Sie sind auf Seite 1von 437

THERMODYNAMICS

An Advanced Textbook for


Chemical Engineers
THERMODYNAMICS
An Advanced Textbook for
Chemical Engineers

GIANNI AST ARIT A


University of Delaware
Newark. Delaware
and University of Naples
Naples. Italy

Springer Science+Business Media, LLC


Library of Congress Cataloging in Publication Data
Astarita, Giovanni.
Thermodynamics: an advanced textbook for chemical engineers / Gianni Astarita.
p. cm.
Includes bibliographies and index.
ISBN 978-1-4899-0773-8
1. Thermodynamics. I. Title.
QD504.A88 1989 89-3521
660.2 / 969-dcl9 CIP

ISBN 978-1-4899-0773-8 ISBN 978-1-4899-0771-4 (eBook)


DOI 10.1007/978-1-4899-0771-4

First Printing-March 1989


Second PrintingMay 1990

1989 Springer Science+Business Media New York


Originally published by Plenum Press, New York in 1989
Softcover reprint of the hardcover 1st edition 1989

All rights reserved


No part of this book may be reproduced, stored in a retrieval system, or transmitted
in any form or by any means, electronic, mechanical, photocopying, microfilming,
recording, or otherwise, without written permission from the Publisher
PREFACE

If a Writer would know how to behave himself with


relation to Posterity; let him consider in old Books,
what he finds, that he is glad to know;
and what Omissions he most laments.
Jonathan Swift

This book emerges from a long story of teaching. I taught chemical engineering
thermodynamics for about ten years at the University of Naples in the 1960s,
and I still remember the awkwardness that I felt about any textbook I chose to
consider-all of them seemed to be vague at best, and the standard of logical
rigor seemed immensely inferior to what I could find in books on such other
subjects as calculus and fluid mechanics. One of the students in my first class
(who is now Prof. F. Gioia of the University of Naples) once asked me a question
which I have used here as Example 4.2-more than 20 years have gone by, and
I am still waiting for a more intelligent question from one of my students. At the
time, that question compelled me to answer in a way I didn't like, namely "I'll
think about it, and I hope I'll have the answer by the next time we meet." I
didn't have it that soon, though I did manage to have it before the end of the
course. That episode convinced me that a discipline in which a sophomore, no
matter how clever, can ask a simple question which leaves the instructor struggling
for an answer is a discipline which needs serious rethinking.
In the early 1970s I had the great fortune of working with Prof. G. C. Sarti
at the University of Naples, and he and I (it was mostly his effort) succeeded in
making some sense, or at least so we thought, of thermodynamics. I was also
.busy writing a book on non-Newtonian fluid mechanics with my colleague
Prof. G. Marrucci, and I dared to volunteer to write a short section on the
thermodynamic aspects of the subject. It was not easy; one cannot deal with the
thermodynamics of non-Newtonian flow without understanding the basic con-
cepts of thermodynamics quite a bit more thoroughly than is required to simply v
vi solve elementary phase and chemical equilibria problems. The work done at that
time resulted in the publication of a short book (An Introduction to Nonlinear
Preface
Continuum Thermodynamics, SpA Editrice di Chimica, Milan, 1975), which deals
mainly with the thermodynamics of viscoelasticity. Chapter 5 in this book is a
spinoff from that book, but the material is presented here without the additional
complications of tensor analysis and differential geometry. Prof. Sarti took over
the teaching of undergraduate thermodynamics in Naples; he chose the approach
taken in this book, and was very successful in doing so. Unfortunately, he moved
to the University of Bologna shortly afterward, and his departure was a great
loss for the University of Naples.
In the late 1970s I taught for two consecutive years the graduate thermo-
dynamics course in the chemical engineering department of the University of
Delaware, at the suggestion of Prof. A. B. Metzner, who was chairman at the
time. I spent about half the time covering the thermodynamics of viscoelasticity,
and the other half on the more classical subjects of phase and chemical equilibria.
For that second part I prepared notes for the students; those notes constituted
the very first preliminary draft of this book. Little, if anything at all, has survived
from those notes to the present text, but the crucial ideas were developed at that
stage. For some time afterward I toyed with the idea of writing a book on
thermodynamics, only to find myself in a state of depression at the idea that
nobody would ever read it-there is no subject like thermodynamics to guarantee
that any two people interested in it will be at odds with each other concerning
how to teach it. But the small book published in 1975 got favorable reviews, I
got older and somewhat less touchy about what other people would think of
what I write, and in the early 1980s I started working in earnest on this project,
at the suggestion of Prof. M. E. Paulaitis of Delaware. Prof. Paulaitis has been
involved in this project for quite some time, has offered criticisms and comments,
and has taught with me a graduate course based on a later draft of the book.
Other people have been helpful in a variety of ways. Graduate students at
Delaware have offered comments on the draft distributed to them. Prof. C. A.
Truesdell read and criticized early drafts of the first chapters. Finally, Dr. R. E.
Rosensweig of Exxon Research and Engineering and Prof. Marrucci have
contributed the chapters on electromagnetic phenomena and on polymers.
Whenever I have taught a course on thermodynamics, I have begun by
presenting an overview of the discipline, and the Introduction is meant to provide
such an overview. I have also always asked the students to read a paper I published
in 1977 (Industrial and Engineering Chemistry, Fundamentals, 16, 138), which
deals with the historical and philosophical background of thermodynamics. A
grasp of the historical background of any discipline is, in my opinion, a useful
starting point.
... there ain't nothing more to write about, and
I am rotten glad of it, because if I'd'a' knowed what
trouble it WIU to make a book I wouldn 't 'a , tackled it,
and ain't a-going to no more.
Mark Twain

Gianni Astarita
Newark and Naples
CONTENTS

INTRODUCTION. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

PART ONE. MACROSCOPIC THEORY

CHAPTER ONE. FIRST AND SECOND LAWS .................. . 11


Notation ................. ...................................... . 11
1.1. Body and State ............................................ . 13
1.2. Energy and Heating ....................................... . 15
1.3. Work and Kinetic Energy ................................... . 16
1.4. The First Law ................................... : ........ . 19
1.5. The Second Law .......................................... . 21
1.6. Local Form of the Second Law .............................. . 23
1.7. Irreversibility and Dissipation ....... , ....................... . 24
1.8. Thermal Engines .......................................... . 26
Examples and Problems ......................................... . 30
Literature ...................................................... . 32

CHAPTER TWO. STATE AND EQUILIBRIUM ................... . 35


Notation .... ................................................... . 35
2.1. State and Site ............................................. . 37 vii
viii 2.2. Pressure and Work. .... ............ ...... .. ...... .......... 40
2.3. Viscosity and Relaxation ... -. . . . . . . . . . . . . . . . . . . . ... . . . . . . . . .. 43
Contents
2.4. The Second Law and External Systems. . . . . . . . . . . . . . . . . . . . . . . . 46
2.5. Internal Systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.6. Baric Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.7. The Maxwell Relations........ ........ ...... ...... ... . . . .... 55
2.8. Absolute Temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.9. Summary of Conceptual Issues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Examples and Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 65

CHAPI'ER mREE. HOMOGENEOUS REACTIONS... . . . . . . . . . . . . 67


Notation...... ....... . ....... ... .. . ....... ..... ........... ...... 67
3.1. A Review of Stoichiometry........... ..... ....... . ... ....... 71
3.2. Thermostatics of Homogeneous Reactions .... . . . . . . . . . . . . . . . . . 75
3.3. Partial Molar Properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.4. Fundamentals of Chemical Equilibrium Theory ................ 81
3.5. Electrochemical Reactions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 84
3.6. Continuous Description of Reacting Mixtures. . . . . . . . . . . . . . . . . . 87
Examples and Problems .............. '. . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 98

CHAPTER FOUR. PHASES ................................. :.... 101


Notation . ..................................................... " 101
4.1. One-Component, Two-Phase Systems ......................... 103
4.2. Phase Equilibria. . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 108
4.3. Heterogeneous Chemical Equilibria. . . . . . . . . . . . . . . . . . . . . . . . . .. 113
4.4. Liquid Crystals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . .. . . . . . . .. 117
4.5. Special Transitions ......................................... 119
Examples and Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 126
Literature .................................................... .- .. 129

CHAPTER FIVE. mERMODYNAMICS OF RELAXATION ........ 131


Notation . ....................................................... 131
5.1. Introduction to Relaxational Systems ......................... 135
5.2. Elementary Thermodynamics of Relaxation ... '" . . . . . . . . . . . . . .. 142
5.3. Equilibrium and Dissipation in Relaxational Systems ........... 145
5.4. Entropic Elasticity and Relaxation. . . . . . . . . . . . . . . . . . . . . . . . . . .. 147
5.5. Propagation of Discontinuities ............................... 150
Examples and Problems .......................................... 152
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 153

CHAPTER SIX. SURFACE mERMODYNAMICS ................. 155


Notation . ....................................................... 155
6.1. Surface Tension. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 157
6.2. Surface Phenomena in Mixtures ............................. . 162 ix
Examples and Problems ......................................... . 165
Contents
Literature ...................................................... . 165

CHAPTER SEVEN. DISSIPATIVE PHENOMENA ................ . 167


Notation . ...................................................... . 167
7.1. Heat Transfer ............................................. . 171
7.2. Hydrostatics .............................................. . 173
73. Diffusion ................................................. . 176
7.4. Momentum Transfer ....................................... . 183
7.5. Unsteady Transport ........................................ . 185
7.6. Coupling ................................................. . 187
7.7. The Symmetry Relations .................................... . 191
7.8. The Curie and Minimum Entropy Production Principles ........ . 193
7.9. Biological Systems ......................................... . 198
Examples and Problems ......................................... . 202
Literature ...................................................... . 202

PART TWO. ENGINEERING THEORY

CHAPTER EIGHT. EQUATIONS OF STATE ...................... 207


Notation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 207
8.1. The Ideal Gas .............................................. 211
8.2. One-Component Systems .................................... 215
8.3. Ideal Mixtures ...................................... -....... 221
8.4. Activity Coefficients. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 226
8.5. Dilute Solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 227
8.6. Constitutive Equations for One-Component Systems .... . . . . . . .. 229
8.7. Constitutive Equations for Mixtures .......................... 232
8.8. Molar Units and Colligative Properties. . . . . . . . . . . . . . . . . . . . . . .. 235
Examples and Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 240
Literature ....................................................... 241

CHAPTER NINE. PHASE EQUILIBRIA .......................... 243


Notation .. ...................................................... 243
9.1. Gas-Liquid Equilibria ...................................... 247
9.2. Liquid-Liquid Equilibria .................................... 253
9.3. Multiphase Systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 256
9.4. Gas-Solid Equilibria ........................................ 260
9.5. Phase Equilibria in Continuous Mixtures . . . . . . . . . . . . . . . . . . . . .. 262
Examples and Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 266
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 267
x CHAPTER TEN. CHEMICAL EQUILIBRIA ....................... 269
Contents Notation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 269
10.1. Homogeneous Equilibria in Ideal and Dilute Solutions ......... 271
10.2. Homogeneous Equilibria in Nonideal Mixtures. . . . . . . . . . . . . . .. 274
10.3. Heterogeneous Equilibria ................................... 276
10.4. Activities................................................. 278
10.5. One-Component Nonreactive Phases .............. , .......... 280
10.6. Nonreactive Mixtures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 284
10.7. Chemical Equilibria in Continuous Mixtures .................. 287
Examples and Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 288
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 289

CHAPTER ELEVEN. ELECTROCHEMISTRy...................... 291


Notation . ....................................................... 291
11.1. Strong Electrolytes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 293
11.2. Electrochemical Potentials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 300
11.3. Weak Electrolytes ......................................... 303
11.4. Electrochemical Reactions .................................. 314
Examples and Problems ........................................ " 318
Literature .......................... " ........................... 318

CHAPTER TWELVE. POLyMERS ................................ 319


G. Marrucci
Notation . ....................................................... 319
12.1. Introduction .............................................. 323
12.2. Chain Conformations. The Random Walk ...... . . . . . . . . . . . . .. 324
12.3. Rubber Elasticity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 327
12.4. Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 333
12.5. Mixtures................................................. 344
12.6. Rigid Polymers .......................................... " 354
Examples and Problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 361
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 362

CHAPTER THIRTEEN. THERMODYNAMICS OF


ELECTROMAGNETISM. . . . . . . . . . . . . . . . . .. 365
R. E. Rosensweig
Notation . ....................................................... 365
13.1. An Overview .............................................. 369
13.2. Electromagnetic Units. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 370
13.3. Electromagnetic Theory ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 371
13.4. Electromagnetic Work with Constant Mass Density. . . . . . . . . . .. 375
13.5. Electromagnetic Work with Variable Mass Density. . . . . . . . . . . .. 382
13.6. The Gibbs Equation and Thermodynamic Relations. . . . . . . . . . .. 389
13.7. Equilibrium in Multiphase and Multicomponent xi
Polarizable Systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 396
Contents
13.8. Applications to Problems of Equilibrium in a Field . . . . . . . . . . .. 405
13.9. Applications to Problems of Transport in a Field . . . . . . . . . . . . .. 420
Literature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 436
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 437

Index ............................................................ 439


INTRODUCTION

A textbook is a sort of table d'hote to which one may


sit down and satisfy his hunger for information,
with no thought of the complex agricultural processes
which gave rise to the raw material, nor of the mills
which converted these raw materials into foodstuffs,
nor of the arts of cookery responsible for
the well-prePared meal which is set before him.
It has not been our desire to
offer such a repast to our reader.
G. N. Lewis and M. Randall

Thermodynamics, like any other branch of engineering science, is really made


up of three parts: fundamental laws, constitutive equations, and engineering
applications. What is somewhat peculiar about thermodynamics is the relative
importance of these three parts, and the way in which they can be approached.
These peculiarities are perhaps best discussed by contrasting thermo-
dynamics with another fundamental discipline of engineering science, i.e., fluid
mechanics.
Let us first consider the fundamental laws. These are general laws of physics
which are postulated to hold for every macroscopic system (by macroscopic we
mean accessible to direct measurement, and hence of a scale large as compared
to molecular dimensions). In thermodynamics, these are the first and second
laws--and perhaps the third. Analogously, in fluid mechanics the physical laws
are the conservation of mass, of linear momentum, and of angular momentum.
The peCUliarity of thermodynamics is that one of its two fundamental laws, the
second, is not a principle of conservation, and is in fact written not as an equality
but as an inequality. There is no analog to that in any other branch of engineering
science.
No engineering science can be developed without the introduction of con-
stitutive equations, i.e., equations which hopefully describe the macroscopic (with
the word used in the same sense as before) behavior of some restricted class of
systems in some restricted class of phenomena, but are by no means regarded as
universal truths. In the fluid mechanics of liquids, the usual constitutive equations
are: (1) density is constant; and (2) stress in excess of the isotropic pressure is
proportional to the symmetric part of the velocity gradient (as a consequence,
2 only two parameters characterizing the specific fluid under consideration are
introduced: density and viscosity; these are easily measured, and the question
Introduction
is settled right there, at the macroscopic level). These two constitutive equations
are then combined with the balance equations, stress is eliminated, and the
Navier-Stokes equations for incompressible fluids are obtained. With this, the
first two parts of classical fluid mechanics are exhausted. The situation is by no
means so simple in thermodynamics, as discussed below.
The constitutive equations of interest in thermodynamics are significantly
complex, and the matter cannot be disposed of by measuring the values of a few
parameters. In fact, even the form of the constitutive equations (which of course
needs to be established before one can even define which parameters are of
interest, let alone establish how to measure their values) is under scrutiny in
thermodynamics. Moreover, while the balance equations (Le., fundamental laws
which express a principle of conservation) are always needed in order to actually
obtain a well-formulated mathematical problem which can in principle be solved
for assigned initial and boundary conditions, that is not the case for the second
law of thermodynamics. This point deserves some detailed comment.
Again let us consider classical fluid mechanics ..The fundamental laws are
the principles of conservation of mass, momentum, and angular momentum. In
the absence of body couples, the latter reduces simply to the requirement that
the stress tensor should be symmetric. Thus the two remaining balance equations,
one scalar and one vectorial, introduce three variables-density, velocity, and
stress--one scalar, one vector, and one tensor. Newton's linear law of friction
(which is the constitutive equation for the stress in excess of pressure), and a
pressure-density constitutive equation, close the problem. This is true for iso-
thermal fluid mechanics. As one moves to nonisothermal fluid mechanics, a new
variable, temperature, is introduced, and a new balance equation is needed,
namely, the balance of energy, say the first law of thermodyna~ics. This in tum
introduces two new variables, the heat flux vector and the internal energy, for
both of which a constitutive equation must be written down (say Fourier's law
for the heat flux, and a simple specific heat constitutive assumption for internal
energy), and the problem is closed again. The second law is not needed to close
the problem.
The second law of thermodynamics has a very peculiar role: it imposes
restrictions on the allowable forms for the constitutive equations, as well as
relationships between the latter. These restrictions are related to the fact that,
for every process which is a solution of the mathematical problem for any
conceivable set of initial and boundary conditions, the second law should never
be violated. It follows that the macroscopic theory includes, in the case of
thermodynamics, what can be called a theory of constitutive equations. The
second law does not tell us what the constitutive equations are, it only tells us
what they cannot possibly be. In the sense just discussed, the constitutive equations
of all branches of engineering science come under the scrutiny of thermodynamics.
This makes the scope of the macroscopic theory of thermodynamics significantly
wider than it is for any other engineering science. Finally, as far as the macroscopic
theory of thermodynamics is concerned, the second law introduces two funda-
mental distinctions: that between reversible and irreversible phenomena, and that
between equilibrium and non equilibrium conditions.
Once the fundamental laws and constitutive equations are written down, it 3
is possible to address the question of the actual solution of specific problems, Introduction
which is the part we refer to as the engineering theory. In most branches of
engineering science the largest fraction of the effort is devoted to this part.
Again, using classical fluid mechanics as an example, the body of the disci-
pline is dedicated to the study of the techniques needed to obtain solutions
to the Navier-Stokes equations for assigned initial and boundary conditions;
the deduction of the Navier-Stokes equations exhausts what we call the macro-
scopic theory. For the reasons discussed before, in thermodynamics the scope
of the macroscopic theory is at least comparable to that of the engineering
theory.
The macroscopic and engineering theories constitute the two parts of this
book; in the engineering part, the constitutive equations are regarded as known.
Thermodynamics, in contrast to many (but not all) other branches of engineering
science, includes, however, a third part, which is usually referred to as statistical
thermodynamics, and which we prefer to call the microscopic theory. In this
third part, molecular-scale models are constructed and constitutive equations
inferred from such models-the fact that the methodology for inferring the
constitutive equations is statistical in nature is incidental. The important point
is that the constitutive equations under scrutiny are often very complex, and
therefore molecular-scale models are needed. Again the example of classical fluid
mechanics is useful as a contrast. The constitutive equation (Newton's linear law
of friction) is very simple, in fact the simplest possible one which does allow a
distinction between the stresses exhibited at rest (an isotropic pressure) and those
exhibited in flow. It is the simplicity of the constitutive equation which results
in the fact that only one easily measurable parameter, the viscosity, is introduced.
A molecular-scale model capable of predicting the viscosity of a liquid is not
impossible, and in fact such models are available; but they are not regarded as
part of classical fluid mechanics, which is content with being restricted to the
macroscopic formulation. However, even in fluid mechanics, when materials such
as molten polymers are considered and which do not obey the simple linear law
of friction, the question arises about how to write down more complex constitu-
tive equations. The macroscopic theory of thermodynamics may tell us how
not to write them, but that by itself is not all that is needed. Molecular models
for the mechanical response of polymeric fluids are indeed common in the
literature.
In thermodynamics, should all common systems, or at least a large majority
of them, behave as ideal gases and ideal solutions, there would be no need for
a microscopic theory (and yet, curiously enough, the only microscopic model
which may be regarded as rigorously developed, the Maxwellian theory of gases,
has to do with ideal gases). However, this is unfortunately not the case. Hence
the need for a microscopic theory by means of which one may hope to infer the
more complex constitutive equations from whatever understanding we may have
of the molecular-scale phenomena involved.
There is very little microscopic theory in this book, and whatever there is is
concerned with somewhat unusual aspects of it. There are many excellent books
on statistical thermodynamics, and whatever we could have written would have
been a poor substitute for these.
4 Scope

Introduction
"Would you tell me please,
which way I ought to go from here?"
"That depends a good deal on where you want to go,"
said the cat.
Lewis Carroll

So far, we have discussed the architecture of this book; let us now tum our
attention to its scope. Writing a book implies that one has a certain type of reader
in mind. The reader is seen as a student of chemical engineering who has already
taken a beginner's course in classical chemical engineering thermodynamics,
where the skills of computing such things as simple phase and chemical equilibria
have been acquired. Those skills could be improved substantially with access to
modem computing facilities, and this would certainly be a fruitful area of writing,
but it has not been included in this book for two reasons. First, it is not the
author's area of expertise; second, the new edition of S. I. Sandler's book is a
masterpiece on the subject. This book is based on a different set of considerations.
Chemical engineering graduates will more and more apply their skills to nontradi-
tional problems: polymers, biological systems, and materials for the electronic
industry are just some of the most conspicuous examples. This being so has been
recognized in many chemical engineering departments, and special courses on
such subjects have been introduced covering these nontraditional areas. However,
we believe that also the core courses should reflect the new challenges facing the
profession; to quote from the dedication of the Lewis and Randall 1923 book
on thermodynamics, "the fascination of a growing science lies in the work of the
pioneers at the very borderland of the unknown, but to reach this frontier one
must pass over well traveled roads; of these one of the safest and surest is the
broad highway of thermodynamics." This means that the thermodynamics of
such nontraditional phenomena will need to be analyzed, and this poses a twofold
challenge. On the one hand, thermodynamic theory needs to be developed with
a methodology of broad enough generality in order to support its application to
nontraditional areas; on the other, some .elements of such applications need to
be discussed. As for the first challenge, it is hoped that the first part of this book
succeeds in presenting the appropriate methodology. As for the second challenge,
an attempt has been made at presenting some of the relevant material; we do
not claim to have exhausted the area (and probably it never will be), but we
hope to have achieved a first step in the right direction. One chapter is dedicated
to the thermodynamics of relaxation, which is crucial in the case of polymeric
materials, and one each to surface phenomena, dissipative processes, polymers,
and the thermodynamics of electromagnetic phenomena. In addition, some dis-
cussion is given of such matters as liquid crystals, biological systems, the con-
tinuous description of reacting mixtures, etc.
Consideration of biological systems brings into relief another important
point. Biological systems are never in equilibrium in a thermodynamic sense: the
equilibrium condition for a biological system is simply death. It follows that
either one abandons any hope of applying thermodynamic analysis to biological
systems, or one must regard thermodynamics as applying to nonequilibrium
conditions. The macroscopic theory of thermodynamics, which in large measure
is a theory of constitutive equations, imposes restrictions on the constitutive 5
equations describing irreversible processes just as it does for the eqUilibrium Introduction
constitutive equations. In both cases, it does not tell us what the constitutive
equations are, but only what they cannot possibly be. In this book, dissipative
processes are discussed in some detail, and attention is paid also to the question
of coupling between different dissipative processes, which is crucial in the case
of biological systems. The microscopic theory provides tools for predicting the
constitutive equations relative to dissipative processes, just as it does for the
classical equilibrium properties, and this is discussed at several points in the book.

Units and Dimensions


From midnight on January 1, 1990, the whole country will
convert to metric time. There will be 10 seconds to the minute,
10 minutes to the hour, 10 hours to the day, and so on:
Old time New time
1 second 1 milliday
1 minute 1 centiday
1 hour 1 deciday
1 day 1 milliyear
1 week 1 decaday
1 month 1 hectaday
1 year 1 kiloday
The fortnight will be withdrawn. Lectures will last .5 decidays,
which represents 1.2 old hours; professors are urged to
talk more slowly in order to fill the time.
C. Schulz

An important decision to be made in writing a book is what policy one


should follow as far as dimensions and units are concerned. As for units, the
contemporary tendency in science is for universal use of the SI units: rational
as such a choice might be, it cannot be denied that it presents a certain awkward-
ness. Furthermore, adoption of a universal set of units, by eliminating all problems
of conversion, may in fact make the conceptual content of dimensional analysis
more difficult to grasp.
The choice of an "appropriate" system of units is particularly awkward in
the case of thermodynamics, for a number of reasons, which will be discussed
later; but before such a choice is made, one needs at least to agree on what are
the dimensions of the quantities one is talking about-and even here there are
some subtleties in the case of thermodynamics. A sophisticated subtlety is that
of the dimensions of temperature, which, if one takes the Maxwellian theory of
gases seriously, should be those of a specific energy. A more earthly subtlety is
concerned with the widespread habit, in thermodynamics, of writing equations
which contain the logarithm of a dimensional quantity, like pressure P or con-
centration c. What are the dimensions of, say, In p? Or of the pH of an electrolyte
solution?
As far as dimensions are concerned, we have chosen to regard temperature
and number of moles as having their own fundamental dimension; consequently,
molecular weight is not dimensionless. The quantity In p is of course dimension-
less, and on those rare occasions where the equation cannot easily be transformed
to a form containing In PI - In P2 = In( PI/ P2) where no trouble arises, it simply
means that the units for P have been established once and for all.
6 This question of units could, of course, be solved by adhering to the SI
units-to the benefit of those few people who know what the viscosity of water
Introduction
is when measured in pascals/s, or what atmospheric pressure is in Newton per
square meter, and so on. However, the majority of us are accustomed to activation
energies expressed in kcal/kmol, viscosities in centipoise, and kinetic energies
in whatever units are customary in one's country, but certainly not calories or
BTUs. This makes for a totally inconsistent set of units-but it is the one we
normally use, and when problems arise we have to keep track of conversion
factors-which is a useful exercise anyhow.
The problem of units in thermodynamics is connected, of course, with the
fact that thermodynamics deals with so many different phenomena. People who
restrict themselves to one branch of science have an easier life; those working
in mechanics never use calories as a unit for energy, but whatever unit they use
is not usually the same as that used by people working in electromagnetism;
mass units are grams, or pounds, or even, on occasion, stones, but moles are
commonplace in chemistry; and when one is dealing with interconversion of
mechanical or electrical energy into thermal energy, one is in a bind. What units
should one choose for the gas constant, or, even worse, for the Faraday constant?
Coulombs per equivalent may not be esthetically elegant, but those are the units
in which we remember the value of F-96,500. And what are the dimensions,
let alone the units, of Avogadro's number? molecules/kmol, or kmol- 1 ?
Another choice to be made is that of notation. One could try to use a different
symbol for every different quantity dealt with, but the available number of letters,
even including the Greek alphabet, is by far too small for that, and one would
need a variety of subscripts and superscripts, in addition to all those which arise
anyhow when dealing with thermodynamics. Also, of course, one would like to
have a different symbol for chemical potentials and viscosities, but both are so
commonly indicated by J.L that the result would be quite awkward. Symbol A is
free energy and S is entropy and a is activity; what is the surface area going to
be?
On the basis of these considerations, the following choice has been made.
Within each chapter, one symbol is used for one and only one quantity; but the
same symbol is allowed to mean different things in different chapters. Thus every
chapter begins with a Notation section. In this section, units are given and are
chosen as those most commonly used for that quantity. The units are not consistent
with each other: they are simply given to help the reader keep track of the
dimensions of all quantities. Pressure, for instance, is always listed as having
units of atm, and volume as m3 ; but energies are often listed as kcal, and of
course 1 atm x 1 m3 is not 1 kcal. How many kcal it is, the reader should be able
to work out without help.
Finally, it should be remembered that there are quantities for which either
the dimensions are not even defined, or for which there is room for choice in
their definition. For instance, what are the dimensions of the state of a body?
Even for the simplest possible case, the state is the ordered pair volume, tem-
perature, and of course the dimensions of such a pair are not defined. Also,
sometimes a symbol stands for any member of a collection of quantities which
do not all have the same dimensions (e.g., any extensive property). In these cases,
no units at all are indicated (while for dimensionless quantities we indicate units
of -). Finally, in the case of functions of one or more dimensional arguments, 7
whose value is a dimensional quantity, I have chosen to indicate the units of the
Introduction
value.

Problems

Some persons have contended that mathematics ought to be


taught by making the illustrations obvious to the senses.
Nothing can be more absurd or injurious: it ought to be our .
never ceasing effort to make people think, not feel.
C. Coleridge

There is an infinite variety of number-crunching kind of problems in ther-


modynamics, and many of these are easily available in standard books on the
subject. No attempt has been made to include any problems of this type, in view
of the reader we have in mind. The problems included are more thought-
provoking; they tend to be nontraditional, and are often open-ended. Indeed, it
may be possible that a course based on this book should include an oral
examination, since that is the best form to test deep comprehension of the
concepts, rather than ability to solve specific quantitative problems.
Part One

MACROSCOPIC THEORY
Chapter One

FIRST AND SECOND LAWS

We might reason, a priori, that such absolute


destruction of living force cannot possibly take place;
because it is manifestly absurd to suppose that the
powers with which God has endowed matter can be
destroyed, any more than that they can be created by
man's agency; but we are not left with this argument
alone, decisive as it must be to any unprejudiced
mind. . .. Experiment has enabled us to answer these
questions in a satisfactory manner, for it has shown
that wherever living force is apparently destroyed, an
equivalent is produced which, in process of time, may be
reconverted into living force.
James P. Joule, 1847

NOTATION

A Free energy density kcal kg-I


lEI A body
aIEl External surface of the body
b Field acceleration ms- 2
If ) Constitutive function for pressure atm
F Force kgms- 2
J, Rate of entropy supply kcal kg- 1S-1
J Radiant entropy supply kcal kg-I S-I
j Entropy flux kcal m -2 S-I K- 1
K Kinetic energy kcalkg- 1
L Any constitutive quantity
l( Constitutive function for L
m Mass kg
II, Total number of moles kmol
p Pressure atm
p Power kcals- 1
IP( A process 11
12 Q Radiant heat supply kcal kg-I S-I
Chapter One Q. Rate of heating kcal8- 1
Q+ Rate of heat supply kcal S-I
Q- Rate of heat removal kcal S-I
q Local rate of heating kcal kg-I S-I
q Conductive heat flux kcal m- 2 S-I
R Gas constant kcal kmol- I K- 1
S Entropy density kcal kg-I K- 1
ds Surface element m2
Time s
T Temperature K
T+ Highest temperature at which heat is supplied K
r Lowest temperature at which heat is removed K
V Internal energy density kcal kg-I
V, Total internal energy kcal
V; Volume m3
v Velocity ms- I
loY. Net work rate kcal S-I
w Local work rate kcal kg-I S-I
Z Dissipation rate kcal kg-I S-I
ZT Thermal dissipation rate kcal kg-I S-I
ZM Nonthermal dissipation rate kcal kg-I S-I
r Efficiency of a heat engine
<I> Density kgm- 3
a State
Potential of body force S-2
cf>
cf>. Potential energy kcal

Subscripts Superscripts

B From body forces Time derivative


E From external forces + Where heat flows in
cond By conduction Where heat flows out
rad By radiation o Adiabatic
t Total
1.1. BODY AND STATE

We have previously called attention to the fact, which is


se/fevident, or at least becomes so if we take into consideration
the changes of volume occasioned by heat, that whenever there
is a difference of temperature the production of motive power is
possible. Conversely, whenever this power can be employed, it is
possible to produce a difference of temperature or to destroy the
eqUilibrium of the caloric. Percussion and friction of bodies are
means of raising their temperature spontaneously to a higher
degree than that of surrounding bodies, and consequently
destroying that equilibrium' in the caloric which had
previously existed.
S. Carnot, 1824

Thermodynamics, like any other science, makes use of a few primitive undefined
concepts, and both body and state are such concepts. A body B is endowed with
a fixed mass m" and occupies some finite region of space of volume V.. In
general, the volume V. will change in time, and so will the region of space
occupied by the body considered. [The subscript t (for "total") is used to indicate
that the whole body is being considered.] In thermodynamics, a body is often
referred to as a "closed system," i.e., .a system which does not exchange mass
with its surrounding, as contrasted to an "open system" which is some region of
space which may be occupied by different bodies at different instants in time.
We will use the symbol aD to indicate the instantaneous external surface of the
body.
Every branch of physical science is based on two sets of fundamental
equations. The first set is that of basic laws of physics, which are postulated to
hold valid for all bodies under all conceivable circumstances; the principles of
conservation of mass, of linear momentum, and of angular momentum are typical
examples. In thermodynamics, the basic laws are the first and seoond laws, which
are introduced in this chapter. The large majority of basic laws. of physics are
principles of conservation of some quantity (mass, linear momentum, etc.); the
first law of thermodynamics falls into this category, but the second law is an
exception, since it is not a principle of conservation.
The second set of fundamental equations are the constitutive equations:
these are relationships which are not supposed to hold for all bodies, but only
to describe the behavior of some restricted class of bodies, or possibly of a larger
class of bodies for a more restricted class of phenomena. A good example is that
of the mechanics of rigid bodies; it is of course obvious that there are many 13
14 bodies in nature which are not rigid (and perhaps one could argue that there are
in actual fact no bodies which are truly rigid); however, the theory of rigid bodies
Chapter One
is a useful abstraction which describes satisfactorily some phenomena as observed
in nature.
Constitutive equations are assumptions which may, or may not, adequately
describe the behavior of real bodies. In writing down constitutive equations, there
are several levels of assumptions which need to be made. The conceptually most
abstract level is the one for which it is postulated that some quantity is a physical
property of the body considered, and therefore its value depends only on the
physical condition of existence of the body' considered, i.e., on its state CT,. Such
quantities are called functions of state, or constitutive properties. If L is a
constitutive property, then the assumption that L is a function of state is equivalent
to the assumption that a mapping 1() exists which maps the state into the value
of L:

L = l(CT,) (1.1.1)

At this stage, l( . ) has to be regarded as a mapping, since the mathematical


nature of the state is still unspecified.
The next lower level of abstraction is to make an assumption about which
quantities determine the state of the body considered, i.e., to assign a mathematical
structure to the state CT,. There is a restriction imposed on quantities which may
contribute to determine the state: their values must in principle be measurable
by measurements made only on the body itself. Thus, for instance, the geometrical
position of the body with respect to other bodies cannot contribute to the
determination of the state. This of course results also in a restriction about which
quantities can be assumed to be functions of state: for example, the kinetic energy
of a body cannot be a function of state, since it depends on the velocities of all
points of the body considered, and such velocities in tum are only defined with
respect to other bodies.
The lowest level of abstraction is that where one assumes a specific functional
form for the constitutive mapping. The three levels of abstraction are perhaps
best understood by considering a simple example: a mass of a gas at a spatially
uniform temperature T. The highest level of abstraction is the assumption that
pressure is a function of state, say that a mapping f( . ) exists such that

p = f(CT,) (1.1.2)

The next lower level is to assume that the state is uniquely determined by
the value of the total volume of the gas, V;, and the value of temperature. With
this assumption, the mapping f( . ) reduces to a function of two variables:

p = f(V;, T) (1.1.3)

The lowest level of abstraction is that where one assigns a specific mathemati-
cal form to the function f( .). For instance, one may restrict attention to ideal
gases; one can write

p = RTn,/V; (1.1.4)
where n, is the total number of moles of the mass of gas considered and R is 15
the gas constant. While it is generally recognized that there may well be gases
First and
for which equation (1.1.4) does not hold (i.e., it is clearly recognized as a Second Laws
constitutive assumption), it is less commonly appreciated that also equation
(1.1.3) is a constitutive assumption, i.e., that it may well be possible that a gas
exhibits a pressure which is not uniquely determined by its total volume and its
temperature. Incidentally, even saying 'that the body considered is a gas is a
constitutive assumption, since of course bodies which are not gases do exist in
nature.
In this chapter, attention is focused on those results of thermodynamic theory
which can be obtained by restricting oneself to only the highestlevel of abstraction,
i.e., by simply making assumptions about which quantities are functions of state.
These results are few: essentially, only the highest possible efficiencies of thermal
engines can be calculated. Few as such results may be, they are infinitely more
than in any other branch of science, where no results whatsoever can be obtained
without going down to at least the second level.
It is of course obvious that the state of a body may change in time, and that
therefore also the values of all functions of state will change in time. A process
is defined as the mapping IP'( ) which maps time into the state:

O',=IP'(t) (1.1.5)

where t is time.

1.2. ENERGY AND HEATING

Now King David was old and stricken in years; and they
covered him with c/othes, but he could get no heat. Wherefore
his servants said unto him: "Let there be sought for my lord
the king a young virgin; and let her stand before the king, and
be a companion to him; and let her lie in thy bosom, that my
lord the king may get heat. .. So they sought for a fair damsel
throughout the borders of Israel, and found Abishag the
Shunnamite, and brought her to the king.
First Kings, 1.1-1.3

Given a body IB, at anyone time it is assumed to be endowed with an amount


of energy U, which is a function of state. No claim is made that U, is the total
energy of the body, since kinetic energy is not a function of state, as discussed
in the previous section. The value of U, will in general change in time.
It is furthermore postulated that U, is an absolutely additive function of
mass (in suggestive physical terms, this means that the internal energy of a body
is the sum of the internal energies of all its parts). Quantities which are absolutely
additive functions of mass are called "extensive." A local internal energy density,
U, can thus be defined such that

U,= L Udm (1.2.1)


16 Energy can flow from one body to another one by only two mechanisms.
The first mechanism is where one body does work on the other one; work is
Chapter One
defined in some underlying theory which is assumed to be known (typically,
mechanics, though also electrical work will be considered); some concepts are
reviewed in the next section. The other mechanism is that of heating, and the
symbol Qt will be used to identify the instantaneous rate of energy' influx into
the body due to heating. Energy is conserved in a global sense because any work
done on one body is done by some other bodies, and the same applies to heat:
any amount of heat flowing from body A to body B decreases the energy of body
A by exactly as much as it increases the energy of body B.
Heating may take place by two different mechanisms: conduction and radi-
ation. In the conductive mechanism, heat flows only between bodies, or parts of
a body, which are in contact with each other. At any point within a body a heat
flux by conduction vector q can be defined (boldface will always be used to
identify vectors). Since heat conduction occurs only by direct contact, the total
rate of heating due to conduction equals the conductive flux of heat through the
external surface of the body,

(Qt)cond = faB q ds (1.2.2)

where ds is the inward-pointing area vector on aBo


In the radiative mechanism, heat can flow from one body to another without
the two bodies being on contact with each other; in particular, radiative heat can
flow directly to interior parts of a body. If Q is the local rate of radiative heat
inflow per unit mass, the total rate of heating by radiation is

(Qt)rad = L Qdm (1.2.3)

and the total rate of heating is

Qt = (Qt)cond + (Qt)rad (1.2.4)

Since both q and Q are local quantities, it does not make much sense to ask
oneself whether they are or are not functions of the state of the whole body. As
will be seen later on, under some mild restrictive assumptions one can define a
local state u at any point of a body. When this is done, the heat flux q is taken
to be a function of the local state but the radiation density Q is not, as was
discussed in the preceding section. Equation (1.2.4) shows that therefore the
instantaneous rate of heating Qt is not a function of state.

1.3. WORK AND KINETIC ENERGY

The gentleman who, when I was young, bathed me at wisdom's


font for nine guineas a term-no extras--used to say he never
knew a bay who could do less work in more time. ..
J. K. Jerome
In mechanics, work is defined as the scalar product of force and the displace- 17
ment of the point where the force acts. Thus, given a force F acting on a material First and
point which moves with velocity v, the rate at which it is doing work, called the Second Laws
power P, is given by

P=Fov (1.3.1)

Forces acting on a body can be classified into contact forces, which act on
the external surface of the body, and body forces, which act directly on elementary
masses within the body (of course, there are also contact forces where one part
of the body does work on other parts, but these do not contribute to the influx
of energy into the body). The classical example of body force is, of course,
gravity. The total power Pt exerted on the body is the sum of the powers exerted
by all the forces cited above, and it can be regarded as the sum of the external
power PtE exerted by all contact forces acting on the surface of the body, and
the body power PtB exerted by all body forces. Ifb is the field acceleration (gravity
in the classical case), the body force acting on the elementary mass dm is b dm,
and therefore the body power is

(1.3.2)

An important subcase is when b can be expressed as the gradient of a scalar,


b = -grad I{>, with I{> independent of time (gravity is such a case, with I{> = gh, h
being height over a fixed horizontal plane). In this case, one has

bov= -v grad I{> = _1{>0


0
(1.3.3)

where I{> represents the rate of change of I{> as observed by the elementary mass
0

dm. (A superscript will always be used to indicate this type of rate of change,
called in general the substantial time derivative.) Thus one has

(1.3.4)

where I{>t is called the "potential energy" of the body:

I{>t= L
I{>dm (1.3.5)

We note that, should b not admit a time-independent potential, which it well


might, one could not even define a potential energy, but the power of body forces
would still be defined by equation (1.3.2).
The total kinetic energy of a body, K" is defined as

Kt=~ L
vovdm (1.3.6)
18 and thus its rate of change is
Chapter One
(1.3.7)

A fundamental theorem of classical mechanics states that for a rigid body,


or for any body which is instantaneously undergoing a rigid-body motion,

(1.3.8)

[We are here referring for simplicity to what in continuum mechanics is


called the nonpolar case, i.e., there are no body couples and thus the resultant
torques on elementary masses are zero. This restriction implies equation (1.3.8)
for rigid-body motions. The argument could be generalized by allowing for body
couples, and one would need to include work done by such couples in rotations.
The argument becomes rather complex, but the thermodynamic implications
would be essentially the same as obtained by considering equation (1.3.8).]
Equation (1.3.8) is a purely mechanical result, i.e., it is a consequence of
only Eulerian mechanics. In spite of this, it appears superficially as an energy
balance, which it absolutely is not, since no independent principle of conservation
of energy has been invoked in its derivation. Equation (1.3.8) applies only to
rigid-body motions, and as a consequence rigid-body motions can never result
in the transformation of energy absorbed as heat into work, or vice versa;
rigid-body motions are thermodynamically degenerate.
For a body which is not undergoing a rigid-J:lOdy motion, but is also under-
going some deformation, equation (1.3.8) does not apply. One can thus define a
net rate of work done on the body:

(1.3.9)

It is noteworthy that body forces contribute nothing to w.,


i.e., the work of
body forces goes entirely into increasing kinetic energy (see the Appendix to
Section 2.2). The net rate of work W. can be expressed as

W,= L wdrn (1.3.10)

where w is the local net rate of work per unit mass.

Appendix

It is important to realize that, in analogy with the total rate of heating Q"
the total net rate of work W. is not a function of state. The analogy is closer than
may appear at first sight. The reason why W. cannot be a function of state is that
body forces contribute to it, and body forces are active owing to the influence
of bodies other than the one considered (for instance, gravity is due to the pull
of the Earth). Contact forces are determined by the internal stresses in the body
considered (in particular, internal stresses on the surface of the body are the only 19
ones which contribute to W,), and internal stresses are functions of state, just as
First and
the heat flux by conduction; body forces are the analog of the radiant heat supply. Second Laws
Body forces do not perform any net work, but, by entering the momentum
balance, they contribute in determining the velocity field and hence the net work
done by contact forces.
In many textbooks on thermodynamics, the fact that the total rate of work
and the total rate of heating are not functions of state is expressed by the statement
that dW, and dQ, are not exact differentials. This statement deserves closer
inspection. Given a differential time increment dt, the corresponding differential
work done and heat supplied are, respectively,

dW, = W,dt and dQ, = Q,dt (l.3.A.I)

and it is difficult to assign any mathematical meaning to the statement that these
differentials are inexact. In mathematics, the concept of an inexact differential
"form" is well defined. In its simplest form, given two quantities M and N which
are functions of two variables x and y, the differential form

M(x, y) dx + N(x, y) dy (1.3.A.2)

is exact if a function F(x, y) exists such that its differential dF equals the
differential form (1.3.A.2). Since

dF = (aF/ax) dx + (aF/ay) dy (l.3.A.3)

the form (1.3.A.2) is exact provided that

aM/ay = aN/ax (1.3.A.4)

We note that the question of exactness arises for the form (1.3.A.2), not for
dF itself; all differentials have full citizenship's rights, and therefore the statement
that dW, and dQ, are not exact differentials is, strictly speaking, meaningless.
Perhaps some meaning could be attached to it by somehow identifying the pair
x, y in the form (1.3.A.2) with the state, but this does not seem to be a fruitful
line of inquiry.

1.4. THE FIRST LAW

In whatever system where the weight attached to the wheel


should be the cause of the motion of the wheel, without any
doubt the center of the gravity of the weight will stop beneath
the center of its axle. No instrument devised by human
ingenuity, which turns with its wheel, can remedy this effect.
Oh, speculators about perpetual motion, how many vain
chimeras have you created in the like quest. Go and take your
place with the seekers after gold.
Leonardo da Vinci, 1494
20 It is now possible to formulate the first law of thermodynamics, which
formalizes the postulate of the conservation of energy. If energy is neither created
Chapter One
nor destroyed, then the net rate at which energy flows into a body must be equal
to the rate at which energy is accumulated in the body,

U;= Qt+ W. (1.4.1)

Equation (1.4.1) shows completely the far-reaching implications of equation


(1.3.8). For a body which undergoes rigid-body motion, the work done results
in accumulation of kinetic energy and, since W. = 0, the heating results entirely
in the accumulation of internal energy: there is no mechanism of interchange
between the thermal and mechanical forms of energy. Since one of the main aims
of a thermodynamic theory is to describe such interchange, one comes to the
conclusion that deformations of bodies need to be considered. The continuum
mechanics of deformable bodies is a rather complex and sophisticated theory
but, fortunately enough, by limiting attention to only the very simplest example
of deformation, i.e., changes in density, a large variety of nontrivial thermody-
namic problems can be analyzed. This is what will be done in most of the balance
of this book.
The first law of thermodynamics can be formulated in a variety of alternate
ways which are obtained from equation (1.4.1) by algebraic manipulation. Substi-
tution of equation (1.3.9) yields

U;+K;= Qt+Pt (1.4.2)

which is often more useful than equation (1.4.1). Also, for those cases where
b = -grad tP, with tP independent of time, substitution of equation (1.3.4) leads
to

U; + K; + tP; = Qt + PtE (1.4.3)

which is very useful when integration over the residence time within an open
system is needed. Of course, equations (1.4.1)-(1.4.3) can be integrated over any
finite interval of time to obtain different macroscopic formulations of the first
law. Finally, equation (1.4.1), which applies to a body of finite size, can be
transformed to a field (oriocal) equation. Substitution of equations (1.2.1 )-( 1.2.4)
and (1.3.10) into equation (1.4.1) yields

r U' dm =
JII
JaR q' ds + r Q dm
JII
+ r w dm
JII
(1.4.4)

By making use of Gauss's theorem, the first integral on the right-hand side
is transformed to a volume integral:

JaD q-ds=_r (divq)dm/<I>


Js
(1.4.5)

where <I> is the local density.


Since equation (1.4.1), and hence also (1.4.4), is postulated to hold for all 21
bodies, equality of integrals implies equality of integrands, and hence First and
Second Laws
U' = -(divq)/cI> + Q + w (1.4.6)

which is the required local form of the first law. In equation (1.4.1), the internal
energy U, is regarded as a function of the state u, of the whole body. An additional
assumption is required in order to make the same statement about equation
(1.4.6), namely, that the internal energy density U at a point is a unique function
of the state in a neighborhood of that point, u. This assumption, which is generally
referred to as the principle of local action, is made in most formulations of
thermodynamic theory, but it is a restrictive assumption, since it excludes from
consideration long-range interactions within II: body.
For future reference, it is useful to define the local rate of heating per unit
mass, q, as

q = Q - divq/cI> (1.4.7)

whence the local form of the first law takes the simple form

U'=q+w (1.4.8)

and the global rate of heating is given by

Q,= L qdm (1.4.9)

1.5. THE SECOND LAW

The idea of a flow of entropy raises a certain conceptual


difficulty of which we are usually oblivious when we speak of
an energy flow. What precisely is meant by a flow of entropy?
The corresponding question in regard to energy flow does not
present itself so forCibly because there is always the comfortable
picture of energy as a kind offluid. It is an idea which remains
over from the caloric theory and obtains mental reinforcement
from the fact of conservation . ...
K. Denbigh

The second law of thermodynamics is a formalization ofthe intuitive concept


of the irreversibility of natural phenomena. In this, it is conceptually very different
from other fundamental principles of physics, which are in general formalizations
of a conservation principle (mass, linear momentum, energy, etc.). A conservation
principle states that some physical quantity is neither created nor destroyed;
hence balance equations such as equation (1.4.1) can be written down. In balance
equations, if the sign of all terms is changed, the equation still holds true; hence
such equations are invariant under a reversal of the direction of time. As a
concrete example, all terms appearing in equation (1.4.1) are rates, and they
22 therefore all change sign upon time reversal. The description of irreversibility
Chapter One
which is required from the second law imposes that it must be written as an
inequality, since inequalities fail to hold true when the sign of all terms is changed.
The second law of thermodynamics can be written by requiring the rate of change
of some quantity to never be less (but it can be more) than what one would
calculate should a conservation principle apply.
The quantity whose rate of change is involved is called the entropy. Just as
is the case for energy in the first law, entropy is a primitive undefined concept;
this being so is less easily acceptable simply because, in contrast to "energy,"
the word "entropy" is not part of familiar, everyday language. Given a body 18,
it is postulated that it possesses a total entropy St which (1) is an extensive
property, and (2) is a function of state. If Jt is the instantaneous rate of supply
of entropy to the body considered, then the second law of thermodymimics is
written in the form

S;~Jt (1.5.1)

and, since St is extensive, an entropy density S is defined as

St= fa Sdm (U.2)

The first and second laws are coupled if, as postulated below, the rate of
entropy supply Jt is related to the rate of heating Qt. In analogy with the conductive
heat flux q and the radiative heat density Q, an entropy flux J and a radiative
entropy density J are defined such that

(U.3)

in perfect analogy with equations (1.2.2)-(1.2.4). The quantities j and J are


postulated to be related to q and Q as follows:

j = q/T (1.5A)

J=Q/T (1.5.5)

where T is the temperature of the point considered. It is interesting to observe


that temperature is also a primitive undefined concept, so that two such concepts
enter into the formulation of the second law. Analogously, energy and heating
are the two primitive undefined concepts entering the formulation of the first law.
Combination of equations (1.5.1)-(1.5.5) yields the following form of the
second law:

S;~f
aB
(q/T)ds+ r (Q/T)dm
JB
(1.5.6)
and two special forms can immediately be deduced. First, if the temperature of 23
all points of the body considered is the same (it may change in time, but is
First and
constant in space at all times), the "isothermal" form of the second law is Second Laws

S;:2: Qt/T (1.5.7)

Under the condition that q is zero throughout alB and Q is zero throughout
IB (note that Qt could be zero under far milder conditions), the "adiabatic" form
is obtained:

(1.5.8)

Occasionally, equation (1.5.8) is said to imply that the entropy of the entire
Universe never decreases-something that equation (1.5.8) does by no means
necessarily imply, and which anyhow is a claim of scarce pragmatic utility. More
concretely, equation (1.5.8) can be used as the starting point of a theory of
equilibrium, which will be called thermostatics. This theory is often regarded as
exhausting the subject of thermodynamics, to the point that the word "thermody-
namic" is sometimes used to mean "at equilibrium." This is too restrictive a
viewpoint.

1.6. LOCAL FORM OF THE SECOND LAW

I do not approve of anything that tampers with natural


ignorance. Fortunately in England, at any rate, education
produces no effect whatsoever.
Oscar Wilde

Applying Gauss's theorem to the first integral on the right-hand side of


equation (1.5.6) and taking into account equation (1.5.2), the second law can be
written as

L S'dm:2: t[Q/T-O/<1diV(eJ/T)]dm (1.6.1)

which has to hold for any body IB; hence

S':2: Q/ T - 0/<1 div(q/ T) (1.6.2)

which is the local form of the second law. If again the principle of local action
is accepted, the entropy density at a point, S, depends only on the state of a
neighborhood of that point, CT.
It is convenient to expand the second term on the right-hand side of equation
(1.6.2) to obtain

S' :2: (1/ T)[ Q - (div q)/<1>] + q . grad T/(<1>T2) (1.6.3)


24 The term within square brackets is the rate of heating per unit mass of the
neighborhood of the point considered, q [see equation (1.4.7)], and so equation
Chapter One
(1.6.3) can be expressed in the simpler form

(1.6.4)

which yields directly some interesting conclusions. First, suppose the body con-
sidered is locally isothermal, say grad T = O. The isothermal local form of the
second law is

(1.6.5)

Now we consider the case of steady heat conduction through a stationary


body. Since the phenomenon is steady, S and U are zero. Since the body is
stationary, w is zero, and hence from the first law q is also zero. Hence equation
(1.6.4) reduces to

qgrad T:5.0 (1.6.6)

i.e., in steady heat conduction through a stationary body the heat flux vector
needs to have a positive component in the direction of decreasing temperatures.
We note, however, that this need not be universally true, not even if heat
conduction is the only phenomenon taking place, since the result has been
obtained by considering a steady-state process. (See Chapter 7 in this regard.)
Finally, the local rate of heating can be eliminated between equations (1.4.8)
and (1.6.5), yielding the following equation:

S - (U - w)/T - q. grad T/(4)T2);:: 0 (1.6.7)

which forms the basis for a theory of dissipation to be discussed in the next section.

1.7. IRREVERSIBILITY AND DISSIPATION

The customer cannot win at this game; this is the first law. In
fact, the customer is likely to lose; this is the second law.
(From the book of instructions for croupiers at the
roulette table.)

We consider a body which is undergoing some process u. = 1P.(t). At any


particular instant in time the se"cond law, in any of the equivalent forms discussed
above, may be satisfied either as a strict inequality ( or as an equality (=). In
the former case, the process is said to be instantaneously irreversible and, in the
latter case, to be instantaneously reversible. Analogous definitions apply to the
local form of a process, u = lP(t).
For a local instantaneously irreversible process, a local rate of energy dissipa-
tion, Z, can be defined. Of course, in view of the first law, "dissipation" should
not be understood as an actual loss of energy. The word is used in the (intuitive)
sense that, in any irreversible process, energy becomes progressively less available.
Since temperature is an intrinsically positive quantity, both sides of equation 25
(1.6.7) can be multiplied by T to yield First and
Second Laws
TS - U + w - q . grad T / (cp T) ~ 0 (1.7.1)

The local rate of energy dissipation per unit mass, Z, is simply defined as

Z = TS - U + w - q . grad T / (cp T) (1.7.2)

and hence the second law assumes the following, very simple form:

Z~O (1.7.3)

It is noteworthy that the first three terms on the right-hand side of equation
(1.7.2) do not contain any heating term, and therefore their sum may be called
the nonthermal dissipation rate, ZM, i.e.,

ZM = TS - U + w (1.7A)

The last term is uniquely related to the conductive heat flux and is called the
thermal dissipation rate, ZT, i.e.,

ZT = -q . grad T/(CPT) (1.7.5)

Hence

(1.7.6)

The second law requires Z to be nonnegative, but it does not require ZM


and ZT to be separately nonnegative. Hence, e.g., q grad T is required to be
nonpositive in steady heat conduction [see equation (1.6.6)], but in principle it
could be locally positive, even at steady state, if ZM is large enough.
This conclusion strikes a cautionary note about applying blindly some
informal expression of the second law, such as "heat flows in the direction of
decreasing temperature" [which is equivalent to equation (1.6.6)]. Such informal
expressions may be correct if the phenomenon considered is the only one taking
place, but they are wrong in general if alternate mechanisms of irreversibility act
simultaneously. For instance, the evolutionary behavior observed in biological
systems, which superficially appears to be contrary to the second law; is in fact
possible because alternate sources of irreversibility are at work at the same time.
If attention is restricted to situations where the temperature is uniform in
space (though it may change in time), the second law requires ZM to be nonnega-
tive. In this regard, it is useful to define the "Helmholtz free energy density," A,
as follows:

A=U-TS (1.7.7)
26 We note that A is an extensive quantity which is a function of state.
Chapter One
Differentiation of equation (1.7.7) with respect to time yields

A' = U' - TS' - ST (1.7.8)

so that the "isothermal" requirement that ZM should be nonnegative becomes

-A'-ST+w~O (1.7.9)

If, in particular, temperature is constant in time as well as in space, equation


(1.7.9) states that the net rate at which work is done on the element considered
per unit mass, w, cannot be less than the rate at which the Helmholtz free energy
density of the element increases. If the process is reversible, the two rates are
equal, which can be interpreted as meaning that the work done is being accumu-
lated in the form of free energy. Apart from this special interpretation for
isothermal processes, the Helmholtz free energy plays a crucial role in the
development of thermodynamic theory, as will be discussed in the next chapter.
Let us consider again the case of an isothermal reversible process, for which
the net rate of work done is equal to the rate of increase of the free energy
density. This is the classical assumption of the theory of elasticity, where the free
energy is generally called the stored elastic energy. Since the process is isothermal,
one has

w=A"= U'-TS' (1.7.10)

which shows that, in fact, elastic energy can be accumulated isothermally either
by an increase of internal energy, or by a decrease of entropy, or a combination
thereof. For certain classes of materials, the microscopic theory suggests that in
fact elastic energy is accumulated entirely by a decrease of entropy. Such materials
are said to have entropic elasticity.
The concept of entropic elasticity is not restricted to elastic materials, i.e.,
to materials which undergo only reversible processes. In fact, consider the
isothermal deformation of any material. The work done can be expressed as

w = ZM + A' = ZM + U' - TS' (1.7.11)

i.e., as the sum of the rate of nonthermal dissipation and the rate of accumulation
of elastic energy; the latter can still be accommodated by the two mechanisms
discussed earlier.

1.S. mERMAL ENGINES

To examine the principles of the production of motion by heat


in all its generality, it must be conceived independently of any
mechanism or of any particular agent; it is necessary to
establish proofs applicable not only to steam engines but to all
other heat engines, irrespective of the working substance and
the manner in which it acts.
s. Carnol, 1824
In this section, theoretical upper bounds for the efficiency of different thermal 27
engines are established. These bounds are totally independent of how the engine First and
actually works, and are direct consequences of the second law. Second Laws
A thermal engine is a machine which absorbs or emits both heat and work,
and possesses the following two properties:

1. In principle, it can work forever.


2. W, is not zero.

First we examine condition (1). An air conditioner, for instance, is a thermal


engine. A block of ice, which could accomplish the same aim of cooling down
a room, is not a thermal engine, since it will work only until the ice is all melted.
The ability to work forever means that the state of the machine, u" is either
constant in time or varies periodically. The analysis presented below is for the
steady-state case, where u, is constant in time, but the results apply also to the
cyclic case (one only needs to substitute time integrals over a cycle of any quantity
appearing in the equations below). Since the state is constant in time, so also
must be the values of all quantities which are functions of state, in particular

(1.8.1)

The first law thus yields

Q,+ w,=0 (1.8.2)

Condition (2) is now seen to be nontrivial, since it requires that Q, and W,


are not separately zero. Heat and work are not separately balanced, and energy
flowing into the machine by one mechanism flows out by another one. If the
body forces admit a time-independent potential (as in practice is always the
case), the total potential energy of the machine and its total kinetic energy are
also constant, and hence

Q,+P'E= 0 (1.8.3)

Some definitions are now needed. The external surface of the machine is
decomposed as follows:

(1.8.4)

where

>o
q. ds { <0 (1.8.5)
=0
28 The physical meaning is obvious. The three terms on the right-hand side of
Chapter One
equation (1.8.4) represent, respectively, that part of the external surface of the
engine through which heat flows into the machine by conduction, that through
which heat flows out, and the adiabatic part.
Analogously, the entire mass of the engine is decomposed in three parts
such that

(1.8.6)

Consideration of the radiant heat is pragmatically relevant: a solar cell is


an example of a thermal engine where radiant heat is of course crucial.
The quantities Q+ and Q- are defined by

(1.8.7)

and

Q- = - L- q ds - L- Q dm (1.8.8)

which have an immediate physical meaning: Q+ is the rate of heat entering the
engine, and Q- is the rate of heat exiting the engine. By definition, both Q+ and
Q- are nonnegative, and

(1.8.9)

Therefore, the first law can be written as

(1.8.10)

Before applying the second law, the following definitions are useful:

T+ = max(T) (1.8.11)

r = min(T) (1.8.12)

Again, the physical interpretation is immediate: T+ is the hottest point at


which heat is entering the engine, and T- is the coldest point at which heat is
exiting. The rate of supply of entropy is obtained from equations (1.5.3)-(1.5.5):

J, = [L+ (q/T)' ds+ L+ (Q/T) dmJ - [- L- (q/T)' ds- L- (Q/T) dmJ


(l.8.13)
Algebraic rearrangement yields 29
First and
(1.8.14) Second Laws

(The equality sign would only apply if all points at which heat is entering the
system were at temperature T+, and all points where it is exiting were at
temperature T-.)
Substitution of equations (1.8.1) and (1.8.14) into equation (1.5.1) yields the
following fundamental result, which is obtained solely from application of the
first and second laws to thermal engines:

(1.8.15)

Thermal engines can be classified in three categories: engines, refrigerators,


and heat pumps. These are briefly discussed in the following.

Engines. The aim of an engine is to produce mechanical power. The engine


is fed energy through heating. Thus the desired quantity is - PtE, the mechanical
power produced by the engine, and the quantity to be expended is Q+, the heating
rate. Correspondingly, the efficiency r is defined as

(1.8.16)

From equations (1.8.10) and (1.8.15) one obtains

(1.8.17)

The right-hand side is the highest possible efficiency of an engine. It j:an


never be larger than unity, and it only approaches unity if T- approaches absolute
zero. Real engines based on the classical Carnot cycle can approach the limit
given in equation (1.8.17).

Refrigerators. The aim of a refrigerator is to cool down some volume in


space, i.e., to extract heat from a cold source and to discharge it into a warmer
sink. In order to do so, the expenditure of some mechanical energy is needed,
and hence PtE> O. The efficiency is the ratio of the desired quantity, Q+, to PtE:

(1.8.18)

and one obtains immediately

(1.8.19)

In a refrigerator T- > T+ by definition, and hence r is positive; it can be


larger than unity. Real refrigerators approaching the theoretical upper bound
can be realized.
30 Heat Pumps. The aim of a heat pump is to furnish heat to some sink, while
absorbing heat from some colder source. Heat pumps work essentially the same
Chapter One
way as refrigerators; the only difference is that the desired quantity is the heat
discharged, Q-, rather than Q+. Correspondingly, the efficiency is defined as

(1.8.20)

and one obtains


(1.8.21)

The highest possible efficiency of a heat pump is always larger than unity.

Appendix

The possibility of establishing theoretical upper bounds to the efficiency of


thermal engines without making any assumption whatsoever about how such
engines may work is conceptually a very important point. In fact, these upper
bounds are the only results in any branch of science that are obtained without
any constitutive assumption whatsoever; for instance, in mechanics no problem
can be solved without some constitutive assumptions, since the fundamental laws
of mechanics do not furnish enough equations to solve any assigned problem.
The second law of thermodynamics is the only fundamental law of physics which
can be used directly, without any constitutive assumption, to obtain some concrete
results.
The pioneers just assumed (based on some metaphysical belief) that such
absolute upper bounds existed; see Carnot's quote at the beginning of this section.
They did more than that: in developing the concept of entropy, Clausius used
the concept that a thermal engine could be reversed to become a heat pump (see
the quote at the beginning of Section 2.1), in spite of the fact that no refrigerator
or heat pump existed in 1852. While the theory of engines was developed a
century after they had come into common industrial use, the theory of refrigerators
was developed before anyone was ever built.
The fact that absolute upper bounds exist has, of course, to do with the fact
that there is an absolute temperature scale which is the same for all substances;
this is discussed in Section 2.8.

EXAMPLES AND PROBLEMS

Examples

1. Consider an elevator of mass m which is moving upward with velocity U. There


are three forces acting on the elevator and they are all vertical, so that their vectorial
nature need not be considered: the pull P exerted by the rope, the weight W of the
elevator, and the friction against the shafts, D. An overall momentum balance yields
P = W + D + m dU / dt (I.E.l)
The total power exerted by all forces is given by 31
Pt = (P - W - D) U (I.E.2) First and
Second Laws
since P has the same direction as U, but Wand D have opposite direction. Substitution
of equation (l.E.1) into expression (l.E.2) gives

Pt = mUdU/dt = K; (1.E.3)

and hence W t = 0: there is no net work done on the elevator, which is consistent with the
fact that the elevator can be regarded as a rigid body. Hence the first law, in the form of
equation (l.4.2), reduces to

U;= Qt (I.E.4)

(with presumably both the right- and left-hand sides being zero).
Now suppose one wants to use the concept of potential energy. The potential energy
with respect to a horizontal plane located at h = 0 is Wh (with h measured upward), and
U = dh/ dt. Thus

</;= wu (I.E.S)

The power of contact forces is

PtE = (P - D) U = K; + "'; (I.E.6)

and thus the first law, in the form of equation (1.4.3), again reduces to equation (l.E.4).
The friction against the shafts, D, is of course a dissipative force. However, dissipation
does not take place in the elevator itself. To clarify this point suppose that, between the
elevator's slides and the shafts, there is a thin layer oflubricant. The surface of the lubricant
in contact with the slide is moving upward with velocity U; the surface in contact with
the shaft is stationary. The two surfaces are parallel to each other, and hence the layer of
lubricant is being sheared. Shearing is of course a deformation, which is possible since
the lubricant is quite obviously not a rigid body. The overall process is irreversible, since
dissipation of energy occurs in the lubricant layer at a rate DU; however, the process
which the elevator itself undergoes is nondissipative.

Problems

1.1. Consider those air conditioners which, in wintertime, are used as heaters. Suppose you
have to calculate the power required, assuming that the actual efficiency is close to the maximum
possible. Would you design based on the coldest day in winter or the hottest day in summer? Is the
answer to this problem going to be the same all over the US, or does it depend on geographical location ?
1.2. You may have seen those little toys shaped as birds which are pivoted on a horizontal
axis. The toy is started by tipping the bird's beak into a glass of water. After a while the bird straightens
up, stays in that position for several seconds, and then dips its beak into the water again; the trick
seems to go on forever. The toy works as follows. Water and air are at the same temperature; when
the beak temperature is the same, the equilibrium position is with the bird standing up, and that's
why the bird straightens out of the water. However, the beak is wet, and it cools down by evaporation
in the air; now the equilibrium position is with the beak down, and the bird moves again after enough
evaporation has taken place.
32 Since friction of the pivot axle is perhaps small but finite, there is energy dissipation. How does
the toy's going up and down apparently undefinitely square with the second law? You are only asked
Chapter One to give a qualitative explanation of the phenomenon.
1.3. Consult any book on refrigeration technique, and you will find a description of an
ammoniabased refrigeration cycle which only uses a heat source to drive it. Analyze such a cycle
and show that the upper limit to its efficiency is still given by equation (1.8.19).
1.4. When a heat pump is needed which pumps heat over a large temperature difference, it
may be impossible to realize it with anyone machine. In this case two machines are used: the first
one pumps heat from the cold source to some intermediate temperature, where it discharges it to the
cold side of the second machine; the latter pumps it up to the desired level.
Is the maximum possible overall efficiency of such a scheme equal to that which would hold in
the case of a single machine?

LITERATURE

Since entropy is an intrinsic property, its magnitude is


dependent only on the nature of the matter under consideration
and the state in which it exists and not on its external position
or motion relative to other bodies. Thus, the entropy of an
elevated object is no different from that of the same object in a
lower position, and the entropy of a rotating flywheel is equal
to that of the same flywheel at rest.
O. A. Hougen, K. M. Watson, and R. A. Ragatz

The formulation of the second law as given in equation (1.5.6) is usually called the
Clausius-Duhem inequality. Admittedly, there is no general consensus on whether the
second law should be formulated in general as the Clausius-Duhem inequality; see, e.g.,
A. E. Green and P. M. Naghdi, Proc. R. Soc. London, Ser. A 357, 253 (1977).
However, the Clausius-Duhem inequality has been very fruitfully employed in' a
number of recent analyses. A discussion of its logical foundations can be found in M.
Feinberg and R. Lavine, "Foundations of the Clausius-Duhem inequality," Appendix
2-A, in: C. A. Truesdell, Rational Thermodynamics, 2nd ed., p. 123, Springer-Verlag, New
York (1984). This appendix also discusses the mathematical structure of thermodynamic
theories. The book including this appendix is a good source for the foundations of the
thermodynamic approach taken in this first part.
It is difficult to pinpoint exactly where the Clausius-Duhem inequality was first stated
explicitly; the best source is B. D. Coleman and W. Noll, Arch. Ratl. Mech. Anal. 13, 167
(1963).
Historically, the Clausius-Duhem inequality has developed as follows. The work of
Planck in 1887 implied that, in general, the second law could be written as dU :5 dw + T dS.
If the differentials are interpreted as rates, this corresponds to U":5 W + TS", which has
been called the Clausius-Planck inequality, and it corresponds to omitting the last term
on the left-hand side of equation (1.6.7) (or, equivalently, it corresponds to the requirement
that ZM is nonnegative). If temperature is uniform in space, the Clausius-Duhem and
Clausius-Planck inequalities are equivalent, but they are not if grad T is different from
zero. If equation (1.6.6) is assumed to be universally true (i.e., Zr is required to always
be nonnegative), then the Clausius-Duhem inequality implies the Clausius-Planck
inequality. The Clausius-Duhem inequality by itself requires the sum ZM + Zr to be
nonnegative, without requiring the two addends to be separately nonnegative.
There has been an argument in the literature as to whether the temperature scale
appearing in the denominator of equations (1.5.4) and (1.5.5) should be the same; see,
e.g., M. E. Gurtin and W. O. Williams, Z. Angew. Math. Phys. 17, 626 (1966).
A mild global assumption, however, has been shown to be sufficient to imply that 33
indeed the same temperature scale should be used; see M. E. Gurtin, Z. Angew. Math.
Phys. 27, 775 (1976). First and
There is general consensus on the formulation of the first law, although in most Second Laws
textbooks it is not expressed in terms of rates as we have chosen to do. The validity of
equation (1.3.8) for rigid-body motions is discussed in any standard textbook on the
mechanics of rigid bodies.
Thermal engines are discussed in detail in most textbooks of thermodynamics for
mechanical engineers. A less known type of thermal engine is the rubber engine, which
makes use of the properties of entropic elasticity exhibited by rubbers (see Sections 5.4
and 8.1). Rubber engines are discussed by R. J. Farris, Polym. Eng. Sci. 17, 737 (1977).
A procedure which turns out to be illuminating, but is nonetheless seldom followed,
is to read the classics. Carnot's work, Reflexions sur la Puissance Motrice du Feu, was
published by Bachelier in Paris, 1824; an English translation can be found in W. F. Magie
(ed.), The Second Law o/Thermodynamics, Harper and Brothers, London (1899). The first
Clausius paper on thermodynamics, "Ueber die bewegende Kraft der Warme, und die
Gesetzte, welche sich daraus fur die Warmelehre selbst ableissen lassen," is in Ann. Phys.
(3) 19,368, 500 (1856). This is cited by J. W. Gibbs in his obituary of Clausius, Proc. Am.
Acad., Vol. xvi, 458 (1889). Clausius published in 1897 a book, Theorie Mechanique de la
Chaleur, Monceaux, Bruxelles, where his thermodynamic work is summarized. So did
Joule in 1872, Das Mechanische Warmeaequivalent, F. Wieweg, Braunschwig; his original
paper, "On Matter, Living Force, and Heat," appeared in the Manchester and Lancashire
General Advertiser, May 5-12, 1847.
Berthelot's contributions are well discussed in his two books, Thermochimie and
Calorimetrie Chimique, published by Gauthier-Villars in Paris in 1897 and 1905, respec-
tively.
Duhem's technical work is summarized in his books Mechanique Chimique,
A. Hermann, Paris (1897) and Energetique, Gauthier-Villars, Paris (1911), and his phil-
osophical views on science in La Theorie Physique, M. Riviere, Paris (1914).
Finally;Gibbs's fundamental thermodynamic work is presented in Trans. Conn. Acad.
2, 309 (1873); 2, 382 (1873); 3, 108 (1876); 3, 343 (1878).
Chapter Two

STATE AND EQUILIBRIUM

The considerations we have presented appear to better


conform to natural effects, and to provide the right
conception of the phenomena considered by evidencing
the restrictions under which the results can be applied.
M. Navier

Sylvester wrote Gilman that ... (Gibbs) at $5,000 would


be dirt cheap ... Until that time Gibbs had served Yale
as an honorary professor, unpaid. The powers there
seem to have been astonished to learn that some other
place might find worthy of pay such an eccentric fellow,
who just sat and thought and wrote equations . ...
C. A. Truesdell

NOTATION

A Free energy density kcalkg- I


a( Constitutive function for A kcalkg
b Body force field acceleration ms- 2
B Friction coefficient kgatmsm- 3
cv Constant volume specific heat kcal kg-I K- 1
cp Constant pressure specific heat kcal kg-I K- 1
D Rate of strain tensor S-I
F( Camot function
dF Elementary contact force kg m S-2
f( ) Constitutive function for p
r l Inverse of f( ) for volume m3 kg- 1
G Free enthalpy density kcal kg-I
g( Constitutive function for G kcal kg-I
H Enthalpy density kcal kg-I
h( Constitutive function for H kcal kg-I
K Kinetic energy kcalkg- I
K[ Equilibrium function for x* 35
36 L Any extensive property
Chapter Two l( Constitutive function for L
M Latent specific heat kcal m- 3
N Sensible specific heat kcal kg-I deg- I
p Total power kcal kg-I
IP' A process
p Pressure atm
g Heating rate kcal kg-I
IR( A transformation
Q Generalized pressure vector
r( Constitutive function for x
S Entropy density kcal kg-I K- I
s( Constitutive function for S kcal kg-I K- I
ds Elementary surface m2
T Temperature K
T Stress tensor kgm- I s- 2
Time
U Internal energy density kcal kg-I
u( Constitutive function for U kcal kg-I
V Specific volume m3 kg- 1
v Velocity ms- I
W Work kcal kg-I
w Rate of work kcal kg-I S-1
x Internal state variable
y Generalized volume vector
Z Dissipation rate kcal kg-I S-I
a Site
{3 Any temperature scale deg
() Affinity
IL Shear viscosity kgm- I S-I
ILl Bulk viscosity kgm- I S-I
cI> Density kgm- 3
u State
1 Unit tensor

Superscripts

* At equilibrium
Substantial time derivative
2.1. STATE AND SITE

It is this maximum of work which must be compared with the


heat transferred. When this is done it appears that there is in
fact ground for asserting, with Carnot, that it depends only on
the quantity of heat transferred and on the temperatures t and
T of the two bodies A and B, but not on the nature of the
substance by means of which the work is done. The maximum
has, namely, the property that by expending it as great a
quantity of heat can be transferred from the cold body B to the
hot body A as passes from A to B when it is produced.
R. J. E. Clausius, 1852

In the preceding chapter, the mathematical nature of the state has been left totally
unspecified. Without such a specification, no results can be obtained in addition
to those given in the previous chapter. However, it should be borne in mind that,
when a specific mathematical structure is assigned to the state, the applicability
of the resulting theory is restricted to classes of materials and! or phenomena for
which the assigned state is a realistic one.
Referring back to the discussion on different conceptual levels of constitutive
assumptions in Section 1.1, in this chapter the second lower level is considered,
i.e., assumptions are made about which variables determine the state, but no
assumptions are made on the specific functional form ofthe constitutive mappings.
Furthermore, the somewhat unusual concept of site is introduced, as distinguished
from state. This point is discussed in detail below; suffice here to say that the
concept of site is essential in the distinction between equilibrium and nonequili-
brium properties.
One needs to begin by assigning as simple as possible a mathematical
structure to the state. It should be borne in mind that, by assigning the state a
very simple mathematical structure, the resulting theory will be comparatively
simple, but it will be applicable only to a restricted class of materials and! or
phenomena. More complex cases will need a more complex formulation, and
the resulting theory is more difficult. Fortunately enough, a nontrivial theory
which is applicable to a rather large variety of pragmatically relevant problems
can be formulated on the basis of an extremely simple mathematical structure
for the state.
A theory which allows for long-range interactions would be hopelessly
complex, and thus it is assumed that the principle of local action holds, i.e., that
the values of the densities of the extensive quantities which are functions of state
(such as the internal energy density U), as well as the local values of any intensive
function of state (such as pressure), are uniquely determined by the local state 37
38 CT. Therefore, all equations will be developed in the local form. Should the local
state be constant throughout the whole body considered, the value of the extensive
Chapter Two
properties, such as U" would simply be equal to the product of U times the
mass of the body. If CT is not constant throughout the body, equation (1.2.1)
should be used.
The first step in assigning mathematical structure to CT is to choose a set of
"state variables." The simplest possible set is as follows. As was discussed in
Chapter 1, a nontrivial thermodynamic theory requires that deformation of bodies
be taken into account. The simplest possible deformation of a body is a change
of density; hence density, or (as commonly done) its inverse, i.e., the specific
volume V, is chosen as one of the state variables. A nontrivial theory requires
also at least one nonmechanica1 state variable, and therefore temperature T is
also included. Both V and T are local quantities which may well not be constant
in space.
Quantities V and T are legitimate state variables, since their values can in
principle be measured by making measurements only on the body itself (for
instance, velocity would not be a legitimate state variable, since its measurement
is based on reference to some other body which is arbitrarily regarded as being
fixed). However, V and T enjoy a special property: their values can be imposed
at will, and independently of each other, at anyone time (proof of this statement
is rather involved, and will be omitted here). Such variables are called external.
State variables which cannot be imposed at will independently of the others,
which may certainly be significant for a variety of systems, are called internal.
It is now possible to define the site a. The concept of site is related to the
consideration of equilibrium, and is best understood by considering how an
equilibrium experiment is to be performed. Suppose, for instance, that we wish
to determine the equilibrium pressure exhibited by a substance. We would run
an experiment where we keep constant as many variables as we can, and measure
the resulting pressure; we would expect the measured pressure to reach, after a
sufficiently long time has elasped, a constant value, which we interpret as the
equilibrium pressure corresponding to the values of the variables which we have
held fixed. That set ofvalues is what will be called the site. Now only external
variables can be assigned at will, and hence only external variables can contribute
to determine the site. However, there is a subtlety involved here, and one needs
to define the site as the set of external state variables which can be held fixed at
any value one wishes and for how long one wishes. At first sight, that seems to
include all external state variables, but that is not the case. Indeed consider, e.g.,
the rate of change of volume, V. Given a triplet V, T, V, its value can be
assigned at will at anyone time, and thus V is recognized as an external state
variable. However, the same triplet can be held constant in time only if V is
zero, and thus V, though being an external state variable, cannot be a member
of the set identified as the site. The question is that keeping the site constant in
time means that all the variables appearing in the site have to be imposed
independently of each other at all times, which is more restrictive a condition
than the one defining the external state variables, Le., that they can be imposed
independently of each other at anyone time-but not necessarily at all times.
The definition of site given above allows us immediately to define what we
mean by equilibrium: the condition which will be observed after the site has been
held constant (at whatever arbitrary value has been chosen) for a sufficiently long 39
time. It follows from this definition that all constitutive properties are, at e.quilibrium,
State and
junctions only of the site (an assumption is implicit here, i.e., that as the site is Equilibrium
held constant from some time onward, the conditions which will eventually be
observed are independent of what happened before the site was held constant).
However, it should be remembered that constitutive properties are defined also
for nonequilibrium conditions; and their values in general will be determined by
the state, not the site.
For the very simple system considered so far, the site is

a = {V, T} (1.1.1)

and thus all constitutive properties are, at least at equilibrium, unique functions
of volume and temperature. Under nonequilibrium conditions, they may have
different values, since the state may in general notcoincide with the site; assump-
tions need to be made about what the state may be. The simplest possible
assumption is that the state and the site coincide with each other:

u=a (2.1.2)

Systems for which equation (2.1.2) holds are called elastic. For elastic systems,
there is no room for a meaningful theory of equilibrium. In fact, any function
of state for an elastic system depends only on the site, and hence has the same
value independently of whether the site is being held constant (eqUilibrium) or
is changing in time.
A process lP(t) has already been defined as the function which maps time
into the local state:

u = 1P(t) (2.1.3)

A transformation lR(t) is the mapping of time into the site:

a = lR(t) (2.lA)

Processes and transformations coincide only in the case of elastic systems.


In the next section, a very simple specific problem is considered to show that a
theory of elastic systems is too restricted in scope.

Appendix

In this section, we have been somewhat cavalier about the question of which
transformations and which processes are admissible. The problems to be con-
sidered are as follows. First, we consider the case of transformations, and in
particular we consider the case where the site is as simple as in equations (2.1.1).
While it is certainly true that V values can be assigned at will at any given time
(it is only necessary to consider a cylinder equipped with a piston, which need
not be frictionless, to convince oneself that indeed one can assign V at will, given
40 arbitrarily large forces acting on the piston's shaft), there is a difficulty if one
wants to consider values of V which are discontinuous in time-so that V', may
Chapter Two
in fact be infinitely large. Values of V', approaching infinity would result in
conceptual problems in a viscous theory such as the one to be discussed in Section
2.3. Even more subtle conceptual problems arise if one considers the question
of which processes are to be regarded as admissible. These conceptual problems,
however, have no relevance to the analysis to be developed in the remainder of
this chapter.

2.2. PRESSURE AND WORK

"But of course you'd rather work-wouldn't you? Course you


would."
"What do you call work?"
"Why, ain't that work?"
"Well, maybe it is, and maybe it ain't. All I know is, it suits
Tom Sawyer. "
"Oh come, now, you don't mean to let on that you like it?"
"Like it? Well, I don't see why I oughtn't to like it. Does a boy
get a chance to whitewash a fence every day?"
That put the thing in a new light.
Mark Twain

Attention is still restricted to bodies for which the only relevant deformations
are changes in density, so that the only nonthermal state variable is V. This means
either that attention is restricted to materials on which it is possible to perform
a nonzero net mechanical work only by changing their density, or that attention
is restricted to phenomena where net work is done only by changing the density
although the material itself could absorb other types of net work should a wider
class of phenomena be considered.
At any point in a continuum body there may be internal stresses. Of these,
only the pressure p does any work in changes of density. Thus, within the
restrictions stated above, one may restrict attention to only pressure as far as
internal stresses are concerned. The instantaneous net rate of work per unit mass
done by pressure is given by

w = -pV' (2.2.1)

so that the local form of the first law becomes, for the case at hand,

U'= q-pV' (2.2.2)

The pressure is assumed to be a function of state, let a mapping f( . ) exist


such that

p = f(u) (2.2.3)
v 41
State and
Equilibrium

FIGURE 2.2.1. Two isothermal transformations.


