISRA, I. PRIGOGINE and M. CouR­<br />

BAGE, "From deterministic Dynamics to probabilistic Description",<br />

Physica 98A (1979) 1-26.; B. MISRA and I. PRIGOGINE<br />

"Time, Probability and Dynamics", in Long-time Prediction in<br />

Dynamics, eds. C. W. Horton, L. E. Recihl and A. G.<br />

Szebehely, (New York, Wiley 1983).<br />

19. I. PRIGOGINE, C. GEORGE, R HENIN, and L. ROSENFELD, ·<br />

Unified Formulation of Dynamics and Thermodynamics,"<br />

Chemica Scripta, Vol. 4 (1973), pp. 5-32.<br />

20. M. CouRBAGE "Intrinsic irreversibility of Kolmogorov dynamical<br />

systems," Physica 1983; B. Misra and I. Prigogine, Letters<br />

in Mathematical Physics, September 1983.<br />

Conclusion<br />

1. A. S. EDDINGTON, The Nature of the Physical World (N_ew<br />

York: Macmillan, 1948).<br />

2. L. LEVY-BRUHL, La Mentalite Primitif (Paris: PUF, 1922).<br />

3. G. MILLS, Hamlet's Castle (Austin: University of Texas Press,<br />

1976).<br />

4. R. TAGORE, "The Nature of Reality" (Calcutta: Modern Review<br />

XLIX, 1931), pp. 42-43.<br />

5. D. S. KOTHARI, Some Thoughts on Truth (New Delhi: Anniversary<br />

Address, Indian National Science Academy, Bahadur Shah<br />

Zafar Marg, 1975), p. 5.


ORDER OUT OF CHAOS<br />

334<br />

6. E. MEYERSON, Identity and Reality (New York: Dover Publications,<br />

1962).<br />

7. Described in H. BERGSON, Melanges (Paris: PUF, 1972),<br />

pp. 1340-46.<br />

8. Correspondence, Albert Einstein-Michele Besso, 1903-1955<br />

(Paris: Herman, 1972).<br />

9. N. WIENER, Cybernetics (Cambridge, Mass.: M.I.T. Press and<br />

New Yo rk: John Wiley & Sons, 1961).<br />

10. M. MERLEAU-PONTY, "Le philosophe et Ia sociologie," Eloge<br />

de Ia Philosophie, Collection Idees (Paris: Gallimard, 1960),<br />

pp. 136-37.<br />

11. M. MERLEAu-PoNTY, Resumes de Cours /952-/960 (Paris: Gallimard,<br />

1968), p. 119.<br />

12. P. VA LRY, Cahiers, La Pleiade (Paris: Gallimard, 1973), p. 1303.<br />

13. For what follows see also I. PRIGOGINE, I. STENGERS, and S.<br />

PA HAUT, "La dynamique de Leibniz a Lucrece," Critique "Special<br />

Serres," Vol. 35 (Jan. 1979), pp. 34-55. Engl. trans.: "Dynamics<br />

from Leibniz to Lucretius," Afterword to M. SERRES,<br />

Hermes: Literature, Science, Philosophy (Baltimore: Johns<br />

Hopkins Univ. Pr. , 1982), pp. 137-55.<br />

14. C. S. PEIRCE, The Monist Vo l. 2 (1892), pp. 321-337.<br />

15. A. N. WHITEHEAD, Process and Reality, pp. 240-41. On this<br />

subject, see I. LECLERC, Whitehead's Metaphysics (Bloomington:<br />

Indiana University Press, 1975).<br />

16. La naissance de Ia physique dans le texte de Lucrece, p. 139.<br />

17. LuCRETIUS, De Natura Rerum, Book II. ·gain, if all movement<br />

is always interconnected, the new arising from the old in a<br />

determinate order-if the atoms never swerve so as to originate<br />

some new movement that will snap the bonds of fate, the everlasting<br />

sequence of cause and effect-what is the source of the<br />

free will possessed by living things throughout the earth?"<br />

18. M. SERRES, op. cit., p. 136.<br />

19. M. SERRES, op. cit., p. 162; also pp. 85-86 and "Roumain et<br />

Faulkner traduisent l' Ecriture," La traduction (Paris: Minuit,<br />

1974).<br />

20. S. MoscoviCI, Hommes domestiques et hommes sauvages,<br />

pp. 297-98.<br />

21. T. KuHN, The Structure of Scientific Revolutions, 2nd ed. incr.<br />

(Chicago: Chicago University Press, 1970).<br />

22. See A. N. WHITEHEAD, Process and Reality, op. cit. and M.<br />

HEIDEGGER Sein und Zeit (Tiibingen: Niemeyer 1977).<br />

23. H. WEYL, Philosophy of Mathematics and Natural Science<br />

(Princeton, N.J.: Princeton University Press, 1949).<br />

24. A. NEHER, "Vision du temps et de l'histoire dans Ia culture<br />

juive," Les cultures et le temps (Paris: Payot, 1975), p. 179.