At time t*, both volume and temperature are the
same in the two transformations. .

Now consider the special case where equation (2.1.1) holds, say the site a
is simply the ordered pair V, T. A pure gas may be a reasonably concrete example
of this. Now consider two distinct transformations of such a system, 1R1 and R2 ;
see Figure 2.2.1. For both transformations T is constant, so that IRI is an isothermal
expansion and 1R2 is an isothermal compression.
At time t*, the value of V is the same in 1R1 and 1R2 The following question
arises: Is the pressure also the same in both transformations at time t*? Such a
question is a physical one; in other words, we are asking whether a real system
will, or will not, exhibit the same pressure. If we do not care about describing
the fact that the two pressures may not be equal, we may proceed with an elastic
theory: If the system is elastic, the state at t* is the same in the two transformations,
and hence the pressure will be the same. However, if we wish to describe the
possibility that the pressure may not be the same, we have to allow for the
possibility that the state may not be the same at t* in the two transformations-i.e.,
we have to move up to a higher level of generality than an elastic theory makes
possible.
Physically, one would expect intuitively that pressure would be approxi-
mately the same provided both the expansion and the compression take place
slowly enough; if on the contrary they take place very rapidly, one would expect
the two pressures at t* not to coincide, and in fact the pressure being larger in
the compression than it is in the expansion. Furthermore, one would expect the
pressure difference to increase with increasing "viscosity" of the gas. As will be
shown later, these expectations can be shown to be deducible from the second law.
However, in order to do so one has to relax the assumption that the system
considered is elastic. As will be seen later, elastic systems do not allow any type
of irreversibility, and therefore the thermodynamic theory of elastic systems is
essentially a reversible theory.

Appendix

Let us consider an infinitesimal area vector ds within a body, and let dF be


the contact force acting through that area. The stress tensor T is defined as the
linear transformation of the area vector into the force vector:

dF=Tds (2.2.A.I)
42 The balance of linear momentum in the neighborhood of a point is
Chapter Two
<l>v = <l>b + div T (2.2.A.2)

where b is the field acceleration of the body force (usually gravity). Taking the
scalar product of equation (2.2.A.2) with the velocity vector v one obtains

K = <l>b v + vdivT (2.2.A.3)

The balance of angular momentum requires the stress tensor to be symmetric


in the absence of body couples (consideration of the case where there are no
body couples excludes magnetic phenomena). Hence the total rate at which work
is being done on the neighborhood of a material point per unit volume is calculated
as

P = <l>bv + div(T v) (2.2.A.4)

On subtracting equations (2.2.A.3) from (2.2.A.4), one obtains the net rate
of work per unit volume:

wi V = T:grad v (2.2.A.S)

where the: sign indicates the scalar product of two tensors. We note that the
body forces do not contribute to the net rate of work.
When the only internal stress is an isotropic pressure, then

T= -p1 (2.2.A.6)

where 1 is the identity tensor. Substitution into equation (2.2.A.5) yields

wlV = p divv (2.2.A.7)

Equation (2.2.1) is obtained from equation (2.2.A.7) by substitution of the


continuity equation

VI V = -divv (2.2.A.8)

The concept of an elastic system can easily be generalized to the case where
there are more than one external state variable to consider in addition to tem-
perature; for instance, in an elastic solid the strain at any given point is a
three-dimensional tensor. Let the site be the pair Y, T, where Y is an n-dimensional
vector called the generalized volume. Correspondingly, there will be an n-
dimensional vector Q, called the generalized pressure, such that the net rate of
work at a point is given by

w = -QY (2.2.A.9)

The system is elastic if the state is also the ordered pair Y, T, and such a
system can only undergo reversible processes.
2.3. VISCOSITY AND RELAXATION 43
State and
One does not insist often enough on the fact that the use of Equilibrium
partial derivatives of thermodynamic functions is based on the
(albeit implicit) assumption that such functions are defined also
for non-equilibrium states. Indeed, it is often impossible within
the set of equilibrium states to change one variable while
keeping the other ones constant. When a partial derivative
implies consideration of non-equilibrium states, its
experimental determination may be difficult, but, nevertheless,
it is a powerful conceptual tool.
R. Defay and I. Prigogine

If one wants a theory capable of predicting that the pressures at t* in the


two transformations in Figure 2.2.1 are different, the simplest possibility is to
allow for pressure to depend also on the rate of change of volume V':
p = f( v, T, V') (2.3.1)

In view of equation (2.2.3), this implies that the state is the ordered triplet
V, T, V':
u = v, T, V' (2.3.2)

However, V' cannot be included in the definition of the site, for the reasons
discussed in Section 2.1. It therefore follows that a system for which equation
(2.3.1) holds is not elastic, since state and site do not coincide; in particular, the
site is still given by equation (2.1.1), while the state also includes the rate of
change of one of the state variables. Systems for which the state is determined
also by the rate of change of some or all of the external state variables are called
viscous.
On the basis of the definition of equilibrium given in the previous section,
it is obvious that a viscous system is at equilibrium when the rate of change of
the external state variables is zero. For the system considered here, eqUilibrium
corresponds to V = O. Correspondingly, an eqUilibrium pressure p* (equilibrium
values will always be identified with an asterisk) is defined as
p* =f(V, T,O) =J*(v, T) (2.3.3)

We note that the equilibrium fuctionf*( . ) maps the site into the equilibrium
pressure. By definition, the eqUilibrium values of the functions of state depend
only on the site.
If one assumes that f( . ) is Taylor-series expandable at V = 0, one can write

p = p* + [af( V, T, 0)/0 V'] V' + O( v' 2 ) (2.3.4)

With this, it is evident that, provided V is sufficiently small (sufficiently slow


expansions or compressions), the actual pressure can be approximated by the
equilibrium pressure, which is indeed the same at t* in the two transformations
in Figure 2.2.1. The approximation p = p* is zero-order; in the first~order approxi-
mation the difference p - p* is linear in V':

p = p* - BV', (2.3.5)
44 where B is minus the partial derivative appearing in equation (2.3.4). Since one
expects pressure to be lower in expansion, B is expected to be positive.
Chapter Two
We now consider the isothermal transformation shown in Figure 2.3.1. The
final site at t> t\ coincides with the initial site at t < to. Both before to and after
t\ the system is at equilibrium, but it is not in the time interval from to to t\. The
total network done in that time interval is (all integrals in this section are intended
as definite integrals over the whole transformation)

w=-f pV'dt (2.3.6)

Should the system be elastic, one would have

w= - f f( V, T) V' dt = - f f( V, T) dV = 0 (2.3.7)

The physical meaning is that an elastic system does as much work on the
surroundings during the expansion part of the transformation as it absorbs during
the compression part: the mechanical behavior of an elastic system is entirely
reversible.
Conversely, if equation (2.3.5) holds true, the total net work is

w= - f f*( V, T) V' dt + B f V' 2 dt =B f V' 2 dt (2.3.8)

and although the final site coincides with the initial one a finite net work has
been done, and in fact a positive one if as expected B > O.
In viscous systems, only the instantaneous rate of change of the external
state variables affects the state: one could say that the external state variables'
values need to be known only at, and immediately before, the actual instant of
observation. In fact, there are materials (notably polymers) which have rather
long relaxation times, so that also the values at instants in time significantly
remote from the instant of observation affect the state. Such systems are called
relaxationaL The thermodynamic theory of relaxational systems is quite complex,
and only the essential features of it will be discussed in Chapter 5.

~
,,
,
'''
' I '
,, '''
,, '
,
I,
I
''
I'
,, ''
,, I' FIGURE 2.3.1. An isothermal transformation
t, with the final volume equal to the initial volume.
Viscous systems allow irreversible processes: for the transformation in Figure 45
2.3.1, the total net work done is dissipated, since the system at the end of the State and
transformation is in the same state as at the beginning. Also, viscous systems Equilibrium
allow a meaningful distinction to be made between equilibrium and nonequili-
brium values of the quantities of interest. However, the type of irreversibility and
of equilibrium considered in this section are not the only possible ones, and in
fact in the field of chemical engineering the pragmatically more important ones
are those which arise in systems with internal state variables.

Appendix

The discussion of viscosity effects in this section should, rigorously, proceed


as follows. The stress tensor in a linearly viscous fluid is given by

(2.3.A.l)

where D is the symmetric part of the velocity gradient tensor, IL is the shear
viscosity, and ILl is the bulk viscosity. Fluids obeying equation (2.3.A.l) are said
to be Newtonian fluids. If the only deformation is compression (or expansion),
D is given by

D = 1(div v)1; tr D = divv (2.3.A.2)

and hence the stress is isotropic:

T = [-p* + (-ILl + 21L/3) div v]1 (2.3.A.3)

The quantity in square brackets is thus recognized as minus the actual


pressure, and it coincides with the equilibrium pressure only if the Stokes
hypothesis holds, i.e., if

(2.3.A.4)

The Stokes hypothesis is just that, a hypothesis, and it need not necessarily
hold true-neither does, for that matter, equation (2.3.A.1) necessarily hold true.
If equation (2.3.A.1) holds, the constant B in equation (2.3.5) is, in fact, identified
with

(2.3.A.5)

It follows that, if the Stokes hypothesis holds, there is no dissipation in


expansion or compression. We note, however, that even if the Stokes hypothesis
holds, actual variations of density are generally not accomplished by pure com-
pression or expansion, but some shear is involved, and therefore dissipation will
in fact occur. Indeed, consider the simple case of a gas, even an ideal gas, flowing
through a tube. There will be a finite pressure drop, and hence the exit gas is at
the same temperature but at a lower pressure than the entering gas; some energy
has been dissipated. The gas has expanded, but not by a pure expansion, since
shear is certain to occur within the tube.
46 The second law imposes two requirements on the viscosity coefficients. The
first is the one already considered, i.e., that B should be nonnegative: a positive
Chapter Two
value of B makes phenomena of pure compression or expansion irreversible.
The second requirement is that f.L should be nonnegative: if f.L is positive,
constant-density deformations of the fluid are irreversible.
In the Maxwellian theory of gases, which at equilibrium yields the constitutive
equation of the ideal gas, it is possible to deduce equations (2.3.A.l) as a small
deformation rate expansion-but not in general, since the Maxwellian gas is not
a Newtonian fluid. However, at the same level of approximation the Maxwell
theory implies that the Stokes hypothesis does indeed hold-so that pure
expansions or compressions of a Maxwellian gas are indeed nondissipative. In
fact, it can be proved in general that pure expansions or compressions are
nondissipative phenomena in a Maxwellian gas (the distribution function is the
equilibrium one in such deformations). It should, however, be borne in mind
that real gases (let alone liquids) need not be adequately represented by the
Maxwell theory.

2.4. THE SECOND LAW AND EXTERNAL SYSTEMS

In the popular treatise, whatever shreds of science are allowed


to appear, are exhibited in an exceedingly diffuse and
attenuated form, apparently with the hope that the mental
faculties of the reader, though they would reject any stronger
food, may insensibly become saturated with scientific
phraseology, provided it is diluted with a sufficient quantity of
more familiar language. In this way, by simple reading, the
student may become possessed of the phrases of the science
without having been put to the trouble of thinking a single
thought about it. The loss implied in such an acquisition can be
estimated only by those who have been compelled to unlearn a
science so that they might at length begin to learn it.
James C. Maxwell

In this section, the standard methodology for deriving consequences of the


second law is introduced. Since the methodology to be used is a somewhat
unusual one, it is important that it should be thoroughly understood. The subtlety
arises, of course, because the second law is an inequality, and therefore only
other inequalities can be deduced from it by a purely algebraic procedure.
However, as will be seen below, one can in fact deduce from the second law,
when coupled with appropriate constitutive assumptions, consequences which
are equations, not inequalities; the procedure is more a logical than an algebraic
one. The validity of the deduced equations is strictly conditional on the validity
of the constitutive assumptions made.
Attention in this section is restricted to external systems which are spatially
isothermal, so that the local form of the second law is equation (1.7.9), namely

w-A-ST"~O (1.4.1)
Since both A and S are functions of state, for a viscous system described 47
by equation (2.3.1) two mappings a() and s() exist such that State and
Equilibrium
A = a(V, T, V) (2.4.2)

and
S = s( V, T, V) (2.4.3)

in addition to the mapping f( ) in equation (2.3.1).


Substitution of equation (2.2.1) into equation (2.4.1), and differentiation
with respect to time, yields

[aajaV + p] V + [aajaT + S]T + [aa/av'] V":s 0 (2.4.4)

It may be worth noting that, since the mapping a() is uniquely identified
by equation (2.4.2), the partial derivatives appearing in equation (2.4.4) are
unequivocal: for instance, aaj a V is the partial derivative of the Helmholtz free
energy density with respect to volume when temperature and rate of change of
volume are held constant. In most books on thermodynamics, partial derivatives
are always written with subscripts identifying the quantities being held constant.
This is due to the fact that the same symbol is used for a mapping and for its
value; since different mappings may have the same value, a partial derivative
without subscripts would be equivocal. This point will be further clarified in the
section on baric derivatives.
Equation (2.4.4) must hold true for every conceivable transformation, say
for any conceivable set of values V, T, T", V', and V". (Since V and Tare
external state variables, their values, as well as the values of all their time
derivatives, can in principle be assigned at will at anyone instant of time-though
not at all times, of course.) As shown below, this requirement implies that the
constitutive mappings are not independent of each other. The second law imposes
restrictions on the form of the constitutive mappings, not on conceivable transfor-
mations.
Suppose that, at some particular instant in time, V, V', T, and T" have been
assigned, so that the value of every term in equation (2.4.4) is assigned, except
V". Since equation (2.4.4) must hold for arbitrary values of V", no matter what
their sign or magnitude, the only possibility is for the coefficient of V to be zero:

aajav'=o (2.4.5)

Equation (2.4.5) should not be understood to imply that the Helmholtz free
energy density depends only on the site in general; in fact, for many systems it
does not. Equation (2.4.5) implies that, if the state is indeed the ordered triplet
V, T, V', then A does not depend on V'.
The second law is now collapsed to

[aajaV + plY' + [aa/aT + S]T":s 0 (2.4.6)


48 We consider the case where V, T, and V have been assigned, so that the
only nonfixed term in equation (2.4.6) is T. Again, equation (2.4.6) must hold
Chapter Two
for arbitrary values of T, and hence the coefficient of T must be zero:

s = -aa/aT (2.4.7)

Combination of equations (2.4.5) and (2.4.7) yields


as/avo = 0 (2.4.8)

It should be noted that equations (2.4.5) and (2.4.8) are restrictions imposed
on the constitutive mappings a( . ) and s( . ), while equation (2.4.7) is a relationship
between these two mappings: Once a() has been assigned, the form of s() is
determined-but not vice versa. This is worded by saying that the Helmholtz free
energy is a potential for entropy.
The second law has now collapsed to
[aa/av + ply s 0 (2.4.9)

We note, however, that now the same argument cannot be duplicated, because
when V is changed, so is p, and one thus cannot fix the value of the square
bracket and yet allow for arbitrary values of V. Now suppose that V and T
have been assigned. Equation (2.4.9) implies that

when V> 0, aa/av+pso (2.4.10)

when V < 0, aa/av+p:20 (2.4.11)

Let us consider the quantity aa / aV + p at assigned values of V and T, so it


depends only on the value of V; see Figure 2.4.1. Equations (2.4.10) and (2.4.11)
imply that the plot is restricted to the odd quadrants; it follows that, if the f( . )
mapping is sufficiently well behaved near V = 0, i.e., near equilibrium, the curve
goes through the origin, and hence
when V = 0, p = p* = -aa/av"= f*(V, T) (2.4.12)

i.e., the Helmholtz free energy is also potential for the eqUilibrium pressure.

aa .p
av
fixed Y and T

FIGURE 2A.I. 8a/8V + p vs. V at fixed V and


T. The curve is restricted to the odd quadrants by
the second law.
Finally, equations (2.4.10) and (2.4.11) imply that the difference p - p* never 49
has the same sign as V'. This in turn means that the pressure in expansion is
State and
always less than that in compression (at equal instantaneous V and T), as Equilibrium
physically expected. In particular, if equation (2.3.5) holds, the viscosity B must
be nonnegative. This shows how the methodology introduced above produces
consequences of the second law which are those expected on the basis of physical
intuition.
If one had considered an elastic system, say a system where (T = v, T, one
would have concluded that p = -aa/av always, and not only at equilibrium;
indeed, for such systems the distinction between equilibrium and nonequilibrium
is irrelevant. Furthermore, the second law would have collapsed to 0:5 0, i.e.,
one concludes that elastic systems can only undergo reversible processes. Since
neither equilibrium nor irreversibility arises in the thermodynamic theory of
elastic systems, such a theory is too restricted in scope to be of practical utility
except in special cases.
Results such as in equation (2.4.12) are called thermostatic results, since
they hold only at equilibrium. It is noteworthy that, had one included T: in the
state, the Helmholtz free energy would have been a potential only for the
equilibrium entropy, i.e., equation (2.4.7) would have been only a thermostatic
result.
A final important consideration is that the results obtained are based on
some initial hypotheses of smoothness for the constitutive mappings, and in
particular assumptions of smoothness near equilibrium. No thermodynamic
theory can be developed without such assumptions, and that is what makes the
thermodynamic theory for relaxational systems very complex. For such systems,
the state includes some functions, and hence the constitutive mappings are
functionals, for which smoothness assumptions can be made only after an
appropriate topology of the space of possible states has been chosen.

2.S. INTERNAL SYSTEMS

The impression [that my dissertational] made upon the public


of physics at that time was naught. Of my teachers at the
University, as I know precisely from conversations with them,
not one understood its contents. ... Even among physicists
more nearly concerned with the subject itself I found no
interest, let alone approval. Helmholtz probably did not read
the paper at all; Kirchoff rejected its contents expressly with the
remark that the concept of entropy, the magnitude of which was
measurable only by means of a reversible process and hence
was definable only for such, could not legitimately be applied to
irreversible processes.
M. Planck

In the preceding section, it has been seen that a crucial role is played by
the fact that T and V are external state variables. The fact that their values, as
well as the values of their time derivatives, can be assigned at will and indepen.
dently is crucial to the logic of deducing consequences of the second law. The
50 type of irreversibility discussed in the preceding section is somewhat special,
since it can be made as small as one wishes by having sufficiently slow transforma-
Chapter Two
tions.
In this section, internal systems are considered. A simple concrete example
is obtained by considering a chemically reacting mixture in a variable volume
box. While volume and temperature can be imposed arbitrarily and independently
of each other, the reaction goes on at a rate determined by its intrinsic kinetics,
and therefore the composition of the gas cannot be assigned at will and indepen-
dently. Given an initial composition, the composition at any subsequent time is
identified by one single scalar parameter, the degree of advancement of the
reaction, x. The quantity x is the internal state variable, and hence the simplest
structure for the state is

u={V,T,x} (2.5.1)

Of course, the site is still {V, T}.


Three fundamental thermodynamic results can be obtained by considering
such a system: the irreversibility con~ected with the evolution in time of the
internal state variable (i.e., for the concrete example, the rate of reaction); the
difference between this type of irreversibility and the viscous one discussed in
the preceding section; and the character of eqUilibrium in an internal system.
The constititive mappings for the system considered are

A = a(V, T, x) (2.5.2)

S=s(V,T,x) (2.S.3)

P = f(V, T, x) (2.5.4)

In addition, one needs an equation for the rate of change of the internal state
variable, x, say for the concrete example one would need to know the kinetics
of the chemical reaction. Indeed, for all internal systems the rate of 'change of
the internal state variables must itself be a function of state: x being internal, its
value cannot be imposed from outside, and hence at anyone time it will change
at a rate governed internally, i.e., at a rate which is a function of state. For the
case at hand

x=r(v,T,x) (2.5.5)

If one performs the same algebra as that leading to equation (2.4.4), one
obtains

[aa/av + p] V + [aa/aT + S]T" + (aA/ax)x:::; 0 (2.5.6)

One can now perform the same analysis as for equation (2.4.4) to obtain
the following results:

p = -aa/av (2.5.7)

S = -aa/aT (2.5.8)

Z = -(aa/ax)x 2: 0 (2.5.9)
The Helmholtz free energy is again a potential for both pressure and entropy, 51
but the second law will in general be fulfilled as an inequality, say the dissipation
State and
rate Z will in general be positive. In fact, except at equilibrium, given a state V, EqUilibrium
T, x, neither aa/ax nor x' will be zero. Furthermore, while the dissipation rate
due to viscosity can be made as small as one wishes by imposing a sufficiently
small value of V', the magnitude of Z as given by equation (2.5.9) cannot be
controlled, since x' is not controllable. In particular, when chemical reactions
occur, energy is dissipated. [There is an important exception to this statement,
namely, reactions taking place in an electrochemical cell, where the reaction rate
depends also on the imposed electrical potential difference, which is an external
state variable.]
The above considerations deserve some additional comment. Suppose that
a viscous gas is brought isothermally from some initial volume VI to some final
volume V2 The total dissipation during the transformation is given by

f Z dt =B f V 2 dt (2.5.10)

in the first-order approximation. By slowing down the process, the interval of


integration grows linearly, but the integrand decreases quadratically, and there-
fore the total dissipation can be made as small as one wishes. On the contrary,
if an internal system evolves from, say, XI to X2, the total dissipation will have
some finite value which cannot be made smaller.
The quantity aa/ax = 8 is called the "affinity" of the system, and is of course
a function of state. Equation (2.5.9) requires the affinity never to have the same
sign as the rate x'. Hence, again under sufficient assumptions of smoothness, 8
will be zero when x' is zero. The condition where x' is zero is an equilibrium
condition. Suppose that V and T are assigned, and kept constant in time. The
corresponding equilibrium values of x is the one for which x' becomes zero, say
it is the solution of the following equation:

0= reV, T, x*) (2.5.11)

The solution of equation (2.5.11) yields the "equilibrium function"

x* = K[V, T] (2.5.12)

and, in view of the above argument,

8(V, T, K[V, T)) = 8*(V, T) = 0 (2.5.13)

We note also that, if V and T are held constant, equations (2.5.6) and (2.5.9)
imply that A can never increase; since A' is zero at an eqUilibrium point, the
value of A at an equilibrium state is the minimum compatible with the assigned
values of V and T. The theory of equilibrium for internal systems is essentially
concerned with the location of the equilibrium functions from the requirement
that free energy is minimized.
52 Now consider an internal system in a nonequilibrium condition, so that x
is different from x*. If the site is held constant, x will change in time at whatever
Chapter Two
rate is implied by equation (2.5.5) for such a condition; hence, the state will
change in time, and possibly quite slowly, so that it may take quite some time
before equilibrium is reached. It is now obvious why our definition of equilibrium
was the condition observed after the site has been held constant in time for a
sufficiently long time. In a viscous system, equilibrium is reached instantaneously
as soon as the site starts being held constant; that is not the case for internal
systems. In some cases, the rate of evolution toward equilibrium may be so slow
that, to all practical purposes, one could say that nothing changes in time, and
yet the system is not at equilibrium.

2.6. BARIC DERIVATIVES

"The round square does not exist" had always been a difficult
proposition; for it was natural to ask "What is it that does not
exist?" and any possible answer had seemed to imply that, in
some sense, there is such an object as the round square, though
this object has the odd-property of not existing.
B. Russell

In this section, attention is restricted to the same system as discussed in the


previous section. However, the results obtained are easily extended to more
general systems.
It is useful to introduce two additional functions of state which are also
absolutely additive functions of mass: enthalpy H and free enthalpy (sometimes
called Gibbs free energy) G. The definitions are

H= U+pV (2.6.1)

and
G=H-TS=A+pV (2.6.2)

so that, in addition to the mappings in equations (2.5.2)-(2.5.4), three additional


ones are considered
U=u(V,T,x) (2.6.3)

H=h(V,T,x) (2.6.4)

G = g(V, T, x) (2.6.5)

The mappings are not independent, and in fact the free-energy one is a
potential for all the functions of state except x. Equations (2.5.7) and (2.5.8)
give s( . ) and f( . ) directly as derivatives of a( . ), 'while from the definitions one
obtains immediately
u() = a() - T iJa/iJT (2.6.6)

h() = a() - T iJa/iJT - V iJa/OV (2.6.7)

g(.) = a() - V iJa/iJV (2.6.8)


In many practical problems, one often knows the value of pressure but not 53
necessarily that of volume, and hence one would prefer to have p rather than V
State and
as an independent variable. Now suppose that the function f( . ) is invertible for Equilibrium
volume, assume a function f-l( . ) exists such that

(2.6.9)

This could then be substituted in the constitutive mappings to obtain new


functions which map the triplet p, T, x into the values of the function of state.
In the classical literature on thermodynamics, the new mappings are given the
same symbols as the previous ones (in fact, the same symbols as their values),
and hence in order to keep track of which mapping is being considered, partial
derivatives have to include subscripts identifying the variables which are being
held constant. Of course, if any of the original constitutive functions is invertible
for volume, the physical property which has its value could be used as an
independent variable instead of V; similar remarks apply to invertibility with
respect to temperature. One ends up with an enormous collection of constitutive
functions, and even more partial derivations, and an almost infinite collection
of relationships between such partial derivatives.
However, the assumption of invertibility is not to be taken lightly. Consider
a very simple case, water at 100C; the internal state variable is the fraction of
vapor phase which may form. Two important curves are drawn in Figure 2.6.1.
The thick line 0257 is the equilibrium pressure, say x = x*. In fact, x* = 0 along
02, x* = 1 along 57, and x* takes values between 0 and 1 along 25. Line 0123
corresponds to x = 0, i.e., to liquid water. Along 23, the water is superheated,
assume it is not at equilibrium; such conditions may be realized in the laboratory.
Line 34 represents an imaginary condition of one-phase H 2 0 which can never
be realized in actual fact, because all points on that line correspond to absolutely
unstable conditions. Finally, line 45 represents an undercooled vapor, and is
again a locus of physically possible non equilibrium conditions. The point to be
made is that there is a whole region between pressures P14 and P36 where the
f( ) function is not invertible for volume. Even the f*( ) function is not
invertible along line 25.
However, we note that the curve for x = 0, 0123, possesses a finite derivative
everywhere except at point 3, and the same applies to the curve for x = 1,4567.

FIGURE 2.6.1. Isothermal pressure-volume


curve through a phase transition. The smooth
curve is the hypothetical one where the system is
always in a one-phase condition. v
54 In fact, at fixed T and x, the /( ) function is locally invertible almost everywhere.
Therefore, the inversion called for in equation (2.6.9) can be done, within the
Chapter Two
restriction of a local inversion called for in this paragraph.
Given differential variations dT and dx, the corresponding dV variation
required in order to keep pressure constant (while staying on the same branch
of the curve) is calculated by differentiating equation (2.5.4) and requiring dp to
be zero:

(af/av) dV = -[(af/aT) dT + (a//ax) dx] (2.6.10)

(which shows that a differential volume change is obtained provided a//av is


finite). Given a du identified by dT, dx, and the corresponding dV calculated
from equation (2.6.9), the corresponding dp will be zero (in the restricted form
discussed before). Given any function of state L with values delivered by a
constitutive mapping 1('), the variation dL corresponding to such du can be
interpreted as a variation at constant pressure, and is given by

dL = [altaT - (al/a V)(a//a T)/(a//a V)] dT


+ [ai/ax - (al/aV)(af/ax)/(a//aV)] dx (2.6.11)

and therefore the two baric derivatives (i.e., within the stated restrictions, partial
derivatives at constant pressure) are as follows:

81/ aT = ai/aT - (alta V) (a//aT)/(a//a V) (2.6.12)

and
131/13x = ai/ax - (al/aV)(af/ax)/(a//av) (2.6.13)

By straightforward substitution one obtains a very important result, i.e., that


the baric derivatives of free enthalpy are equal to the partial derivatives of free
energy, for instance

13g/13 = aa/a (2.6.4)

The equilibrium theory of internal systems (mostly chemically reacting


systems) is mostly developed by minimizing free enthalpy at assigned values of
p and T, rather than minimizing free energy at assigned values of V and T.
A final baric derivative needs to be defined, i.e., 8113p. Indeed, within the
stated restrictions, substitution of equation (2.6.9) into any of the original constitu-
tive mappings would give functions mapping the triplet p, T, x into the value of
L. Such functions would in general possess a partial derivative with respect to
p: this is the baric derivative with respect to pressure. Suppose T and x are
assigned, while V is changed by a differential amount dV, resulting, within the
stated restrictions, in differential changes of pressure and L:

dp = (a//aY) dV (2.6.15)

dL = (alIa V) dV (2.6.16)
Therefore 55
State and
fJl/fJp = (al/aY)/(aj/aV) (2.6.17) Equilibrium

2.7. THE MAXWELL RELATIONS

"Isn't one man as good as another?", demanded a Socialist


orator, addressing an Irish mob. "Av coorse he is" was the
eager response "and a great deal betther. "
This not only contradicts the previous assertion, but also
contradicts itself. ... But self-contradiction, in a proposition, is
not an attribute that would for a moment discredit it in the
Emerald Isle.
Lewis Carroll

The discussion in this section is again restricted to a system of the type


discussed in the previous sections, but the results are easily generalized.
In addition to the three (or more) state variables V, T, and x, there are six
functions of state to be considered: p, S, A, U, H, and O. Under sufficient
assumptions of invertibility for the constitutive functions which map the state
into these quantities, the values of any three of the nine variables could be
regarded as identifying the state: this leads to 84 different sets which could be
interpreted as identifying the state. Given any such set, anyone of the other six
variables could be considered as assigned by a function of the chosen state; thus
504 different constitutive functions can be identified. Each of these functions has
three partial derivatives, for a total of 1512 different partial derivatives.
Let us consider a differential change of state {dV, dT, dx}. The corresponding
change of free energy is calculated from equations (2.5.7) and (2.5.8) and the
definition of affinity:

dA = -pdV - SdT+ 8dx (2.7.1)

and will in general be differential, since p, S, and 8 have finite values. The
assumptions of invertibility for the constitutive functions imply that, for the same
differential change of state, the changes of all other functions of state will also
be differentials. One can therefore simply substitute the definitions to obtain the
following three additional equations:

dU = -pdV + TdS + 8dx (2.7.2)

dH = Vdp+ TdS+ 8dx (2.7.3)

dO = V dp - S dT + 8 dx (2.7.4)

Equations (2.7.1}-(2.7.4) are called the Maxwell relations.


Relationships between the partial derivatives referred to before are obtained
by applying standard rules of differentiation to the definitions and the Maxwell
relations. Which particular function is being considered is identified by the
following nomenclature. The quantity which is being 'differentiated is the value
56 of the function considered. The three variables regarded as identifying the state
are the variables with respect to which the derivative is taken, and those appearing
Chapter Two
as subscripts. Therefore, e.g., the two derivatives

and (2.7.5)

are derivatives of the functions mapping {H, p, x} into G and {T, p, x} into U.
Since the most common case is where one wants to regard {p, T, x} as the
state, we can simplfy the notation by using subscripts only in those few cases
where triplets other than either {V, T, x} or {p, T, x} are intended. We can also
drop the distinction between the symbol identifying a thermodynamic quantity
and the one identifying its constitutive mapping.
Relationships between partial derivatives can be obtained from the Maxwell
relations by a variety of shortcut methods. Some of thesee are discussed below.

1. By direct inspection, e.g., from equations (2.7.1) and (2.7.4):

-s = 8G/8T = aA/aT (2.7.6)

Equation (2.6.14) can be obtained directly by this method.


2. By setting two of the differentials on the right-hand side to zero. As an
example, setting dp and dx to zero in the Maxwell relation for H yields

8H/8T = T8S/8T (2.7.7)

3. By differentiation and substitution of the definitions, e.g., from equation


(2.7.3) and the definition of enthalpy:

aH/OT= TaS/aT+ Vap/OT=aU/aT+ Yap/aT (2.7.8)

Hence

Tas/aT = aU/aT (2.7.9)

4. By using the commutativity of second derivatives,

ap/aT = -a2 A/Ov aT = as/av (2.7.10)

Two partial derivatives playa very important role and are therefore assigned
a special symbol: the "specific heat at constant volume," cv, and the "specific
heat at constant pressure," Cp,

Cv = aU/aT (2.7.11)

Cp = 8H/8T (2.7.12)

From equation (2.7.9) one therefore obtains

as/aT = cv/T (2.7.13)


and with perfectly analogous algebra 57
as/aT = cp/T (2.7.14) State and
Equilibrium

Appendix

There is no general consensus on the definition of the specific heats, and


equations (2.7.11) and (2.7.12) are not universally adopted as definitions. It is
important to realize that, while equation (2.7.11) is generally applicable, there is
some subtlety involved with equation (2.7.12) if phase changes are involved. In
fact, consider, e.g., boiling of water at 1 atm pressure. As boiling occurs, volume,
internal energy, and enthalpy change, but temperature and pressure do not. The
enthalpy vs. temperature plot at constant pressure has the shape shown in Figure
2.7.1; the slope to the left of the boiling point is the constant pressure specific
heat of the liquid, the slope to the right is that of the gas. At the boiling point
itself the curve is not differentiable, and hence Cp is not defined at that point.
In the nineteenth century literature (see in this regard also Section 2.8) the
specific heat concept was generally used in a different form. For a system the
state of which is V, T, the instantaneous rate of heating q can be written as
q=MV+NT (2.7.A.l)

and the coefficients M and N were called the latent and the sensible heat capacity.
The coefficient N is immediately recognized as the constant volume specific heat.
Substitution of the first law in the form of equation (2.2.2) yields
N=p+aUfaV (2.7.A.2)

a quantity which would not be called latent heat today.


The question of the invertibility of the f( ) function does not arise solely
in the case of phase changes. Indeed, consider a system (even an elastic one) in
a one-phase condition, so that ap/av < 0 always, and invertibility with respect
to volume is guaranteed. However, there is still the problem of invertibility with
respect to temperature, as in the question of whether one could use p and V,
rather than V and T, as the independent variables. The question is of course
related to whether ap/aT has always the same sign. Now consider the following
equation:
av/ aT = -(apjaT)/(ap/aV) (2.7.A.3)

FIGURE 2.7.1. Isobaric enthalpy-temperature


plot through a phase transition.
58 which guarantees that, within the stated restrictions, ap/aT and 13V/13T always
have the same sign.
Chapter Two
For most materials under the majority of conditions, 13V/ 13T > 0, i.e., the
material expands with increasing temperature at constant pressure. However,
liquid water (to give the best known example) has a minimum of density at 4C,
and thus water, at 1 atm, exhibits a negative value of both 13V/13T and ap/aT
between 0 and 4C. Thus in this region the f( ) function is not invertible for
temperature, and the pair p, V cannot be used as the pair identifying the state.
This has important implications when adiabatic transformations are con-
sidered. For an elastic system the state of which is V, T, an adiabatic transforma-
tion is one where
(p + aU/aV)v' + cvT" = 0 (2.7.A.4)

which guarantees that the heating rate is zero. A little algebra with the Maxwell
relations shows that
p+au/av= Tap/aT (2.7.A.5)

Thus in an adiabatic transformation one has

'/T" = ap _ cvap/av (2.7.A.6)


p aT Tap/aT

which guarantees that p'/T' has the same sign as ap/aT. Thus in the region of
anomalous behavior of water the rates of change of pressure and temperature
have opposite signs, and this can be shown to result in major anomalies of Carnot
cycles.

2.S. ABSOLUTE TEMPERATURE

Modern mysticism has been much exercised in respect to the


terms sensible heat and latent heat, whether in decrying them,
or in continuing to use them, but with aggravating haziness,
instead of the clear wrongness of the old doctrine.
Lord Kelvin

The way of measuring temperature has obviously a degree of arbitrariness.


Suppose one establishes some arbitrary way of measuring temperatures, say a
temperature scale {J, with the provision that "hotter" bodies are assigned a larger
value of {J. Any two such temperature scales are related to each other by a
continuous function which is monotonously increasing:

d{Jd d{J2 > 0 (2.8.1)

Of all such scales, only one may be such that equations (1.5.4) and (1.5.5)
hold true, namely, the absolute temperature scale. The fact that this absolute
scale happens to coincide with the scale obtained by using an ideal-gas ther-
mometer is largely coincidental, and it has been both helpful in the development
of thermodynamic theory, and a stumbling block to it, as discussed below.
The concept of an absolute temperature scale was introduced by Lord Kelvin 59
in 1852. His argument was essentially as follows. Consider a substance which
State and
undergoes a differential change of state dV, dT. The argument happens to be Equilibrium
correct even if the state has more structure than being simply the ordered pair
V, T, provided the following conditions are satisfied: (1) all other external state
variables are being held constant, (2) there are no internal state variables, and
(3) work can in fact be expressed by equation (2.2.1). If an absolute temperature
scale has not yet been established, the change of state has to be identified by dV,
d~, ~ being whatever arbitrary temperature scale has been chosen. The work
done on the surroundings is p dV, and the amount of heat received by the
surrounding is

dq = N d{3 +M dV (2.8.2)

where N and M were called the sensible anq latent specific heats. In modem
terminology, the quantities Nand M are identified with

N = (iJU/iJ~)v = cvdT/d~ (2.8.3)

and
M = p + (iJU/OV)fj = p + (iJU/iJVh (2.8.4)

Lord Kelvin's argument was based on the consideration of a closed cycle


in the V, ~ plane; see Figure 2.8.1. He argued that, since for such a closed cycle
the total exchange of energy with the surroundings must be zero ("The total
external effect must, according to Proposition I, amount to nothing"), the func-
tions M and N must satisfy the condition

iJN/OV = iJ(M - p)/iJ~ (2.8.5)

which of course we can identify as the proposition that

(2.8.6)

FIGURE 2.8.1. Closed cycle in the V -p plane.


o
60 The second part of Lord Kelvin's argument was based on his Proposition
Chapter Two
II, which basically was the argument of Carnot that the maximum work done
by a heat engine working between two given temperatures f31 and f32 was indepen-
dent of the nature of the substance used in the engine. This can be shown to
require that

F(f3) ap/ af3 = M (2.8.7)

where F(f3) is a function which is the same for all substances. The absoJute
temperature scale was thus established as being T = F(f3). Indeed, if f3 is
identified with T, equation (2.8.7) becomes, in modern terminology,

Tap/aT=p+(au/avh= TaS(aVh (2.8.8)

which is of course equation (2.7.10). As was discussed before, the fact that T
happens to coincide with the temperature measured by an ideal gas thermometer
made matters somewhat easier for the pioneers; however, in an ideal gas au/ aV
is zero, and hence the distinction between the quantity M and pressure itself was
a subtle one to identify.

2.9. SUMMARY OF CONCEPTUAL ISSUES

Happy is the man that findeth wisdom, and the man that
getteth understanding.
For the merchandise of it is better than the merchandise of
silver, and the gain thereof than fine gold.
Proverbs 3, 13-14

The distinction between site and state may, at first sight, look both artificial
and somewhat obscure; it is useful to return to this point after having discussed
some of the implications. It is easiest to discuss the matter by considering a
simple reacting mixture, for which the only difference between site and state is
that the former is the ordered pair V, T, but the latter includes an internal state
variable x; however, all the points made below are more generally applicable.
The site is the set of those variables which (1) can be held constant at any
value one may wish (which is what is done in an equilibrium experiment), and
(b) are tentatively supposed to be the only ones which matter in determining the
equilibrium behavior. The pair V, T satisfies condition (1), but is not guaranteed
to satisfy condition (2), as the following two simple counterexamples show.
Let us consider an elastic spring. It is clearly possible to extend the spring,
without changing either its volume or its temperature. However, if one measures
the force needed to keep the spring elongated at some length I, which is obviously
a function of state, one reaches the conclusion that it is not constant when V
and T are held constant, but its value depends on 1 as well. One can now ask
oneself whether it is possible to keep 1 constant independently at any preassigned
value while keeping V and T constant as well; the answer is obviously yes. For
this system, one thus comes to the conclusion that the site is not simply the
ordered pair V; T, but it must be extended to at least the ordered triplet V; T, I; 61
correspondingly, a new function of state needs to be considered, i.e., the axial State and
force acting on the spring. The latter will eventually attain an equilibrium value Equilibrium
which depends only on the site when the latter is held constant for a sufficiently
long time, but of course there is no guarantee that it will do so instantly as soon
as the site starts being held constant-there may well be some delay.
Another useful example is that of an electrochemical cell, like the battery
of an automobile. One can certainly maintain constant V and T and yet measure
a response which depends on how the external circuit is set: if current is drawn
from the battery, the electrical potential difference measurable between the two
electrodes will not be the same as that measurable while the battery is being
recharged. One thus concludes that it is necessary to regard the applied electrical
potential as an additional variable which determines the site. The thermodynamic
theory of electrochemical cells will be discussed in a later chapter; here the
example is brought forward only as further confirmation of the fact that the pair
V; T need not be the site in all conceivable cases.
Turning attention back to the case of a reacting mixture, the state is the
triplet V; T, x. It is quite obvious that is is possible for the system to be in an
infinite variety of states at any given value of the site-x is a measure of the
composition of the system, and certainly at any given value of the pair V; T
infinitely many different compositions are possible. The reason that x is recognized
as an internal state variable (in contrast to the length of the spring in the first
example above) is that there is no way of keeping x constant at some arbitrary
preassigned value while keeping V and T independently constant as well. It is
also evident that, should one keep V and T constant, x will change in time-the
system is at a constant site, but its state is changing in time. Sometimes the rate
of change will be extremely slow, in fact so slow as to be unmeasurable, and one
could (wrongly) conclude from an experiment that the system is at equilibrium.
For instance, in a mixture of carbon monoxide and oxygen at room temperature
and pressure the rate of oxidation will be unmeasurably small, and so for all
practical purposes x will in fact be constant in time; yet the system is not at
equilibrium. If one fires a spark in such a mixture, the combustion reaction will
in fact proceed to chemical equilibrium, and if the heat of reaction is removed
the system will at the end be at the same volume and temperature, but at a
different composition-there will be quite a bit of carbon dioxide in it, the total
number of moles will be less than it was before, and the pressure will be
correspondingly lower. Unreacted mixtures are not the only example of apparent
equilibria: the glassy state, for instance, is not an equilibrium state, yet its evolution
toward the crystallized equilibrium state is extemely slow and often so slow as
to be unobservable.
It is important to realize that, while all functions of state have equilibrium
values which depend only on the site, there is a distinction between site and
equilibrium state. Again consider the example of a reacting mixture, with the site
being V; T and the state being V; T, x. If the reaction is allowed to proceed until
chemical equilibrium is reached, x will have a value x* which depends uniquely
on the site. The equilibrium state is the triplet V; T, x*, which is uniquely
determined by the site V; T, but does not coincide with it. The statement that
such a system is in an equilibrium state implies that x has indeed the value x*;
62 the statement that the same system is at some site V, T has no such implication-
x may well be different from x* at such a site and, as was discussed in the previous
Chapter Two
paragraph, may in fact stay different from x* for very long stretches of time.
Another conceptual issue which is worth some discussion is the following
one. In mechanics, it is possible to develop a logically coherent theory of statics
without ever appealing to dynamics; this is related to the fact that one can write
down the two basic laws of mechanics directly for the static subcase (the sum
of all forces acting on a system, as well as the sum of all torques acting on it, is
zero). This is not the case in thermodynamics, since there is no satisfactory way
of writing the second law (and perhaps not even the first one) as valid for the
static (or equilibrium) case. In most books on thermodynamics, the "equilibrium"
form of the second law is written down as

dS = dq/T (2.9.1)
but the question which arises is, of course, what one means by dS and by dq. As
far as dS is concerned, one could say that it is the difference between the entropies
at two neighboring states; this only introduces the logical difficulty of what one
means by two equilibrium states being close to each other, but perhaps that could
be circumvented. However, what is dq? The standard answer is that it is the heat
absorbed in a reversible transformation leading from the first to the second state;
however, apart from the difficulty of establishing what is meant by a reversible
transformation without touching on the question of what is meant by an irrevers-
ible one (as would be required if a purely static theory has to be developed),
this answer introduces the concept of a transformation-and, by definition, at
equilibrium no transformation occurs.
Often this point is circumvented by bringing in another difficult concept,
that of a quasi-static transformation, which proceeds "through a sequence of
equilibrium states." Quasi-static is an impressive word, but the only meaning
which can be attached to it is the less impressive word "slow"-and how can
one speak of slowness without implying the concept of time? How slow is slow
enough? If one chooses to develop a thermodynamic theory (rather than a
thermostatic one), the answer is easy. For instance, in the case of a system where
the state is V, T, V, one needs to assume thatf( ) is a Taylor-series expandable
at V = 0 to obtain equation (2.3.4). One then reaches the conclusion that if the
condition

V p* /[af( V, T, 0)/ a V] (2.9.2)

is satisfied, then indeed the difference between p and p* is negligibly small as


compared to p*, and thus the process can be regarded as a quasi-static one.
However, the argument does result from the analysis of a possibly irreversible
transformation; there is no way of developing a logically satisfactory thermostatic
theory other than as a limiting case of a thermodynamic one. The situation is
similar to one which should be familiar to chemical engineers. Suppose we wish
to develop a control system for a process which is to run in steady state. The
aim of the control device is to keep the process at steady state; yet it is impossible
to design the control system without understanding the behavior of the process
in a nonsteady-state condition.
EXAMPLES AND PROBLEMS 63
State and
Equilibrium
Examples

1. Consider a spring which, under the influence of a tensile force f. can be brought
to some length I other than its rest state length. The length of the spring may vary without
any change of volume. One thus comes to the conclusion that there are in fact at least
three external variables which can be imposed arbitrarily and independently of each other:
V, T, and I. We therefore identify the site as ex = {V, T, I}. Correspondingly, there are two
mechanical functions of state, i.e., pressure p and tensile force f. and the rate of work is
-pV+ fl".
First consider the case where the system is elastic, i.e., the state is u = {V, T, I}.
Equation (2.4.1) yields

(p + aAjaV) V + (S + aAjaT)T" + (aAjal- nl" oS 0 (2.E.l)

Since the three rates can be imposed arbitrarily without affecting the values of the
three brackets, one concludes that free energy is a potential not only for pressure and
entropy but also for the tensile force

f = aAjal (2.E.2)

2. Since the state is V, T, I, in general the internal energy U will depend on the
length of the spring, I. However, in the case of a spring made of rubber it turns out that
in fact internal energy is independent of I. This can be ascertained experimentally as
follows. First, the constant volume specific heat is determined experimentally; for simplicity
assume that Cy is independent of temperature. Next, the spring is elongated adiabatically,
and the temperature rise of the spring (if any) is measured. Suppose that U = u( V, T).
Since the spring has been elongated adiabatically, the work done, W (which can be
measured), must be equal to the increase of internal energy. Since V has not been changed
by the elongation, U can only increase if temperature increases. The temperature increase
must be equal to W j Cy, and this can be checked experimentally. For rubbers, the agreement
is quite satisfactory, and one concludes that indeed U = u( V, T). In view of equation
(2.E.2), and of the fact that aU jal = 0, one thus concludes that

f= -TaSjal (2.E.3)

Since free energy is a potential for entropy, one can write

-s=aujaT-S-TaSjaT (2.E.4)

and hence

Tas/aT=au/aT (2.E.5)
and
a2 SjaTal=0 (2.E.6)

This guarantees that aSj al is independent of T, and hence that the force f is proportional
to absolute temperature. The elastic moduli of rubbers are indeed proportional to absolute
temperature.
64 3. Now consider a spring as in the first example, but allow for the possibility of
viscous effects-i.e., that the tensile force required to elongate the spring is different from
Chapter Two the tensile force that the spring exhibits during unloading. This can be described by
assuming that the state is u = {V, T, ~ n.
Thus the second law becomes

(p + iiA/aY) V" + (8 + iiA/iiT)T" + (iiA/iil- f)r + (iJA/iil")r:5 0 (2.E.7)

First, since 1- can be imposed arbitrarily, one concludes that free energy is in fact"
independent of the rate of elongation, so that the last term on the right drops out. The
rates of change of temperature and volume can now still be imposed arbitrarily without
changing the values of the brackets in equation (2.E.7), but the rate of change of length
cannot, since it contributes to the determination of the state. Thus one concludes that free
energy is a potential for pressure and entropy, but not for the tensile force. The second
low reduces to

(f - iiA/iil}l" 2: 0 (2.E.8)

This restricts the values of (f - iiA/iJI), if plotted at constant Ivs. r, to the even
quadrants, and thus, provided!(l") is sufficiently well behaved near r = 0, one concludes
that

f* = !(V, T, 1,0) = iiA/iil (2.E.9)

Equation (2.E.S) now requires the tensile force in loading to be larger, and that during
unloading to be smaller than the equilibrium value corresponding to r = o.

Problems

2.1. Suppose that, for some fluid system, it has been determined experimentally that the pressure
in isothermal expansion is given by an equation of the form

p = p*-Aff (2.P.I)

a. Is this compatible with the second law?


b. Suppose you are willing to assume that the behavior during isothermal compression is of the
same general type. How would you write a generalization of equation (2.P.I) valid in both expansion
and compression?
c. Does the conclusion, that one may effect either expansion or compression slowly enough to
make the total dissipation corresponding to any preassigned change of volume as small as one wishes,
still hold?
2.2. Consider a reacting mixture which, at time zero, is at some equilibrium state V, T, x*. Is
it possible to bring it to some different equilibrium state slowly enough to make the total dissipation
as small as one wishes? Which conditions need to be satisfied, and which assumptions are needed
to reach this conclusion?
2.3. Consider an ideal gas (which is an elastic system) undergoing the socalled Maxwell
expansion. The gas is initially contained in a volume VI" and is separated from an empty volume
VI2 by a membrane. Temperature is held constant, and the membrane is punctured, so that the
available volume suddenly becomes Vu + V.2' In the new situation, pressure is lower, internal energy
is equal, and entropy is larger than at the beginning; hence, the Maxwell expansion is irreversible.
This appears to be in contrast to the result that elastic systems can only undergo reversible processes.
Discuss the reasons for this result, and clarify the paradox. {Hint: This has to do, in some loose
analog sense, with the case of lamination through a valve, which is usually regarded as an irreversible 65
process, as compared to compression in a compressor, which is regarded as a reversible process.)
State and
Equilibrium

LITERATURE

Do you really find no logical difficulty in regarding Nature as


a process of involution, passing from definite coherent
homogeneity to indefinite incoherent heterogeneity?
Lewis Carroll

The concept of "site," as used in this chapter, was introduced in G. Astarita and
G. C. Sarti, Chim. Ind. (Milan) 57, 680, 749 (1975).
The technique of deriving consequences of the Clausius-Duhem inequality used in
this chapter makes crucial use of the concept that external state variables, as well as their
time derivatives, can be imposed arbitrarily and independently of each other. A formal
proof of this is presented in G. Astarita, An Introduction to Nonlinear Continuum Thermody-
namics, SpA Editrice di Chi mica, Milan (1975).
The question of the constant pressure specific heat at phase transitions is discussed
in more detail in Section 4.5.
The question of admissibility of processes and transformations, discussed in the
Appendix to Section 2.1, is conceptually very important. See in this regard the paper by
Feinberg and Levine cited in the Literature section of Chapter 1.
Dissipation in a Maxwellian gas, and the lack of it in pure expansions and compress-
ions, is discussed in detail in the following two references: H. Grad, "Principles of the
Kinetic Theory of Gases," in Encyclopaedia of Physics, Vol. XII, pp. 205-294, Springer-
Verlag, Berlin (1958); c. A. Truesdell, Rational Thermodynamics, 2nd ed., Chapters 8, 9,
and 10, Springer-Verlag, Berlin (1984).
The concept of an absolute temperature is one which has a long tradition and a very
rich literature. Possibly the best recent discussion of the subject, which includes a historical
perspective, is M. Pitteri, in: C. A. Truesdell, Rational Thermodynamics, 2nd ed., Appendix
G6, Springer-Verlag, Berlin (1984).
We have used the idea that entropy is a perfectly legitimate concept also when the
system considered is not at equilibrium. This is always a moot point, but the reader is
simply asked to consider a mixture of hydrogen and oxygen at ambient conditions, a
system for which nobody would deny an entropy is a perfectly acceptable concept, and
yet the system is not at equilibrium, since water would be present at equilibrium in
such a quantity as to make the concentration of either oxygen or hydrogen negligibly
small. That entropy is a perfectly legitimate concept in nonequilibrium situations was so
obvious to the pioneers that they didn't feel the need to make the point explicitly: the
only one who made the point explicitly was J. W. Gibbs, "Graphical methods in the
thermodynamics of fluids," Trans. Conn. Acad. 2, 309 (1873).
The question of invertibility of the constitutive mappings is delicate. If one considers
the mapping delivering the entropy, s( V, T), this is invariably invertible for temperature
(even in highly more complex cases), and hence entropy, rather than temperature, could
be used as an independent variable; this has, e.g., to do with the speed of sound, to be
discussed in Chapter 8. Should one use the pair V, S as the state, temperature can easily
be shown to be (aU/aS)v, which of course guarantees that the derivative (au/aS)v exists
and is always positive. Indeed, Truesdell and Toupin, The Classical Field Theories, Springer-
Verlag, Berlin (1960), develop the classical thermodynamic theory by requiring that internal
energy depends, in addition to mechanical variables such as V, on another variable S
which is dimensionally independent of the mechanical variables; the latter is then identified
66 with entropy. In the mechanical engineering literature, often the pair S, T is used as the
state; this implies, albeit implicitly, that the mapping s( V, T) is invertible for volume,
Chapter Two which is almost always, but not invariably, true. In fact, at constant T, S is generally an
increasing function of V; however, for instance, when ice melts S increases and so does
1/ V-a very simple counterexample.
The question of invertibility of the f( ) function with respect to temperature, which
has been discussed in the Appendix to Section 2.7, casts doubt on the significance of the
axiomatic approach of C. Caratheodory, Math. Ann. 67, 355 (1909), who regards the pair
V, p (or generalizations thereof) as identifying the state. Other difficulties with
Caratheodory's approach have been discussed by G. Whapples, 1. Ratl Mech. Anal I,
302 (1952); B. Bernstein, 1. Math. Phys. 1,222 (1960); J. B. Boyling, Proc. R. Soc. London
Ser. A 329, 35 (1972); J. L. B. Cooper, 1. Math. Anal Appll7, 172 (1967).
The related problem of the anomalies in Carnot cycles resulting from adiabates along
which the rates of change of pressure and temperature have different signs is discussed
by C. Truesdell and S. Bharatha, The Concepts and Logic of Classical Thermodynamics as
a Theory of Heat Engines, Springer-Verlag, Berlin (1977), and by J. S. Thomson and T. J.
Hartka, Am. 1. Phys. 30, 26, 388 (1962).
The concepts of affinity and of "extent of reaction" were first put forward by De
Donder and van Rysselberghe, Thermodynamic Theory of Affinity, Stanford University
Press (1936).
A purely thermostatic theory, which is strictly constrained to equilibrium values of
all functions of state, has in fact been developed by J. W. Gibbs within the restricting
assumption that the state is V, S: Trans. Conn. Acad. 3-16, 108,343 (1876).
Chapter Three

HOMOGENEOUS REACTIONS

It is very easy but perhaps impudent for me, an


academic, an engineer on paper, to criticize those who
have successfully created a large-scale commercial
reactor, but I ask you to ponder on this point. This is a
simple reaction between two molecules at 360C. Surely
the chemical rate is not terribly small, and at 15 cm/sec
in a 10 m deep bed the reactants are in contact for more
than a minute. Surely you would be surprised if the
reaction was not nearly complete in such a deep bed.
P. N. Rowe

NOTATION

a( Constitutive function for free energy kcalkmol- I


A Kinetic matrix arJ/axK S-I

B Any component
B( ) Component distribution
c( ) Concentration distribution kmolm- 3
D( ) Difference operator
E Electrical potential between electrodes volt
E' Applied electrical potential volt
F Faraday's constant coulomb/ equiv
f( ) Constitutive function for pressure atm
g( ) Constitutive function for free enthalpy kcal kmol- I
G{ } Constitutive functional for G
G'{ } Functional derivative of G{ }
h Displacement from equilibrium kmolm- 3
I Current intensity ampere
k( Kinetic constant distribution function S-I

k* Kinetic constant near equilibrium S-I


67
68 kkj First-order kinetic constants S-1

Chapter Three K Equilibrium constant


K[ Equilibrium function kmol
L Any extensive property
lp( Constitutive function for L
L' Partial molar L vector
m( Mass distribution kg
M Number of independent chemical reactions
M( Molecular weight distribution kgkmol- 1
M' Number of chemical reactions
n Number of moles vector kmol
n( Number of moles distribution
P Pressure atm
Pk Multipliers for independence check, discrete
p( Multipliers for independence check, continuous
q Arbitrary affinity kcalkmol- 1
R Gas constant kcal kmol- 1 K- 1
R' External electrical resistance ohm
r Rate of reaction kmols- 1
r Rate of reaction vector kmols- 1
r( Rate of reaction distribution kmols- 1
S Entropy kcal kmol K- 1
T Temperature K
Time s
u Continuous description species variable
u*( Functional derivative of IR{ } at x( ) = x*( S-1

V Specific volume m3 kmol- 1


W Work rate kcals- 1
w,w' Continuous description reaction variable
x Extent of reaction vector kmol
x( Extent of reaction distribution
y Mole fraction vector
Z Dissipation rate kcals- 1
z Electrovalence
a,a' Arbitrary scalars
f3 Coefficient volt ampere- 1
T Number of atoms
6 Affinity kcalkmol- 2
8 Affinity vector kcalkmol- 2 69
6( Affinity distribution Homogeneous
Chemical potential vector kcalkmol- I Reactions
....
p.( Chemical potential distribution
II Product operator
u Stoichiometric coefficient
u( Stoichiometric coefficient distribution

Subscripts

j Componentj
k Reaction k
p Atomp
Total
0 Initial

Superscripts

* At equilibrium
f For formation reaction
Substantial time derivative
0 At standard state
3.1. A REVIEW OF STOICHIOMETRY

To smash the little atom


all mankind was intent.
Now every day
the atom may
return the compliment.
N. Bohr

We consider a one-phase system (so that the assumptions of invertibility discussed


in the preceding chapter are justified) in which a single chemical reaction is
taking place, say, e.g.,

(3.1.1)

Equation (3.1.1) does not, of course, imply addition in the ordinary algebraic
sense of the quantities CO and O2 The meaning of equation (3.1.1) is to recall
the experimentally observed fact that, when a given number of moles of CO is
consumed by the reaction, the same number of moles of CO 2 is produced, and
half as many moles of O 2 are consumed. Let the components appearing in the
reaction be numbered consecutively, say CO is Bt. O2 is B2 , and CO 2 is B 3 , and
let Uj be the corresponding stoichiometric coefficients, say UJ = -1, U2 = -!, and
U3 = 1. We note that the stoichiometric coefficients of the species appearing on
the left-hand side (the "reactants") are assigned negative stoichiometric
coefficients. Equation (3.1.1) can now be written as follows:

(3.1.2)

Let nj be the number of moles of component Bj at time t. The meaning of


equation (3.1.2) is that the quantity nj/ Uj is independent of the index j:

(3.1.3)

where x is the rate of reaction. We note that the meaning in equation (3.1.3) is
left unchanged if all the stoichiometric coefficients are multiplied by the same
nonzero scalar. In particular, considering the case where the scalar is -1, one
sees that the distinction between reactants and products is arbitrary.
The set of stoichiometric coefficients of a reaction is subject to the restriction
of the fundamental axiom of chemistry, the permanence of atoms. In the specific 71
72 case of reaction (3.1.1), the atoms involved are carbon, C, and oxygen, O. The
symbols CO, O 2 , and CO2 are the brute chemiciu formulas of the three components
Chapter Three
involved. The subscripts represent the number of atoms in the component, with
no subscript being interpreted as a subscript of 1, and the absence of an atomic
symbol being interpreted as a subscript of zero. Let 1jp be the subscript of the
pth atom in the brute chemical formula ofthejth component. For reaction (3.1.1),
the matrix 1jp has the following form:

x
CO
C

1
0

J
All others

0
CO2 1 2 0
O2 0 2 0

Note that, for a component which is the molecule of one of the elements
(in this case, O2 ), the corresponding row has only one nonzero entry. That entry
will be identified by Tp , the number of atoms in the molecule of the pth element.
The axiom of permanence of atoms can be written in the form

(3.1.4)

Equation (3.1.4) is a sufficient, but not a necessary, condition for conservation


of mass, and hence the axiom is not empty (see the Appendix).
Let t = 0 be an instant in time when the number of moles of each component
is known:

t = 0, (3.1.5)

One can now define an "extent of reaction" x at time t as

x= J: x dt (3.1.6)

and one obtains immediately

(3.1.7)

This latter equation deserves some comment. Given a system with N com-
ponents participating in the reaction, the composition space has N - 1
dimensions. In fact, since the total mass is fixed, the composition is identified
by the N mol fractions, only N - 1 of which are independent, since they must
add up to unity. For the special case of reaction (3.1.1), three components are
involved and hence the composition space is two-dimensional: the usual triangular
diagram could be used to represent possible compositions (see Figure 3.1.1). 73
However, equation (3.1.7) shows that only one parameter, X, is needed to describe Homogeneous
all compositions which can be reached from the initial one through the occurrence Reactions
of the chemical reaction, and hence the reaction subspace is one-dimensional.
For example, in the case of reaction (3.1.1), if the CO to 02 number of moles
ratio is 2 to 1 at any given time, it will stay at that value at all times; the reaction
subspace would be the straight line going through the CO2 corner.
For a system where only one chemical reaction may take place, i.e., a system
with a one-dimensional reaction subspace, the composition space would also be
one-dimensional only if the number of components were 2. This would require
the chemical reaction to be an isomerization, for which there is only one reactant
and only one product.
We now consider a system where more than one chemical reaction can take
place:
(3.18)

with 1 :S j :S N, 1 :S k :S M', with C1jk an N x M' matrix. Let M be the rank of


the C1jk matrix (the order of the largest square submatrix having a nonzero
determinant). If the number M' is larger than M, some of the reactions which
have been written are linear combinations of others, and can thus be deleted
from the list. When this is done, k = 1, ... , M, and M < N always. The value
of M may be equal to N - 1 only if all the chemical reactions are isomerizations
or polymerizations, since the brute chemical formula of all the Bis would need
to be the same (see the Appendix). Except in such a (rare) case, M < N - 1,
i.e., the reaction subspace has fewer dimensions than the composition space.
In practice, writing a list of independent reactions is easy, and checking the
rank of the matrix can be omitted. Each new reaction added after the first should
contain one and only one component which has never appeared in the previous
list. When this rule is followed, a sufficient set of independent reactions has been
written, provided all components which are expected to be present iii significant
amounts appear in the list (see, in this regard, the Appendix to Section 3.4).
For a system where M independent reactions can take place, the nj values
are given by
nj = njO + L UjkXk (3.1.9)
k

where Xk is the extent of the kth reaction. The XkS are internal state variables,

reaction
subspace

FIGURE 3.1.1. Composition space and reaction


subspace for carbon monoxide oxidation. co
74 and their ordered set is an M-dimensional vector x. Correspondingly, an M-
Chapter Three
dimensional vector affinity 8 can be defined, the components of which are the
partial derivatives of the a() function with respect to the homologous com-
ponents of x:

8 = aa/ax = 8g/8x (3.1.10)

Appendix

If the rank of the Ojk matrix is M, the set of M equations

(3.1.A.l)

for the M unknowns x~ has only the trivial solution x~ = O. Should the rank be
less than M, nontrivial solutions would exist. Let x~ be such a nontrivial solution.
One would have

(3.1.A.l)

and hence equation (3.1.9) would not determine the values of the extents of
reaction univocally. This shows how reduction to a set of independent reactions
is necessary for the theory of chemical equilibrium to be developed.
The question of the axiom of permanence of atoms being stronger than the
mass conservation requirement is as follows. The molecular weight of component
j is given by

(3.1.A.3)

where Mp is the atomic weight of the pth element. If the extent of reaction
changes by an amount dx, the corresponding mass change is

dm =Ldnj~ = dxLOj~
k k

(3.1.AA)

which shows that the mass conservation is implied by equation (3.1.4). However,
dm could be zero under milder conditions, since the sum over j could well be
zero without each individual addend being zero.
Suppose M = N - 1. The axiom of permanence of atoms requires equation
(3.4.1) to be satisfied for all values of k and p. If equation (3.4.1) is divided by
'TNP, with aJ = 'TJp/ 'TNP it reduces to

(3.1.A.5)
This is a system of M equations for the N - 1 unknowns aJ which has a unique 75
solution. Thus the brute chemical formulas of all components are multiples of
Homogeneous
each other, i.e., the components are either isomers or polymers of each other. Reactions

3.2. THERMOSTATICS OF HOMOGENEOUS REACTIONS

Calcination, sublimation, dissolving, putrefaction, distillation,


coagulation, coloration. Whoever shall now ascend and pass
these seven steps shall come to such a wonderful place that he
shall see and experience many secret things in the
transmutation of all natural things.
Paracelsus, 1527

In this section, the fundamental basis for a theory of eqUilibrium of


homogeneous reactions is presented. Since attention is restricted to thermostatic
results, viscous dissipation is not considered; the state is {V, T, x} and equation
(3.1.10) applies. The rate of reaction is a K-dimensional vector which is a function
of state:
x' = r( V, T, x) (3.2.1)

Following a procedure similar to that in Section 2.5, one obtains again


equations (2.5.7) and (2.5.8), i.e., free energy is a potential for both pressure and
entropy. The second law collapses to
z = -8'x'~0 (3.2.2)

Generally, one wishes to determine equilibrium conditions at fixed values


of p and T, rather than V and T. For a homogeneous system, the f( . ) function
is invertible for volume, and hence the baric derivatives can be used. The affinity
vector 8 can thus be regarded as
8 = 8g/8x (3.2.3)

An "equilibrium state" {p*, T*, x*} is one at which the reaction rate vector
is zero, i.e.,
r(p*, T*,x*) = 0 (3.2.4)

What is sought is an "equilibrium function" which, at given p and T, assigns


the corresponding equilibrium value for x*:
x* = K[p, T]; 0= r(p, T, K[p, TD (3.2.5)

Values of all functions of state at an equilibrium state are called equilibrium


values, and are identified by an asterisk.
At any eqUilibrium state, the scalar product of 8 and x is zero and, in view
of equation (3.2.2), at such an equilibrium state that scalar product has a
maximum:

x = K[p, T], d(8' r) = 0 = 8 . dr (3.2.6)


76 where the last equality holds because the rate of reaction is zero at an equilibrium
point.
Chapter Three
Again a smoothness hypothesis is needed to proceed. The assumption is that
r( . ) is invertible for x at x*, say

[det( 8rl 8x)]* F- 0 (3.2.7)

Since the determinant of the transpose of a square matrix is equal to that


of the matrix itself, equation (3.2.6) implies that

8* = 0 (3.2.8)

Equation (3.2.8) only follows from the assumed invertibility, and hence it
is a result which is derived from the consideration of kinetics, albeit only in a
neighborhood of equilibrium. The significance of the assumption of invertibility
is discussed in some detail in the Appendix.
We note that equation (3.2.2) implies that, in all processes for which p and
T are constant, the free enthalpy can never increase. It follows that of all states
corresponding to assigned values of p and T, the equilibrium state has the lowest
free enthalpy. It also follows that cyclic reactions are impossible in the absence
of a simultaneous and independent source of irreversibility. Indeed, consider the
M-dimensional reaction subspace. Equation (3.2.2) guarantees that free energy
is a Liapunov function over this subspace for the rates of change of the space
coordinates. The cyclic reactions acting as internal clocks in biological systems
can occur only because such simultaneous alternate sources of irreversibility are
present as a result of the metabolism. This conclusion also shows that the
thermodynamics of biological reactions cannot be developed on the basis of
simple systems such as that considered in this section.
From a conceptual viewpoint, the subject matter of the thermostatics of
homogeneous reactions is exhausted by equation (3.2.8). In practice, one wishes
to solve the following problem. Given an initial composition, the reaction subspace
is identified by equation (3.1.9). Given p and T, one wishes to find the point in
this reaction subspace for which equation (3.2.8) is satisfied, say the point
corresponding to the lowest free enthalpy. In order to solve such a problem, one
needs to know the constitutive function which delivers 8 as a function of the
state. Some fundamental concepts of the theory of constitutive equations for
mixtures are discussed in the next section.
It is noteworthy that the approach discussed in this section is not restricted
to the case of chemical reactions, but applies to any system for which internal
state variables can be identified. For any such system, the second law reduces to
the requirement that the scalar product of the affinity vector with the rate of
change of the internal state variables vector should be nonpositive. An assumption
of invertibility near equilibrium, analogous to equation (3.2.7), will always pro-
duce the equilibrium condition that the affinity vector is zero. Conversely, if the
second law for any system can be reduced to the requirement that the scalar
product o( two vectors, one of which is a vector of partial derivatives of the
constitutive equation for free energy with respect to some variables while the
other is the vector of the rate of change of those same variables, should be
non positive, the first vector is identified with the affinity and the second with the 77
internal state variables' rate of change. Homogeneous
From the rate of evolution viewpoint, there is an important aspect to be Reactions
discussed concerning equation (3.2.2). As it stands, this equation implies that the
simultaneous occurrence of several chemical reactions never results in an increase
of free energy. It does not imply that each individual chemical reaction takes
place in such a way that, by itself, it causes free energy to decrease; in other
words, the individual products 9i<Xk are not required to be separately nonpositive,
but rather only their sum. This could also be worded in a different way: The
individual products -9i<Xk represent the individual rates of dissipation; the second
law does not exclude the possibility that some individual rates of dissipation may
be negative, through coupling with simultaneous chemical reactions, provided
the total rate of dissipation is nonnegative. In actual fact, the literature on
chemistry almost always makes the assumption that each individual chemical
reaction takes place in such a way as to result by itself in a nonincreasing free
energy; this poses a more restrictive constraint on possible paths in the reaction
subspace than the second law by itself would imply.

Appendix

The assumption ofinvertibility in equation (3.2.7) should be considered with


care. First we consider the case where there is only one reaction; the assumption
is now that ar/ax = k* is nonzero at x = x*, i.e., that in a neighborhood of the
equilibrium point the rate can be expressed as a linear function of the displacement
from equilibrium:
r = k*(x - x*) + O[(x* - X)2] (3.2.A.l)

This equation determines an intrinsic time scale for the reaction, 1/ k*. Seen
in this light, the assumption of invertibility may appear to be very strong, since
a simple counterexample comes immediately to mind: an irreversible dimerization
reaction, for which one would write r as kc 2, with k a "second-order" kinetic
constant. Quantity kc 2 is of course not invertible at equilibrium (c = 0). However,
this is an artifact of the assumption that the reaction is irreversible. In fact, let
the reaction be reversible, so that its rate would be given by kc 2 - kC2/ K, where
C2 is the concentration of the dimer and K is the concentration-based equilibrium
constant. The equilibrium concentration of monomer, c*, is of course given by
C*2 = e2/ K, and hence the rate can be written as k(e 2 - C*2). Now let the actual
concentration be very close to equilibrium, say e = c* + h, with h c*. One
obtains
r = 2ke*h + O[(h/e)2] (3.2.A.2)
which is indeed linear in h, the displacement from equilibrium, and is invertible.
The constant k* is thus identified with 2kc*, which is clearly degenerate for a
truly irreversible reaction (e* = 0). However, real reactions are never truly
irreversible, and thus no problem arises. Of course, for an almost irreversible
reaction, one would not use k* in the description of the kinetics, but that does
not mean that there is not a finite (if very small) value of k*.
78 Next we consider the case of two reactions and suppose that, in a neighbor-
hood of equilibrium, the reaction rates are given by
Chapter Three

(3.:U.3)

and
(3.l.A.4)

Now the invertibility condition is violated, in spite of both rates being linear
in the displacements from equilibrium, and the condition in equation (3.2.6) is
satisfied with

of=q and or=-q/a' (3.2.A.5)

for arbitrary nonzero q. This seems to be a quite likely counterexample, but in


fact there seems to be no known case of such a situation.
Extension of this result to the case of M independent reactions is easy. Let
A stand for the M x M matrix iJrJ / iJXK at equilibrium. If det A is zero, there is
at lea~t one nonzero vector of displacement from equilibrium dx such that

A dx= 0 (3.l.A.6)

(Of course, any multiple of such a vector satisfies the same equation.) For instance,
in the case of equations (3.2.A.3) and (3.2.A.4), any vector proportional to
(1, -1/ a') gives displacements from equilibrium for which the rates are zero.
The set of vectors satisfying equation (3.2.A.6) is called the kernel of A. If A is
invertible, its kernel contains only the zero vector. Any affinity vector lying in
the kernel of A would satisfy equation (3.2.6), and hence equation (3.2.8) is the
result obtained when A is invertible. Extension to the case of infinitely many
reactions is discussed in Section 3.6.
The question of free energy being a Liapounov function is briefly reviewed
in the following. Consider two variables x and Y. and let F(x, y) be a function
of them; F(x, y) = C describes a curve in the x- y plane. Let F( ) be such that
all such curves are closed, no two of them cross each other, and if C) < C2 the'
curve corresponding to C) lies within C2 For one particular value of C, which
by normalization can be set to zero, the ciosed curve degenerates to a point xo, Yo,
say F(xo. Yo) = o.
Now we allow the x and y variables to vary in time, where x(t) and y(t)
are the parametric equations of a curve in the x-y plane which is called a path;
the curve has an intrinsic direction identified by increasing time. Now we assume
that the path is restricted by the following requirement: if the point x(t), y(t)
lies on F = C), then x(t + a), y(t + a) for any positive a lies on F = C2 < C).
This guarantees that any path will eventually end at point Xo, Yo. and that no
path can close on itself. If these conditions are satisfied, F(x, y) is said to be a
Liapounov function for the paths.
We now let x and y be the extents of two independent reactions, and let
F(x, y) be the free energy (if V and T are held constant). The affinity vector at
any point x, y is orthogonal to the F curve through that point, and points outward. 79
Equation (3.2.2) now guarantees that the rate of reaction vector, which is tangent Homogeneous
to the path and points in its intrinsic direction, has a positive component on the Reactions
inward-pointing direction, which guarantees that free energy is a Liapounov
function for the paths in the reaction subspace. The argument is easily generalized
to the case of more than two independent reactions.

3.3. PARTIAL MOLAR PROPERTIES

Mathematicians are like Frenchmen: whatever you say to them


they translate into their own language and forthwith it is
something entirely different.
Goethe

We raise a great dust and complain we cant see.


Bishop Berkeley

The theory of mixtures is best developed by using moles, rather than grams,
as the unit of mass. This is due to the fact that the stoichiometric coefficients of
reactions turn out to be very simple small integers if mole units are chosen.
Therefore, from this point on we will use mole units unless otherwise stated; no
confusion should arise from the fact that the same symbol, say V, is used for the
molar specific volume as has been used so far for the mass specific volume.
The fundamental concepts of the theory of mixtures are simple consequences
of the fact that most of the quantities of interest are extensive, say they are
absolutely additive functions of mass. Let n be the number of moles vector, with
components n l , ... , nN which are the number of moles of each component
present in the system at some time t, and let n be the total number of moles at
time t:

(3.3.1)

The mole fraction vector y is also an N-dimensional vector defined by

y = n/n (3.3.2)

and of course its components add up to unity.


The whole analysis in this section is based on the assumption that the f( )
function is locally invertible, and hence baric derivatives can be used. Thus one
can regard the state as being

u = p, T, n (3.3.3)

Correspondingly, any extensive property L is given by a constitutive equation of


the form

(3.3.4)

with the partial derivatives of fp( ) being the baric derivatives.


80 Since L is an absolutely additive function of mass, the following equation
Chapter Three
must hold true for an arbitrary positive scalar a:

(3.3.5)

On subtracting 1., from both sides, one obtains

(a - 1)lp(p, T, n) = (a - 1)6L/6n' D + O[(a _1)2] (3.3.6)

Taking the limit as a approaches unity yields the fundamental result

Lt=nL' (3.3.7)

where the Lj are the partial molar properties

L' = 6L /6n
t (3.3.8)

If the baric derivative of equation (3.3.7) is taken with respect to njo we obtain

(3.3.9)

and hence

n' 6L'/6nj = 0 (3.3.10)

Equation 3.3.10 is known as the Gibbs-Duhem equation.


As was discussed, for instance, in the last section, equilibrium calculations
are generally based on minimization of the free enthalpy lj.t assigned p and T.
Hence the partial molar free enthalpy vector is assigned a special symbol, fl. The
components of fL are called the chemical potentials. In general, the relationships
established in earlier sections between extensive properties also hold for partial
molar quantities; in particular, the following two are of importance:

6fL/6T = -S' (3.3.11)

and
6fL/6p = V' (3.3.12)

The analysis given above duplicates that usually encountered in textbooks,


although a somewhat different viewpoint has been taken. The approach followed
here, however, leads easily to generalizations, and in particular to the analysis
of continuous mixtures to be discussed in Section 3.6.
The important point to be stressed at this stage is that the whole analysis is
simply a consequence of the fact that extensive properties are absolutely additive
functions of mass, and it therefore has in reality no thermodynamic content
whatsoever. Any extensive property offers the possibility of defining its partial
molar values, and equations (3.3.7) and (3.3.10) would hold for any such quantity,
such as the electrical capacity of a mixture.
Another important point is that there is nothing magic about mole units, 81
and exactly the same analysis can be carried out in mass units, thus defining HomogeT!eous
partial mass properties, Lj. In fact, the choice of partial mass units is appropriate Reactions
when considering the thermodynamics of diffusion phenomena, to be examined
in Chapter 7.
The question of mass vs. mole units also pertains to the theory of dilute
solutions, which is discussed in Chapter 8. Finally, it should be remembered that,
in the case of polymers, mole units are entirely out of the question, because the
molecular weight of polymers is very large, and in the case of crosslinked polymers
it is in fact infinity.

3A. FUNDAMENTALS OF CHEMICAL EQUILIBRIUM mEORY

Formerly it was supposed that all chemical reactions took place


completely. It was conceived that the stronger affinity caused
the reaction to be complete at the expense of the weaker
affinity. ... Thermochemists, more particularly Berthelot, have
striven to uphold this conception, which has no strict scientific
foundation.
Arrhenius, 1897

The classical theory of chemical equilibria in homogeneous systems is based


on equation (3.2.11). Taking into account equation (3.1.9), the kth component
of the affinity vector can be calculated as follows:

8k = aa/aXk = I (aa/anj)(anj/aXk)
j

(3A.I)

and therefore equation (3.2.11) can be written as

(3A.2)

Since the quantities ILj depend on composition, in principle equation (3.4.2)


is sufficient to determine the equilibrium composition. However, in order to
actually calculate the eqUilibrium composition, constitutive equations for ILj need
to be written down. Such equations in general make use of the concept of a
standard state. The standard state is chosen for all components once and for all;
constitutive property values in the standard state are identified by a superscript
O. The standard state is always a pure component state, and hence the values of
the molar density and of the partial molar value of the same property coincide.
Furthermore, the standard state is always chosen at the same temperature as the
mixture under consideration. For any given reaction, the standard change of any
extensive constitutive quantity L is defined as follows:

Dk(LO) = I UjkLJ (3A.3)


j
82 while the constitutive equation fQr the chemical potential has the form
Chapter Three
(3.4.4)

where aj is the activity of component j. Substitution of the definitions given above


into equation (3.4.2) and rearrangement yields

IT a'P' = exp[ -Dk(GO)/ RT] = Kk (3.4.5)


j

where Kk is called the (activity based) equilibrium constant for the kth reaction.
The task of calculating equilibrium compositions is now reduced to the following
two steps:

1. Write constitutive equations which relate the activities to the concentra-


tions in the reacting mixture.
2. Determine experimentally the values of the equilibrium constants for the
reactions of interest.

The first task will be discussed later. As for the values of quantities K k , it
is evident that they depend on the particular choice of the standard state. For
chemical reactions, the standard state is chosen as that where the pressure is one
atmosphere; this of course means that Kk is independent of the actual pressure
of the reacting mixture, and in fact depend only on temperature.
Given a chemical reaction with stoichiometric coefficients which are a linear
combination of those of other reactions, its value of Kk can be calculated from
the values for the other reactions, as can easily be seen from the definitions. This
point allows one to construct a simplified way of tabulating Kk data, discussed
below.
For any given component Bj , its "formation reaction" is defined as the
reaction which transforms the molecules of the elements into the component
considered. For instance, in the case of carbon dioxide the formation reaction is

(3.4.6)

The formation reaction is identified by superscript f. Referring to the


nomenclature for stoichiometric coefficients in Section 3.1, the stoichiometric
coefficients for the formation reaction are seen to be

uj = 1 (3.4.7)

and
(3.4.8)

Substitution into the definition of the standard change, while making use of
equation (3.1.4), yields

D(LO) = ~ Up(LJf) (3.4.9)


j
which shows that the task of tabulating values for all reactions of interest is 83
reduced to the task of tabulating values for the formation reactions of the Homogeneous
components of interest. For components which are molecules of elements, D( L Of) Reactions
is of course zero.
If constitutive equations for the chemical potentials are available, as well as
values of the relevant equilibrium constants, the calculation of the eqUilibrium
composition can be based on the solution of equations (3.4.5). These are a system
of coupled polynomial equations, and if the number of independent chemical
reactions is not rather small, the calculational problem may become rather difficult.
However, with the same amount of information, it is possible to actually evaluate
the free enthalpy of the system for all points in the reaction subspace, and
therefore the equilibrium point can be obtained by directly minimizing the free
enthalpy in the reaction subspace. The latter procedure is numerically preferable
when the number of independent chemical reactions involved is rather large.
With regard to the procedure based on minimizing free enthalpy over the
reaction subspace, two important questions arise. The first is as follows: Since
the free enthalpy of any given pure component is defined to within an arbitrary
additive constant, the free enthalpy of a mixture is defined to within an arbitrary
additive hyperplane over the composition space, and thus the location of the
minimum may appear to be unspecified. This, however, turns out not to be the
case over the reaction subspace, as will be discussed in Section 10.2.
The second problem is the possibility of the existence of more than one local
minimum of the free enthalpy. As will be discussed in the next chapter, for
mixtures which are homogeneous over the whole composition space the free
enthalpy surface can present only one minimum, and thus the question does not
arise. More than one local minimum over the reaction subspace may be exhibited
only by mixtures which also exhibit the possibility of phase separation, and this
will be discussed in Section 4.3.