INDEX<br />

Note: Page numbers given in boldface indicate location of definitions<br />

or discussions of terms or concepts indexed here.<br />

Acceleration, 57-59<br />

Affinity, 29, 136<br />

Against the Current (Berlin), 2<br />

Agassiz, Louis, 195<br />

Alchemy, 64; affinity in, 136;<br />

Chinese, 278<br />

Alembert, Jean Le Rond d', 52;<br />

Diderot and, 80-82;<br />

opposition to Newtonian<br />

science of, 62, 63, 65, 66<br />

Ambidextrous Universe, The<br />

(Gardner), 234<br />

Amoebas, 156-60<br />

Ampere, Andre Marie, 67, 76<br />

Anaxagoras, 264<br />

Antireductionists, 173-74<br />

Archimedes, 39, 41, 304<br />

Aristotle, 39, 40, 71, 85, 173,<br />

305, 306; on change, 62;<br />

notion of space of, 171;<br />

physics of, 39-4 1; and<br />

theology, 49-50<br />

Arrow of time, xx, xxvii, 8, 16;<br />

Boltzmann on, 253-55; and<br />

elementary particles, 288;<br />

and entropy, 119, 257-59; and<br />

heat engines, 111-15; Layzer<br />

on, xxv; meaning of, 289;<br />

and probability, 238-39; roles<br />

played by, 30 I<br />

Atomists, 3, 36; conception of<br />

change of, 62, 63; on<br />

turbulence, 141<br />

Attractor, 121, 133, 140, 152<br />

Bach, J. S., 307<br />

Bachelard, Gaston, 320n<br />

Bacterial chemotaxis, I 75<br />

Baker transformation, 269,<br />

272-76, 278-79, 283, 289<br />

Being and Becoming, 310<br />

Belousov-Zhabotinsky reaction,<br />

151-53, 168<br />

Benard instability, 142-44;<br />

transition to chaos in, 167-68<br />

Bergson, Henri, 10, 79, 80,<br />

90-94, 96, 129, 173-74,<br />

301-2; on dynamics, 60; on<br />

time, 214, 294<br />

Berlin, Isaiah, 2, 11, 13, 80<br />

Bernoulli, Daniel, 82<br />

Berry, B., 17<br />

Berthollet, Claude Louis,<br />

Comte, 319n<br />

Besso, Michele, 294<br />

Bifurcations, xv, 160-61, 176,<br />

275; cascading, 167-70; in<br />

evolution, 171-72;<br />

fluctuations and, 177, 180; in<br />

reaction-diffusion systems,<br />

260; role of chance in, xxvi,<br />

170, 176; social, 313; and<br />

335


ORDER OUT OF CHAOS 336<br />

Bifurcations ( cont' d)<br />

statistical model, 205-6;<br />

theory of, 14<br />

Big Bang, xxvii, 288; and arrow<br />

of time, xxv, 259<br />

Biology, 2, 10; catalysts in,<br />

133-34; chemical reactions<br />

in, 131-32; "communication"<br />

among molecules in, 13;<br />

concepts from physics<br />

applied to, 207; and<br />

conversion, 108; evolution<br />

and, 12, 128; logistic<br />

equation in, 193-96;<br />

molecular, see Molecular<br />

biology; reductionistantireductionist<br />

conflict in,<br />

174; technological analogies<br />

in, 174-75; time in, 116;<br />

Whitehead on, 96<br />

Birchoff, 266<br />

Blake, William, 30<br />

Boerhave, Hermann, 105<br />

Bohr, Niels, 2, 74, 220, 224-25,<br />

228, 229, 292-93<br />

Boltzmann, Ludwig, xvii, 15,<br />

16, 122-27, 219, 227, 234-36,<br />

258, 259, 274, 286-87, 297,<br />

329n, 33ln; and arrow of<br />

time, 253-55; on ergodic<br />

systems, 266; on evolution<br />

toward equilibrium, 240-43;<br />

objections to theories of,<br />

243-46; and theory of<br />

ensembles, 248, 250<br />

Boltzmann's constant, 124<br />

Boltzmann's order principle,<br />

122-28, 142, 143, 150, 163,<br />

187<br />

Bordeu, 321n<br />

Born, Max, 220, 235<br />

Boundary conditions, 106,<br />

120-2 1, 125, 126, 138-39,<br />

142, 147, 151<br />

Braude!, xviii, xix<br />

Bridgmann, P. W., 316n<br />

Brillouin, 216<br />

Broglie, Louis de, 220<br />

Bruno, Giordano, 15<br />

Bruns, 72, 265<br />

Brusselator, 146, 148, 151, 152,<br />

160<br />

"Brussels school," xv<br />

Buchdahl, Gerd, 316n<br />

Buffon, Ge<strong>org</strong>es Louis Leclerc<br />

de, Comte, 65-67, 319n<br />

Butts, Thomas, 30<br />

Caillois, Roger, 128<br />

C<strong>alvin</strong>, John, xxii<br />

Cancer tumors, onset of, 188<br />

Canonical equation, 226<br />

Canonical variables, 70, 71,<br />

107, 222<br />

Cardwell, D., 322n<br />

Carnap, Rudolf, 214, 294<br />

Carnot, Lazare, 112<br />

Carnot, Sadi, 111-15, 117, 120,<br />

128, 140, 323n, 324n ; Carnot<br />

cycle, 112- 114, 117<br />

"Carrying capacity" of<br />

systems, 192-97<br />

Catalysis, 133-35, 145, 153<br />

Caterpillars, strategies for<br />

repelling predators of, 194-95<br />

Catherine the Great, 52<br />

Cells: Benard, 143; chemical<br />

reactions within, 131-32<br />

Chance, concepts of, xxii-xxiii,<br />

14, 170, 176, 203 ; see<br />

Randomness<br />

Change: motion and, 62-68;<br />

nature of, 29 1; of state, 106;<br />

in thermodynamic system,<br />

120-2 1; Whitehead on, 95<br />

Chemical clock, xvi, 13,<br />

147-48, 179, 307;<br />

communication in, 180; in<br />

glycolysis, 155; in slime mold<br />

aggregation, 159<br />

Chemical reactions, 127; in<br />

biology, 131-32; diffusion in,<br />

148-49; fluctuations and<br />

correlations in, 179-8 1;<br />

kinetic description of,<br />

132-34; self-<strong>org</strong>anization in,


337<br />

144-45; thermodynamic<br />

descriptions of, 133-37; see<br />

also specific reactions<br />

Chemistry, xi, I 0; Bergson on,<br />

91; and Buffo n, 65, 66;<br />

conceptual distinction<br />

between physics and, 137;<br />

and conversion, 108; Diderot<br />

on, 82, 83; fluctuations and,<br />

177-79; in<strong>org</strong>anic, 152, 153;<br />

irreversibility in, 209;<br />

Newtonian method in, 28;<br />

relation between order and<br />

chaos in, 168; and "science<br />

of fire," 103; temporal<br />

evolution in, 10-1 1<br />

China, 57; alchemy in, 278;<br />

social role of scientists in,<br />

45-46, 48<br />

Chiral symmetry, 285<br />

"Chreod," 172<br />

Chris taller model, 197, 203<br />

Christianity, 46, 47, 50, 76<br />

Chronogeography, 272<br />

Chuang Tsu, 22<br />

Clairaut, Alexis Claude, 62, 65<br />

Clausius, Rudolf Julius<br />

Emanuel, 114, 115, 233, 234,<br />

240, 304; entropy described<br />

by, 117- 19<br />

Clinamen, 141, 303 , 304<br />

Clocks: invention of, 46; as<br />

symbol of nature, Ill; see<br />

also Chemical clocks<br />

Closed systems, xv, 125<br />

Collective phenomena, xxiv; in<br />

amoebas, 156-160; in insects,<br />

181-86; in human geography,<br />

197-203 ; in social<br />

anthropology, 205, 317n<br />

Collisions, 63, 69, 132, 240-42,<br />

270-7 1 ' 280-85<br />

Combinatorial analysis, 123<br />

Communication: description as,<br />

300; in dissipative structures,<br />

13, 148; and entropy barrier,<br />

295-96; and fluctuations,<br />

187-88; and irreversibility,<br />

INDEX<br />

295; molecular basis to, xxv,<br />

180; stabilizing effects of, 189<br />

Compensation, 107; Clausiuson<br />

Carnot cycle, 114; statistical,<br />

124, 133, 240<br />

Complementarity, principle of,<br />

225<br />

Complexions, 123, 124, 127,<br />

150; in Benard instability,<br />

142-43<br />

Complexity: dynamics and<br />

science of, 208-9; limits of,<br />

188-89; modelizations of,<br />

203-7<br />

Comte, Auguste, 104-5<br />

Condillac, Etienne Bonnot de,<br />

66<br />

Condorcet, Marie Jean Antoine<br />

Nicolas Caritat, marquis de,<br />

66<br />

Conservation, life defined in<br />

terms of, 84<br />

Conservation of energy, I 07- 1 1 ;<br />

and Carnot cycle, 114, 115;<br />

and entropy, 117, 118;<br />

principle of, 69-7 1<br />

Convection, 127; in Benard<br />

instability, 142<br />

Copernicus, 307<br />

Correlations: dynamics of,<br />

280-85 ; fluctuations and,<br />

179-8 1<br />

Cosmology, xxviii, 1, 10;<br />

entropy and, 117; mysticism<br />

and, 34; and<br />

thermodynamics, 115-17;<br />

time and, 215, 259;<br />

Whitehead's, 94<br />

Counterintuitive responses, 203<br />

Critical threshold see Instability<br />

threshold<br />

Critique of Pure Reason (Karit),<br />

86<br />

Cybernetics (Weiner), 295-96<br />

Darwin, Charles, xiv, xx, 128,<br />

140, 215, 240, 24 1, 25 1


ORDER OUT OF CHAOS 338<br />

Darwinian selection, 190, 191,<br />

194, 195<br />

David, Ben, 318n<br />

Democritus, 3<br />

Deneubourg, J. L., 181<br />

Density function p, 247-50,<br />

261, 264; with arrow of time,<br />

277, 289; or distribution<br />

function, 289; in phase space,<br />

265-72, 274, 279<br />

Deoxyribonucleic acid (DNA),<br />

154; dissymetry of, 163<br />

Desaguliers, J. T. , 27<br />

Descartes, Re ne , 62, 63 , 81<br />

Destiny, 6, 257<br />

Determinism, xxv, 9, 60, 75 ,<br />

169-70, 176, 177, 216, 226,<br />

23 1 '<br />

264, 269-72, 304;<br />

concepts of, xxii-xxiii<br />

Dialectics of Nature (Engels),<br />

253<br />

Dialogue Concerning the Two<br />

Chief World Systems<br />

(Galileo), 305<br />

Dictionnaire Philosophique<br />

(Voltaire), 257<br />

Dictyostelium discoideum, 156,<br />

157<br />

Diderot, Denis, 79-85, 91, 136,<br />

309, 321n<br />

Differential calculus, 57, 222<br />

Diffusion, 148-49, 177<br />

Dirac, Paul, 34, 220, 230<br />

Disorder, xxvii, 18, 124, 126,<br />

142, 238, 246, 250, 286-87,<br />

293<br />

Dissipation (or loss), 63, 112,<br />

I 15, 117, 120, 125, 129,<br />

302-03<br />

Dissipative structures, xii, xv,<br />

xxiii, 12-14, 142-43, 189,<br />

300; coherence of, 170;<br />

communication in, 148;<br />

cultural , xxvi<br />

Dissymmetry, 124; 163; in time,<br />

125; see also Symmetrybreaking<br />

Distribution function se<br />

Density fu nction p<br />

Dobbs, B. J. , 319n<br />

Dander, Theophile de, 136, 325<br />

Donne, John, 55<br />

Driesch, Hans, 171<br />

Drosophila, 172<br />

du Bois Reymond, 77, 97<br />

Diierrenmatt, E, 21<br />

Duhem, Pierre Maurice Marie,<br />

97<br />

Duration, xxviii-xix; Bergson's<br />

concept of, 92, 294<br />

Dynamics, II, 14-15, 58-62,<br />

107; baker transformation in,<br />

276-77; basic symmetry of,<br />

243 ; change in, 62-68;<br />

concept of order in, 287; of<br />

correlations, 280-85;<br />

incompatibility of<br />

thermodynamics and, 216,<br />

233-34, 252-53; language of,<br />

68-74; and Laplace's demon,<br />

75-77; objects of, 306;<br />

operators in, 222; probability<br />

generated in, 274;<br />

reconciliation of<br />

thermodynamics and, 122;<br />

reversibility in, 120; and<br />

science of complexity, 208-9;<br />

static view of, xxix;<br />

symmetry-breaking in,<br />

260-6 1; and theories of<br />

irreversibility, 25 1; theory of<br />

ensembles in, 247-5 1;<br />

twentieth-century renewal of,<br />

264-72<br />

Dyson, Freeman, 1 17<br />

Ecology, logistic equation in,<br />

192-93, 196, 204<br />

Eddington, Arthur Stanley, xx,<br />

8, 49, 119, 233, 291<br />