Appen~ix

The question of the number of independent chemical reactions to be con-


sidered in the solution of a chemical eqUilibrium problem should be analyzed
carefully. Suppose a set of M independent reactions has been written down and
includes N components. An (N + l)th component could be formed by a reaction
which has not been considered in the set; let the stoichiometric coefficient of this
new component be +1 in the new reaction (this is always possible by normali-
zation), and suppose that the eqUilibrium constant for the new reaction is a very
small number. This implies that, at equilibrium, the concentration of the new
component is guaranteed to be very small, so that its number of moles is going
to be negligible as compared to that of other components. Now it should be
considered that'the coupling between the original and new set of equations occurs
only through the linear equation (3.1.9) withj = N + 1. It follows that the solution
of the reduced problem, where the new component is not considered at all, is
approximately correct, since the linear equation will be approximately satisfied
anyhow (nN+1 being much smaller than other numbers of moles appearing in
the same equation). This consideration allows one to first solve the M-order
84 problem and then, if the concentration of the (N + l)th component needs to be
known, it can be calculated from the eqUilibrium equation for the (M + l)th
Chapter Three
reaction, in which all the other concentrations are calculated from the M-order
problem. Often in fact the concentration of components which are present in
very small amounts does not need to be calculated at all; thus the M-order
problem is all that is needed. An important exception is that of electrolytic
reactions, to be discussed in Chapter 11.

3.5. ELECTROCHEMICAL REACTIONS

In the mutual contact oJ, say, silver with tin a force is


generated, so that silver gives electrical fluid to tin.... This
force produces, if the circuit is completed, a cu"ent which goes
from silver to tin, and from this through the wet conductor goes
back to the silver. ...
A. Volta, 1796
Electrochemical reactions constitute a thermodynamically special class of
chemical reactions, as discussed below. Let us consider an electrochemical circuit
as drawn in Figure 3.5.1. The electrical potential difference between the two
electrodes is E. The part of the circuit which is external to the cell includes a
resistance R' and an applied potential difference E'. These two quantities can
be imposed at will, independently of the state of the system in the cell. In
particular, R' can be set to infinity by switching open the external circuit.
Within the cell, a chemical reaction takes place, at a rate x'. This rate is
related to the intensity I of the electrical current which circulates through the
circuit by Faraday's law:
1= zFx' (3.5.1)

where F is a universal constant, and the electrovalence of the reaction, z, is a


fixed number for any given reaction. Equation (3.5.1) shows that the extent of
reaction is not an internal state variable in the usual sense: its rate of evolution
can always be set to zero by opening the external circuit.

CELL
FIGURE 3.5.1. Sketch of the circuit of an
electrochemical cell.
The net rate at which work is being done on the system within the cell is 85
the usual mechanical power plus the electrical power, say Homogeneous
w. = -pV;- El (3oS.:!)
Reactions

The state of the system is determined by volume, temperature, extent of


reaction, and applied electrical potential. Hence, in particular,

(3.5.3)

Substitution of equations (3.5.1)-(3.5.3) into the second law yields

[(aAJaT) + St]T + [(aAJay,) + plY;


+ (8 + EzF)x" + (aAtjaE)E" 05 0 (3.5.4)

Since E" can be imposed at will without influencing the state of the system,
just as T" and V;, equation (3.5.4) shows that pressure and entropy are given by
the usual partial derivatives of the free energy function, and that the free energy
itself is in fact independent of the applied electrical potential. With this, the
second law reduces to

(8 + EzF)x" 05 0 (3.5.5)

where the affinity 8 is independent of the electrical potential, since so is the free
energy. We note that, since it is always possible to open the external circuit and
thus make x" zero, equilibrium can be achieved at any value of x. We also note
that, when one opens the external circuit, E is no longer an external state variable,
since it cannot be imposed at will in an open circuit. Thus, when R' = 00 (open
circuit), E will have some value E* which is called the electromotive force of
the reaction. The reaction rate is a function of state, i.e., it is given by

x" = r(Y" T, E,x) (3.5.6)

and E* is the solution of the equation

r(Y"T,E*,x) =0 (3.5.7)

Under sufficient assumptions of smoothness, equation (3.5.5) implies that,


when x" = 0,

8* = -E*zF (3.5.8)

Since 8 is independent of E, equation (3.5.8) implies that

() = -E*zF (3oS.9)

Substitution of equation (3.5.9) into (3.5.5) yields

(E - E*)1 05 0 (3oS.tO)

since zF> o.
86 Equation (3.5.10) describes the irreversibility connected with electrochemical
Chapter Three
reactions. It implies that, when the cell is being charged (I < 0), the electrical
potential is larger than when the cell is discharged. This in tum implies that the
amount of electrical work needed to charge the cell is more than the amount that
the cell yields upon discharge.
Equation (3.5.8) is, as usual, based on an assumption of smoothness. This
should be considered in some detail. Consider a fixed triplet Yt, T, x, so that E*
is fixed at the solution of equation (3.5.7). Equation (3.5.10) now imposes a
restriction on the constitutive function r( . ), the value of which can never have
the same sign as E - E*. Thus in an I vs. E - E* plot such as in Figure 3.5.2
(at fixed Yt, T, and X, I is a unique function of E), the curve is restricted to the
two odd-sign quadrants, and thus, if sufficiently smooth, it will pass through the
origin, and hence equation (3.5.8) holds.
In actual fact, the curve does pass through the origin, but its slope at the
equilibrium point is very small, as shown qualitatively in Figure 3.5.2. Only at
comparatively large absolute values of E - E* does the intensity become reason-
ably large for practical use, and in fact the curve becomes almost vertical. Two
regions can thus be identified. The first corresponds to very small intensities, and
the curve can be approximated by a straight line through the origin:

f31 = (E* - E) (3.5.11)

In this region, the rate of dissipation is given by

I(E* - E) = f312 (3.5.12)

Thus it is concluded that, by having a sufficiently large value of R', one can
charge or discharge the cell very slowly, and the rate of dissipation is quadratic
in the intensity. The time interval needed to charge or discharge the cell by a
given amount is inversely proportional to the intensity, and hence the total
dissipation is proportional to the absolute value of the intensity, and can be made
as small as one wishes.
The situation described above, however, holds only for extremely low
intensities, and in actual operation a cell will work on one of the two almost

FIGURE 3.5.2. Intensity vs. difference between


actual and equilibrium electrical potential.
vertical branches of the curve. Under these conditions, E* - E is almost indepen- 87
dent of I, and therefore the rate of dissipation is approximately proportional to Homogeneous
the absolute value of the intensity. This means that charging or discharging the Reactions
cell by a given amount results in a total dissipation which is approximately
independent of the intensity: in any given charge-discharge cycle, a constant
amount of electrical energy is dissipated. The irreversibility just described is
different in character from that of ordinary reactions.

3.6. CONTINUOUS DESCRIPTION OF REACTING MIXTURES

There will be those of us who will complain that anything so


mathematical cannot be of much practical use, and there will
be those of us who will, il,J fact, do theoretical work of little
worth. ... In the meantime we might gain perspectioe by
looking past our boundaries: we need only look to work of
modem economic theorists (social scientists) to find that
chemical engineers are not quite so mathematical after all.
M. Feinberg

It may frequently happen that one deals with mixtures of very many com-
ponents, which may even not all be individually known. Crude oil is a good
example of such a situation. In these cases it is useful to describe the mixture
by means of one or more distribution functions, i.e., to have a continuous
description, rather than a discrete one where some finite number of components
is kept track of. Even if the individual components of a mixture are known, if
their number is large one may wish to substitute a few pseudocomponents for
ease of analysis. A continuous description furnishes a rational method for choos-
ing the pseudocomponents, in that one can look for the appropriate quadrature
points of the resulting integrals. The mathematics involved with the continuous
description of mixtures are rather involved, and only a brief outline of the theory
is given in this section. Some mathematical technicalities are dealt with in the
Appendix.
In the case of a discrete description, jndividual components are identified
by the subscript j, which takes integer values from 1 to N; in a continuous
description, components are labeled by a variable u which takes all values over
some finite; or possibly infinitely large, range. So the specific species is B(u),
just as it is Bj in the discrete description. In the continuous description, the
composition space is infinite-dimensional.
The fundamental distribution function is the mass distribution function m(u),
which is defined as follows: m(u) du is the mass of species identified by values
of the label between u and u + dUo It is important to observe that, if one allows
any distribution function to contain also a sum of a finite number of terms each
of which is proportional to a Dirac delta function I) (u - j), the discrete description
is subsumed by the continuous one.
Let M(u) be the molecular weight of species B(u). The distribution function
of the number of moles is

n(u) = m(u)/ M(u) (3.6.1)


88 and the concentration distribution function is
Chapter Three
c(u) = n(u)/V, (3.6.2)

Now suppose one wants to describe a reaction in such a mixture. This is


accomplished with equation (3.1.2) in the discrete case. The obvious extension
to the continuous case is

f u(u)B(u) du =0 (3.6.3)

where u( u) is a stoichiometric coefficient distribution function. Of course, it is


quite likely that only a f~w components participate in the reaction considered,
and thus u( u) is likely to be expressible as a sum of a few delta functions. The
u( u) distribution function must obviously satisfy the conservation of mass condi-
tion, i.e.,

f u(u)m(u) du =0 (3.6.4)

(Permanence of atoms imposes in fact a stricter constraint; see the Appendix.)


The interesting case, however, is where many reactions may take place, and
one may begin by considering the case of M reactions identified by the subscript
k, in analogy with the discrete case:

f uk(u)B(u) du = 0, k = 1,2, ... ,M (3.6.5)

with equation (3.6.4) satisfied for all k.


The question of whether the M reactions are independent pertains, in the
discrete case, to the rank of the stoichiometric coefficients matrix, which is M x N
with N > M. The matrix has rank M if the equation

(3.6.6)

has only the trivial solution Pk = o. We note that equation (3.6.6) must hold for
all k from 1 to N.
In the continuous description we have M stoichiometric coefficient distribu-
tion functions, and the independence condition becomes that the equation

(3.6.7)

has only the trivial solution Pk = 0 over the whole range of u values.
However, since in essence one is dealing with an infinite number of species, 89
it makes sense to consider the possibility of describing an infinite set of chemical Homogeneous
reactions. Just as the component label j of the discrete description carries over Reactions
to a variable label u, the reaction label k carries over to a variable label w which
identifies the reaction. Now consider a piecewise-continuous distribution function
of stoichiometric coefficients u(w, u) which satisfies the mass conservation
requirement for all values of w:

f u(w, u)m(u) du = 0 (3.6.8)

Then

f u(w, u)B(u) du =0 (3.6.9)

is a continuum of reactions over some range of the variable w.


The requirement of independence for the reactions in this continuum is the
straightforward generalization of that relative to equation (3.6.7). The functional
equation

f p(w)u(w, u) dw = 0 (3.6.10)

must, over the whole range of u values, have only the trivial solution p( w) = o.
It is now possible to introduce a distribution function of extents of reaction,
x(w, t), which is defined by the analog of equation (3.1.9):

n(u, t) - n(u,O) = f u(w, u)x(w, t) dw (3.6.11)

Correspondingly, the rates of reaction distribution function is

r( w, t) = ilX( w, t)/ ilt (3.6.12)

Here, equation (3.6.11) restricts the values that n(u, t) may assume, given
the values of n(u, 0); i.e., it identifies the reaction subspace. However, now also
the reaction subspace is infinite-dimensional.
We now consider a mixture where, say, a whole series of cracking reactions
may take place, and suppose that one has chosen the carbon number as the label
w. A typical cracking reaction could be written as

B(u + u') = B(u) + B(u') (3.6.13)

but now the product species could react again. In this case, a bicontinuous
distribution of reactions would need to be considered, say a distribution function
u( w, w', u) such that

f u(w, w', u)m(u) du = 0 (3.6.14)


90 which defines a bicontinuous distribution
Chapter Three
f u(w, w', u)B(u) du = 0 (3.6.15)

with the condition of independence now being that the equation

ff pew, w')u(w, w', u) dwdw' (3.6.16)

has only the trivial solution p( w, w') = O.


For the cracking reaction considered above, u(w, w', u) is given by

u(w, w', u) = 8(u - w) + 8(u - w') - 8(u - w - w') (3.6.17)

So far, only the definitions have been laid down. In order to develop a
thermodynamic theory, one needs to face several issues. First, one must decide
what is the state of a continuous mixture. Since attention is restricted to
homogeneous reactions, the development below is given in terms of free enthalpy,
and p and T are regarded as independent variables. Thus in the discrete case
the state would be p, T, D. The obvious extension to the continuous case is that
the state is p, T, n(u). This however implies that the state is not identified by a
finite number of parameters, but by two parameters and a function. It follows
that the constitutive mappings are functionals (a few fundamental concepts and
definitions of functional analysis are given in the Appendix). Thus the free
enthalpy is given by

Gt = G{p, T; n(u)} (3.6.18)

where G{ } is a functional of the mole number distribution function n(u), which


also depends explicitly on -p and T.
One now needs to define chemical potentials, i.e., to develop the analog of
the formalism in Section 3.3. Partial molar properties have the structure discussed
in Section 3.3 because extensive properties are absolutely additive functions of
mass, say at constant composition, equation (3.3.4) holds. That condition can be
written, in the continuous description, as follows:

G{p, T; an(u)} = aG{p, T; n(u)} (3.6.19)

This can be shown to imply that a well-defined functional derivative (see


the Appendix) of G{ }, G'{p, T, u; n(u)}, exists; it appears in the following
relationship:

G{p, T; n(u)} = f G'{p, T, u; n(u)}n(u) du (3.6.20)

which is the equivalent of equation (3.3.13). Thus G'{ } is identified as the


chemical potential distribution function J.t(u):

J.t(u) = G'{p, T, u; n(u)} (3.6.21)


and 91

G, = f~(u)n(u) du (3.6.11)
Homogeneous
Reactions

which is the analog of equation (3.3.7).


Let us now consider a differential variation of composition, dn(u), resulting
in a variation dG, of the total free enthalpy. Equations (3.6.20) and (3.6.21) yield

dG, = f~(u) dn(u) du (3.6.13)

while differentiation of equation (3.6.22) gives

dG, = f[~(u) dn(u) + n(u) d~(u)] du (3.6.24)

It follows that

f n(u) d~(u) du = 0 (3.6.15)

which is of course the continuous formulation of the Gibbs-Duhem equation.


Now we consider one particular reaction, say a specific value of the label
w, and let its extent change by an amount dx(w). The corresponding change of
the number of moles distribution function is obtained from equation (3.6.11) in
the form

dn(u) = u(w, u) dx(w) (3.6.26)

while the corresponding change of free enthalpy is given by

dG, = f G'{p, T, u; n(u)}u(w, u) dx(w) du

= dx(w) f u(w, u)~(u) du (3.6.27)

It follows that the affinity distribution function is

6(w) = f u(w, u)~(u) du (3.6.l8)

which is the analog of equation (3.4.1).


The rate of change of free enthalpy can now be calculated:

f
G; = (aG/ap)p' + (aG/aT)T + ~(u)n'(u) du (3.6.29)
92 The usual argument can now be used to infer again that free enthalpy is a
potential for volume and entropy, and that the second law reduces to the
Chapter Three
requirement that the integral on the right-hand side of equation (3.6.29) should
be nonpositive. If the rate of change of n(u) is calculated from equation (3.6.11)
and substituted into the integral, one obtains for the rate of dissipation

Z(t) = - JJ p.(u)u(w, u)r(w, t) dudw

= - J (J(w)r(w, t) dw ~0 (3.6.30)

which is the generalization of equation (3.2.2). The usual assumption in chemistry


that each individual reaction proceeds in the direction of decreasing free enthalpy
would correspond to the requirement that the integrand in equation (3.6.30) is
identically nonpositive; the second law requires only the integral to be nonpositive.
We now assume that, given an initial composition n(u, 0), equation (3.6.11)
determines the composition n(u, t) at any later time in terms of the extent of
reaction distribution x(w, t). It follows that all constitutive properties can be
written in terms of functionals of the extent of reaction distribution. In particular,
the affinity and the rate of reaction distributions are, respectively,

(J(w, t) = 8{x(z, t); w} (3.6.31)

and
r( w, t) = R{x(z, t); w} (3.6.32)

where z is a dummy label of reaction. The affinity of reaction w depends on the


whole extent of reaction distribution x(z, t) and, of course, parametrically on w
itself. Since wand z span the same range, the question of the invertibility of
R{ } is a legitimate one.
The fact that Z(t) has a minimum at the equilibrium point x*(z, t), where
r(w, t) is zero, implies that the following equation needs to be satisfied:

JJ 8{x*(z, t); w}u*{z; w}s(z) dzdw =0 (3.6.33)

where u*(z; w} is the functional derivative ofR{ } at x(z, t) = x*(z), which of


course depends parametrically on w, and s(z) is any arbitrary displacement from
equilibrium, s(z) = x(z) - x*(z). Now let I[) be the set offunctions s(z) to which
u*(z; w) is orthogonal, say functions for which the following equation is satisfied:

f u*(z; w)s(z) dz = 0 (3.6.34)

Now any (J(w) in [J) would satisfy equation (3.6.33). However, if R{ } is


invertible at x(w) = x*(w), [J) contains only the zero function, and hence the
eqUilibrium condition is established as

(J*(w) = 0 (3.6.35)
i.e., the equilibrium affinity is zero for all values of w. Comparing this with 93
equation (3.6.28), the equilibrium condition can be written as
Homogeneous

f u(w, u)#*(u) du = 0 (3.6.36)


Reactions

over the whole range of w values. This is the analog of equation (3.4.2).
The theory of the kinetics of reactions in continuous mixtures is still very
primitive, and very few problems have been tackled outside of the case where
all the kinetics involved are first-order. This case would be described, in the
discrete formulation, as follows:

rd . = L kkjCj (3.6.37)
j

The generalization to the continuous case introduces a distribution function


k(w, u):

r(w,t)/.= f k(w,u)c(U, t) du (3.6.38)

A slightly more general result can be obtained by the following consideration.


Given a differential displacement from equilibrium ds(z), the corresponding rate
of reaction distribution function dr( w) is given by

dr(w) = f u*(z; w) ds(z) dz (3.6.39)

where u*(z; w) is the functional derivative appearing in equation (3.6.34). That


functional derivative has been assumed to exist and to be orthogonal only to th~
zero function in order to obtain the result in equation (3.6.36), i.e., what is
universally regarded as the correct description of chemical equilibrium. It follows
that the assumption is presumably a very good one.
We now note that u*(z; w) has units of a frequency. Also, if there is only
one reaction to worry about, e.g., w = W, equation (3.6.39) reduces to
dr(W) = u*(W; W) ds(~) (3.6.40)

Equation (3.6.40) allows u*(w; w) to be interpreted as the pseudo-first-order


kinetic constant of reaction w in a neighborhood of eqUilibrium, i.e., the analog
of the k* discussed in the Appendix to Section 3.2.

Appendix

A functional can be regarded as a transformation of an argument function


into a value. The simplest example of a functional is a definite integral. We

r
consider the following equation:

y= f(u) du (3.6.A.l)
94 The value of y depends on which particular functionf(x) is used in perform-
Chapter Three
ing the integration, but it does not depend on u-a functional is not a composite
function. The following notation is used for a functional:

y = 1F{f(u)} (3.6.A.2)

and, as in the case of equation (3.6.A.1), the argument function need only be
defined over some range of u values [a < u < b in the case of equation (3.6.A.2)]
which is not indicated in the notation adopted unless it is not obvious from the
context.
A linear functional is one which satisfies the usual condition of linearity:

(3.6.A.3)

where the difference between two functions is defined as that function which has
value equal to the difference between the values of the two functions. Of course,
the integral in equation (3.6.A.l) is a linear functional.
From this point on, notation is simplified by using f, g, etc., for the argument
functions. The Frechet differential 811' of a functional is defined in the following
way:

1F{f + g} = IF{J} + 81F{fl g} + IR (3.6.A.4)

and it is a functional of both f and g which is linear in g. The residual IR is


required to be o(llgll), i.e., it goes to zero faster than Ilgll, where Ilgll is the norm
of g. The norm Ilgll is best understood by considering that the domain of F{ }
is a set of functions, and this set can be visualized as a space. Quantity IIgll is
then interpreted as the distance, in this space, between f and f + g.
A little more formally, the domain of IF{ } can be assigned a metric by
assigning a norm, i.e., by assigning a rule for calculating the distance between
any two functions f and h, namely Ilh - fll. The norm of h - f must satisfy the
triangular condition: it is to be a nonnegative scalar, II h - fll cannot be larger
than II h - gil + II g - fll, II h - h II must be zero, and II h - fll must be equal to
Ilf - hll We note, however, that the condition that IR be o(llgll) only requires
identification of infinitesimal distances, i.e., only the topology of the domain
need be known. Given a metric, a corresponding topology is induced; however,
different metrics could induce the same topology. A simple example of a norm is

Ilf(u)11 = f If(u)1 du (3.6.A.5)

where If(u)1 is the absolute value of feu).


Given a functionallF{J}, its first Frechet differential, if it exists with respect
to the appropriate topology, is given by

8F{f1 g} = dlF{f + eg}j del E _ O (3.6.A.6)


Since the Frechet differential is linear in its second argument, it can in general 95
be expressed as an integral: Homogeneous

IlF{flg} = f k(u)g(u) du (3.6.A.7)


Reactions

but the kernel k( u) still depends on f, since so does 1l1F{ }. Hence the kernel is
a functional itself, which also depends explicitly on u as a parameter:

k(u) = k{u;f(u)} (3.6.A.8)

k{u;f(u)} is called the functional derivative of IF{ }.


The condition in equation (3.6.19) (which, incidentally, does not imply that
the functional G{ } is linear) can be used to infer that

(a -1)G{p, T; n(u)} = G{p, T; an(u)} - G{p, T; n(u)} (3.6.A.9)

The right-hand side can be expressed as in equation (3.6.A.4):

(a -l)G{p, T; n(u)} = IlG{p, T; n(u)l(a -1)n(u)} + O[(a _1)2] (3.6.A.I0)

Dividing equation (3.6.A.10) by a-I and then taking the limit as a approaches
unity produces equation (3.6.20), with G '{ } being the functional derivative of
G{ } with respect to the argument function n (u). This is the continuous generali-
zation of the III Ilnj derivative which produces the partial molar properties.
Permanence of atoms places the following restriction on the stoichiometric
coefficient distribution:

f u(w, u)'Tp(u) du = 0 (3.6.A.ll)

for all wand all P, where 'Tp(u) is the number of P atoms in species u. The
molecular weight of species u is

M(u) = L 'Tp(u)Mp (3.6.A.12)


p

and the change in the number of moles of component u due to reaction w is

dn(u, w) = u(u, w) dx(w) (3.6.A.13)

The corresponding change of mass is

dm(w) = f M(u) dn(w, u) du = dx(w) f M(u)u(w, u) du

= dx(w) f [~'TP(U)Mp ]U(W, u) du

= dx(w) ~ [ Mpf 'Tp(u)u(w, u) dU] = 0 (3.6.A.14)


96 which guarantees that equation (3.6.4) is satisfied. However, the converse is not
true: equation (3.6.4) is satisfied provided the sum in equation (3.6.A.14) is zero,
Chapter Three
without need for every term in it to be zero. Thus permanence of atoms imposes
a stronger constraint than the one in equation (3.6.4).

EXAMPLES AND PROBLEMS

Examples

1. Consider a mixture of methane and ethane which is sent to a reactor where the
water shift reaction is to occur, for the production of hydrogen and carbon dioxide.
Oxidation, however, can also occur to the carbon monoxide form, so that one may start
by writing the following reactions:

CH. + H2 0 = CO + 3H2
CH. + 2H20 = CO 2 + 4H2
C2H6 + 2H 20 = 2CO + 5H 2
C2H6 + 4H 20 = CO2 + 7H2

By inspection of the stoichiometric coefficients matrix, one could now find out that
its rank is 3, i.e., that one of the reactions is a linear combination of the others. This is,
however, a tedious procedure. It is easier to realize immediately that the fourth reaction
does not introduce any new component, and hence it can be dropped.
The set of the first three reactions is not, however, the best one for setting up the
problem. In fact, even in a gaseous mixture in which there are no activity coefficients to
deal with, the equilibrium condition for a reaction results in a polynomial equation with
degree equal to the largest of the sums of all positive and all negative stoichiometric
coefficients; hence the three reactions result in polynomials of degree 4, 5, and 7, respec
tively. The system of equilibrium equations is thus equivalent to a polynomial of degree
4 x 5 x 7 = 140, for which even numerical techniques would be difficult to use to find the
only significant root. It takes a little bit of juggling around to find the following equivalent
but much simpler set:

1. CH. + H 20 = CO + 3H2
2. CO + H2 0 = CO 2 + H2
3. 2CH. = C2H6 + H2

which gives polynomials of degree 4, 2, and 2, and hence a system degree of 16.
There are six components, so that the composition space is five-dimensional. The
reaction subspace is three-dimensional. Let the components be numbered consecutively
as CH., C2H6 , H20, H2, CO, CO 2. Initially, there are M moles of ethane and R moles
of steam per mole of methane. The number of moles vector is

The solution has to be found under the constraint that all components of the number
of moles vector must be nonnegative.
Problems 97
Homogeneous
3.1. Consider a box of volume V,(t) in which reaction (3.1.1) takes place. The forward rate is Reactions
proportional to oxygen concentration, i.e.,

r F = k[02]

and equilibrium is reached when

[C0 2]/[CO]v'[02] = K


At time zero, there is one mole of CO, one mole of 2, and no CO2, Temperature is constant
in time.
a. Derive the equation for the reverse rate.
b. What is the stoichiometrically possible range of x values?
c. Draw the curve representing the reaction subspace in the triangular diagram representing the
composition space.
d. Derive the equation relating x' to x and V,.
e. Determine the relationship between x* and V,. Can more than one equilibrium point exist
at some value of V;?
f. Suppose the function V,(t) is assigned. How do you calculate x(t)? Now consider the case
where the initial composition is different from the one given above, but does, however, lie in the
reaction subspace you have calculated. Does x(t) at very large t become independent of the
composition at time zero?
3.2. Consider a chemical reaction

aA+ bB = pP+ qQ ...

for which the kinetics are of the standard mass action type, say the forward rate is given by

The net rate could be expressed as the rate of change of the extent of reaction x,

dx/dt=f(x), f(x*) =0

where x* is the equilibrium extent of reaction.


Which conditions should be satisfied if one wants the fIx) function to be linear, say fIx) =
k'(x* - x)?
3.3. Consider the following simple model for a predator prey population. Let P be the number
of predators and Q the number of prey. The following equation is written for the predator:

dP/dt = k,Q - k 2P

with the first term describing the higher probability of predators to reproduce when more prey is
available. For the prey, one writes

dQ/ dt = -k,P + k2Q + k.


with the three terms on the right representing the prey being devoured, the natural reproduction rate,
and a fixed supply of food.
a. Determine the steady-state values of P and Q.
b. Given an initial condition for both P and Q, will the solution of the differential equations
evolve toward the solution of part (a)?
98 c. Can the system be identified as a standard internal system, with P and Q the internal state
variables? In other words, are the given kinetic equations legitimate for a closed chemically reacting
Chapter Three system? Should the answer be no, does that imply that the given predator-prey model is forbidden
by the second law?
3A. If the M x M matrix in equation (3.6.6) is invertible, its kernel contains only the zero
vector and its rank is M. Now suppose it is not invertible, so that the rank is M - K The kernel will
now contain nonzero vectors, and in fact classes of them (each class containing vectors all proportional
to each other). How many such classes are there in the kernel? (Hint: The answer is K)
3.S. Consider the following functional:

F{f(u)} = f f2(u) du

and let the norm be given by equation (3.6.A.5). Calculate the first Frechet differential, verify that R
is o<llgll>, and calculate the functional derivative. If you have done the algebra correctly, the functional
derivative should be an ordinary function, not a functional. Why is that? Construct the simplest
functional you can think of such that its functional derivative is not an ordinary function.
3.6. Consider the following functional:

F{f(u)} = f2(a)
where a is a fixed number. Now consider two metrics: one with the norm as in equation (3.6.A.5)
and one where it is given by

Ilf(u)11 = f f2(u) du

Calculate the Frechet differential via equation (3.6.A.6). Is the residual R of o<llgll> with respect to
both topologies? Comment on your result.

LITERATURE

Most of the crackpot papers which are submitted to The


Physical Review are rejected not because it is impossible to
understand them, but because it is possible. Those which are
impossible to understand are usually published.
F. J. Dyson

The method of determining the rank of the stoichiometric matrix presents some
subtlety when different isomers may be present in the reacting mixture; see J. C. Whitwell
and R. S. Dartt, AIChEJ. 19, 1114 (1973).
Proof that the reaction subspace and the composition space coincide only in the case
of isomerization and polymerization reactions is given by G. Astarita, Chem. Eng. Sci. 31,
1224 (1976).
Direct coupling between different chemical reactions is generally excluded as a
possibility in the chemical literature; the point is discussed by M. J. Boudart, 1. Phys.
Chem. 87, 2286 (1983).
The implications of the strong constraint, that every individual chemical reaction
proceeds with a nonnegative dissipation rate on possible reaction pathways, are discussed
by R. Shinnar and C. A. Feng, Ind. Eng. Chem., Fundam. 24, 153 (1985).
There is ample literature on the mathematics of stoichiometry and mass action kinetics;
two important references are E. H. Kerner, Bull Math. Biophys. 34, 243 (1972) and M.
Feinberg, Arch. Ratl Mech. Anal. 49, 187 (1972).
The general case of linear kinetics is formalized by J. Wei and C. D. Prater, Adv. 99
Catal13,203 (1962).
Homogeneous
Point (e) of Problem 3.1 raises the possibility that there may be more than one
Reactions
equilibrium point in the reaction subspace. This is possible in nonideal mixtures; see H.
G. Othmer, Chern. Eng. Sci. 31, 993 (1976).
The whole question of independence of chemical reactions, the influence of
stoichiometry on the equilibrium behavior of reacting systems, and the stability of chemical
equilibrium was discussed in detail long ago by E. Jouguet, J. Ecole Poly., II Ser. 21, 61,
181 (1921); a more recent formal discussion was given by R. Aris, Arch. Ratl. Mech. Anal.
19, 81 (1965).
The general theory of homogeneous reacting mixtures, expressed in terms of internal
state variables, has an ample literature. Some important references are M. E. Gurtin and
A. S. Vargas, Arch. Ratl Mech. Anal. 43,179 (1971); B. D. Coleman and M. E. Gurtin,
J. Chem Phys. 47, 597 (1967); Phys. Fluids 10, 1454 (1967); C. A. Truesdell, Rational
Thermodynamics, 2nd ed., Chapter 5, Springer-Verlag, Berlin (1984).
The best source for the concept of affinity is Th. De Donder and P. Van Rysselberghe,
Thermodynamic Theory and Affinity, Stanford University Press (1936); the concept of
affinity, however, predates this by quite some time. The word "affinity" was used as far
back as 1888 by H. Le Chatelier, Les Equilibres Chimiques, Carre et Naud, Paris (1888),
and it was attributed to earlier workers such as Berthelot, if in derogatory terms.
The question of writing the stoichiometric equation for a reaction in the form of
equation (3.1.2), with stoichiometric coefficients which can be both positive and negative,
also has an old tradition; see, e.g., K. G. Denbigh, The Thermodynamics of the Steady
State, Methuen, Landon (1951).
This leads to the question of what is meant by the so-called "principle of microscopic
reversibility." This is discussed in the Denbigh book cited above, as well as by J. Wei, J.
Chem Phys. 36, 1578 (1962) and by G. Astarita, Quad. Ing. Chim. Ital. 14, 1 (1978).
The continuous description of mixtures was first developed by Th. De Donder,
L'affinite-Seconde Partie, Gauthier-Villars, Paris (1931). The presentation in Section 3.6,
as well as most modem analyses of continuous mixtures, is based on the methodology
discussed in the fundamental work of R. Aris and G. R. Gavalas, Phi/os. Trans. R. Soc.
London, Ser. A 260, 351 (1966); however, the second-law implications are based on the
methodology presented by Truesdell for the case of a finite number of reactions in the
book cited above. The question of substituting pseudocomponents for a mixture, by finding
the optimal quadrature points for the integrals appearing in a continuous description,
is discussed by S. K. Shibata, S. I. Sandler, and R. A. Behrens, Chern. Eng. Sci. 42, 1977
(1987).
Kinetic analyses for continuous mixtures (mostly linear kinetics) have been reviewed
by V. W. Weekman, Chern. Eng. Prog., Symp. Ser. No. 11, 75, 3 (1979) and by K. B.
Bischoff and P. Coxson, Proc. 2nd ICREC Meeting, Pune (1987). Some preliminary
analysis of the nonlinear case is discussed by G. Astarita and R. Ocone, AIChEJ. 39,1299
(1988) and by G. Astarita, AIChEJ. (to be published).
A good source for studying functional analysis is F. Riesz and B. Nagy, Functional
Analysis, Ungar, New York (1955).
Liapounov functions are discussed in all textbooks on stability analysis; a good source
is M. M. Denn. Stability of Reactions and Transport Processes, Prentice-Hall, Englewood
Cliffs, NJ (1975).
Problem 3.3 is a variation of the classical predator-prey model of A. J. Lotka, J. Am
Chem Soc. 42, 1595 (1920); Proc. Natl. Acad. Sci. U.S.A. 6, 410 (1920). In the Latka
formulation the kinetic equations are
Chapter Four

PHASES

A good many times I have been present at gatherings of


people who by the standards of the traditional culture
are thought highly educated and who have with
considerable gusto been expressing their incredulity at
the illiteracy of scientists. Once or twice I have been
provoked and have asked the company how many of
them could describe the second law of thermodynamics.
The response was cold; it was also negative. Yet I was
asking something which is about the scientific equivalent
of: "Have you read a work of Shakespeare's?"
C. P. Snow

NOTATION

A Free energy density kcalkmol- '


a( Constitutive function for A kcal kmol- '
cp Constant pressure specific heat kcal kmol- ' K- '
D Difference operator
G Free enthalpy density kcal kmol- '
G o( Arbitrary distribution kcalkmol- '
Go Arbitrary constant vector kcalkmol- '
G{ Constitutive functional for free enthalpy kcal
GM1X Constitutive functional for mixing free enthalpy kcalkmol- '
k Thermal conductivity kcal m- I S-I K- '
Position of freeze surface m
L Any extensive property
m Mass kg
M Number of independent chemical reactions
n Number of moles vector kmol
n( Number of moles distribution kmol
N Number of components
N' Mass transfer vector kmol S-I 101
102 p Pressure atm
Chapter Four p* Equilibrium phase transition pressure atm
S Entropy density kcal kmol K- 1
Time s
T Temperature
V Specific volume
x Internal state variable
x Extent of reaction vector kmol
X Internal state variable
y Distance from exposed surface m
y( Mole fraction distribution
y Mole fraction vector
z Mass fraction of phase 1
U (l/V)6V/6T K- 1
UK Amount of Kth mixture (fraction of total moles)
f3 -(l/V)6V/6p atm- I
T 6V/6x
T' 6S/6x
6( Dirac delta function
(J 6G/6x
6 Affinity vector kcal kmol- 2
/.I. Chemical potential vector kcalkmol- I
(J' Stoichiometric coefficients matrix
fl( Heaviside step function

Subscripts

p Baric derivative with respect to pressure


Total
T Baric derivative with respect to temperature
X Baric derivative with respect to internal
state variable

Superscripts

T Transpose
1,2 For phase 1,2
* At equilibrium
Substantial time derivative
MIX Of mixing
Partial molar
4.1. ONE-COMPONENT,
TWO-PHASE SYSTEMS

Believe nothing, merely because you have been told it,


or because it is traditiona~
or because you have imagined it.
Gutarna Buddha

A phase is defined as a part of a body within which the local state is a smooth
function of position (the gradient of any state variable is finite at all points). A
multiphase system is therefore a body which contains surfaces of discontinuity
for the state; such surfaces are called interfaces. The principle of action and
reaction forbids pressure to be discontinuous at interfaces, and so pressure is
continuous even in multiphase systems, except for the effect of surface tension
on curved interfaces. There is no fundamental law of physics which forbids
temperature discontinuities, but in heat transfer theory the usual assumption of
a continuous temperature distribution has been generally successful, and therefore
that assumption is retained here (see the literature section in this regard).
In this chapter, attention is focused essentially on the equilibrium behavior
of multi phase systems, and therefore an additional simplification is possible,
namely, that the state is constant in space within a phase. This in tum implies
that temperature and pressure are constant throughout a multiphase system. This
simplification is certainly valid at equilibrium (except, as far as pressure is
concerned, for the hydrostatic pressure distribution), and therefore the equili-
brium equations derived below are valid. However, the nonequilibrium results
should be regarded as only indicative of trends.
If attention is restricted to a one-component elastic system, the only state
variable is the local specific volume, and all functions of state can be regarded
as uniquely determined by the value of Y; in particular

A = a(Y) (4.1.1)

Since pressure is the same in the two phases, the same is true of the derivative
da/ dY. It follows that two phases may exist only if the a( Y) function has at
least two distinct points with equal derivative. This in tum implies that the second
derivative d 2a/ dy2 must change sign at least once. In fact, two phases will exist
under conditions where d 2 a/ dy2 changes from positive at low Y to negative and
then to positive again, as discussed below. 103
104 Let us consider the a (Y) curve in Figure 4.1.1. The derivative is always
negative, so that pressure is always positive (in actual fact, a negative pressure
Chapter Four
is not a physical impossibility if the system considered has a finite tensile strength).
The following equation holds along the entire curve:

(4.1.2)

It follows that, in the "spinodal" region, i.e., in the region where d 2al dy2 is
negative, pressure increases with increasing volume. It is easy to convince oneself
that this is a mechanically unstable condition.
Suppose the system considered exists in a two-phase condition, with the two
phases identified by points 1 and 2 in Figure 4.1.1; dal dY is indeed the same at
these two points. Let a fraction z of the total mass be at condition 1, and a
fraction 1 - z at condition 2. The total mass of the system, m" and the total
volume, Yo, are fixed. The following equation is obvious:
V;lm t = zyl + (1- z) y2 (4.1.3)

Equation (4.1.3), and the condition that dal dY be the same at points 1 and
2, poses two constraints on the values of z, Y I , and Y2 It follows that both YI
and Y2 are determined once z is assigned. Thus, by differentiating equation (4.1.3)
with respect to z one obtains
(4.1.4)

We now consider the total free energy of the system, which is given by

(4.1.5)

Differentiation of equation (4.1.5) and substitution of equation (4.1.4) yields

(4.1.6)

A
SPINODAL
REGION

FIGURE 4.1.1. Isothermal free energy vs. volume


V4 v plot for a single-component system.
The system considered can be identified as an internal one, with z playing 105
the role of the internal state variable, hence aatl az being the affinity. It follows that Phases
(aa tlaz)z':5 0 (4.1.7)

For points 1 and 2, it is easy to convince oneself that the right-hand side of
equation (4.1.6) is negative. It follows from equation (4.1.7) that z will tend to
increase, i.e., that the two points will tend to move to the right. Equilibrium is
of course reached when the two points have moved to points 3 and 4, the only
couple of points for which the right-hand side of equation (4.1.6) is zero.
Whenever the average specific volume of the system, V,I mt , is in the interval
y3 _ y\ the system at equilibrium will exist in a two-phase condition, with y3
and y4 being the specific volumes of the two phases. The equilibrium value of
z is determined by the lever rule, say it is calculated from equation (4.1.3). We
note that both y3 and y 4 lie outside the spinodal region, since d 2al dy2 is positive
at both of them. An equilibrium two-phase system can only exist if the a( Y)
curve has a spinodal region, but the specific volumes of the two phases at
equilibrium are not spinodal points.
If Figure 4.1.1 is regarded as relative to a liquid-vapor system, the branches
of the curve are related to the branches of the curve in Figure 2.6.1; indeed, the
latter is nothing else than minus the derivative of the former. For Y values up
to y3, the curve is an equilibrium curve, with z = 1 (pure liquid). For Y values
between y3 and the first spinodal point, the curve is the nonequilibrium curve
of a superheated liquid; for values between y4 and the second spinodal point,
it is the nonequilibrium curve of a supercooled vapor. Finally, at values from y4
upward, the curve is the equilibrium curve for a gas (z = 0).
The right-hand side of equation (4.1.6) is the difference between the free
enthalpies of the two phases. It follows that, at equilibrium, the free enthalpies
of the two phases are equal. Pressure and free enthalpy are the only two constitu-
tive properties which, at equilibrium, are equal in the two phases .
. Now consider a one-component, one-phase system at some fixed pressure.
Since the baric derivative of free enthalpy with respect to temperature is minus
the entropy, it is intrinsically negative, i.e., the free enthalpy decreases with
increasing temperature (see Figure 4.1.2). At any given temperature, the equili-
brium phase is the one endowed with the lower free enthalpy. It follows that, as
temperature is increased, a phase transition will occur when the two curves

PH SE I

FIGURE 4.1.2. Isobaric free enthalpy YS. tern


perature plot for a single-component system.
106 representing two different phases cross each other; the phase which at equilibrium
exists at the higher temperature is the one with the larger entropy, as required
Chapter Four
by the fact that the curves, both of which have negative slope, must cross each
other. The symbol DL will be used to represent the difference between the value
of L in the high entropy phase and that in the low entropy one at the phase
transition point, so that DS > 0 by definition. By making use of the appropriate
Maxwell relations one obtains

DG=Dp=O (4.1.8)

DH= TDS> 0 (4.1.9)

DU=DH-pDV (4.1.10)

DA=pDV (4.1.11)

Generally, DV is also positive (it is always positive for a liquid-vapor phase


transition). However, this is not universally true. For instance, for the ice-water
transition, DV is negative, since the phase with a larger density, water, also has
a larger enthalpy.
Of course, the a( V) curve may present more than one spinodal region, and
in fact it often does. Given a curve with two spinodal regions, in general one
can identify two separate phase transitions; see Figure 4.1.3. However, as tem-
perature changes, so does the shape of the a( V) curve, and at one particular
temperature it may have the shape in Figure 4.1.4, with the low-volume equilibrium
point of the second transition coinciding with the high-volume one of the first
transition. At that particular temperature, the equilibrium condition may be the
simultaneous existence of three phases. The temperature at which this occurs,
and the pressure corresponding to the tangent passing through all three equili-
brium points, form a pair which is called the triple point (reference being made
to a point in a pressure-temperature plot).

FIGURE 4.1.3. Free energy vs. volume plot


showing two successive phase transitions of a
V single-component system.
At any given temperature, the equilibrium pressure of the two-phase system 107
is uniquely determined:
Phases
p* = p*(T) (4.1.12)

(in the case of a liquid-vapor transition, p* is called the vapor pressure of the
liquid). Let us consider a differential change in temperature dT. If the system is
to remain in an equilibrium two-phase condition, the corresponding dp is given
by (dp* / dT) dT, and hence the free enthalpy change in both phases can be
written as
dO = [50/ ilT + (50/5p) dp*/ dT] dT
= [-8 + Vdp*/dT] dT (4.1.13)

But the dO for the two phases must be the same, since they are kept in
equilibrium with each other. It follows that

dp*/dT = D8/ DV = DH/(TDV) (4.1.14)

which is called the Clausius-Clapeyron equation. For those transitions (like ice
melting) for which DV is negative, so is dp* / dT: at high pressures, ice melts at
temperatures lower than OC. For vapor-liquid transitions, the vapor pressure
always increases with increasing temperature (see Section 4.5 for a more detailed
discussion of this point).
At very low temperatures (lower than the triple-point temperature), the a( V)
curve may present one or more spinodal regions corresponding to solid-solid
transitions. These are followed by a spinodal region corresponding to a solid-gas
transition, which is called a sublimation. At temperatures exceeding the triple-
point temperature, a solid-liquid transition occurs between the last solid-solid
one and the liquid-vapor one. (It will be seen in Section 4.4 that liquid-liquid
transitions are sometimes observed as well.) As the temperature is further
increased, the spinodal region of the liquid-vapor transition becomes progress-
ively smaller (DV decreases with increasing temperature) until a temperature is

FIGURE 4.1.4. Free energy vs. volume plot at


the triple-point temperature for a single-
component system. v
108 reached where d 2a/ dy2 is always positive except at one particular point, where
it is zero (that is the point to which the spinodal region has shrunk). The
Chapter Four
temperature at which this occurs is called the critical temperature. The Y value
for which d 2a/ dy2 is zero is called the critical volume, and the pressure at that
point is called the critical pressure. At temperatures exceeding the critical tem-
perature, no liquid-gas transition occurs. Systems at temperatures slightly in
excess of the critical temperature, and with volumes not very different from the
critical volume, are said to be in a supercritical state.
It is worth noting that, as far as liquids and vapors are concerned, one
normally assumes that two-phase equilibria are reached almost instantaneously.
This is not strictly true, since superheated liquids and supercooled vapors do
occur in nature-evolution to the equilibrium state requires nucleation of the
second phase. However the assumption is a reasonable one, since when nucleation
does occur the evolution toward the equilibrium state is very fast. The situation
is completely different for solids, which can be quenched to a low temperature
fast enough that some solid-solid phase transitions do not occur at all. Such
solids may then exist practically indefinitely in a nonequilibrium condition, i.e.,
in a solid phase other than the equilibrium one corresponding to low temperature.
Metallurgy makes major use of such quenching techniques. In addition to this,
glasses are nonequilibrium phases which transform to the equilibrium (crystalline)
form over time scales of the order of hundreds of years.

4.2. PHASE EQUILIBRIA

Ask a scientist what he conceives the scientific method to be,


and he will adopt an expression that is at once solemn and
shifteyed: solemn, because he feels he ought to declare an
opinion; shifteyed, because he is wondering how to conceal the
'fact that he has no opinion to declare. If taullled he would
probably mumble something about "Induction" and
"Establishing the Laws of Nature," but if anyone
working in a laboratory professed to be trying to establish
the laws of nature by induction we should begin to
think he was overdue for leave.
P. B. Medewar

In this section, we consider the case of a mixture in which no chemical


reactions may occur, but which may exist in a two-phase condition. Again,
pressure and temperature are the same in the two phases, since the arguments
of the preceding section apply again.
The two phases will be identified by superscripts 1 and 2, since numerical
subscripts are used to identify components. Given any extensive quantity L., its
value is simply the sum of the values for the two phases:

(4.2.1)

If this is applied in particular to the free energy, its rate of change is


obtained as
(4.1.1)
and the second law reduces [after applying equation (4.2.1) to eritropy as well] to 109
(4.2.3)
Phases

The usual argument shows again that the first two terms in equation (4.2.3)
are identically zero. Furthermore, we note that, if no chemical reactions occur,
the only way for n' to change in time is by transferring components to or from
the other phase. Hence
(4.2.4)

so that the second law reduces to

(4.2.5)

Equation (4.2.5) imposes a restriction on the mass transfer vector n'. This
equation does not require that mass transfer of each component individually
should be toward the phase for which the chemical potential is lower, but only
that the scalar product of the affinity (or driving force) vector ... ' _ ... 2 with the
mass transfer vector should be nonpositive.
The vector n' can be regarded as the internal state variable, and therefore
the driving force vector is the affinity. Again under sufficient assumptions of
smoothness near equilibrium [perfectly analogous to those leading to equation
(3.2.8)], one concludes that the equilibrium affinity is zero, i.e.,

(4.2.6)

Again, in principle equation (4.2.6) exhausts the subject of multi component


phase equilibria. In practice, in order to calculate the composition of the two
phases at equilibrium, constitutive equations relating the chemical potentials to
the composition need to be written down.
Equation (4.2.5) implies that, of all two-phase states corresponding to assig-
ned values of temperature and pressure, the equilibrium state has the lowest total
free enthalpy. Now consider the special case of a two-component system, for
which the composition space is one-dimensional, since only the mole fraction of
one component need be known. Suppose the system is in a one-phase state and,
in analogy with the discussion in Section 3.3, we consider a collection of such
systems, all characterized by the same value of the total number of moles, n (such
a collection forms a one-dimensional space). If two systems within the collection
are considered with y, values differing only by a differential dy, , the corresponding
difference of any extensive quantity L is given by

(4.2.7)

since
(4.2.8)

Hence, since the system is in a one-phase state, we calculate the following


baric derivative:
(4.2.9)
110 or, since the density L is simply 1-, divided by the constant value n,
Chapter Four
(4.2.10)

Now suppose that, as long as the system is in a one-phase state, the free
enthalpy density as a function of YI is as plotted in Figure 4.2.1. Since free
enthalpy is only defined to within an arbitrary additive quantity which need not
be the same for two different one-component systems, the curve in Figure 4.2.1
is to be understood as significant to within addition of an arbitrary straight line
(or, in other words, the values of G atYI = oand atYl = 1 are arbitrary). Therefore,
any conclusion drawn from the figure must be invariant with respect to such an
additive straight line. For instance, statements about the curvature are invariant.
The statement that the slope is equal at two points is invariant, but the actual
value of the slope is not.
Now let us assume that, for any reason whatsoever, the system unmixes to
a two-phase state. The free enthalpy of each phase will be a point on the curve,
but the average free enthalpy density of the two-phase system will not, and it
will be given by the usual lever rule construction from the points representing
the two phases and the assigned average value of YI. This leads, for example, to
the average free enthalpy density at point A for a two-phase condition character-
ized by points Band C and the average mole fraction D. We note that, in the
interval of YI values over which the curvature of the G curve is negative, this
construction leads to a free enthalpy of the unmixed system lower than that of
the one-phase state, even if the two phases have compositions which are only
differentially diverse from each other. This means that even an infinitesimally
small unmixing (which can always occur simply by Brownian motion) will result
in a decrease of free enthalpy. It follows that the one-phase state is locally
unstable to unmixing (the term '''diffusional instability" will- be used in the
following), in the sense that it is unstable to infinitesimal disturbances.
The two points of inflection of the G curve are again !lalled spinodal points,
and the region between them is called the spinodal region. Consider now point
E, which is outside the spinodal region. It represents a locally stable one-phase

SPINODAL
REGION

FIGU~ 4.:1.1. Free enthalpy vs. composition


o plot for a two-component mixture exhibiting a
o miscibility gap.
condition. However, it does not represent a globally stable condition, since 111
two-phase states, characterized by Yl values differing by a finite amount, do exist
Phases
which would give an average free enthalpy lower than that at point E. In fact,
all the points on the curve between points F and H have this property, and in
fact a two-phase system identified by points F and H results in the lowest possible
free enthalpy for all such points. It follows that, at equilibrium, whenever the
average value of Yl is intermediate between the two values corresponding to
points F and H, the system will be in the F/H two-phase state. It is noteworthy
that the equilibrium condition is satisfied at points F and H. In fact, if equation
(4.2.10) is applied to the free enthalpy, one obtains

(4.2.11)

The left-hand side of equation (4.2.11) is simply the slope of the G curve,
while the right-hand side must have the same value in the two phases at equili-
brium. Of course, couples of points with equal slope other than F and H do exist,
but only the pair F/H is such that not only is the difference between the two
chemical potentials equal, but also their individual values.
Points on the G curve outside the spinodal region but inside the F I H interval
are metastable. If the system exists in a metastable one-phase state, in order to
unmix toward the equilibrium state the second phase needs to nucleate, i.e., a
finite disturbance is needed. Nucleation is often the governing step in unmixing
phenomena, and in fact it normally would not take place at all, or with great
difficulty, in the absence of nucleation sites (microscopic solid grains, irregularities
ofthe surface of a bounding solid, and so on). In contrast with this, points within
the spinodal region can only exist in nature long enough to be observable if
growth of nuclei is sufficiently slow (as is the case in solids), since nucleation
would occur spontaneously under the influence of Brownian motion.
It is comparatively easy to generalize the analysis given above to the case
of a mixture with an arbitrary number of components. Let Go be the vector whose
components are the free enthalpy densities of the pure components. Should free
enthalpy be additive, the free enthalpy density of a mixture with composition
identified by the mole fraction vector y would be given by

G = yG o (4.2.12)

Since free enthalpy is defined to within an arbitrary additive constant for


each pure component, the vector Go is entirely arbitrary, and equation (4.2.12)
simply states the fact that, should free enthalpy be additive, it would be a linear
function of y. In actual fact free enthalpy is not additive; it can be expressed in
the form

(4.2.13)

where the quantity G M1X is in principle accessible to measurement. Since Go is


arbitrary, any statement concerning the function G( ), in order to make sense,
should be invariant under addition of an arbitrary linear function of y [i.e., it
must be formulated as a statement concerning the G M1X ( ) function].
112 The condition of complete miscibility in all proportions can be determined
Chapter Four
as follows. Consider any number of mixtures, each one in an amount of a K moles
and characterized by a mole fraction YK, with

(4.2.14)

The free enthalpy for the system consisting of such separated mixtures is
simply the sum of their free enthalpies, and is thus obtained by the multi-
dimensional generalization of the lever rule:

(4.2.15)

If all sucli mixtures are mixed together, the resulting mole fraction vector is

(4.2.16)

and hence complete miscibility will occur if

(4.2.17)

A function which satisfies equation (4.2.17) over a connected domain ofYK


values is said to be convex over that domain; if the left-hand side of equation
(4.2.17) is smaller than the right-hand side, the function is said to be concave
(the properties of concaveness and convexness are invariant under addition of
an arbitrary linear function). Hence one concludes that the condition for complete
miscibility is that the free enthalpy density function is convex over the whole
composition space.
If the free enthalpy function iii concave over a region of the composition
space, that region constitutes the spinodal region. The existence of a spinodal
region guarantees the existence of a (larger) region over which phase separation
will occur at equilibrium. The number of phases which wili be observed ilt
eqUilibrium may well be more than two. Indeed, consider a system for which the
average mole fraction vector falls into the multiphase equilibrium region. The
average mole fraction can be decomposed in infinitely many ways in the form
of equation (4.2.16). Of all such decompositions, the one which will be observed
at equilibrium is that resulting in the lowest free enthalpy as calculated from
equation (4.2.15).

Appendix

It takes some tedious but straightforward algebra to extend the results in


this section to a mixture described by continuous distribution functions. The
fundamental result for two-phase systems is the obvious analog of equation
(4.2.6), i.e., the chemical potential distribution is the same in both phases. It
should be obvious that this does not imply that the concentration distribution at
equilibrium has the same shape in the two phases; however, this point has been
overlooked more than once in the literature (see Section 9.5 for a more detailed
discussion of this point).
The free enthalpy per mole of a mixture is given by 113

G f n(u) du = G{p, T; n(u)} (4.2.A.l)


Phases

In view of the property expressed by equation (3.6.19), this reduces to

G = G{p, T; y(u)} (4.2.A.2)

where the y(u) functions are of course restricted by the requirement that

f y (U)dU=1 (4.2.A.3)

The free enthalpy of mixing is obtained as

G = J Go(u)y(u) du + GMIX{p, T; y(u)} (4.2.A.4)

In a phase where no chemical reactions occur, Go( u) is any arbitrary function.


The condition of complete miscibility in equation (4.2.17) becomes

(4.2.A.S)

for any set of Cl!K satisfying equation (4.2.14); i.e., the functional G MIX{ } has
to be convex over the space of argument functions satisfying equation (4.2.A.3).

4.3. HETEROGENEOUS CHEMICAL EQUILIBRIA

As society gradually recesses toward a level of communication


last utilized in the Pleistocene, those means of expression still
remaining to us assume increasing importance. One of these is
the much-maligned clichl!. Once scorned as the verbal
equivalent of TV dinners, these universally understood,
prefabricated constructions enable vast numbers of individuals
to communicate without having to undergo the painful
experience of attempting to produce an original thought.
L. A. Abel

In this section, two-phase systems in which chemical reactions may occur


are considered. Again attention is focused mainly on the equilibrium behavior,
and therefore pressure and temperature are assumed to be constant throughout
the system considered.
In two-phase systems, the actual composition at anyone time is not con-
strained to be in the reaction subspace. This is due to the fact that, for each one
of the phases, the number of moles vector may change in time by mass transfer
exchange with the other phase; it follows that the entire composition space is
accessible in such systems, whatever the initial composition may be.
114 It is useful to introduce some nomenclature at this stage. Let 0' represent
the stoichiometric coefficient matrix, and x the extent of reaction vector, so that
Chapter Four
equation (3.1.9) can be written as

net) = nCO) + O'x (4.3.1)

We note that the matrix multiplication in equation (4.3.1) is between an


N x M matrix and an M x 1 matrix whose elements are the components of x,
since in equation (4.3.1) the summation is over the reaction index. The 0' matrix
also appears in equation (3.4.1), but here summation over the component index
is intended, and hence equation (3.4.2) can be written as

(4.3.2)

and the multiplication involved is that of the M x N matrix 0' T and the N x 1
matrix of the components of .....
The second law reduces again to equation (4.2.3), and the usual argument
shows that the first two terms are identically zero, so that

(4.3.3)

Both number of moles vectors can change in time, as a consequence of the


chemical reactions and also because of mass transfer. Let N be the mass transfer
vector from phase 2 to phase 1. Thus

(4.3.4)

and
(4.3.5)

Substitution of the latter two equations into equation (4.3.3) yields

(4.3.6)

By making use of the properties of transpose matrices, equation '(4.3.6) can be


expressed as follows, after substitution of equation (4.3.2):

(4.3.7)

There are in fact 2M + N terms on the left-hand side of equation (4.3.7)


which are not required by the second law to be individually nonpositive (only
their sum is). Indeed, in the theory of mass transfer with chemical reaction,
instances occur where the mass transfer of one component is toward the phase
where it has a higher chemical potential.
The mass transfer vector can be identified from equation (4.3.4) as the time
derivative of a vector:
(4.3.8)
It follows that the left-hand side of equation (4.3.7) can be identified as the scalar 115
product of two vectors, one of which is the time derivative of another vector.
Phases
Therefore the internal state variable for the system considered is the vector

(4.3.9)

and, correspondingly, the affinity is

(4.3.10)

with 0 1 and 02 being the chemical affinities in the two phases.


Again, under sufficient assumptions of smoothness near equilibrium, one
concludes that equation (4.3.7) implies that the equilibrium affinity is zero. This
means that, at equilibrium, the chemical affinities must be zero in both phases,
and the chemical potential of each component must be the same in the two
phases. In other words, the equilibrium conditions for the chemical reactions
and for phase equilibrium hold jointly. It takes some very tedious algebra to
show that the same result holds also for continuous mixtures.
The occurrence of heterogeneous chemical equilibria is crucially related to
the fact that the reaction subspace, but for exceptional cases, has fewer dimensions
than the composition space. This point is best illustrated by considering a system
where only one reaction takes place, so that the reaction subspace is simply
o :5 x :5 1 (it is always easy to normalize the definition of x so that its values
range in that interval). If the reaction considered is an isomerization, and the
two isomers are the only components present, the reaction subspace coincides
with the composition space. Now consider the case where the free enthalpy in
the reaction subspace has the shape as in Figure 4.3.1. Since a chemical reaction
may transform one isomer into the other one, the difference between the free
enthalpies of the two pure isomers is measurable (at least in principle), and
therefore both are known to within the same arbitrary additive constant. It follows
that while the absolute position of anyone point of the curve in Figure 4.3.1 has
no objective meaning, the slope of the curve has.
Now suppose that the system is initially at x = O. The reaction would take
place until the free enthalpy reaches a local minimum, say x reaches the value
Xl. We note, however, that the system at x = Xl is globally unstable to unmixing,
and thus a second phase could nucleate at point X2. However, X2 is not a local

FIGURE 4.3.1. Free enthalpy vs. extent of reaction


plot for an isomerization reaction.
116 minimum for free enthalpy, and therefore the reaction would start proceeding
forward toward X3 in the nucleated phase. The phase at X3 is not in physical
Chapter Four
equilibrium with that at Xl, and thus mass transfer would occur, until phase 1
completely disappears. The final equilibrium condition is at point X3, which is
stable to unmixing. This shows that, in a system for which the reaction subspace
coincides with the composition space, stable equilibrium will invariably
correspond to a one-phase condition.
However, let us consider the slightly more complex case of a single chemical
reaction taking place in a three-component mixture (say A + B = C, with no
other components present). The composition space is the area in the triangle in
Figure 4.3.2. The average composition of the system is of course restricted to the
reaction subspace, say to the vertical straight line. However, any point in the
reaction subspace may be unstable to unmixing in the composition space, provided
the direction of unmixing forms a finite angle with the vertical straight line. The
final equilibrium condition may therefore well be a two-phase one, with the two
phases being at, say, points A and B. These two points must satisfy the following
criteria:

1. They must both lie on the two-phase envelope curve, and be on the same
tie line (i.e., they are in physical equilibrium with each other).
2. The tie line connecting them must cross the vertical line, since the average
composition of the system is restricted to that line.
3. Of all tie lines satisfying (1) and (2), the equilibrium one corresponds to
the lowest total free enthalpy, as calculated from the lever rule at the
intersection with the vertical line.

Again, it is easy to extend the analysis to mixtures with an arbitrary number


N of components within which an arbitrary number M N) of independent
reactions may occur. In such a mixture, the vector Go defined in Section 4.2 has.
N components; however, only N - M of those are arbitrary, since the standard
free enthalpy change of each of the independent reactions is in principle accessible
to measurement. This implies that, over the restricted reaction subspace, the
G( ) function is defined to within only one additive constant, and therefore it
makes sense to identify local minima of that function in the reaction subspace.

FIGURE 4.3.2. Composition space for a


system exhibiting heterogeneous chemical
equilibrium.
The geometry of the problem is best understood by considering again Figure 117
4.3.2. Over the two-dimensional triangular diagram, a linear function of y is a
Phases
plane, and hence if there were no reactions the G( ) function would be deter-
mined to within addition of an arbitrary plane. However, knowledge of the
standard free enthalpy change of the reaction assigns the slope of the intersection
of that plane with the vertical plane rising from the reaction subspace line. Thus,
to within one arbitrary additive constant, the arbitrary plane is restrained to pivot
around a line in that vertical plane.
Now suppose that more than one local minimum exists within the reaction
subspace. If the composition reaches a local minimum which is not the lowest
one of all such minima, nucleation of a phase corresponding to a lower local
minimum may occur. However, the nucleated phase is not at chemical equilibrium,
and hence the chemical reaction will proceed within it, with mass transfer of the
reactants occurring from the other phase. Final equilbrium will be reached only
when the lowest minimum is reached, which corresponds to a one-phase condition.
The considerations in the preceding paragraph, however, only apply if the
system is restricted to stay in the reaction subspace; such a restriction would
hold only in the case where M = N - 1, i.e., only for isomerization or polymeriz-
ation reactions. In all other cases, unmixing into two or more phases may occur
in some direction in composition space which forms a finite angle with the reaction
subspace, alld hence heterogeneous equilibria will in fact be observed.
The above discussion clarifies the statement made in the preceding chapter,
that in mixtures which are miscible in all proportions there can be only one
equilibrium composition. In fact, if the G( ) function is convex over the whole
composition space, it can, with any legitimate choice of the arbitrary components
of Go, have only one minimum over the whole composition space, and hence a
fortiori over the reaction subspace.

4.4. LIQUID CRYSTALS

The converse (of heat) cause produces crysta~ which is nothing


but water frozen by excessive cold; one never finds it except
where snow is frozen into ice: and so it is certain it is indeed
ice, as witnessed by its Greek hame.
Pliny the Elder, Natural History

A large number of organic substances, characterized by strongly anisotropic


molecules, may form phases which are called liquid crystals. A liquid crystal may
flow (though not quite as an ordinary liquid) and yet it possesses a certain degree
of crystallinity, as evidenced by X-ray diffraction patterns. In a liquid crystal,
some degree of long-distance alignment of molecules is present.
There are several types of liquid crystals. These are classified into three broad
categories: nematic (N), cholesteric (C), and smectic (S); the smectic crystals
can in tum be subclassified in a variety of special crystalline structures (SA, SB,
etc.). Nematic liquid crystals are the best understood ones; in their phenomeno-
logical description, the local state is characterized also by a unit vector (which
may be interpreted as the local orientation of molecules).
118 Single-component liquid crystals are usually referred to as thermotropic,
since the usual way of inducing phase transitions in such systems is by changing
Chapter Four
temperature. In many cases, the crystalline solid melts into a liquid crystal at
some temperature; at higher temperatures phase transitions between different
liquid crystals may occur, until at some high temperature a transition to an
ordinary isotropic liquid phase occurs. A few examples are given below.

4-n-pentyl-cyanobiphenyl: solid-N-isotropic
Cholesteryl benzoate: solid-C-isotropic
Ethyl-p-azoxycinnamate: solid-SA-isotropic
Cholesteryl myristate: solid-SA-C-isotropic
Terephthal-bis-4-n-butylaniline: solid-SB-SC-SA-N-isotropic

For all the examples given above, the phase transitions are "reversible" in
the same sense as ordinary phase transitions are: the indicated sequence is traveled
from left to right on heating, and from right to left on cooling; however, supercool-
ing is commonly observed, particularly in the liquid crystal-solid transition. Such
reversible behavior is referred to in the specialized literature as enantiotropic.
For other one-component systems, a behavior is observed which is called
monotropic. The typical example is cholesteryl nonanoate (CN). Solid CN melts
into a cholesteric liquid crystal at 78.6 DC; transition to the isotropic form occurs
at 91.2 DC. Upon cooling at a sufficiently slow rate, the isotropic liquid transforms
to the cholesteric crystal at 91.2 cC. However, when the C form is cooled, no
matter how slowly, it does not solidify at 78.6 DC, and undercooling occurs down
to a temperature of 75.5 DC, when a transition to a smectic A liquid crystal is
observed. The solid can only be obtained by further cooling the smectic phase.
The behavior of monotropic systems can be understood with reference to
Figure 4.4.1. The uppermost curve represents the free energy function at some
high temperature where two ordinary phase transitions would be observed. As
the temperature is lowered, the intermediate valley moves up with respect to the

FIGURE 4.4.1. Free energy vs. volume plots at


three different temperatures. The lowest curve
suggests an explanation for monotropic behavior
V of liquid crystals.
other two, until a triple-point temperature is reached as shown in the middle 119
curve. At an even lower temperature, the shape of the curve is likely to be as
Phases
shown in the lowest curve. Consideration ofthe latter curve helps in understanding
monotropic behavior, as discussed below.
An undercooled cholesteric liquid crystal may not solidify because nucleation
of the solid form does not take place, due to the long-range order of the cholesteric
form. As the temperature is further decreased, the smectic form may nucleate-and
a C-S phase transition occurs, as identified by the straight line marked A. This
corresponds to a decrease of free energy, though not as large as would correspond
to solidification. We note that there is a finite range of V values where the
cholesteric phase is globally unstable to solidification, but is stable to nucleatidn
of the smectic phase: this explains why there is an interval of temperatures where
the supercooled cholesteric crystal is stable. There is also a finite range of V
values in which the smectic phase is stable to solidification, though globally
unstable to a cholesteric-solid two-phase condition. Finally, at some lower
temperature the smectic phase is globally unstable to solidification, and the
freezing phase transition may occur by nucleation of the solid from the smectic
phase., On heating, the solid undergoes directly the true equilibrium transition
to the cholesteric phase.
Liquid crystals are also possible in mixtures; in this case, the simplest way
to induce phase transitions is by changing the concentration, and such systems
are called lyotropic. The main area of interest for lyotropic liquid crystals is that
of polymers; liquid-crystalline polymers are easily processed because of their
comparatively low viscosity, and the finished product has a very high degree of
orientation and hence extremely good mechanical properties. Liquid-crystalline
polymers are often processed as solutions, since in their pure form they cannot
be made to melt (they degrade thermally at temperatures lower than the melting
point).

4.5. SPECIAL TRANSmONS

LORD, when thou wentest out of Seir, when thou marchedst


out of the field of Edam, the eanh trembled, and
the heavens dropped, the clouds also dropped water.
The mountains melted from before the LORD,
even that Sinai from before the LORD God of Israel.
Judges 5, 4-5

Some systems, particularly in the solid state, exhibit phenomena which can
be regarded in some sense as phase transitions, but are however quite different
from the ordinary ones. Such phenomena have been called higher-order phase
transitions, lambda transitions, and so on. Theoretical analyses of such transitions
have been presented in the literature, and these can be broadly classified in three
fundamental categories: theories of n-order phase transitions, nonequilibrium
theories, and supercritical theories. A brief discussion of the whole subject is
given in this section.
Let us consider an ordinary phase transition of a one-component system. At
equilibrium, pressure, temperature, and chemical potential are continuous
120 through the phase transition, and one can therefore regard the equilibrium free
enthalpy G* as a function of T and p. At constant pressure, G* is a continuous
Chapter Four
function of temperature, but its first derivative (which is minus the entropy)
suffers a discontinuity at the phase transition. Analogously, G* is, at constant
temperature, a continuous function of p, but its first derivative (the volume)
suffers a discontinuity at the phase transition. For the special phase transitions
considered in this section, volume does not suffer a discontinuity, but the second
derivative of the free enthalpy with respect to volume, namely

(4.5.1)

is discontinuous. The question of whether entropy does or does not suffer a


discontinuity is a moot point, from both the experimental and theoretical view-
point.
Theories of n-order phase transitions are basically as follows. Analogously
to volume, the assumption is made that entropy also does not suffer a discon-
tinuity, but the second derivative of the chemical potential with respect to
temperature does:

(4.5.2)

We note that, if entropy is not discontinuous, neither is enthalpy, and hence


the latent heat of a transition such as that described above is zero.
There are only two well-established instances of transitions which indeed
show no discontinuity of either volume or enthalpy: the rubber-glass transition
in polymers, and the onset of superconductivity in a zero magnetic field. Few as
such instances may appear to be, a theory of such phenomena is clearly desirable.
The first step of an n-order type theory is to establish the equivalent of the
Clausius-Clapeyron equation, i.e., a relationship for p*( T), where p* is the
pressure of the equilibrium two-phase system at temperature T. The Clausius-
Clapeyron equation becomes indeterminate, since both the numerator and the
denominator are zero. However, one can duplicate the procedure leading to
equation (4.1.14) by requiring that both volume and entropy do not suffer
discontinuities at any temperature. This will of course lead to two equations for
dp* / dT, as shown below.
The volume requirement is

Da = D{3(dp*/dT) (4.5.3)

where
Va = t3V/t3T and V{3 = -t3V/t3p (4.5.4)

It is interesting to note that equation (4.5.3) is based on only one assumption,


namely, that in both phases volume is a unique function of pressure and
temperature. The entropy requirement is

DCp/T = VDa(dp*/dT) (4.5.5)


where also the Maxwell relations liS 1 liT = cpl T and liS1 lip = - Va have been 121
used. Equations (4.5.3) and (4.5.5) are called the Ehrenfest equations. Their ratio Phases
produces the result
(4.5.6)

which is called the Prigogine-Defay relationship.


It turns out that while equation (4.5.5) is sometimes confirmed reasonably
well by data, equation (4.5.3) [and of course equation (4.5.6)] is not. (Most of
the available data are for the rubber-glass transitions of polymers.) This leads
to considering the possibility that there may be internal state variables, say that
volume and entropy may depend, in addition to T and p, also on some internal
state variable x. With this, equations (4.5.3) and (4.5.5) become
Da = D{3 (dp* 1 dT) + DT( dx* 1 dT) (4.5.7)

and
Dcp = Da(dp*ldT) + DT'(dx*ldT) (4.5.8)

where
T = liV/lix and T' = liSllix (4.5.9)

This seems to dispose of the paradox, but there is a subtlety involved. The
affinity 6 = liG / lix should, at equilibrium, be zero. Now T = li6/ lip and T' =
-li61 liT, and hence if the system were at equilibrium one would again obtain
the Ehrenfest relations. Indeed, it is quite likely that glassy polymers are not
equilibrium phases.
If one allows for more than one internal state variable, one could get by
with an equilibrium theory by postulating that the kinetic equation is not invertible
at the equilibrium point (see the discussion in Section 3.2). However, this seems
to be a rather artificial way of dealing with the matter. In spite of these conceptual
difficulties, n-order theories have been generalized to orders higher than two, by
considering the possibility that derivatives of G of higher order than the second
are the first to suffer a discontinuity.
There are two other types of difficulties with the n-order transition kind of
theory. The first is as follows. In ordinary phase transitions one can, as discussed
in Section 4.1, simply assume that the state for a homogeneous system is the
ordered pair V, T, and then proceed to show that, if a spinodal region exists in
the isothermal A- V plot, the homogeneous state may be unstable to unmixing,
and therefore equilibrium states corresponding to two phases may exist. In
contrast with this, if one accepts the second-order phase transition theory as
sketched above, the assumption that for a homogeneous system the state is the
ordered pair V, T simply leads to consideration of the case where the A- V curve
may present a discontinuity of the second derivative. This would not lead to the
consideration of instabilities possibly resulting in phase separation, but simply
to the fact that, at any given pressure, the state is either that of phase I or of
phase II according to whether temperature is larger or smaller, even by an
infinitesimal amount, than the equilibrium transition temperature, and vice versa.
There would be no equivalent of the supercooled vapor or superheated liquid
concept. This point deserves some detailed discussion.
122 Consider a system which is just at the transition condition as identified by
values v'o and To of total volume and temperature; the latter is uniform in space.
Chapter Four
Let the superscripts I and II identify the behavior at temperatures immediately
above and below To, and suppose that the system exists in a "two-phase"
condition, i.e., part ofit exhibits type-I behavior and another part type-II behavior.
Since DV = 0, the specific volume is constant in space. Hence, if indeed the state
is simply V, T, the state is constant in space-and therefore one cannot identify
the existence of two phases in the sense of the definition given in Section 4.1,
since there is no surface of discontinuity of the local state. The only properties
which are different in the two "phases" are those which are not defined at that
state-the A- V curve has a discontinuous second derivative at the transition,
and the two regions simply exhibit the values corresponding to the second
derivative to the right and to the left. It is easy to convince oneself that the
relative amount of the two "phases" is in fact indeterminate. If even a differentially
small amount of heat is added to the system, the region of type-II behavior will
simply disappear since, if indeed the latent heat is zero, addition of heat will
result in an increase of temperature. The same is true if temperature is held
constant and the volume is changed by even only a differential amount.
The above point is clarified by consideration of the rate of "phase transition."
Consider first the case of an ordinary phase transition such as the freezing of
water. Suppose a semi-infinite body of water located at y > 0 is initially at the
freezing point (OC), while at time zero the plane y = 0 is brought to some
temperature lower than 0 C and thereafter maintained at that temperature. A
layer of ice will start to grow, through wtpch unsteady heat conduction carries
heat from the freezing surface located at y = l(t) to the cold surface at y = o.
This is the classical Stefan problem, which is formulated by assuming that the
temperature at the freezing surface is the equilibrium freezing temperature, i.e.,
oC. The freezing surface thus advances at a rate governed by the rate at which
the latent heat DH is removed by heat conduction, say a situation in which the
following boundary conditions are imposed on the differential equation for
unsteady heat conduction:

y = I(t), kaT/ay = (DH/V) dl/dt; T= T* (4.5.10)

Should a transition of the type discussed above take place the problem could
not be formulated in the same way, since equation (4.5.10) would yield an infinite

4
Cp I
I
I
I
I

7
I

II
FIGURE 4.5.1. Specific heat vs. temperature for
To T a system exhibiting an ordinary phase transition.
rate of advance of the freezing surface-the whole semi-infinite body would be 123
predicted to instantaneously undergo the transition as soon as the surface y = 0
Phases
is cooled below the transition temperature. The difficulty can be circumvented
by allowing the temperature at the transition surface to differ from its equilibrium
value, while assuming that the rate at which the transition takes place is governed
by the local undercooling. This corresponds essentially to regarding the system
as one endowed with an internal state variable, the rate of change of which is
the rate of phase transition.
The question of whether the latent heat is in fact zero in the special transitions
discussed here is a delicate one. In ordinary phase transitions, the equilibrium
enthalpy H* suffers a discontinuity. For instance, its value at a given pressure
could be expressed as

H* = 11(T) + DHfl(T - To) + fiT) (4.5.11)

where fl( ) is the Heaviside step function, 11 is zero at T > To, and 12 is zero
at T < To. Correspondingly, the specific heat at constant pressure is

(4.5.12)

where 8( ) is the Dirac delta function; the equilibrium specific heat exhibits a
delta spike at the phase transition. The behavior is of the type shown in Figure
4.5.1. We note that, on both sides of the transition, the specific heat is a quite
smooth function of temperature.
In some special transitions, the specific heat curve has the shape shown in
Figure 4.5.2 (which justifies the name of lambda transitions sometimes given to
the transitions). The shape of the curve as given in Figure 4.5.2 is, qualitatively,
what is observed for the transition that liquid helium undergoes at 2.2 K, and
for one of the solid-state special transitions of quartz. Whether the specific heat
actually becomes infinitely large at To is open to question, and the point is of
course not possible to settle by experiment; however, it certainly becomes very
large indeed, and the behavior is thus similar to that observed in supercritical
conditions. This is the basis of the supercritical theories of special transitions.
Such theories are essentially theories where the two "phases" are not regarded
~s truly different phases, but simply as conditions on the two sides of a region

FIGURE 4.5.2. Specific heat vs. temperature for


a system exhibiting a special phase transition.
124 where physical properties change very rapidly with slight changes of the state
Chapter Four
variables. Thus such theories are not truly phase-transition ones, since the abrupt-
ness of the transition is not retained.
If we wish to retain the idea of an actual phase transition, it is perhaps best,
in view of the difficulties with the n-order type of equilibrium theory, to simply
abandon the idea of an equilibrium phase transition and assume that the phase
observed at the lower temperature (the "glass") is a nonequilibrium phase. This
of course implies that internal state variables would need to be taken into account,
and this would bring kinetics into the picture.
There are only two ways of avoiding the issue of kinetics, i.e., of the rate of
evolution of internal state variables toward their equilibrium values. One is to
assume that it is infinitely fast, so that the internal state variables XI are always
equal to their equilibrium values xt. The other way is to assume that the rate
of evolution is infinitely slow, so that the internal state variables remain indefinitely
at whatever value they may have. If one makes the first assumption for the
high-temperature phase and the second for the glass, then quantities XI in the
glass have the value they had at the glass transition itself-the literature refers
to this situation by saying that the internal state variables get frozen in at the
transition.
Given any quantity L, it is in general expressed by an equation of the form

(4.5.13)

and its equilibrium value L* is given by

L* = l(p, T, Xr, ... , X~) = l*(p, T) (4.5.14)

since, of course,
(4.5.15)

with functions xf( ) being determined by the condition that G is minimized at


eqUilibrium.
In the following, a shorthand notation IT = al/aT, etc., will be used for the
partial derivatives of lower-case functions.
Let us now consider in particular the compressibility {3. If one makes the
two assumptions discussed above, in the rubber phase one would measure the
quantity
(4.5.16)

while in the glass phase one would measure

(4.5.17)

These quantities will of course differ without any need for the v( ) function to
be anything but perfectly smooth. The same argument applies also to a and cpo
The analysis below is based on the hypothesis that indeed all l( ) functions of
interest are smooth, and that any observed discontinuity at the transition is simply
related to the fact that one measures derivatives of the 1*( ) functions above it
and derivatives of the l( ) functions below.
We now let [L] represent the difference between the measured values of L 125
in the rubber phase and the glass phase. The notation DL is avoided, because
Phases
[L] is a difference between measurable quantities while DL represents an actual
difference in values of a function l( ) at the transition, and such differences
have been assumed not to exist. For the case of two internal state variables X
and Y (the reason for choosing two internal state variables will become clear in
the following) equations (4.5.16) and (4.5.17) yield

(4.5.18)

and analogous equations can be written for [cp ] and [a]. For the latter, in fact,
two equations can be written, since VT = -sp; so we obtain

[Cp ] / T = sxx} + Syy:} (4.5.19)

V[ a] = Vxx:} + Vyy:} (4.5.20)

V[a] = -sxx; - Syy; (4.5.21)

However,
gxxx; = -vx (4.5.22)

and
gxxx:} = Sx (4.5.23)

and also analogous equations for Y. On substituting equations such as (4.5.22)


and (4.5.23) into (4.5.19)-(4.5.21), we obtain

V[f3] = v~/gxx + v~/gyy (4.5.24)

(4.5.25)

V[a] = vxsx/gxx + VySy/gyy (4.5.26)

where equation (4.5.26) is obtained from both equations giving [a]. Since all
quantities are to be calculated at the transition itself, the second derivatives of
g are positive, and hence equations (4.5.24) and (4.5.25) imply that [f3] and [cp ]
are positive, as indeed is observed experimentally. No constraint is predicted
concerning the value of [a], though the latter is in general also measured as
positive. However, one cannot exclude that a special transition with [a] < 0 may
exist, just as ordinary phase transitions with DV < 0 exist.
By substitution and algebraic rearrangement one obtains

(4.5.27)

which of course implies that 71' ~ 1, as indeed is observed experimentally.


126 The reason for choosing two internal state variables is now perhaps clear.
Chapter Four
With only one internal state variable, the right-hand side of equation (4.5.27)
would be zero, i.e., the result 7T = 1 would be obtained again. The analysis could
of course be extended to more than two internal state variables, but the same
inequalities would be obtained, though with more cumbersome algebra. Since
these inequalities are all that the theory can provide, there is no reason to assume
the existence of more than two internal state variables.
It is noteworthy that the above results have been obtained without requiring
any of the constitutive functions to suffer even mild discontinuities at the transition
point. It follows that one cannot establish equations for dp* / dT (of the Ehrenfest
type), since such equations are simple mathematical consequences of assumed
discontinuities of the constitutive functions.