Edge of Objectivity, Th<br />

(Gillespie), 31<br />

Ehrenfest model, 235-38, 240,<br />

246


339<br />

Eigen, M., 190-91<br />

Eigenfunctions, 22 1-23; of<br />

Hamiltonian operator, 227; of<br />

operator time T, 289;<br />

superposition of, 227-28<br />

Eigenvalues, 22 1, 222; of<br />

Hamiltonian operator, 226,<br />

227<br />

Einstein, Albert, xiv, 76, 242,<br />

271, 301, 307, 310; on basic<br />

myth of science, 52-53;<br />

demonstration of<br />

impossibility by, 296;<br />

dialogue with Tagore, 293;<br />

ensemble theory of, 247-5 1,<br />

26 1; God of, 54; on<br />

gravitation, 34; on<br />

irreversibility, 15, 258, 259,<br />

294-95 ; Mach's influence on,<br />

53; and quantum mechanics,<br />

218-20, 224; on scientific<br />

asceticism, 20-2 1; on<br />

simultaneity, 218; special<br />

theory of relativity of, 17;<br />

thought experiments of, 43;<br />

on time, 214-15, 251 ;<br />

"unified field theory" of, 2;<br />

use of probabilities rejected<br />

by, 227<br />

Electrons, 287, 288; stationary<br />

states of, 74<br />

Elementary particle physics,<br />

xxviii, l, 2, 9, 10, 19, 34,<br />

230, 285-88; "bootstrap"<br />

philosophy in, 96; quantum<br />

mechanics and, 230-3 1;<br />

T-violation in, 259; wave<br />

behavior in, 179<br />

Eliade, Mircea, 39-40<br />

Elkana, Y. , 323n, 33 1 n<br />

Embryo: development of,<br />

81-82; formation of gradient<br />

system in morphogenesis of,<br />

150; internal purpose of,<br />

171-73<br />

Encyclopedie, 83<br />

Energeticists, 234<br />

INDEX<br />

Energy, 107; dissipation of,<br />

302-3; and elementary<br />

particles, 287; and entropy,<br />

118- 19; exhaustible, Ill, 114;<br />

as invariant, 265; for living<br />

cells, 155; in quantum<br />

mechanics, 220-2 1; of<br />

unstable particles, 74; of<br />

universe, 117; see also<br />

Conservation of energy<br />

Energy conversion, 12, 108,<br />

114<br />

Engels, Friedrich, 252-53, 332n<br />

Engines, 12, 103, 105-07,<br />

111-15<br />

Enlightenment, the, 67, 79, 80,<br />

86<br />

Ensemble theory (Einstein­<br />

Gibbs), 247-5 1, 26 1;<br />

equilibrium and, 265<br />

Entelechy, 171<br />

Entropy, xix-xx, 12, 14-18,<br />

117-22, 227; and arrow of<br />

time, XXV, 253-54, 257-59;<br />

and atomism, 288; as barrier,<br />

277-80, 295-97; in evolution,<br />

131; flux and force and, 135,<br />

137; law of increase of, xxix;<br />

in linear thermodynamics,<br />

138-39; mechanistic<br />

interpretation of, 240-43 ;<br />

probability and, 124, 126,<br />

142, 234, 235, 237-38, 274;<br />

production of, 119, 131, 133,<br />

135, 137-39, 142; as<br />

progenitor of order, xxi-xxii;<br />

as selection principle,<br />

285-86; subjective<br />

interpretation of, 125, 235,<br />

25 1-52; thought experiment<br />

on, 244; universal<br />

interpretations of, 239<br />

Enzymes, 133-34; feedback<br />

action of, 154; in glycolysis,<br />

155; resembling Maxwell's<br />

demon, 175<br />

Epicurus, 3, 305


ORDER OUT OF CHAOS 340<br />

Equilibrium, xvi; and baker<br />

transformation, 273;<br />

chemical, 133; chemical<br />

reactions in, 179-80; and<br />

entropy, 120, 131; evolution<br />

toward, 24 1-43; flux and<br />

force at, 135-37; in future,<br />

275, 276; and matter-light<br />

interaction, 219; maximum<br />

probability at, 286; and<br />

theory of ensembles, 265;<br />

thermal, 105, 116; thermal<br />

chaos in, 168; in<br />

thermodynamics, 12, 13,<br />

125-29, 138; velocity<br />

distribution in state of, 241;<br />

see also Far-fromequilibrium;<br />

Nonequilibrium<br />

Ergodic systems, 266<br />

Espagnat, B. d', 329n<br />

Esprit de systeme, 83<br />

Euclid, 171<br />

Euler, Leonhard, 52, 65, 82<br />

Everett, 228<br />

Evolution, xx, 12, 128; and<br />

arrow of time, xxv;<br />

bifurcations in, 171-72;<br />

biological, 153; Boltzmann .<br />

on, 240; chemical, 177;<br />

concepts from physics<br />

applied to, 207-9; cosmic,<br />

215, 288; Darwinian, 128;<br />

from disorder to order, xxix;<br />

entropy in, 119, 131; toward<br />

equilibrium, 241-43;<br />

feedback in, 196-203 ;<br />

logistic, 192-96, 204;<br />

·paradigm of, 297-98; in<br />

quantum mechanics,<br />

226-228, 238; toward<br />

stationary state, 138-39;<br />

structural stability in, 189-91<br />

Existentialism, xxii<br />

Expanding universe, 2, 19, 215,<br />

259<br />

Experimentation, 5, 41-44; and<br />

global truth, 44-45 ; Kant<br />

and, 88; universality of<br />

language postulated by, 51;<br />

Whitehead on, 93, 95 ; see<br />

also Thought experiments<br />

Falling bodies, Galileo's laws<br />

for, 57, 64<br />

Far-from-equjlibrium, xxvi,<br />

xxvii, 13-14, 140-45, 300;<br />

chemical instability in,<br />

146-53; in chemistry, 177;<br />

dissipative structures in, 189;<br />

in molecular biology, 153-59;<br />

prebiotic evolution in, 191;<br />

self-<strong>org</strong>anization in, 176<br />

Faraday, Michael, 108<br />

Faust (Goethe), 128<br />

Feedback, 153; in biological<br />

systems, 154; in evolution,<br />

191, 196-203 ; between<br />

science and society, xiii<br />

Feigenbaum sequence, 169<br />

Feuer, 329n<br />

Feynman, Richard, 44<br />

Fluctuations, xv, xxiv-xxv,<br />

xxvii, 124-25, 140-41, 143;<br />

amplification of, 141, 143<br />

181-89; and chemistry,<br />

177-79; and correlations,<br />

179-8 1; in Markov process,<br />

238; on microscopic scale,<br />

23 1-32<br />

Fluid flow, 141<br />

Fluxes, 135-37; random noise<br />

in, 166-67; in reciprocity<br />

relations, 137-38<br />

Forces: generalized, 135-37; in<br />

reciprocity relations, 137-38<br />

Fourier, Baron Jean-Joseph, 12,<br />

104, 105, 107, 115-17<br />

Fraser, J. T. , 214<br />

Frederick II, King of Prussia,<br />

52<br />

Free particles, 70-72<br />

Free will, xxii<br />

Freud, Sigmund, 17<br />

Friedmann, Alexander, 215<br />

Fundamental level of<br />

description, 252-53


341<br />

Galileo, 40, 41, 50, 51, 305; on<br />

cause and effect, 60; and<br />

global truths, 44; and<br />

mechanistic world view, 57;<br />

thought experiments of, 43<br />

Galvani, Luigi, 107<br />

Gardner, Martin, 234, 259<br />

Gassendi, Pierre, 62<br />

Generalized forces, 135-37<br />

Geographical time, xviii<br />

Geography, 197; internal time<br />

in, 272<br />

Geology, 121; time in, 116, 208<br />

Gibbs, J. W., 15, 238, 247-5 1,<br />

261<br />

Gillispie, C. C., 31, 321n<br />

Glycolysis, 155-56<br />

Goethe, Johann Wolfgang von,<br />

128<br />

Gould, Stephen J. , 204<br />

Grasse, 181, 186<br />

Gravitation: Comte on, 105;<br />

and determination of motion,<br />

59; in early universe, 298;<br />

Einstein's interpretation of,<br />

34; explanatory power of, 28,<br />

29; in far-from-equilibrium<br />

conditions, 163-64; universal<br />

law of, I, 12, 66<br />

Guldberg and Waage's law (also<br />

Mass action, law oO, 133<br />

Hamilton, William Rowan, 68,<br />

%; see Hamiltonian<br />

Hamiltonian: equation, 249;<br />

function, 68, 70-7 1, 74, 107,<br />

220-2 1; operator, 221,<br />

226-27; and T-violation, 259<br />

Hankins, Thomas, 318n<br />

Hao Bai-lin, 151, 152<br />

Hausheer, Roger, 2<br />

Hawking, 117<br />

Heat, 12, 79, 103; conduction<br />

of, 104, 135; electricity<br />

produced by, 108; and heat<br />

engines, 12, 103, 106-7; heat<br />

engines, arrow of time and,<br />

111-15; propagation of,<br />

INDEX<br />

104-5; repelling force of, 66;<br />

specific, 106; transformation<br />

of matter by, 105<br />

Hegel, G. W. R, 79, 89-90, 92,<br />

93, 173<br />

Heidegger, Martin, 32-33, 42,<br />

79, 310<br />

Heisenberg, Werner, xxii, 220,<br />

292, 2%<br />

Heisenberg uncertainty<br />

relations, 178, 222-26<br />

Helmholtz, Hermann Ludwig<br />

Ferdinand von, %, 109-11<br />

Herivel, J. , 322n<br />

Hess, Benno, 155<br />

Hirsch, J. , 151<br />

History: of ideas, 79; open<br />

character of, 207; reinsertion<br />

of, into natural and social<br />

sciences, 208; of science, 307<br />

(cosmology) 208, 215,<br />

(geography) 197, 272,<br />

(geology) 116, 121, 208<br />

Holbach, Paul Henri Thiry,<br />

baron d ' , 82<br />

Hooykaas, R., 317n<br />

Hopf, 266<br />

Hubble, Edwin Powell , 215<br />

Humanities, schism between<br />

science and, 11, 13<br />

Hume, A. Ord, 317n<br />

Huss, John, xxii<br />

Huyghens, Christiaan, 60<br />

Hydrodynamics, 127; far-fromequilibrium<br />

phenomena in,<br />

141<br />

Hypnons, 180, 287, 288<br />

Hysteresis, 166<br />

Idealization, 41-43, 69,<br />

112-1 14, 115, 120, 216, 248,<br />

252, 305-06<br />

Impossibility, demonstrations<br />

of, 17, 216- 17, 296, 299-300<br />

Individual time, xviii<br />

Industrial Age , 111; combustion<br />

and, 103


ORDER OUT OF CHAOS<br />

342<br />

Information, 17-18, 250,<br />

278-79, 283, 295, 297-98<br />

Initial conditions (or state), 61 ,<br />

68, 75, 121, 124, 128, 129,<br />

139-40, 142, 147, 248,<br />

261 -67, 270-7 1, 276, 278-79,<br />

295, 310<br />

Innovation, psychological<br />

process of, xxiv<br />

Innovative becoming,<br />

philosophy of, 94<br />

Instability, dynamic, 73,<br />

268-72, 276, 300; chemical,<br />

144-53; thermodynamic,<br />

141-42; threshold, 146, 147,<br />

160<br />

Insects, self-aggregation of,<br />

181-86<br />

Integrable systems, 71-72, 74,<br />

264-65, 302<br />

Internal time, 272-73, 289<br />

Intrinsically irreversible<br />

systems, 275-77, 289<br />

Intrinsically random systems,<br />

274-76, 289<br />

Intuition, 80, 91, 92<br />

Irreversibility, xx, xxi, xxvii,<br />

xxviii, 7-9, 63, 115;<br />

acceptance by physics of,<br />

208-9; and biology, 128, 175;<br />

in chemistry, 131, 137, 177;<br />

controversy over, 15-16;<br />

cultural context of<br />

incorporation into physics of,<br />

309- 10; and dynamics of<br />

correlations, 280-85 ; Einstein<br />

on, 294-95; and ensemble<br />

theory, 250; in evolution, 128,<br />

189; formulation of theory of,<br />

105, 107, 117-2l; and limits<br />

of classical concepts, 261-64;<br />

and matter-light interaction,<br />

219; measurement and, 228;<br />

microscopic theory of, 242,<br />

257-59, 285-86, 288-90, 310;<br />

probability and, 16, 124, 125,<br />

233-40; quantitative<br />

expression of, Ill; from<br />

randomness to, 272-77; rate<br />

of, 135; in reciprocity<br />

relations, 138; as source of<br />

order, 15, 292; subjective<br />

interpretation of, 25 1-52; as<br />

symmetry-breaking process,<br />

260-6 1; in thermodynamics,<br />