EXAMPLES AND PROBLEMS

Examples

1. Consider water at its triple point, in a condition where ice, water, and steam are
simultaneously present. Two examples concerning such a system will be discussed. First,
consider the case of a very damp winter morning; there is some partially melted ice on
the ground, and let us assume that the humidity is 100%. We are therefore in the presence
of all three phases, and yet, quite obviously, pressure is 1 atmosphere, i.e., it is not the
triple-point pressure for water. Furthermore, the ambient temperature may well not be
the triple-point temperature. How do we explain the situation?
First consider the case where the temperature is indeed the triple-point temperature.
This is trivial: it is an equilibrium situation, with the partial pressure of water vapor in
the air being equal to the triple-point pressure. If the temperature is, say, higher than the
triplepoint temperature, the system is not in equilibrium, and heat transfer from the
ambient to the ice-water interface will result in a finite rate of melting; it is still reasonable
to assume that the temperature at that interface is the triple-point temperature (see Chapter
7 for more details on this point).
2. Now consider the case where a closed constant-pressure box contains only ice,
water, and steam at the triple-point temperature and pressure. Suppose some heat is added
very slowly, not enough to make any of the three phases disappear. Is ice going to melt,
water to evaporate, or some combination thereof?
Should the constraint be that the box has a constant volume, the problem would be
trivial. Sayan amount Q of heat is added. Let DVM be the volume change of melting
(which is negative for water) and DVv the volume change of vaporization; and let DHM
and DHv be the corresponding enthalpy changes (or latent heats). The two equations

(4.E.l)

and
(4.E.Z)

determine uniquely the values of the number of moles melted, nM , and of the number of
moles vaporized, n v. However, the constraint is that of constant pressure, and it is easy 127
to convince oneself that one cannot write down two independent equations which determine
the values of n M and n v. It is possible to fool oneself that one has written two such Phases
equations, or some set which solves the problem, but on close inspection it will tum out
that the equations which have been written are not independent of each other.
The problem is related to the fact that pressure is not an external state variable but
a function of state. One should ask oneself how would one actually conduct a constant-
pressure experiment. An approach would be to imagine that one has a frictionless piston
with a fixed weight on it. However, frictionless pistons do not exist in reality. Another
way would be to consider a piston with friction, which is controlled by a pressure transducer
signal. The control would never move the piston unless pressure changes from the set
value; as long as heat is added slowly, so that the system is always at equilibrium, pressure
would always be equal to the triple-point pressure, and therefore the control would never
move the piston. Hence the real constraint is a constant-volume one, and the solution is
trivial.
This example is meant to illustrate the actual meaning, of pragmatical engineering
significance, of the fact that volume is indeed an external state variable, while pressure
is not. The following example stresses the same point.
3. In the late 1970s, the Seveso accident took place. A chemical reactor exploded
during the night, when the plant was shut down, and a very toxic chemical was spread
over a vast area, resulting in very major damage. The reactor was supposed to be a
constant-pressure one, and hence one which could not possibly explode. The constant-
pressure device was extremely simple: a two-inch-diameter tube which vented the reactor
to the atmosphere.
What happened was that, in the liquid left in the vessel, some very fast chemical
reactions occurred and produced a large amount of gaseous components. These could, of
course, escape through the two-inch tube; however, the highest possible velocity through
a tube is the velocity of sound, and the rate of production of gaseous components exceeded
the corresponding highest possible ftowrate through the tube. This is known because an
acute whistling noise was heard immediately before the explosion. Pressure within the
vessel therefore increased well above atmospheric pressure, and the explosion could occur.

Problems

4.1. In the usual pressure-temperature plot, phase transitions are identified by lines, since
pressure is the same in the two phases. The lines represent the equilibrium pressure at which the
transition occurs. One can construct also a temperature-volume equilibrium plot. In such a plot,
curves would represent loci of equilibrium points of the isothermal a (V) curves. These curves delimit
regions in the V - T space where, at equilibrium, the system is in either a one-phase or a two-phase
state, such as a gas region, a gas-liquid region, and so on.
Consider two different one-component systems. Both have only one solid equilibrium phase; for
one of them the DV of the solid-liquid transition is positive, for the other it is negative. Draw
qualitative pressure-temperature and temperature-volume equilibrium plots for both systems.
4.2. Consider a two-phase system in which both heat and mass transfer are taking place. Only
one component is transferred between phases. In developing this problem, use the same approach
as taken in this chapter: the local state is uniform in both phases, but it suffers a discontinuity at the
interface, and in particular both the chemical potential of the transferring component and the
temperature are discontinuous across the interface. (This corresponds to neglecting the temperature
and concentration boundary layers near the interface.) The quantities of interest are: the heat transfer
rate q, the mass transfer rate N, the two chemical potentials of the transferring component in the
two phases, and the temperatures ofthe two phases. The system as a whole is maintained at constant
volume and is adiabatic.
128 a. What form does the second law take?
h. When the two temperatures are equal, is the heat transfer rate necessarily zero? Conversely,
Chapter Four if the two chemical potentials are equal, is the mass transfer rate necessarily zero?
4.3. It has heen observed for some gas-liquid systems that the dynamic surface tension s is
different from the static one s*. The static surface tension is that observed when the interface area
l: is constant in time. The following equation has been proposed for the dynamic surface tension:

s = s* + /L,l:'

with /L, a constant called the surface viscosity.


What restrictions are imposed by the second law on the possible values of the surface viscosity?
4.4. In ordinary fluid mechanics of liquids, the assumption of incompressibility is common.
Now suppose both the "liquid" and the glass phase of a polymer are truly incompressible, i.e., their
density at any given temperature is independent of pressure. Does that lead to problems for the
analysis of secondary phase transitions in Section 4.5? Comment on your results in detail. (Hint:
First start by considering a simple, ordinary one phase system, and assume it is truly incompressible.
Think of the definition of baric derivatives.)

LITERATURE

"For each hotness there is some body whose latent heat


at that hotness fails to vanish for some volume."
This is called the "Thermometric axiom "; and, since it has
to do with latent heat, it seems appropriate that it
should have been formulated by J. B. Boyling
[Proc. R. Soc. London 329, 35 (1972)].

The idea that a one-component system has in principle properties which are continuous
through the phase change, embodied in Figure 4.1.1, is originally due to Van der Waals,
Die Continuitat des Gas!ormingen und Flussigen Zustandes, J. A. Barth, Leipzig (1881).
The question of pressure and temperature being continuous across an interface
between two phases presents some subtlety. For instance, shock waves in gases are
phenomena where both temperature and pressure are discontinuous; however, the discon-
tinuity travels through the gas and is sustained by a nonequilibrium phenomenon. The
point is discussed in detail in any book on gas dynamics. Discontinuities of temperature,
chemical potential, and velocity can also occur in nonequilibrium phenomena; the whole
question is discussed in a recent review paper by G. Astarita and R. Ocone, Adv. Chern.
Eng. (to be published).
Phase equilibria are discussed in all thermodynamics textbooks. For the case of
mixtures, a very complete discussion is given by M. B. King, Phase Equilibria in Mixtures,
Pergamon Press, Oxford (1969).
Phase equilibria in continuous mixtures have been discussed in a number of papers;
the recent one by S. K. Shibata, S. I. Sandler, and R. A. Behrens, Chern. Eng. Sci. 42,
1977 (1987) is a good guide to the relevant literature.
The possibility of heterogeneous isomerization equilibria has been considered by H.
S. Caram and L. E. Scriven, Chern. Eng. Sci. 31, 163 (1976); see also G. Astarita, Chern.
Eng. Sci. 31, 1224 (1976), and the literature cited in these papers.
A comparatively recent review on liquid crystals is to be found in G. W. Gray and
P. A. Winsor (Eds.), Liquid Crystals and Plastic Crystals, Ellis Horwood, Chichester (1974).
Special transitions are discussed by H. B. Callen, Thermodynamics, Wiley, New York 129
(1960); L. Tisza, Phase Transformations in Solids, Chapter 1, Wiley, New York (1951);
and K. Denbigh, The Principles of Chemical Equilibrium, Cambridge University Press, Phases
Cambridge (1964). The specific case of the rubber-glass transition in polymers is discussed
in detail by R. N. Haward, The Physics of Glassy Polymers, Applied Science Publ., London
(1973). The literature which deals specifically with the problem of predicting inequalities
such as 1T;;" 1 is extremely confusing, and much of it is best left unread.
Chapter Five

THERMODYNAMICS OF
RELAXATION

I might proceed to apply the theory to the phenomena of


combustion, the heat of which consists in the living force
occasioned by the poweiful attraction through space of
the combustible for the oxygen, and to a variety of other
thermochemical phenomena; but you will, doubtless, be
able to pursue the subject further at your leisure. I do
assure you that the principles which I have very
impeifectly advocated this evening, may be applied very
extensively in elucidating many of (he abstruse as well
as the simple points of science; and that patient inquiry
on these grounds can hardly fail to be amply rewarded.
J. P. Joule, 1847

NOTATION

A Free energy density kcalkmol- 1


o{ Constitutive functional for A kcalkmol- 1
0' Functional derivative of 0
B,B' "Viscosity". coefficients
D( Distance operator
E Tangent modulus
E' Secant modulus
F( ) Function delivering x* kmol
f( ) Constitutive function for p* atm
~{ } Constitutive functional for pressure atm
g Objects in domain of functional
k Inverse time constant S-1

L Any thermodynamic property


"T Total number of moles kmol
p Excess stress tensor atm
p Pressure atm
q Retardation parameter S-1
131
132 R Gas constant
Chapter Five r( Constitutive function for rate of change
IR Residual
IR[ Mapping delivering the site
S Entropy density kcal kmol- l K- l
s Time measured backward from t s
s( ) Function delivering S* kcal kmol- l K- l
s{ } Constitutive functional for S - S* kcal kmor l K- l
T Temperature K
T Stress tensor atm
Time s
U Internal energy kcalkmol- l
u Velocity ms- l
u( Constitutive function for U kcalkmol- l
V; Total volume m3
x Extent of reaction kmol
z Direction of propagation m
a Site
a'( ) History of site
a~( ) Retarded history
~ Sum of stoichiometric coefficients
(J Relaxation time
<I> Density
U' State
U'/ Stochiometric coefficient
T Time s
1 Unit tensor

Subscripts

q For retarded history


o Initial

Superscripts

t, T History at time t, T
* At equilibrium
Substantial time derivative
Operators 133
Thermodynamics
11 Modulus (or absolute value)
of Relaxation
S Frechet differential
1111 Norm
.
[ ] Jump
S/St Invariant time derivative
D( ) Distance
5.1. INTRODUCfION TO RELAXATIONAL
SYSTEMS

I am part of all that I have met;


Yet 'all experience is an arch wherethrough
Gleams that untraveled world whose margin fades
For ever and for ever when I move.
Alfred Lord Tennyson
Relaxation phenomena are of particular importance in the mechanics of polymeric
systems, and the corresponding thermodynamic theory presents very significant
mathematical complexities since, in addition to the need for functional analysis,
continuum mechanics introduces the additional complexity related to the fact
that an appropriate three-dimensional description requires a basis in tensor
analysis. The purpose of this chapter is not that of presenting the complete
thermodynamic theory of relaxation, but only the essential aspects of it, and as
much of the mathematical complexities as possible are left out of the picture.
There is no way of avoiding functional analysis, but tensors can be avoided by
restricting attention to systems described only by scalars. Of course, the systems
considered below are so simple that, in actual fact, one would describe them
with a theory based on internal st'ate variables; however, they do exhibit relaxation
phenomena, and the essential features of the thermodynamic theory of relaxation
can hopefully be grasped by consideration of such simple systems.
In particular, we begin by considering an extremely simple but concrete
case: a box fitted with a piston which allows the total volume to be imposed
arbitrarily at anyone time. Temperature is regarded as spatially uniform, but it
can also be imposed arbitrarily at anyone time. The box contains a gaseous
mixture within which a nonequimolar chemical reaction may take place; thus
the system is endowed with one internal state variable, i.e., the extent of reaction
x. The Feader is asked to consider the situation where the fact that a chemical
reaction may be taking place in the box is not known.
First let us consider the case where temperature is held constant. The
equilibrium function is then
x* = F(Y,) (5.1.1)

with F( . ) a known function. Consider the simple case where F( . ) is single-valued;


the values of x* range in a restricted interval, say 0:$ x* :$ 1. Furthermore,
consider a very simple constitutive equation for the rate of evolution of the
internal state variable, namely,
dx/dt = (x* - x)/(J (5.1.2) 135
136 This equation guarantees that at all times x is approaching the instantaneous
equilibrium value x*, at a rate which is linear in the "driving force" x* - x; 6
Chapter Five
is the intrinsic time scale of the evolution of the internal state variable. Indeed,
should the volume be held constant (so that also x* is constant in time), equation
(5.1.2) (subject to the boundary condition x = Xo at t = 0) has the following
trivial solution:

x = x* + (xo - x*) exp(-tI6) (5.1.3)

which shows that equilibrium will be reached over a time scale of order 6.
Now consi4erthe case where V; is changing in time in some arbitrary fashion.
Equation (5.1.2) can be rearranged as follows:

x + 6 dxl dt = F( V;(t; t = O,x = Xo (5.1.4)

Equation (5.1.4) has the obvious integrating factor exp(tI6), so the solution is

x= Xo exp( -tl 6) + (116) J: exp( -sl 6)F( Vb - s ds (5.1.5)

This equation should be examined in great detail. First of all we note that,
when t is significantly larger than 6, the value of x becomes independent of its
initial value Xo. Furthermore, since the value of F is bound to be between 0 and
1, at large values of t the value of the integral is dominated by the values of the
integrand near the lower limit of integration. What is implied is that, at any given
time t (provided it is large as compared to 6), the value of x is determined
entirely by the values of V; in the last several units of time, and is independent
of the values that V; may have had in the very distant past. Indeed, s is a time
variable which measures, at any given instant of observation t, "how long ago"
a given value of V; was assigned. In this interpretation, exp( -sl 6) is seen to
represent how well events occurring s time units ago are remembered. The system
has a "fading memory": volumes imposed shortly ago do influence the present
state, but those imposed in the distant past do not.
Since equation (5.1.4) admits a simple integrating factor, its solution is in
terms of an explicit integral; this is, however, related to the very simple form of
the kinetic constitutive equation (5.1.2). More complex kinetics would result in
a differential equation, which may well not admit an integrating factor. Now the
integral on the right-hand side of equation (5.1.5) is a very simple functional of
F[ ] (in fact, it is its Laplace transform). For more complex kinetics, the solution
may not be expressible as an integral, but it is perhaps obvious that it will in any
case be representable as a functional of the F[ ] function. This clarifies why
functional analysis is crucial to the thermodynamics of relaxation.
We now consider the following possibility. For the system under consider-
ation, we do not know that there is an internal state variable, and even less do
we have equations for its rate of evolution in time and for the equilibrium function.
We are undertaking isothermal experiments, and we are simply trying to find out
what the pressure-volume relationship may be. Since x is a state variable, the
instantaneous value of pressure depends also on X, and therefore the consider-
ations made in the preceding paragraph apply to pressure as well.
At the beginning of our experiment, the results will not make any sense at 137
all since they will depend on what volumes were imposed on the system prior
Thermodynamics
to the beginning of the experiment; these volumes are not known (should they of Relaxation
be known, the actual experiment has really been started earlier). In the following,
we discuss the kind of results we would obtain in different experiments.
The simplest experiment would be one where we simply keep the volume
constant. In such an experiment, x would vary in time according to equation
(5.1.3), and therefore we will record a pressure that changes in time and
approaches asymptotically some final value p*; the time scale of this phenomenon
will be 8. It is obvious that the system cannot be described by an elastic theory,
since at a given value of volume different pressures are exhibited. However, also
a viscous theory is inadequate: at assigned values of both volume and its rate of
change (the latter being assigned a value of zero), different pressures are exhibited.
Neither is there any hope by considering some extension of a viscous theory
where higher-order time derivatives of volume are included into the state, since
all of those are kept at a value of zero while pressure is changing. Hence the
need for a theory of higher complexity is made evident (if the existence of an
internal state variable is either unknown or difficult to model).
Another possible experiment is one where, after the volume has been assigned
in some arbitrary fashion for some time, it is suddenly changed by a finite amount
over a time interval negligibly small as compared to 8. Such a volume change
could be regarded as being instantaneous, and consideration of equation (5.1.4)
shows that, in response to such a change, x would not vary, but only dx/ dt would
suffer a discontinuity. However, pressure does depend on the instantaneous value
of V as well as on x, and thus we would observe that, in response to an
instantaneous change of volume, pressure also undergoes an instantaneous
change, followed by a relaxation phenomenon. In this section we try to avoid
mathematical technicalities; however, in some sense the result of the experiment
just described would convince us that the instantaneous value of V has special
relevance, and thus we would conclude that the isothermal equation of state for
pressure is of the following form:

p(t) = HY,(t); Y,(t - s)} (5.1.6)

where ~ is a functional of the function Y,(t - s), whose value also depends
parametrically on the instantaneous value Y,(t).
The fact that the instantaneous value of volume at time t appears explicitly
as an argument of the functional needs to be discussed in some detail. Suppose
that, at some time t - s previous to the instant at which pressure is being measured,
the volume had been given a spike such as is shown in Figure 5.1.1. Such a spike
would have no effect on the pressure at time t, since it would contribute nothing
to the integral in equation (5.1.5). In contrast, a volume spike at time t would
result in a pressure spike, and this is the reason why the instantaneous value of
the volume appears explicitly as an argument in equation (5.1.6) [the exact
mathematical definition is that the functional in equation (5.1.6) is smooth except
for an atom at time t; see also the Appendix for more details on this point].
By additional experiments we would convince ourselves that, no matter
which Y,(t) function we impose on the system, the pressure exhibited at any
138 v
Chapter Five

I
I
I
I
I
S I
----1I
I

FIGURE 5.1.1. A spike of volume in the past.

given instant only depends on the volumes we have assigned over the past several
time units. This observation could be incorporated mathematically in equation
(5.1.6) by techniques of functional analysis which will be discussed in the next
section.
So far, only isothermal experiments have been considered. Now suppose we
start conducting experiments where also temperature is allowed to change. We
would begin by again undertaking the simple equilibrium experiment at some
other fixed temperature. For the concrete system we are considering, the value
of (J would now be different, and presumably smaller if the temperature has
been increased. So we would conclude that the "relaxation time"-the time
scale needed to reach equilibrium isothermally-is a decreasing function of
temperature.
In some other experiment, we will change temperature with time in some
arbitrary fashion while keeping the volume constant. Now the values of both x*
and (J on the right-hand side of equation 5.1.2 are changing in time, and therefore
the value of x at anyone time would depend on which temperatures we have
assigned in the last several units of time. A sudden change of temperature after
some fixed preassigned T(t) again would not result in a discontinuity of x, but
only of dx/ dt; however, pressure depends on T as well as on x (when V. is held
constant), and thus such a sudden change of temperature will result in a sudden
change of pressure. We would thus finally conclude that the constitutive equation
for pressure has the following form:

pet) = ~{V.(t), T(t); V.(t - s), T(t - s)} (5.1.7)

with the dependency on both v.(t - s) and T( t - s) having the fading memory
property.
We note that the ordered pair v.,
T would be identified as the site a.
Therefore, equation (5.1.7) suggests that the state of our system is

(T(t) = aCt), aCt - s) (5.1.8)

with the fading memory property.


It is important to realize two points about the discussion in this section.
First, had one started from more complicated kinetic equations instead of equation
(5.1.2), one would have reached the same qualitative conclusions. Second, con- 139
sideration of systems for which we would not know 9f the existence of an internal
Thermodynamics
state variable is not an abstract exercise. The internal state variable need not be of Relaxation
the extent of a chemical reaction, a phenomenon of which we would presumably
be aware. For instance, in the case of polymeric materials, the morphology may
change in time due to whatever deformations are imposed on the material, and
such morphological changes have their own intrinsic kinetics. Thus, qualitatively,
the discussion in this section would apply with x being the instantaneous mor-
phology; and indeed we do not understand well at all the intrinsic kinetics of
morphological changes in polymers. Biological systems may also be endowed
with some fading memory for past events, as related to some biological internal
state variable of which we understand even less than we do the morphology of
polymers. For systems such as these, one would want indeed to describe their
thermodynamic behavior in terms of the state given by an equation of the form
of equation (5.1.8), rather than being the ordered triplet V, T, X, or some appropri-
ate analog thereof.
Now let us consider again the physical system discussed in this section, with
x actually being the ex;tent of reaction, and suppose that the system is in fact a
gas mixture which behaves as an ideal gas. The pressure at any instant in time is

(5.1.9)

where nT is the total number of moles in the system. As the chemical reaction
proceeds, nT changes in time, which is obviously the reason why p depends on
x (this would not be true, of course, for an equimolar reaction, and therefore
the discussion here applies to a nonequimolar reaction). At equilibrium, the value
of nT would depend on temperature and on total volume. Therefore one would
conclude that the equilibrium pressure is given by an equation of the type

p* = f(T, v,) (5.1.10)

with f( .) nonlinear in both T and 1/ v" in spite of the fact that the mixture
behaves as an ideal gas at any given composition.
Let f3 be defined as follows:

(5.1.11)

(unless the reaction is equimolar, f3 is nonzero). If no is the total number of


moles corresponding to x = 0, the total number of moles at any given time is

(5.1.1l)

and therefore the equilibrium pressure is

p* = RT(no + x*f3)/ V, = RT[no + F( V,)f31/ V, (5.1.13)

The nonequilibrium pressure is

p = RT(no + xf3)/ V, (5.1.14)


140 If attention is restricted to the case where temperature is being held constant
in time, x can be eliminated between equations (S.1.2) and (S.1.14) to yield
Chapter Five
p + (Jdp/dt = p* (5.1.15)

This equation is called the Maxwell equation, since it is analogous to an equation


derived by Maxwell from his kinetic theory of gases. In its proper three-
dimensional formulation, with the stress tensor playing the role of pressure and
the deformation tensor that of volume, it has been used extensively to describe
the isothermal mechanical behavior of polymeric materials. The three-dimen-
sional formulation will not be discussed here, since it involves significant
mathematical complexities.

Appendix

We consider the more general case where the gas can still be regarded as
ideal, but the kinetics are not described by the simple linear expression in equation
S.1.2. Here
dx/ dt = r( v" T, x) (5.I.A.I)

Suppose x has been so chosen that its initial value is zero. Then the total
number of moles at any time is related to the initial value by

(S.I.A.2)

and pressure is given by

p = RTno(1 + {3x)/V, (S.I.A.3)

Differentiation with respect to time yields

p dV,/ dt +Vt dp/ dt = RTno{3 dx/ dt + (3RnoxdT/ dt (5.l.AA)

while the solution for x gives

(5.I.A.5)

Substitution of equations (S.1.A.I) and (S.1.A.S) into equation (S.1.A.4) leads


to an equation of the following form:

p dV,/ dt + Vt dp/ dt = F(p, v" T) + G(p, v" T) dT/ dt (5.I.A.6)

where the F( ) and G( ) functions are determined by the r( ) function.


Now let us assume that the total volume is changed by a finite amount in a
zero time, i.e., it is forced to undergo a jump discontinuity [V,]. At such an
instant, dVt / dt is infinitely large, and this will result in an infinitely large value
of dp/ dt as well, i.e., pressure will also undergo a jump [p]. The same argument
also applies to the case of a temperature jump [T]. It is noteworthy that such
jumps would not be allowed for in a viscous theory, where an infinitely large
rate of change of volume would result in an infinitely large pressure.
The possibility of such jumps is important in two respects. First, it is crucial 141
to the conceptual development of the theory since, as will be seen in the next
Thermodynamics
section, it makes the partial derivative in equation (5.2.6) legitimate. Second, it of Relaxation
represents the basis for a theory of propagation of discontinuities, which is
sketched in Section 5.5.
The arguments in this appendix can be generalized to nonideal gases, though
the algebra becomes rather complex. In the thermodynamic theory of polymers,
two approaches are present in the literature. One is based directly on functionals
such as in equation (5.1.7). The other approach is a statistical mechanics one; a
molecular model for the polymeric material is formulated, which results in the
consideration of a distribution function cp for the molecular parameters (orienta-
tion, end-to-end distance, and the like). This is followed by an analysis analogous
to that presented in this section, with cp playing the role of the internal state
variable, and the kinetic function r being deduced from the modeling of the
molecular scale mechanics.
The equation derived by Maxwell is as follows. In a gas maintained at
constant volume and temperature, the pressure is constant and is given by the
ideal gas law. The stress tensor, however, is not constant, and equation (5.1.15)
applies to the difference P between the stress and the isotropic pressure:
P =T+ p1 (!I.I.A.7)

P+ (JdP/dt = 0 (S.I.A.S)

Of course, P decays exponentially in time toward its equilibrium value of


O. The time scale is (J, which in the same theory is the ratio of the shear viscosity
of the gas to pressure. For air at atmospheric conditions, that corresponds to
about lO-IO s, i.e., the decay is so rapid that for practical purposes it can be
regarded as instantaneous.
The Maxwell equation often used in the description of the flow behavior of
polymeric materials is as follows. For an incompressible Newtonian fluid,
equation (2.3.A.1) can be rewritten as

P = T + p1 = 2p. D (!I.l.A.!)

where P is called the extra stress (for an incompressible fluid, pressure is defined
only to within an arbitrary additive constant, and hence only the extra stress is
determined by its constitutive equation). The Maxwell equation is
P+ (JSP/St = 2p.D (!I.I.A.IO)

where S/ St is an appropriately invariant time derivative. The appearance of time


derivatives in the constitutive equations for a system exhibiting relaxation is one
of the main reasons for the mathematical complexity of the theory. In fact, the
substantial time derivative is not frame-invariant, and thus it cannot appear in a
constitutive equation; however, several different invariant time derivatives can
be formulated. Thus, upon generalization ofa simple one-dimensional (or scalar)
model containing time derivatives to the full three-dimensional formulation, there
is an arbitrary choice of which particular time derivative to use, a choice that
must be guided by experimental results.
142 5.2. ELEMENTARY THERMODYNAMICS
OF RELAXATION
Chapter Five
I remember, I remember
In the days of chill November
How the blackbird on the . ..
I forget the rest. It is the beginning of the first piece of poetry
I ever leamt; for
Hey, diddle diddle,
the cat and the fiddle,
I take no note of, il being of a frivolous character, and lacking
in the qualities of t;"'e poetry. I collected fou.rpence by the
recital of "I remember, I remember." I know it was fourpence,
because they told me that if I kept it until I got twopence more
I should have sixpence, which argument, albeit undeniable,
moved me not, and the money was squandered, on the very
next moming, though upon what memory is a blank.
J. K. Jerome

In this section, we consider a system the state of which is


u(t) = a(t), a(t - s) (5.2.1)

endowed with the fading memory property. The site a is the ordered pair (or
vector) V, T.
The first question to address is the mathematical formalization of the concept
of fading memory. This can be accomplished as follows. First, the space of
possible past histories of the site (i.e., of functions mapping the positive quantity
t - s into the ordered pair V, T) is assigned a topology such that two histories
which differ only in the distant past are close to each other. Second, the functionals
mapping the state into the. values of the constitutive quantities are assumed to
be smooth with respect to such a topology. This guarantees that the theory will
deliver values of the constitutive quantities which are close to each other for two
such histories. In the Appendix to Section 3.6, the concepts of topology, norm,
and metric have been introduced; however, they are not crucial to the theory of
continuous mixtures, where a norm of the simple type

11/(u)11 = f I/(u)1 du (5.2.2)

is adequate. In the description of relaxation phenomena, the topology must be


so chosen as to satisfy the fading memory property. This can of course be
accomplished by choosing an appropriate metric, as discussed below, but it
should be kept in mind that while the topology induced by the metric has physical
significance, the metric itself does not.
In order to assign a metric, one needs first to have a definition of subtraction
for the objects in the space considered. Since this is available for functions, we
may proceed. Let the norm of the objects g, namely Ilgll, be a transformation of
the gs into nonnegative real numbers. Then the distance between two objects g1
and g2 is defined as
(5.2.3)

provided 11'11 is such that D( . ) satisfies the triangular rule.


One is now in a position to assign a topology to the space of states which 143
will produce the fading memory property. Let the norm be Thermodynamics
of Relaxation
lIa. a(t - s)1I = lal + too exp(-ks)la(t - s)1 ds (5.2A)

where la(t - s)1 is the modulus of the vector v..


T at time t - s.
Then. given two quantities a( ) which are different only in the distant past
[at least not so different as to overcome the negative exponential in equation
(5.2.4)], at equal instantaneous quantities a the two states will be close to each
other, as required. [It should be noted that the constant k appearing in equation
(5.2.4) is not a property of the material considered: The same topology could be
induced by a different norm. and since only the topology is a property of the
material, the quantity k is not.]
In order to construct a thermodynamic theory, one needs not only the
assumption of smoothness, but also that of differentiability, since in order to
deduce consequences of the second law time derivatives have to be calculated.
As time progresses, both a and its history change. The history changes not only
because some more events are added to it, but also because previous events are
perceived as having happened longer and longer ago. A change in notation is
perhaps useful to clarify this point. Suppose the function a'(s) is defined as
follows:

a'(s) = a(t - s) (5.2.5)

The function a'(s) maps the time elapsed since a given site was imposed into
that site: It indicates how long ago that site was imposed. The domain of a'(s)
is the set of all positive real numbers; zero is not included, since a'(O) is simply
the present site a(t). which appears explicitly in the expression of the state. The
quantity a'(s) is called the history of the site at time t.
The free energy is given by an equation of the form

A = GJ{a; a'(s}} (5.2.6)

and its rate of change can be expressed as

A = (aGJ/iJa)a + 8o{ (5.2.7)

where the partial derivative is the ordered pair of partial derivatives with respect
to V and T, while the second term on the right-hand side is the Frechet differential,
the mathematical nature of which has been discussed in the Appendix to Section
3.6. Which particular Frechet differential is involved here is discussed at the end
of this section.
The partial derivative is of course to be understood in the sense that the
history is being kept constant. This can only be achieved by an instantaneous
change of site (albeit a differential one), but, as was discussed in the previous
section, such instantaneous changes do indeed make sense. (They would not
144 make sense for a viscous system, because a", which is included in the state for
viscous systems, would be infinite at such a change. In this sense, the theories
Chapter Five
of viscous and relaxational systems are seen to be mutually exclusive.) The Frechet
differential is to be understood as the initial rate of change of free energy which
would be observed should the site be held constant at the instant of observation.
This shows that 80{ } describes an essential feature of the behavior of relaxa-
tional systems.
If equation (5.2.7) is substituted in the second law, one obtains the results

p = -ao/av (5.2.8)

s = -ao/aT (5.2.9)

80:s 0 (5.2.10)

Equation (5.2.10) shows that the Frechet differential is indeed minus the
rate of energy dissipation. It also implies that, if volume and temperature are
held constant, free energy can only decrease in time-and hence that of all states
corresponding to an assigned site, the equilibrium state has the lowest free energy.
We note that the partial derivatives on the' right-hand side of equations
(5.2.8) and (5.2.9) are themselves functionals of the state, and therefore both
pressure and entropy are expected to change in time even if the site is held
constant; in other words, the equations give the instantaneous values of entropy
and pressure, which need not be equilibrium values.
The Frechet differential has been introduced in the Appendix to Section 3.6.
In the special case of the free energy functional, if the site is being held constant,
one can write
af+df(s) = af(s) + (aaf(s)/at) dt (5.2.11)

In view of equation (5.2.5), one has

aaf(s)/at = -af'(s) (5.2.12)

with the prime on the right-hand side indicating the ordinary derivative. Thus
the Frechet differential for the free energy functional in equation (5.2.10) is

80 = 80{a f (s)l-a f'(s)} (5.2.13)

Of course, the underlying assumption is that the free energy functional is


Frechet differentiable. The Frechet differential could be written as

80 = f o'{afs; s}af'(s) ds (5.2.14)

with (!]'{ } being the functional derivative of the free energy functional.
5.3. EQUILIBRIUM AND DISSIPATION IN 145
RELAXATIONAL SYSTEMS
Thermodynamics
of Relaxation
It was a past that kept changing while he advanced in his trip,
because the traveler's past changes according to the path
followed, and I don't say only the recent past to which each
passing day adds a new day, but also the distant past.
I. Calvino

In a system which exhibits relaxation phenomena, equilibrium is reached


only after the site has been held constant for a sufficiently long time. Let a'*(s)
be the following function:

a'*(s) = a(t) for all s> 0 (5.3.1)

The history a'*(s) corresponds to the situation where the site has always been
held constant in the past at its present value, and therefore it obviously corre-
sponds to equilibrium. Hence the eqUilibrium values of all constitutive properties,
identified with an asterisk, are given by equations of the form'

A* = lIlI[a(t); a'*(s)] (5.3.2)

However, one can never in actual fact know that the site has always been
kept constant in the past: what can be realized experimentally is only to keep
the site constant for a sufficiently long time. This situation can be described
mathematically as follows. Let a'(s) be the (possibly unknown) history of the
site at the time when one begins holding the site constant. At some later time T,
the history is

0< s:5 T - t: aT(s) = a(t)


(5.3.3)
s> T - t: aT(s) = a'(s + t - T)

Such a history is called the static continuation of the history a' (s ). It is seen
that, as T - t increases, the distance between the static continuation and the
equilibrium history a'*(s) decreases. Thus the assumptions of smoothness for
the constitutive functionals guarantee that indeed equilibrium will be approached
asymptotically in static continuations.
Since a'*(s) is uniquely determined by the present site, the equilibrium
properties are unique functions of the site. Furthermore, in view of equations
(5.2.8) and (5.2.9), one concludes that the relationships between equilibrium
properties are the same as would apply for an elastic system. In this sense, it is
concluded that as long as one is only interested in the determination of equilibrium
properties, one can regard a relaxational system as elastic.
The rate of dissipation in relaxational systems is simply minus the Frechet
differential in equation (5.2.13). At any given instant in time of any transforma-
tion, one could start holding the site constant, and the free energy would begin
decreasing at a rate given by that Frechet differential. This gives a physical
146 interpretation of the dissipation rate in relaxational systems: it is the initial rate
Chapter Five
at which free energy would decrease should the site suddenly be held constant.
An important question to be addressed is the fOllowing: Is it possible to
perform a transformation in a relaxational system slowly enough to make the
total dissipation as small as one wishes? The answer, as shown below, is a quali-
fied yes.
Given a transformation a = lR(t), the history of the site at anyone time is
determined:
a'(s) = lR(t - s) (5.3A)

Now suppose that another transformation is accomplished, which scans the same
sites but at a slower rate. Such a transformation is called a retardation of the
original one, and the corresponding history of the site is

a~(s) = a'(qs) (5.3.5)

where the retardation parameter q is a positive number smaller than unity.


The restriction to the result below is that the original history of the site (and
hence also the retarded history) needs to be Taylor-series expandable. If that is
the case, one can write

a~(s) = a(t) + a~(O)qs + O(l) (5.3.6)

The two terms on the right-hand side can be written as follows:

a(t) = a'*(s) (5.3.7)

and
(5.3.8)

i.e., the second term is minus the instantaneous rate of change of the site in the
retarded transformation. Thus one obtains

p = p* + .s~[a; a'*{s) I-s]qa" + O(q2) (5.3.9)

The latter equation shows that, provided q is small enough (i.e., provided
the transformation is slow enough), a viscous theory is recovered asymptotically.
Since for viscous systems the total dissipation can be made as small as one wishes
by effecting the transformation slowly enough, the result has been proved.
The above analysis could be extended by retaining higher-order terms. This
would produce a series of higher- and higher-order approximations to the pressure
response for sufficiently slow transformations. In such a scheme, the elastic theory
(p = p*) is then identified as the zero-order approximation, the viscous theory
as a first-order approximation, and so on. Second-order constitutive equations
for the stress have frequently been used in the analysis of the deformation of
polymeric materials.
While the analysis given above provides proof that sufficiently smooth
transformations leading from one equilibrium state to another can be accom-
plished slowly enough to make the total dissipation as small as required, the 147
approximation procedure in higher- and higher-order constitutive equations is Thermodynamics
useful in practice only if the finite rate transformations to be analyzed are of Relaxation
reasonably well represented by a Taylor series truncated after a few terms for s
ranging from zero to some value well in excess of the relaxation time. For the
concrete case discussed in Section 5.1, volume and temperature changes would
need to be small and smooth over a time range of order () for any approximation
to be valid.
It may at first sight seem that the conclusion, that transformations in relaxa-
tional systems can be accomplished slowly enough to make the total dissipation
as small as required, is in contrast with the earlier conclusions, that the dissipation
connected with the occurrence of chemical reactions does not have that property.
However, the above conclusion is only that the dissipation can be made as small
as required if one starts from an equilibrium state, and indeed under such a
restriction the conclusion is valid for chemical reactions. However, in actual
practice the latter are not carried out starting from an equilibrium state, but the
reactants are simply brought together and allowed to react at whatever rate is
dictated by the kinetics of the reaction, and thus the total dissipation is necessarily
finite.
A final observation is in order. When developing a linear viscous theory, the
restriction imposed by the second law (the viscosity is to be nonnegative) is
somewhat intuitive, and thus the constitutive equation for pressure can easily be
expressed in a form which does not violate the second law. That is not the case
for relaxational systems, for which the restriction imposed is obtained by combin-
ing equations (5.2.8) and (5.2.10): the constitutive functional for pressure must
be expressible as the partial derivative of a constitutive functional for free energy
which enjoys the property in equation (5.2.10).

5.4. ENTROPIC ELASTICITY AND


RELAXATION

When the body is not in a state of thermodynamic eqUilibrium,


its state is not one of those which are represented by our
surface. The body, however, as a whole has a certain volume,
entropy, and energy, which are equal to the sums of the
volumes, etc., of its parts.
w. Gibbs

As discussed before, one of the important areas of application of the theory


of relaxational systems is the description of the behavior of polymers. For
polymeric materials, it is often plausible to assume that the elasticity is entropic
in nature, i.e., that mechanical work done on polymers by deformation may be
accumulated as elastic energy via a decrease of conformational entropy, without
the internal energy being affected. This idea, which is based on molecular
considerations, can be incorporated formally in the theory as discussed below.
148 A relaxational system is said to have entropic elasticity if the internal energy
depends only on the site:
Chapter Five
U = u(a) (SA.l)

Since the equilibrium properties depend only on the site, equation (5.4.1)
implies that the internal energy of a system with entropic elasticity is always
equal to the equilibrium internal energy:

U = U* = u*(a) (SA.2)

The constitutive equation for entropy can be written as

S = s*(a) + s{a; a'(s)} (SA.3)

The value of the functional in equation (5.4.3) represents the excess entropy in
non equilibrium states, and is zero at equilibrium.
The free energy is given by

A = u*(a) - Ts*(a) - Ts{a; a'(s)} (SAA)

Partial differentiation with respect to T yields

S = -au*/aT+ Tas*jaT+ s* + s + Tas/aT (S.4.5)

The first two terms on the right-hand side cancel, because at eqUilibrium equation
(2.7.9) holds. The third and fourth terms cancel with the left-hand side, hence

as/aT = 0 (SA.(j)

i.e., the excess entropy is independent of the present temperature (but does
depend on the history of the temperature):
S = s*(a) + s{V; a'(s)} (SA.7)

Pressure is calculated from equations (5.2.7) and (5.5.4) in the form

p = p* + T as/ av (SA.8)

and, in view of equation (5.4.7), one concludes that the excess pressure p - p*
is proportional to absolute temperature. Rubbery solids are not elastic, since they
can dissipate mechanical energy, and in fact they are best described as relaxational
systems; however, they have entropic elasticity, and extra stresses are indeed
proportional to absolute temperature.
The second law for relaxational systems with entropic elasticity reduces to

8s{a; a'(s)l-a"(s)};:: 0 (SA.9)

This implies that, if the site is held constant, the entropy can never decrease, and
hence that of all states corresponding to any given site, the equilibrium state has
the largest entropy (the excess entropy is never positive).
It is possible to verify experimentally that a given polymeric system is indeed
endowed with entropic elasticity; in an adiabatic deformation it will heat up by
an amount corresponding to the total work done on it, and this can be checked. 149
Also, in an adiabatic stress relaxation experiment, temperature is predicted to
Thermodynamics
stay constant, and this can again be verified. Polymeric melts do in general exhibit of Relaxation
the predicted behavior.
A polymer melt undergoing deformation has, on the basis of the discussion
above, a lower entropy and hence a higher chemical potential than the same
polymer at rest. This is shown schematically in Figure 5.4.1, where the chemical
potential is plotted as a function of temperature at a constant pressure. The melt
at rest is represented by the lower curve on the right, and the melt undergoing
deformation by the upper curve at the right. The latter curve has a lower absolute
value of the slope, since the latter is in fact the entropy. The curve on the left
represents the free enthalpy of the solid; its intersection with the lower melt curve
identifies the eqUilibrium crystallization temperature TM. The curve for the melt
undergoing deformation intersects the solid curve at a temperature TJ > TM : as
a result, a polymer melt may crystallize at a temperature exceeding the equilibrium
one. This phenomenon of flow-induced crystallization has been observed experi-
mentally in polymer melts and in polymer solutions.
The dashed curve in Figure 5.4.1 represents the free enthalpy of a crystalline
form (II) other than that (I) of the full curve on the left. It should be noted that,
if the melt is at rest, when T > TM the liquid is the equilibrium phase; when
T < TM, the crystalline form I is the equilibrium one, and therefore the melt will
crystallize to form I-there is never a driving force for formation of the type-II
crystal. However, when the melt is undergoing deformation, if the temperature
is between TJ and T2 , a driving force for the formation of the type-II crystal
exists, and therefore the melt may indeed solidify into type-II crystals. At tem-
peratures lower than T3 , there is a driving force for a solid-solid phase transi-
tion from type-II to type-I crystals; however, solid-solid phase transitions are
extremely slow in polymers, in fact so slow that to all practical purposes one can
regard them as not taking place at all. It follows that, if indeed the melt has
solidified into type-II crystals, these will then be stable when cooled. Indeed, the

flowing melt

quiescent
FIGURE 5.4.1. Illustration of the phe- melt
nomenon of flow-induced crystalliza-
tion in a polymer. Crystalline form II
would never be observed in crystalliz-
ation from a quiescent melt. T
150 crystalline form obtained by flow-induced crystallization of polymers is often
Chapter Five
different from that obtained by solidification of the quiescent melt.

5.5. PROPAGATION OF DISCONTINUmES

Errors seem to propagate nowadays more rt!adily than truths,


the medium being less resistant, or, to use an archaic term,
more sympathetic, because more like in kind to that which
perturbs it.
C. A. Truesdell

The theory of elastic systems immediately yields the equations governiqg


the propagation of discontinuities. The concrete example of major interest 'is
related to the propagation of sound, which is discussed in the Appendix to Section
8.2. For a one-dimensional problem where z is the only coordinate of interest,
if ell is the density and u the velocity, the linearized equations of motion yield
by cross differentiation

(5.5.1)

where L is the property which is constant across the discontinuity. Equation


(5.5.1) is hyperbolic, and it admits discontinuous solutions, which propagate with
velocity (iJP/iJcIl)1/2. The usual assumption is that the velocity of sound is high
enough that heat flux can be neglected, so that L is identified with entropy, i.e.,
the velocity of sound is equal to the square root of the inverse of the isoentropic
compressibility.
The question of interest here is not whether the isoentropic, or the isothermal,
compressibility is of importance. The important point is that, as long as an elastic
theory is accepted, one obtains equations which are hyperbolic in nature; and
thus they admit discontinuous solutions, and an acoustic theory is thus available.
However, since the theory is elastic, it does not allow for dissipation, and the
damping of discontinuities cannot be described. The situation is analogous to
that encountered in the hydrodynamics of perfect fluids (i.e., fluids endowed
with zero viscosity): the Euler equations of motion do not allow for ellergy
dissipation (because the constitutive equation assumed for the internal stress is
essentially elastic), and hence paradoxical results are obtained whenever one
seeks the value of a quantity related to energy dissipation, like the drag force for
flow around a submerged object, or the presl\ure drop for flow in a constant
section tube. The best example is that of an aerofoil: the Euler equations will
predict rather well the lift force (which is nondissipative, since it is orthogonal
to the direction of motion), but will predict a zero drag: one could design the
shape of the wings of a plane rather well, but as far as designing the engine is
concerned the Euler equations would predict that the airplane doesn't really need
an engine to fly on a horizontal course, since all the engine's power on a horizontal
course is dissipated.
Of course, it would appear to be easy to allow for the damping phenomenon
by moving from an elastic to a viscous theory. The conceptual difficulty with the
problem of the propagation of discontinuities, however, is that, in many cases,
recourse to a viscous theory results in the total loss of the ability of the governing 151
equations to describe the propagation of discontinuities. For example, if, in the
Thermodynamics
derivation given in the Appendix to Section 8.2, one chooses a linearly viscous of Relaxation
constitutive equation for pressure, say
p = p*(If>, L) + BIf> (5.5.2)

the same linearization procedure produces the differential equation


iilf>/at 2 = (ap*/alf iilf>/az 2 + B a31f>/az 2 at (5.5.3)

which does not admit discontinuous solutions. This was observed as long ago as
1901 by Duhem, who introduced the concept of a quasi-wave, a region of very
steep gradients but no discontinuity. In effect, while the Navier-Stokes equations
of classical fluid mechanics indeed do not allow for discontinuities, it is not true
that any dissipative theory (not even any linear one) smoothes out discontinuities.
The equation for a flexible string with linear friction has a structure which, in
terms of pressure and density, could be written as
(5.5.4)

which is hyperbolic and does admit discontinuous solutions. The velocity of


propagation is still (ap/iJlfi!2, but now a damping mechanism arises from the
second term on the left.
Now one should bear in mind that systems with zero relaxation time do not
in reality exist: as was discussed in Section 5.1, even the Maxwellian theory of
gases results in a finite (albeit exceedingly small) relaxation time. It follows that
theories which do not allow for relaxation can only be regarded as valid when
the time scale of the phenomena under consideration is very large as compared
to the intrinsic relaxation time, so that relaxation phenomena may be regarded
as essentially instantaneous. When the phenomenon considered is the propagation
of a discontinuity, the time scale is essentially zero-and hence relaxation
phenomena cannot be excluded from consideration, no matter how small the
relaxation time of the material under consideration. This argument can be quan-
tified by examining flame fronts, for which an estimate of the residence time of
the gas in the shock layer (which is mathematically described as a discontinuity)
is available. The estimate is about 10-9 s, which is not very different from the
estimate of the relaxation time of an ideal gas obtained from the Maxwellian
theory of gases.
The theory of the propagation of discontinuities in relaxational systems is
very complex, and only some essential results will be presented here. Before
doing so, it is useful to recall some general results of the theory of elastic systems.
Let p = f(lf be the constitutive function for an elastic system, where entropy is
regarded as constant (to within accepting the argument that the discontinuity
propagation is isoentropic). The quantities p and If> need not necessarily be
pressure and density; p could be some appropriate measure of stress and If> of
strain. The first derivative df/ d <I> is called the tangent elastic modulus, E. Given
a finite variation If>' - <1>, the value of p will correspondingly change by an amount
p' - p. The ratio (p' - p )/(<1>' - <1 is called the secant elastic modulus E', which
of course depends both on <I> and on <1>'. The tangent modulus is the limit of the
secant modulus when <1>' - If> approaches zero. The tangent and secant modulus
152 coincide with each other only if the p(eII) function is linear; even for an ideal
gas, at constant entropy p(eII) is not linear.
Chapter Five
Two types of discontinuity can be considered. The first, which is called a
shock wave, is where ell itself is discontinuous by an amount ell' - ell. The second,
which is called an acceleration wave, is where the time derivative of ell is
discontinuous, but ell itself is not. The square of the speed of propagation of a
shock wave in an elastic system is equal to the ratio of the secant modulus
corresponding to the density jump to the density, while the square of the speed
of propagation of an acceleration wave is equal to the ratio of the tangent modulus
and the density.
These results hold also for relaxational systems, provided the tangent
modulus is interpreted as the partial derivative ofthe functional for p with respect
to the instantaneous value of ell, and the secant modulus is interpreted
analogously-i.e., at constant past history of ell. However, a damping mechanism
is now present, and in addition to this acceleration waves are predicted to possibly
grow into shock waves in a finite time, provided the initial amplitude of the wave
is sufficiently large, and provided certain coefficients in the constitutive functionals
have appropriate signs.

EXAMPLES AND PROBLEMS

Examples
1. Consider a mixture of two optical isomers A and B, and suppose that it behaves
as an ideal gas. Since the isomerization reaction is equimolar, volume has no influence
on the eqUilibrium composition. Furthermore, since the total number of moles is indepen-
dent of composition, so is pressure, and hence as long as only pressure is measured the
observed behavior is indistinguishable from that of a pure component. Suppose, however,
that the polarizability m is measured; this does depend on composition.
We choose a basis of one mole of A + B, and we define the extent of reaction as the
number of moles of B, so that 0 s x s 1. The equilibrium condition is
x"/(l- x*) = K(T) (5.E.I)

and assuming that the kinetics of the isomerization reaction is of mass action type
dx/ dt = k(1 - x) - kx/ K = (kx*)(x* - x) = F(T)[K(T) - x] (5.E.2)

This can be rearranged to the form


x + f(T) dx/dt = K(T) (5.E.3)

where f( T) = 1/ F( T). Equations (5.E.2) is of the same form as equation (5.1.15). If one
makes the (quite reasonable) assumption that m is a linear function of X, a similar equation
will apply to m as well, i.e., the mixture will exhibit relaxation behavior, as far as the
polarizability is concerned, in response to variations of temperature.

Problems

5.1. A box of variable volume V initially contains I mole of component A. Component A may
dissociate according te,
A=2B
(the nitrogen oxide case discussed in Chapter 11 is a good example). Temperature is held constant 153
at all times. The following assumptions are justified:
Thermodynamics
I. The gas in the box is ideal. of Relaxation
2. The activity-based equilibrium constant of the reaction is unity.
3. The reaction is regulated by mass action kinetics.

a. Define an extent of reaction x such that its value is zero at time zero and its maximum possible
value is I.
b. Determine the value of the equilibrium extent of reaction x* = x*( V, T).
c. Determine the value of the equilibrium pressure,p* = p*( V, T). What is the value you calculate
when V approaches infinity? Explain.
d. Express the kinetics as dx/dt = reV, T,x).
e. Express pressure as p = f( v, T, x).
f. Eliminate x between the results in (d) and (e) to obtain a differential equation for pressure.
This equation may contain, of course, time derivatives of p and of V, but it should not contain x.

LITERATURE

Nostalgia isn't any more what it used to be.


Simone Signoret
The literature on the thermodynamics of systems endowed with finite relaxation times
is generally difficult to read, because of its mathematical complexity, which is related to
the fact that such systems generally exhibit internal stresses that are very significantly
different from a simple isotropic pressure. The basic references are B. D. Coleman and
W. Noll, Arch. Ratl. Mech. Anal. 13, 167 (1963), and B. D. Coleman, Arch. Ratl. Mech.
Anal. 17, 1 (1964).
A somewhat simplified introduction to the subject is given in G. Astarita, An Introduc-
tion to Non-Linear Continuum Thermodynamics, SpA Ed. di Chimica, Milan (1975).
In this chapter, the subject has been approached by considering phenomena where
the finite relaxation time is related to a reasonably well understood kinetic process which
is taking place. The relationship between a phenomenological, memory-type description
and one which chooses internal state variables is discussed in G. Astarita and G. C. Sarti,
Polym. Eng. Sci. 16, 490 (1976), and G. C. Sarti and G. Astarita, Trans. Soc. Rheol. 19,
215 (1975).
Relaxation phenomena are of particular importance in the case of polymeric materials.
An ample literature where the statistical mechanics of polymeric media is developed, and
the thermodynamic consequences thereof are analyzed, is available. The following referen-
ces are of importance: M. Doi and S. F. Edwards, "The Theory of Polymer Dynamics,
Clarendon Press, Oxford (1986); R. B. Bird et al., Dynamics of Polymeric Liquids, Wiley,
New York (1977); and G. Marrucci, Advances in Transport Processes, Vol. 5, Wiley, New
York (1988).
When speaking of systems endowed with finite relaxation times, the question of what
is meant by entropy outside of an equilibrium (or possibly "quasi-static") situation is
bound to emerge. First of all, the idea that one Can write a "balance" equation for entropy
(though of course including the requirement of a nonnegative generation term) is by now
almost 100 years old. See J. Jaumann, Wiener Akad. Sitz. 120,385 (1911); Wiener Akad.
Wiss. Denkshr. 95, 461 (1918); and K. Lohr, Wiener Akad. Wiss. Denkshr. 93, 339 (1917);
99, 59 (1924). On these papers K. Denbigh, The Thermodynamics of the Steady State,
Methuen, London (1951), remarks that: "Their discussion covers some five hundred pages
of the German text and it is perhaps for this reason that their work has been somewhat
overlooked."
154 Inclusion of a generation term of course implies not only that a "balance" equation
for entropy can be written under nonequilibrium conditions, but also, and more funda-
Chapter Five mentally, that entropy has a meaning outside of equilibrium. The literature on the statistical
mechanics of polymeric systems cited above is of help in visualizing such entropy as being
related to the conformational statistics of the systems, which may be nonequilibrium.
The question of entropic elasticity is well known in the equilibrium theory of rubbers.
Its extension to nonequilibrium situations of general polymeric materials is discussed in
G. Astarita and G. C. Sarti, in: Theoretical Rheology (J. F. Hutton, J. R. A. Pearson, and
K. Walters, eds.), Applied Science, Barking (1975).
Propagation of discontinuities in relaxational systems is discussed in B. D. Coleman
et al., Wave Propagation in Dissipative Materials, Springer-Verlag, New York (1965).
The discussion in this chapter shows that, for all the systems considered, it is possible
to move from one equilibrium state to another via a sufficiently slow transformation in
order to make the energy dissipation as small as one wishes. While this supports the
success of thermodynamic analyses which use some handwaving argument of the "quasi-
static process" type, it does not guarantee that such arguments are bound to be invariably
successful.
Chapter Six

SURFACE THERMODYNAMICS

The boundaries between phases are not surfaces in a


strict mathematical sense, but are very thin regions, in
which the properties change with great abruptness from
the properties of the one homogeneous phase to those of
the other. Ordinarily these thin transition regions
between phases contain relatively so small an amount of
substance that they may be entirely ignored. But when
we study surface tension, adsorption, and kindred
phenomena, they become of great importance.
O. N. Lewis and M. Randall

NOTATION

A Free energy density kcal kmol- I


a,( Constitutive function for At kcal
D Difference operator
d Surface rate of strain S-I

f( Function delivering V in surface layer m3 kmol- 1


df Differential surface force kgms- 2
L Any extensive property
dl Line element on surface m
m Mass kg
n Coordinate orthogonal to interface m
n Number of moles vector kmol
p Pressure atm
h Capillary pressure atm
q Defined in equation (6.2.5) kmol
R 1 ,R2 Curvature radii m
s Surface tension kgs- I
s* Equilibrium surface tension kgs- I ISS
156 s Surface tension tensor kgs-2
Chapter Six tl; Tangential stress difference kgm- I S-2
V Specific volume m3 kmol- 1
v Surface velocity ms- I
W Work rate kgm2 S-3
y Mole fraction vector
z Mass fraction of phase I
r Relative adsorption vector kmolm- I
e Thickness of surface layer m
J.L Chemical potential vector kcal kmol- I
I-l; Surface shear viscosity kgm- I S-I
I-l; Second surface viscosity kgm- I S-I
I. Surface area m2
cf> Contact angle rad
1 Two-dimensional unit tensor

Subscripts

L Liquid
G Gas
S Solid
t Total
1 Phase 1
2 Phase 2

Superscripts

I. Surface
I Phase I
II Phase II
Substantial time derivative
6.1. SURFACE TENSION

That such pygmies should cast such giant shadows only shows
how late in the day it has become.
L. Chatgaft

When a system exists in a two-phase state, surface tension is exhibited at interfaces.


Let us consider in particular a gas-liquid system consisting of a single liquid
drop in contact with a gas. Let s be the surface tension, and l: the total interface
area. It is possible to do work on such a system without changing its volume,
but simply changing its interface area. For instance, the total rate of work done
on the system can be expressed as

~= -pV; + sl:" (6.1.1)

One is now tempted to generalize the definition of the state of, say, the liquid
phase by including the surface area among the external state variables, for
instance, to write a constitutive equation for free energy as follows:

At = a.( V;, T, l:) (6.1.2)

However, this cannot be done because free energy is an absolutely additive


function of mass, and the interface has no volume and hence no mass. The
problem here is that the concept of an interface as a surface of discontinuity,
while useful for the description of many properties of multiphase systems, is not
in fact realistic, and it leads to paradoxes if applied to the interface itself.
Actually, in the vicinity of what is perceived as an interface within two
phases, the density does not suffer a discontinuity but only exhibits a very steep
gradient over a very short distance e in a direction orthogonal to the interface.
The distance e is so small as compared to the radius of curvature of interfaces
encountered in practical cases that, in describing the surface layer, one can always
regard it as being plane.
In the sense just discussed above, the mass of the surface layer is small but
finite, and hence it is legitimate to consider quantities such as the free energy of
the surface, or the number of moles of each component in the surface layer, and
so on. All such quantities will be expressed per unit interface area; e.g., if ml'. is
the mass of surface layer per unit area, L l'. is the surface density of the extensive
quantity L, to be interpreted essentially as the actual average density in the surface
layer mUltiplied by ml'.. In this section, attention is limited to one-component
systems, so that the question of the composition of the surface layer does not arise. 157
158 Given two phases at equilibrium with each other, and characterized by
volumes VI and V2 such as in Figure 4.1.1, we arbitrarily choose to consider the
Chapter Six
entire surface layer as part of phase 1. Let n be the coordinate orthogonal to the
interface, with n = 0 being the edge of phase 2, and the surface layer extending
to n = E. The specific volume in the surface layer is given by

V=f(n); f(O) = VI and (6.1.3)

The mass per unit interface area of an element located between n and
n+dnis

dm~ = dn/f(n) (6.1.4)

The free energy in the surface layer is different from (and, in view of Figure
4.1.1, larger than) that of phase 1. If one is willing to assume that the a( V)
function is the appropriate constitutive equation for free energy also in the surface
layer, the free energy per unit area, A~, is calculated as

A~ = t [a(f(n - a(VI)] dn/f(n) (6.1.5)

and the total free energy of the two-phase system is thus given by

A, = m,[za(VI) + (1- z)a(V2 )] + A~l: (6.1.6)

The important point is that the quantity A~ may depend on temperature and
pressure, but not on the interface area l: itself. In other words, the fact that free
energy is an absolutely additive function of mass guarantees that, insofar as it
may depend on interface area, it does so linearly.
One can now go through the usual argument, and conclude that in fact the
surface tension is the derivative of free energy with respect to l::

s = aa,/al: = A~ (6.1.7)

if the system is elastic.

LIQUID GAS

FIGURE 6.1.1. Contact angle at a liquid-solid


SOLID interface.
Now consider the case where the interface is curved. The two radii of 159
curvature are to be large as compared to e for the surface tension concept to Surface
apply; however, this is an easily satisfied condition. A simple mechanical argument Thermodynamics
shows that, at mechanical equilibrium (no driving force for motion), pressure in
the two phases is not the same. The pressure difference p~ is called the capillary
pressure and is given by

(6.1.8)

We consider a liquid phase in contact with both a gas phase and a solid
phase (such as a liquid drop on a solid plate), and let the solid surface be plane.
As can be seen from Figure 6.1.1, a contact angle c/J can be identified. At
equilibrium, the contact angle is given by
(6.1.9)

Angle c/J can be both larger and smaller than 90, according to the relative
magnitude of the solid-gas and solid-liquid surface tensions. When c/J > 90, the
liquid is said to wet the solid. If a capillary is inserted vertically into a pool of
stagnant liquid, if the liquid wets the capillary it will rise to reach an equilibrium
height where the capillary pressure is balanced by gravity; if it does not wet the
capillary, it will dip to a depth determined by the same condition. If a capillary
is inserted horizontally into a liquid pool, a wetting liquid will move indefinitely,
if at a progressively lower rate, into the capillary.
Now we consider a liquid drop (phase 1) floating on the surface of another
liquid (phase 2); see Figure 6.1.2. At equilibrium, the following relationship
applies:

(6.1.10)

This can be sati~fied provided that

(6.1.11)

If condition. (6.1.11) is not satisfied, the drop will tend to spread indefinitely
on the surface of liquid 2. In actual fact, the spreading is not indefinite: As the
thickness of the surface film approaches that of a molecular monolayer, the
tendency to spread out disappears (this was observed by Benjamin Franklin, who
allowed a drop of olive oil to spread over the surface of the Clapham Common
pond in 1765. The final thickness was calculated at about 2.5 x 10-9 m, which is
close to the molecular dimensions of olive oil).

GAS

FIGURE 6.1.2. A liquid drop floating on 2


the surface of another liquid.
160 The phenomenon of the spreading of a film down to possibly monolayer
thicknesses can also take place on the surface of a solid phase. Equation (6.1.9)
Chapter Six
cannot be satisfied if

(6.1.12)

and if that is the case, the film-spreading tendency will be observed.


The considerations developed up to this point were limited to static (Le.,
equilibrium) conditions. Under dynamic conditions, Le., when the superficial
velocity is neither zero nor uniform over the whole interface, the surface
phenomena are remarkably more complex.
Just as a fluid phase in the bulk cannot at equilibrium sustain any stress
other than an isotropic pressure, the interface between two fluid phases at
equilibrium cannot sustain any other surface stress than an isotropic surface
tension. However, under dynamic conditions that is no longer true, and a surface
tension tensor s needs to be considered. The latter is defined as follows. Consider
a point on the surface, and a differential line element dl through that point; dl
is a (two-dimensional) vector, orthogonal to the line element and with modulus
proportional to its length. A force df (which is tangential to the surface) acts
through the line element considered, which need not be (as it would at equilib-
rium) orthogonal to it. The'surfce tension tensor is defined as

df=sdl (6.1.13)

If s* is the equilibrium surface tension, at equilibrium the surface tension


tensor is isotropic and is given by

s = s*1 (6.1.14)

where 1 is the two-dimensional unit tensor.


Let v be the two-dimensional surface velocity vector. Under dynamic condi-
tions, grad v is not zero. Let d be the surface rate of strain, Le., the symmetric
part of grad v. Since equation (6.1.14) holds when d is zero, the simplest constitu-
tive equation for the surface tension tensor is the following linear one, perfectly
analogous to equation (2.3.A.1):

(6.1.15)

where JL~l and JL~ are the two surface viscosities.


Let us now consider the case where the motion of the surface is a pure
two-dimensional dilatation or compression, so that d is given by

d = !(divv)1 (6.1.16)

and thus the surface tension tensor is given by

(6.1.17)
The quantity in square brackets could be regarded as a dynamic surface 161
tension, and it is seen to coincide with the equilibrium surface tension only if Surface
the rate of dilatation is zero. The second law imposes the following requirement: Thermodynamics

(6.1.18)

i.e., the work done on the system when the interface area is being increased is
no less, and may be more, than the work the system does when the interface is
being decreased. There is no analog of the Stokes hypothesis that the left-hand
side of equation (6.1.18) may be zero; indeed, finite values of it have been
measured.
Now consider a pure shear motion of the interface, say one where div v = O.
The surface tension tensor is now given by

s = s*1 + 2J.t1:d (6.1.19)

and the second law requires that

(6.1.20)

Equations (6.1.18) and (6.1.20) exhaust the requirements imposed by the


second law on the linear constitutive equation for the surface tension tensor.
When a thin film of a second liquid phase is spread over the surface, J.t1: has
been interpreted as the bulk viscosity of the film multiplied by the film thickness.
Since values of J.t1: as large as 10-6 kg S-1 have been measured with very thin
films, bulk viscosities as large as 102 kg m- 1 S-1 would be estimated from such
measurements.
Not only pressure may suffer a discontinuity at an interface because of
capillarity [see equation (6.1.8)], but also the tangential stress. The difference in
tangential stress across the interface, t1:, is given by

t1: = grad s (6.1.2)

and is thus nonzero whenever the surface tension is not constant over the interface.
When grad s is nonzero, the tangential stress at the interface must be nonzero in
at least one of the two phases. Since a fluid cannot sustain a tangential stress at
equilibrium, when grad s is nonzero at least one of the fluid phases will be set
in motion.
The surface tension may be nonuniform over the surface because of several
reasons, the two fundamental ones being as follows. The first is when temperature
is nonuniform; the corresponding motion is called thermo capillary motion. The
second is when concentration is nonuniform on the interface; as may well be the
case during mass transfer phenomena; the corresponding motion is called the
Marangoni effect.
162 6.2. SURFACE PHENOMENA IN MIXTURES
Chapter Six
Certainly, one is building on an insecure foundation, who rests
his work on hypotheses concerning the constitution of matter.
Difficulties of this kind have deterred the author from
attempting to explain the mysteries of nature, and have forced
him to be contented with the more modest aim of deducing
some of the most obvious propositions. ... The only error into
which one can fall, is the want of agreement between the
premises and the conclusions, and this, with care, one may
hope, in the main, to avoid.
J. E. Gibbs

In this section, we consider the surface conditions of a system where there


is more than one component, and two phases. Again, one should consider the
fact that the surface is not a two-dimensional surface in the mathematical sense,
but a very thin layer of extremely small but finite mass. There are, however,
distinct advantages in regarding it as a surface, and the thermodynamic theory
is developed by substituting for the real system, which may be viewed as composed
of two phases and a surface layer, with an "ideal" model system consisting of
two bulk phases and a surface-two-dimensional in the analysis, but yet endowed
with finite mass.
With such a philosophical approach, the total number of moles vector can
be thought of as decomposed in three parts, say

(6.2.1)

where nl: is the vector of the number of moles at the surface. The whole analysis
is developed by assuming that

(6.2.2)

so that all the results obtained previously for the bulk phases still hold. In
particular equation (4.2.4), which plays a crucial role in the establishment of the
phase eqUilibrium conditions, is still regarded as essentially valid because of
condition (6.2.2).
Equation (6.2.1) can be expressed in the equivalent form

(6.2.3)

which shows that there is a fundamental ambiguity in the meaning of the vector
nl:. Indeed, if equation (6.2.2) holds true, the left-hand side of equation (6.2.3)
is the difference between two large quantitites which are almost equal to each
other. It follows that any even very minor ambiguity in the choice of how many
total moles to assign to phase I and phase II results in a very major difference
in the value of nl:.
The ambiguity can be resolved by a number of different arbitrary conventions.
The usual one is suggested by the considerations of dilute solutions, so that one
particular component (the solvent) plays a special role. Let 1 be the component
chosen as a reference. Furthermore, let I be the phase to which the surface layer
is attributed (usually the more dense phase, say the liquid in a gas-liquid system),
and let D be the operator defined as the difference between the quantity operated 163
upon in phase I and in phase II. Equation (6.2,3) can be rewritten as
Surface
Thermodynamics
(6.2.4)

Let q be defined as follows:

q = n;/ DYI (6.2.5)

Then equation (6.2.5) can be rearranged to yield


nl: - qDy = (n - nyI) - (n l - nyDDy/ DYI (6.2.6)

The right-hand side of equation (6.2.6) is independent of the number of


moles assigned to each individual bulk phase, and hence so is the left-hand side.
One can therefore regard the vector on the left-hand side as an unambiguous
number of moles at the interface vector, referred to component 1. In particular,
if that number of moles is divided by the interface area ~, one obtains the vector
of the relative adsorptions at the interface:
r = (nl: - qDy)/~ (6.2.7)

By definition, the relative adsorption of component 1 is zero.


By analogy, given any quantity L, its relative surface density is defined as
Ll: = (L~ - qDL)/~
(6.2.8)

which again has a value independent of the number of moles assigned to each
individual bulk phase.
Of course, different conventions could be used to define properly unam-
biguous surface properties, and such different conventions would of course result
in different values for the surface quantities. It is therefore important to keep in
mind that, although the equilibrium surface tension is still the partial derivative
of the total free energy with respect to interface area, it is not equal to Al:, except
for one particular convention which is not the one usually adopted. Indeed,
chemical potentials at the interface can be defined as follows:

(6.2.9)

which can be shown to be independent of the convention used [both Al: and r
depend on the convention chosen, but the derivative in equation (6.2.9) does
not]. Some cumbersome algebra, which essentially duplicates that relative to
mixtures discussed in Section 3.3, yields the following equilibrium results:

s=Al:-r"l: (6.2.10)

r = -as/a"l: (6.2.11)

"l: = "I "II


= (6.2.12)

Equations (6.2.11) and (6.2.12) together afford a means of measuring the relative
adsorptions at the interface.
164 It is important to realize that, in the case of mixtures, there are in fact two
types of phenomena which may result in non-equilbrium surface properties. The
Chapter Six
first type is connected wit~ the actual possibility of motion of the interface, and
is thus of the same type as discussed in the previous section. The second type is
connected with the fact that, as new interface area is formed, it may well in
general form at a composition quite different from that at equilibrium. This in
turn implies that equilibration now requires individual components to diffuse
toward or from the interface, a process which may well be rather slow; of course,
as long as the interface composition is different from the equilibrium one, so are
all other surface properties. This is particularly relevant in the case of solids,
where diffusion from and to the interface may be extremely slow, so that the
actual surface composition may differ from the equilibrium one for substantial
lengths of time. The catalytic properties of solid surfaces are strictly related to
this type of phenomenon.

Appendix

The whole question of the arbitrariness of the definition of excess surface


properties is perhaps best illustrated by considering a simple example of a binary
dilute solution. The concentration profiles in the interface region for both solute
and solvent are sketched in Figure 6.2.1. The (arbitrary) position of the
" equivalent" surface can be chosen (n = s) in such a way that the hatched area
for the solvent concentration on both its sides is the same; with that choice, the
solvent excess concentration is of course zero.
Since the two concentration profiles do not need to have the same shape,
such a choice will make the hatched area for the solute concentration on both
sides of the surface unequal. For the case drawn in Figure 6.2.1, one obtains a
positive excess concentration of solute at the interface, since the area on the right
(positive) is larger than the area on the left (negative).
While the choice of solvent as the reference component is natural in the case
of dilute solutions, for concentrated solutions there is an obvious arbitrariness.
Furthermore, one could also choose to place the surface at such a position that

c
50lvent

solute

5 pOSition n
the sum of all areas on the right equals the sum on the left. In addition, one 165
could use mass units rather than mole units, and this selection entails a whole
Surface
new series of possible choices for the arbitrary position of the "equivalent Thermodynamics
surface."

EXAMPLES AND PROBLEMS

Problems

6.1. Suppose a one-component system is brought to a temperature where it exhibits a gas-liquid


transition. The system is held in a constant-volume box, and the volume is such that indeed two
phases will exist at equilibrium. On the earth, the system will separate in two phases, and because
of gravity the liquid will settle at the bottom of the container, with a horizontal interface separating
the two phases. Now suppose the same phenomenon takes place in space in the absence of gravity.
The container is a rectangular box measuring 3 x 2 x 1 feet. What shape will the interface have at
true equilibrium? Do you suppose this true equlibrium condition will in fact be quickly attained?
Discuss the sequence of events you think would take place.
6.2. There is a castle built on a small island off the coastline of Naples; the walls are of tuff,
a porous volcanic rock which is plentiful in the area. The tuff walls are visibly darker up to' a height
of about 15 m above sea level.
Estimate the average pore size of tuff rock.
6_3. Suppose a supertanker of the type built by Japanese shipyards during the oil crisis gets
sunk in the Pacific Ocean, spilling its whole cargo in the process. Estimate the fraction of the Pacific
Ocean area which could be covered by the spill.
6_4. Consider a horizontal capillary tube inserted into a pool of a wetting liquid. Develop the
equations describing the flow inside the capillary. In particular, develop an expression for the flowrate
as a function of time and of the appropriate geometrical and physicochemical parameters.

LITERATURE

Yet what are all such gaieties to me


whose thoughts are full of indices and surds?
x 2 + 7x +53
= 11/3
Lewis Carroll

The classical paper by J. W. Gibbs, "On the equilibrium of heterogeneous substances,"


Trans. Conn. Acad. 3, 108, 343 (1878), includes a long section entitled "Theory of
capillarity-surfaces of discontinuity between fluid masses" which is still today the best
discussion of the equilibrium behavior of interfaces.
The book by R. Defay and I. Prigogine, Tension Superficielle et Absorption, Dunod,
Paris (1951) is difficult to read but covers the subject completely. A recent reference dealing
with both equilibrium and dynamic properties is C. A. Miller and P. Neogi, Interfacial
Phenomena: Equilibrium and Dynamic Effects, Dekker, New York (1985)..
A good reference book, where most problems connected with surface thermodynamics
(including adsorption and catalysis) are discussed thoroughly, is A. W. Adamson, Physical
Chemistry of Surfaces, Interscience, New York (1960).
The book by B. Levich, Physicochemical Hydrodynamics, Prentice-Hall, Englewood
Cliffs (1964), is a good source for dynamic phenomena at interfaces.
Chapter Seven

DISSIPATIVE PHENOMENA

Prigogine and Mazur: "... all coupling between


quantities of different tensorial character being
forbidden . ... "
Kirkwood and Crawford: "We must treat scalars,
vectors and tensors separately, for entities of differential
tensorial character cannot interact (Curie's theorem}."
"What this theorem is, we may have some difficulty in
divining, since the terms 'interact' and 'couple' are not
found in books on algebra, although they do appear
frequently in the The Arabian Nights."
(Truesdell)

NOTATION

A Constant parameter
At Total free energy kcal
a Constant parameter kmolm- I S-I
a "Thermodynamic coordinate"
B Constant parameter K- 1
b Arbitrary vector
b Constant parameter kmolm- 3 s- 1
c Concentration vector kmolm- 3
Cv Constant volume specific heat kcal kmol- I K- 1
D (Section 7.3) Diflusivity matrix m2 s- 1
D (Section 7.4) Rate of deformation tensor 8- 1

E RNA concentration kmolm- 3


e Constant parameter S-I

F Tangential force kgm- I S-2


Fjk Mass-based mobilities kg2 kcal- I m- I S-I

f( Constitutive function for F kgm- I S-2


G Constant symmetric matrix 167
168 G Free enthalpy density kcal kmol- 1
Chapter Seven g Gravity acceleration ms- 2
H Enthalpy density kcalkmol- 1
h Vertical height m
J,J' "Fluxes"
j Entropy flux kcal m- 2 S-1 K- 1
K Kinetic energy density kcalkg- 1
K Thermal conductivity tensor kcal m- 1 S-1 K- 1
k Thermal conductivity kcal m- 1 S-1 K- 1
L Any extensive property density
L' Partial molar L
L" Partial mass L
Ljk Mole-based mobilities kmoe kcal- 2 m- 1 S-1
L "Matrix of phenomenological coefficients"
M Molecular weight kgkmol- 1
M Molecular weight vector kgkmol- 1
N Molar diffusive flux in z direction kmolm- 2 s- 1
N' Total molar flux in z direction kmolm- 2 s- 1
N Diffusive flux vector in z direction kmol m- 2 S-1
Po Matrix defined in equation (7.2.9) kcalkmol- 1
p Pressure atm
Q Diffusive mass flux in z direction kg m-2 S-1
Q' Total mass flux in z direction kg m- 2 S-1
Q* Talandic partition function
q Heat flux in z direction kcalm- 2 S-1
q Heat flux vector kcal m- 2 S-1
R Gas constant kcal kmol- 1 K- 1
r Distance between plates m
S Entropy density kcal kg- 1 K- 1
T Temperature K
T Stress tensor kgm- 1 S-2
Time s
tp Residence time s
U Internal energy density kcal kg- 1
u Relative molecular velocity ms- 1
V Specific volume m3 kmol- 1
V; Total volume m3
v Velocity ms- 1
v' Mole average velocity 169
v Talandic velocity Dissipative
w (no subscript) Work rate Phenomena
w Weight fraction
x Ordinate orthogonal to planes m
X Length scale of hydrostatic distribution m
X "Forces"
y Enzyme concentration
y Mole fraction vector
Z Dissipation rate
z Direction of diffusive flux
a Constant parameter
{3 Constant parameter
(J Relaxation time s
.... Chemical potential vector kcalkmor 1
~ Density kgm-3
u State
<f>J Total chemical potential kcal kg- 1
O( Local talandic internal energy kmoem-6 s- 1

Subscripts

j, k For component j, k
SS Steady state

Superscripts

+ Symmetric part
Antisymmetric part
* Talandic
T Transpose
Substantial time derivative
7.1. HEAT TRANSFER

Such a theorem as "the square of the hypothenuse of a right


angled triangle is equal to the sum of the squares of its sides"
is as dazzlingly beautiful now as it was in the day when
Pythagoras first discovered it, and celebrated its advent, it is
said, by sacrificing a hecatomb of oxen--a method of doing
honor to science that has always seemed to me slightly
exaggerated and uncalled for. One can imagine oneself, even in
these degenerate days, making the epoch of some brilliant
scientific discovery by inviting a convivial friend or two, to join
one in a beefsteak and wine. But a hecatomb of oxen-it would
produce a quite inconvenient supply of beef
Lewis Carroll

As was discussed in Chapter 2, heat transfer is a dissipative phenomenon, and


if it represents the only source of dissipation the second law requires that, at
steady state, the scalar product of the heat flux and the temperature gradient
should be nonpositive:

qgrad T$.O (7.1.1)

[We shall see in more detail in Section 7.5 that the second law reduces to the
requirement in equation (7.1.1) only for steady-state phenomena. In other words,
for unsteady-state phenomena, the second law does not forbid heat to flow in
the direction of increasing temperature, if only for short intervals of time.]
At equilibrium, temperature is uniform in space (Le., grad T = 0), and there
is no heat flux. It is therefore reasonable to assume that the heat flux vector
depends on the temperature gradient. The heat flux is a function of state, and
thus the simplest possible form of constitutive equation for it is

q = q(grad T, V, T) (7.1.2)

This, however, implies that the state is {grad T, V, T}, and hence one would need
to assume that all functions of state depend on grad T. With this assumption, the
second law reduces to

[aA/iiV + plY' + [iiA/iiT + S]T'


+ [iiA/ii grad T] (grad T)' + q' grad T/(iPT) $. 0 (7.1.3) 171
172 The usual argument now shows that the free energy has in fact to be
independent of grad T. Since furthermore the free energy is still a potential for
Chapter Seven
both pressure and entropy, the latter are also independent of grad T, and thus
one concludes that in fact the heat flux vector is the only function of state which
depends on the temperature gradient. We note however that, as will be discussed
in the section on coupling, this conclusion is connected with the peculiarly simple
form assumed for the state, and in more general cases other thermodynamic
properties may well depend on the temperature gradient.
Since q is known to be zero when grad T is zero, the simplest form for
equation (7.1.2) is Fourier's law for isotropic materials:

q = -k(V, T) grad T (7.1.4)

with k( V, T) being the (nonnegative) thermal conductivity.


Matters are considerably simplified by assuming that V is constant, and that
there is no radiant heat, so that equation (1.4.8) reduces to

U = cvT = div[k(T) grad T] (7.15)

which, if the dependency of k and Cv on T is neglected, reduces to the usual


equation stating that the Laplacian of temperature is proportional to the substan-
tial time derivative of temperature itself.
Equation (7.1.4) is based not only on the assumption that the heat flux is
a linear function of grad T, but also that it is an isotropic function of it, i.e., that
the conductivity is the same in all directions. That may very well not be true in
solids, for which Fourier's law becomes, in general,

q = -Kograd T (7.1.6)

where K is the conductivity tensor. Any tensor can be decomposed into the sum
of a symmetric and an antisymmetric tensor:

and (7.1.7)

and the symmetric part K+ is required by the second law to be positive-definite,


i.e., such that for any vector a

(7.1.8)

It is noteworthy that the second law places no restriction on the antisymmetric


part of K. because b (K- b) = 0 identically (K- being a skew matrix). A question
0

which has caused a great deal of discussion in the technical literature is whether
K is necessarily symmetrical for all solids or not. According to the so-called
thermodynamics of irreversible processes (hereafter referred to as TIP), K is
necessarily symmetric. However, this conclusion is debatable, as discussed below.
First of all, if temperatures are prescribed on the bounding surface of the 173
body considered, the resultant temperature distribution will depend only on the
Dissipative
symmetric part of K, and hence no experiment of this type could settle the issue. Phenomena
However, experiments could be carried out where boundary conditions are
imposed on the heat flux rather than on temperature itself (for instance, some
part of the bounding surface could be kept adiabatic), and therefore the issue
could in principle be resolved by experiment. However, there are several experi-
mental difficulties. First, the solid should not be a randomly microcrystalline one,
which would be macroscopically isotropic, but should be macroscopically non-
isotropic, say ideally it should be a single crystal. Symmetries of single crystals
are well known, and for a large variety of crystalline classes the symmetry of K
is simply a consequence of the symmetries of the crystal; however, for some
crystals (like gypsum, erythrite, dolomite, etc.) this is not so. Stokes laid down
the hypothesis that indeed K is symmetric, i.e., that K- = 0, and in fact he wrote
that "it is only among crystals which possess a peculiar sort of asymmetry, that
we should expect to find traces of [its] existence." Very few experiments which
could yield an answer have been performed, and these seem to indicate that
indeed K- = 0; however, there is no thermodynamic reason why it should
necessarily be so for all solids.
While for the case of single crystals the question of whether the conductivity
tensor is or is not necessarily symmetric may seem rather academic, that is not
so in the case of polymers. Polymers, even if isotropic when at rest, may well be
anisotropic when flowing, as evidenced by their exhibiting flow birefringence.
Whether this anisotropy has any effect on the heat transfer in polymers is not
known; it is however potentially of importance, because heat transfer in flowing
polymers occurs in almost all polymer processing operations.

7.2. HYDROSTATICS

To the agency of heat may be ascribed those vast disturbances


which we see occurring everywhere on the earth: the movements
of the atmosphere, the rising of mists, the fall of rain and other
meteors, the streams of water which channel the surface of the
earth, of which man has succeeded in utilizing only a small
part.
s. Carnot
While in the case of heat transfer there is no difficulty as regards the
equilibrium condition, i.e., that temperature is uniform in space, the matter is
not so straighforward in the case of mass transfer, as discussed in this and the
next section.
Thermodynamic equilibria have so far been discussed on the basis of the
hypothesis that, at equilibrium, pressure is uniform in space (as well as in time).
However, this is only true if gravity forces are neglected: in a fluid at rest, pressure
is in actual fact not uniform in space, but it has the hydrostatic distribution.
We begin by considering the case of a one-component ideal gas. The hydro-
static pressure distribution is given by
dp/dh = -Mgp/ RT (7.2.1)
174 where h is vertical height above a reference horizontal plane, g is the gravity
Chapter Seven
acceleration, and M is the molecular weight. A subscript zero identifies values
at the reference plane h = O. Equation (7.2.1) integrates to

RTln(p/po) = -Mgh (7.2,2)

The left-hand side is the difference in the free enthalpy density, or


equivalently the chemical potential. Thus for a single-component ideal gas the
chemical potential has the hydrostatic distribution

(7.2.3)

We note that, when the chemical potential has the hydrostatic distribution,
no mass flux will occur, in spite of the fact that the gradient of chemical potential
is not zero. This point will be discussed in more detail in the next section.
The result in equation (7.2.3) is in fact not restricted to ideal gases. The
hydrostatic pressure distribution is given in general by

dp/dh = -Mg/V (7.2.4)

For anyone-component system in equilibrium (and thus with temperature


being uniform in space), the vertical distribution of chemical potential can be
calculated as follows:

dp,fdh = (Bp,/Bp)(dp/dh) = (V)(-Mg/V) = -Mg (7.2.5)

which of course integrates to equation (7.2.3). Two comments are important in


regard to equation (7.2.5). First, the hydrostatic pressure distribution is linear
only for incompressible fluids, but the chemical potential distribution of a one-
component system is always linear. Second, the chemical potential distribution
is independent of the density of the system-in liquids, the pressure gradient is
much steeper than in gases, but the pressure dependency of the chemical potential
is much weaker, and the two effects cancel out.
Let us now consider the significantly more complex case of mixtures. The
hydrostatic pressure distribution is still given by equation (7.2.4) provided Mis
intended as the average molecular weight:

M=My (7.2.6)

where M is the molecular weight vector, the components of which are the
molecular weights of the individual components.
A philosophical problem arises at this stage. There are in fact two distinct
possibilities: the first is to assume that equation (7.2.5) holds for the free enthalpy
of the mixture, say

dG/dh = -Mg (7.2.7)


This assumption leads to the conclusion that the composition of the mixture is 175
uniform in space. However, it makes more sense to ~uppose that the composition Dissipative
is not uniform, with heavier components tending to settle low, and vice versa. Phenomena
Hence we have the alternate assumption, i.e., that equation (7.2.5) holds for each
individual component, say

dfJ./dh = -Mg (7.2.8)

In the next section, a proof is given that indeed equation (7.2.8), rather than
equation (7.2.7), is the correct one for mixtures.
It is now useful to define the matrix P G as follows:

P G = 8p./8y (7.2.9)

so that the Gibbs-Duhem equation can be written as

P~'y = 0 (7.2.10)

The vertical distribution of the chemical potential vector can be expressed


as follows:

dp./dh = (8p./8p)(dp/dh)+PG dy/dh (7.2.11)

Combination of equations (7.2.4), (7.2.8), and (7.2.11) yields

PGdy/dh = [(V'M/V) - M]g (7.2.12)

This equation determines the vertical distribution of composition. We note that,


if one takes the scalar product of equation (7.2.12) with y, the following condition
is obtained:

y-[PG(dy/dh)] =0 (7.2.13)

By making use of the properties of transpose matrices, equation (7.2.13) becomes

dy/dh(P~y) = 0 (7.2.14)

which, in view of the Gibbs-Duhem equation, is identically satisfied.


It is of interest to estimate the length scale X over which the composition
changes significantly. Consider in particular a binary mixture, for which equation
(7.2.12) simplifies considerably:

(7.2.15)

Since partial molar volumes and average molar volumes are of the same order
of magnitude, the right-hand side is of the order of a molecular weight difference
8M times the gravity acceleration. For an ideal solution, the baric derivative of
the chemical potential with respect to the mole fraction is RT. For nonideal
solution that is of course not true, but for the purpose of an order of magnitude
analysis RT is still a good estimate. One thus concludes that X satisfies

X ""2RT/g8M (7.2.16)
176 At room temperature, 2RT/g is about 5 x lOSmkgkmol- l Therefore, even
with 3M values of order 100 kg kmol-I, the significant length scale is of the order
Chapter Seven
of a few kilometers. One may thus conclude that, except for special cases (like
the distribution of composition in the atmosphere or iIi deep oil wells), the
hydrostatic effect can be negected altogether, and the equilibrium distribution of
chemical potentials is uniform in space.

7.3. DIFFUSION

Since the tum of the century anyone who has set pen to paper
in an attempt to advance thermodynamics has come under
attack from one quarter or another, and the only thing upon
which we all agree is that Gibbs was a very smart fellow. So,
not knowing what to make of the battles raging around us, we
, opt for neutrality: we confine our teaching to the substance and
style of 19th century thermodynamics. Although this course of
action has served us reasonably well and, incidentally, lends to
the subject an undeniable charm, at some point we must ask if
such a state of affairs is to prevail forever.
M. Feinberg
The nonequilibrium theory of mixtures is remarkably complex, and only a
conceptual sketch will be given here. The complexity is twofold: conceptual and
mathematical. Therefore, in order not to cloud the conceptual issues involved,
attention is restricted to a geometrically very simple case.
The conceptual difficulties are related to two problems, which are connected
with each other. The first problem is the following one. In a mixture where
diffusion occurs, the components are in relative motion with each other, and so
there is some ambiguity about what is meant by the average velocity of the mixture
as a whole. As will be discussed below, this ambiguity can in fact be resolved,
based on the requirement that the equations of motion for the mixture as a whole
should be written in the usual form, say for a Newtonian fluid in the form of
the Navier-Stokes equations. 'However, when the ambiguity is resolved in this
way, the diffusive fluxes need to be defined in a way which is at odds with that
usually adopted in the theory of mass transfer.
The second problem is related to the fact that, in the presence of diffusive
phenomena, the concept of a body (as a closed system which does not exchange
matter with its surroundings) is in a sense lost: even if one considers the neighbor-
hood of a material point as in describing the motion of the mixture as a whole,
no matter how one allows such a neighborhood to deform in time, it still exchanges
matter with its surroundings. This implies that, when writing balance equations
(or trended imbalances like the second law) in the Lagrangian form (Le., with
substantial rather than partial time derivatives), one still has to take into account
convective fluxes. Indeed, the literature rooted in mechanics chooses to regard
what chemical engineers would consider convective fluxes as conductive fluxes,
so as to retain the characteristic that, in the Lagrangian form of balance equations,
only conductive fluxes are involved.
The simplifications introduced are as follows. First, a one-dimensional prob-
lem is considered, Le., all quantities are regarded as variable only along one
direction, say z, Second, temperature is assumed to be uniform in space. Third, 177
some of the conceptual points are discussed by restricting attention to two special
Dissipative
cases: a binary mixture of ideal gases, and a binary liquid mixture the density Phenomena
of which is independent of composition. Finally, at the end of the analysis the
further simplification of considering only steady-state phenomena is made.
Let us consider first the very simple case of a one-component system moving
at some velocity v(z) (if v indeed varies along z, the density needs to be changing
in time). The mass flux Q', i.e., the mass crossing the plane orthogonal to z at z
per unit area and per unit time, is simply the product of the density times the
velocity, Q' = cl>v. The molar flux N' is of course simply Q'/ M, with M the
molecular weight. Of course, one would not regard Q' or N' as manifestations
of what we intend by diffusion: they simply represent the actual motion of the
system.
The next step is to consider a mixture which moves as the single component
considered above, with nothing else happening. In other words, all the com-
ponents of the mixture move with the same velocity. The mass of component j
per unit volume of the mixture, cl>j' is

(7.3.1)

where Wj is the weight fraction (the values of Wj naturally sum to unity). Thus
one calculates the fluxes:

and (7.3.2)

where cj = cl>wj/ ~ is the concentration of the jth component. The total mass
flux is

Q' = L Q; = cl>v (7.3.3)

N' = L N; = v L cj (7.3.4)

Again, one would not regard these fluxes as being diffusive fluxes, since they
simply reflect the motion of the mixture as a whole.
Diffusive fluxes arise when the individual components move with different
velocities. Let Vj be the velocity of component j; one has

(7.3.5)

Thus the average velocity of the mixture is recognized as

(7.3.6)

The diffusive mass flux of component j is the difference between its actual
mass flux and the flux it would exhibit if the mixture simply moved as a whole,
say

(7.3.7)
178 and of course the diffusive mass fluxes add up to zero. The quantities Uj are the
relative diffusion veloities.
Chapter Seven
The molar fluxes are

(7.3.8)

(7.3.9)

We note that while the quantities Qj sum to zero, the ~ do not. In order to
understand this point, consider a binary liquid mixture the density of which is
independent of composition. Let such a mixture rest in a long cylinder with
sealed ends, with diffusive fluxes taking place. Since density is constant, no net
mass flux takes place, and hence v = o. The two mass fluxes must necessarily be
equal and opposite to each other. However, the total molar flux is not zero, since
the two molar fluxes will be in inverse proportion to their molecular weight.
The point discussed above is related to the interpretation of the velocity of
the mixture as in equation (7.3.6), i.e., as the mass-average velocity. Of course,
one could have chosen the alternate approach, say one could regard the molar-
average velocity as the velocity of the mixture as a whole:

(7.3.10)

in which case the molar diffusive fluxes would add up to zero but the mass
diffusive fluxes would not.
For this case it is useful to consider a binary mixture of ideal gases in a long
cylinder with sealed ends, with pressure constant along z. Since the molar density
of an ideal gas at constant temperature and pressure is independent of composi-
tion, the net molar flux is zero in such a system, say now v' = 0; the two molar
fluxes are necessarily equal and opposite (this is called equimolar counterdiffusion
in classical mass transfer theory). However, the mass flux is not zero, since mass
would actually move in the direction in which the heavier component is diffusing.
Hence, from a mass viewpoint, the mixture would actually be moving.
The substantial time derivative of any quantity L is, for the simple geometry
considered here,

L" = aL/at + vaL/az (7.3.11)

and it represents the rate of change observed as one moves together with the
mixture as a whole. Therefore, any ambiguity with respect to what is meant by
the velocity of the mixture as a whole results in an ambiguity in what is meant
by the substantial time derivative.
The ambiguity in fact must be resolved in favor ofthe mass average velocity,
since the equations of motion for the mixture as a whole must have the same
form as they would for a single-component system. Let us consider, for example,
the case of classical fluid mechanics, where the Navier-Stokes equations are
regarded as correctly describing the flow behavior of the fluid, independently of
whether the fluid is a single-component one or a mixture. The density appearing
in the Navier-Stokes equations is undoubtedly the mass density, not the molar
density, and therefore the velocity appearing in them is the mass-average velocity. 179
It is of course still perfectly legitimate to write the mass balance of each component Dissipative
in terms of molar fluxes, say cj = -a~/ az, but the ~ must be interpreted as in Phenomena
equation (7.3.9), and not as Cj(Vj - v'), and consequently they will not add up
to zero. In the following, the analysis is carried out in terms of the mass fluxes,
and only the final results are recast also in terms of the molar fluxes, to be
interpreted in the sense of equation (7.3.9).
If one wishes to express the relevant equations in terms of mass fluxes, it is
necessary to consider partial mass properties, as distinguished from partial molar
properties. Since the whole matter of partial molar properties is simply related
to the fact that extensive properties are absolutely additive functions of mass,
partial mass properties are legitimate, and in particular all the relationships
holding for partial molar properties apply to partial mass ones as well. The
notation L" will be used for partial mass properties. Also, the density of any
extensive quantity L is to be understood as the mass density, not the molar density.
Now consider the first law as it applies to our system. In particular, suppose
that the frame of reference has been chosen so that v = 0, uj = Vj; it follows that
the partial derivatives coincide with substantial derivatives in the usual sense. It
is easiest to follow the argument by having in the back of one's mind the picture
of a liquid mixture in a long cylinder sealed at the ends, with density being
independent of composition, so that in fact the actual net mass flux is zero. With
this, substantial derivatives are simply partial derivatives with respect to time in
the frame of reference fixed to the cylinder walls. We consider a differential
volume between two unit area planes orthogonal to z separated by a distance
dz. This is of course an open system, into which energy can flow not only by
heating and by work, but also by convection of energy. Indeed, the flux of energy
in the positive z direction equals the sum of the flow of heat, the rate of work
being done, the convective flux of internal energy, and the convective flux of
kinetic energy, and therefore the first law is

[~(U + K)r = -dq/dz - d[I (Uj +!uj)Qj]/dz + ~w (7.3.11)

where q is the heat flux in the positive z direction, which may be nonzero in
spite of temperature being constant since coupling of diffusion and heat transfer
may occur.
If, as is usual in the mechanics literature, one wants to retain the fact that
in a Lagrangian formulation only conductive fluxes appear, equation (7.3.12) is
to be interpreted in the sense that, in a mixture, the conductive flow of heat is
q + I (Uj + !uJ)Qj. This may at first sight appear as very artificial, but in actual
fact it is not. Indeed, in the Maxwellian kinetic theory of gases, the heat flux is
by definition the flux of kinetic energy, say it is the equivalent of the uJ term in
equation (7.3.12).
The work term is calculated as follows. Let us consider componentj entering
the system at z, and let (PJ be its volume fraction. The pressure is p, but the
component only crosses a fraction CPj of the unit cross-sectional area, and hence
the force is PCPj, and the rate at which the component is doing work on the system
is PCPjVj. The velocity is in fact Qj/~j; see equation (7.3.5). The ratio CPj/~j is
recognized to be the volume of component j per unit mass of the same component,
180 Le., it is in fact the partial mass volume. Therefore the net rate of work is given
by
Chapler Seven

I>w = -dEL (pV'j + gh)Qj]ldz (7.3.13)

where g is the gravity acceleration and h is vertical height.


The first law is thus

[I>( U + K)r = -dgl dz - d{L [(H'j + gh + !uJ)Qj]}1 dz (7.3.14)

Had one chosen to write the equations in terms of the molar fluxes, the
density appearing on the left-hand side of equation (7.3.14) would have been the
molar density. The latter would depend on composition in a constant-density
liquid; conversely, in a mixture of ideal gases the molar density is independent
of composition but the mass density is not.
We now turn our attention to the second law. There are again two mechanisms
for the flow of entropy j, namely,

(7.3.15)

and, in the absence of radiant heat, the second law can be written as

(<<I>S)" 2!: -d(qIT)ldz - deL S'j Qj)1 dz (7.3.16)

Since temperature is constant in space, q can be eliminated between the first


and second law to obtain

[I>(A + K)r + dEL (cp'j Qj)]1 dz !S 0 (7.3.17)

where

cp'j = O'j + gh + !uJ = JLjl ~ + gh + !uJ (7.3.18)

The second term in equation (7.3.17) can be expanded as

L (cp'j dQjl dz + Qj dcp'j I dz) = L (-cp'jI>j + Qj dcp'j I dz) (7.3.19)

where the mass balance of component j has been used.


The significance of quantities cp'j will be discussed later. Attention is now
restricted to the steady-state case (physically, this would require appropriate
constant rate supplies of the components at the ends of the cylinder). All time
derivatives become zero, and equation (7.3.17) reduces to

(7.3.20)
The usual assumption of smoothness near equilibrium now implies that at equi- 181
librium (i.e., when the diffusive fluxes are zero) the 4>; are uniform in space.
Dissipative
When the diffusive are also zero, the kinetic energies of diffusion are also zero, Phenomena
and hence one recovers the conclusion that the chemical potential has the
hydrostatic distribution at equilibrium. The assumption of smoothness near
equilibrium can also be used to yield a linear constitutive equation for the fluxes:

(7.3.21)

where ~k may be called the mobilities.


Correspondingly, the molar fluxes are

(7.3.22)

where

4>j = JLj + ~gh +!uJ~ (7.3.23)

The quantity JLj + ~gh is seen to be the difference between the actual
chemical potential and the value it would have under hydrostatic conditions. The
simplest way of dealing with the unusual term ~gh is to simply redefine the
chemical potential as JLj + ~gh, so that gravitational effects are automatically
taken into account (but see the Appendix in this regard). It is perhaps surprising
that, even with this redefinition, the chemical potential is not the potential for
diffusion, because of the uJ term appearing in equation (7.3.23).
The kinetic energies of diffusion are almost always very small, and if they
are neglected the chemical potential is indeed the potential for diffusion. It is
seen that equation (7.3.20) does not require every diffusive flux to be in the
direction of decreasing chemical potential, i.e., coupling between diffusive fluxes
is possible, and they are indeed observed experimentally. In the case of diffusion,
coupling must always be taken into account, since it is not possible to have only
one chemical potential nonuniform in space.
Let N be the molar flux vector (i.e., the vector whose components are the
fluxes of the N components of the mixture in the z direction) and L be the matrix
whose components are Ljk Equation (7.3.22) can be written in the form

N = Ld,../dz (7.3.24)

and we return to the question of the symmetry of L, which again is asserted as


universally true in TIP. The same symmetry was laid down as a hypothesis by
Stefan, analogously to the Stokes hypothesis for heat transfer. The mobilities of
individual components are related to the drag that each component exerts on the
others. It can be shown that, if the drag exerted by component j on component
k is independent of the concentration of the other components, the matrix L is
indeed symmetrical. This condition is satisfied trivially for binary mixtures, but
it need not necessarily be satisfied in mixtures with more than two components.
Indeed, there is some experimental evidence that in some systems the L matrix
is not symmetrical.
182 At any given temperature and pressure, both L and I' depend on composition,
say on the concentration vector c. One can therefore rewrite equation (7.3.24) in
Chapter Seven
the form of Fick's law:
N = Ddc/dz (7.3.25)

where the diffusivity matrix D is given by


D = Ldl'/dc (7.3.26)

In the proper three-dimensional formulation, N is to be interpreted as a


3 x N matrix whose rows are the components of the molar flux vector of each
component, and d / dz should be interpreted as the gradient. The advantage of
writing Fick's law in the form of equation (7.3.25) is that the diffusivity tensor
is more nearly independent of concentration than the mobility tensor. (See Section
7.8 for a detailed discussion of the heat transfer analog of this point.)
Turning our attention to the kinetic energies of diffusion, while it is true that
they are in general quite small, their neglect has some important consequences.
If they are not neglected, substitution of the constitutive equation for the molar
flux into the mass balances of the individual components yields hyperbolic
differential equations; if they are neglected, the equations are parabolic. Hyper-
bolic equations admit discontinuous solutions, and hence any discontinuity
imposed on a boundary propagates at some finite speed. On the other hand,
parabolic equations do not admit discontinuous solutions, and hence if a discon-
tinuity is imposed on a boundary it propagates at infinite speed. In the classical
theory of mass transfer between fluid phases, the transfer in the more viscous
phase is modeled according to the so-called penetration theory, where a discon-
tinuity is imposed on the interface at time zero. Since the kinetic energies of
diffusion are neglected, the discontinuity propagates at infinite speed: indeed,
the solution to the penetration theory equations is in terms of the error function,
so that if a surface element has a zero concentration at time zero, when its
interface is brought to some finite concentration, at any time larger than zero the
concentration is nonzero everywhere, no matter how far from the interface.

Appendix

The question of whether one should regard I-' or I-' + Mgh as the chemical
potential is not simply a matter of taste. If one wants to regard the chemical
potential as a function of state, then it cannot include the term Mgh, since the
latter is defined only in reference to an externally imposed gravitational field
which is not a legitimate state variable. Let us consider one gram of air at 20C
contained in a one-liter box; its chemical potential is expected to be the same
whether the box is on the Earth or on the Moon-thus one should not, in principle,
callI-' + Mgh the chemical potential. In the Maxwellian theory of gases, one can
formally add a gravitational term to the Hamiltonian, and the resulting chemical
potential will in fact be uniform in space at equilibrium, since itreally is I-' + Mgh.
However, this is only formal; in substance, the Hamiltonian should only depend
on properties of the collection of particles considered, and therefore it should
not contain a gravitational term.
7.4. MOMENTUM TRANSFER 183
Dissipative
... when different strata of a gas slide upon one another with
Phenomena
different velocities, they act upon one another with a tangential
force tending to prevent this sliding, and similar in its results to
the friction between two solid sUrfaces sliding over each other
in the same way. The explanation of gaseous friction, according
to our hypothesis, is, that particles having the mean velocity of
translation belonging to one layer of gas, pass out into another
layer having a different velocity of translation, and by striking
against the particles of the second layer, exert upon it a
tangential force which constitutes the internal friction of the gas.
J. C. Maxwell

Momentum is a vector, and its flux is therefore a tensor; the geometrical


simplifications considered so far (i.e., that the specific volume is the only relevant
kinematic variable) cannot be carried over to the case of momentum transport.
Indeed, in the fluid mechanics of liquids, it is usually assumed that the liquid is
incompressible, so that one entirely loses the possibility of doing net work on
the liquid considered by changing its density.
In order to discuss some fundamental aspects of the thermodynamics of
momentum transport, attention is restricted to an extremely simple geometrical
system, namely, a fluid contained within two parallel plates, one of which is held
fixed while the other is allowed to move in the z direction (see Figure 7.4.1). It
should be noted that in this configuration net work can be done on the fluid by
the action of a tangential force in the z direction, and that the corresponding
deformation is not a change in density but a shearing of the fluid (the density
does not change, but the shape of the fluid sample changes in time). Net rate of
work can be done if indeed the fluid can sustain a tangential stress which, by
definition of a fluid, it cannot do at equilibrium. However, when the relative
velocity of the two plates is not zero, the fluid is not at equilibrium, and it may
in fact sustain a tangential stress.
Let r be the distance between the two plates, and consider the case where
the velocity v of the upper plate is held constant in time. The tangential force
per unit surface, F, in the z direction must, at steady state, be the same at all
values of x, since the fluid is not being accelerated and therefore the sum of all
forces acting on any slice parallel to the plates must be zero. The velocity of the
fluid is presumably a function of x, since it has the value v at x = r and is zero
at the lower plate. It is the existence of such a velocity gradient which shows
that the system considered is indeed not at equilibrium; the tangential stress is
known to be zero at equilibrium, and hence it is reasonable to assume that its
value depends on the velocity gradient itself. It follows that, since F is constant
along x, so must the velocity gradient, and therefore the velocity profile is linear.

FIGURE 7.4.1. Shear flow between two parallel


plates.
184 The velocity gradient is simply vI r, and the rate at which work is being done on
the system as a whole is Fv.
Chapter Seven
A constitutive equation for the tangential force can be written as

F = f(vlr) (7A.1)

with f(O) = o. This implies that vi r is to be regarded as a state variable, say the
state is

CT, = V;, T, vir (7.4.2)

(The whole system is a square prism of unit cross-sectional area and height r.)
By reasoning analogous to that in Section 7.1 leading to the conclusion that
free energy cannot depend on the temperature gradient, it is now concluded
that for the case at hand it cannot depend on vI r. The free energy is still a
potential for entropy, and since pressure is constant in space the second law
reduces to

[aAJav; + p]V;:5 Fv (7.4.3)

Since the total volume of the system is constant in time, one concludes that Fv
must be positive, say that the force F must point in the same direction as the
velocity v. If it is assumed that the f( . ) function in equation (7.4.1) is linear, the
proportionality constant is the viscosity of the fluid, and the second law requires
the latter to be nonnegative.
The discussion in the last paragraph was somewhat simplified, but in the
proper three-dimensional formulation the conclusions which can be rigorously
deduced are indeed those stated. The constitutive equation, which is the three-
dimensional analog of equation (7.4.1) in its linear form, is equation (2.3.A.l);
for an incompressible fluid, tr D is zero identically, and so the second term on
the right-hand side drops out. We note that an incompressible viscous fluid cannot
accumulate elastic energy, and hence all the net work done by contact forces is
dissipated. This is not necessarily so for all fluids, and viscoelastic fluids which
are capable of accumulating some elastic energy are known to exist. In this regard,
see Chapter S.
The linear relationship between stress T and the symmetric part of the velocity
gradient D implies only a scalar parameter, the viscosity. This is due to the fact
that requirements of invariance require the mapping of D into T to be isotropic.
Since every isotropic matrix is necessarily symmetrical (though of course not vice
versa), the question of the symmetry of the matrix appearing in the linearized
constitutive equations for momentum flux is disposed of by a requirement of
invariance.
It is important to realize that the requirement that Fv ~ 0 is only imposed
at steady state. Indeed, even in the Maxwellian theory of gases, equation (S.1.A.8)
applies, and hence if the gas is being sheared steadily in one direction, and at
time t the shearing direction is suddenly reversed, the tangential stress will, for
very short times, point in a direction opposite to that of shearing, say Fv < O.
7.5. UNSTEADY TRANSPORT 185
Dissipative
We are all agreed that your theory is crazy. The question which Phenomena
divides us is whether it is crazy enough to have a chance of
being correct. My own feeling is that it is not crazy enough.
N. Bohr, commenting on a paper presented by W. Pauli

It is usual in the theory of transport phenomena to assume that the constitutive


equation for the flux established under steady-state conditions also applies to
unsteady-state conditions. This, however, is not guaranteed to necessarily hold
true. The discussion in this section is limited to one very simple example, namely,
heat conduction through a stationary body. However, the considerations discussed
below could be applied to a variety of other transport phenomena.
Let us consider the local form of the second law as given by equation (1.6.4),
which is

(7.5.1)

For steady-s~e conditions, this implies that the heat flux vector forms an obtuse
angle with the temperature gradient; the simplest constitutive equation satisfying
this constraint is the isotropic form of Fourier's law with a positive heat conduc-
tivity Ie. Now consider unsteady-state conditions. If the body examined is also
undergoing deformation, one would have two parallel mechanisms ofirreversibil-
ity, and this will be discussed in the sections on coupling. Here we treat the case
where the body is stationary (or in rigid-body motion), so that w = 0, and the
first law reduces to

U'=q (7.5.2)

Should equation (2.7.13) apply, the first two terms in equation (7.5.1) would
cancel out; however, equation (2.7.13) does not necessarily apply when the system
is not at steady state, and hence the heat flux and the temperature gradient vectors
do not necessarily form an obtuse angle. (See the Appendix in this regard.)
Even if the isotropic form of Fourier's law has been established experi-
mentally for steady-state conditions, it need not hold also under unsteady-state
conditions. Indeed, consider the following constitutive equation for the heat flux,
analogous to the Maxwell equation for the stress discussed in Section 5.1:

q+ 8iJq/iJt = -kgrad T (7.5.3)

with 8, the relaxation time for heat flux, a positive constant. This equation is
guaranteed to deliver a heat flux vector which, in steady state, forms an obtuse
angle with the temperature gradient vector, and its validity (or lack thereof)
cannot be ascertained by steady-state experiments.
At unsteady state, equation (7.5.3) does allow the heat flux vector to form
an acute angle with grad T, as the following simple example shows. Suppose
grad T has been held constant at some value, and correspondingly the heat flux
forms an obtuse angle with it. At some time t = 0, the temperature gradient is
suddenly reversed. The heat flux will also reverse in a time scale of order 8, but
186 at t larger than 0 by an amount negligibly small as compared to 8 it will still
Chapter Seven
have the direction it had at negative times-and hence it will form an acute angle
with grad T.
This, however, does not contradict the second law, since equation (7.1.3)
requires the heat flux to form an obtuse angle with grad T only at steady state.
If a finite relaxation time 8 is allowed for, the usual Maxwell relations do not
hold at unsteady state, and hence the othenerms appearing in equation (7.5.1)
may well compensate for a positive value of the last term on the left-hand side.
Indeed, some experimental results on the rate of crystallization in polymers
suggest that a constitutive equation for the heat flux of the type of equation
(7.5.3) is needed in order to model the data. From a microscopic viewpoint, one
could simply consider the possibility that, as a temperature gradient is suddenly
applied to the system, the morphology tends to undego some change with some
intrinsic kinetics dominated by a time scale of order 8; thus one would have a
relaxational system for the heat conduction phenomenon. Let us consider the
classical Maxwellian gas with u the molecular velocity with respect to the mixture
as a whole and (I) the average of any quantity lover the distribution function.
The heat flux vector is given by
(7.5.4)

while the stress tensor is


T = <I>(uu) (7.5.5)

where uu is the dyadic product. Since stress in a Maxwellian gas does exhibit
the phenomenon of relaxation [see equation (5.1.17)], so does the heat flux.
Indeed, the constitutive equation for the heat flux obtained from the theory of
the Maxwellian gas is significantly more complex than equation (7.5.3).
For the Maxwellian gas, the relaxation time for heat flux is ~ of that for
momentum flux, and it can be shown to be the largest possible of all relaxation
time scales. All such relaxation times are negligibly small for gases at ambient
conditions.

Appendix

If the density is constant, and one assumes the system to be elastic, then
any function of state depends only on T [say U = u(T), S = s(T), etc.]. It
follows that, since no work is done,
q/T = U/T = cvT/T (7.5.A.I)

and
S" = (ds/dT)T = q/T (7.5.A.2)

since Tds/dT= du/dT.


However, since there is relaxation involved, the equation for entropy would
be, e.g., something like
S+ 8S = s(T) (7.S.A.3)
and analogously for V. While it is still true that T dsl dT = dV I dT, the first two 187
terms in equation (7.5.1) would not cancel out, even if the relaxation time is the Dissipative
same for all functions of state. An analysis similar to that presented in Section Phenomena
5.1 would yield

s = too exp(-a)(cv/T)(dTlda) da (7.S.A.4)

and

V'IT = (liT) too exp(-a)cv(dTlda) da (7.S.A.S)

where a is (t - t')/6, t being the instant of observation and t' any previous
instant. Since T depends on a, the two expressions above do not cancel out.

7.6. COUPLING

We like to regard as useless what we do not know; it is a kind


of revenge; and since mathematics and physics are rather
generally unknown, they rather generally pass for useless.
B. Fontenelle

When more than one dissipative phenomenon is taking place there is always
a possibility of coupling, and the second law requires the total dissipation rate
to be nonnegative without restricting the individual ones. It is useful to distinguish
between two types of coupling, indirect and direct coupling; these are discussed
concisely in the following.
As an example of indirect coupling, consider the case of simultaneous heat
and momentum transfer. The usual way of dealing with such problems is to
assume that the constitutive equations which have been separately established
for the two transport phenomena when each is the only one taking place hold
also under conditions where the two take place simultaneously. This assumption
consists essentially in assuming that there is no direct coupling and, as will be
discussed later, it is an assumption which is not necessarily good, and is certainly
not dictated by the second law of thermodynamics. Nonetheless, even with such
an assumption, the two phenomena may still be coupled in an indirect way.
Indeed, some examples have already been considered explicitly: if temperature
is nonuniform on the interface between two fluid phases, a surface tension gradient
will induce thermo capillary motion in at least one of the two phases. Heat transfer
by free convection is another example of a flow induced by temperature differen-
ces, i.e., of an indirect coupling between heat and momentum transfer.
It is useful in this regard to consider the operation of a refrigeration cycle.
It is obvious that, on a macroscopic scale, heat flows from a cold source to a
hotter sink in a refrigerator: thus the idea that the second law requires heat to
invariably flow from hot to cold is belied by the fact that refrigerators do work.
In fact a refrigeration cycle can be analyzed on the basis of extremely simple
188 constitutive assumptions for the working fluid. In the discussion that follows,
reference is made to the classical Camot cycle analysis of a refrigeration cycle;
Chapter Seven
see Figure 7.6.1. The cycle starts at, say, point 1, where. the fluid is compressed
to some higher pressure. The fluid is regarded as bei~g an elastic gas, since
compression takes place sufficiently slowly that viscous effects can be neglected
(it should be borne in mind that viscous effects in compression and expansion
are negligible unless the rate of change of density is extremely high). Thus the
compression is reversible, and if it can regarded as adiabatic as well (which is
reasonable for sufficiently large units: as the size of the unit increases, its
surface-to-volume ratio decreases, and thus it approaches adiabatic behavior)
the compression will be isoentropic, thus resulting in the vertical segment 12 (in
Figure 7.6.1) of the Camot cycle. We note that here the characteristic of the
compression step (i.e., its being isoentropic) is deduced from the second law, by
making use of the independently arrived at conclusion that the step takes place
reversibly. The horizontal segment 23 corresponds to a change of phase taking
place at temperature T1 , again regarded as reversible. Here the fluid discharges
heat to the warm source, which in actual fact will be somewhat colder than TI
but which can in principle be at a temperature as close to TI as one wishes.
The segment 34 presents some subtlety. It corresponds to an expansion which
again can be regarded as adiabatic. It would therefore appear that one can again
regard it as isoentropic, but in fact it is generally assumed to be isoenthalpic if
it takes place through a nozzle valve. This is connected to the fact that expansion
through a valve is so sudden that viscous effects cannot be neglected, and therefore
the expansion is not reversible: indeed, the fluid does no work at all on the
surrounding if it expands through a nozzle valve. In this case, the characteristic
of the step (its being isoenthalpic) is deduced from the first law, by making use
of the fact that the fluid does no work on the surrounding. It is, however, possible
to effect the expansion not through a valve but through some kind of device such
that the fluid does do work on the surrounding, and expands slowly enough to

I Heat discharge
3
ExpanSionD
device
4 1
f Heat pickup

T1 ------02
T2 ------4
3

1 FIGURE 7.6.1. A classical refrigeration cycle


(top) and the corresponding Carnot plot
5 (bottom).
again make viscous effects negligible. In this case, the expansion can again be 189
regarded as reversible and hence isoentropic, resulting in the vertical leg 34 of Dissipative
the Carnot cycle. The fluid does work on the surrounding in the device, but of Phenomena
course it may well not be feasible to make use of such work; this irreversibility
would however be related to the outside mechanism, not to the working fluid
itself. Finally, the horizontal leg 41 is again a change of phase at temperature
T2 , regarded as reversible, during which the fluid subtracts heat from the cold
source. The latter would need to be at a temperature somewhat higher than T2 ,
but in principle it could work at a temperature as close to T2 as one wishes.
It is now of interest to consider the overall dissipation connected with the
operation of a refrigeration cycle. There are, of course, external sources of
irreversibility (not all the electrical power used to drive the compressor is trans-
mitted to the fluid; the hot sink is somewhat cooler than T 1 , and the cold source
somewhat hotter than T2 ; the work done by the fluid in expansion is not totally
recovered, if at all), but as far as the analysis given above is concerned there are
no irreversible processes taking place in the working fluid itself. The external
sources of irreversibility can in principle be made as small as one wishes, and
thus the actual refrigeration cycle could approach as closely as required the highest
possible efficiently established in Section 1.8. That highest possible efficiency
corresponds to the case where the positive dissipation connected with the transfor-
mation of the difference between the mechanical work done on the fluid in the
compressor and that done by th~ fluid in the expansion device into heat discharged
at TI exactly balances the negative dissipation of the transfer of heat from the
cold source to the hot sink: the total dissipation is zero, through a coupling of
two processes effected in as efficient a way as allowed by the second law.
If such coupling is possible on the macroscopic level of a refrigeration cycle,
one concludes that it may also be possible on the local scale of constitutive
equations-the second law is supposed to apply on both scales. A simple special
case is treated below. If viscous effects (which account for the possibility of
nonthermal dissipation) and heat transfer must be taken into account simul-
taneously, the simplest possible structure for the state is

u = v. T, V-, grad T (7.6.1)

and, by an argument strictly analogous to that developed in Sections 2.4 (as far
as V' is concerned) and 7.1 (as far as grad T is concerned), we can conclude
that free energy depends in fact only on V and T. It can also be concluded that
free energy is a potential for entropy, so that also the latter depends only on V
and T. However, free energy is not a potential for the heat flux, and it is a
potential only for the equilibrium pressure p* (corresponding to both V and
grad T being zero) but not for the actual pressure p. The second law reduces to

[p - p*]V + qgrad T/(iPT) $ 0 (7.6.2)

(where V is the mass specific volume). Should grad T be zero, p - p* would


need to have its sign opposite to that of V. Since however p may in principle
depend on grad T, this need not to be true in general, and hence the mechanical
190 dissipation rate [p* - p] V could well be negative, provided the thermal dissipa-
tion rate is positive and its absolute value not less than that of the mechanical
Chapter Seven
dissipation rate. Conversely, the heat flux must form an obtuse angle with grad T
if V is zero, but since in principle q could depend on V this need not be true
in general-hence heat could flow in the direction of increasing temperatures if
there is a concomitant sufficiently large positive mechanical dissipation.
With regard to the above discussion, some of the TIP literature would negate
the possibility of such a coupling based on the so-called Curie theorem, since q
is a vector and V' is a scalar. However, there is in fact no such thing as a Curie
theorem in algebra: scalar quantities may well depend on vectors, and vice versa,
though there are some restrictions imposed on such dependencies if one assumes
linearity. We do not claim here that in systems so simple as to be describable by
equation (7.6.1) local coupling of viscous dissipation and heat transfer is likely
to actually be of importance; the point is simply that such coupling is not forbidden
by the second law.
It is of interest to observe that, in the theory of relaxational systems discussed
in Chapter 5, free energy is a potential for pressure, and therefore within the
scope of such a theory one would conclude that pressure cannot depend on
grad T. It would still, however, be possible for the heat flux not to form an obtuse
angle with grad T in the presence of a sufficiently large positive mechanical
dissipation rate.
A case of coupling at the level of constitutive equations has already been
discussed in Section 7.3, where the diffusive flux of each component was shown
to depend in principle on the chemical potential gradients of all components. If
diffusion takes place simultaneously with heat transfer, an additional coupling
arises in that the diffusive fluxes depend also on the temperature gradient (this
is called the Soret effect) and, vice versa, the heat flux may depend on the chemical
potential gradients (this is called the Dufour effect). Both effects have been
observed experimentally. Again, TIP asserts that, in the linear description of
these two effects, the cross coefficients should be equal to each other, but this
can be shown to be necessarily true only under some additional simplying
assumptions which are independent of that of linearity.
Chemical reactions may also in principle be coupled to each other, so that
in a system where more than one chemical reaction takes place the product of
the affinity and the reaction rate need not separately be nonpositive for each
component reaction; see equation (3.2.2). However, the literature on chemical
reactions is generally based on the assumption that there is no such coupling.
This assumption may not be justified in the case of chemical reactions taking
place in biological systems. Chemical reactions may also be coupled with heat
and mass transfer phenomena on the constitutive equations level; they certainly
are on the macroscopic level, and in the theory of mass transfer with chemical
reactions based on the assumption of no coupling at the constitutive equation
level cases arise where the diffusive flux of some component takes place in the
direction, of increasing chemical potential. The ability of the kidney to concentrate
impurities is believed to be connected with such coupling phenomena.
The possibility of coupling of chemical reactions offers another interesting
example of the question of symmetry of the relevant matrix of coefficients in the
linear form of constitutive equations. Let us consider the following very simple
example: two reactions may occur, and the system is sufficiently close to equilib- 191
rium that the kinetics can be written as linear in the affinities:
Dissipative
Phenomena
and (7.6.3)

It is seldom useful to write chemical kinetics in this way, but this is irrelevant
to the present discussion-a neighborhood of the eqUilibrium point exists where
equations (7.6.3) hold if indeed both affinities are zero at equilibrium, as was
discussed in Chapter 3. Now the second law requires that

(7.6.4)

which is satisfied provided that

kll 2:: 0, (7.6.5)

These conditions certainly do not require k12 to be equal to ~1' and there is no
reason whatsoever to think it should be. (The argument could easily be generalized
by introducing a kinetic constant tensor with components klJ , and it would then
be exactly analogous to that developed in Section 7.1, with no crystal symmetry
available to be invoked for providing the appropriate symmetry.)
We now note that, if one makes the usual assumption in chemistry, i.e., that
the two terms on the left in equation (7.6.4) are separately nonpositive, one would
conclude that both k12 and k21 are zero, i.e., that the kinetic constant matrix is
diagonal. Since every diagonal matrix is symmetric (though the converse is not
true), one would obtain a stronger result than the symmetry asserted by TIP.
Thus two possibilities remain: either the system considered is simple enough for
the usual assumption in chemistry to hold, and then the matrix is not only
symmetric, but also diagonal; or the system is more complex than that, in which
case there is no reason to suppose it is symmetrical, but only that it is positive
definite.

7.7. THE SYMMETRY RELATIONS

A combination of impressions regarding the nature of heat with


imperfectly developed apprehension of the new theory has
somewhat liberally perplexed the modem student of
thermodynamics with questions unanswerable by theory or
experiment, and propositions which escape the merit of being
false by having no assignable meaning.
Lord Kelvin

In the preceding sections, some concrete examples of linear constitutive


equations for fluxes have been considered, and the question of the symmetry of
the matrix of coefficients for such examples has been discussed. In patticular
heat flux, momentum flux, diffusive fluxes, and chemical reactions have. been
considered. The TIP literature usually takes a seemingly more general approach,
which is, however, open to very serious questions. This is discussed concisely in
the following.
192 The TIP literature considers the general case where there are, say, M "fluxes"
Chapter Seven
which constitute an M -dimensional vector J. Correspondingly, there are M
"forces" or "affinities" (which are perhaps to be understood in the sense that
driving forces are understood in chemical engineering), which constitute another
M -dimensional vector X. If the constitutive equations are linearized, one obtains

J=LX (7.7.1)

where L is an M x M matrix of "phenomenological coefficients." The assertion


of TIP is that L is symmetric, an assertion offered as universally true.
Though this formulation appears as very general, the 'problem which arises
is that one needs to know what the components of J and of X might be. The TIP
literature gives two alternate definitions in this regard, the first one being as
follows. If Z is the local total rate of dissipation, the local rate of "entropy
production" is Z / T. Vectors J and X are identified by the fact that the rate of
entropy production is given by

Z/T=JX (7.7.2)

This is, however, not an acceptable definition, since infinitely many different
couples of vectors may well have the same scalar product. Indeed, let us consider
a new set of fluxes, J', given by

J' = (1 + WL-1).J = (L+ W)X (7.7.3)

where W is any nonzero skew matrix. If L is symmetric, the vector J' is not related
to X by a symmetric matrix, since W is nonzero. Yet J' could well be taken as
the set of fluxes, since it satisfies equation (7.7.2). In fact, by an elementary
theorem of matrix algebra

(WX)X = 0 (7.7.4)

and therefore

Z/T = J . X = J' . X (7.7.5)

Conversely, suppose sets J' and X are given for which L is not symmetrical.
Then L can be decomposed into its symmetrical part D and skew part W, and a
new set J can be defined as

J = (1- W'L-1)'J' = DX (7.7.6)

Vector J again satisfies equation (7.7.2), and is related to X by a symmetric


matrix. It follows that, if equation (7.7.2) is accepted as the definition, the asserted
symmetry of L reduces to the following statement: "There exists a linear combina-
tion of the components of J which is related to X by a symmetric matrix." This
is a trivial result in matrix algebra, and not a thermodynamic statement.
The alternative definition of J and X to be found in the TIP literature is as 193
follows. A set of "thermodynamic coordinates" a is considered, such that the Dissipative
entropy production can be written in the form Phenomena

Z/ T = !(G a) a (7.7.7)

where G is a (constant) symmetric matrix. The sets J and X are now defined as

J= a' (7.7.8)

and
X=G'a (7.7.9)

Equation (7.7.2) does follow from equations (7.7.7)-(7.7.9), though the


converse is not true. Moreover, any linear transformation of J and X which does
not violate equations (7.7.7)-(7.7.9) preserves the symmetry of L. Therefore, to
within our ability to identify what the set a may be, the statement that L is
symmetric is not simply a matrix algebra result, but a thermodynamic statement
which could be disproved or confirmed by experiment.
However, identification of the set a is open to serious problems. For instance,
the heat flux is certainly a legitimate flux to consider, and yet what is the
thermodynamic quantity whose rate of change is the heat flux, as equation (7.7.8)
would require? In fact equation (7.7.8) does not give any clue, because given
any J one could find the thermodynamic coordinates a by integration:

a = f J dt (7.7.10)

which automatically satisfies equation (7.7.8).


Equations (7.7.7) and (7.7.8) appear to be suggested by the consideration
of chemical reactions: a would thus be identified with the extent of reaction
vector, J with the rate of reaction vector, and X, perhaps, with the affinity vector.
There does not seem to be any physically significant way in which equations
(7.7.8) and (7.7.9) can be applied to phenomena involving gradients, such as heat
transfer, diffusion, electrical current, and the like.
It follows from the above arguments that the question of the symmetry of
L can only be discussed if in fact the sets of fluxes and driving forces to be
considered have been identified concretely, as was done in the preceding sections.

7.8. THE CURIE AND MINIMUM ENTROPY PRODUCTION


PRINCIPLES

Everybody believes it completely because mathematicians think


it is an experimental fact, and experimentalists think it is a
theorem in mathematics.
H. Poincare
194 Another difficulty with the TIP literature, which has been hinted at in the
Chapter Seven
previous sections, is the common reference to the so-called Curie theorem,
sometimes called the Curie principle (which introduces a first difficulty, since
the words theorem and principle are not synonyms). Two different formulations
of the theorem can be found in the literature, namely:
1. Strong form. "Fluxes and thermodynamic forces of different tensorial
character do not interfere."
2. Weak form. "Forces and fluxes that correspond to a coupling of tensors
whose orders differ by an odd number do not occur."

Trivial counterexamples to both statements can be found. For example,


consider the isotropic form of Fourier's law, with a conductivity which depends
linearly on temperature:
q = ko(1 + BT) grad T (7.8.1)

where a vector flux (order-one tensor) is coupled with the scalar (order-zero
tensor) quantity, temperature T.
It is generally a frustrating exercise to try to understand what statements in
the TIP literature really mean. However, equation (7.7.8) must have been formu-
lated by somebody who was thinking of chemical reactions; the flux would then
indeed be the rate of reaction, the thermodynamic coordinate the extent of
reaction, and the force the affinity. Of course, chemical kinetics are rarely linear,
but we shall forget about that for the time being. Chemical reactions do occur
often (if not invariably) in the presence of heat transfer, and so the question of
coupling between the two arises. At a given temperature gradient, mayor may
not the heat flux in a reacting mixture depend on the extent of reaction? This
seems to be the question addressed by TIP, and the answer, with the Curie
principle in either form, is no-the heat flux is a vector, and the extent of reaction
is a scalar. This conclusion is clearly nonsense, since the thermal conductivity of
a mixture may very well depend on composition, and thus on the extent of reaction.
An even more dramatic example is that of a polymerization reaction. As the
degree of polymerization increases, the viscosity increases by several orders of
magnitude and therefore, at an assigned value of the shear rate tensor, the stress
tensor changes dramatically. This would be forbidden by the Curie principle in
its strong form-though, admittedly, not in its weak form, since stress is a tensor
and the extent of reaction is a scalar, i.e., the orders differ by two.
Returning to the example of coupling between heat flux and extent of
reaction, one can convince oneself that the Curie principle is not only wrong, it
is meaningless. Let us consider a mixture in which three chemical reactions take
place simultaneously. Now the extent of reaction is a three-dimensional vector,
and according to the Curie theorem in such a case the heat flux could depend
on it. Does it make any sense to think that the heat flux can depend on the extent
of reaction vector only when there are exactly three independent reactions going
on? One could argue oneself out of this paradox by saying that by "vectors" one
means three-dimensional vectors in physical space, and not ordered triplets of
numbers, and that therefore the extent of reaction is not a vector. However, this
leads to another paradox. In fact, either the Curie principle is a theorem in
algebra, or it is a principle of physics. In algebra, the distinction between ordered 195
triplets of numbers and vectors in physical space disappears. So the Curie principle Dissipative
needs to be a principle in physics. From which fundamental laws of physics is Phenomena
it deduced?
Finally, it is of interest to observe that, if one undertakes the painstaking
exercise of following the Curie principles through the chain ofliterature references
in the TIP literature back to its origin, one does reach a work by Curie in which
nothing even remotely similar to the Curie principle is stated.
We now tum our attention to another difficult point in the literature on TIP,
which is the question of a vaguely formulated principle of minimum entropy
production. Again an unambiguous statement of the principle is hard to pin
down, but the following formulation is perhaps a reasonable synthesis of what
can be found in the literature:
"Given a dissipative phenomenon which takes place under conditions such
that all boundary conditions imposed on the system are steady-state ones, the
spatial distribution of variables is such as to result in the minimum global rate
of entropy production (or energy dissipation) compatible with the imposed
boundary conditions."
A trivial counterexample is easily found. Flow of a Newtonian fluid through
a circular pipe takes place, at Reynolds numbers exceeding 2100, in turbulent
flow. It could, in principle, equally well take place under laminar flow conditions,
and the latter would result in a lower pressure drop and hence a lower rate of
energy dissipation. Conversely, at Reynolds numbers less than 2100 a turbulent
flow condition would result in a lower pressure drop than laminar flow, and yet
the latter is the one which is actually encountered. Thus it is concluded that in
fact the system always chooses the flow regime where the rate of energy dissipation
is highest, rather than the flow regime where it is smallest.
This difficuty could be overcome by restricting the formulation of the prin-
ciple to situations where not only the boundary conditions are steady state, but
the actual distribution of variables is also steady state. This of course excludes
from consideration turbulent flow. One is thus left, for the case of flow of
Newtonian fluids, with only steady-state velocity distributions, and yet counter-
examples could still be easily constructed for high Reynolds number flows. One
now could restrict the scope of the principle even more, and require it to apply
only in the zero Reynolds number limit, i.e., for creeping flow.
The classical variational principle for creeping flow of Newtonian fluids,
known as the Helmholtz-Korteweg theorem, indeed produces the result that a
steady-state creeping flow solution is one which minimizes the volume integral
of a quantity proportional to the local rate of energy dissipation, and hence for
this example, and in this restricted form, the principle does indeed hold. However,
the Helmholtz-Korteweg theorem is a purely mechanical result, which is obtained
with no appeal to any thermodynamic principle whatsoever. The proportionality
of the quantity to be minimized with the rate of energy dissipation is related to
a simple dimensional argument, and is thus fortuitous. Indeed, a modified form
of the Helmholtz-Korteweg theorem can be constructed for non-Newtonian fluids
and-except in the dimensionally degenerate case of a power-law constitutive
equation-the quantity to be minimized is not proportional to the rate of energy
dissipation.
196 The point discussed in the previous paragraph is best illustrated by consider-
Chapter Seven
ing a simple one-dimensional problem: flow of a fluid between two parallel plates,
which both move with constant velocity and parallel to each other in the x
direction. If v is the velocity in the fluid (parallel to the plate velocity) and r the
tangential stress, then the momentum balance is
<I> av/at + ar/ax = 0 (7.8.2)

Since this equation implies that av/at and ar/ax have opposite signs, one has
0;0: (av/at)(ar/ax) = a(rav/at)/ax - ra(av/at)/ax (7.8.3)

If this is integrated over the distance between the two plates, the first term
vanishes, since av/at is zero on both plates. Hence

f r[a(av/at)jax] dx ;0: 0 (7.8.4)

Now let us assume the fluid is Newtonian, i.e., r = -11- av/ax, where 11- is
the viscosity. We define the quantity n as
(7.8.5)

so that
r[a(av/at)/ax] = -an/at (7.8.6)

Equation (7.8.4) now reduces to

(7.8.7)

and a variational principle has been established: the solution is that which
minimizes the integral of n. This of course means that the velocity gradient is
constant, which is known to be the correct solution. The He1mholtz-Korteweg
theorem is nothing else but the generalization of this result to the general
three-dimensional formulation. Since the rate of energy dissipation is proportional
to (av/ax)2, the minimum entropy production principle happens to hold true for
this case. However, if r is a nonlinear function of av/ax, a variational principle
of the Helmholtz-Korteweg type can still be established, but the quantity to be
minimized is not proportional to the rate of entropy production unless r is
proportional to a power of av/ax.
However, let us consider the analogous heat transfer problem, with q the
heat flux. The equivalent of equation (7.8.4) now becomes

f q[a(au/at)/ax] dx (7.8.8)

If it is assumed that Fourier's law applies with a constant conductivity k, and


that the internal energy is proportional to temperature, namely U = cT (which
are certainly plausible hypotheses), the quantity n is, for this case,
(7.8.9)
and the variational principle is again established, with (aT/ax)2 the quantity to 197
be minimized. This again produces the well-known result that aT/ax is constant.
Dissipative
However, the energy dissipation is Phenomena

(7.8.10)

which is not minimized at the steady-state solution.


Of course, it is always possible to fix up things by juggling the algebra
around. According to TIP, the appropriate driving force for heat flux is not grad T
but grad f3, where f3 = 1/ T. It would follow that the linear constitutive equation
for the heat flux in the one-dimensional case would be

q = Ldf3/dx (7.8.11)

and for the flat slab geometry a linear distribution of f3 would be expected. We
assume density is constant. Hence U = u(f3), with.

(7.8.12)

where c is the specific heat. Thus the balance equation becomes

(7.8.13)

i.e., aq/ax and af3/at have necessarily equal signs. It follows that

0;;:: (aq/ax)(af3/at) = a(q af3/at) - Lf3 a(af3/at)/ax (7.8.14)

and analogous algebra would lead to !l being proportional to (af3/ax)2. With


this, a variational principle of the Helmholtz-Korteweg type would have been
established, the quantity to be minimized being indeed the rate of entropy
production.
This seems to dispose of the problem, but there is an important conceptual
difficulty, i.e., does one really expect the distribution of f3 to be linear, rather
than that of T? In other words, is L in equation (7.8.11) more likely to be
independent of temperature than the thermal conductivity k as usually defined?
The question is usualy brushed aside by maintaining that, if one is close enough
to eqUilibrium to justify linearization (i.e., if the temperature gradient is small
enough), Land k are proportional to each other:

(7.8.IS)

This argument requires T to span a range small enough that the proportional-
ity factor 1/ T2 may be regarded as constant. However, one can have steady heat
conduction with as small a temperature gradient as one wishes and still have T
ranging over a wide range, provided the slab is thick enough. Thus, no matter
how close to equilibrium one may be, the question of whether L or k is more
nearly independent of temperature is legitimate, and experimental evidence is
rather in favor of k being nearly independent of temperature.
198 Be that as it may, the important point is that, by writing linear constitutive
equations in some appropriate form, and within the restrictions discussed earlier,
Chapter Seven
it is sometimes possible to work things out in such a way that a Helmholtz-
Korteweg type of variational principle can be obtained which requires the rate
of entropy production to be minimized. However, the type of constitutive
equations required may be rather unusual, and there is no experimental support
to the idea that they are appropriate.

7.9. BIOLOGICAL SYSTEMS

At the centre of biological science lies the concept of


organization, undefined but indispensable. This organization is
not only one in space, as ref/ected in the morphology of the
organisms, but also one in time, revealed by their behavior.
There is a constant challenge to the biologist to try to
understand and explain these organizational aspects of
biological systems in terms of simpler and better
defined concepts.
B. C. Goodwin

Biological systems are not "bodies," since they invariably exchange matter
with the environment. Any thermodynamic analysis of biological systems must
therefore be based on the formulation of the first and second law for open systems.
Let us consider in particular the very simple open system shown in Figure 7.9.1,
which has only one entering stream and only one exiting stream. (Real biological
systems are much more complex; however, the single entering stream may be
regarded as the sum of all entering streams, and the same applies to the exiting
stream.) The second law, in the form of equation (1.7.2), can be applied to a
differential mass entering the system at some time to and exiting at time to + t p ,
where t p is the residence time within the system. This requires integration of all
terms over the residence time. However, before performing the integration, it is
useful to recast equation (1.7.2) into the following form:

A' + ST - w - Zr ,.;; 0 (7.9.1)

Temperature in living biological systems is almost exactly constant in both


space and time, so that the second and fourth term may, as a quite reasonable
first approximation, be dropped. If the work term is expressed as -pV', equation
(7.9.1) becomes
0"- Vp''';; 0 (7.9.2)

OP EN
SYSTEM

FIGURE 7.9.1. An open system.


Finally, pressure is also very nearly uniform, and thus equation (7.9.2) 199
reduces to the requirement that the free enthalpy must be decreasing:
Dissipative
Phenomena
(7.9.3)

We first consider the case where the system may be regarded as being at
steady state. For this case, equation (7.9.3) implies that the total convective flow
of free enthalpy out of the system must be no more than the total convective
flow of free enthalpy into the system. This condition is always satisfied trivially
for biological systems, since the entering streams are typically high free. enthalpy
nutrients, while the exiting streams are low free enthalpy metabolic products.
Indeed, for a variety of fermentation processes the thermodynamic efficiency has
been measured; this is essentially the ratio of the total outflow and inflow of free
enthalpy. For those biological systems for which data are available, the thermo-
dynamic efficiency is generally of the order of 60%, and is often less than that.
The steady-state approximation is, however, open to considerable question,
for a number of reasons. Cyclic processes are known to take place in all living
systems; the same systems respond dynamically to variation in the external stimuli;
and finally, evolution is of course a crucial feature of biological systems. These
three types of phenomena, however, take place over very different time scales,
so that a hierarchy of phenomena can be considered in the temporal organization
of biological systems if the relaxation times of different phenomena can be
determined. In such a hierarchy, biophysics is at the bottom and population
genetics at the top.
The analysis of open metabolic systems is usually carried out by regarding
the concentrations of macromolecular compounds as constant (or at most slowly
changing in time) environmental parameters. This implies that the time scales
involved need to be Significantly less that those of macromolecular synthesis. The
processes considered are diffusion, interaction, and transformation by enzyme
catalysts of low molecular weight molecules, and the relevant time scale is of the
order of 10- 1_10+ 1 s. This is indeed much shorter than the relevant time scale for
the next higher system (the epigenetic one). However, a difficulty arises owing
to the very real possibility of chemical oscillations in the metabolic system. Even
if these oscillations have intrinsic time scales of the stated order of magnitude,
they are nonlinear and hence can exhibit subharmonic phenomena, which can
then interfere with the time scales involved in the epigenetic system.
Oscillations are in fact of crucial importance in biological systems, and their
analysis can be carried out on the basis of a microscopic theory which has strong
analogies with the classical statistical theory of thermodynamics. Only a brief
sketch of such a theory is given in the following. In regard to this theory, the
following points which distinguish it from the classical thermodynamic theory,
and which should always be kept in mind, are to be noted:

1. The "elementary particles" of the theory are individual cells, rather than
molecules as in the classical case. Thus the scale of description at the
microscopic level is significantly larger than in the classical case, and
statistical considerations are more loosely justified.
200 2. The miscoscopic theory is developed (and its development is still today
in a rather elementary stage) before a macroscopic formulation is avail-
Chapter Seven
able. This is in reverse order with respect to the development of classical
thermodynamics.
3. The concept of equilibrium is substituted by the concept of steady state-
and, as will be seen, in fact there is no tendency to evolve toward such
a state.

Consider the very simple feedback control loop in Figure 7.9.2, where Lis
a genetic locus which synthesizes messenger RNA at concentration, say, E. The
mRNA signals a ribosome R where an enzymatic protein is synthesized at
concentration Y. The protein influences the metabolic state at C resulting in a
metabolite M, a fraction of which loops back to L and represses the protein
synthesis. An extremely simplified model of the kinetics involved results in the
following equations for the rates of change of E and Y:

dY/dt = aE - (3 (7.9A)

dEl dt = [a/(A + eY)] - b (7.9.5)

where a, (3, a, A, e, and b are constant parameters. The system of equations


(7.9.4) and (7.9.5) admits the following trivial steady-state solution:

E = (3/a (7.9.6)

Y = (a - Ab)/(be) (7.9.7)

However, the steady-state solution can never be reached from any initial
condition other than the steady state itself. In fact, equations (7.9.4) and (7.9.5)
can be easily combined into a single equation:

(aE - (3) dEl dt + {b - [a/(A + eY)]} dY/ dt = 0 (7.9.8)

mRNA,E

Metabolite FIGURE 7.9.2. A very simplified sketch


of a biological feedback control system.
which integrates to 201
Dissipative
!aE2 - f3E + bY - (ale) In(A + eY) = G(E, Y) = constant (7.9.9)
Phenomena
which, in the E - Y plane, describes a closed loop around the steady-state solution
for any positive value of the constant on the right-hand side (the value of the
constant is obviously determined by the initial condition). This is of course a
case of cyclic chemical reactions, which are possible in metabolic systems because
of the concomitant source of irreversibility arising from the degradation of
nutrients to metabolic products.
We note that the rates of change of the two variables E and Y can be
expressed as

dEldt = -aGlaY (7.9.10)

and
dYldt = aGlaE (7.9.11)

which shows that they are in fact in the Hamiltonian form. A variational principle
analogous to Hamilton's principle in mechanics can be established, the variational
quantity being

!(EdYldt - YdEldt) - G(E, Y) (7.9.12)

In actual fact, there are several different proteins and corresponding


messenger RNAs, so that one should in fact consider vectors E and Y. If one
now considers a large number of cells, and a phase space a point of which is the
ordered pair E, Y, a density <P(E, Y) can be defined in phase space, and a velocity
v which is the ordered pair dEl dt, dYI dt. With this, a continuity equation in
phase space can be written which reduces simply to the requirement that the
substantial derivative of the density is zero: in fact, the appropriate generalization
of equation (7.9.10) and (7.9.11) implies that the divergence of v is zero.
With that much background, it is now possible to introduce a canonical
ensemble in perfect analogy with the statistical theory of classical thermo-
dynamics, a talandic partition function Z*, and a talandic temperature T*. (The
word talandic, from the Greek Talanthosis, "oscillation," is used as a reminder
that the quantities involved have only a formal and rather loose analogy with
the analogous quantities arising in classical thermodynamics.) The units of the
talandic temperature are molecules per cell per second, and it is a measure of
the energy of oscillations, just as in the case of the Maxwellian gas temperature
is a measure of the kinetic energy. The talandic internal energy (G) (the ( ) sign
is used to indicate phase-space averages) tends to flow from regions of high T*
to regions of low T*, and is related to the talandic partition function Q* by

U* = (G) = -a In Q* /0(11 T*) (7.9.13)

Analogously, a talandic entropy S* can be defined as

S* = In Q* - (11 T*) aIn Q* I a(11 T*) (7.9.14)


202 and indeed at fixed U* the system will tend to evolve in the direction of
highest S*.
Chapter Seven
How useful a "talandic thermodynamics" theory may tum out to be in the
description of the macroscopic behavior of large collections of living cells is still
open to some question. Probably, biology today is at best at the stage of scientific
evolution which classical phenomena of physics had reached in the late 19th
century, when the classical statistical theory of thermodynamics was being
developed.

EXAMPLES AND PROBLEMS

Problems

7.1. Consider a solid for which the heat flux is given by equation (7.5.3). A semi-infinite body
of that solid, located at z > 0, is initially at equilibrium at a temperature To. At time zero, the plane
at z = 0 is suddenly brought to temperature T[ and thereafter maintained at that temperature.
Determine the temperature in the solid as a function of distance from the external surface, z, and time.
7.2. Obtain an estimate of the partial pressures of oxygen and carbon dioxide at the top of
Mount Everest.
7.3. As one walks up a mountain, one meets at a certain altitude the so-called tree line: no
trees grow above that altitude. Give one or more possible reasons for this phenomenon.
7.4. Estimate the significant length scale for vertical composition changes in the atmosphere
of Jupiter. You may assume that the atmosphere is made up essentially of methane and ammonia.
7.5. Describe in detail how you would set up an experiment aimed at establishing whether the
heat conductivity tensor is or is not symmetric.
7.6. Consider a system where three reactions may take place, and assume that the kinetics are
linear in the affinities. Determine the conditions imposed on the kinetic constant matrix by the second
law.
7.7. Develop a variational principle for steady one-dimensional diffusion of a dilute solute
through a solvent. Suppose that Fick's law holds with a constant diffusivity. Is the variational principle
one which minimizes entropy production? How would the constitutive equation for the flux need to
be written in order to obtain a variational principle which requires the rate of entropy production
to be minimized?

LITERATURE

When a Man observes the Choice of Ladies nowadays, in the


dispensing of their Favours; can he forbear paying some
Veneration to the Memory of those Mares mentioned by
Xenophon; who, while their Manes were on, that is, while they
were in their Beauty, would never admit the Embraces of
an Ass.
Jonathan Swift

In this chapter, most of the dissipative phenomena which have been considered belong
to the class usually identified with transport phenomena, the driving force being the
gradient of some potential such as temperature. The classical textbook for these is R. B.
Bird, W. E. Stewart, and E. N. Lightfoot, Transport Phenomena, Wiley, New York (1960).
However, it should be borne in mind that all phenomena of evolution of internal state 203
variables (such as, typically, chemical reactions) are also dissipative, although no gradients
of any quantity are required for such processes. The TIP literature does not clearly Dissipative
distinguish between these two types of dissipative phenomena. Phenomena
Data showing that in multi component mixtures the mobility matrix is not invariably
symmetrical are presented by P. J. Dunlap and L. J. Gosting, 1. Am. Chern. Soc. 77,5238
(1955).
The possibility that a relaxation time may appear in the constitutive equation for the
heat flux, and the consequences thereof, have been discussed by R. Ocone and G. Astarita,
AIChE 1. 33, 423 (1987); the idea goes back to the work of C. Cattaneo, Atti Sem. Mat.
Fis. Univ. Modena 3, 83 (1948). Relaxation in heat flux in Maxwellian gases was first
analyzed by J. C. Maxwell himself, Scientific Papers, Cambridge University Press (1890).
See in this regard also E. Ikenberry and C. Truesdell, 1. Ratl. Mech. Anal. 5, 1 (1956) and
C. Truesdell, J. RatL Mech. AnaL 5, 55 (1956).
The possibility that, on sudden reversal of the direction of shearing, a Maxwellian
gas may exhibit for a short time a tangential stress pointing in a direction opposite to that
of shearing is discussed by C. Truesdell, Rational Thermodynamics, Springer-Verlag, Berlin
(1984).
The fact that, if the kinetic energies of diffusion are not neglected, the diffusion
equations become hyperbolic and hence discontinuities propagate at finite speeds is
discussed by I. Muller and P. Villaggio, Meccanica 11, 191 (1976).
The literature on TIP is very abundant; it is generally difficult to read and overwhelm-
ingly confusing. The best presentation is given by K. Denbigh, The Thermodynamics of
the Steady State, Methuen, London (1951). This is also the only presentation to clarify
that, if at all, the theory is applicable only to steady-state phenomena.
The earliest, and possibly still today the clearest discussion of coupling is the analysis
of the thermoelectric effect presented by Lord Kelvin, Mathematical and Physical Papers,
Cambridge University Press, London (1922), Vol. 1, p. 232.
The viewpoint expressed in this chapter follows that of C. A. Truesdell, Rational
Thermodynamics, 2nd Ed., Springer-Verlag, New York (1985).
For a discussion of the difficulties related to the Onsager reciprocal relationships, see
B. D. Coleman and C. A. Truesdell, 1. Chem. Phys. 33, 28 (1960); and G. Astarita and
G. C. Sarti, Chim. Ind. (Milan) 57, 680, 749 (1975). The paper by J. Wei, I.E.C., Int. Ed.
58, 55 (1966) is also of interest.
While the symmetry relations are usually regarded as following from the so-called
principle of microscopic reversibility, some authors regard them as a postulate of the
theory; see, e.g., E. A. Desloge, Thermal Physics, Holt, Rinehart, and Wilson, New York
(1968). From this viewpoint, the question of proof does not arise; however, the question
of the validity of the postulate becomes crucial.
The question of the relationship between the Helmholtz-Korteweg variational prin-
ciple and entropy production is discussed by G. Astarita, 1. Non-Newtonian Fluid Meeh.
1, 343 (1977) and 13, 223 (1983).
The talandic theory of the thermodynamics of biological systems is discussed in detail
by B. C. Goodwin, Temporal Organization in Cells, Academic Press, New York (1963).
Part Two

ENGINEERING THEORY
Chapter Eight

EQUATIONS OF STATE

BARBERINI: You think in terms of circles and


e/lypses, of uniform velocities and simple motions, that
is of things similar to your mind. But, suppose the
Almighty decided stars should move like this (makes a
strange gesture). Where would your calculations go then?

GALILEO: Then, Your Eminence, God would have


made our minds like this (makes the same gesture), so
we could believe a motion like this is simple. I believe in
the human mind.
B. Brecht

NOTATION

A Free energy density kcalkmol- 1


ak Coefficients in series expansion
ao Entropy at absolute zero kcal kmol- 1 K- 1
a( Constitute function for A kcalkmol- 1
B Henry's constant atm
b Cv at absolute zero kcal kmol K- 1
bk Series expansion coefficients
C Curvature of gEX curve
C Specific heat kcal kmol- 1 K- 1
cp Constant pressure specific heat kcal kmol- 1 K- 1
CV Constant volume specific heat kcal kmol- 1 K- 1
D Difference operator
f Fugacity atm
F( Corresponding states function m
G Free enthalpy density kcalkmor 1
g(T) Thermal free enthalpy of ideal gas kcal kmol- 1
gEX Dimensionless excess free enthalpy 207
208 H Enthalpy density kcal kmol- 1
Chapter Eight L Density of any extensive property
Lt Any extensive property
L' Partial molar L
L" Partial mass L
M Molecular weight kgkmol- 1
n (no subscript) Total number of moles kmol
n Number of moles kmol
PL Matrix defined in equation (8.3.7)
p Pressure atm
pos Osmotic pressure atm
R Gas constant kcal kmol K- 1
S Entropy density kcal kmol- 1 K- 1
s( ) Constitutive function for S kcal kmol- 1 K- 1
Sk( ) (k = 1,2) Compoments of s( kcal kmol- 1 K- 1
T Temperature K
U Internal energy density kcalkmol- 1
u( Constitutive function for U kcalkmol- 1
u Velocity ms- 1
V Specific volume m3 kmol- 1
Vs Velocity of sound ms- 1
v Fugacity coefficient
W Total weight fraction of solutes
w Weight fraction
y Total mole fraction of solutes
y Mole fraction
a Coefficient of cubic expansion K- 1
~ Isothermal compressibility atm- 1
T Activity coefficient
TD Activity coefficient for dilute solution
IL Chemical potential kcalkmor 1
<I> Mass density kgm- 3

Subscripts

o At reference state
00 At infinite dilution
C At critical point
G For gas 209
~j For component i,j Equations of
k Index in series expansion State

L For liquid
S For solvent

Superscripts

0 At standard state
* For mass-based formalism
EX In excess of ideal mixture value
ID For ideal gas
1M For ideal mixture
MIX In excess of additive part
S At saturation
T Transpose
8.1. THE IDEAL GAS

It is true that Regnault has lately shown, by a very careful


investigation, that this law [of ideal gases] is not strictly
accurate, yet the departures from it are in the case of
permanent gases very small, and only become of consequence
in the case of those gases which can be condensed into liquids.
From this it seems to follow that the law holds with greater
accuracy the more removed the gas is from the condensation
point with respect to pressure and temperature. We may
therefore, while the accuracy of the law for the permanent
gases in their ordinary condition is so great that it can be
treated as complete in most investigations, think of a limiting
condition for each gas, in which the accuracy of the law is
actually complete.
R. J. E. Clausius

Let us consider a one-component system such as discussed in Section 4.1. At any


given temperature, the a( V) curve has a last phase transition region; at volumes
larger than those, no more phase transitions occur and the system is said to be
a gas. At temperatures above the critical temperature, this definition implies that
the gas phase is reached by sublimation from the solid phase, and the pressure
at. the sublimation point may be extremely high, so that the resulting gas may
well have a density comparable to that of liquids.
As the volume of gas is increased, the pressure steadily decreases, until at
sufficiently high volumes the behavior approaches that of an ideal gas. It turns
out that the most common gas in nature, air, behaves as an ideal gas at ambient
conditions. While historically this is the reason why the ideal gas law was first
formulated, the definition of an ideal gas would be useful even if the behavior
it describes could in actual fact be observed only under very extreme conditions
(one need only consider thermodynamic theory as developed by hypothetical
scientists living on Jupiter, where the atmosphere on the surface certainly does
not behave as an ideal gas). For mechanical stability, pressure must increase as
volume decreases, and hence the simplest assumption is that they are inversely
proportional to each other, say that the product pV is a unique function of
temperature. Experimentally one observes that p increases with increasing tem-
perature, and hence the simplest assumption is that p V is proportional to T, and
that the proportionality constant is the same for all substances (this turns out to
be reasonable if one chooses the molar volume). Finally, one needs an equation
for internal energy, and the simplest possible assumption is that it depends only
on temperature and is independent of volume. This leads to the following 211
212 definition of an ideal gas:
Chapter Eight
pV=RT (8.1.1)

U = u(T) (8.1.2)

A system for which pressure is a unique function of volume and temperature


(such as an ideal gas) is an elastic system, and therefore all processes are reversible.
It follows that, when work is done isothermally on such a system, it is entirely
accumulated in the form of free energy:

-pV=A=U-TS (8.1.3)

In this sense, free energy can be regarded as elastic energy in isothermal


transformations. Indeed, in the classical theory of elasticity, where isothermal
transformations are considered, the stored elastic energy is the free energy.
Equation (8.1.3) shows that, in the isothermal deformation of an elastic system,
elastic energy can be accumulated by increasing the internal energy, or by
decreasing the entropy, or by a combination thereof.
Now suppose attention is restricted to systems for which elastic energy can
be accumulated only by decreasing the entropy. Such systems are said to be
endowed with entropic elasticity. For such systems, an isothermal deformation
results in no change of the internal energy. For the case of a gas, this means that
internal energy is independent of volume, and depends only on temperature.
Equation (8.1.2) is thus seen to identify an ideal gas as a system with entropic
elasticity.
If equation (8.1.2) holds, one can write

A = u(T) - Ts(V, T) (8.1.4)

and
8 = s(V, T) (8.1.5)

Differentiation of equation (8.1.4) with respect to temperature yields

-8 = du/dT - s(V, T)"': Tas/aT (8.1.6)


and therefore

as/aT = (l/T) du/dT (8.1.7)

The right-hand side of equation (8.1.7) is a unique function of temperature, and


therefore

(8.l.8)

It follows that the constitutive equation for entropy is of the following form:

(8.1.9)
When this is substituted into equation (8.1.4) and the partial derivative taken 213
with respect to volume, one obtains Equations of
State
(8.1.10)

and thus the conclusion is reached that in a system with entropic elasticity pressure
is proportional to absolute temperature. Therefore, equations (8.1.1) and (8.1.2)
are not entirely independent of each other: if equation (8.1.2) holds, p is
necessarily linear in T.
Ordinary elastic solids are not systems with entropic elasticity. However, the
elasticity of rubbers is, to a very good approximation, entropic, and the stresses
exhibited by rubbers at any given level of deformation are proportional to
temperature. This means that rubbers appear to be stiffer at higher temperatures,
contrary to ordinary elastic solids for which the elastic moduli are generally
decreasing functions of temperature.
When work is done adiabatically on a system with entropic elasticity, internal
energy does increase, since the first law requires that. The only way for U to
increase is for T to increase, and hence the system will heat up. This does not,
however, imply that dissipation of energy is taking place, since the phenomenon
of heating upon work being done on the system may well correspond reversibly
to a phenomenon of cooling of the system when it does work, and of course that
is the case for ideal gases.
While the phenomenon of a gas with behavior not very different from an
ideal gas heating up on adiabatic compression and cooling down on adiabatic
expansion is well known, the analogous phenomenon for rubbers is somewhat
more surprising. It is easy to convince oneself of the phenomenon by stretching
a rubber band rapidly and then touching it with the lips (which are sensitive to
small temperature variations). The rubber band will be seen to be distinctly hotter
than before stretching. If the rubber band is kept stretched for a sufficiently long
time, so that by heat transfer it reaches the temperature of the surrounding air,
and is then rapidly unstretched, it will cool down well below the ambient
temperature.
Entropic elasticity is observed in materials which, on a molecular scale, are
capable of significant variations of conformational entropy. Rarefied gases are
in this category because molecule-molecule interactions are negligible, and
changes of density simply result in changes of conformational entropy. Rubbers
(and, more generally, polymers in the molten or amorphous states) accommodate
deformations by changes of the conformational entropy of the macromolecular
segments, which are long and thin enough to allow coiling and uncoiling without
significant variations of internal energy.
We now tum our attention back to the ideal gas. Mter substitution of equation
(8.1.1), equation (8.1.10) can be integrated to give

S1 = So+Rln V (8.1.11)

where So is a (so far) unspecified constant. The function u(T) is left arbitrary
in the definition of an ideal gas. For any given system, the specific heat at con~
stant volume, i.e., the derivative of the u(T) function, can be determined
214 experimentally. Thus the internal energy is given by
Chapter Eight
U = u(T) = IT cvdT+ Uo
To
(8.1.12)

and its value is thus determined to within an arbitrary additive constant. Integra-
tion of equation (8.1.7) yields, after substitution of equations (8.1.9) and (8.1.11),

S = So + IT (C
To
v/ T) dT + R In V = si T) + R In V (8.1.13)

The other thermodynamic quantities are now easily calculated as

H = RT+ u(T) (8.1.14)

A = u(T) - TSo - Ts 2 (T) - RTln V (8.1.15)

G = RT+ u(T) - TSo - Ts 2 (T) - RTln V (8.1.16)

Since for an ideal gas the f( . ) function is invertible, it is useful to express


the free enthalpy in terms of pressure rather than volume:

G = IL = RT[l -In(RT)] + u(T) - TSo - TsiT) + RT lnp


= g(T) + RTlnp (8.1.17)

We note that, as long as So and Uo are regarded as unspecified constants,


the g( T) function is determined to within an arbitrary additive linear function
of temperature. However, as will be discussed later, the absolute entropy can be
determined experimentally, so that in fact So is not an arbitrary constant. This
in tum implies that quantities U, H, A and G are all determined to within the
same arbitrary additive constant (Uo). This does not pose any problem, since in
thermodynamic calculations one always needs only the difference between the
values of these four properties at two different conditions, and hence the arbitrary
constant Uo invariably cancels out.
The partial and baric derivatives of the thermodynamic functions for an
ideal gas are easily calculated as follows:

iJu/iJV = 8u/8p = iJh/iJV = 8h/8p = 0 (8.1.18)

iJs/iJV=R/V, 8s/8p=-R/p (8.1.19)

iJa/iJV = iJg/iJV = -RT/V, 8a/8p = 8g/8p = RT/p (8.1.20)

Finally, the difference between the two specific heats is

(8.1.21)
It is important to note that the definition of ideal gas chosen in this section 215
(which is the usual one) allows the function u(T) to be arbitrary, subject to the
Equations of
requirement that it should be monotonously increasing. The Maxwellian gas is State
a special case of an ideal gas: in fact, u(T) is, for the Maxwellian gas, a linear
function of temperature, and the specific heat at constant volume is 3R/2 at all
temperatures. From this viewpoint, the Maxwellian gas is more specific than the
ideal gas as defined in this section. However, while for a Maxwellian gas equation
(8.1.1) holds at all conditions, the stress tensor is not required to be isotropic
except at equilibrium. A Maxwellian gas can sustain tangential stresses under
non equilibrium conditions, and in fact it behaves mechanically in a way which
is significantly more complex than that of a Newtonian fluid as described by
equation (2.3.A.1); see, e.g., the discussion at the end of Section 5.1. The Maxwel-
lian gas model also predicts other nonequilibrium properties in addition to
tangential stresses, such as the heat flux and the diffusive fluxes; indeed, all
molecular models can predict both equilibrium and nonequilibrium properties,
although they are most commonly used only to predict the former ones.

8.2. ONE-COMPONENT SYSTEMS

In particular, though many eminent theorists have overlooked


it, considerations of molecular theory are less suited to co"ect
established thermodynamical laws than to be themselves
accommodated thereto.
W. Nemst

The equilibrium thermodynamic properties of anyone-component system


for which pressure is the only relevant stress can be calculated, provided the
following information is available:

1. The cv ( T) function at a pressure sufficiently low for ideal gas behavior.


2. The f() function in all possible phases of the system.
3. The values of DV and DH at every equilibrium phase transition.

This information can be obtained experimentally by calorimetry (cv and


DH) and by pressure-volume-temperature measurements at eqUilibrium. Two
parameters of particular importance are the isothermal compressibility {3 and the
coefficient of cubic expansion a:

V{3 = -I3V/l3p = -l/(ap/aV) (8.2.1)

and

Va = I3V/I3T (8.2.2)

For an ideal gas, {3 = 1/P and a = 1/ T.