12; see also Arrow of time;<br />

Entropy<br />

Isomerization reaction, 165<br />

Jammer, M., 329n<br />

Jordan, 220<br />

Joule, James Prescott, 108-9<br />

Kant, Immanuel, 79, 80, 85-89,<br />

93, 99, 214<br />

Kauffman, S. A., 172<br />

Kepler, Johannes, 49, 57, 67,<br />

307<br />

Keynes, Lord, 319n<br />

Kierkegaard, Soren, 79<br />

Kinetic energy, 69-7 1, 90, 107,<br />

261<br />

Kinetics, chemical , 132-34<br />

Kirchoff, Gustav, %<br />

Knight, 321 n<br />

Koestler, Arthur, 32, 34-35<br />

Kolmogoroff, 266<br />

Kothari, D. S., 293<br />

Koyre, Alexandre, 5, 32, 35-36,<br />

62, 317n, 319n<br />

Kuhn, Thomas, 307-9, 320n,<br />

329n<br />

Lagrange, Comte Joseph Louis,<br />

52, %, 104, 324n<br />

Laminar flow, 141-42, 303<br />

Laplace, Marquis Pierre Simon<br />

de, xiii, 28, 52, 54, 66, 67,<br />

115, 323n; Fourier criticized<br />

by, 104<br />

Laplace's demon, 75-77, 87,<br />

27 1<br />

Large numbers, law of, 14, 178,<br />

180<br />

Lavoisier, Antoine Laurent, 2&,<br />

109


343 INDEX<br />

Layzer, David, xxv<br />

Lebenswe/t, 299<br />

Leibniz, Gottfried von, 50, 54,<br />

302, 303; formulae for<br />

velocity and acceleration, 58;<br />

monads of, 74<br />

Lemaitre, Ge<strong>org</strong>es, 215<br />

Lenoble, R., 3<br />

Levi-Strauss, Claude, 205, 317n<br />

Levy-Bruhl, L., 292<br />

Lewis, G. N., 233<br />

Liebig, Baron Justus von, 109<br />

Life: Bergson on, 92;<br />

compatibility with far-fromequilibrium<br />

conditions, 143;<br />

as expression of self<strong>org</strong>anization,<br />

175-76; and<br />

order principle, 127-28;<br />

origin of, 14; Romantic<br />

concepts of, 85; Stahl's<br />

definition of, 84; symmetrybreaking<br />

as characteristic of,<br />

163; temporal dimensions of,<br />

208; see also Molecular<br />

biology<br />

Light: velocity of, 17, 55,<br />

217-19, 278, 295, 296; waveparticle<br />

duality of, 219-20<br />

Limit cycle, 146-47<br />

Linear thermodynamics,<br />

137-40<br />

Liouville equation, 249, 250,<br />

266<br />

Logistic evolution, 192-96,<br />

203-04<br />

Look at the Harlequins<br />

(Nabokov), 277<br />

Lorentz, Hendrik Antoon, 270<br />

Loschmidt, 244, 246<br />

Louis XIV, King of France, 52<br />

Love, Milton, 195<br />

Lucretius, 3, 141, 302-5, 315n,<br />

334n<br />

Luther, Martin, xxii<br />

Lyapounov, 151<br />

Mach. Ernst. 49. 53-54. 97.<br />

318n<br />

Machines: Archimedes's, 41;<br />

ideal, 63, 69-70; mathematics<br />

and, 46; using heat, 103<br />

Macroscopic system, 106-07<br />

Many-worlds hypothesis, 228<br />

Markov chains, 236, 238, 240,<br />

242, 273-76; and dynamics of<br />

correlations, 283 ; and entropy<br />

barrier, 278<br />

Marx, Karl, 252<br />

Mass action, law of, 133, 136,<br />

23 1<br />

Materialistic naturalism, 83<br />

Mathematization, 46; in<br />

Hamiltonian function, 71;<br />

Hegel's critique of, 90;<br />

Leibniz on, 50; of motion, 60<br />

Matter: active, 9, 286-90, 302;<br />

anti-matter, 230-3 1; Diderot<br />

on, 82; effect of heat on, 105;<br />

in far-from-equilibrium<br />

conditions, 14; interaction of<br />

radiation and, 219; new view<br />

of, 9; nonequilibrium<br />

generated by, 181; perception<br />

of differences by, 163, 165;<br />

properties of, 2; Stahl on,<br />

84-85; transition to life from,<br />

84; wave-particle duality of,<br />

221<br />

Maxwell, James Clerk, 54, 73,<br />

122, 160, 240, 24 1, 266<br />

Maxwell's demon, 175, 239<br />

Mayer, Julius Robert von, 109,<br />

Ill<br />

Measurement, irreversible<br />

character of, 228-29<br />

Mechanics, II, 15;<br />

generalization of, Ill; Hegel<br />

on, 90; and probability, 125;<br />

see also Dynamics; Quantum<br />

mechanics<br />

Medicine: Bergson on, 91;<br />

Diderot on, 82, 83<br />

Merleau-Ponty, M., 299<br />

Metternich, Clemens We nzel<br />

Nepomuk Lothar, Fiirst von,<br />

xiii


ORDER OUT OF CHAOS 344<br />

Meyerson, Emile, 293<br />

Microcanonical ensemble, 265<br />

Minimum entropy production,<br />

theorem of, 138-4 1<br />

Minkowski, H., 230<br />

Mole, 121n<br />

Molecular biology, 4, 8; farfrom-equilibrium<br />

conditions<br />

in, 153-59; vitalism and, 84<br />

Molecular chaos assumption,<br />

246<br />

Monads, 74, 302-03<br />

Monod, Jacques, 3-4, 22, 36,<br />

79, 84<br />

Morin, Edgar, xxii-xxiii<br />

Morphogenesis, 172, 189<br />

Moscovici, Serge, 22, 306-7<br />

Motion: and change, 62-68;<br />

complexity of, 75 ; instability<br />

of, 73; in mechanical engine<br />

vs. heat engine, 112;<br />

positivist notion of, 96;<br />

productton of, in heat engine,<br />

107; universal laws of, 57-62,<br />

83; see also Dynamics<br />

Nabokov, Vladimir, 277-78<br />

Napoleon, 52, 67<br />

Natural laws: belief in<br />

universality of, 1-2;<br />

mathematical concepts of, 46;<br />

Newton on, 28; primary and<br />

secondary, 8; social structure<br />

and views of, 48-49; timeindependent,<br />

2, 7; trials of<br />

animals for infringements of,<br />

48<br />

Nature of the Physical World,<br />

The (Eddington), 8<br />

Needham, Joseph, 6-7, 45, 48,<br />

49, 278, 322n<br />

"Neolithic Revolution," 5-6, 37<br />

Neumann, von, 266<br />

New Science, The (Vico), 4<br />