For any extensive property L, the baric derivative with respect to pressure is

I3L/l3p = (aL/aV)/(ap/aV) = - V{3 aLja V (8.2.3)


216 Let us first consider entropy. From equations (2.7.10) and (8.2.3) one obtains
Chapter Eight
a8/iJV = ap/aT and 88/8p = - V,8 apjaT = - Va (8.2.4)

Equation (2.7.2) immediately yields

aU/av = -p + T as/av = -p + Tap/aT (8.25)

and
8U/ 8p = V,8(p - Tap/aT) (8.2.6)

From the definition of enthalpy one obtains

aH/aV = au /av + V apjaV + p = Tap/aT - 1/,8 (8.2.7)

and
8H/8p = V(l- ,8Top/aT) (8.2.8)

while equation (2.7.4) gives

8G/8p = V and aG/av= -1/,8 (8.2.9)

For free energy

iM/av =-p and M/8p = ,8pV (8.2.10)

We now first consider a gaseous system. The difference between the equili-
brium value of any thermodynamic quantity L and the value the same quantity
would have if the system were an ideal gas, LID, can be calculated in two different
ways:

(8.2.11)

(8.2.U)

The integrands are available from equations (8.2.4)-(8.2.10) and (8.1.18)-


(8.1.20). The integrals are always finite, since the integrands become zero in the
lower limit. Since LID is known, equations (8.2.10) and (8.2.11) yield the value
of L for any gaseous system in terms of only the information in (1) and (2)
above. The free enthalpy is of particular importance, and it is obtained from
equations (6.1.17), (8.2.9), and (8.2.12) in the form

G = geT) + RTlnp + J: (V - RT/p) dp (8.2.13)


The fugacity! and the fugacity coefficient v are defined as follows: 217
Equations of
G = geT} + RTln! = geT} + RTln (vp) (8.2.14) State

and RT In v is thus identified with the integral in equation (8.2.13). For an ideal
gas, the fugacity is equal to the pressure, and the fugacity coefficient is unity.
The above equations can be used as long as the system is a gas, i.e., up to
the pressure (or down to the volume) at which the first equilibrium phase transition
occurs. If DH and DV are known, the thermodynamic quantities in the nongas-
eous phase at eqUilibrium can be calculated from equations (4.1.8}-(4.1.11). For
example, at temperatures above the triple-point temperature, the properties of
the saturated liquid can be calculated. Equations (8.2.3}-(8.2.9) can then be used
to calculate the values of the equilibrium properties in the liquid at any other
volume smaller than the saturation volume, down to the volume at which freezing
occurs. The procedure can be repeated at every subsequent phase transition.
The compressibility of phases other than gaseous is invariably very small,
and often the approximation {3 = 0 is quite reasonable for solids and liquids.
However, a caveat is useful here. No real system is really incompressible, and
there are phenomena like the propagation of sound which are governed solely
by the compressibility. The velocity of sound is in fact equal to the inverse of
the square root of the isoentropic compressibility:

(8.1.15)

where cI> is the mass density. For an ideal gas, the right-hand side of equation
(8.2.14) is (cp / cv)(RT/ M). For liquids, the molar volume is in fact much smaller
than for gases, but the compressibility is so much smaller that sound actually
propagates in liquids faster than it does in gases. Of course, if one sets the
compressibility of a liquid to zero, one would calculate an infinite velocity of
sound. Propagation of sound is not the only phenomenon for which one cannot
regard the compressibility of a liquid as zero; the classical water hammer problem
of hydraulic engineering is another example of a phenomenon dominated by the
small, but finite compressibility of liquids.
The specific heats are given by

Cp = 8H/8T = T8S/8T (8.1.16)

and
Cv = aU/aT = TaS/aT (8.2.17)

Therefore, the specific heat difference is calculated from equation (2.6.11):

cp - Cv = V{3T(ap/aT)2 = VaT(ap/aT) (8.2.18)

For almost incompressible phases, the specific heat difference is best


expressed as p Va, and is thus seen to be often negligible. Therefore, for liquid
and solid phases the distinction between the two specific heats is often irrelevant,
218 and the symbol c will be used for both. Since c is a measurable quantity, the
following additional information is available for one-component systems:
Chapter Eight

4. The specific heat c as a function of temperature.

By making use of such information, equation (8.2.16), coupled with equation


4.1.10 at any equilibrium phase transition, can be integrated at any pressure down
to absolute zero. If one then accepts the "third law of thermodynamics," i.e.,
that S = 0 at T = 0, the absolute value of entropy can be extracted. However,
this procedure presents four important problems, as discussed below.
In order to carry out the procedure, the value of c must be measured down
to T = O. This is in practice very difficult experimentally, since at very low
temperatures c becomes a strong function of temperature, approaching 0 itself
as T approaches 0 [if it were not so, the integral of equation (8.2.16) would
diverge near T = 0].
The second problem is that the foundation of the third law ofthermodynamics
resides in the consideration of perfect crystals. Many one-component systems
may not transform into a perfect crystal at low temperatures, no matter how
slowly they are cooled down: a good example is that of polymers, which can
never be brought to a condition of total crystallinity.
The third problem is that determination of an absolute entropy is hardly
important from a pragmatic viewpoint. For chemical reactions, the value of DHo
can be measured calorimetrically, and the value of DOo can (if reliable constitutive
equations for the chemical potentials are available) be extracted from equilibrium
composition measurements and equation (3.4.5). It follows that a value for DSo
can be calculated, and if for all components the absolute entropy has been
obtained with the procedure discussed above, it can be checked against the
predicted value. The check is usually successful, which gives support to the third
law; yet no pragmatically important result has been obtained from it, though of
course the result is conceptually a confirmation of the validity of the third law.
Other, and more indirect experimental checks of the third law are possible, and
in all instances where such checks have been performed they have been successful.
The last problem is philosophical. The first and second laws are not based
on a molecular model argument, but the third one is. While molecular models
are useful in predicting with reasonable approximation the values of thermody-
namic quantities, they are inevitably only rough approximations, and there is
certainly some conceptual problem in basing what is regarded as a fundamental
law of physics (which the third law is supposed to be) on a molecular argument.
Let us now turn our attention to the free enthalpy density of a liquid. At
the equilibrium saturation pressure, it is equal to the saturated vapor chemical
potential, say

(8.2.19)

The definition in equation (8.2.14) can be extended to nongaseous phases,


so that the fugacity of the saturated liquid is

(8.2.20)
At any pressure higher than pS, the chemical potential can be calculated by 219
integrating equation (8.2.9). If the liquid is regarded as incompressible, this leads
Equations of
to State
(8.2.21)

where VL is the liquid molar volume. The chemical potential of a one-component


solid can be calculated analogously.

Appendix

The third law of thermodynamics was proposed by N ernst; his argument is,
except for differences in formalism, as follows.
Given a system for which the free energy is a potential for entropy, the
following equation holds:

A=U+TaAjaT (8.2.A.l)

where, for simplicity, one may regard the partial derivative as being at constant
volume (i.e., attention is restricted to systems for which the state is V, T; however,
the argument could be generalized). Now suppose that, in the neighborhood of
absolute zero, the internal energy at constant volume can be Taylor-series
expanded:

(8.2.A.2)

The constant b is recognized as the constant volume specific heat at absolute


zero. Equation (8.2.A.1) can now be integrated to yield

(8.2.A.3)

where ao is the constant of integration. This in turn yields

S = -aAjaT = ao + b In T + b + OCT) (8.2.A.4)

The first argument advanced by Nernst is that, unless b = 0, one would


conclude that the A( T) curve starts with an infinite derivative, and entropy would
be -00 at absolute zero. This in turn would imply that entropy becomes negative
below some temperature, while thermodynamic theory is based on the assumption
that it is an intrinsically nonnegative quantity. Thus it is concluded that b = 0,
i.e., that the constant volume specific heat approaches zero as T approaches zero,
and equations (8.2.A.2) and (8.2.A.3) become

U = Uo + O(T2) (8.2.A,5)

and
(8.2.A.6)
220 while ao is thus seen to be the entropy at absolute zero. At this point, based on
a series of considerations deriving from molecular arguments, Nemst makes the
Chapter Eight
additional hypothesis that A and U should be equal to each other as temperature
approaches absolute zero, and this of course implies that a o = O.
As discussed before, experimental verification of the third law can only be
accomplished unequivocally for some chemical reactions, for which the con-
clusion is that the standard entropy change at absolute zero is in fact zero. This
guarantees that the entropy per atom at absolute zero is the same for all those
components which participate in chemical reactions for which the check has been
made; it does not guarantee that the common value at absolute zero is zero.
It is interesting to observe that the molecular arguments used in support of
the third law are based on consideration of perfect crystals; yet Nemst thought
that only for liquids which vitrify rather than crystallize could the third law be
validated experimentally. ("Liquids, when strongly supercooled, are as a rule
frozen, and so the examination of them down to very low temperatures is rendered
impossible. We know, however, a large number of exceptions, in particular the
glasses; quartz glass may be mentioned as a striking example. Molten quartz, if
cooled sufficiently rapidly, does not crystallize, but passes continuously into the
condition of amorphous, solid quartz; the specific heat of this can be measured
without any special difficulty down to temperatures as low as may be desired.")
Note that the words "if cooled sufficiently rapidly" imply that in fact the glassy
state is not a true equilibrium state. In this regard, see also the discussion in
Section 4.5.
Another point worth some discussion is the question of the velocity of sound,
i.e., the velocity of propagation of a pressure discontinuity. Let us consider a
one-dimensional problem where the only component of velocity, U, is in the x
direction, and everything is uniform along the y and z directions. The momentum
balance, in the absence of gravity and viscosity, becomes

(8ol.A.7)

while the mass balance is

aiP/at + a(iPu)/ax = 0 (8.2.A.8)

Now if there is a relationship between density and pressure, p = p(iP), then


equation (8.2.A.7) becomes

iP(au/at + u au/ax) = -(ap/aiP)(aiP/ax) (8.2.A.9)

For a small perturbation about a state of rest, u = u' and iP = iPo + iP', one
may neglect all terms which are quadratic in the disturbance (primed) quantities,
and it is thus easy to eliminate u' by cross differentiation to obtain

(8.l.A.tO)

where the subscript 0 reflects the fact that the derivative is evaluated at the rest
state, i.e., at equilibrium. Equation (8.2.A.I0) is hyperbolic and admits discon-
tinuous solutions, with discontinuities propagating at a velocity (a p / aff) ) 1/2. Hence 221
the velocity of sound has been established-to within having established what
Equations of
the p(ff)) relationship may be. The usually accepted rationale for choosing the State
isoentropic compressibility is as follows: at the surface of discontinuity, there is
no heat flux, hence it is adiabatic, and if reversible the process is isoentropic.
There are important conceptual difficulties with this argument (suffice it to
consider that, at the traveling discontinuity, since pressure undergoes a discon-
tinuity, so does temperature, and hence the temperature gradient is infinitely
large), but it does lead to the correct result as far as the velocity of sound is
concerned. Whether the same argument can be duplicated for other types of
shock propagations is open to question.

8.3. IDEAL MIXTURES

For the writer ojJantastic stories to help the reader to play the
game properly, he must help him in every possible unobtrusive
way to domesticate the impossible hypothesis. He must trick
him into an unwary concession to some plausible assumption
and get on with his story while the illusion holds.
H. G. Wells

Section 8.1 was dedicated to the ideal gas involving a constitutive equation
for one-component systems that has two important properties: first, it is extremely
simple, and second, it is appropriate for the description of real one-component
systems at least under some limiting but realistic conditions. Section 8.2 dealt
with the introduction of correction coefficients for real systems which do not
necessarily behave as ideal gases. One may now follow a similar approach for
the case of mixtures. First, in this section we introduce the definition of a special
class of mixtures, which are called "ideal." The constitutive equation for ideal
mixtures is both very simple and is adequate for real systems at least under some
limiting but realistic conditions. In the next section we define corrections
coefficients for mixtures which deviate from the behavior of ideal mixtures.
As was discussed in Chapters 3 and 4, equilibrium calculations for both
chemical reactions and phase equilibria require knowledge of constitutive
equations for the chemical potentials as a function of composition, i.e., as a
function of the mole fraction vector y. Consideration of quantities which are
functions of composition presents some subtlety. Of course, one can write a'n
equation of the following form for the density L of any extensive quantity which
depends on composition:

(8.3.1)

with the partial derivatives being of course baric derivatives. However, partial
derivatives with respect to the mole fractions have no meaning, since if one of
the mole fractions changes, the others cannot all remain constant.
222 Let us now consider the total quantity L" which can be written as
Chapter Eight
(8.3.2)

If nj is changed by an amount dnj and all the others are kept constant, the
corresponding dL, is

dL, = Lj dnj (8.3.3)

while the mole fraction is changed by

(8.304)

In this sense, a baric derivative of L with respect to Yj can be defined as


follows:

(8.3.5)

It is important to realize that the baric derivative so defined represents the


derivative with respect to Yj when the ratios of all other quantities yare kept
constant. The point is best understood by considering a ternary mixture, the
composition of which is identified by a point in the usual triangular diagram;
see Figure 8.3.1. Let A be the point at which the baric derivative in equation
(8.3.5) must be evaluated. This is the slope of the L surface at point A along the
dashed line, which identifies the locus of compositions such that the ratio Yk/ YI

.~
A
..

L
... K

FIGURE 8.3.1. Direction along which the 13/I3Yi derivative is taken.


is constant. It can easily be verified that the Gibbs-Duhem equation also holds 223
for the baric derivatives defined above, i.e.,
Equations of
State
y. 8L'/8yj = 0 (8.3.6)

We note that there are N independent Gibbs-Duhem equations. It is useful


to define the following N x N matrix for any extensive quantity L t :

P L = 8L'/8y (8.3.7)

and the Gibbs-Duhem equation can be written as

p[. y = 0 (8.3.8)

One is now in the position of writing the simplest possible constitutive


equation for a mixture, Le., as simple as possible a form for the matrix 8ILd 8Yj.
The first requirement is as follows. The chemical potential of component i depends
only on its own mole fraction, and is independent not only of the mole fractions
of the other components, but of their number and their nature as well. This
requirement sets a constraint on the off-diagonal elements of the matrix, as
discussed below.
Since ILi depends only on Yi, an arbitrary change of the mole fractions of
all other components j should result in a zero change of ILi. Such an arbitrary
change is, however, subject to the following requirement:

(8.3.9)

It follows that ILi will depend only on Yi provided that

(8.3.10)

with f(Yi) an as yet unspecified function. The Gibbs-Duhem equation can now
be used to express 8ILd 8Yi in terms of f(y;):

(8.3.11)

This equation shows that one cannot simply assign f(Yi) = O. Now the value of
f(y;) has dimensions of an energy density and can depend at most on temperature
(see the discussion in Section 8.8). Thus the simplest possible form for f(y;) is

(8.3.12)

and correspondingly

(8.3.13)

Solutions described by equations (8.3.12) and (8.3.13) are called ideal. The
strength of this definition of ideal solutions is not only its simplicity, but the fact
224 that gaseous mixtures at moderate pressures and some liquid solutions are in
fact adequately described by it.
Chapter Eight
Equation (8.3.13) can be integrated directly between any two mole fractions
Yjl and Yj2 to yield

(8.3.14)

In particular, if Yj2 = 1, one has

(8.3.1S)

where f.L J is the chemical potential of the pure component, Le.,


f.LJ = gj(T) + RTlnfjo (8.3.16)

The fugacity of a component of a mixture is defined as follows:

(8.3.17)

so that equation (8.3.16) assumes the following very simple form:

(8.3.18)

Equation (8.3.18) is often referred to as Raoult's law.


Ideal solutions exhibit some interesting properties. Equations (3.3.17) and
(8.3.16) yield

V'=VO (8.3.19)

Le., the partial molar volumes are constant and equal to the pure components'
volumes. This in turn implies that the volume of an ideal solution is additive, Le.,

(8.3.20)

From equations (3.3.16) and (8.3.12) one obtains

S' = SO - R lny (8.3.21)

which implies that entropy is not additive. For nonadditive extensive properties,
the mixing value is defined as follows:

L MIX = L - Y L (8.3.22)

and hence the entropy of mixing for ideal solution is

SMIX = -y . R In y > 0 (8.3.23)


The free enthalpy of mixing is 225
Equations of
a M1X = RTy lny < 0 (8.3.24) State

It is noteworthy that a M1X is always concave upward, and hence no phase


separation can occur in an ideal mixture. Straightforward algebra leads to the
conclusion that internal energy and enthalpy are additive, and that the free energy
of mixing is equal to the free enthalpy of mixing.
Gaseous mixtures are ideal solutions, unless the pressure is very high. This
is the reason why gases are always miscible in all proportions, and no gas-gas
phase transition is known to occur. In fact, one-component gases start deviating
from ideal gas behavior at pressures significantly lower than those at which
gaseous mixtures start deviating from ideal solution behavior; thus there is a
rather wide range of pressures where a gas mixture behaves as an ideal solution
of nonideal gases. It follows that the pure component fugacity appearing in
Raoult's law may well be different from pressure. However, its value is still
determined by only pure component properties (specifically, its equation of state).
Since no "interaction" parameter appears in the constitutive equations of ideal
solutions, the thermodynamic properties of such solutions can be calculated by
knowing only the equations of state of the individual components.
There is one conceptual problem with the meaning offt Since the integration
called for in equation (8.3.14) is a baric integration, it needs to be performed
while staying in the same phase, say in the case of a gaseous mixture fjo is the
fugacity of component j pure, at the same temperature and pressure as the
mixture, and in a gaseous state. Such a state may well not be an equilibrium
state, or not even a possible nonequilibrium one. A simple example clarifies this
point: moisture in the atmosphere. For water,fjO would be the fugacity of gaseous
H2 0 at 1 atm and 20C-but at such a condition water is liquid at equilibrium,
and water vapor at 1 atm cannot possibly be undercooled down to 20C. This
difficulty is circumvented as follows. Humid air at atmospheric conditions can
be regarded as an ideal gas. Hence the fugacity of water vapor in the air is simply
its partial pressure. Now consider liquid water at 20C and 1 atm; its fugacity is

(8.3.25)

and that must be equal to the equilibrium partial pressure of water vapor in the
saturated air, py, which can be measured. Thus it is concluded that the constant
f O appearing in equation (8.3.18) is

(8.3.26)

This point will be discussed in more detail in the section on dilute solutions.
In contrast to gases, liquid mixtures are generally not ideal solutions, unless
all components have closely analogous molecular structures.
226 8.4. ACfIVITY COEFFICIENTS
Chapter Eight
In the course of its development, a physical theory is free to
choose whatever way it pleases, provided logical contradictions
be avoided; in particular, it is free to take no account of
experimental facts.
It is not the same when the theory gains its full development.
When the logical structure has reached its full height, then we
must compare the whole of its mathematical propositions,
obtained as conclusions from these long deductions, with the
whole of the facts of experience.
P.Duhem

Many mixtures do not behave as predicted by the constitutive equation of


ideal mixtures. The behavior of nonideal mixtures is described in terms of activity
coefficients. The fugacity is still defined as in equation (8.3.17); however, in
nonideal solutions its value is not given by equation (8.3.18), and the following
definition of activity coefficient 'Tj is introduced:

(8.4.1)

so that the activity coefficients are all unity in ideal solutions.


Another useful definition is that of the excess of the density of any extensive
property:

(8.4.2)

where LIM is the value of L one would calculate from ideal mixtures theory. For
those properties which are additive in ideal mixtures, LEX coincides with L MIX.
The excess properties can all be expressed in terms of the activity coefficients:

.... EX = RT In 'T (8.4.3)

G EX = RTy In'T (8.4.4)

VEX = VMIX = RTy l3ln 'T/ l3p (8.4.5)

SEX = -RTy' l3ln 'T/ l3T (8.4.6)

HEX = H MIX = RTy (In 'T - Tl3 In 'T/ l3T) (8.4.7)

UEX = U MIX = RTy (In'T - Tl3 In 'T/ l3T - pl3 In 'T/ l3p) (8.4.8)

A EX = RTy (In'T - pl3 In 'T/ l3p) (8.4.9)

Equation (8.4.3) implies that the Gibbs-Duhem equation holds for the In'T
vector, say

Pl:..y = (l3ln'T/l3y)T.y = 0 (8.4.10)

By definition, the activity coefficient of a pure component is unity, i.e.,

limln'Tj=O (8.4.11)
'Tj=l
The Gibbs-Duhem equation for componentj implies that 227
Equations of
lim(S In 'T) SyJ = 0 (8.4.12) State
Tj=l

If the mole fraction of a component is close to unity, that component is


called the solvent. Equation (8.4.121 guarantees that the activity coefficient of the
solvent can be taken as unity to within second order in 1 - Yi> say

'Tj = 1 + 0[(1 - yYJ (8.4.13)

Equation (8.4.3) can be rewritten in the following form:

(8.4.14)

and therefore G EX approaches zero linearly when any of the mole fractions
approaches unity.

8.5. DILUTE SOLUTIONS

Jack Sprat could eat no fat;


His wife could eat no lean:
and so, between them both,
they licked the platter clean.
Lewis Carroll

Constitutive equations for chemical potentials simplify very considerably in


the case of dilute solutions. A dilute solution is defined as a mixture in which
the mole fraction of one component, called the solvent, is close to unity. We will
indicate the properties of the solvent with no suffix, so that the definition of
dilute solution is y "" 1. The theory of dilute solutions is essentially a zero-order
approximation, where all quantities of order 1 - yare neglected with respect to
quantities of order unity.
Let us first consider the case of a binary mixture, for which the Gibbs-Duhem
equation can be written in the form

yS In 'T/Sy = (1- y)S In 'Tddy (8.5.1)

Equation (8.5.1) implies equation (8.4.12) provided the baric derivative appearing
on the right-hand side does not approach infinity when y approaches unity, at
least not as rapidly as 1/(1 - y). The theory of dilute solutions is based on the
assumption that indeed such baric derivatives always satisfy the stated restrictions.
It should be noted that the conclusion that the solvent's activity coefficient is
unity is a first-order approximation, since deviations are of the order of (1 _ y)2.
This, as will be seen in Section 8.8, is a very important result. In fact, while
dilute solution theory is a zero-order approximation as far as the solutes are
concerned, it happens to be a first-order approximation for the solvent, and
228 within the first order the solvent behaves as if the solution were ideal. This makes
possible experimental verification of the theory.
Chapter Eight
The activity coefficient of the solute can be expressed as

(8.5.2)

and hence, to within the zero-order approximation, it can be taken constant and
equal to its value at infinite dilution, 'Tleo'
Now we suppose that another solute is added to the system, but in small
enough an amount to maintain the solution dilute, so that Y2 is of order 1 - y.
The activity coefficient for the first solute can be written as

(8.5.3)

and therefore, again to within the zero-order approximation, it can be taken equal
to its infinite dilution value. The procedure can obviously be reiterated, and one
concludes that all the activity coefficients can be regarded as constant:

T = Teo (8.5.4)

Equation (8.4.1) thus becomes

(8.5.5)

which is called ~enry's law.


As long as a mixture is regarded as a dilute solution, the values of fO and
'Teo for any given solute need not be known separately, but only their product,
i.e., the Henry constant B, must be known. The latter is often more easily accessible
to measurement than the former two. In particular, fO may well correspond to a
state which is neither an equilibrium state nor an attainable nonequilibrium state,
analogously to what was discussed at the end of Section 8.1. Consider, for
example, the case of a sparingly soluble gas in a liquid, say CO 2 in water at room
temperature and 1 atm. The value of fO for CO2 in the liquid phase would be
the fugacity of pure liquid CO 2 at those conditions. But at those conditions CO2
is always a gas, since liquid CO 2 cannot be superheated at 1 atm up to room
temperature (an even more extreme case would be the case of O2 solubility at
the same conditions, since the critical temperature of O 2 is well below room
temperature). However, consider now gaseous CO 2 , which at 1 atm can be
regarded as an ideal gas, and hence its fugacity is equal to the pressure. The
fugacity of CO 2 dissolved in water in equilibrium with pure gaseous CO2 at 1 atm
is therefore also 1 atm. Since the mole fraction of CO2 in water under such
equilibrium conditions can be measured, the Henry constant can be determined
experimentally.
For the same case as discussed above, the value offO could also be calculated
in the following way. Consider the gas-liquid equilibrium of pure CO 2 at room
temperature, where the two phases have the same fugacity. The pressure is high
enough for CO 2 not to be an ideal gas, but the gas fugacity can be calculated
from the equation of state of the gas. Thus the fugacity of pure liquid CO 2 at
this high pressure can be calculated. The fugacity at 1 atm can now be formally 229
calculated from equation (8.2.21). The calculation is only a formal one, because
Equations of
it is based on the consideration of liquid volumes at pressures where the liquid State
cannot exist in reality. However, this does not pose any problem, because any
artificiality in the meaning of fO in equation (8.5.5) is automatically adjusted by
regarding it as part of the definition of Too. We note that this second procedure
would not be possible for the O 2 case.
As the mole fractions of the solutes become larger, the behavior of the
mixture starts deviating from dilute solution behavior. In this case, it is useful
to define a second type of activity coefficient, which will be identified by super-
script D, as follows:
(8.5.6)

The relationship between the two activity coefficients is

(8.5.7)

Consider the case of a binary mixture, and let the fugacity of one of the
components vary with its mole fraction as in Figure 8.5.1. Raoult's law corresponds
to the lower straight line, and Henry's law to the upper one, which by definition
is tangent to the curve at the origin. The Raoult's law line is tangent to the curve
at y = 1 because of equation (8.4.12).

8.6. CONSTITUTIVE EQUATIONS FOR ONE-COMPONENT SYSTEMS

Why is the equivalence of the practically. rigid body and the


body of geometry-which suggests itself so readily-rejected by
Poincare and other investigators? Simply because under closer
inspection the real solid bodies in nature are not rigid. ...
A. Einstein

So far, constitutive equations for one-component systems have not been


written down except for the ideal gas. The fugacity coefficient has been defined,
and it has been shown that it is a measurable quantity, but no actual equations
for it have been given.

FIGURE 8.5.1. Fugacity of a component


of a binary mixture vs. its mole fraction.
230 As the pressure approaches zero, and therefore the volume approaches
infinity, ideal gas behavior is approached:
Chapter Eight

limpVIRT=1 (8.6.1)
t/v"'o

If it is assumed that the equation of state can be expanded as a power series


in the molar density II V. then we have

pVIRT=I+IadVk (8.6.2)

which can be employed with the sum truncated after some finite number of terms
as the equation of state for a one-component system. Equation (8.6.2) is called
the virial equation of state, where ak are the virial coefficients. The latter can be
obtained either by fitting experimental data, or from an appropriate molecular
model. The series expansion in equation (8.6.2) is in terms of II V. and hence is
at some fixed T; consequently, the coefficients ak could well depend on T.
Of course, the power-series expansion cannot be valid over a phase transition,
where the molar density suffers a discontinuity. Therefore, in order to use the
virial equation of state for nongaseous systems, one has to assume that it represents
the one-phase behavior through the phase transition, say a continuous curve such
as in Figure 2.6.1. This in turn implies that the derivative apjaV must change
sign twice when going through any phase transition.
The derivative is

(8.6.3)

and therefore every additional change of sign requires two more terms in the
series. If the equation must be used through several phase transitions, the number
of terms needed becomes large enough for calculations to be cumbersome. If the
equation is used only for the liquid phase (or the solid phase in the case of
sublimation at temperatures low enough for the intermediate bump in Figure
4.4.1 to have disappeared), two nonzero virial coefficients are the minimum
requirement, since a third-order polynomial may have the shape in Figure 2.6.1.
The virial equation of state has as many parameters as there are terms in the sum
retained.
The Van der Waals constitutive equation is a slight modification of the
two-parameter vi rial equation of state, since the first term in the series is written
as all (V - at). This of course implies that a l is the smallest possible volume of
the system, since pressure approaches infinity when the volume approaches a l
For the classical virial equation of state, pressure approaches infinity only when
the volume approaches zero.
It is of interest to consider the dimensionality of the equation of state. Since
p, Vand T are dimensional quantities, the mathematical form ofthe f( . ) function
depends on the choice of the units of measurement. It can always be made
dimensionless (Le., invariant with respect to a change of units) by introducing
three dimensional parameters having units of pressure, volume, and temperature:

PiPe = F(VIVe , TITd (8.6.4)


In the case of the ideal gas, the three parameters collapse to only one (R), 231
and the latter is in fact a universal constant, Le., it has the same value for all
Equations of
substances. State
The equation of corresponding states is based on the hypothesis that the
function F( . ), though not the values of the three parameters appearing in it, is
the same for all substances. If this assumption is accepted, the behavior of any
one substance is totally identified by the values of the three parameters. Since
all one-component systems exhibit the critical-point behavior, the choice of the
critical pressure, critical volume, and critical temperature as parameters is obvious.
The equation of corresponding states is generally used in the form of tabulated
values of pV/ RT as determined by PiPe and T/Te; it is remarkably successful
for gases, but of little practical utility for liquids and solids.
As discussed before, for many phenomena liquids and solids can essentially
be regarded as incompressible, so that their equation of state reduces to

v= VeT) (8.6.5)

The coefficient of cubic expansion is defined as

a = 81n V/8T (8.6.6)

and it is generally a weak functiqn of temperature.

Appendix

The Van der Waals equation of state can be expressed in the form

P = RT/(V - b) - a/V2 (8.6.A.l)

and is thus a cubic with two adjustable parameters (a and b). The values of the
two parameters can be chosen in such a way as to satisfy the conditions required
at the critical point:

ap/av = a2p/av2 = 0 (8.6.A.2)

Hence
b = Ve /3 (8.6.A.3)

and
a = 9RTe Ve /8 (8.6.A.4)

Now, of course, whether one third of the critical volume is really the lowest
possible attainable volume is open to question. However, this is not the important
point, since the Van der Waals equation is not supposed to be valid at extremely
high pressures. It is of more interest to consider variations of the Van der
Waals equation which have been used with varying degrees of success. The
most commonly used variation is the Redlich-Kwong equation, which can be
232 written as
Chapter Eight
p = RT/(V - b) - a/[JT(V2 + bY)] (8.6.A.5)

and again has only two (constant) parameters, but is more successful than the
original Van der Waals equation in fitting data. Addititional refinements have
been considered in the literature. Values of the parameters and functions appear-
ing in such equations have generally been constrained by the requirement that
equation (8.6.A.2) should hold at the critical point.

8.7. CONSTITUTIVE EQUATIONS FOR MIXTURES

Mathematics is not a science. It can form an elegant and


imaginative exercise for those to whom painting is too concrete
and crossword puzzle too trivial. Sometimes there is the
impression that the mathematician can be kept as a priest of the
oracle, to make an infallible pronouncement on the infrequent
occasions when he is consulted, but normally to be ignored
while the scientist gets on with the real work.
B. E. Noltingk

As for one-component systems, constitutive equations for mixtures are


expressed in the form of corrections to the ideal mixture case-fugacity coefficients
in the former case, activity coefficients in the latter. In particular, in view of
equation (8.4.14), constitutive equations are best written down for the excess free
enthalpy. We shall first consider binary mixtures.
Any constitutive equation for the excess free enthalpy must satisfy two
conditions. First, the excess free enthalpy must go to zero when either one of the
two mole fractions goes to zero. Second, it must do so linearly, since the success
of the theory of dilute solutions guarantees that the baric derivatives of the activity
coefficients with respect to the mole fractions never become infinitely large. The
simplest possible constitutive equation for the excess free enthalpy of a binary
mixture which satisfies this requirement is

(8.7.1)

with a being the only parameter. From equation (8.4.14) one obtains

(8.7.2)

and
In T2 = ayi (8.7.3)

Equations (8.7.2) and (8.7.3) are the "one-suffix" Margules equations for the
activity coefficients. A few binary mixtures are adequately described by them.
Let us return to the discussion in Section 4.2. In order to have a two-phase
equilibrium system, the curve of gEX vs. Yt must have a spinodal region, i.e., its
second derivative or curvature, C, must be negative for some values of Yt. For
an ideal solution 233
Equations of
(8.7.4) State

which is intrinsically positive-and in fact infinity when either of the two mole
fractions approaches zero. Since a EX approaches zero linearly in both limits, e
is necessarily positive in those limits-implying that complete immiscibility can
never occur. For a binary mixture adequately described by equation (8.7.1), the
value of e is given by

e = e lM -2a (8.7.5)

and thus a spinodal region will occur provided a> 2 (the minimum value of
elM is 4). Of course, the value of a may be either positive or negative. Mixtures
for which gEX is positive are said to exhibit positive deviations from ideality,
and in actual fact represent the most common case. Mixtures which exhibit
negative deviations from ideality generally do not exhibit a miscibility gap, though
mathematically it would not be impossible to write a constitutive equation for
gEX such that gEX < 0 but e < 0 for some YI'

It is easy to construct constitutive equations of increasing order of approxima-


tion for gEX, with equation (8.7.1) being the first-order approximation. One may
choose a formal multiplicative expansion to obtain, in the second-order approxi-
mation,

gEX = aOYIY2 + a lyiY2 + a2YIY~ (8.7.6)

which yields

In'Tl = y~(blO + blYI) (8.7.7)

and
In'T2 = yi(b20 - b1Y2) (8.7.8)

where
blO = ao + a2 (8.7.9)

b20 = ao + al (8.7.10)

bl = 2(a l - a2) (8.7.11)

which are called the "three-suffix" Margules equations.


Equation (8.7.1) implies that the activity coefficients at infinite dilution are
equal for the two components of the binary mixture; equations (8.7.7) and (8.7.8)
allow for two different values ofthe infinite dilution activity coefficients. However,
the latter equations contain three parameters; a lower-order expansion, which
allows for different infinite dilution activity coefficients and contains only two
234 parameters, is as follows:
Chapter Eight
(8.7.12)

which results in
(8.7.13)

and an analogous equation for In 72' Equation (8.7.12) is the Van Laar equation
for activity coefficients; higher-order Van Laar expansions, as well as higher-order
Margules equations, are of course possible.
Higher-order Margules-type approximations simply lead to the following
type of constitutive equation for the activity coefficient of component 1:

(8.7.14)

Constitutive equations different in form from equation (8.7.14) do occasion-


ally arise from a molecular model. Indeed if, in order to be represented by an
equation of the form of equation (8.7.14), the behavior of any given system needs
more than a very few terms in the sum, some other form may be preferable for
purposes of calculation. It should be noted, however, that any constitutive
equation containing a y~blO term must be of the form

(8.7.15)

with 1(0) = 0 in order to be consistent with the theory of dilute solutions. The
value of blO is of course the value of In 7100'
Let us now consider ternary mixtures, for which the simplest form of
constitutive equation for gEX is

(8.7.16)

Equation (8.4.14) yields

(8.7.17)

and the equations for the other activity coefficients are obtained by cyclic permuta-
tion of indices.
Although Equation (8.7.16) is legitimate, it contains a conceptual problem.
While the properties of ideal mixtures can be calculated by knowing only proper-
ties of the pure components, that is not true for nonideal mixtures: in the binary
case, even the simplest possible constitutive equation (the Margules equation)
contains a parameter which can only be measured by experiments conducted
with the mixture itself. Now for equation (8.7.16), consider the situation where
Y3 is zero. The last two terms drop out, and thus the value of al2 could be
determined by experiments conducted on the binary mixture of components 1
and 2. A similar argument holds for the other two parameters, and therefore one
would conclude that the properties of the ternary mixture can be predicted by
measurements carried out only on the three binary component mixtures. That
does not appear very convincing. Indeed, it is natural to regard the au parameters 235
as interaction parameters, and thus equation (8.7.16) only accounts for binary
Equations of
interactions, while in a ternary mixture ternary interactions are possible. State
In fact, the problem discussed above disappears in the next-order approxima-
tion, which allows for ternary interactions:

gEX = a12YIY2 + a13YIY3 + a23Y2Y3 + a122YIY~


+ a 133YIyj + a 112 yiY2 + a113yiY3
+ a233Y2yj + a223Y~Y3 + a123YIY2Y3 (8.7.18)

This produces for the activity coefficients equations of the following form:

In 71 = b22Y~ + b33yj + b23Y2Y3 + b222Y~


+ b333Y~ + b122Y~YI + b 133yjYI + b123YIY2Y3 (8.7.19)

Clearly, the b123 coefficient can only be obtained from experiments conducted
on the ternary mixture itself.
The above argument can be reiterated as any new component is added: with
only binary and ternary interactions, the behavior of a 4-component mixture can
be predicted from the results of experiments performed only on the 4 ternary
component mixtures, and so on. In general, experiments on the N-component
mixture itself will be needed whenever N-ary interactions are allowed for. If one
allows only interactions up to, say, M-ary interactions, then experiments are
needed only for mixtures up to M components, and the behavior of mixtures
with more components can be predicted from these.
Since there is always some conceptual problem with the idea that the behavior
of a mixture can be predicted without any experiment being conducted on the
mixture itself, one would like to allow for N-ary interactions at least. But, as the
examples in this and the preceding section have shown, the algebraic complexity
and the number of parameters grow very rapidly as either the order of the
interactions or the number of components grow. The algebraic complexity could
be handled by modern computing facilities, but the large number of parameters
to be determined experimentally creates problems of accuracy of measurement
which cannot be circumvented. It is therefore clear that predictions obtainable
from molecular models are particularly important in this area.

S.S. MOLAR UNITS AND COLLIGATIVE PROPERTIES

In most questions the analyst assumes, at the beginning of his


calculations, either that matter is continuous, or the reverse,
that it is formed of atoms [molecules). In either case, the
results would be the same.
H. Poincare

Except for the discussion in Section 7.3, the thermodynamic theory developed
so far is all based on the use of molar units. It should be kept in mind, however,
236 that it could equally well be developed in mass units. The preference for molar
units is, for the material discussed so far, essentially a matter of convenience.
Chapter Eight
With molar units, the stoichiometric coefficients for chemical reactions are small
integers, which is preferable to the noninteger values which would be needed for
mass units. The ideal gas law has the universal constant R when molar units are
used, while it would have different values for different gases if mass units were
used (the mass-based constant being R/ M). Partial mass properties L" can be
defined in perfect analogy with the partial molar ones L'; the former are simply
the latter divided by the molecular weight of the component considered, i.e.,
L" = L'/ M. Since the Gibbs-Duhem equation is simply a consequence of the
fact that extensive properties are absolutely additive functions of mass, it holds
for partial mass properties as well.
A crucial difference, however, arises in the definition of ideal mixtures. The
mole-based definition is given in equations (8.3.9) and (8.3.10), which reduce to
equation (8.3.13). One could perform the same algebra in mass-based units and,
if w is the weight fraction, one would obtain for the mass-based chemical potential
(identified by an asterisk)

(8.8.1)

It should be noted, however, that equation (8.8.1) satisfies the Gibbs-Duhem


equation only if A is a constant which may depend on temperature but not on
the nature of the component considered; hence it is legitimate only if such is the
case.
Suppose one chooses this mass-based definition of an ideal solution, and
one defines mass-based activity coefficients T* so that, in general (i.e., for solutions
which are not mass-based ideal),

JL* = JL*o + A In(wT*) (8.8.2)

Since the Gibbs-Duhem equation holds for JL * and for In w, with this definition
it also holds for In T*.
Let us consider a solution which is mole-based ideal, say equation (8.3.13)
holds. Since JL * = JL/ M, one obtains

In T* = (RT/ MA) In(y/w) (8.8.3)

i.e., a mole-based ideal solution is not ideal in the mass-based formulation, and
vice versa. This is the crucial reason for choosing the mole-based formulation:
gases at moderate pressures are mole-based ideal solutions, while there is no
concrete case described by the mass-based ideal solution definition. Of course,
this conclusion is supported by molecular-type arguments: in the Maxwellian
theory of dilute gases, the "particles" are molecules, endowed with a mass
proportional to their molecular weight.
The point made in Section 8.3 about the strength of the theory of mole-based
ideal mixtures is now reinforced. A mass-based ideal solution is one where the
chemical potential of every component depends only on its own weight fraction,
and is independent of what the weight fractions of other components may be,
and of their number and nature. This is as simple as a mole-based ideal solution, 237
but is not realistic-no real system is mass-based ideal. Equations of
One could also choose an alternate approach for the mass-based formalism: State
one could omit entirely the definition of ideal solution and use equation (8.8.2)
as a definition of the mass-based activity coefficients T*. One would now obtain
the following relationship between mass-based and mole-based activity
coefficients:

Aln(wT*) = (RT/M)ln(YT) (8.8.4)

and one is tempted to choose the constant A as RT/ M. Such a selection would
make the constant in equation (8.8.2) dependent on temperature and on the
nature of the component considered, but not on the nature of the other com-
ponents. However, there are two problems with such a choice. First, since now
A does depend on the nature of the component considered, the Gibbs-Duhem
equation would hold for A In( WT*) and for In w, but not for In T*. Second, and
even more important, with such a choice (or with any choice where A is not a
universal constant at any given temperature) under no conditions could T* be
unity, since if it were the Gibbs-Duhem equation would not be satisfied.
The point is particularly relevant to the behavior of the solvent in a dilute
solution, since this is assumed to be what would be observed in a mole-based
ideal solution (in the first-order approximation)-and hence it is not what would
be observed in a mass-based ideal solution. For the solvent, T = 1, and if Y is
the sum of the. mole fractions of all solutes and W the sum of their mass fractions,
in the first-order approximation equation (8.8.4) reduces to

W + In T* = (RT/ MA)Y (8.8.5)

where M is the molecular weight of the solvent. We note that even the choice
A = RT/ M does not make T* = 1, except when the average molecular weight
of the solutes, M', is equal to M. In the first-order approximation, WM = YM',
and hence the choice A = RT/ M' would indeed give T* = 1; however, this would
imply that the definition of the solvent's activity coefficient depends on the nature
of the solutes-which is of course totally unacceptable. Hence the only sensible
way of developing a mass-based formalism would be to define ideal solutions
by equation (8.8.1), with A depending only on temperature.
Of course, from a purely descriptive viewpoint the mass-based formalism is
perfectly equivalent and, based on the same sequence of assumptions, one would
obtain for the fugacity of the solvent in a dilute solution

(8.8.6)

while in the mole-based approach one obtains

(8.8.7)

It is noteworthy that, in the zero-order approximation (which is the only


one used for solutes in the absence of activity coefficients), both equations (8.8.6)
238 and (8.8.7) simply reduce to f = fO. In the second-order approximation, again
any difference could be included in the definition of the solvent's activity
Chapter Eight
coefficient. However, in the first-order approximation equations (8.8.6) and (8.8.7)
give different results, and only one of them can be right. The point is related to
the fact that the sequence of assumptions leading to the equations given includes
that the derivative on the right-hand side of equation (8.5.1) does not approach
infinity (at least not as fast as 1/ Y) when Y approaches zero, and the analog
assumption in the mass-based formalism is incompatible with it.
The above discussion essentially implies that the validity of equation (8.8.6)
within the first-order approximation has been established on the basis of experi-
mental evidence. The experimental' evidence is based on mixtures where the
molecular weights of the different components are of the same order of magnitude,
and for such mixtures therefore the choice of the molar units approach is not
simply a matter of algebraic convenience for the theory of dilute solutions. In
the case of mixtures of polymers and low moleclar weight compounds, the
condition of molecular weights of the same order of magnitude is violated, and
indeed in the thermodynamic theory of such mixtures formalisms other than the
mole-based one are generally used.
Since the question of the difference between mass and molar units arises
only in connection with the fugacity (or the chemical potential) of the solvent,
the whole discussion above is pragmatically relevant only to those cases where
some measurable quantity depends on the chemical potential of the solvent. Such
quantities are called colligative properties. If equation (8.8.6) indeed holds within
the first-order approximation, the chemical potential of the solvent is given by

(8.8.8)

The important point about equations (8.8.6) and (8.8.8) is that the right-hand
side depends on Y, but is independent of the nature and the relative amounts of
the solutes. This statement is not equivalent to that obtained by substituting W
for Y in it, and the theory of colligative properties discussed below can only
apply to one formulation-under the restrictions discussed above, it applies to
the molar-based formulation. The most important colligative properties are (under
appropriate qualifying conditions) the boiling temperature, the freezing tem-
perature, and the osmotic pressure. These are discussed in the following.
The boiling temperature is a colligative property if all the solutes have
negligibly small volatilities, so that when boiling occurs the vapor phase consists
of the pure solvent. Let us first examine the case when the solvent vapor can be
regarded as an ideal gas, so that at the equilibrium boiling point its fugacity is
equal to the pressure on the system. Quantity f O is, for the same case, the vapor
pressure of pure solvent. The equilibrium boiling point corresponds to the
temperature at which f becomes equal to the total pressure: since f is less than
f O, this corresponds to a higher boiling temperature than that of the pure solvent.
Without any approximation, the boiling-point increase can be calculated as
,follows. Considering equation (3.3.16), the partial molar enthalpy of the solvent
can be written as

(8.8.9)
An analogous equation can be written for the pure vapor. Since the properties 239
of the solvent are the same as would be observed in an ideal solution, H' is the Equations of
enthalpy of the pure solvent. Substitution of equation (8.8.2) and integration yields State

RY = fT (DH/T2) dT (8.8.10)
Ts

where T is the boiling point of the solution and Ts that of the pure solvent.
Equation (8.8.10) is seen to contain no properties of the solutes, and hence it
can be used to extract the value of Y from the measurement of T.
The freezing temperature is a colligative property if the solid phase which
forms upon freezing essentially consists of the pure solvent. This is in fact often
the case. Equilibrium freezing occurs when the chemical potential of the solvent
is equal to that of the pure solid. The latter decreases with temperature at constant
pressure, and hence freezing occurs at a temperature lower than it would for the
pure solvent (that is the reason why salt is sprayed on snow to melt it: the freezing
point is lowered by the addition of salt). Equation (8.8.7) is obtained again, with
the limits of integration reversed, and with DH being of course the latent heat
of fusion.
Finally, we consider the case of osmotic pressure. The pure solvent fugacity
f O increases with pressure at fixed temperature, and one can ask at what pressure
the fugacity of the solvent in the solution will equal that of the pure solvent at
pressure p; the difference is called osmotic pressure, pOs. A straightforward
calculation for the case where the liquid is incompressible yields

(8.8.lt)

Osmotic pressure is often introduced by considering semipermeable mem-


branes which allow only the solvent to permeate. The membrane is thought of
as separating the solution from the pure solvent, and the osmotic pressure is then
the pressure difference across the membrane at equilibrium. If indeed truly
semipermeable membranes could be found, osmotic pressure could in fact be
measured in such a way. The definition given above for pOs avoids the issue of
whether truly semipermeable membranes actually exist. Membranes which
approach semipermeability do certainly exist, and if two solutions with a common
solvent are on the two sides of such a membrane, the solvent will tend to permeate
in whatever direction is needed in order to equalize the osmotic pressure on the
two sides. These types of phenomena playa major role in biological systems.
The theory of colligative properties can be used for three different aims.
First, suppose a colligative property is measured for a system for which Y is
known (Le., the amount of solutes and their molecular weights are known). In
this case, experiments are used to validate the theory, Le., to ascertain that indeed
equation (8.8.6) holds in the first-order approximation. Second, suppose the mass
of a solute in the mixture is known, but not its molecular weight. In this case,
the measurement of a colligative property can be used to infer the molecular
weight of the solute. Finally, suppose that the molecular weight of a solute is
known, but the measurement of the colligative property does not give the expected
value. One can then infer (to within belief in the theory) the effective molecular
240 weight of the solute in the solution; this may differ from its molecular weight in
the pure form because either oligomerization or dissociation of the solute occurs
Chapter Eight
when it is dissolved in the solvent. Indeed, this is the experimental basis originally
responsible for the formulation of the theory of dissociation of electrolytes in
aqueous solutions. (Upon dissolution of, e.g., NaCI in water, the value of Y
calculated from the colligative properties measured is almost exactly twice what
one would expect from the molecular weight of NaCI; this suggests that NaCI
dissociates completely into two ions. In the case of CaCI 2 , the ratio is almost
exactly 3, suggesting that it dissociates into three ions; and so on.)

EXAMPLES AND PROBLEMS

Problems

8.1. Consider a dilute solution for which the gas and solid phases consist of the pure solvent,
with the solutes present only in the liquid phase. Draw a pressure-temperature phase diagram for
the case where also the liquid phase consists of the pure solvent, and for the case of the solution,
and infer graphically the boiling point increase and the freezing temperature decrease. For the same
system, draw a chemical potential vs. temperature diagram at a fixed pressure, and again infer the
two colligative properties from such a plot.

8.2. Consider the gas represented by (a) the virial equation of state, (b) the Van der Waals
equation of state, and (c) the corresponding states equation. Which, if any, are systems with entropic
elasticity? Which assumptions did you make in getting your answer, and can some of these be proved?
Does the fact that gases usually behave as ideal mixtures support your conclusion?

8.3. Consider the quantity g(T) defined in equation (8.1.17). Is the sign of the derivative dgldT
fixed, and if so, what is it? Does the result hinge on the application of the third law?

SA. Does the corresponding states equation as defined in Section 8.6 imply that the quantity
Pc Vcl RTc is a universal constant?

8.5. Write the appropriate second-order equation corresponding to the expansion in equation
(8.7.12).
8.6. Can equation (8.7.12) predict phase separation? If it can, what constraints must the
parameters A and B satisfy in order to have a spinodal region?
8.7. Calculate the mole fraction of styrene in a 1% by weight solution of styrene in polystyrene.
The polystyrene has a molecular weight of 2 million. Do you think it realistic that Henry's law, or
some modification thereof, could be used to describe the solubility of styrene vapor in polystyrene?
If you do, write down the modification that you think is realistic.
8.8. If you add salt to water before it boils, will the time needed to bring it to boiling be more
than, less than, or equal to the time needed without salt? Give an estimate of the percentage difference
if you think there is one.

8.9. Calculate the osmotic pressure of seawater. Do you think it realistic that it could actually
be measured by an equilibrium experiment, even if a truly semipermeable membrane could be found?
How would you set up the experiment?
LITERATURE 241
Equations of
Mathematics is an obscure field, an abstruse science,
complicated and exact; yet so many have attained peifection in State
it that we might conclude almost anyone who seriously applied
himself would achieve a measure of success.
Among [the Greeks] geometry was held in highest honour;
nothing was more glOriOUS than mathematics. But we have
limited the usefulness of this art to measuring and calculating.
Cicero

The best presentation of Nemst's argument on the third law of thermodynamics is


in his book, The New Heat Theorem, published originally in German in 1917; the first
English translation was published by Methuen and Co. in London in 1926. A good
discussion of the possible experimental validations of the third law is given by Taylor and
Glasstone, Physical Chemistry, Chapman and Long, London (1953).
That the velocity of sound is (ap/a<f>}'/2 was deduced as early as 1687 by Newton,
Philosophiae Naturalis Principia Mathematica, J. Straeter, London (1687). Newton used
(in the first edition of the Principia) the isothermal value of ap/a<f>, thus underestimating
the speed of sound in ambient air by about 30%. The issue was resolved in 1822 by
Laplace [Oeuvres, Imprimerie Royale, Paris (1846)] who, in order to obtain the correct
result, simply made use of the known fact that the pressure of air, when compressed very
rapidly (and hence adiabatically), is proportional to the 5/3 power of density. Since at
those times there was no thermodynamic theory whatsoever available, the result is clearly
a purely mechanical one, with no thermodynamic content at all. The thermodynamic
analysis (i.e., the application of the first law to the phenomenon considered) was developed
by W. J. M. Rankine, Phil. Trans. R. Soc. London 160,277 (1870), and more completely
by H. Hugoniot, J. Ecole Polito 57, 3 (1887); 58, 1 (1888); 1. Math. Pures Appl. 3, 477
(1887); 4, 153 (1887). The whole matter is discussed in detail in any textbook on gas
dynamics. Propagation of discontinuities in elastic systems is discussed in general by C.
Truesdell and R. A. Toupin, "The Classical Field Theories," in: Encyclopaedia of Physics,
Vol. 3/1, Springer-Verlag, Berlin (1960); this theory is still a thermodynamically reversible
one. For the more general case of dissipative systems, see B. D. Coleman et ai, Wave
Propagation in Dissipative Materials, Springer-Verlag, Berlin (1965).
The statement in Section 8.3 that no gas-gas phase transitions are known to occur
should be qualified. In the case of a shock wave through which a combustion reaction
proceeds to completion, a surface of discontinuity exists which separates two gases which
are at different pressure, temperature, and composition, and hence should, by all reasonable
definitions, be recognized as two phases. However, this is not an equilibrium condition.
The existence of interfaces between phases sustained by a nonequilibrium phenomenon
is discussed by G. Astarita and R. Ocone, Adv. Chem. Eng. (to be published).
The Van der Waals analysis has been discussed in the literature section of Chapter
4. The idea that parameter a may be a decreasing function of temperature was considered
as long ago as 1880 by Clausius himself, and a simpler version of Clausius' result was
given by Berthelot in 1900; the Redlich-Kwong form is in o. Redlich and J. N. S. Kwong,
Chem. Rev. 44, 233 (1949). Several later variations are due to Wilson; Soave; Lee, Erbar,
and Edminster; Peng and Robinson; and Tang. The latter in Huagong Xuebao 2, 149
(1984) [English Transl.: Int. Chem. Eng. 27, 148 (1987)] gives a general overview of
equations of state for single-component systems of the Van der Waals type.
There is a very ample literature on constitutive equations for nonideal mixtures. A
good general presentation is given by J. H. Hildebrand, J. M. Prausnitz, and R. L. Scott,
Regular and Related Solutions, Van Nostrand, New York (1970); Perry's handbook is also
a good, if very concise source.
242 Constitutive equations for nonideal mixtures are best written as equations giving G EX
as a function of composition. Equation (8.4.14) then produces equations for the activity
Chapter Eight coefficients which automatically satisfy the Gibbs-Duhem equation, as well as the so-called
Gibbs-Helmholtz equation, i.e., 8(G Ex/T)/8(1/T) = HEX. The generality of this pro-
cedure is discussed by K. Wohl, Trans. Am. Inst. Chem. Eng. 42, 215 (1946).
Special constitutive equations which have been commonly used are in G. M. Wilson,
1. Am. Chem. Soc. 86, 127 (1964). The two-parameter Wilson equation has only one major
disadvantage, namely, its inability to predict liquid-liquid phase separation. This is
overcome by the nonrandom, two-liquid (NRTL) model; see H. Renon and J. M. Prausnitz,
Ind. Eng. Chem. 57(5), 18 (1965). Other equations which have been used extensively are
discussed in 1. Am. Chem. Soc. 57, 1805 (1935); Ind. Eng. Chem. 40, 341 (1935); Ind. Eng.
Chem. SO, 391 (1958); Ind. Eng. Chem. 51, 211 (1959); AIChE 1. 5, 249 (1959).
A model which has been used extensively is the UNIQUAC (Universal Quasi Chemical
Theory); for an introduction to the basic concepts of UNIQUAC, see D. S. Abrams and
J. M. Prausnitz, AIChE J. 21, 116 (1975).
Chapter Nine

PHASE EQUILIBRIA

One could even conceive the bodies A and B such that


they would remain themselves at a constant temperature
though losing or gaining quantities of heat. If, for
example, the body A were a mass of vapor ready to
condense and the body B a mass of ice ready to melt,
these bodies, as is well known, could give out or receive
caloric without changing their temperature.
s. Carnot, 1824

NOTATION

A Constant in Langmuir equation atm- 1


B Henry's constant atm
b Vapor pressure label of components atm
C Curvature of dimensionless free enthalpy curve
C Constant in BET equation
Cp Specific heat kcal kmor 1 K- 1
D Difference operator
F( Vapor pressure function atm
f Fugacity atm
/;( Enthalpy functions kcal kmol- 1
G Free enthalpy kcal kmol- 1
gEX Dimensionless excess free enthalpy
g{ } Constitutive functional for gEX kcalkmol- 1
g'{ } Functional derivative of g{ }
K Partition coefficient
M Molecular weight kgkmol- 1
M' Second moment of mole fraction distribution atm 243
244 M" Defined in equation (9.5.12) atm
Chapter Nine n Total number of moles kmol
n(u) Number of moles distribution kmol
N Avogadro's number molecules kmor l
p Pressure atm
P' Capillary pressure atm
Q Average vapor pressure atm
R Vapor pressure ratio
r Radius of embryo m
S Entropy kcal kmol- I K- I
s Surface tension matm
T Temperature K
To Critical temperature K
TE Three-phase temperature K
U, u' Component labels
V Specific volume m3 kmol- 1
VL Liquid specific volume m3 kmor l
V. Total volume m3
v Fugacity coefficient
w Phase label
x Weight fraction of sorbate
XM Value of x for monomolecular layer
y Mol fraction
y(b) Mol fraction distribution atm- I
y(u) Mol fraction distribution
Yo Value of y at three-phase equilibrium
a Relative volatility
aK Fraction of phase K
a( Phase fraction distribution
{3 Pressure correction coefficient
T Activity coefficient
T( Activity coefficient distribution
TD Dilute solution activity coefficient
T{ } Constitutive functional for T(
r( ) Gamma function
8( ) Boiling temperature function K
1 Chemical potential kcalkmol- I
1( Chemical potential distribution kcal kmol- I
Specific surface area of solid m 2 kg- 1 245
u Surface area of one molecule m2 molecule-I Phase Equilibria
Parameter of distribution

Subscripts

1,2 For component 1,2


AVG Average for the system
j For component j
K Of phase K
o At phase transition
S At special phase transition

Superscripts

I, II For phase I, II
* At equilibrium
EX In excess of ideal mixture
G For gas phase
ID For ideal mixture
L For liquid phase
S At saturation
9.1. GAS-LIQUID EQUILIBRIA

Thermodynamics gives me two strong impressions: first of a


subject not yet complete or at least of one whose ultimate
possibilities have not yet been explored, so that perhaps there
may still be further generalizations awaiting discovery; and
secondly and even more strongly as a subject whose
fundamental and elementary operations have never been
subject to adequate analysis.
P. W. Bridgman

In one-component systems, gas-liquid equilibria simply occur when the average


volume of the system is intermediate between those of the saturated liquid and
the saturated vapor at the temperature T of the system. Correspondingly, the
pressure is equal to the vapor pressure p' at temperature T.
As one moves up to mixtures, the simplest possible case is that where the
liquid phase may be regarded as a pure component A (this implies that all other
components have negligibly small mole fractions in the liquid phase). The gas
phase is however a mixture. In this case, at equilibrium the fugacity of A in the
gas phase is that of the saturated liquid, for instance

/ = p'v s exp[ VL(p - pS)/ RT] = pv s(3 (9.1.1)

where v' is the fugacity coefficient of A at a pressure equal to p', and the fugacity
correction (3 for pressure in the liquid has been given based on the assumption
of incompressibility of the liquid phase. If the pressure is low enough for the
gas mixture to be ideal (a much milder requirement than the condition that the
fugacity coefficients should be unity), one can write

/=pvy (9.1.2)

where v is the fugacity coefficient of A at pressure p, and y is the mole fraction


of A in the gas phase. Elimination of/between equations (9.1.1) and (9.1.2) yields

y = p'v'(3/pv (9.1.3)

If all fugacity coefficients can be taken as unity, and if the pressure correction
for the liquid phase is neglected, this reduces to the classical elementary result

y = pS/p (9.1.4) 247


248 Let us now consider the phenomenon of boiling. Of course, if a system is
Chapter Nine
boiling it is not at equilibrium; however, by considering it to be at equilibrium,
one can establish conditions for the existence of a driving force for boiling.
Although the gas above the boiling liquid is a mixture, the gas in the bubbles is
pure A, and hence the liquid-bubble eqUilibrium is a one-component setup and
the equilibrium condition is that pressure should be equal to pS. Thus, indepen-
dently of what the fugacity coefficients may be, the boiling point condition is
p = pS. Often the boiling point is expressed in terms of a boiling temperature,
say the temperature at which the vapor pressure equals the total pressure on the
system. The pressure of the liquid-bubble system is in principle larger than the
pressure in the gas phase, due to the hydrostatic pressure head.
In actual fact, the phenomenon of boiling requires nucleation of the vapor
bubbles. If a very small vapor bubble is nucleated, the pressure inside it will be
larger than that in the surrounding liquid, due to the capillary pressure h. It
follows that the condition for growth of a nucleated bubble is that the temperature
should be high enough for pS to be p + h. It is true that P:t becomes rapidly
negligible as the bubble size increases. However, nucleation does initially result
in very small bubbles, and therefore significant liquid overheating above the
boiling temperature will be observed before the actual boiling phenomenon starts.
Conversely, we consider the case where the mixture is initially all in the gas
phase and at equilibrium; again only component A can be regarded as present
in any liquid phase which may form. The condensation condition, i.e., the
threshold at which a driving force for formation of liquid drops begins to exist,
is now (neglecting capillarity) equation (9.1.3), and the temperature at which this
equation is satisfied (usually called the dew point) will be equal to that at which
the vapor pressure equals the partial pressure in the gas phase only if fugacity
coefficients are taken to be unity and the pressure correction for the liquid phase
is neglected. Even under these conditions, significant undercooling of the gas
phase may occur as a consequence of capillary phenomena, as discussed in the
Appendix to this section.
Let us now tum our attention to mixtures, where all components are present
in non-negligible amounts in both phases. The superscripts Land G will be used
to identify the two phases. For each component j, the fugacity in the liquid phase
is given by

(9.1.5)

If the gas phase is an ideal mixture, the gas phase fugacity is given by
equation (9.1.2), and therefore the equilibrium condition in the absence of
capillary effects can be written as

(9.1.6)

Several types of equilibrium problems can be envisaged. The simplest is


when temperature and composition of the liquid phase are given. In this case
equation (9.1.6), together with the condition that quantities Y7
should add up to
unity, yields almost directly the composition of the gas phase and the pressure.
More complicated cases arise when, of the four variables (pressure, temperature,
and composition of both phases), the two given ones are other than temperature 249
and liquid phase composition.
Phase Equilibria
Of the several correction coefficients cluttering equation (9.1.6), most impor-
tant are the activity coefficients, since they are often nonunity even at moderate
pressures. Therefore, subsequent attention is restricted to the simpler case where
all fugacity coefficients can be taken as unity, and the pressure correction in the
liquid phase is negligible, so that equation (9.1.6) reduces to

(9.1.7)

First we consider the simplest possible case where the liquid phase is an
ideal solution, so that equation (9.1.7) reduces to Raoult's law:

pyy = pSy; (9.1.8)

The boiling point of the liquid mixture at any given pressure p is determined
by the requirement that equation (9.1.8) must be satisfied for all components
with quantities yy summing to unity, and the dew point by the requirement that
it should be satisfied with all the quantitiesy;summing to unity. The dew points
and boiling points of the mixture are always intermediate between those of the
highest boiling and lowest boiling pure components. A plot of dew and boiling
points for an ideal binary mixture is presented in Figure 9.1.1. Given an average
composition of the mixture such as y, and a temperature T, such that the point
y, T lies within the lense-shaped region, the equilibrium condition will be a
two-phase one, with the liquid- and gas-phase mole fractions being determined
by the intersection of the horizontal through T with the dew and boiling point
curves, and the relative amount of the two phases being given by the usual lever
rule applied to that horizontal segment. This is true also for nonideal solutions,
where the two-phase envelope may have a shape quite different from the simple
lense in Figure 9.1.1.
Next let us consider the case of a binary nonideal mixture. It is useful to
define a relative volatility a as follows:

(9.1.9)

FIGURE 9.1.1. Dew point and boiling point plots


for a binary mixture obeying Raoult's law. o y
250 which, for the case of an ideal solution, is simply given by
Chapter Nine
(9.1.10)

and is thus a weak function of temperature. Since temperature at two-phase


eqUilibrium for an ideal mixture ranges between the boiling points of the two
components, if the latter are not very different a will be either larger or smaller
than unity over the whole range. This is not, however, necessarily the case for
nonideal mixtures.
Indeed, for a nonideal mixture, a is given by

(9.1.11)

and it may well change from larger to smaller than unity if a liquid-phase
composition exists for which

a=l (9.1.2)

A liquid-phase composition for which equation (9.1.12) holds is called an


azeotrope; the gas-phase composition in eqUilibrium with an azeotrope is the
same as the liquid-phase one. Which component is at eqUilibrium more abundant
in the gas phase than in the liquid phase depends on which side of the azeotropic
composition the liquid phase is situated. Distillation of a binary mixture can
never surpass an azeotropic point from either side, since the driving force for
liquid-gas composition change vanishes at an azeotropic point.
Equation (9.1.12) can be satisfied at some particular composition with either
positive or negative deviations from ideality (i.e., if the quantities 'T' are either
larger or smaller than unity). If deviations from ideality are positive, the
azeotrope's temperature will be higher than the boiling point of either component
.(this is called a high-boiling azeotrope), and vice versa with negative deviations
from ideality (see the Appendix in this regard). Figure 9.1.2 is a sketch of a dew
point and a boiling point diagram for a binary mixture with a high-boiling
azeotrope.
There is an intrinsic difficulty with equation (9.1.5) which is similar to that
discussed in Section 8.4. We begin by considering the simple case of a binary
mixture for which Raoult's law applies, i.e., the liquid mixture is ideal. At

FIGURE 9.1.2. Dew point and boiling point plots


for a binary mixture exhibiting a high-boiling
o azeotrope.
two-phase equilibrium, the temperature is intermediate between the boiling tem- 251
peratures of the two components, and thus in particular is higher than that of
Phase Equilibria
the low boiling one. Thus, the pressure correction term implies consideration of
the pure-liquid specific volume at pressures where the pure component would in
fact exist as a gas at equilibrium. This essentially implies extrapolation of liquid
volume data beyond the equilibrium region of pressures. As far as the liquid is
almost incompressible, this does not pose much of a problem.
A more serious problem arises if the temperature of the two-phase system
is above the critical temperature of one of the components, or below the triple-
point temperature of one of them, as it may well be. In fact, in such cases the
vapor pressure p* is undefined, unless the p* vs. T curve is extrapolated (some-
what arbitrarily) beyond the range over which it is defined. The arbitrariness of
the extrapolation is compensated by the fact that the activity coefficient is in turn
defined with respect to such an arbitrary extrapolation.
If the temperature of the two-phase mixture is very significantly above the
critical temperature of say component A, the Henry's law formalism discussed
in Section 8.5 is preferable. In fact, in such cases, the liquid-phase mole fraction
of A is likely to be quite small. If the liquid-phase fugacity of A is written as in
equation (8.5.6), its eqUilibrium condition becomes

(9.1.3)

If the mole fraction of A in the liquid phase is small enough so that TD = I,


one obtains

(9.1.14)

Often, the quantity in parentheses is called the Henry's law constant. The
distinction becomes irrelevant if the pressure is low enough so that v = 1.

Appendix

Consider a gas phase containing a condensable component which is super-


cooled, i.e., the chemical potential of the condensable component in the gas
phase, p'0, is larger than the chemical potential of the same component in the
form of a pure liquid, p. L. By Brownian motion, aggregations of molecules of
the condensable component may form; these are called embryos. For simplicity,
we assume that embryos are of spherical shape, so that the capillary pressure is
given by

h = 2s/r (9.1.A.l)

where r is the radius of the embryo. The chemical potential within the embryo
is thus

(9.1.A.Z)
252 The number of moles in the embryo is ~1Tr3I VL The total free enthalpy of
Chapter Nine
the embryo, including the surface excess free enthalpy, is thus

Gt = ~1Tr3(I-'L/VL + 2slr) +4s1Tr2


= ~1Tr3 I-' L + 1Tr 2 (9.1.A.3)

The same amount of condensable component in the gas phase has a total
free enthalpy ~1Tr3 I-' G I VL , and hence the driving force for growth of the embryo
is

(9.1.AA)

The embryo will tend to grow, provided the right-hand side of equation
(9.l.A.4) decreases with increasing radius. We note, however, that at small values
of r the negative second term will predominate, and therefore embryos formed
by Brownian motion will tend to redisperse again in the gas phase. The critical
embryo radius is that for which the derivative of 8Gt with respect to r is zero:

r = (3 VLsl 10)(1-' G _ I-' L)

= (3 VL s/10)[RTln(plp')] (9.1.A.5)

where p is the partial pressure of the condensable component in the gas phase.
An estimate of the actual value of pip needed for significant growth. of
embryos can be obtained by absolute rate theory, with the relevant energy being
the value of 8G, corresponding to the r value in equation (9.1.A.5). The rate of
growth thus calculated exhibits almost a sharp discontinuity at a critical value
of pip" which may be as large as 5 for growth of water embryos in cool
atmospheric air.
The above discussion is concerned with homogeneous nucleation. In actual
fact, nucleation occurs preferentially on seeds, which in the case of atmospheric
formation of fogs and clouds are microscopic salt grains. If one considers a very
small drop of water containing the original salt seed, two opposing effects take
place. On the one side, as r is increased, capillary effects have the same behavior
as discussed above. On the other side, as r grows the mole fraction of water in
the embryo increases, and thus its equilibrium vapor pressure increases. Indeed,
at very small radii the latter effect is predominant, so that the chemical potential
of water in the concentrated salt solution corresponding to the original embryo
is very low. This is the reason why nucleation on salt seeds predominates over
homogeneous nucleation.
Let us consider a binary nonideal mixture which exhibits an azeotrope at
some fixed pressure p. Gas-liquid equilibrium occurs when

(9.1.A.6)

If this equation is differentiated with respect to Yl, we obtain after rearrangement

(9.1.A.7)
which, in view of equation (9.1.12), shows that the boiling temperature of the 253
azeotrope is an extremum. The second differentiation shows that it is a maximum Phase Equilibria
for positive deviations from ideality, and vice versa. The argument is easily
generalized to more than two components.

9.2. LIQUID-LIQUID EQUILIBRIA

Out of whose womb came the ice? And the hoary frost of
heaven, who hath gendered it? The waters are hid as with a
stone, and the face of the deep is frozen.
Job 38, 29-30

First we consider the case of binary mixtures. As was discussed in Section


8.7, if two components form an ideal mixture, that mixture will never exhibit a
miscibility gap, and hence the question of liquid-liquid equilibria does not arise.
Mixtures with negative deviations from ideality (i.e., mixtures for which G EX is
negative), while they could in principle exhibit miscibility gaps, in fact practically
never do. So attention will be focused on systems with positive deviations from
ideality.
In binary systems, a miscibility gap will occur if there exists a spinodal
region where the curvature of the G(Yt) curve is negative. In the case where gEX
is given by equation (8.7.1), the condition for the existence of a miscibility gap
is given by equation (8.7.5). In the case of more complex constitutive equations
for gEX, the condition is more complex, and it can always be determined by
considering that e ID is given by equation (8.7.4). The actual value of e is given
by

(9.2.1)

and thus the existence of a spinodal region can be ascertained. It is noteworthy


that, as any ofthe mole fractions approaches unity, the right-hand side of equation
(9.2.1) approaches some constant value, while the first term approaches infinity.
This naturally implies that complete immiscibility can never occur, though in
some cases the miscibility gap is so extended that it can in practice be regarded
as complete immiscibility.
Binary mixtures may exhibit a miscibility gap only over a limited range of
temperatures; systems may be miscible both above and below some critical
temperature. A T vs. Yt plot for a mixture exhibiting an upper critical solution
temperature T. is sketched in Figure 9.2.1. The continuous line is the locus of
compositions of two phases in eqUilibrium with each other; the dashed curve is
the locus of spinodal points. The two curves are tangential at the critical solution
temperature and composition Y. If a one-phase system at a composition other
than Y. is cooled down, it will reach the two-phase envelope outside the spinodal
region, and hence it will tend to unmix into two phases by the phenomenon of
nucleation and growth of the other phase's droplets. If, however, a one-phase
solution of composition Y. is cooled down, it will enter the spinodal region, and
homogeneous nucleation may take place.
254 Let us now turn our attention to ternary mixtures. The composition space
is two-dimensional, and can be represented in the usual triangular diagram form
Chapter Nine
such as in Figure 9.2.2. The free enthalpy is a surface over the composition space,
and two-phase equilibrium conditions will appear if the surface has a spinodal
region, i.e., a region where its curvature is negative. The simplest case where a
miscibility gap occurs is represented in Figure 9.2.2a, where two of the component
pairs are miscible in all proportions, while the third presents a miscibility gap;
this extends into the three-component region as shown. Correspondingly, there
exists a spinodal region bounded by the dashed curve, which is tangent to the
miscibility gap curve at D. Two phases at equilibrium are represented by two
points on the miscibility gap envelope connected by a tie line. One of the tie
lines corresponds to the AlB binary mixture; at the other extreme, point D
represents a degenerate tie line. The curves AD and BD are the phase-I and
phase-II curves. Neglecting pressure corrections for the chemical potential, the
equilibrium condition for component C is
(9.2.2)

As long as the mole fraction of C is small in both phases, the activity


coefficients can be approximated by their infinite dilution values, and the equili-
brium condition becomes linear:

(9.2.3)

where K is called the partition coefficient.


More complicated miscibility behaviors are represented in the other plots.
In particular, Figure 9.2.2c represents a case where all three component binary
mixtures are miscible in all proportions, but the ternary mixture is not; this is
clearly possible only if ternary interactions are significant. Figure 9.2.2d represents
a case where all three component binary mixtures exhibit a miscibility gap, while
the one-phase regions are restricted to neighborhoods of the three corners. The
triangular region in the center is a three-phase one.
An interesting point concerning Figure 9.2.2d is the following. To simplify
the discussion, let us assume that equation (8.7.13) holds. Since all three com-
ponent pairs exhibit a miscibility gap, all three coefficients are larger than 2. Now
suppose the system is constrained to the vertical line bisecting the triangular

FIGURE 9.2.1. Miscibility gap at different tem-


peratures for a binary mixture exhibiting an upper
critical miscibility temperature.
diagram (d). This can be accomplished in practice if components A and Bare 255
copolymerized in stoichiometric 1: 1 proportions. Hence Phase Equilibria

YA = Yo = (1 - Yd/2 (9.2.4)

The curvature of the ideal-solution free enthalpy curve along the vertical
line is

e lD = (1 + Yd/[Yd1 - yd] (9.2.5)

and has a minimum of 5.828 at Yc = .J2 - 1. The curvature of the excess free
enthalpy curve along the vertical line is

(9.2.6)

and thus it may well be positive or, if negative, its absolute value may well be
less than 5.828, so that no miscibility gap will occur. Indeed, the stronger the
immiscibility between A and B, the more likely it is that no miscibility gap will
be observed along the vertical line. This point is very important for realizing
miscible polymer blends, since polymers very often exhibit strong immiscibility;
however, if the least miscible pair is copolymerized, it becomes miscible with the
third polymer.
The algebra becomes rapidly very cumbersome as the number of components
is increased, and hence a continuous description of the type discussed in Section
9.5 becomes useful. For a three-component mixture, the curvature of the excess
free enthalpy surface over the Yl-Y2 plane is given (to within a multiplicative
constant) by

(9.2.7)

c c

a B A B
b

A
C

FIGURE 9.2.2. Ternary diagrams


for systems exhibiting miscibility
gaps. A c B A B
d
256 and therefore in an ideal mixture
Chapter Nine
(9.2.8)

which is always positive, with a minimum value of 15 when all three mole fractions
equal ~.

9.3. MULTIPHASE SYSTEMS

Falsifiability marks the distinction between, on the one hand,


statements that belong in science and to the world of common
sense, and on the other hand statements which, though they
belong to some other world of discourse, are not to be
dismissed contemptuously as nonsense. Metaphysics is a
compost that can nourish the growth of scientific ideas. But if
we accept falsifiability as a line of demarcation, we obviously
cannot accept into science any system of thought (for example,
psychoanalysis) which contains a built-in antidote to disbelief:
to discredit psychoanalysis is an aberration of thought which
calls for psychoanalytical treatment.
P. B. Medawar

In this section, cases are considered where, at equilibrium, there may be


more than two phases present. Let us first consider a binary mixture which exhibits
a miscibility gap in the liquid phase. Here, / and y" are the mole fractions of
component 1 in the two liquid phases at equilibrium with each other (labeling
of the two phases having been chosen so that / > /'), y the mole fraction of
component 1 in the gas phase, and YAva the average mole fraction of component
1 in the system as a whole. For the sake of simplicity, pressure corrections to the
fugacity in the liquid phase are neglected, and the gas phase is regarded as ideal.
The conclusions reached hold qualitatively also for the case where the quantities
f3 and v differ from unity.
Consider the plot of Y vs. YAva in Figure 9.3.1; the plot is at an assigned
pressure, and coherently temperature is variable. The diagram exhibits a region
where the gas-liquid equilibrium line is horizontal; this corresponds to values
of YAva in the interval Y' - y". In this interval, there are either two or no liquid
phases present. When there are two liquid phases, a change in YAVO corresponds

0"----'-_ _ _--'_--' FIGURE 9.3.1. Phase diagram for a system exhibit-


ing a miscibility gap in the liquid phase and an
o "azeotrope."
to a change in the relative amount of the two liquid phases, but their composition 257
remains constant. Consequently the gas phase, which is in equilibrium with both Phase Equilibria
liquid phases, has a constant composition Y = Yo. Under the simplifying assump-
tions made, the equilibrium condition is

(9.3.1)

where the second equality is guaranteed to hold because of equation 9.2.2.


The equilibrium condition for component 2 is

(9.3.2)

It is evident that equations (9.3.1) and (9.3.2) can be satisfied simultaneously


at only one temperature, say TE , and therefore temperature along the horizontal
branch of the gas-liquid equilibrium line is constant. For YAVG values outside
the miscibility gap interval, only one liquid phase may form.
For the plot as given in Figure 9.3.1, the equilibrium line crosses the main
diagonal, and in a restricted sense one could say that there is an azeotrope at
YAVG = Yo It is, however, an azeotrope only in a formal sense, since only the
average composition of the two liquid phases is equal to that of the gas phase.
It is possible to have systems exhibiting a miscibility gap for which the gas-liquid
equilibrium line does not cross the main diagonal; see Figure 9.3.2.
Now we consider the dew point-boiling point plot for a system which exhibits
a three-phase "azeotrope." The behavior is sketched in Figure 9.3.3. The tem-
perature of three-phase equilibrium, TE , is lower than the boiling point of both
components. Only if the composition of the gas is Yo will the dew point result
directly in two liquid phases; at any other composition, condensation from the
gas phase will initially result in only one liquid phase. The two lines separating
the liquid-phase region are only approximately vertical, since the extent of the
miscibility gap varies with varying temperature.
If the miscibility gap is very large, the two liquid phases at equilibrium with
each other can both be regarded as dilute solutions. With this equations (9.3.1)
and (9.3.2) simplify to:

PYo = P~y' (9.3.3)

FIGURE 9.3.2. Phase diagram for a system exhibit- o "--.L....-_-'-_ _ _---'


ing a miscibility gap in the liquid phase but no
"azeotrope. "
o
258 and
Chapter Nine (9.3.4)

in the first-order approximation (Le., with the activity coefficient of the solvent
taken to be unity). The value of TE is determined by the following condition:

(9.3.5)

In the zero-order approximation (Le., yl "" 1 and 1 - yll "" 1) this reduces
to the condition that the sum of the vapor pressures of the two components
equals the total pressure.
It is of interest to determine the conditions under which TE is lower than
the boiling temperature of both components. Since the vapor pressure is a
monotonously increasing function of temperature, it can be inverted to yield a
function (J:

pS = P(T); T = p-.(pS) = lJ(pS) (9.3.6)

and the boiling point of the two components are simply 1J.(p) and 1J2 (p). If R
is the vapor pressure ratio p~/ p~, with the labeling of components chosen so that
R> 1, and thus (J.(p) < (J2(P), then TE is obtained from equation (9.3.5) in the
form

(9.3.7)

Since IJ is a monotonously increasing function, the condition is

(9.3.8)

which is, of course, more easily satisfied the larger the miscibility gap.
The essential feature of the phenomenon discussed above is the following.
When a large miscibility gap exists in the liquid phase, the fugacity of component
1 in phase I (as well as that of component 2 in phase II) increases rapidly with
temperature because so does the vapor pressure; the effect of temperature on the
activity coefficients and on the fugacity coefficients of the saturated vapor is much
weaker. In the gas phase, the fugacity of each component decreases with decreas-
ing mol fraction, and is at most a weak function of temperature. It follows that

I
lI+I1 II
I
I
FIGURE 9.3.3. Phase diagram for a system
exhibiting a miscibility gap in the liquid phase.
T
259
Tt.11
Phase Equilibria

II

FIGURE 9.3.4. Liquid-solid phase diagram for


o --L
a system exhibiting a eutectic. Eutectic

equilibrium of either liquid phase with the gaseous mixture is reached at a


temperature lower than that at which equilibrium with a pure gas phase would
be reached.
Now let us consider the solid-liquid equilibrium behavior of a binary mixture.
Solid phases generally exhibit very large miscibility gaps, so that in fact the
zero-order approximation is justified. The typical phase diagram is of the type
shown in Figure 9.3.4, which is exactly analogous to that in Figure 9.3.3. In the
case of solid-liquid equilibria, TE and Yo are called the eutectic temperature and
composition.
In this case, the fugacity of both solid phases is the fugacity of the saturated
vapor at the sublimation point, since the temperature is lower than the melting
temperature of the pure component; see Figure 9.3.5 (for the purposes of the
present discussion, the melting curve can be regarded as vertical, and the pressure
correction to the fugacity of the solid is neglected). The fugacity of, say, com-
ponent 1 in the liquid mixture is given by

(9.3.9)

where p. is the vapor pressure of the liquid-vapor phase transition, to be read


from the extrapolation of the equilibrium curve into the low-temperature range
(dashed line in Figure 9.3.5). The temperature dependence of fL is dominated
by that of p., which is weaker than that of the sublimation vapor pressure. Hence
as temperature is decreased, freezing will occur at a temperature lower than the
single-component melting temperature.

FIGURE 9.3.5. Vapor pressure of the solid at the --L p


Eutectic
eutectic temperature. Temperature
260 9.4. GAS-SOLID EQUILIBRIA
Chapter Nine
Our next subject must be the ores of iron, a metal which is at
once the best and the worst servant of humanity, for to bring
death more speedily to our fellow men, we have given wings to
iron and taught it to fly.
Pliny, Natural History

The phenomenon of adsorption of gases on solid surfaces is well known.


The adsorbed layer can, in a restricted sense, be regarded as somewhat analogous
to a liquid phase. An important distinction to be made is between the cases where
the temperature is higher or lower than the critical temperature of the adsorbing
component, Te. If T> Te, no liquid phase of the adsorbing component could
form whatever its pressure in the gas phase. Thus if indeed a liquid-like layer is
formed on the surface of the solid, it must be due entirely to the distorting effect
of the solid surface, and therefore it has to be of molecular thickness. If T < Te ,
the phenomenon of adsorption needs to be considered only for gas-phase partial
pressuresp lower than the vapor pressure p'; when p = p', an actual liquid phase
will form at equilibrium.
Adsorption equilibrium data are generally presented as constant-temperature
plots of p vs. the grams of adsorbate per gram of solid, x. Such plots are called
adsorption isotherms. Five basic shapes of adsorption isotherms have been
observed; they are shown schematically in Figure 9.4.1. The adsorption process
is always an exothermic process, i.e., at a given p the value of x decreases with
increasing temperature. The heat of adsorption (which can be measured directly,

x x

x x

FIGURE 9.4.1. Different types of


p adsorption isotherms.
or can be obtained from an Arrhenius plot of In x vs. 1/ T) may vary very 261
considerably with varying x. Often the adsorption isotherm exhibits hysteresis, Phase Equilibria
Le., the curve as measured during adsorption is different from that measured
during desorption; of course only one of the two branches of a hysteresis loop
(and possibly neither one) may represent actual equilibrium.
A "molecular monolayer" capacity XM can be defined. This corresponds to
the value of x which would be obtained with a surface layer of adsorbate one
molecular layer thick. If l is the surface area of the solid per unit mass, u the
area occupied by one molecule on the surface, N is Avogadro's number, and M
is the molecular weight of the sorbate, then XM is given by

(9.4.1)

A distinction can be made between "mobile" and "fixed" surface layers. In


the first case, adsorbed molecules can be anywhere on the surface of the solid,
and hence u is expected to depend on the size of the sorbate molecule and
possibly on its orientation on the surface; the latter may depend on the type of
solid. In the second case, the sorbate molecules are fixed to specific sites of the
surface lattice, and u is thus expected to depend only on the properties of the
solid itself.
Adsorption isotherms of type I (which is the only type exhibited at T > TO>
are generally well correlated by the Langmuir isotherm equation:

X/XM = Ap/(l + Ap) (9.4.2)

where A may be regarded as an adjustable parameter the value of which can be


estimated from the appropriate value of which can be estimated from the appropri-
ate molecular theory. If XM in equation (9.4.2) is interpreted as the true monolayer
capacity, equation (9.4.2) implies that at no pressure is the adsorbed layer thicker
than one molecule, which is indeed the molecular assumption on which the
Langmuir equation is based.
Type II and type III isotherms can be correlated by the BET equation, which
in its molecular formulation considers the possibility of more than one molecular
layer being formed:

p/X(pS - p) = l/xMc + [(c -l)/XMC](P/pS) (9.4.3)

with type III being obtained if c < 1. The BET equation contains two parameters
(x M and c), and it does predict the obvious result that, as p approaches p", x
approaches infinity (an actual liquid phase is formed). If T> Te, the BET
equation could be regarded as having three parameters, though the value of pS
is expected to be close to what one would calculate from extrapolation of the
vapor-liquid equilibrium curve beyond Te. Type IV and type V adsorption
isotherms can be correlated by modifications of the BET equation which contain
additional parameters.
Hysteresis has been explained in terms of capillary effects (hysteresis of the
contact angle, or different geometries of the menisci in adsorption and desorption)
and in terms of swelling phenomena. The latter are particularly important in the
262 case of polymers, and since swelling in polymers takes place at a rate governed
by its intrinsic kinetics, hysteresis is clearly seen to be in fact a kinetic
Chapter Nine
phenomenon.
Finally, adsorption phenomena are often separated into two classes: physical
and chemical adsorption; an actual chemical bond between sorbate and solid is
assumed to be formed in the second case. The separation is somewhat arbitrary,
but three macroscopically observable phenomena can be employed. First, the
heat of chemical adsorption is in general significantly larger, being of the order
of magnitude of the enthalpy change of a chemical reaction. Second, chemical
adsorption may proceed at rather slow rates, since it is essentially a chemical
reaction, while physical adsorption equilibria are generally achieved very quickly.
Finally, chemical adsorption is always confined to a molecular monolayer. In
some cases, chemical adsorption prevails at low partial pressures, while physical
adsorption is observed at higher values of p.

9.5. PHASE EQUIUBRIA IN CONTINUOUS MIXTURES

If you do not expect it, you will not find the unexpected,
for it is hard to find and difficult.
Heraclitus, On Nature, VII

The description of phase equilibria can easily be extended to the case of a


continuous mixture of the type discussed in Section 3.6. Attention in this section
is focused only on gas-liquid equilibria, and in fact only on comparatively simple
instances of it; however, the methodology is obviously liable to generalization.
Let us first consider the free enthalpy functional as given in equation (3.6.18).
If in equation (3.6.19) the scalar a is chosen as the total number of moles in the
system,

a = f n(u) du (9.5.1)

then one obtains for the total free enthalpy density


G = G{p, T; y(u)} (9.5.2)

where y( u) is the mole fraction distribution function.


The case of a liquid phase which is an ideal solution, in equilibrium with a
gas phase which behaves as an ideal gas, will now be examined. This is the
continuous generalization of Raoult's law. To simplify notation, the vapor pres-
sure of each component will be indicated by b, and b itself (rather than u) will
be used as the label identifying the components. This is possible, because in the
case of Raoult's law the vapor pressure is the only constitutive property of interest
[note, however, that y(b) has dimensions of an inverse pressure]. Given a liquid
phase characterized by a mole fraction distribution function yL(b), the partial
pressure of component b in the gas phase is given by
(9.5.3)
The total pressure is given by 263
Phase Equilibria
(9.5.4)

where M' is the second moment of the liquid-phase mole fraction distribution
(the first moment is unity by definition). Hence one obtains

(9.5.5)

Now consider a particular distribution in the liquid phase, say

(9.5.6)

where Q is a normalization factor which guarantees the first moment to be unity,


r is the gamma function or generalized factorial, and n ~ 1 is a parameter which
determined the shape of the distribution function. The valhe n = 1 corresponds
to an exponential distribution function, with the largest mole fraction being that
of the components having a negligible vapor pressure:

yL(b) = exp( -bl Q)I B (9.5.7)

As n approaches 00, the distribution in equation (9.5.6) approaches a delta


function:

(9.5.8)

i.e., the liquid phase is a one-component phase with vapor pressure Q. As n


increases from 1 to 00, the distribution becomes sharper and sharper.
The corresponding vapor-phase mole fraction distribution is calculated from
equation (9.5.5):

(9.5.9)

and is thus seen to be sharper than the liquid phase quantity. Even in the case
of Raoult's law, the shape of the mole fraction distribution in the vapor phase
differs from that in the liquid phase.
We next consider the case of a nonideal liquid phase in equilibrium with a
gas phase which behaves as an ideal gas, and suppose that the pressure correction
in the liquid phase is negligible. One cannot now use b as a label, since two
components with the same vapor pressure b may well have different activity
coefficients, and thus a true label u is needed. The activity coefficient of component
u, r(u), is defined by

JL(u) = RTln[yL(u)r(u)] (9.5.10)

The partial pressure of component u is given by

(9.5.11)
264 and the total pressure is

f f
Chapter Nine
p = pyG(u) du = r(u)b(u)yL(U) du = M" (9.5.12)

where now, however, M" is not the second moment of the liquid-phase mole
fraction distribution. The mole fraction distribution in the gas phase is

(9.5.13)

and will of course in general have a shape different from that of the liquid-phase
distribution. However, this is not necessarily so, and in fact it is possible to have
a continuous azeotrope. Indeed, the activity coefficient distribution will in general
depend on the liquid-phase composition:

r(u) = T{p, T; yL(U)} (9.5.14)

and therefore the equation

M" = r(u)b(u) for all u (9.5.15)

may have a solution for a particular distribution yL(U). That distribution then
corresponds to a continuous azeotrope, since the gas phase and the liquid phase
will have the same composition.
In order to actually carry out calculations for nonideal solutions, it is of
course necessary to write down constitutive equations for the activity coefficients.
These must be obtained from some appropriate generalization of equations written
for mixtures of a few components. A brief sketch of the methodology is given
below for a particularly simple case.
In an ideal solution, the free enthalpy of mixing is given by (the superscript
L is omitted for simplicity)

OlD = RT fy(u) In[y(u)] du <0 (9.5.16)

One can therefore define an excess free enthalpy OEX,

(9.5.17)

and write constitutive equations for it. We now consider the special case where,
for a two-component mixture, the one-parameter Margules equation is adequate,
say equation (8.7.1) holds. For an N-component mixture, if only binary interac-
tions are allowed for, one would write

(9.5.18)
where the second sum is over K different from I, alK = aKI, all = 0, imd 265
gEX = OEX/ RT. The continuous generalization is
Phase Equilibria

gEX = ff a(u, u')y(u)y(u') du du' (9.5.19)

where a(u, u') = a(u', u) and a(u, u) = o.


More generally, the excess free enthapy can be expressed as a functional of
the mole fraction distribution:

gEX = g{p, T; y(u)} (9.5.20)

and again using the property in equation (3.6.19) one obtains

ngEX = g{p, T; n(u)} (9.5.21)

The activity coefficient distribution is obtained in the form

In T(U) = g'{p, 1; u; n(u)} (9.5.22)

where g'{ } is the functional derivative of g{ }; equation (9.5.22) is the generaliz-


ation of equation (8.4.14).
The condition of complete miscibility in the liquid phase is again a condition
of convexity for the free enthalpy of mixing. Let the functional G M1X { } be
defined as

(9.5.23)

Given any set of scalars aK which satisfy

(9.5.24)

the condition of complete miscibility is satisfied provided

(9.5.25)

which defines a convex functional.


Equation (9.5.24) is appropriate if a finite number of phases is considered.
The generalization to the case of infinitely many phases is to consider a phase
amount distribution function a(w), with w the phase label, which satisfies

f a(w)dw=1 (9.5.26)

with y(u; w) being the mole fraction distribution in phase w. The convexity
condition becomes

f a(w)GM1X{p, T; y(u; w)} > G M1X { p, T; f a(w)y(u; w) dW} (9.5.27)


266 EXAMPLES AND PROBLEMS
Chapter Nine

Examples

1. Consider a temperature swing adsorption process. An adsorb able component is


separated from a gas mixture by adsorption on a solid, and desorbed from it at some
higher temperature. For simplicity, assume that Langmuir's isotherm is an acceptable
constitutive equation, and that the highest partial pressure is still such that the term Ap
is negligible in the denominator of equation (9.4.2), i.e., that the adsorption equilibrium
equation is linear, x = Kp. Let Q be the heat of adsorption, so that

K = Kcexp(-Q/RT) (9.E.l)

One may now ask the following question: if one had a choice of sorbant solids, what
value of Q would one want to have in terms of minimizing the energy requirements of
the process?
It would seem at first sight that one requires as low a value of Q as possible, since
one certainly has to provide the heat of desorption in the regeneration step. However,
things are more subtle than that. Let x be the value reached in the adsorption step, and
choose a unit mass of sorbate-free solid as the basis for calculation. The temperature
swing DT required (i.e., the temperature difference between the adsorption and regener-
ation steps) is determined by the partial pressure ratio between extract and reject, which
is a specification of the problem. Linearization of equation (9.E.1) in the temperature
difference DT yields the result that DT is in fact inversely proportional to Q:

DT= a/Q (9.E.2)

There are in fact two heat requirements: the amount needed to heat a mass 1 + x
over a temperature swing DT, which is roughly proportional to (1 + x)DT, and the heat
of desorption, which is proportional to xQ, so that the total energy requirement E is

E = AxQ+ B(l +x)/Q (9.E.3)

This has a minimum at a value of Q which is proportional to [(1 + X)/X]'/2. One


concludes that, if x is large (strongly sorbable gas), one would wish rather low heats of
adsorption, and vice versa.

Problems

9.1. Draw a dew point-boiling point plot for a system exhibiting a miscibility gap in the liquid
phase but no three-phase azeotrope.
9.2. Consider a binary mixture for which the activity coefficients are given by equations (8.7.2)
and (8.7.3). Suppose that the ratio R = p~/ p~ is constant with temperature. Establish conditions for
the values of a and R which will result in the existence of a three-phase azeotrope.
9.3. Consider the plot in Figure 9.3.1. Is component 1 necessarily the one with a larger vapor
pressure, as Figure 9.3.3 implies?
9.4. You have certainly observed that when you add some watery food such as tomatoes in
a pan where some oil has been heated up, violent boiling results, which however ends very rapidly.
When the boiling is observed, is the temperature of the oil necessarily more than 100C? Why do 267
you think the boiling ends rapidly to be substituted by gentle simmering?
9.S. There is no general consensus in the literature on the definition of the r function. Let the Phase Equilibria
argument 0 be an integer; then the two definitions found are such that nO) is either the factorial
of 0 or the factorial of 0 - 1. Which is the definition which applies to equation (9.5.9)?
9.6. Prove that, when 0 approaches 00 in equation (9.5.6), one correctly recovers the one-
component result p = Q.

LITERATURE

It is absurd to have a hard and fasl rule aboul what one


should read and what one shouldn'l. More than half of modern
culture depends on what one shouldn'l read.
Oscar Wilde

There is such an ample literature on phase equilibria that it is difficult to provide any
reasoned guide. Three classical books are M. B. King, Phase Equilibrium in MixlUres,
Pergamon Press, Oxford (1969), A. Reisman, Phase Equilibria, Academic Press, New York
(1970), and J. E. Ricci, The Phase Rule and Heterogeneous Equilibrium, Van Nostrand,
New York (1955).
Of the books more specifically dedicated to special topics, the following are useful
references: A. W. Francis, Liquid-Liquid Equilibriums, Wiley, New York (1963) and
W. Malesinski, Azeotropy, Wiley New York (1965).
With modem computing facilities, constitutive equations for activity coefficients in
mixtures can be used for phase equilibrium calculations even if they are of rather
overwhelming complexity; see J. M. Prausnitz et al., Computer Caculations for Multicom-
ponent Vapor-Liquid and Liquid-Liquid Equilibria, Prentice-Hall, Englewood Cliffs (1980)
and S.1. Sandler, Chemical Engineering Thermodynamics, 2nd Ed., Wiley, New York (1988).
The whole question of the continuous description of gas-liquid equilibria is discussed
in several recent works. The simple case where Raoult's law applies is discussed by J. R.
Bowman, Ind. Eng. Chem. 41, 2004 (1949), and W. C. Edminster and D. H. Buchanan,
Chem. Eng. Prog., Symp. Ser. 6, 69 (1953). More recent literature is reviewed by S. K.
Shibata, S. I. Sandler, and R. A. Behrens, Chem. Eng. Sci. 42, 1977 (1987). Some of the
published literature fails to recognize that the shape of the distribution function in the
gas phase is different from that in the liquid phase [see, e.g., R. L Cotterman, D. Dimitrelis,
and J. M. Prausnitz, Ber. Bunsenges. Phys. Chem. 88, 796 (1984)] and should therefore be
read with care.
Chapter Ten

CHEMICAL EQUILIBRIA

The ten books on Pyrotechnics, where one amply


discusses not only each kind and variety of mining, but
also what is being researched on the practice of those
things which pertain to the art of melting and casting
metals, as well as any other similar thing. Typeset for
Mr. Vanoccio Biringuccio of Siena. With Apostolic
Privilege, and of the Caesarean Majesty, and of the
very Illustrious Venitian Senate.
Venice, 1540

NOTATION

a Activity
B Henry's constant atm
c Concentration kmol m- 3
CT Average molar density kmolm- 3
Cp Specific heat kcal kmol- 1 K- 1
D(L) Operator defined in equation (10.1.3)
fO Standard state fugacity atm
Gt Total free enthalpy kcal
gEX Dimensionless excess free enthalpy
g Vector of reference free enthalpies kcalkmol- 1
K Equilibrium constant vector
K( Equilibrium constant distribution
L Density of any extensive property
L( L distribution
M Number of independent reactions
m Initial concentration of amine kmolm- 3
n Total number of moles kmol 269
270 n( Number of moles distribution kmol
Chapter Ten n Number of moles vector kmol
N Total number of components
p Pressure atm
q Parametric vector kmol
r Steam/methane ratio
R Gas constant kcal kmol- 1 K- 1
T Temperature K
u Species label
VL Liquid specific volume m3 kmol- 1
v Fugacity coefficient
w Reaction label
x Extent of reaction vector kmol
XT Total extent of reaction vector kmol
y Mol fraction
y( Mol fraction distribution
z Parameter vector in Section 10.6
~ Pressure correction coefficient
f3( ). ~ distribution
p Vector defined in equation (10.1.10)
'T Activity coefficient
'TD Activity coefficient for dilute solution
fA. Chemical potential vector kcalkmor 1
p.( Chemical potential distribution kcal kmol- 1
a Stoichiometric coefficient matrix
u( Stoichiometric coefficient distribution

Subscripts
0 Initial value
j Speciesj
k Reaction k
T System as a whole

Superscripts

* At equilibrium
S For saturated vapor
K For nonreactive phase K
10.1. HOMOGENEOUS EQUILIBRIA IN
IDEAL AND DILUTE SOLUTIONS

It is known that in the blast furnace the reduction of iron oxide


is produced by carbon monoxide, but die gas leaving the
chimney contains a considerable proportion of carbon
monoxide, which thus carries away an important quantity of
unutilized heat. Because this incomplete reaction was thought
to be due to an insufficiently prolonged contact between carbon
monoxide and the iron are, the dimensions of the furnaces have
been increased. In England they have been made as high as
thirty meters. But the proportion of carbon monoxide escaping
has not diminished, thus demonstrating, by an experiment
costing several hundred thousand francs, that the reduction of
iran oxide by carbon monoxide is a limited reaction.
Acquaintance with the laws of chemical equilibrium would have
permitted the same conclusion to be reached more rapidly and
far more economically.
Le Chatelier, 1882

The theory of chemical equilibria consists essentially in the determination of the


eqUilibrium composition in a system for which the reaction subspace is determined
by the knowledge of an initial composition or of sufficient mass balance con-
straints. In the case of homogeneous (Le., one-phase) equilibria, the thermody-
namic equilibrium conditions is given by equation (3.4.2), which is rewritten
below.in terms of the N x M matrix of stoichiometric coefficients a (N being
the number of components and M the number of independent reactions, with
M < Nand M being the rank of the a matrix):
aT ..... * = 0 (10.1.1)

It is useful at this stage to introduce some shorthand notation. Given a


quantity L which is defined for each component, the operator D(L) has already
been introduced in Section 3.4. For each reaction Ie, one has

DK(L) = L (uJKLJ ) (10.1.2)


J

or, in vector notation,

D(L) = aTL (10.1.3)

where vectors are M-dimensional. Equation (10.1.1) can thus be written as

(10.1.4) 271
272 Another useful operator is defined as follows for each reaction:
Chapter Ten
(10.1.5)

The M quantities 71'K can also be interpreted as an M-dimensional vector. The


operators D and 11 are related to each other as follows:

In[1I(L)] = D(In L) (10.1.6)

Equation (10.1.1) represents a system of M equations in the N unknowns


which are the components of the chemical potential vector fI. *. The mathematical
problem is closed by the following equations:

1. A set ()f N constitutive equations for the chemical potentials relating


these to the N number of moles of the components present in the system
at equilibrium, say to the equilibrium number of moles vector n*.
2. The stoichiometric constraint for the n* vector to lie in the reaction
subspace, say

n* = Do + a . x* (10.1.7)

where x* is the M-dimensional extent of reaction vector.

The general approach to the solution of this type of problem has been
discussed in Section 3.4; see equation (3.4.5). It is based on the definition of a
standard state, with respect to which the M-dimensional equilibrium constant
vector .K is defined. In this section, attention is restricted to the simple cases
where the activity coefficients are either unity (ideal solutions) or independent
of composition (dilute solutions), so that the constitutive equations for step (1)
assume a particularly simple form.
The very simplest case is where the system is a gaseous mixture at a pressure
low enough not only for the solution to be ideal, but also for all components to
behave as ideal gases. In this case, the constitutive equations for the chemical
potentials are

fI. = g(T) + RTln(py) (10.1.8)

Equation (10.1.8) still presents an ambiguity, since pressure could be


measured in different units, and (coherently) the function g(T) for every com-
ponent is in fact defined to within an arbitrary additive constant. The ambiguity
is resolved by the conventional choice of measuring pressure in atmospheres, so
that g( T) is the free enthalpy of the component, should it behave as an ideal
gas, at a pressure of 1'atm (this is the standard state). Consequently, the value
of the equilibrium constant defined in equation (3.4.5) is independent of pressure:
the equilibrium constant vector is given by

.K = exp[ -D(g)/ RT) (10.1.9)


Let the M-dimensional vector IJ be defined as 273
Chemical
13K = L, (TJK (lO.1.10) Equilibria

The kth component of IJ is zero if the kth reaction is equimolar. Equation (3.4.5)
reduces to

1T(Y*) = Kp-Il (10.1.11)

Equation (10.1.11) should be discussed in some detail. If the reactions


involved are all equimolar (IJ = 0), the equilibrium composition is entirely
independent of pressure, since so is the right-hand side. Other things being equal,
reactions which result in a decrease of the total number of moles (13K < 0) are
favored (in the sense that the product mole fraction is increased at equilibrium)
by an increase in pressure, and vice versa.
Equation (10.1.11) is a set of M equations for the N unknowns y,. Upon
substitution of equation (10.1.2), the unknowns are reduced to the M values of
the extents of reactions. The physically significant solutions are those which,
upon substituting back the x* value into equation (10.1.7), yield all n, values
which are both real and positive. In the simple case considered here, only one
such solution will exist. However, it should be remembered that, unless M is a
rather small number, solution of the coupled system of polynomial equations
may present significant difficulties, and it may be preferable to directly seek the
minimum of the mixture free enthalpy in the reaction subspace.
In the case of gaseous mixtures at moderately high pressures, where the
mixture still behaves as an ideal solution but the individual components exhibit
deviations from ideal-gas behavior, equation (10.1.8) becomes

(10.1.12)

where v are the fugacity coefficients.


It should be noted that, even iflJ = 0, the right-hand side of equation (10.1.12)
now depends on pressure, though it is still independent of composition.
We now consider the case of dilute liquid solutions. The chemical potentials
of the solutes are given by

(10.1.13)

If only the solutes participate in the reactions, then equation (10.1.4) reduces to

(10.1.14)

where B is Henry's constant. To within the zero-order approximation on which


the theory of dilute solutions is based, the chemical potential of the solvent is a
constant the value of which can be included in the definition of K, and hence
equation (10.1.14) also applies to the case where the solvent participates in the
chemical reactions. [In regard to equation (10.1.14), see Section 10.4.]
274 In all three cases discussed so far, the equilibrium condition reduces to
1I'(y*), being equal to a vector which depends only on pressure and temperature.
Chapter Ten
The number of moles vector at equilibrium is given by equation (10.1.7), and the
total number of moles is given by

n* = ~ n1 = no + IS' x* (10.1.15)
J

so that

Y* = (110+ (J" x*)/(no+ IS x*) (10.1.16)

and therefore the equilibrium equations reduce to a system of M coupled


polynomial equations.
The equilibrium constant vector K depends only on temperature. Straightfor-
ward application of the appropriate Maxwell relations yields

15 In K/ 15(1/ T) = -D(H)/ R (10.1.17)

Equation (10.1.14) can be used directly to calculate K at temperatures other


than that (usually 25C) for which data are tabulated. The values of D(H) at
all temperatures can be calculated if the value at one particular temperature is
known, provided values of Cp as a function of temperature are available.

10.2. HOMOGENEOUS EQUILIBRIA IN NONIDEAL MIXTURES

Nothing is created either in the operations of the laboratory, or


i1' those of nature, and one can affirm as an axiom that, in
every operation, there is an equal quantity of matter before and
after the operation; that the quality and quantity of the
principles are the same, and that there are only alterations and
modifications. On this axiom is founded the whole art of
making experiments in chemistry; we must suppose in all of
them a true equality or equation between the principles of the
body one examines, and those that we extract by analysis.
A..L. Lavoisier

Let us consider a nonideal mixture in which chemical reactions may take


place. The chemical potential of every component is given by

(10.2.1)

where fO is the fugacity of the pure component at the same pressure and
temperature of the mixture, and in the same (in most cases, liquid) phase. The
value of f O is given by

(10.2.2)
where the superscript S identifies the saturation conditions of pure-component 275
gas-liquid equilibrium. If P is defined as
Chemical
Equilibria
(10.1.3)

then equations (10.2.1) and (10.2.2) can be combined to yield

(10.1.4)

which identifies g(T) with in the usual convention if the vapor pressure pS is
measured in atmospheres (see also Section 10.4). The equilibrium condition
becomes

(10.1.5)

There is a very substantial difference between equation (10.2.5) and the


eqUilibrium equations established in the previous section: the right-hand side in
the present case depends on composition through the activity coefficient term T*.
This makes the solution of the equilibrium problem significantly more complex,
since the system of coupled polynomial equations (10.2.5) has to be coupled with
the constitutive equations for the activity coefficients. This additional difficulty
can be dealt with in two ways:

1. Some simplified model for the activity coefficients can be used. In par-
ticular, it should be noted that the activity coefficients need not be known
individually, but only in the combination appearing in the operator 'IT;
this allows on occasion considerable simplifications of the constitutive
equations.
2. The eqUilibrium composition is sought by directly looking for the
minimum of free enthalpy in the composition subspace. If an equation
for gEX is available, this can be accomplished generally with numerical
techniques which are less cumbersome than those required for the solution
of the problem as formulated by equation (10.2.5).

The mathematical structure of the problem as resulting from the second


approach is as follows. The total free enthalpy of the mixture is given by the sum
of three terms: an additive part, an ideal solution mixing part, and an excess
part accounting for deviations from ideality. Hence

GJ RT = n g+ n .lny+ ngEX(y) (10.1.6)

Each component of g is defined to within an additive arbitrary constant.


However, knowledge of the value of D(g ) sets M constraints on these N arbitrary
constants, and this is all that is required, as shown below. The number of moles
vector is

n=no+ax (10.1.7)
276 and therefore the additive term in equation (10.2.6) can be expressed in the form
Chapter Ten
n' g = Do' g + (u x) g
= Do g + x (u T g)
= Do' g + X D(g) (10.2.8)

The unknown term Do g is constant and thus contributes nothing to the location
of the minimum of the free enthalpy in the reaction subspace. All other terms in
equation (10.2.6) can be expressed explicitly in terms of x.
Since attention is being restricted here to mixtures which do not exhibit
miscibility gaps, the free enthalpy surface over the composition space (and thus
a!ortiori its restriction to the reaction subspace) is concave upward everywhere,
and therefore only one minimum can exist; in fact, it is also the only extremal
point (there can be no saddle points and no local maxima). It follows that the
numerical search for the equilibrium point presents no major difficulties.

10.3. HETEROGENEOUS EQUILIBRIA

Think of the image of the world in a convex mirror. ... I do


not see how men in the mirror are to discover that their bodies
are not rigid solids and their experiences not good examples of
the correctness of Euclidean axioms. ... If two inhabitants of
the two different worlds could communicate with one another,
neither, as far as I can see, would be able to convince the other
that he had the true, the other the distorted, relation. Indeed I
cannot see that such a question would have any meaning at all,
so long as mechanical considerations are not mixed up with it.
H. Helmholtz

The analysis of heterogeneous chemical equilibria is particularly complex


and is best carried out by considering different special subcases; some of these
will be discussed in Sections 10.5 and 10.6. The difficulties encountered are
essentially the following two:

1. Given an overall initial composition of the system considered as a whole


(or any equivalent set of mass balance constraints), the number and type
of phases which will in fact be present at equilibrium is not known a priori.
2. Stoichiometry restricts the average composition of the system to lie within
the reaction subspace, which is identified by the initial average composi-
tion, but it does not restrict the composition of any individual phase
present at equilibrium.

As was discussed in Section 4.3, the conditions for heterogeneous chemical


equilibrium are that the affinity vector is zero in all phases present at equilibrium,
and that the chemical potential of each component should be the same in all
phases. These conditions are redundant in the sense that, once the affinity vector
is zero in one particular phase, equality of the chemical potentials implies that
it is zero also in all other phases. One therefore has a degree of freedom in setting
up the analysis of any given problem, in that one can choose one particular phase 277
and regard it as the "reactive" phase, i.e., one chooses to impose the condition
Chemical
that the affinity vector is zero in only one particular phase. This must then be Equilibria
coupled with the phase equilibrium condition, that the chemical potentials of all
components are the same in all phases.
The choice of "reactive" phase is a matter of convenience of calculation.
One general criterion is that the reactive phase should be that which is guaranteed
to exist at equilibrium. A second is that the components which participate in the
reactions should be present in the reactive phase in nonnegligible amounts at
equilibrium; it will be seen that it is not always possible to satisfy this condition.
Finally, if there is still room for choice, it is best to select as the reactive phase
that for which the constitutive equations for the chemical potentials have a
particularly simple form. Thus, e.g., the choice of a gaseous phase (which can
usually be regarded as an ideal solution) as the reactive phase is to be preferred
whenever possible.
Some general notation is introduced in the following. The subscript T will
be used to identify the system as a whole; an "initial" composition, i.e., a total
number of moles vector DTO, is regarded as known. This is essentially one point
of the reaction subspace that may be either an actual initial composition, or a
composition which is "admissible" in the sense that it is compatible with the
mass balance constraints imposed on the system. Given the matrix (T of
the stoichiometric coefficients, the total number of moles vector DT is given
by the usual equation

(10.3.1)

where XT is the "total extent of reaction" vector. Equation (10.3.1) determines


the reaction subspace.
Quantities in the phase chosen as the reactive phase will bear no superscript,
while quantities in other phases (which mayor may not be present at equilibrium)
will be identified by superscripts K = 1,2, .... These phases will be called
"nonreactive," though of course there is no implication that the chemical reactions
may not take place in them as well. It follows that DT is given by

(10.3.2)

If N is the total number of components, then the best way to formulate the
problem is to seek a parametric solution, where the parameter is an N-dimensional
vector q, which represents the "initial number of moles" of the components in
the reactive phase, say (in a purely formal sense)

(10.3.3)

We note that each component of q does not represent the number of moles of
that component present in the reactive phase at equilibrium. The components of
q must of course satisfy the obvious condition

(10.3.4)
278 If q is regarded as known, one can write
Chapter Ten
n=q+ax (10.3.5)

and hence the problem is reduced to a homogeneous chemical equilibrium


problem in the reactive phase. This problem can be solved by the methods
discussed in the preceding sections (and the solution is computationally easiest
if the reactive phase is that for which the constitutive equations for the chemical
potentials are simple). Such an approach yields the chemical potentials and the
number of moles at equilibrium in the reactive phase, respectively,

n* = n*(q) (10.3.6)

and
IJ.* = 1J.*(q) (10.3.7)

The condition of physical eqUilibrium now requires that, for every other
phase which may conceivably exist at equilibrium,

(10.3.8)

and, given constitutive equations for the chemical potentials in the nonreactive
phases, this can be transformed to

(10.3.9)

. Substitution of equations (10.3.6) and (10.3.9) into (10.3.2) will in principle


result in the elimination of die parameter q and hence in the complete solution
of the problem. In practice, two difficulties may arise:

1. There is no way to satisfy all the equations. This implies that either one
or more of the phases assumed to exist at equilibrium does in fact not
exist, or vice versa.
2. The solution is likely to be nonanalytical, and hence elimination of q may
be computationally difficult.

10.4. ACfIVITIES

A Crocodile had stolen a Baby off the banks of the Nile.


The Mother implored him to restore her darling.
"Wel~ .. said the Crocodile, "if you say truly what
I shall do I will restore it: if not, I will devour it. ..
"You will devour it," cried the distrocted Mother.
If he devours the Baby, he makes her speak truly, and so
breaks his word; and if he restores it, he makes her speak
falsely, and so breaks his word. His sense of honour being thus
hopeless of satisfaction, we cannot doubt that he would act in
accordance with his second ruling passion, his love for babies.
Lewis Carroll
In many instances where some of the components in a reactive system are 279
in phases other than gaseous, it is sometimes difficult to determine experimentally
Chemical
the value of g( T), the free enthalpy of the component considered at the tem- Equilibria
perature of the system should it be an ideal gas at a pressure of one atmosphere.
In such cases, it is more convenient to redefine the equilibrium constants in a
different way, and accordingly to rewrite the constitutive equations for the
chemical potentials coherently with the new definition of equilibrium constants.
As was discussed in Section 3.4, the constitutive equation for the chemical
potential of any component is written in the following way, which is to be regarded
as a definition of the activity a:

(10.4.1)

where p, is the chemical potential of the component considered in the "standard


state." Correspondingly, the equilibrium constant is redefined as

K = exp[-D(p,)/RT] (10.4.2)

and the eqUilibrium condition becomes

1T(a) = K (10.4.3)

The choice of standard state is, of course, arbitrary. In practice, it is chosen


in such a way that values of K can be determined experimentally, and can be
tabulated most easily by the methods discussed in Section 3.4. The standard state
is always chosen at a pressure of one atmosphere, so that the value of K is
independent of pressure. Furthermore, the standard state is always chosen at the
same temperature as the system, so that K does depend on temperature. The
usual choices for the standard state are listed below.

Gases. The standard state is the pure gas regarded as an ideal gas. With
this, p, is simply g(T) as established in Section 10.1, and hence equations
(10.1.9)-(10.1.11) hold with p measured in atmospheres.

Dilute Solutions. In a dilute solution (even if only moderately dilute, so that


quantities TD diflerfrom unity), the chemical potential ofa component is given by

(10.4.4)

where, if g( T) is to be interpreted in the usual sense, B is to be measured in


atmospheres and, coherently, f3 is the correction factor for pressure between the
actual pressure and a pressure of one atmosphere (the value of f3 is taken as
unity in most cases). If CT is the average molar density of the mixture and C the
concentration of the component considered, both measured in moles per liter,
then equation (10.4.4) can be expressed in the form

(10.4.5)
280 The chemical potential in the standard state is defined as
Chapter Ten
,.,.0 = geT) + RTln(B/ CT) (1004.6)

i.e., the standard state is the component considered at a concentration of one


mole per liter if 'To is still unity at such a concentration. Correspondingly, the
activity is given by

a = cr D {3 (1004.7)

and the equilibrium condition becomes

(1004.8)

which, neglecting pressure corrections, becomes in the zero-order approximation

1T(C) = K (1004.9)

In the zero-order approximation, CT is simply the molar density of the solvent


(55.5 moles per liter for aqueous solutions at standard conditions) and that is
the value actually appearing in equation (10.4.6), with any difference of CT at
higher solute concentrations being included in the value of 'To.

One-Component Phases. For one-component phases (mostly solids, but


liquids as well when they are known to be essentially one-component) the standard
state is the pure component itself at a pressure of 1 atm:
,.,.0 = G(T) (1004.10)

and hence the activity is unity to within the pressure correction term {3.

10.5. ONE-COMPONENT NONREACfIVE PHASES

That is how, for instance, that calcareous spath, produced in


water by the combination of mephitic acid with simple earth,
which Mr. Sage calls "adsorbing," is not soluble in water as
long as so combined: but if one decomposes it by calcination,
the igneous acid thus replacing the mephitic one pushed out by
fire, it forms, with the base earth of spath, a new compound
which is called qUicklime, which is per/ectly soluble in water.
However, if the igneous acid is displaced again by mephitic
acid, the calcareous spath so regenerated will, as before,
be insoluble in water. [Mephitic acid: since that is the acid
which in nature produces lethal and suffocating vapors,
one couldn't indicate it with a better name.
"Savamque exhalat opaca mephitim," VergiLJ
M. De Rome de I'Isie, 1783
In this section, the following class of equilibrium problems is considered: 281
in addition to a reactive phase which is a mixture, there may at equilibrium be Chemical
one or more nonreactive phases which are one-component solid or liquid phases. Equilibria
For simplicity, pressure corrections in solid and liquid phases will be neglected.
It follows that the equilibrium activity of all the components present at equilibrium
also as a one-component phase is unity in the reactive phase-independently of
the amount of one-component phases present. While this condition sets the
activity of a certain number of components (hence decreasing the number of
unknowns in the problem), it produces an equal number of new unknowns, i.e.,
the amounts of the one-component phases present.
We first examine the very simple case where there is only one chemical
reaction to consider, and it involves two one-component phases. Two specific
examples are, first, the decomposition of limestone

CaO + CO 2 = CaC0 3 (10.5.1)

and second, the flue gas desulfurization process where a lime slurry is used for
absorbing sulfur dioxide:

(10.5.2)

Let 1 be the component which does not form a one-component phase (C0 2
and S02 for the two specific examples). In both examples, the activity of com-
ponent 1 is, at equilibrium, equal to the equilibrium constant of the reaction, i.e.,
it is a unique function of temperature:

at = K(T} (10.5.3)

This means that, at any given temperature, there is only one concentration of
component 1 in the reactive phase which, at equilibrium, is compatible with the
presence of both one-component phases. If the concentration is larger than this
value, the reaction will proceed to the right until the solid phase on the left
disappears, and vice versa. This has important practical consequences, such as
discussed subsequently for the two examples given above.
In the limestone decomposition case, coexistence of CaO and limestone is
possible at equilibrium only at one particular partial pressure of CO 2, That partial
pressure at room temperature is extremely low, in fact lower than the value in
atmospheric air; that is why limestone is a perfectly stable compound commonly
found in rocks. As the temperature is increased, the equilibrium CO 2 partial
pressure increases, and as soon as it exceeds the partial pressure of CO 2 in the
atmosphere a driving force for limestone decomposition appears. However, the
reaction is very slow, since it could only take place on the exposed surface of
the limestone. As temperature is increased further, the eqUilibrium CO2 partial
pressure reaches one atmosphere, and now a driving force exists even within the
limestone, so that decomposition will actually take place. At a temperature even
slightly larger than that, the equilibrium partial pressure of CO2 is larger than
one atmosphere, and therefore in a furnace open to the atmosphere the equilibrium
condition can never be realized until the limestone has decomposed completely
and the only solid left is CaO.
282 For the desulfurization process, the equilibrium concentration of S02 in the
Chapter Ten
liquid phase at room temperature is extremely low so that, in the presence of a
gas phase containing S02 in nonnegligible amounts, absorption continues and
lime is continuously transformed into gypsum, and equilibrium is reached only
when either all the lime has been consumed, or the partial pressure of S02 in
the gas phase has been reduced to the extremely low value corresponding to
coexistence of lime and gypsum. In practice, the process is run in such a way
that the first condition can never be realized (the entering flowrate of lime is
slightly in excess of the stoichiometric value needed to eliminate all the S02 from
the gas phase), and hence the process can in fact reduce the S02 content in the
gas to very low values. We note that, as long as the process is run, the slurry
contains both lime and gypsum, and hence the S02 concentration in the liquid
phase is kept at a constant low value. This means that there is no need to run
the slurry and the gas phases in countercurrent, but any flow arrangement is
equivalent to any other as far as the distribution of driving forces is concerned.
Indeed, the limestone process is always run with configurations other than the
countercurrent one.
More complex problems may arise when there is more than one reaction to
be considered. A good example is the steam reforming reaction of hydrocarbons;
for simplicity, consider the case of methane. The following two homogeneous
reactions are to be considered:
(10.5.4)

and
(10.5.5)

In addition to these, the possibility of methane cracking should be con-


sidered:
(10.5.6)

and the technical problem is typically that one wants to operate under conditions
where no solid carbon is formed. There are six components to be considered,
and these are numbered consecutively in the following order: CH 4 , H 20, CO,
H 2, CO 2, C. If the reactor is fed with a mixture of H 20 and CH 4 in a ratio r,
the initial total number of moles vector is
DTO = 1, r, 0, 0, 0, 0 (10.5.7)

Should solid carbon be present, its activity in the gaseous phase would be
unity, which corresponds to a negligibly small amount of carbon vapor (the
sublimation pressure of carbon is negligibly small). One may therefore assume
that, if no solid carbon forms at equilibrium, in fact no carbon at all forms; in
the terminology of Section 10.3, one chooses q to coincide with DTO. Reaction
(10.5.6) is tentatively excluded from consideration, and the actual number of
moles vector is expressed as
(10.5.8)

where Xl and X2 are the extents of reactions (10.5.4) and (10.5.5), respectively.
One now can solve this homogeneous equilibrium problem with the methodology
discussed in Sections 10.2 and 10.3 to yield the equilibrium activities of the 283
first five components; their value will, of course, depend on the value of the
Chemical
parameter r. Equilibria
Reaction (10.5.6) will in fact not take place provided the following condition
is satisfied:
(10.5.9)

where K is the equilibrium constant of reaction (10.5.6) [if condition (10.5.9) is


satisfied, the activity of carbon is less than unity]. The left-hand side of equation
(10.5.9) is an increasing function of r, and hence a minimum value of r needed
to avoid the formation of solid carbon at equilibrium can be calculated. This
does not by itself guarantee that no solid carbon will form at all; for example,
if reaction (10.5.6) is much faster than the others, solid carbon may form initially.
The calculation only guarantees that no solid carbon will be present at equilibrium.
In the steam reforming example, the sublimation pressure of solid carbon
is so low that one can assume that, if no solid carbon is formed, no carbon at
all is formed. This is not generally the case, as the following example shows.
Consider the reduction of a metal oxide with carbon. The reaction is carried out
in a high-pressure vessel, and only liquid metal can be withdrawn from the vessel.
One aims at complete reduction of the metal oxide, and hence the reactor will
be charged initially with an amount of carbon in excess of the stoichiometric
amount required for complete reduction; this guarantees that solid carbon will
always be present. The following two reactions are selected for consideration:

MO+CO = M+C0 2 (10.S.10)

and
C+C02 =2CO (10.S.U)

while the five components are numbered in the following order: MO, CO, M,
CO2 , C. Of these only 2, 3 and 4 may be present in significant amounts in the
reactive gaseous phase. Since solid carbon is guaranteed to be present at equili-
brium, the following condition is satisfied:

(10.5.1:!)

where K is the equilibrium constant of reaction (10.5.11).


As long as there is still solid MO present, the following condition is also
satisfied:

(10.5.13)

where K' is the equilibrium constant of reaction (10.5.10).


Since the reactive phase is gaseous, activities are related to mole fractions by

a = Pyv (10.S.14)

and hence equations (10.5.12) and (10.5.13) can be written, respectively, as

Y'VY4 = K/pTr(v) = F(T,p) (10.5.1S)


284 and
Chapter Ten (10.5.16)

The three nonzero mole fractions in the gaseous phase are related to each
other by two additional equations. The first is simply the requirement that they
add up to unity:

(10.5.17)

The form of the second relationship depends on whether liquid metal does or
does not form. First we consider the case where it does not form. In this case,
for every gmole of metal in the reactive phase, there must also be one gatom of
oxygen, and hence

(10.5.18)

If some liquid metal does form, there are more gatoms of oxygen in the
reactive phase than there are gmoles of metal; however, the phase equilibrium
condition with the liquid metal now imposes

PY3 = P~ (10.5.19)

Since there are in both cases four equations relating the three nonzero mole
fractions, equations (10.5.15) and (10.5.16) can be satisfied, for any given tem-
perature, at only one particular pressure. Thus for the two subcases considered
one can solve the homogeneous reaction problem and obtain both the mole
fractions in the gaseous phase and the pressure.
In practice, it is best to first treat the case where no liquid metal is formed
and solve the problem described by equations (10.5.15)-(10.5.18). The calculated
pressure increases with increasing temperature, and so does the fugacity of the
metal vapor. For the special case where the metal is zinc, the left-hand side of
equation (10.5.19) increases with temperature more rapidly than the right-hand
side, so that at low temperatures no liquid zinc will form. It is therefore possible
to establish a minimum operating temperature (and a corresponding equilibrium
pressure) at which liquid zinc will form at equilibrium. If the reactor is run at
any higher temperature, liquid zinc is guaranteed to form, and can thus be
withdrawn. As it is withdrawn, new zinc oxide will be reduced, until reduction
is complete.

10.6. NONREACfIVE MIXTURES

Dry carbonic aCid, mixed with dry ammonia, gives a solid


white compound formed by the union of one molecule of
carbonic acid with two of ammonia. If heated, this solid gives
off gases which, from their properties, the chemists have
identified as a mixture of carbonic acid and ammonia.
P. Duhem, 1895
Attention is now focused on the case where two phases, both of which are 285
mixtures, are known to exist at equilibrium. As specific examples, two cases of
Chemical
gas-liquid systems are considered, but the method can be generalized. The Equilibria
existence of both a gaseous and a liquid phase at equilibrium is guaranteed, for
example, if attention is focused on a system where there is an aqueous solution,
and the temperature is low enough that all the water present cannot possibly be
vaporized; hence the existence of a liquid phase is guaranteed. In addition to
this, the system contains a relatively large amount of a chemically inert component
which has a low solubility in the aqueous phase (e.g., methane), so that
unless the pressure is enormously high the existence of a gas phase is guaranteed.
The methodology illustrated in the two examples considered is essentially that
sketched in Section 10.3.
The choice of which phase to regard as reactive is a matter of convenience.
From the viewpoint of simplicity of the constitutive equations for the chemical
potentials, it would be preferable to choose the gas phase. However, the examples
considered are cases where this is not possible, since some of the reactive species
are ions, which are essentially confined to the liquid phase. Therefore, the liquid
phase is chosen as the reactive one and the problem is reduced to a homogeneous
equilibrium problem by introducing the parameter q discussed in Section 10.3.
The first example is that of a system containing methane and carbon dioxide
as the main components in the gas phase; the liquid phase is initially an aqueous
solution of an amine RNH 2. The six reactive components to be considered are
numbered consecutively in the following order: CO 2, RNH 2, RNH;, HC03",
C03"-, RNCOO-. All components except CO 2 are regarded as confined to the
liquid phase. Water is regarded as being present in large enough an excess that
its chemical potential is essentially constant; coherently, water itself is not written
in the reactions. The density of the liquid phase is regarded as constant, and
therefore, if the basis of calculation is one liter of aqueous solution, the number
of moles is equal to the concentration.
If m is the moles per liter of amine initially present in the liquid phase, then
the six-dimensional vector q is chosen as

q = zm, m, 0, 0, 0, 0 (10.6.1)

and is thus identified by the value of one scalar parameter, z. This parameter
represents the ratio of the total number of moles of CO 2 (in both physically
dissolved and chemically combined form) in the liquid to the initial number of
moles of amine.
The reactions to be considered are

CO2 + RNH2 = RNH; + HC03" (10.6.2)

RNH2 + HC03" = RNH; + C03"- (10.6.3)

RNH2 + HC03" = RNCOO- (10.6.4)

Hence the number of moles vector is

(10.6.5)
286 where Xt. X2 and X3 are the extents of reactions (10.6.2)-{10.6.4), respectively.
The problem is reduced to a homogeneous equilibrium problem, with three
Chapter Ten
independent reactions. The solution of this problem will yield the equilibrium
chemical potentials of all components as functions of the parameter z:

Jl.*=Jl.*(z) (10.6.6)

In particular, the chemical potential of CO 2 is obtained as a function of z:

(10.6.7)

and, since CO 2 is the only component which can be present in both phases, the
phase equilibrium condition for CO 2 yields the partial pressure of CO 2 in. the
gas phase as a function of z. This constitutes essentially the required informa-
tion.
The second example is where there are two components which maybe present
in both phases, say CO 2 and H 2 S. To simplify the chemistry involved, attention
is restricted to the case where the liquid phase is an aqueous solution of a base
B, which is not alkaline enough for any significant amount of either S-- or CO)-
to form. The components to be considered are numbered in the following
order: H 2 S, CO 2 , B, BH+, HCO), HS-. With m again the number of moles per
liter of B initially present in the liquid phase, the parameter q is chosen in the
form

q = z,m, z2m, m, 0, 0, 0 (10.6.8)

i.e., it is here determined within the two-dimensional vector z. The reactions to


be considered are

CO 2 + B = BH+ + HCO) (10.6.9)

H 2 S + B = BH+ + HS- (10.6.10)

and thus the number of moles vector is given by

(10.6.11)

and the problem is reduced to a two-reaction homogeneous equilibrium problem.


The solution will yield

(10.6.12)

and the phase equilibrium conditions for CO 2 and H 2S will thus yield
the equilibrium partial pressure of the latter in the gas phase as functions
ofz.
10.7. CHEMICAL EQUILIBRIA IN CONTINUOUS MIXTURES 287
Chemical
An Algument to prove, that the Abolishing of Christianity Equilibria
in England, may, as Things now stand, be attend~d
with some Inconveniences, and perhaps, not produce
those many good Effects proposed thereby.
Jonathan Swift

The methodology of the analysis of chemical equilibria in continuous mix


tures is rather straightforward, and some essential points are discussed below.
The actual solution of chemical equilibria problems may, however, involve subtle
difficulties.
The fundamental equation which, in principle, determines the equilibrium
composition is of course equation (3.6.36), namely

f u( w, u)# *(u) du = 0 (10.7.1)

Given a quantity L which is defined for each component [i.e., the function
L(u) is assigned], the operator D in equation (10.1.3) is generalized to

D(w, L) = f u(w, u)L(u) du (10.7.2)

Equation (10.7.1) can thus be written as the following generalization of


equation (10.1.4):

D(w, #*) = 0 (10.7.3)

The operator 7T in equation (10.1.5) can also be generalized as follows:

7T( w, L) = f L(u)u(w,u) du f
= exp[u( w, u) In L(u)] du (10.7.4)

but this is not very useful. Equation (10.1.7) becomes

n*(u) = no(u) + f u(w, u)x*(w) dw (10.75)

Let us now consider the case where the mixture is an ideal gas. The chemical
potential distribution is

#(u) = g(T, u) + RTln[py(u)] (10.7.6)

and the equilibrium constant distribution is

K(w) = exp[ -D(w, g(T, u/ RT] (10.7.7)


288 The nonequimolarity factor f3 has the following distribution:

f
Chapter Ten
f3(w) = u(w, u) du (10.7.8)

by means of which equation (10.1.11) is generalized to

7T(W,y*(U = K(w)p-Il(w) (10.7.9)

Analogous equations could be developed for dilute solutions and for nonideal
mixtures. However, solving this type of equation for y*(u) may be extremely
difficult, even for the simple ideal-gas case, and thus the equilibrium composition
is best found by directly minimizing free enthalpy. The discussion in Section 10.2
concerning this technique is easily generalized. Equation (10.2.6) becomes

G.I RT = f n(u)g( T, u) du

+ f n(u) In y(u) du + nGEX{p, T; y(u)} (10.7.10)

The unknown term is of course the first integral. However, that term can be
expressed as follows:

f n(u)g(T, u) du = f no(u)g(T, u) du + ff u(w, u)g(T, u)x(w) dwdu

= f no(u)g(T, u) du + f x(w)D(w, g) dw (10.7.11)

It follows that knowledge of the D( w, g) function determines completely the


minimum, since the first integral on the right-hand side of equation (10.7.11) is
a constant.

EXAMPLES AND PROBLEMS

Problems

10.1. If any fuel is burned to heat up a house, there is a risk that CO may form. This means,
as compared to the case when only carbon dioxide is produced, both a decreased efficiency (oxidation
of CO is an exothermic reaction) and a risk, since CO is poisonous. Discuss methods available for
minimizing the risk of CO production.
10.2. Lime is a comparatively expensive chemical to use for flue gas desulfurization, and it is
therefore proposed to use a limestone slurry as the absorbing medium. Discuss the chemistry involved
and establish the eqUilibrium conditions for such a process. Can the sulfur dioxide mole fraction in
the gas phase still be reduced to a negligible amount?
10.3. Does Duhem, in the quote of Section 10.6, refer to ammonium carbonate or to ammonium 289
carbamate?
10.4. In an initial analysis of a methanol synthesis process, the following chemical reactions Chemical
have been considered: Equilibria

co + 2H2 = CH 3 0H
CO + 3H 2 = CH. + H 20
CO + H 20 = CO 2 + H,
2CO = CO,+C

CO 2 + 4H2 = CH. + 2H 20
CH 3 0H = C + H 20 + H2
4CO + 2H 20 = CH. + 3CD 2

a. How many independent reactions are there?


b. Choose the simplest set of independent reactions.
c. If coking (formation of C) is seen as a potential problem in methanol synthesis, how would
you set up a calculation for deciding under which conditions coking will not occur at equilibrium?
d. Will coke form, other things being equal, below or above some critical pressure, or is pressure
irrelevant?

LITERATURE

We cannot but wonder at the fact that fire is necessary for


almost every operation. By fire minerals are disintegrated and
copper is produced, in fire is iron born and
by fire it is subdued, by fire gold is pUrified.
Pliny the Elder, Natural History

Of course, the literature section in Chapter 3 is relevant to the material discussed in


this chapter. Most textbooks on chemical engineering thermodynamics discuss in detail
the calculation procedure for mixture compositions at chemical equilibrium; a lucid
presentation is given by K. G. Denbigh, The Principles of Chemical Equilibrium, 3rd ed.,
Cambridge University Press (1971). Gas-liquid heterogeneous equilibria in systems involv-
ing ionic species are discussed by G. Astarita, D. W. Savage, and A. L. Bisio, Gas Treating
with Chemical Solvents, Wiley, New York (1983).
Chapter Eleven

ELECTROCHEMISTRY

The principles of thermodynamics occupy a special place


among the laws of nature. For this there are two
reasons: in the first place, their validity is subject only to
limitations which, though not, perhaps, themselves
negligibly small, are at any rate minimal
in comparison with many other laws of Nature; and
in the second place, there is no natural process to
which they cannot be applied.
W. Nemst

NOTATION

A Constant in equation (11.1.27) m1.s kmol-o.s


At Total free energy kcal
a Activity
B Constant in equation (11.1.24) m1.S kmol-o.s
c Concentration kmol m- 3
c* "True" concentration kmolm- 3
D Diffusivity m2 s- 1
E Electrical potential difference V
e The electron
F Faraday's constant kcal kmol- I V-I
G Free enthalpy of solid kcal kmol- I
Gt Total free enthalpy kcal
I Ionic strength kal m- 3
Current intensity A
K Concentration-based equilibrium constant
m Concentration of electrolyte kmol m- 3
m* Concentration of buffer kmolm- 3 291
292 n Number of electrons
Chapter Eleven n (With suffix) number of moles
N Diffusive lIux kmolm-2 s- 1
N' TotallIux kmolm- 2 s- 1
p Electrovalence of cation
pH =-log[H+]
pK =-logKp
q Electrovalence of anion
Q Electrical charge density coulombm- 3
R Gas constant kcal kmol- I K- I
S Solubility kmolm- 3
T Temperature K
Transport number
u Mobility kmol m- I cal-I s-I
U Constant defined in equation (11.1.31) kmoem-6
l' Velocity ms- I
x Extent of reaction kmolm- 3
y Mole fraction of solute
z Electrovalence
a Degree of dissociation
~ =m/m*
~ (With subscript) interaction parameter m3 kmol- 1
T Activity coefficient m3 kmol- 1
e Dielectric constant coulomb V-I m- I
p. Electrochemical potential kcalkmol- I
4> Electrical potential V
n Electrical conductivity ohm-I

Subscripts S For solvent


w For water
+ Of cation
o At potential 4>0
Of anion
Average for ion pair
b For base
Superscripts
j For component j
p For protonation o At standard state
11.1. STRONG ELECTROLYTES

Besides the known fiftyfour chemical elements,


there exists in nature only one agent more,
and this is called force; it can under suitable conditions
appear as motion, cohesion, electricity, light,
heat, and magnetism.
F. Mohr, 1837

When an electrolyte is dissolved in water (or in other polar solvents) it dissociates


into ions. The ions themselves may always be regarded as present essentially only
in the aqueous phase. The theory of electrolytic solutions is based essentially on
the approach of dilute solution theory, since the existence of a solvent is guaran-
teed. However, the theory of electrolytic solutions is peculiar in three aspects:
first, ionic reactions are extremely fast, and thus may be regarded as being always
at equilibrium. Second, the condition of electrical neutrality must always be
satisfied. Third, the behavior of solutes in electrolytic solutions, even at infinite
dilution, is somewhat different from what it is in ordinary dilute solutions of
nonelectrolytes.
Let us first consider the simple case of an aqueous solution which has been
obtained by adding to water a uni-univalent electrolyte MX, with M being the
cation and X the anion. The dissociation reaction can be written as

(11.1.1)

Let a be the "true" degree of dissociation, and let m be the kmoles of MX


added per cubic meter of solution. Quantities referring to the cation will be
indicated with a + subscript, those referring to the anion with a - subscript, and
those referring to the undissociated electrolyte with no subscript. Then the "true"
concentrations (identified by an asterisk) in the solution are

c*=(1-a)m (11.1.2)

and
c! = c~ = am (11.1.3)

Since reaction (11.1.1) is an ionic one, it can be regarded as being always


at equilibrium, and hence the equilibrium condition can be written to yield the
following equation for a:

(11.1.4) 293
294 where K is the concentration-based equilibrium constant for reaction (11.1.1).
Chapter Eleven (As discussed in Chapter 10, K is not a true constant since it contains activity
coefficients; however, this is irrelevant in the present context.) The important
point is that, as m approaches zero, the solution of equation (11.1.4) approaches
a = 1, i.e., provided the solution is sufficiently dilute, the electrolyte is completely
dissociated. The same conclusion is reached also in the general case of an
electrolyte which is not uni-univalent. The dissociation reaction can be written as

(11.1.5)

where p is the (positive) electrovalence of the cation and q the (negative)


electrovalence of the anion, with the electroneutrality of the electrolyte itself
requiring that
Pp = Qq (11.1.6)

Equation (11.1.4) now becomes

(11.1.7)

and, since P + Q > 1, the solution for m approaching zero is always a = 1.


Although the conclusion that at infinite dilution any electrolyte is completely
dissociated is valid no matter what the value of the equilibrium constant K may
be, provided it is nonzero, from a practical viewpoint there is an important
distinction to be made. Let us consider again equation (11.1.4); a will approach
the value unity when m becomes significantly smaller than K. Now if K has a
value which is very small as compared to the lowest values of m for which
experimental data can be collected easily (or the value below which one could
also regard the solution as being simply the pure solvent), complete dissociation
at infinite dilution is scarcely of any pragmatic relevance (what this lower limit
for m may be is of course somewhat arbitrary, but 10- 5 kmol m -3 is not an
unreasonable value). Thus the theory of electrolytic solutions is somewhat
arbitrarily subdivided by considering strong electrolytes (those for which complete
dissociation is in fact approached at pragmatically relevant values of m) and
weak electrolytes (for which it is not). (In actual fact, in the case of aqueous
solutions, there is a nonarbitrary lower bound of m below which it becomes
irrelevant whether the electrolyte dissociates or not. This, as will be seen later,
is due to the fact that water itself dissociates and becomes the pH-determining
electrolyte when m drops below approximately 10- 7 kmol m- 3 .)
The theory of strong electrolytes is developed by formally assuming that the
degree of dissociation is unity, and then ascribing any observed deviation from
the behavior that such an assumption would predict for the activity coefficients.
The formal assumption guarantees correctness in the limit of infinite .dilution,
and in fact in a limit which is accessible to measurement. An analogy is useful
here to fully understand this approach.
Consider the equilibrium p- V- T behavior of N 20 4 , a gas which can
dissociate according to the reaction

(11.1.8)
It is easy to convince oneself that the equilibrium degree of dissociation 295
approaches unity as pressure becomes very low and temperature very high-which
Electrochemistry
are the conditions at which ideal-gas behavior is expected. Hence, if one assumes
the gas to be completely dissociated, the observed equilibrium p- V - T behavior
will indeed correctly approach the ideal-gas behavior; it would not if one assumed
the gas to be undissociated, since in the limit the actual number of moles would
be twice that assumed. As temperature is lowered and/ or pressure increased,
deviations from ideal-gas behavior would be observed. These could be used to
infer the "true" degree of dissociation, if one is willing to assume that in fact
the mixture still behaves as an ideal gas-but there is no guarantee that that is
the case. Alternately, one could regard the gas as always completely dissociated
and ascribe all observed deviations to the fugacity coefficient (this is the approach
of the theory of strong electrolytes). Third, if one has some independent means
of ascertaining the actual degree of dissociation, "true" fugacity coefficients could
be extracted from the p- V - T data. This approach is hardly feasible in the case
of electrolytic solutions.
We now note that the discussion in the above paragraph would apply to any
gas capable of dissociation, say for oxygen, where one could consider its dissoci-
ation into atomic oxygen. This is never done, because the pressure and temperature
at which molecular oxygen approaches the behavior of an ideal gas are much
less extreme than those at which it dissociates significantly into atomic oxygen.
The difference is not qualitative but simply quantitative; the equilibrium constant
for the oxygen dissociation reaction is extremely small. The analog of N 20 4 is a
strong electrolyte; the analog of molecular oxygen is a weak electrolyte. Salts are
invariably strong electrolytes; acids and bases may be both strong and weak.
Let us first consider the theory of strong electrolytes, in particular the
problems imposed by the condition of electrical neutrality. This condition requires
that, at any point and at any time, the total concentration of positive charges
must equal that of negative charges. A rationale for this condition is discussed
in the next section. The important point is that the condition of electrical neutrality
makes the chemical potential of an ion unaccessible to measurement. The deriva-
tive 5Gt /5nj can never be measured, since it is experimentally impossible to have
two solutions which differ only in the number of moles of one particular ion; at
least one of the two would violate the condition of electrical neutrality. However,
from a formal viewpoint the concept of the chemical potential of an ion is
perfectly legitimate.
We shall examine again the simple case of a solution for which only
reaction (11.1.1) needs to be considered. The equilibrium condition can be
written as

/L = /L+ + /L- (l1.t.9)

and, as will be discussed later, /L is accessible to measurement. The constitutive


equation for the ionic chemical potentials is expressed as follows (when the same
equation applies to both cation and anion, only the first is written);

(H.t.tO)
296 with the standard state chosen so that, at infinite dilution, the activity coincides
with the concentration measured in kInol m- 3 :
Chapter Eleven

(11.1.11)

We note that c+ is the "formal," not the "true," concentration, i.e., it is


calculated on the basis of the formal assumption of complete dissociation [and
hence it simply coincides with m in the case of reaction (11.1.1)]. However, in
the limit called for in equation (11.1.11) the distinction is irrelevant. Activity
coefficients are defined as follows:

(11.1.12)

and the chemical potential in the standard state is

(11,1.13)

It is now useful to define the following average values for the two ions:

(11.1.14)

(11.1.15)

(11.1.16)

The geometric average is chosen for the concentration and the activity
coefficient, since these contribute to the chemical potential through their
logarithms. With these definitions, equation (11.1.9) becomes

(11.1.17)

which shows that the average activity coefficient of the ion pair is accessible to
measurement, though the two individual ones are not. In the case of an electrolyte
which dissociates according to reaction (11.1.5), the averages are defined as
follows:

(11.1.18)

(11.1.19)

(11.1.20)

while equation (11.1.17) becomes

(11.1.21)
Since only the average activity coefficient of an ion pair is accessible to 297
measurement (and of pragmatic interest), one can arbitrarily assign the value Electrochemistry
unity to the activity coefficient of one particular ion (the hydrogen ion is the
usual convention), and then define individual activity coefficients based on such
a convention. For instance, the activity coefficient of the chlorine ion is simply
the average coefficient for dissociated hydrogen chloride. Once this is established,
the activity coefficient of the sodium ion is such that, when multiplied by that of
the chlorine ion, it gives the square of the average coefficient of dissociated
sodium chloride; and so on. Values of individual activity coefficients of ions
reported in the literature are based on this convention.
Now let us turn our attention to the question of the behavior of the solute
in a dilute electrolytic solution. Again considering the simple case of an uni-
univalent electrolyte, the mole fraction Y of the solute is, within the first-order
approximation, proportional to its concentration:

Y = m/cs (11.1.22)

where Cs is the molar density of the solvent (55.5 kmol m- 3 in the case of water).
If'Ts is the activity coefficient of the solvent, the Gibbs-Duhem equation can be
expressed in the form

81n'Ts/8m = -[Y/O- Y)] 81n'T/ 8m (11.1.23)

As far as the solvent is concerned, the classical theory of dilute solutions


requires the left-hand side to be zero as Y approaches zero. This is guaranteed
if, as assumed in the classical theory, the derivative appearing on the right is
finite at Y = O. However, it is also true under the milder condition that that
derivative becomes infinity, provided it does not do so as rapidly as 1/ Y. The
latter is the situation encountered in electrolytic solutions. In the case of a
uni-univalent strong electrolyte, the following behavior has been well established
at sufficiently strong dilution:

In'T = -BJm (11.1.24)

where B is a constant. It should be noted that, as required, 'T = 1 at m = O.


However, the derivative is

(11.1.25)

and hence it becomes infinity as Y approaches zero, though not as rapidly as


1/ Y. Thus the behavior of the solvent is still that of the classical theory of dilute
solutions (and, in particular, it is legitimate to infer the dissociation of electrolytes
from the measurement of colligative properties). However, for the solute, devi-
ations from the infinite dilution behavior ('T = 1) are of order J Y rather than
of order Y.
Equation (11.1.24) has rather strong support in the molecular theory of dilute
electrolytic solutions, since it is based on the Debye-Huckel model of such
solutions. The leading deviations from infinite dilution behavior are predicted to
298 be first order in Y when only short-range interactions are considered, because
Chapter Eleven
the short-range interactions to which a solute is subjected are only solute-solvent
ones at sufficiently strong dilution. In contrast with this, deviations may be other
than first order when long-range interactions are taken into account; the Debye-
Huckel model is essentially a model of such long-range ion-ion interactions in
dilute solutions.
Let I be the ionic strength of the solution, which is defined as

(11.1.U)

where the quantities Cj are the concentrations of the ions and the Zj the correspond-
ing electrovalences (for the case of a solution containing only a uni-univalent
electrolyte, 1= m). The Debye-Huckel model results in the equation

(11.1.27)

where A is a universal constant. Equation (11.1.24) is, of course, a special case


of equation (11.1.27).
The Debye-Huckel equation generally correlates well experimental data up
to ionic strengths of about 10-3 lemol m- 3 In less dilute solutions deviations from
equation (11.1.21) are observed, and these can be described with techniques
analogous to those discussed in Section 8.7. In particular, constitutive equations
generally account only for binary interactions, say

(11.1.28)

where the first term on the right is the Debye-Huckel one. The interaction
parameters are often expressed as follows:

(11.1.29)

with the terms on the right regarded as constant for any given ion.
Average activity coefficients of ion pairs are accessible to measurement
through a variety of techniques: solubility measurements (in the case of elec-
trolytes which are solid if pure, as indeed is the case with most salts), vapor
pressure measurements (in the case of volatile electrolytes such as Hel), measure-
ment of the electromotive force of electrochemical cells (i.e., of the val~e E*
discussed in Section 3.5), and nonequilibrium measurements such as diffusion.
A specific case is discussed here, that of AgI03 If a solution of AgI0 3 is in
equilibrium with the pure solid salt, the chemical potential of AgI0 3 in solution
is equal to the free enthalpy density of the pure solid, say

(11.1.30)

Even if the value of'O is known, that of p.~ is not, so equation (11.1.30)
may be rewritten as
(11.1.31)
with U an unknown constant. Let S be the solubility of AgI0 3 , i.e., the number 299
of moles of salt dissolved into the solution at equilibrium; S is measurable. With Electrochemistry
the formal assumption of complete dissociation, one obtains

(11.1.32)

and hence
(11.1.33)

Now suppose the solubility is low enough that the resulting ionic strength
is within the limits of validity of the Debye-Huckel model. Equation (11.1.33)
can be expressed as

log S = Av'I + log v'U (11.1.34)

It is now possible to perform experiments at different ionic strengths, by


adding another salt which does not have any ion in common with AgI03 , such
as KN0 3 At sufficiently low ionic strengths, a plot of log S vs. v'I should yield
a straight line of slope A; this is indeed observed experimentally (hence the
experimental confirmation of the Debye-Huckt,l model). This straight line can
be extrapolated to zero ionic strength to yield the value of U. At any finite ionic
strength, the average activity coefficient can now be calculated from equation
(11.1.33).
It is noteworthy that the experimental point at the lowest possible ionic
strength corresponds to the case where there is no KN0 3 , say the point at which
the ionic strength in fact coincides with the solubility. Since an extrapolation to
zero ionic strength is needed, the method described is not suitable for highly
soluble salts, for which the data point corresponding to the lowest ionic strength
would be outside the validity of the Debye-Huckel model.
The added salt needs to be one which does not have any ion in common
with AgI0 3 , since otherwise equation (11.1.32) would be violated. For example,
let us consider the case where the added salt is KI0 3 and m the kmol m- 3 of
KI0 3 added. Hence

(11.1.35)

c=S+m (11.1.36)

I=!(S+m) (11.1.37)

and the equation relating S and I becomes

! 10g[SI/ U] = Av'I (11.1.38)

The quantity S2 is called the solubility product of AgI0 3 ; it represents the


product of the concentrations of the two ions in equilibrium with the pure solid.
If one of the ions is present also as a result of the dissociation of other electrolytes
(such as 10; in the previous example), the concentration of the coion at equili-
brium must be less than S; an electrolyte with a common ion depresses the
solubility.
300 11.2. ELECTROCHEMICAL POTENTIALS
Chapter Eleven
"r know what you're thinking about" said Tweedledum;
"but it isn't so, nohow." 'Contrariwise" continued Tweedledee
"if it was so, it might be; and if it were so, it would be;
but as it isn'l, it ain't. That's logic.
Lewis Carroll

The Poisson equation for the spatial distribution of the electrical potential
q, is

divgrad q, = -Q/ E (11.2.1)

where Q is the local electrical charge density (coulomb m -3) and E is the dielectric
constant (coulomb V-I m- I ). The Faraday constant F is the charge of a
kequivalent of ions, i.e., the charge of one kmole of a univalent ion. Hence the
local charge density is given by

(11.2.2)

and the Poisson equation becomes

(11.2.3)

The value of F is 9.65 x 107 coulomb per kequivalent. The dielectric constant
of aqueous electrolytic solutions is of the order of 10-9 coulomb V-I m- I Hence
the coefficient F / E in equation (11.2.3) is at least 1017 V m per kequivalent.
Now if the sum appearing in equation (11.2.3) is of the order of, say, to- 1O
kequivalent per m\ then such a value would be extremely small as compared to
the values of any term in the sum-unless the ionic concentrations are so small
as to be of no practical importance whatsoever. It follows that, if the concentra-
tions of all the ions except one are known, the latter could be calculated by
assuming that the sum is zero, i.e., one would make use of the condition of
electroneutrality, as discussed in the previous section. However, the right-hand
side would still be of the order of t07 V m- 2 , i.e., a very steep curvature of the
electrical potential distribution could exist. It follows that, while the condition
of electroneutrality can in fact be regarded as universally true (except near
electrode-solution interfaces), the electrical potential does not have a Laplacian
distribution (i.e., divgrad q, is not zero). The electroneutrality condition is violated
significantly [say the sum appearing in equation (11.2.3) is of the order of
to-I kequivalent per m 3 ] when divgrad q, is of the order of tO I6 V m -2. Since the

electrical potential differences in electrochemical systems are of the order of a


few volts, this implies a distance of the order of 10-8 m over which the elec-
troneutrality condition may be violated. This indeed happens near electrode-
solution interfaces, and is generally described by appropriate boundary conditions
at such interfaces (the literature usually refers to an "electrical double layer"
near such interfaces). In the bulk of the electrolytic solution, the electrQneutrality
condition can always be used in calculations involving the actual ionic
concentrations.
Now let us consider a solution of electrolytes at some uniform reference 301
electrical potential CPo. Quantities in this condition are indicated with a 0 subscript. Electrochemistry
Now consider the same solution at some other electrical potential cpo Other
conditions being equal, its free energy is increased with respect to the reference
potential by an amount of electrical energy which is the product of the electrical
charge times the potential difference:

(11.2.4)

and of course the same equation applies also to free enthalpy. Hence, the
electrochemical potential of the jth ion is given by

(11.2.5)

where the second term is as legitimate a concept as the first, since both imply
only consideration of a differentially small variation of nj, which is not forbidden
by the electroneutrality condition since the latter holds in the sense discussed in
the previous paragraph.
Now we consider a situation where there are finite gradients of the
electrochemical potentials, so that we expect finite diffusive fluxes of the ions.
Let Nj be the total flux of the jth ion, which includes the diffusive flux as well
as the convective flux,

(11.2.6)

where v is the average velocity discussed in Section 7.3. The vector of current
intensity is
(11.2.7)

and, in view of the condition of electroneutrality, the convective fluxes contribute


nothing to the current intensity (the flow of an electrically neutral solution does
not by itself carry electricity):

(11.2.8)

In order to proceed, one needs to write constitutive equations for the diffusive
fluxes. The theory of transport processes in electrolytic solutions is generally
developed by neglecting the possibility of direct coupling at the constitutive
equations level (as will be seen, strong coupling arises anyhow), so that the flux
of the jth ion is given by

(11.2.9)

where the quantities Uj are called the mobilities. Substitution into equation 01.2.8)
gives

(11.2.10)
302 Let us first consider the case where there is no net electrical current flowing
through the solution (electrochemical cells are thus excluded from consideration).
Chapter Eleven
If the concentrations of the ions are not uniform in space, the second term on
the right-hand side of equation (11.2.10) is nonzero, and hence also the first term
must be nonzero. In an electrolytic solution, when nonzero concentration
gradients are present, a gradient of electrical potential may develop. The physical
interpretation of this conclusion can be better understood by considering equation
(11.2.7) which, for the case of zero intensity, yields the condition

(11.2.11)

If, for instance, the solution only contains a strong uni-univalent electrolyte
MX, then for this case equation (11.2.11) simply requires the fluxes of the two
ions to be equal. The condition of electroneutrality requires the concentrations
of the two ions to be the same everywhere, and hence so are the concentration
gradients. If the diffusivities of the two ions are not equal to each other, the first
term on the right-hand side of equation (11.2.9) would be different for the two
ions, and hence the two fluxes can only be equal if a gradient of electrical potential
develops which slows down the flow of the ion with a higher diffusivity, and
accelerates that of the other one. This shows that, even neglecting direct coupling
at the constitutive equation level, strong coupling between the diffusive fluxes of
different ions arises anyhow.
The relationship between diffusivities and mobilities can be established as
follows. Equations (11.2.9) can be rewritten as

(11.2.12)

and thus one has

(11.2.13)

which, for solutions which are dilute enough for activities to coincide with
concentrations, reduces to

(11.2.14)

This is known as the Nernst-Einstein equation.


Let us now return to the general case where there is a finite current intensity
through the solution. First we consider the special case where there are no
concentration gradients in the solution (this can be achieved in practice by forcing
an alternate current through the solution, so that significant concentration
gradients have no chance of developing). Equation (11.2.10) now reduces to the
ordinary formulation of Ohm's law, with the electrical conductivity n identified
with

(11.2.15)

It is customary to define a transport number tj as follows:

tj = p2ZJUJ'j/il (11.2.16)
which, in the case where there are no concentration gradients, can be interpreted 303
as the fraction of the total current carried by the jth ion. The definition, of course,
Electrochemistry
is. still legitimate even when there are concentration gradients, but the interpreta-
tion is more vague.
Experiments carried out with alternating current afford a means of measuring
the value of n. For dilute solutions, the values of the mobilities can be estimated
from equation (11.2.14), and hence the measurement of n yields an estimate of
the order of magnitude of the ionic concentrations. It is this type of measurement
which leads to the consideration of weak electrolytes, for which the measured
electrical conductivity is very significantly less than what one would estimate
from the assumption of complete dissociation. Hence the conclusion is reached
that such electrolytes are only very weakly dissociated, resulting in ionic con-
centrations several orders of magnitude smaller than the amounts of electrolyte
present in the solution, even for very dilute solutions.

Appendix

In order to convince oneself of the impossibility of measuring individual


activity coefficients of ions, it is useful to consider how an actual experiment
would have to be set up. One would need at least one solution which violates
the electroneutrality condition by some measurable amount. Let us assume the
lowest level of measurable concentration differences is 10-7 kmol m- 3 Now it
turns out that the best one can do is to have as small a sample as possible; say
one conducts an experiment with a 1 cm3 sample. This corresponds to a total
unbalance of charges of 10- 10 gions which, when multiplied by the Faraday
constant, gives a total electric charge of 10- 5 coulomb. If the sample is regarded
as a sphere in a vacuum (which cannot be too far from reality), one calculates
a potential of 107 V, which is.hardly attainable. We note that even higher electrical
potentials would be obtained if the sample is larger, or if the ionic concentration
unbalance is higher.

11.3. WEAK ELECfROLYTES

With characteristic German thoroughness they have classified


everything as masculine, feminine and neuter, and they have
made a horrible mess. German girls (Madchen) and unmarried
women (Fraulein) are neuter, walls (Wand) are feminine.
What is there for a young man to do in Germany? All young
and unmarried women are sexless. If he wants feminine
company he must find himself a married woman or go to the
wall. Is it any wonder that German youth has been responsible
for so many upheavals in the past century?
H. J. Lipkinsey

In this section, the problem of calculating the composition of aqueous


solutions of weak electrolytes is considered. We begin by restricting attention to
the case of a solution containing only one electrolyte, which in a sense is a more
difficult problem than that of mixed electrolytes. As discussed in Section 11.1,
304 ionic reactions can be regarded as being always at chemical equilibrium, and
hence in principle the problem to be examined is nothing else but a standard
Chapter Eleven
problem of chemical equilibrium. There are, however, two peculiarities in the
case of solutions of weak electrolytes:
1. By the very fact that the electrolytes considered are weak, the concentra-
tions of at least some of the ions will be several orders of magnitude smaller
than those of other dissolved species. In a classical problem of chemical equili-
brium of nonionic species, the components known to be present at equilibrium
in very small concentrations could be neglected altogether. This is not the case
for electrolytic solutions, where the concentrations of some ions, even if negligibly
small as compared to those of other species in solution, however, have major
influence on the macroscopic behavior of the system.
2. Not only are the equilibrium constants of the reactions involved often
very small [which is essentially what was discussed in (1) above], they often
differ from each other by several orders of magnitude. This implies that the
complete system of equilibrium equations may be very stiff and numerical solution
may present major difficulties. It follows that techniques of approximation are
often needed before the problems can actually be subjected to numerical solution.
Since both peculiarities discussed above are related to order of magnitude
effects, the discussion in this section is restricted to the situation where concentra-
tion-based equilibrium constants are regarded as constant. It follows that the
results obtained are strictly valid only for very dilute solutions, for which activities
and concentrations coincide. For less dilute solutions, numerical solutions are
needed which take into account activity coefficients, and hence the fact that the
concentration-based equilibrium constants are not really constant. However, the
latter effect is not an order of magnitude effect, and therefore the results in this
section provide an initial guideline in any event.
We begin by treating an aqueous solution containing m kmol m- 3 of a weak
acid HX, such as acetic acid. Water is itself a weak electrolyte which dissociates
according to
(11.3.1)

with equilibrium constant given by


(11.3.2)

where square brackets indicate concentrations measured in kmol m- 3 (It should


be noted that, consistently with the dilute solution approximation, the water
concentration is regarded as constant.) The dissociation of the acid itself is

(11.3.3)

and it is regarded as weak if the eqUilibrium constant for the dissociation

(11.3.4)

is very small, e.g., less than 10-2 kmol m- 3 The equilibrium constant Kp is called
the protonation constant and, as will be seen, it is also defined, with appropriate
qualifications, for bases and salts.
Since concentrations of different species may have very different orders of 305
magnitude, it is useful to work with their logarithms. For the special case of the Electrochemistry
hydrogen ion the definition of pH is

(11.3.5)

where the minus sign is chosen so that values of pH are almost invariably positive.
The pK of the acid is defined as

pK = -logKp (11.3.6)

The problem could be set up in perfect analogy with classical chemical


equilibrium problems by introducing the extents of the reactions (11.3.1) and
(11.3.3). However, it should be noted that every ionic reaction is electrically
neutral, so that one of the extent of reaction conditions can be substituted by
the requirement of electroneutrality. In general, the reaction subspace is identified
by electroneutrality and by a sufficient number of mass balance constraints; this
is perfectly equivalent to the extent of reaction approach. For instance, for the
case at hand there are four components to be considered: HX, H+, X-, and OH-,
hence the composition space is four-dimensional. The extent of reaction approach
would be set up as follows:

[HX] = m -XI (11.3.7)

[H+] = XI + X3 (11.3.8)

[OW] = X3 (11.3.9)

[X-] = XI (11.3.10)

which restrict the possible composition to the two-dimensional reaction subspace


of the two extents of reaction. Equivalently, one can write the electroneutrality
condition and a balance of X atoms to obtain

[H+] = [X-] + [OW] (11.3.11)

[HX] + [X-] = m (11.3.12)

an equivalent set of linear equations which reduce the composition space to the
reaction subspace. The algebra involved is often simpler with the second approach.
The equations which identify the reaction subspace are linear in the con-
centrations, but the equilibrium equations are not; conversely, if the logarithms
of the concentrations are used, the equilibrium equations become linear, but the
former ones become nonlinear. Suppose that, for the example at hand, the extent
of reaction approach has been chosen, and consider the case where pK = 4. The
two equilibrium equations become

(1l.3.13)

and
(11.3.14)
306 This corresponds to a fourth-order polynomial, and the case considered is
sufficiently simple for the solution to present no particular difficulties. We note,
Chapter Eleven
however, that the fact that the coefficients appearing in the two equations differ
by ten orders of magnitude introduces some stiffness even in this very simple case.
The method of approximate solution which is easiest to follow is to draw a
"characteristic plot" of the logarithms of the concentrations of interest vs. pH,
i.e., to seek a solution that is parametric in pH. Consider such a plot shown in
Figure 11.3.1. The lines for H+ and OH- are immediately obtained as simple
straight lines. Now we let a be the degree of dissociation of the weak electrolyte
(which is expected to be small), say

(11.3.15)

The equilibrium equation for the acid dissociation can be written as

10g[a/(1- a)] = pH - pK (11.3.16)

and hence, as long as pH is significantly less than pI{, a 1, and equation


(11.3.16) can be approximated by

log a = pH-pK (11.3.17)

which gives the straight line on the left of Figure 11.3.1. We note that 10g[OH-] =
pH - 14 and hence, provided pK is significantly less than 14, as long as equation
(11.3.17) holds it will result in a value of [X-] significantly larger than [OH-].
Hence the intersection of the 10g[X-] and 10g[H+] lines, that indeed takes place
at a pH significantly less than pK, is the solution of the problem since the
electroneutrality condition is (approximately, but very closely) satisfied at that
point. The intersection is given by
2pH = pK - log m (11.3.18)

u
CI
pH::V2 (pK-Iogm)::2.25
~
logm
-1

-3

-s
-7
0 2 4 6 8 10 12 14 pH

FIGURE 11.3.1. Characteristic plot for a weak acid.


and hence a plot of measured pH vs. log m should yield, at sufficiently low values 307
of m, a straight line of slope! from which the value of pK can be extracted.
Electrochemistry
The case of a weak base B, such as ammonia, can be analyzed along very
similar lines. A weak base may not have a sheddable hydroxyl ion (ammonia
does not), and hence in general the dissociation reaction can be written as

(11.3.19)

with equilibrium constant K b , which is expected to be small in the same sense


as Kp was expected to be small in the previous case, if the base is weak. Instead
of reaction (11.3.18), one can consider the equilibrium of the following reaction:

(11.3.20)

which is called the protonation reaction of the base, with equilibrium constant
Kp given by

(11.3.21)

If Kb is 10-4 kmol m -3, so that in a sense the base is as weak in this example
as the acid was weak in the previous one, the pK would be 10. This shows how
the distinction between weak bases and weal acids is somewhat arbitrary. One
could regard as a base any weak electrolyte with a pK in excess of 7, and as an
acid anything with a pK less than 7, water being left to represent the borderline;
this is, however, totally arbitrary.
Again a characteristic plot such as in Figure 11.3.2 is useful. The degree of
dissociation a can be introduced again, with [B] = m(1- a) and [BH+] = rna,
so that the equilibrium condition for reaction (11.3.20) becomes

10g[a/(1- a)] = pK - pH (11.3.22)

u
Ol
pH=7+1/2 (pK+logm)=11.75
~
log m
-1

-3 BH+

-5 :
I
I

10<:
:0.
-7 I
I

0 2 4 6 8 10 12 14 pH

FIGURE 11.3.2. Characteristic plot for a weak base.