Newton, Isaac, xv, xxviii, 12,<br />

27-29, 76, 98, 104, 120, 124,<br />

234, 305, 319n; alchemy and,<br />

64; on change, 62, 63 ;<br />

eighteenth-century opposition<br />

to, 65 ; laws of motion, 70;<br />

and mechanistic world view,<br />

57; objectivity defined by,<br />

218; objects chosen for study<br />

by, 216; presentation of<br />

Principia to Royal Society, 1;<br />

second law, 58; see also<br />

Newtonian science<br />

Newtonian science, xiii, xiv,<br />

xix, xxv, xxvi, 37-40, 213;<br />

absence of universal constant<br />

in, 217; concept of change in,<br />

63-68; Diderot and, 80, 82;<br />

Koyre on, 35-36;<br />

incompleteness of, 209;<br />

instability and, 264;<br />

instability of cultural position<br />

of, 30; Kantian critique of,<br />

85-87; laws of motion of,<br />

57-59; limits of, 29-30;<br />

positivism and, 96; prophetic<br />

power of, 28; spread of,<br />

28-29; Voltaire and, 258;<br />

world view of, 229<br />

. Nietzsche, Friedrich, Ill, 136<br />

Nisbet, R., 79<br />

N onequilibrium: cosmological<br />

dimension of, 23 1; difference<br />

between particles and<br />

antiparticles in, 285 ;<br />

fluctuations in, 178-80;<br />

innovation and, xxiv; and<br />

origin of structures, xxix; as<br />

source of order, 287; see also<br />

Far-from-equilibrium<br />

Non-linearity, 14, 134, 153,<br />

154-55, 197 see Catalysis<br />

Non-linear thermodynamics,<br />

137, 140<br />

Noyes, 152<br />

Nucleation, 187, 188<br />

Oersted, Hans Christian, 108<br />

Old Te stament, xxii<br />

Onsager, Lars, 137, 138<br />

Operators, 22 1-22, 225;<br />

commuting, 223


345<br />

Opticks (Newton), 28<br />

Optimization, 197, 207<br />

Order, 12, 18, 126, 131, 143,<br />

171-75, 238, 246, 250-5 1,<br />

286-87<br />

Order through fluctuation, 159,<br />

178; models based on<br />

concept of, 206<br />

Organization theory, xxiv<br />

Oscillating chemical reactions,<br />

19, 147-49<br />

Oscillations: glycolytic, 155;<br />

time- and space-dependent,<br />

148<br />

Ostwald, Wilhelm, 324n<br />

Pascal, Blaise, 3, 36, 79<br />

Pasteur, Louis, 163<br />

Pattern selection, 163, 164<br />

Pearson, Karl, 49<br />

Peirce, Charles S., 17, 302-3<br />

Pendulum, 16, 73, 216, 261-62<br />

Phase changes, 187<br />

Phase space, 247-50, 261, 264;<br />

delocalization in, 289;<br />

unstable systems in, 266-72<br />

Photons, 230, 288<br />

Physicists, The (DOerrenmatt),<br />

21<br />

Physics: application of concepts<br />

to evolution, 207-9; Bergson<br />

on, 91, 92; changing<br />

perspective in, 8-9;<br />

complementary developments<br />

in biology and, 154; and<br />

concepts of change, 63;<br />

conceptual distinction<br />

between chemistry and, 137;<br />

Diderot on, 80-83;<br />

evolutionary paradigm in,<br />

297-98; inspired discourse of,<br />

76; introduction of<br />

probability in, 123; and laws<br />

of motion, 57; Lucretian,<br />

141 ; macroscopic, xii;<br />

objectivity in, 55 ; positivist<br />

view of, 97; of processes,<br />

INDEX<br />

243 ; and theology, 49; time<br />

in, 116; vitalism and, 84;<br />

Whitehead on, 95, 96<br />

Planck, Max, 121, 219, 242,<br />

324n, 329n, 33 ln; on second<br />

law of thermodynamics,<br />

234-35<br />

Planck's constant, 217, 219,<br />

220, 223, 224<br />

Planetary motion: Kepler's laws<br />

for, 57; in Newtonian<br />

dynamics, 59, 64<br />

Plato, 7, 39, 67<br />

Poincare, Jules Henn, 68, t2,<br />

97, 151, 236, 243, 253, 265<br />

27 1<br />

Poisson distribution, 179-8 1<br />

Pope, Alexander, 27, 67<br />

Popper, Karl, 5, 15, 254-55,<br />

258-59, 276, 317n<br />

Populiire Schriften<br />

(Boltzmann), 240<br />

Positive-feedback loops, xvit<br />

Positivism, 80, 96-98; Comte<br />

and, 104-5; German<br />

philosophy and, 109<br />

Potential: dynamic, 69-70;<br />

thermodynamic, 126, 138-40<br />

Potential energy, 69-70, 73, 107<br />

Prebiotic evolution, 190-91<br />

Pre-Socratics, 38-39<br />

Principia (Newton), l, 28<br />

Probability, 122-24; Einstein<br />

on, 259; at equilibrium, 286;<br />

in far-from-equilibrium<br />

conditions, 143; entropy and<br />

142, 274, 297; and<br />

fluctuations, 178, 179; and<br />

irreversibility, 233-40; in<br />

quantum mechanics, 227;<br />

subjective vs. objective<br />

interpretations of, 274; in<br />

unstable systems, 271-72<br />

Process: physics of, 12, 105,<br />

107, 243; Whitehead's<br />

concept of, 258, 303<br />

Process and Reality<br />

(Whitehead), 93, 96, 310


ORDER OUT OF CHAOS 346<br />

Proust, Marcel , 17<br />

Pulley systems, 41-42<br />

Quantization, 220<br />

Quantum mechanics, 9, II, 15,<br />

34, 218-22; causality in, 31 1;<br />

correlations in, 286; cultural<br />

background to, 6;<br />

delocalization in, 289;<br />

demonstrations of<br />

impossibility in, 217, 296-97;<br />

Hamiltonian function in, 70;<br />

Heisenberg's uncertainty<br />

relations in, 222-26; and<br />

Newtonian synthesis, 67-68;<br />

and probability, 125, 178-79;<br />

and reversibility, 61; temporal<br />

evolution in, 226-29, 238;<br />

thought experiments in, 43<br />

Quetelet, Adolphe, 123, 241<br />

Radiation: black-body, 209, 215;<br />

interaction of matter and, 219<br />

Randomness, xx, 8, 9, 126,<br />

23 1-32, 236; to irreversibility<br />

from, 272-77<br />

Rationality, 1, 29, 32, 36, 40,<br />

42, 92, 306<br />

Reaction-diffusion systems, 260<br />

Real ity, conceptualization of,<br />

225-26<br />

Reciprocity relations, 137-38<br />

Reduction of the wave function,<br />

227-28<br />

Reductionism, 173-74<br />

Reichenbach, H., 97<br />

Relation, philosophy of, 95<br />

Relativity, 9, 34, 215, 307; in<br />

astrophysics, 116; Bergson's<br />

misunderstanding of, 294;<br />

demonstrations of<br />

impossibility in, 217, 296;<br />

Einstein's special theory of,<br />

17; and elementary particles,<br />

230; and Newtonian<br />

synthesis, 67, 68, 229; static<br />

geometric character of time<br />

and , 230; and thermal history<br />

of universe, 23 1; thought<br />

experiments in, 43 ; and<br />

universal constants, 217 -18;<br />

and velocity of light, 295<br />

Religion: ancient Greek, 38, 39;<br />

resonance between science<br />

and, 46-5 1<br />

Residual black-body radiation,<br />

209, 215<br />

Respiration, physiology of, 109<br />

Reversibility: of canonical<br />

equations, 71; of<br />

thermodynamic<br />

transformation, 12, 112-13,<br />

120; of trajectories, 60-61<br />

Revolution, concept of, xxiv<br />

Reynolds' number, 144<br />

Ritz, W, 259<br />

Rosenfeld, Leon, 264, 329<br />

Sakharov, A. D., 230<br />

Sartre, Jean-Paul, xxii<br />

Schlanger, J. , 321n<br />

Schrodinger, Erwin, 18-19 220,<br />

242, 329n<br />

SchrOdinger equation, 226-29<br />

Science and Civilization in<br />

China (Needham), 278<br />

"Scientific revolution," 5, 6<br />

Scott, W, 322n<br />

Sein und Zeit (Heidegger), 310<br />

Self-<strong>org</strong>anization, xv; in Benard<br />

instability, 142; and<br />

bifurcations, 160-67; in<br />

chemical clock, 148; in<br />

chemical reactions, 144-45;<br />

and dynamics, 208; as<br />

function of fluctuating<br />

external conditions, 165-67;<br />

life as expression of, 175-76;<br />

in slime-mold aggregation,<br />

156; in turbulence, 141-42<br />

Serres, Michel, 104, \41, 303,<br />

304, 320n, 323n<br />

Shakespeare, William, 293<br />

Signals, propagation of, 217,<br />

218


3


ORDER OUT OF CHAOS 348<br />

Time (cont'd)<br />

277-78, 295-%; in everyday<br />

life, 16-17; global judgments<br />

of, 17; in Hamiltonian<br />

function, 70, 71; Hegel on,<br />

90; and human symbolic<br />

activity, 312; internal,<br />

272-73, 289; involved in<br />

turbulence, 141; meaning of,<br />

in physics, 93 ; as measure of<br />

change, 62; in nineteenthcentury<br />

physics, 117;<br />

oscillations dependent on,<br />

148; positivist view of, 97;<br />

progressive rediscovery of,<br />

208; in quantum mechanics.<br />

229-30; revision of<br />

conception of, 96; roots of, in<br />

nature, 18; social context of<br />

rediscovery of, 19; in<br />

thermodynamics, 12, 129;<br />

varying importance of scales<br />

of, 301; see also Arrow of<br />

time; Irreversibility<br />

Time Machine, The (Wells), 277<br />

Toftler, Alvin, xi-xxvi, xxxi<br />

Trajectories, 59-60, 68, 121,<br />

177; intrinsically<br />

indeterminate, 73 ; limits of<br />

concept of, 261-64; in phase<br />

space, 247-48; and<br />

probability, 122; in unstable<br />

systems, 270-72; variations<br />

in, 75<br />

Transition probabilities, 274<br />

Thrbulence, 141<br />

Turbulent chaos, 167-68<br />

Turing, Alan M., 152<br />

Unidirectional processes,<br />

258-59<br />

"Unified field theory," 2<br />

Universal constants, 217-19,<br />

229<br />

Universe: age of, 1; aging of,<br />

xix-xx; disintegration of,<br />

116; energy of, 117; entropy<br />

of, 118; expanding, 2, 19,<br />

215, 259; history of, 215; in<br />

Newtonian dynamics, 59;<br />

nonequilibrium, 22932;<br />

Pierce on, 302-3; pluralistic<br />

character of, 9; thermal<br />

history of, 9; time-oriented<br />

polarized nature of, 285<br />

Unstable particles, 74, 23 1, 288<br />

Urbanization, model of,<br />

198-202<br />

Urn model, 235-38, 246, 273<br />

Valery, Paul, 16, 301<br />

Velocity, 57-59; distribution of,<br />

240-42; instantaneous<br />

inversion of, 61, 243-46,<br />

280-85; of light, 17, 55,<br />

217-19, 278, 295, 2%<br />

Velocity distribution function,<br />

242-46, 248-50<br />

Velocity inversion experiment,<br />

280-84<br />

Venel, 83, 309<br />

Vico, G., 4<br />

Vienna school, 97<br />

Vitalism, 80, 83-84; vs.<br />

scientific methodology, llO<br />

Volta, Count Alessandro, 107-8<br />

Voltaire, 257-58<br />

Waddington, Conrad H., 172,<br />

174, 207, 322n, 326n<br />

Watanabe, S., 330n<br />

Watt, James, 103<br />

Wave behavior, 179; see also<br />

Chemical clocks<br />

Wave functions, 226-28; time<br />

and, 229-30<br />

Wave-particle duality, 219-20,<br />

226<br />

Wealth of Nations (Smith). 103<br />

Weiss, Paul, 174<br />

Wells, H. G., 277<br />

Wey l, Herman, 311


34Q<br />

Whitehead, Alfred North, 10,<br />

17, 47, 50, 79, 93-96, 212,<br />

216, 258, 302, 303, 310, 322n<br />

Wiener. Norbert, 295-96<br />

Wycliffe , John, xxii<br />

INDEX<br />

Zermelo, 15, 244, 253, 254<br />

"Zero growth" society, 116<br />

Zhang Shu-yu, 151<br />

Zola, Emile, 323n

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!