308 As long as pH is significantly in excess of pK, a I, and hence one obtains
again a straight line (this time on the right of the figure, and with negative slope)
Chapter Eleven
to represent the BH+ concentration. Its intersection with the OH- line again
represents very closely the solution to the problem, since the line for BH+ is well
above that for H+. The point of intersection is represented by

2pH = 14 + pK + log m (11.3.13)

which again is the basis for the experimental determination of the pK


Next we consider the case of a salt MX resulting from the neutralization of
a strong base MOH and a weak acid HX. Since both MX and MOH are strong
electrolytes, the cation M+ will never be present in any nondissociated form, say

(11.3.24)

and it simply represents a fixed concentration of positive charges that needs to


be neutralized by as many negative charges. The anion X-, however, can reassoci-
ate according to

(11.3.25)

which is the exact analog of reaction (11.3.19), i.e., the dissociation reaction of
a weak base. In the presence of the cation of a strong base, the anion X- behaves
as a weak base.
Instead of reaction (11.3.25), one could consider the equilibrium of reaction
(11.3.3). With this, the problem would be set up exactly as in the case of a weak
acid, except that the electro neutrality condition is now

(11.3.26)

Since the term [M+] is fixed, it is best to eliminate it between equations


(11.3.12), (11.3.24), and (11.3.26) to obtain

(11.3.27)

It is now an easy matter to construct again a characteristic plot and go


through the same argument as before, with now a the degree of reassociation of
the anion, [HX] = am, to convince oneself that the solution occurs at a very low
value of a, yet at conditions where [HX] [H+]; one thus obtains again equation
(11.3.23). The behavior is strictly analogous to that of a weak base, except that
in equation (11.3.23) the pK is now that ofthe weak acid HX. At m = 1 kmol m- 3,
a weak base with a pK of 10 would give, according to equation (11.3.23), a pH
of 12, while a salt MX with HX having a pK of 4 would result in a pH of 9~
lower (hence less basic) but still in excess of the "neutral" value 7. The characteris-
tic plot is given in Figure 11.3.3.
The examples discussed so far could be solved more easily by the following 309
"rule of thumb." Given an electrolyte, one immediately knows whether its
Electrochemistry
dissolution in water will result in an acidic (pH < 7) or basic (pH> 7) solution.
If it is a salt of a strong base and a weak acid, it will act as a weak base; if a
salt of a weak base and a strong acid, as a weak acid. For acidic solutions, the
OH- concentration can be neglected in the electroneutrality equation, and for
basic solutions the H+ concentration can be neglected. For salts, it is always
useful to eliminate the strong ion concentration from the governing equations.
A significantly more complex case is that of polyfunctional weak acids, such
as phosphoric acid. Let us consider a three-functional acid H3X (in the case of
phosphoric acid, the anion X--- would be identified with PO;--), which can
undergo three subsequent dissociations, each characterized by its own pK:

(1l.3.28)

and so on. For phosphoric acid, the three successive pK values are 2.23, 7.21,
and 11.32. In principle, in a solution containing m kmol m- 3 of H3X, there are
thus four chemical equilibria to consider: the three acid dissociations and the
water dissociation. This is equivalent to a sixteenth-order polynomial equation,
containing four coefficients the values of which span 12 orders of magnitude.
This is an extremely stiff nonlinear problem, and an approximate preliminary
solution is not only desirable but indispensable.
Again, it is useful to consider a solution scheme which is parametric in the
pH. Let ao, ... , a3 represent the fractions of m present in the form of
H 3X, ... , X---. The following three relationships between the quantities a are
easily established:

(1l.3.l9)

which, together with the condition that the quantities a sum to unity, gives the
desired parametric solution; this is plotted, for the case of phosphoric acid, in

u
t7I
.

-1

-3

-5

-7
0 2 4 6 8 10 12 14 pH

FIGURE 11.3.3. Characteristic plot for the salt of a weak acid and a strong base.
310 Figure 11.3.4. It can be observed that at no value of the pH is the value of more
than two quantities a nonnegligible; hence in every pH range the acid can in
Chapter Eleven
fact be analyzed as a monofunctional one. If a calculation is conducted for an
m value of 0.1 lanol m- 3 (or larger), we can start by assuming that there is only
H3X and H2X-, and indeed the calculation confirms this hypothesis. In this sense,
phosphoric acid can be regarded as a weak electrolyte. However, if the same
calculation is performed at a concentration level of 10-4 lanol m-3, the same
hypothesis leads to a paradox, and the solution actually results in a pH range
where the only species of interest are H2X- and HX--. There is no undissociated
phosphoric acid at this low concentration, i.e., phosphoric acid behaves as a
strong electrolyte. This confirms the earlier statement that the distinction between
strong and weak electrolytes is somewhat arbitrary, and the classification of any
given electrolyte depends in some measure on the level of total concentrations
under investigation.
Indeed, the above considerations apply to any solution of a weak acid or
weak base. For instance, in the case of HX discussed at the beginning, equation
(11.3.18) in fact holds only as long as pH is less than pK, or equivalently, as
long as log m is larger than -pK. At lower concentrations, the acid is in fact
completely dissociated (i.e., it behaves as a strong electrolyte), and pH = -log m.
In contrast with this, the anion X- of a salt MX behaves always as a weak base;
equation (11.3.23) holds down to log m = -pK. At even lower concentrations,
water itself becomes the electrolyte which determines the acidity of the solution,
and pH = 7. (This is a general result: whenever the total concentration of elec-
trolytes in an aqueous solution is less than 10-7 lanol m- 3, their presence can be
entirely neglected. However, this is an exceedingly low concentration, which is
seldom encountered except in the laboratory.)
The shape of the curves in Figure 11.3.4 is related to the fact that the
successive pK values of the phosphoric acid dissociations differ significantly
from each other. While this is always true in the case of polyfunctional inorganic
acids, it is not in the case of polyfunctional organic acids. For the latter, the full

1.0
OV i

0.5

FIGURE 11.3.4. Fractional dissociations of a three-functional acid like phosphoric acid.


system of equilibrium equations must be solved. However, the numerical problem 311
is not a stiff one, because the coefficients appearing in the equilibrium equations
Electrochemistry
do not differ from each other by several orders of magnitude.
Possibly the most complex calculation is that relative to a solution of the
salt of a strong alkali and a polyfunctional acid, such as, say, MH 2 X, with H3X
being an inorganic polyfunctional acid and MOH a strong alkali. For such a
calculation, it is useful to first construct the characteristic plot for the acid, such
as in Figure 11.3.4. The pH of the salt solution is most likely in the range where,
in such a characteristic plot, the concentration of the salt's anion (for the example
at hand, H 2 X-) is highest. This value can be assumed in a preliminary calculation,
and corrections applied later.
Before discussing the case of solutions of more than one electrolyte, the
following observations are useful. In the characteristic plots we have typically
two curves for the concentrations of a dissociated and an undissociated form
which sum to a value m. These two curves are approximately straight lines in
the region sufficiently far from the "characteristic point" (which has ordinates
pK and log m) where the straight lines cross each other. At one pH unit to the
right or left of the characteristic point, the actual curve which represents the full
solution of the problem is displaced from the straight line by 0.04 units on the
log c scale; at the characteristic point itself, the displacement is at its maximum
(0.3 units). It follows that, for an order of magnitude analysis, the two straight
lines could often be taken to hold up to the characteristic point, and the actual
solution curve can thus be approximated by those two straight lines.
Let us now consider again the case of a weak acid HX (for which, by
definition of weak acid, pK < 7). Equation (11.3.18) holds down to a log m value
equal to -pK; at lower acid concentrations, HX can in fact be regarded as a
strong electrolyte, and the pH is simply equal to -log m. As the value of m is
further decreased down to log m < -7, the pH is dominated by the water dissoci-
ation, i.e., pH = 7. This shows that even an aqueous solution of only one electrolyte
is in actual fact a solution of two electrolytes, water being one of the two. The
pH vs. log m curve has the shape sketched in Figure 11.3.5a.
Conversely, in the case of a weak base B equation (11.3.23) holds down to
a log m value of pK - 14 (we note that 14-pK is a measure of the basicity of B
just as pK is a measure of the acidity of HX). Again, at lower m values the base
is a strong electrolyte (pH = 14 + log m), and at log m < -7 water becomes the
dominating electrolyte. The pH vs. log m curve is sketched in Figure 11.3.5b.
In the case of a weak base MX, equation (11.3.23) still holds, and in principle
the electrolyte would become strong at log m < pK-14. However, since pK < 7,
this condition is reached only after water has become the dominating electrolyte
at log m = -4, so that the pH vs. log m curve has the shape sketched in Figure
11.3.5c-MX never becomes a strong electrolyte. Analogous considerations apply
to the salt resulting from the neutralization of a weak base and a strong acid
(which acts as a weak acid, but never becomes a strong electrolyte).
The possibility that, in a solution of a single weak electrolyte, the electrolyte
may actually behave as a strong one at sufficiently low concentrations, implies
that in the analysis of such solutions the concentrations of the hydrogen ion and
of the hydroxyl ion can never both be assumed to be negligible a priori. In the
case of solutions of mixed electrolytes, on the other hand, it is generally possible,
312
pH (PH =7; dominated by water
Chapter Eleven

6 iPH=-IOg m; strong electrolyte

4
r PH =1I2(pK-logm) i
weak electrolyte
2

-10 -8 -6 -4 -2 o

., , ,., ,
pH

14 pH=7+1/2 (pK+logm) j
~U.
12

-10 -8 -6 -4 -2 o logm

pH

11 C. . .
pH =7+1/2 (pK+ log m);

",,'~.

I
9 pH = 7; dominated
by water

7 I-~ _ _ _ _ _"'-

-10 -8 -6 -4 -2 o logm

FIGURE 11.3.S. pH vs. concentration plots: (a) a weak acid; (b) a weak base; (c) a salt
of a weak acid and a strong base.
313
u
.9 rPH=PK \og[M+] r\Og(2m.) Electrochemistry

-1
-::::-------~---------------~-----------------
-3

-5

FIGURE 11.3.6. Characteristic plot for a buffer solution.

except at extremely low concentrations, to do so; this results in a very considerable


simplification of the calculation.
Indeed, we first consider the case of a solution of a salt MX and of its acid
HX, both present at the same concentration m*. The characteristic plot is given
in Figure 11.3.6 for the case where log(2m*) > -pK. The solution is obviously
at pH = pK, i.e., such a solution has (to within neglecting activity coefficients)
a pH which is independent of the concentration m*. When log(2m*) < -pK the
solution is simply pH = 7, and the pH vs. log m curve has the shape sketched in
Figure 11.3.7. It is noteworthy that, when log(2m*) > -pK, [HX] "" [X-] "" m*,
i.e., MX is completely dissociated and HX not at all, and both [H+] and [OH-]
are negligibly small. Such a composition could be obtained almost by inspection
as a result of considering the rule of thumb that, in solutions of mixed electrolytes,
the concentrations of hydrogen ion and hydroxyl ion are negligible. Solutions of
this type are called buffer solutions, since the pH is approximately constant even
upon addition of a third electrolyte, provided the concentration of the latter is
somewhat less than 2m*. We may consider the case where a concentration m of

pH
8

pH= pK I
4 ---------------+-~---------
1
I
I
1
2 1
",I
o.i
I
o
log (2m)
-10 -8 -6 -4 -2 o
FIGURE 11.3.7. pH vs. concentration for a buffer solution.
314
u 1+~
Chapter Eleven C1> logm
('PH=PK+I09 1-/3
.2 ("IOg(2m)

-1

-3
BH+

-5
><:
0.,
-7 ,
I

0 2 4 6 8 10 12 14 pH

FIGURE 11.3.S. Characteristic plot for a buffer solution with some base added.

a weak base B is added to the MX/HX solution, and (3 is the ratio m/m*. The
characteristic plot is sketched in Figure 11.3.8, and the pH is calculated as

pH = pK + log[(l + (3)/(l- (3)] (11.3.30)

and is thus seen to differ little from the pH of the pure buffer as long as {3 is
significantly less than unity. We note also that the pH depends on {3, but not on
the basicity of the third electrolyte, B.
Since in solutions of mixed electrolytes the concentrations of both the
hydrogen ion and the hydroxyl ion are negligible, the calculation of their composi-
tion can be set up as a standard chemical equilibrium problem, and only ionic
reactions which do not include H+ or OH- need be considered. The cases
discussed in Section 10.6 are of this type. There is no problem of stiffness of the
resulting equations, since all components involved are present in appreciable
concentrations. Once the composition has been obtained, the pH can be calculated
from the equilibrium of any ionic reaction involving two ions whose concentration
has been determined as well as H+. Let us consider, for instance, the first example
in Section 10.6. Once the chemical equilibrium problem has been solved, the
CO:;-- and HCO:;- concentrations are known. The pH can then be calculated from
the equilibrium condition of the reaction

(11.3.31)

11.4. ELECTROCHEMICAL REACTIONS

One does not find in the circuit all the heat developed by
chemical action ... part of it apparently serves to overcome
some resistance, on the exact nature of which I do not dare
to advance any hypothesis. One must therefore admit
that part of the motive power arising from the chemical agents
does not contribute to the production of that useful
electrical work one tries to realize in electromotors.
P. A. Favre, 1854
By electrochemical reactions we mean those reactions which can only take 315
place when a finite lliectrical current is flowing through the electrolytic solution. Electrochemistry
In particular, the electrode reactions are those which take place at the interface
between the solution and an electrode. Such reactions need not be electrically
neutral, since the electric charge liberated (or absorbed). by them can be carried
away by the electrode through the external electrical circuit of the cell. Formally,
the electroneutrality of these reactions can be maintained by considering the
electrons themselves as reactants and/ or products.
Electrode reactions can be classified in two main categories. The first is that
of redox reactions, which liberate or absorb one or more electrons, but do not
exchange metal ions with the electrode. The stoichiometry of a redox reaction
can be written as follows:

(ItA.I)

where n is the number of electrons released by the reaction. The balance of


electrical charges imposes the following restriction on the stoichiometric
coefficients:

(ItA.2)

The second category of electrode reactions comprises those where metal ions
are actually exchanged with the electrode. They can always be expressed as a
sequence of redox reaction plus a reaction of the type

M = MZ++ze (ItA.3)

where z is the (positive) electrovalence of the metal ion. We note that, at any
one time, the overall condition of electroneutrality of the solution needs to be
respected, and hence for any electron released by the solution to an electrode,
one must be released by the other electrode to the solution. Furthermore, except
in a layer of molecular size near the metal-solution interfaces, the electroneutrality
condition must be satisfied everywhere in the solution. .
Electrons and metallic ions are the only species which can be exchanged
between the solution and the electrodes. In this sense, the electrode interface is
an ideal semipermeable membrane, which allows only these two species to
permeate. (The statement should be qualified in the case of ~ydrogen, which can
in fact.permeate some metals.) The electrode-solution equilibrium conditions
are therefore easily established by requiring the electrochemical potential of the
permeating species to be the same in the two phases. Furthermore, the electrode
has a fixed composition, and hence the electrochemical potentials in the electrode
phase are simply given by

I-'ion = 1-'0 + ZFcPM (llAA)

and
(llA.5)

where subscript M identifies the metal and subscript E the electron.


316 In particular, an ion-exchange reaction is examined. The electrochemical
potential of the metal ion in solution is given by
Chapter Eleven

(11.4.6)

where the subscript S identifies the solution. Hence the equilibrium condi-
tion is

cf>M - cf>s = cf>~/s + (RTf zF) In(cr) (11.4.7)

where the term cf>~/s is called the Galvani potential and, at any given temperature,
is a constant characteristic of the electrode. It should be noted, however, that
the electrical potential difference between an electrode and a solution can never
be measured: in order to carry out a measurement, one would necessarily need
another metal to pick up the solution potential, and this metal would act as an
electrode and establish its own electrical potential difference with the solution.
It will be seen later that this difficulty is circumvented by arbitrarily assigning
the value zero to one particular electrode-solution pair.
The junction between two (different) metals is now treated. The only species
which can permeate through the junction are the electrons, and thus the electrical
potential difference between two metals is, at any given temperature, a constant
depending only on the nature of the two metals. It is easy to convince oneself
that this has an important consequence. Given a chain of several different metals,
the equilibrium electrical potential difference between the first and the last is the
same as would exist if the first and last metal were directly linked to each other.
It follows that one need not worry, at equilibrium, about the details of the external
circuit of an electrochemical cell; only the nature of the two metals constituting
the electrodes is of importance, and these two can be regarded as being directly
linked together. Quite obviously, these considerations do not apply to nonequili-
brium conditions, Le., when a finite current is flowing through the metals. Suffice
it to state that the electrical potential within a given single metal is not constant
when a current is flowing. The argument is also restricted to an isothermal metal
chain; thermocouples and Peltier effect coolers work on the basis of the fact that,
in a nonisothermal metal chain, potential differences develop which are not
balanced when the chain is closed into a loop.
Let us now consider the case of an electrode at which a redox reaction may
take place. The only permeable species are the electrons, and hence the equili-
brium condition is simply the equality of the electrochemical potential of the
electron in the metal and in the solution. This leads to the equation

(11.4.8)

Equation (11.4.8), in analogy with equation (11.4.7), gives the relationship


between electrical potential difference and solution composition; again, the actual
electrical potential difference can never be measured. Also, we note that both
equations (11.4.7) and (11.4.8) give the electrical potential difference for a given
electrode-solution pair, or (equivalently) for a given electrode reaction.
The electrode solution pair which is chosen arbitrarily as the reference is a 317
platinum electrode at which the following redox reaction may take place: Electrochemistry

(1l.4.9)

The solution is an aqueous solution of HCI, at a concentration of 0.1 kmol m- 3 ,


a temperature of 25 C, and a hydrogen partial pressure in the gas phase of 1 atm.
The electrical potential difference at this "standard hydrogen electrode" is
assigned the value zero.
Once this convention is established, it is possible to perform the following
equilibrium experiment (equilibrium is easily accomplished by setting the resist-
ance of the external circuit essentially equal to infinity). Let S' be the solution
of the standard electrode potential and S the solution with respect to which the
standard electrical potential of some electrode M must be established. The circuit
consists (apart from other metals which may be present in the external circuit
but contribute nothing to the electrical potential difference at equilibrium) of the
platinum electrode connected through an infinite resistance voltmeter to the M
electrode, the solution S, a connection between the latter and the solution S' (the
connection will be discussed later), and solution S'. The voltmeter reading E* is
given by

E* = EPt/M + E M/S + E s/s' (1l.4.10)

since ES'/Pt has been assigned the value zero. If the experiment can be performed
under conditions where the SIS' electrical potential difference is zero, the
measured value of E* is called the standard electrochemical potential of the MIS
pair.
Since the Sand S' solutions are different, current (though microscopically
small under the conditions of the experiment) can flow between them only by
actual flux of ions, and this could give rise to an SIS' electrical potential difference
due to the diffusive coupling discussed earlier. The problem is serious, since one
of the ions involved is the hydrogen ion, which has a much higher mobility than
that of most other ions. In addition, if the two solutions are brought into direct
contact, the irreversible neutralization of metal ions by molecular hydrogen could
take place. Therefore, Sand S' need to be kept separated, and this is usually
accomplished by what is called a salt bridge. A salt bridge is a tube containing
a gel in which a salt is dissolved whose anion and cation have very close mobilities.
With this device, the SIS' electrical potential difference can be made negligibly
small as compared to the other potential differences involved.
Once the standard electrochemical potential of an MIS pair has been estab-
lished, the value of E* for a cell with two different electrodes in the same solution
is obtained by the difference. It is important to realize that two different conven-
tions concerning the sign to be assigned to standard electrochemical potentials
are found in the literature. Of course, the whole discussion above is only concerned
with the equilibrium behavior of electrochemical reactions.
318 EXAMPLES AND PRO~LEMS

Chapter Eleven
Problems

11.1. Draw the characteristic plot, and a plot of pH vs. In, for ammonium sulfate.
11.1. Consider solutions of potassium carbonate and potassium citrate, with m = 0.1 kmol m-3.
A solution of 0.2 kmol m-3 HCI is slowly added to 10-4 m3 of the solutions considered. Draw a plot
of pH vs. m3 of solution added.
11.3. Estimate the electrical conductivity (for AC current) of solutions of sodium chloride,
sodium acetate, and ammonia, as a function of m.
11.4. It has been observed that the measured electrical conductivity of an electrolyte solution
depends on the frequency of the applied AC current. Describe the possible reasons for this, and a
method for extracting the true conductivity from the data.
11.5. Consider the first example in Section 10.6. Which condition should the pK of the amine
satisfy for the carbonate ion concentration to be negligibly small? Can an analogous condition be
determined for which the bicarbonate ion concentration is negligibly small?

LITERATURE

This chapter is heavily indebted to the notes for a course in electrochemistry which
Prof. F. Gioia gave for several years in the 19708 at the University of Cagliari, Italy.
Textbooks on electrochemistry abound, and most of them cover the material presented
here (and much more).
Chapter Twelve

POLYMERS

GIUSEPPE MARRUCCI
University of Naples
Naples, Italy

All human knowledge thus begins with intuitions,


proceeds then to concepts, and ends up with ideas.
I. Kant

NOTATION

A Free energy kcalkmol- I


a Diameter of tube of constraints m
al Activity of solvent
b Kuhn length m
c Concentration molecule m- 3
c* Critical concentration molecule m- 3
CI Lower-limit concentration molecule m- 3
C2 Upper-limit concentration molecule m- 3
C' Concentration by weight kgm-3
c" Molar concentration kmolm- 3
D Diffusivity m2 s- 1
DM Center-of-mass diffusivity m2 s- 1
DR Rotational diffusivity 8- 1

d Diameter of rods m
E Potential energy kcal kmol- I
F Force N
G Free enthalpy kcaIkmol- 1 319
320 G Elastic modulus (Section 12.3) Nm- 2
Chapter Twelve g =G/RT
g* Dimensionless G per unit volume
H Spring constant Nm- I
H M1X Enthalpy of mixing kcalkmol- I
J Total flux of dumbbells m- 2 S-I
JD Diffusion-induced contribution to J m- 2 S-I
JE Elastic-induced contribution to J m- 2 s- 1
Jv Velocity-induced contribution to J m- 2 S-I
k Boltzmann constant kcal molecule-I K- I
L Curvilinear tube length m
1 Rod length m
M Molecular weight kgkmol- I
Mc Critical molecular weight kgkmol- I
N Number of "tube" random walk steps
n Number of Kuhn segments
nJ Number of moles of J component kmol
p Order parameter
pOs Osmotic pressure atm
R Gas constant kcal kmol- I K- I
r End-to-end vector m
S Entropy kcal kmol- I K- I
So Cross-sectional area of unstretched sample m2
S' Partial molar entropy kcal kmor l K- 1
s'( Constitutive function for S' kcal kmol- I K- I
T Temperature K
T* Equilibrium phase transition temperature K
Time s
u Unit vector
V Free volume m3
VEXCL Excluded volume m3
v Velocity ms- 1
W( Distribution function m- 3
X Ratio of polymer to solvent volume
y Ratio of polymer to monomer volume
x,y,z Components of r m
YJ Mole fraction of component J
a Expansion ratio
Expansion ratio in J direction 321
Magnitude of velocity gradient Polymers
Shear magnitude
Angle
Viscosity kgm-1s- 1
Viscosity increment due to polymer kgm- I S-I
Chemical potential of solvent kcal kmol- I
Chemical potential in standard state kcalkmol- I
Viscosity of solvent kgm- I S-I
Intrinsic viscosity m 3 kg- 1
Friction coefficient Nsm- I
u Axial stress Nm- 2
(J' Extra stress tensor Nm- 2
Components of extra stress tensor Nm- 2
Relaxation time s
4> Volume fraction
4>J Volume fraction of component J
4>* Limit volume fraction for concentration
solution
4>1 Limit volume fraction for two-phase system
4>2 Limit volume fraction for total anisotropy
X Interaction parameter

Superscripts

MIX Of mixing
SW Of swelling
EL Elastic

Operators

D Difference across phase change


{) Baric derivative
() Ensemble average
12.1. INTRODUCTION

... I shall demonstrate the laws of motion of an indefinite


number of small, hard, and perfectly elastic spheres acting on
one another only during impact. If the properties of such a
system of bodies are found to correspond to those of gases, an
important physical analogy will be established, which may lead
to more accurate knowledge of the properties of matter. If
experiments on gases are inconsistent with the hypothesis of
these propositions, then our theory, though consistent with
itself, is proved to be incapable of explaining the phenomena of
gases. In either case it is necessary to follow out the
consequences of the hypothesis.
J. C. Maxwell

The thermodynamics of polymeric systems is dominated by a feature which is


the essential characteristic of most polymers, i.e., that they are chainlike molecules
with an enormous length-to. width ratio. This "geometric" property has con
sequences of a general nature which are often much more important than the
details of the specific chemical structure of the polymer. Therefore, a sort of
"universal" behavior is frequently observed which is characteristic of polymeric
substances and which justifies that these systems be studied per se. In this chapter
some of these universal properties are discussed. The individual aspects of the
particular chemical structure will appear only indirectly, i.e., through the numeri-
cal values of parameters which enter the equations describing the universal theory.
The fact that universal theories of polymeric behavior are essentially based
on geometrical features of the polymer molecules has the same logic underlying
the classical Maxwellian theory of gases. In that theory, molecules are regarded
as volumeless mass points, i.e., as will be done in the present case, they are
assigned a special geometrical structure. The statistical analysis then produces
results which are typical of that structure, and are independent of the chemical
details of individual molecules. Some of the latter only enter in determining the
parameters appearing in the theory: in the case of Maxwellian gases, their
molecular weight. Other details are omitted entirely. They may sometimes be
used to calculate "corrections" to the universal behavior.
Since the fundamental geometrical structure assigned to macromolecules is
significantly more complex than that assigned to molecules in the Maxwellian
theory of gases, the resulting theory is both more complex and richer as regards
results. Some of the predictions may be so distantly related to the basic assump-
tions that they are no longer easily grasped intuitively. In some cases, the
theoretical predictions may even be difficult to test experimentally. To mention
just one difficulty which is frequently encountered, though not necessarily the 323
324 major one, it should be remembered that polymers are usually polydisperse, i.e.,
they have a spread of molecular weights. Conversely, most theoretical results are
Chapter Twelve
obtained for the monodisperse case, and their extension to a poly disperse situation
is not always easy, not even conceptually. All results reported in this chapter
refer to monodisperse polymers only.
Most of the chapter deals with flexible polymers, which are by far the vast
majority. In some cases, the tacit assumption is also made that the polymer is a
linear chain, i.e., that there are no long branches in its chemical structure. This
again is the most frequent occurrence. The last section of the chapter, however,
is devoted to rigid rod-like polymers, which have recently acquired some import-
ance in the context of polymeric liquid crystals. It should also be clear that, with
the exception of rubbers, the chapter only deals with the liquid state of polymers.
Thus nothing will be said about the crystalline and glassy states, or about the
so-called thermosetting polymers, which do not admit a liquid state at all.
Throughout this chapter, the mathematical derivations are kept to a
minimum, and most complications are purposely left out, the scope being merely
that of introducing the reader to some of the main concepts used in polymer
theory, without any attempt at either completeness or even accuracy in the details.
Most equations are given without the numerical factors, even when they are
known, the emphasis being on showing trends rather than on calculating absolute
values. The latter are seldom predictable anyway, in view of the uncertainties in
determining the values of most parameters. It should finally be mentioned that
comparisons between theoretical and experimental results are only occasionally
alluded to, in spite of their obvious importance. In a single chapter, there is not
enough room for a vast subject, and drastic choices become unavoidable.

12.2. CHAIN CONFORMATIONS. THE RANDOM WALK

Accordingly, he turned (the world's shape) rounded and


spherical, equidistant every way from center to extremity-
a figure the most perfect and uniform of all; for he judged
uniformity to be unmeasurably better than its opposite.
Plato, Timaeus, 33b

Let us consider one of the simplest polymeric chains, that of polyethylene.


Chemically, we might represent the chain as follows:

H H H H
I I I I
.. -C-C-C-C-
I I I I
H H H H

the actual polymer being obtained when the basic unit, -CH 2 - , is repeated
thousands or tens of thousands of times.
Such a seemingly dull chain is in fact capable of assuming an astronomically
large number of different conformations (or shapes), as discussed below. The
possibility of a very large number of different molecular configurations is what
sets apart polymers from ordinary low molecular weight materials, the molecules 325
of which have an essentially fixed shape, capable of only minor distortions. Polymers
To understand the possibility of a large number of conformations, one should
remember the tetrahedral structure of the bonds of the carbon atom. Thus, if
three consecutive carbon atoms are placed in the plane of the paper (see Figure
12.1.1), the fourth one can move along the circle sketched in the figure without
altering bond angles and distances. A more detailed analysis shows that, along
the circle, there exist three energy minima (corresponding to one trans and two
gauche conformational isomers), and that the energy barriers between them are
small enough to be easily overcome by thermal energy. The existence of three
alternative positions for each consecutive carbon atom implies that a chain made
up of n such atoms can assume a number of different configurations of the order
of 3", which is indeed astronomically large already for n of order 100.
This conclusion, in its qualitative implications, is by no means altered if a
different "chemistry" is considered. For example, polystyrene will differ from
polyethylene insofar as some conformations will be more or less impeded by the
bulky benzene groups. Nevertheless, an enormous number of them will still
remain available. Similarly, the presence of double bonds, by prohibiting rotation
around them, will decrease the degrees of freedom to some extent. A chain like
that of polybutadiene, however, although containing a double bond every four
carbon atoms, still remains extremely flexible (the word flexible is used here to
indicate the capability of the chain to change its conformation by rotations around
bonds; it has nothing to do with the flexibility of, say, an elastic beam). Flexible
chains in the liquid state (or in the rubbery state, as will be seen later) will in
fact continuously change their conformation as a consequence ofthermal motion.
We shall see in the following the far-reaching effects of this peculiar property.
Polymers which have a single, fixed conformation are by far the exception.
These rigid polymers are also of interest, however. They will be discussed in
Section 12.6.
Returning to Figure 12.2.1, we may define a local orientation of the chain
by the vector which links alternate (not consecutive) carbon atoms. Thus, starting
from the left, the first and third atoms in the figure define a specific orientation
of the chain, U 1 (which is horizontal in the figure). The second and fourth atoms,
depending on the position of the latter on the circle, define in general a different
orientation U2. Of course, for a given Ulo U2 is restricted to a range of values. If
additional atoms are considered, however, soon the orientational correlation is
completely lost. In other words, if I and J are not too close, UJ and UJ are
randomly oriented with respect to each other. (If the fourth atom in Figure 12.2.1

w)
,,//,
" ..;/'/
/' / /
FIGURE 12.2.1. The carbon atom backbone ,,"
I
/1 / '
,
of a polymeric chain. After placing the first I
I
I
,
I
,

three atoms in the plane of the figure, the fourth


, ,
I
I
one may be placed anywhere along the circle I, ' ,

"" ,,
I

depicted. Unit vectors UI and U2 specify the


local chain direction. t/
326 is regarded as capable of occupying indifferently any position on the circle,
randomness is complete in a continuum sense; if only three positions with equal
Chapter Twelve
probability are considered, the randomness is also complete, but is restricted to
a discrete space.)
This consideration, which remains valid for all flexible chains, allows one
to consider, in place of the actual chain, a fictitious one made up of freely jointed
links of an appropriate length. Such a chain is depicted in Figure 12.2.2; all
segments of that chain are randomly oriented. The length of a link, called the
Kuhn length, depends on the details of the "chemistry." Thus, for example, we
expect polystyrene to have a larger Kuhn length than polyethylene insofar as,
because of the benzene groups somehow hindering rotations, a larger distance
along the polymer chain is required before loosing the orientational correlation.
Conversely, the statistical properties of the chain do not depend on the chemistry;
they will hopefully describe the universal properties of all flexible chains.
It is readily apparent that in a system of free chains (i.e., in the absence of
external fields inducing anisotropy) any conformation of a freely jointed chain
can equivalently be interpreted as the path of a three-dimensional random walk
of assigned step length b. Thus all the well-known results which describe the
statistics of a random walk apply also to this chain. In particular, if the total
number n of links in the chain is large enough, it is known that the distribution
W(r) of the end-to-end vectors r is well approximated by the Gaussian

(12.2.1)

and the average of the square end-to-end distance is given by

(12.2.2)

One implication of equation (12.2.2) is as follows. Let us assume that flexible


polymers in the liquid state are well represented by the model chain in Figure
12.2.2, and take the root-mean-square end-to-end distance of the chain as a
measure of the "size" of the coiled macromolecule. Then, since for a given
chemistry n is obviously proportional to the molecular weight M of the polymer,
equation (12.2.2) predicts that the size of the random coil grows proportionally
to the square root of M. Compare this prediction with the obvious consideration
that the extended length of chain grows with the first power of M, and that,
conversely, if the molecule is made to collapse into a solid globule, the size of
such a globule would grow as MI / 3.

FIGURE 12.2.2. The chain made of Kuhn


segments. All segments are randomly oriented.
The square-root dependence predicted by equation (12.2.2) compares favor- 327
ably with neutron scattering experiments of coil size conducted in polymer melts
Polymers
and concentrated solutions. The case of dilute solutions is more complicated. In
fact, the model chain depicted in Figure 12.2.2, i.e., the random walk, ignores
effects due to excluded volume. The chain in Figure 12.2.2 is volumeless and
therefore it allows for overlapping conformations which would be prohibited for
the actual chain.
Accounting for excluded volume effects results in the prediction that the
coil is more expanded than predicted by equation 12.2.2. Also, the M dependence
of the coil size changes to a somewhat larger power. Long ago, Flory calculated
that
(12.2.3)

a prediction essentially confirmed by experiments, by more sophisticated recent


analyses, as well as by computer simulations of SAWs (self-avoiding walks).
The result in equation (12.2.3) holds for the case of a dilute solution in good
solvents. By decreasing the solvent quality, i.e, by increasing the attraction forces
among polymer segments relative to those between polymer and solvent, the
chain "contracts" and a situation can be envisaged where the contraction due to
these energetic effects exactly compensates for the expansion due to excluded
volume. Such a special situation is obtained in relatively poor solvents, at an
appropriate temperature called the 8 or Flory temperature. At the 8 temperature,
the chain behaves "ideally," i.e., it essentially obeys the statistics of the random
walk, and equations (12.2.1) and (12.2.2) apply.
As mentioned above, equations (12.2.1) and (12.2.2) also apply to the
technologically important case of polymer melts and concentrated solutions in
solvents of arbitrary quality. This seemingly paradoxical result is due to "screen-
ing" of excluded volume effects of a given chain brought about by the surrounding
ones. In other words, bumping of a chain into itself, which in a dilute solution
gives rise to an expansion, is now exactly compensated by the collisions of the
given chain with the others. A chain segment does not distinguish whether it is
bumping against a segment of the same chain or of a different one. Therefore,
the excluded volume effect vanishes and the ideal situation is recovered.

12.3. RUBBER ELASTICITY

But in the present century . .. we have come to see


that the unproved postulates with which we start
are purely arbitrary. They must be consistent,
they had beller lead to something interesting.
C. Coolidge

Equation (12.2.1) describes the probability that the chain may be found in
the neighborhood of a given value of the end-to-end vector r. Therefore, as long
as the only variable considered is r, one may associate with it (to within an
arbitrary additive constant) the entropy S,
S(r) = kin W(r) = -3kr2/2nb 2 (12.3.1)
328 where k is the Boltzmann constant. Correspondingly, and consistently with the
fact that in the ideal chain there are no energy effects, the free energy is given by
Chapter Twelve

A(r) = -TS(r) = 3kTr2/2nb 2 (12.3.2)

Finally, by assuming that no dissipative processes take place, we can equate


a change in free energy to the work done on the system, and hence obtain the
force F given by

F(r) = dA(r)/dr = 3kTr/nb 2 = Hr (12.3.3)

where H = 3kT/nb 2 is a constant which, at temperature T, is characteristic of


the chain considered.
Several comments are in order concerning equation (12.3.3). First, let us
understand the physical meaning of the equation. Suppose we imagine an
experiment in which we hold one end of the chain fixed at a point. We already
know that, because of thermal motion, the other end of the chain will move about
and explore the space around the origin, with the probability of visiting each
point described by equation (12.2.1). It is worth noting, incidentally, that the
neighborhood of the origin is visited most frequently.
We shall assume now that also the second end of the chain is held fixed at
some point, thus specifying a given value of r. The effect of thermal motion will
be that of pullil1g on the constraints which hold the ends of the chain fixed and,
although this force will vary with time in both direction and magnitude, its
average value is that given by equation (12.3.3). For obvious reasons of symmetry,
the direction of this average force is the same as that of r. It is also apparent
that, by increasing the end-to-end separation, this average force must increase
insofar as thermal agitation of the chain will more and more frequently attempt
to pull its ends toward a more probable relative position. The force will only be
zero, on the average, when the two ends coincide and symmetry is restored.
The gedanken experiment just described forms the basis of a characteristic
behavior of flexible polymers, i.e., that they are elastic objects. (Note that the
word elasticity is used in its classical mechanical sense of a force-deformation
relationship, not in the specific sense of "state coinciding with site" used in earlier
chapters, though rubbers are elastic also in the latter sense.) As detailed later in
this section, elasticity of rubbers is one macroscopic manifestation of this property,
another being the viscoelasticity of polymeric liquids. In fact, in order to move
one step toward macroscopic behavior, it is noteworthy that the force given by
equation (12.3.3), which so far has been interpreted as an average over time for
a single chain, can equivalently be viewed as the mean instantaneous response
of a large collection of equal chains, all of which possess the same value of the
end-to-end vector r. (This is the essence of the so-called ergodic description.)
Before leaving equation (12.3.3), a few more comments are in order. The
linearity ofthe F(r) relationship, which is an obvious consequence ofthe Gaussian
distribution, appears to imply that an arbitrarily large extension of the chain can
be achieved, even larger than the fully extended length nb, provided a large
enough force is applied. This apparent paradox is due to the fact that the Gaussian
function becomes inadequate for large values of the end-to-end separation, where
it should be replaced by an inverse Langevin function which correctly predicts 329
the force to approach infinity when r approaches nb. The use of the linear Polymers
relationship, however, is not a serious limitation, except in rather exceptional
cases. In fact, it can be shown that deviations become important at about one
third of the fully extended length. Now, since the ratio of the fully extended
length to the size of the random coil is [see equation (12.2.2)]

lrMAX = nb/b.jn =.jn (12.3.4)

and n is typically a very large number, enormous deformations can be applied


to the polymer molecule before its elastic response becomes nonlinear.
The very fact that large deformations of the chain are possible, as reflected
macroscopically in a similar behavior of rubbers, characterizes the kind of
elasticity considered here which, as has been shown, has an en tropic origin. The
elasticity of most solid materials has to do with a change in internal energy due
to deformation, the applied strain reflecting a change in the distance between
atoms linked together by one form or other of interatomic potential. An energetic
elasticity is characterized by very small deformations and large elastic moduli.
As opposed to this, an entropic elasticity implies small elastic moduli and the
possibility of large deformations. It does not involve a change of atomic bond
lengths or bond angles but, rather, changes of conformations, all of which, at
least in the ideal case, are energetically equivalent. The elastic force is merely a
manifestation of thermal motions. Indeed, as shown by the elastic constant value
in equation (12.3.3), should the absolute temperature go to zero, the elastic force
would vanish. We mention in passing that another good example of entropic
elasticity is offered by an ideal gas contained in an isothermal cylinder-piston
system (see Chapter 8). Pushing the piston in yields no change of internal energy,
and yet the piston is ready to bounce back upon removal of the force on the
piston. The elastic force, i.e., the increased pressure in the gas, results again from
thermal motions of the gas molecules and is of entropic origin. Actually the
analogy between the ideal chain and the ideal gas is useful at a more general,
logical level and will be further discussed later in this chapter. (See Section 8.1
for a discussion of entropic elasticity in ideal gases.)
Let us now consider rubber elasticity in a strict sense. A rubber is composed
of polymer chains chemically linked to each other at some points to form a
three-dimensional network. Thus, a macroscopic piece of rubber is a single giant
macromolecule and is, of course, a solid. (See the discussion in Section 8.8 in
this regard. Obviously, it does not make sense to speak of the molecular weight
of a cross-linked rubber, since it is essentially infinitely large. The relevant
molecular weight in a network is that of a segment between consecutive cross-
links). Because the nodes of the network (i.e., the cross-links) are relatively
sparse, however, the chain segments between consecutive cross-links are long
enough to maintain all the conformational degrees of freedom as if they were in
a liquid. Therefore, all the results previously reported for a free chain also apply
to chain segments of the network. We will use in particular equation (12.3.2) for
the free energy.
Rubbers are incompressible materials, in the same sense that liquids are
incompressible, i.e., while it is easy to change their shape at constant volume,
330 changing their volume at constant shape requires very large pressures. Therefore,
in most cases volume changes can be neglected, and the thermodynamic state of
Chapter Twelve
a rubber is specified by any two of the three variables: temperature, deformation,
and stress, just as in a gas we have temperature, volume, and pressure. (Sufficient
assumptions of invertibility of the isothermal deformation-stress relationship are
of course involved here; see by analogy Section 2.6.) The constitutive relationship
between these three variables is the equation of state of the rubber, although the
name rubber elasticity is more frequently used to indicate the relationship between
stress and deformation at some given temperature. If the property of incompressi-
bility is taken literally, only isochoric (i.e., volume preserving) deformations are
possible, and the equation of state will predict stresses only to within an arbitrary
pressure.
An arbitrary deformation is completely specified by three mutually
orthogonal principal directions and by three positive numbers at, a2, and a3
which represent the stretch (a> 1) or compression (a < 1) ratios along the
principal directions. Restriction to isochoric deformations requires that

(12.3.5)

We now consider any chain segment in the network between two consecutive
cross-links. If x, y, z are the components of the end-to-end vector before the
deformation has been applied, after the deformation we may write

(12.3.6)

By averaging over the distribution of r vectors, with averages indicated by


), equation (12.3.6) becomes

(r2) = ai(x 2) + a~(y2) + ai(z2)


= (ai + a~ + ai)(x 2) (12.3.7)

where the rightmost equality is obtained by assuming that, in the undeformed


rubber, the r distribution is isotropic and therefore (x 2) = (y2) = (Z2). Moreover,
if one temporarily assumes that all chains in the network are equal, i.e., that they
are made up of the same number of Kuhn segments, one has (x 2) + (y2) + (Z2) =
nb 2 Thus equation (12.3.7) becomes

(12.3.8)

We are now ready to calculate the free energy per unit volume of a deformed
rubber. By ensemble averaging equation (12.3.2) and substituting for (r2) from
equation (12.3.8), we obtain

(12.3.9)

where c is the number of chains per unit volume forming the network.
It should be noted that n does not appear explicitly in equation (12.3.9). 331
Thus, as shown in the following, the assumption that all chains have the same
Polymers
value of n is unnecessary. As long as equation (12.3.7) remains valid for any
subset of chains having given values of n[, equation (12.3.9) is obtained, with
the corresponding CI in place of c, for the Ith contribution to free energy. Summing
all these contributions, equation (12.3.9) with the overall concentration C is again
obtained. We note, however, that the average value of n appears indirectly in
the result, through C itself. Indeed, imagine adding further cross-links to the
network, thus reducing the average value of n. The number of chains per unit
volume correspondingly increases. In fact, the prefactor in equation (12.3.9)
essentially depends on cross-link density and, weakly, on temperature. We shall
see shortly that this prefactor is the modulus of rubber elasticity.
An important assumption embodied in the result of equation (12.3.9) is that
of affinity of the deformation. When writing equation (12.3.6), we have assumed
that the chain end-to-end vector deforms as if it were imbedded in the continuum,
i.e., that it deforms affinely with the macroscopic deformation. Since equation
(12.3.7) rather than (12.3.6) has been used to obtain our result, it is only required
that the affinity assumption holds on the average. Even for the average, however,
affinity is by no means assured, and different possibilities can be, and have been,
envisaged. Here we ignore all such complications and proceed to show some
results of rubber elasticity theory in its basic form.
Stresses are obtained from equation (12.3.9) by equating the work done by
them upon deforming the material to the corresponding change in free energy.
(This is essentially why the theory considered is one of elasticity, in the sense
used in previous chapters, so that mechanical processes are reversible.) This can
be done for a general deformation, thus obtaining the general constitutive equation
for the stress of an ideal rubber. For the sake of simplicity, we only show here
the procedure for a particularly simple deformation, uniaxial elongation. Con-
cretely, we may think of a cylindrical specimen of rubber being uniformly stretched
in the axial direction under the action of a force F.
Let us assume the principal stretch ratio along the axial direction is a.
Because of symmetry, the other two principal ratios are equal and, in view of
equation (12.3.5), are given by 1/.ja. Thus for the case considered equation
(12.3.9) becomes

A(a} = ckT(a 2 + 2/ a}/2 (12.3.10)

with the derivative given by

dA/da = ckT(a - l/a 2 } (12.3.11)

If L is the current length of the specimen and Lo the undeformed length,


the elementary work done by the force F is FdL = FLo da, and the work per
unit volume is FLo da/ V = Fda/ So, where So is the area of the undeformed
cross section of the sample. Thus, on comparison with equation (12.3.11), one
obtains

(12.3.12)
332 Quantity F I So is usually called the engineering stress. The true stress (T is
the ratio F I S, with S the area of the current cross section. Since Sol S = a, the
Chapter Twelve
stress is given by

(T = ckT(a 2 -1/a) (12.3.13)

This equation should apply to arbitrary values of a (i.e., it is not a small


deformation approximation). It should be remembered, however, that the
Gaussian distribution breaks down at very large deformations, i.e., when the
chains are extended close to their fully extended length. Therefore, one should
expect deviations to appear when a approaches aMAX as given by equation
(12.3.4). At the opposite extreme (and there is a significant range in between),
i.e., for very small deformations, equation (12.3.13) can be linearized. By setting
a = 1 + e, e 1, one obtains

(T = 3ckTe (12.3.14)

Therefore, the Young modulus of the ideal rubber is 3ckT. Consequently, if the
rubber is incompressible, i.e., if it has a Poisson ratio 0.5, the shear modulus G
is given by

G= ckT (12.3.15)

For a shear deformation y of arbitrary magnitude, the predictions of the


theory are as follows:

(Tl2 = Gy (12.3.16)

(T11 - (T22 = Gy2 (12.3.17)

(12.3.18)

where the directions 1, 2, 3 as well as the definition of yare reported in


Figure 12.3.1.

FIGURE 12.3.1. A cubic element of rubber


subjected to a shear deformation. 'Y = AA'jOA.
All the above predictions, which are certainly not trivial, are in reasonably 333
good agreement with experiments. Systematic, albeit minor deviations are also
Polymers
found, however. The source of these discrepancies has formed the object of warm
debate for many years and is not fully settled even now. As will be further
discussed later, one possible origin of these deviations is the presence of entangle-
ments among the chains of the network, or, as previously mentioned, some partial
breakdown of the affinity assumption. On the whole, however, the rubber elasticity
theory, as developed in the 1940s, has been a real success and has influenced
considerably subsequent advances in polymer theory.

12.4. TRANSPORT PROPERTIES

In 1866 I made some attempts to ascertain whether the state of


strain in a viscous fluid in motion could be detected by its
action on polarized light. ... I was unable to obtain any
result with a solution of gum or sirup of sugar,
though I observed an effect on polarized light
when I compressed some Canada balsam which had
become very thick and almost solid in a bottle.
J. C. Maxwell, 1873

An area where molecular modeling of polymers has been, and still is,
particularly active is the prediction of transport properties in the liquid state,
such as the diffusion coefficient or the viscosity. To avoid possible misunderstand-
ings, it is better to specify from the beginning that molecular modeling does not
predict values of either the diffusivity or the viscosity, but rather their dependence
on the molecular weight of the chain, just as the random walk analysis in Section
12.2 predicts the M dependence of the molecular size, not its value, the latter
depending also on the Kuhn length b, which is determined by the chemical details
of the chain. Similarly, in dealing with transport properties, one makes use of
an additional chemistry-dependent parameter, i.e., the friction coefficient of a
chain segment, the purpose of the theory being that of predicting how this "local"
transport property scales up to the level of the whole chain or of a collection of
them.
On the other hand, the scope of the theory is also more ambitious than
sketched above. In fact, momentum transfer in polymeric liquids is a much more
complex phenomenon than in ordinary liquids. While for the latter viscosity is
all one needs to know, polymeric liquids are non-Newtonian and viscoelastic
(or, in other words, they exhibit relaxation phenomena on a time scale which is
by no means negligibly small as compared to experimental time scales). Thus,
the purpose of the theory is actually that of predicting a constitutive equation for
the transport of momentum, i.e., for the stress. There is no room in this chapter
for such a scope, to which whole books are dedicated. It may be useful, however,
to gain some insight on how the problem is tackled, at least at an elementary
level. We note further that the concept of viscosity remains meaningful also for
polymeric materials because it can be shown that, for sufficiently slow flows,
their complex behavior approaches that of a Newtonian, purely viscous liquid.
(Of course, the same is true for the Maxwellian gas, only in the case of the latter
334 a flow of any pragmatic interest is almost invariably "sufficiently slow." In the
Chapter Twelve case of polymeric liquids, on the other hand, deviations from Newtonian behavior
take place well within conditions of practical interest.) All values of viscosity
reported subsequently refer to slow flow conditions only.
We shall first consider the case of extremely dilute solutions, where each
polymer chain may be thought of as being in contact only with the solvent.
Subsequently, we shall briefly touch upon the case of concentrated solutions and
polymer melts. In either case, the first property to be examined is the diffusivity.
Let us start by defining the friction coefficient of a molecule, &, and by
showing its universal relationship to the diffusivity D. Assume that a molecule
of a fluid is acted upon by a force F. That molecule will then move, relative to
the others, with a mean velocity v. The friction coefficient is defined as the ratio

(1:ZA.l)

and is assumed to be independent of F (or v). We note that equation (12.4.1) is


nothing else but Stoke's law, valid for macroscopic objects at very low Reynolds
numbers, extrapolated down to molecular level. At that level, it is a hypothesis
which, however, appears to be consistent with experimental evidence.
Assume now that a constant field of force acts upon the molecules of a solute
in a dilute solution, driving them toward an impassable barrier. The tendency of
these molecules to concentrate against the barrier will be counteracted by diffusion
in the opposite direction. Under stationary conditions, we may establish a balance
between the flux in the direction of the force and that due to diffusion, i.e.,

cFj&-Dac/ax=o (12.4.2)

where c is the concentration, D the diffusion coefficient, and x is oriented in the


direction of the acting force field. Therefore, the equilibrium distribution is given
by

c(x) = Co exp(Fx/ D&) (12.4.3)

where Co is a normalization constant. On the other hand, an eqUilibrium distribu-


tion must be Boltzmannian, i.e., of the form exp(-Elk1), with kthe Boltzmann-
ian constant and E the potential energy. Since in our case E = Fx. the following
identity is derived:

(1:ZAA)

This equation is known as the Einstein relationship. It provides a simple


link between a frictional property & and a parameter arising from Brownian
motion, namely, the diffusivity D.
Equation (12.4.4) is of general validity (within the stated assumptions) and
applies to polymer molecules as well. Thus, in order to make predictions on the
polymer diffusivity, one may equivalently speculate about the friction coefficient
of the polymer chain. Two extreme cases are easily analyzed. One of them is the
so-called "free draining" case. It assuines that, in any relative motion of the chain
with respect to the solvent, the latter moves freely inside the volume (having, on 335
the average, a spherical shape) containing the polymer coil. Thus, the friction Polymers
on like segments of the chain is the same, and the overall friction coefficient is
proportional to chain length or molecular weight. In other words, if Co is the
friction coefficient of a chain segment and the polymer molecule comprises n
such segments, the friction coefficient of the polymer molecule is nCo. Therefore

C=M; (llA.5)

The other case is the opposite extreme. It assumes that the sphere containing
the polymer molecule is dynamically impermeable to the solvent, i.e., the polymer
segments close to its external surface "shield" the more internal ones, which thus
exert no friction. Since the polymer molecule now behaves as a solid sphere of
radius (r2)1/2, by considering Stokes's law for a sphere one writes

(ll.4.6)

where the two indicated proportionalities make use of either equation (12.2.2)
or (12.2.3) and are therefore valid for a 8 solvent or for a good one, respectively.
As one might expect, reality falls somewhere in between the two extreme
cases treated above. On the other hand, a more precise calculation would require
,accounting explicitly for the hydrodynamic interactions between polymer seg-
ments, i.e., solving for the flow field inside the sphere. We omit these complications
here and move on to consider the next property, i.e., the viscosity. More precisely,
we want to estimate the increase of viscosity over that of the solvent due to
addition of a polymeric solute.
A simple way of obtaining a prediction makes use of a result, again due to
Einstein, which holds for a dilute suspension of solid spheres. Einstein's result is

(IL - ILs)/ ILs = 2.5e/> (ll.4.7)

where ILs and IL are the viscosities of the solvent and of the suspension, respec-
tively, and e/> is the volume fraction occupied by the spheres. By assimilating the
polymer molecules to spheres with radius (r2)1/2, one obtains

(ll.4.8)

where c is the number of polymer molecules per unit volume and, again, the
exponent of M depends on whether we consider a 8 solvent or a good one.
Postponing a discussion of this result, we now consider a more general
approach. Indeed, the use of Einstein's formula for solid spheres neglects a
feature of the polymer molecules which we envisage might play some role, i.e.,
that they are deformable objects. The effect of a velocity gradient, by dragging
different parts of the same polymer molecule in different directions (relative to
the molecule center), is in fact "deformational." Moreover, the rigid-sphere model
is, as we have seen, an extreme case of friction behavior, and we would like to
explore the free draining extreme as well or, possibly, something in between.
336
Chapter Twelve
FIGURE 12.4.1. The elastic
dumbbell model, which may be
used to roughly simulate the
dynamic response of a polymer
in dilute solutions.

To account for deformations of the polymer molecules due to flow we must


distinguish the motion of different parts of a chain. On the other hand, different
parts of the same chain are obviously linked to each other by the physical
continuity of the chain itself. Thus the simplest mechanical model which incorpor-
ates the essential features under scrutiny here is the elastic dumbbell depicted in
Figure 12.4.1. The two friction beads are meant to represent different parts of
the same molecule. It is apparent that, if the solvent has a velocity gradient, the
two beads are dragged differently and a deformation of the dumbbell will result.
The deformation, however, is opposed by the spring, which is meant to portray
the effect of chain continuity. Furthermore, and in the same spirit, it appears
sensible to perform the following identifications: (1) the end-to-end vector of the
chain is identified with that of the dumbbell; (2) elasticity of the spring is made
to obey equation (12.3.3), Le., we take a linear spring (unless we must proceed
to extreme deformations); (3) we assign to the beads a friction coefficient ( of
the same order of magnitude as that ofthe whole polymer molecule. Consistently,
we also assign to the beads a diffusion coefficient D = kT/ (. Thus, although the
spring attempts to bring the two beads together, Brownian motion will keep them
apart. In fact, in the quiescent solvent, Le., at equilibrium, the two opposing
effects will balance, giving rise to an end-to-end vector distribution which exactly
matches that given in equation (12.2.1).
If the dumbbell model appears too crude, insofar as the friction distributed
along the chain has been concentrated in only two beads, the remedy is clearly
at hand. We may consider the model chain, depicted in Figure 12.4.2, with N + 1
beads connected by N springs. In this case, of course, we must assign to each
bead a friction coefficient of order (/ N, and each spring will have an elastic
constant which, consistently with equation (12.3.3), is N times larger. This
generalized model, known as the Rouse chain, can be treated in two ways. In

FIGURE 12A.2. The Rouse chain. As long as the


number N of elastic segments is large, the results
are independent of N.
the solution obtained by Rouse himself, hydrodynamic interactions among the 337
beads are ignored, Le., the free draining case is considered. In the solution
Polymers
due to Zimm, hydrodynamic interactions are also accounted for, albeit only
approximately.
The nice thing about these models is that, as long as the elasticity of the
springs is kept linear and no other complications are introduced, they give rise
to solutions in closed form for arbitrary velocity gradients, i.e., they generate
explicit constitutive equations for the viscoelastic response of these liquids. Since
the dumbbell model, which has the simpler mathematics, already contains most
of the essential physics, we say a few more words and show briefly how the
predictions of this model are worked out.
The only internal coordinate of the dumbbell model is the end-to-end vector
r. Thus, as in all statistical treatments, we must consider a population of dumbbells
generally exhibiting a spread of r values, as described by a distribution function.
In a general time-dependent situation, this distribution is indicated by W(r, t),
where time t acts as a parameter. At any fixed time, the function describes how
many dumbbells are found in the neighborhood of a given r, in the sense that
W(r) d 3r represents the differential fraction of the population in the neighborhood
d 3r of r.* An example of this function is offered by equation (12.2.1), which
applies at equilibrium. Of course, since at equilibrium the distribution is isotropic,
vector r only appears in it through its (scalar) modulus r.
The effect of a velocity gradient will be that of "distorting" the distribution,
making it anisotropic. The distribution function W(r, t) obeys an equation of
change which looks like a mass balance or continuity equation, indeed expressing
a similar conservation concept. The equation is written in the form

aw/at = -divJ (12.4.9)

This equation states that if, at a given value of r, W is changing in time, the
change must be due to the fact that the flux J of dumbbells toward that value of
r is different from the flux of the dumbbells leaving the same value, the difference
being measured by the flux divergence (in r space).
In equation (12.4.9), J is the overall flux. It is readily recognized that J is
made up of three contributions, i.e.,

(12.4.10)

where J v is the flux due to the velocity gradient, J E that due to the spring elasticity,
and J o that due to diffusion.
The explicit expressions for these fluxes are not given here because we are
only interested in stating the problem in general terms. Suffice it to say that, once
the expressions for the three fluxes are substituted into equation (12.4.9), there
results a differential equation for W(r, t) which admits explicit solutions for

* More explicitly, if X, y, z are the Cartesian components ofr, W(x,y, z, t) dx dy dz gives the fractional
population at time t of dumbbells with components between x and x + dx, y and y + dy, z and z + dz.
338 arbitrary velocity gradients (as long as the spring elasticity is linear). Thus, the
Chapter Twelve
anisotropy induced by any flow can be evaluated, and the next question immedi-
ately poses itself. How do we obtain values of, say, the stresses induced by the
flow from this information?
The answer to this question or to similar ones (for instance, one might want
to get information about the optical anisotropy, Le., the birefringence) is as
follows. Macroscopically observable quantities, like a stress or a birefringence,
can be expressed as appropriate averages of some functions of the end-to-end
vector r. Thus, since the distribution of r is known, these averages are easily
calculated.
For example, the average which is required to calculate the extra stress a
(Le., the stress to within an arbitrary additive isotropic tensor, which is all that
the constitutive equation delivers for an incompressible fluid) is that over the
dyadic product rr. Specifically .

a = cH(rr) (11.4.11)

where H is the spring constant and c the number of dumbbells (Le., of polymer
molecules) per unit volume. Equation (12.4.11) is obtained from the definition
of stress, Le., by ideally cutting through the material with a plane surface, counting
how many dumbbells are cut per unit area, and summing over the elastic forces
in those dumbbells. Of course, equation (12.4.11) only gives the stress contributed
by the polymer molecules, which adds to that due directly to the solvent. Unless
the solution is extremely diluted, however, the polymer contribution turns out
to be the dominant one.
An important comment on the structure of equation (12.4.9) follows. Since
the equation is first-order differential in time, integration from time zero up to
the current time requires an initial condition (Le., the initial distribution needs
to be known), as well as knowledge of the velocity gradient throughout the time
interval from 0 to t. In other words, W(r, t) depends on the previous history of
the velocity gradient (see Chapter 5), not just on its value at time t, i.e., the
material has memory for past kinematics. In view of equation (12.4.11), this
property of the distribution function is transferred to the stress, Le., the material
is viscoelastic. A loose analogy with the discussion in Chapter 5 can be obtained
by regarding the distribution function W(r, t) as the internal state variable;
equation (12.4.9) is then recognized as the kinetic constitutive equation for the
rate of change of the internal state variable. The analogy is not as loose as it
appears at first sight: the stress (a generalization of pressure) depends on the
internal state variable, and the value of the latter cannot be imposed externally
(like the velocity gradient) but is determined by its initial value and its intrinsic
kinetics of evolution in time. Finally, the property of fading memory discussed

* For those not familiar with tensor notation and r space averages, the meaning of equation (12.4.11)
is as follows. Assume one wishes to calculate, say, the xy component of a. Then, by calling X, y
and z the components of r, one has

U xy = cH(xy) = cH fff xyW(x,y, z) dxdydz


in Chapter 5 is guaranteed by the stochastic process of Brownian motion, i.e., 339
by the J D and J E components of the flux. Polymers
Actually, there is more which can be said about the thermodynamics of
systems such as those discussed above, in the spirit of Chapter 5. The conforma-
tional entropy of an anisotropic distribution can be calculated in a similar way
as was done for rubber in the previous section. The corresponding free energy
is also readily obtained, because internal energy is not affected by the distribution
function for ideal chains. (In the latter sense, these materials are recognized as
materials with entropic viscoelasticity.) Now, since an equation equivalent to
equation (5.2.8) can be derived from the appropriate three-dimensional generaliz-
ation of the theory presented in Section 5.2, the stress can be expressed as the
instantaneous partial derivative of free energy with respect to strain. This alternate
derivation of the stress would give the same result as in equation (12.4.11). In
fact, this alternate route was followed in the preceding section for the case of
rubbers, which are materials with entropic elasticity (no viscoelasticity). Since
by means of the affinity assumption the free energy was obtained directly in that
case, there was no need to derive explicitly the anisotropic distribution resulting
from deformation. It should be noted finally that, by calculating the excess free
energy due to flow, an estimate of the flow-induced crystallization effects discussed
in Section 5.4 can be obtained. (See also the next section.)
Returning to the models discussed earlier, it should be mentioned that the
procedure for calculating the stress outlined above, which passes through two
subsequent steps, can in fact be shortened and considerably simplified. Instead
of first calculating the distribution function from the assigned kinematics of
motion, and then the stress from the distribution, a direct link between stress
and kinematics can be established. By suitable manipulation, equation (12.4.9)
can be transformed into an equation for the average (rr), i.e., into a constitutive
equation for the stress tensor. The following two material parameters will appear
in the latter equation:

T = ~/2H (12.4.12)

and
/-I: = ckTU2H (12.4.13)

which have the meaning of a relaxation time and a shear viscosity, respectively.
They will be discussed separately.
By using the expression for the spring constant H given by equation (12.3.3),
we first rewrite equation (12.4.13) in the form

(12.4.14)

where IL - ILs has been written instead of IL' since equation (12.4.13) only gives
the polymer contribution to the viscosity. Also, (r2) has been introduced instead
of nb 2 for a more direct comparison with equation (12.4.8). It is understood that
(r2) is an equilibrium average.
Again, two extreme cases can be considered. In one of them, by assuming
that the macromolecule is dynamically impermeable to the solvent, the friction
340 coefficient is taken to be proportional to (r2)1/2. Since it is also proportional to
the solvent viscosity, equation (12.4.8) is obtained again. At the opposite extreme,
Chapter Twelve
i.e., in the free draining case, , is proportional to M, and equation (12.4.13)
becomes

(12.4.15)

We finally express equations (12.4.8) and (12.4.15) in terms of intrinsic (or


inherent) viscosity. If c' is the polymer concentration by weight (c' = cM), we
obtain

[JL] = (JL - JLs)/ JLsC' = MI/2( = M4/5) (12.4.8')

[JL] = M( = M6/5) (12.4.15')

Although the predictions span a range of possible powers of M, depending


on the importance of the hydrodynamic interactions and on solvent quality, the
common feature of the above equations, i.e., that [JL] grows with increasing M,
should be emphasized. The meaning of this prediction is as follows. We assume
that, by using samples of the same polymer differing in molecular weight, solutions
with the same concentration by weight are prepared. Then, although these
solutions have exactly the same monomer concentration, their viscosities are
different, the larger viscosity being obtained from the polymer with larger molecular
weight. In other words, the "geometrical" organization of the material has a profound
influence on viscosity: fewer but longer polymer molecules give rise to a larger viscosity
than that obtained with more numerous but shorter ones, the total amount being the
same in the two cases. The experiments fully confirm this prediction, showing
an M-power dependence which is typically around 0.7-0.8. In fact, measurement
of the intrinsic viscosity is a simple way of determining the molecular weight of
a polymeric substance.
The other important parameter arising from the theory is the relaxation time
T. As the name implies, it measures the time required for the perturbed system

to relax back to equilibrium. More precisely, the dumbbell model predicts that,
after a flow or a deformation has been stopped, the stress (or any other measure
of the anisotropy induced by the motion) decays exponentially, with a time
constant T as given by equation (12.4.12). More detailed theories, like the
Rouse-Zimm model, predict a sum of exponentials instead of a single one, i.e.,
a set of relaxation times. The largest of them, however, is what really matters in
most cases, and it essentially coincides with that predicted by the dumbbell model.
The relaxation time depends more strongly on molecular weight than the
intrinsic viscosity. Again depending on hydrodynamic interactions and solvent
quality, the predictions vary somewhat, but roughly one finds

(12.4.16)

Thus, if M is large enough, T may well approach values of the order of, say,
seconds, the more so the more viscous is the solvent. This has important con-
sequences on the "deformation" of the polymer molecules. Indeed, as we have
already mentioned, a flow has a tendency to deform the molecular coils which, 341
however, is counteracted by Brownian motion tending to reestablish equilibrium. Polymers
The strength of the flow is measured by the magnitude r of the velocity gradient,
while the rate by which equilibrium is approached is measured by T. Thus, which
of the opposing effects will prevail is determined by comparing r with 1/ T. In
particular, if we have

r l/T (12.4.17)

we may expect that the flow is effective in stretching and aligning the polymer
molecules. In some cases, these effects may have important practical con-
sequences.
We now move on to briefly consider the case where the polymer molecules
are concentrated. We may as well jump to the other extreme of a polymer melt,
the case of concentrated solutions not differing in essence from that situation.
Indeed, the essential difference in the dynamics between the dilute and the
concentrated case is the following. While in a dilute solution the interaction
between any polymer molecule and the surrounding medium, i.e., the solvent, is
purely frictional, in a melt each polymer molecule must also take account of the
fact that, when moving about, it cannot cut across the other chains. In other
words, the interaction is not merely frictional but, in some sense, topological.
To go directly to an extreme case, let us examine a melt of very long chains.
Each chain is still a random coil, i.e., it is an extremely convoluted object which,
on the whole, occupies a spherical volume the size of which is determined by
equation (12.2.2). This sphere, however, is not filled up by that single molecule
which actually only takes up a small portion of the available space. The rest,
which in the dilute case was occupied by the solvent, is now occupied by segments
of other long chains. We have therefore an extremely intertwined situation and
the molecules are said to be entangled.
We consider now the dynamics of that chain. By thermal motion, the chain
attempts to move about but, in so doing, it bumps repeatedly against impassable
topological obstacles constituted by the surrounding chains. At first sight, it might
seem that the diffusivity is zero, i.e., that the system is solid-like. Closer inspection,
however, reveals that the chain has retained the possibility of diffusion in spite
of the obstacles, but the price to be paid has increased enormously.
To see this possibility we refer to Figure 12.4.3, which gives a two-dimensional
picture of the situation. The chain is shown together with a number of dots,
which are meant to represent cross sections of surrounding chains with the plane
of the figure. The dots are impassable obstacles and therefore the chain is restricted
in its sideway motion to a space which, in three dimensions, looks like a tube
(the shaded region in Figure 12.4.3). On the other hand, the obstacles do not
prohibit the chain moving along its own length or, to be more precise, along the
length of this tube of lateral constraints. This particular kind of diffusion of
polymer chains in entangled systems has been called reptation, or snake-like
motion.
Let us now estimate the diffusion coefficient under these conditions. We
must distinguish between a diffusion process along the tube, regulated by a
diffusivity D, and the diffusion in space of the center of mass of the polymer
342 molecule, D M , which is observed macroscopically in diffusion experiments of
labeled chains.
Chapter Twelve
As far as the snake-like motion along the tube is concerned, it can be argued
that the polymer molecule only experiences an ordinary friction, regulated by a
friction coefficient C, D and C being linked by equation (12.4.4). We can also
envisage C strictly proportional to the length of the chain (there are no hydro-
dynamic interactions to account for in this case), and therefore

D= 11M (12.4.18)

Before estimating D M , we need to introduce a characteristic dimension of


the tube of constraints, a, which is the spacing between obstacles. It should be
noted that a represents both the tube diameter and the distance over which the
tube has changed direction randomly. As long as the molecular weight of chains
is large enough, this characteristic length a is not expected to vary with M but
only with the density of polymeric material, i.e., to have its minimum value in
the melt and to increase progressively when diluting with a solvent. We note
further that, even in the melt, a is expected to be significantly larger than the
Kuhn length of the chain b, i.e., the tube is significantly wider than the chain
itself. In other words, in its wriggling motion due to thermal agitation, the chain
explores sideways a space which goes well beyond the distance of the nearest-
neighboring material. It is the collective action of so many other chains which
ultimately encages sideways the given one. A theoretical prediction of the link
between the values of a and b has not been found yet; the information that a is
significantly larger than b is obtained indirectly through comparison with experi-
ments of a-related measurable quantities.
Once a has been assigned as a parameter of the model, the average length
L of the tube is determined by the relationship

L= Na; (12.4.19)

In fact the tube itself is a random walk of step length a and, of course, the
end-to-end square distance of the tube must be the same as that of the chain.

FIGURE 12.4.3. A two-dimensional


picture of an entangled chain. The
dots are obstacles due to other
chains. The shaded region is the
space explored by the chain in its
short-time thermal motion.
The number N of tube steps is called the number of entanglements per chain. 343
It should not be confused with the number n of Kuhn segments of the chain. It
Polymers
follows from what was said above about the relationship between a and b that
n N.
We assume now that the chain has moved by diffusion a distance of order
L along its own tube. Since the diffusivity along the tube is D, the time required
to perform such a motion is of order

(12.4.20)

During such a time, Le., as a consequence of a curvilinear displacement of order


L, the center of mass of the molecule has been displaced by a straight distance
which is only of order (r2)1/2. Therefore, the diffusivity of the center of mass is
given by

(12.4.21)

where the M dependence is obtained by using equations (12.2.2) and (12.4.18)


and by considering that L is proportional to M. The molecular-weight dependence
predicted by equation (12.4.21) is well confirmed by experiments.
Note the price paid because the molecule is entangled. Imagine immersing
the same polymer molecule in a solvent such that the friction coefficient ~ is the
same as in the melt considered. Since the chain can move in whatever direction
it likes, D would directly be its center-of-mass diffusivity. Thus, as shown by
equation (12.4.21), the entanglements per se decrease the diffusivity by the
following factor:

(12.4.22)

Equation (12.4.20) can also be used to obtain another important result. We


assume that the polymer melt has been deformed. As a consequence, the system
has become anisotropic. Both the chain and its tube of constraints are no longer
random walks. How much time must elapse before equilibrium is restored? In
other words, how large is the relaxation time of this system?
The answer to this question is equation (12.4.20) itself, at least as an order
of magnitude. Indeed, after the chain has diffused a curvilinear distance of order
L, the old tube will have been abandoned by the chain. Now, since the chain
ends, as they diffuse out of the tube, choose their direction randomly, once the
deformed tube has been abandoned the new one is again a random walk. Thus
equation (12.4.20) also gives the relaxation time T:

(12.4.20')

The prediction that the relaxation time grows with the third power of M in
a concentrated system, as compared to about the second power of the dilute case,
is also in reasonably good agreement with experiments. Actually, the experimental
results indicate a power slightly larger than 3, and a debate is ongoing about the
source of this minor difference.
344 Building upon the idea of reptation, which was first proposed by De Gennes,
Doi and Edwards have constructed a complete dynamical theory which leads to
Chapter Twelve
a constitutive equation for concentrated polymeric liquids. There is no room here
to discuss this theory. Suffice it to say that the chain is modeled as a Rouse chain
constrained in a tube, and the assumption is made that the tube segments, during
a motion, deform affinely with the continuum. At the same time, the chain attempts
to reestablish equilibrium by reptating out of the deforming tube, thus con-
tinuously generating at the extremities new, randomly oriented, tube segments.
A simple prediction of the theory is that, in slow flows, the viscosity scales with
M in the same way as the relaxation time, say

(12.4.23)

It is noteworthy that, for viscosity, the difference with respect to dilute solutions
is even stronger than that found for either the diffusivity or the relaxation time.
In fact, since p. in a polymer melt is directly an intrinsic property (the melt
density is essentially independent of molecular weight), equation (12.4.23) is to
be compared with equation (12.4.15'), i.e., switching from the dilute to the
concentrated case implies a change from first- to third-power dependence on
molecular weight. Experiments essentially confirm this prediction, showing a
power slightly larger than 3 in concentrated systems.
It remains to discuss what happens in a polymer melt if the molecular weight
is progressively decreased. Obviously there must be a transition at some value
of M because, in the limit, the liquid even ceases to be polymeric. The transition
to a different kind of behavior is actually observed to occur at a critical molecular
weight Me which is still relatively large (order of thousands or tens of thousands,
depending on the polymer). Below the transition the behavior appears to be well
described by the Rouse theory or, roughly, by the dumbbell model. In other
words, the entanglements appear to have vanished.
The fact that Me is not too small is consistent with what was said before
about the relative magnitude between a and b. We should expect that there are
no entanglements if the end-to-end distance of the chains drops to values of
order a. The measured value of Me and the independent estimate of the value
of a are indeed mutually consistent. We note therefore that the entangled situation
is not obtained for all cases when polymers are concentrated; it is also required
that their molecular weight be large enough. If M < Me, a friction coefficient is
all that is needed to describe the dynamics, just as in the case of dilute solutions.
Finally, it is worth mentioning that the value of Me in concentrated solutions
increases with decreasing polymer concentration, as might be expected.

12.S. MIXTURES

Unerring Minos doomed me for the art


a/chemi/:, which I practiced among men
Dante, Inferno XXIX 119-120
Trans!. by D. L. Sayers
In the preceding sections, we have often considered polymers in solution 345
without as yet examining a fundamental question, namely, whether or not the
Polymers
entropy change associated with polymer mixing, either with a solvent or with
another polymer, obeys the ideal solution law which, we recall, is given by

(U.S.I)

where nl and n2 are the number of moles of the two components of the mixture
and Yt and Y2 the corresponding mole fractions.
Before examining this question, we must specify the state of the pure
components with respect to which SMIX is defined. Indeed, in dissolving a solid
polymer to form the solution, if the state of the pure polymer is, at least, partially,
crystalline, there is a positive entropy contribution due to phase change, which
it is appropriate to distinguish from that due to mixing in a strict sense.
The main part of the entropy change, which occurs when a polymer melts,
i.e., when it goes from the crystalline, ordered state to the amorphous state of
the liquid, is conformational. It has to do with the fact, discussed in the previous
sections, that many conformations are available to a chain in the amorphous
state, while a single one will exist in the crystal (excluding defects of the crystal
structure). It is readily estimated that the entropy change for a chain comprising
n segments is of order kn. We may use, for example, equation (12.3.1) giving the
entropy of a chain for which the end-to-end vector is assigned.
If it is assumed that, in the crystal, the conformation of the chains is the
fully extended one, for which r = nb, then equation (12.3.1) gives

SCRYST = -3kn/2 (12.5.2)

In the amorphous state, we can use equation (12.3.1) with the average value
of r2. Thus, since (r2) = nb 2, we get

SAMORP = -3k/2 (12.5.3)

The meaningful quantity is the difference between the two, i.e., the entropy
change due to disorientation. Since n is a large number, we obtain

SmSOR = 3kn/2 (U.5A)

Of course, equation (12.5.4) should not be taken literally for at least two
reasons. It is based on the random walk chain model and, furthermore, we have
used equation (12.3.1) with r = nb, i.e., beyond its range of validity. The result
is correct, however, in an order-of-magnitude sense. A similar result was obtained
long ago by Flory with the aid of the lattice model. Another calculation, referring
to the aliphatic chain, is given as an example at the end of the chapter. All of
these calculations predict a disorientation entropy which, as anticipated above,
is of order kn per chain, i.e., of order k per chain segment. Thus, although the
entropy of melting is large on a chain basis, reflecting the large increase in internal
degrees of freedom in the liquid state, it is of a comparable magnitude to that
of small molecules when referred to a chain segment or "monomer." This is
346 reflected in the fact that the melting point of polymers shows no unusual behavior,
the enthalpy of melting also being proportional to n, when referred to a chain
Chapter Twelve
made up of n monomers.
We now proceed to estimate the entropy of mixing in the proper sense. It
is understood that the disorientation entropy is omitted and that we take a
standard state for the pure polymer which is already amorphous, such as a
polymer melt (a possibly imaginary state at the temperature of interest).
Following Hildebrand, a simple way of dealing with the entropy of mixing
is to use the concept of free volume which, by definition, is the volume available
to the center of mass of any molecule in the system, properly accounting for the
volume occupied by all other molecules. Thus, the increase in entropy which
takes place upon mixing results from the fact that the molecules of each species
will, in the mixture, also enjoy the free volume made available to them by all
other species.
Let us first consider the case of ideal gases for which the free volume coincides
with total volume since, by definition (if one interprets ideal gases as being
Maxwellian gases), the molecules are volumeless mass points. When, at constant
temperature and pressure, we form a mixture of gases, each one of them expands
from the original volume to that of the mixture. Thus, in the case of only two
species, the change of entropy is given by

(12.5.5)

where V. and V2 are the original volumes of the pure gases, V = V. + V2 being
the mixture volume. Now, since in an ideal gas the volume fraction coincides
with the mole fraction, equation (12.5.5) is the same as equation (12.5.1), i.e.,
ideal gases obey the ideal mixing law.
Equation (12.5.5) is more general, however, as long as V, V., and V2 are
interpreted as free volumes. It can be applied to liquids as well, polymeric or
otherwise. The problem, of course, becomes that of estimating the free volume,
by no means an easy matter in the general case.
To proceed with the calculation. and, what is more important, in order to
show the essential peculiarities resulting from the polymeric nature of at least
one of the components of the mixture, we make the simplifying assumption that,
in the liquid state, all species have the same free volume fraction at any given
temperature and pressure. In other words, we assume that the larger the volume
occupied by a molecule, the larger the accompanying free volume, so that the
ratio of the two is constant. We further extend this assumption to mixtures, i.e.,
the free volume fraction in the mixture is the same as that in the pure components.
This extension is, of course, equivalent to the assumption of volume additivity
upon mixing (V M1X = 0).
With the assumptions stated above, free volume is proportional to volume
and equation (12.5.5) becomes

(12.5.6)

where quantities cf>J are the volume fractions of the two components in the mixture.
Needless to say, equation (12.5.6) coincides with equation (12.5.1) if the
molar volumes of the two species are equal (as for ideal gases). This may well
be the case for a liquid mixture of small molecules, at least in some instances, 347
and at least approximately more often. Conversely, when one component is Polymers
polymeric and the other is not, the difference between the two equations is very
significant. We will subsequently call "solvent" the low molecular weight com-
ponent, over the whole concentration range.
We note that, although equation (12.5.6) is based on a somewhat arbitrary
approximation, it is nevertheless expected to portray the essential features of
polymeric behavior, i.e., to be much closer to reality, than equation (12.5.1) could
ever be. In fact, in order to obtain equation (12.5.1) from equation (12.5.5) also
in the polymeric case, we would need to assume that a molecule of polymer, in
spite of its enormous size, carries with it the same absolute amount of free volume
as does a small molecule of solvent. This would clearly be an absurdity. The
assumption that the relative amount remains unchanged is obviously much better,
though corrections may be required.
Postponing further discussion of equation (12.5.6), we proceed to consider
also the heat of mixing so as to reach an expression for the free enthalpy of
mixing. Following Flory, the enthalpy change which occurs when mixing a
polymer (subscript 2) with a solvent (subscript 1) can be expressed in the form

(12.5.7)

The argument used to write this equation is as follows. By definition of the


dimensionless quantity X, kTX is the difference in energy when a single molecule
of solvent is dispersed in the pure polymer with respect to that of the pure solvent.
Since in the actual mixture the solvent-polymer contact is approximately reduced
in proportion to the volume fraction, the energy per solvent molecule becomes
kTXCP2. Then equation (12.5.7) is obtained for n l moles of solvent.
Combination of equations (12.5.6) and (12.5.7) gives for the free enthalpy

(12.5.8)

which is the celebrated Flory-Huggins equation.


Equation (12.5.8) needs to be discussed in several respects, primarily as
regards its essential message which is: polymers are more difficult to bring into
solution than small molecules. Very synthetically the reason is that, because of the
large molecular weight, the entropic contribution n2 In CP2 is smaller than for
ordinary molecules and therefore it becomes harder to overcome adverse energetic
contributions.
In order to appreciate this point in greater detail, let us first convert equation
(12.5.8) to the equivalent specific form, e.g., in terms of free enthalpy per unit
volume g*. If X is the ratio of the molar volumes of polymer and solvent (X is
a large number), then the volume fractions can be expressed as

and (12.5.9)

Finally, by setting CPI = cP (the volume fraction of the low molecular weight
component), CP2 = 1 - cP, and one obtains

g*(cp) = XcP In cP + (1- cp) In(1- cp) + XXcp(1- cp) (12.5.10)


348 Thus, the polymeric solution will be miscible throughout the 4> range if, and only
if, the following inequality is satisfied in the whole range (so that there is no
Chapter Twelve
spinodal region);

(12.5.11)

or, equivalently, if and only if

(12.5.12)

This equation sets an upper limit to the energy parameter X for achieving
complete miscibility. Now if X is, say, equal to I, i.e., if the solute molecule is
as small as the solvent molecule, this upper limit is X = 2. Conversely, if the
solute is polymeric, X I, and the upper limit becomes essentially X = !. This
fourfold decrease in the maximum allowable value of X makes it much more
difficult to find solvents for polymeric substances. Note how the effect is purely
entropic; it is brought about by the large size of polymer molecules.
The situation is much worse if both components of the mixture are polymers.
In such a case, the entropic term is almost negligible and even a very small
positive contribution from the energy term makes the polymers immiscible. To
see this clearly, we need to reformulate somewhat the energy term. Since now
both components are polymeric, we redefine X on the basis of a polymer segment
of the size of a monomer, which is the correct equivalent of a solvent molecule.
Thus, if the molecules of polymer 1 are made up of Y monomers, equation
(12.5.8) essentially becomes

(12.5.13)

The latter equation already tells the whole story. Since Y is a very large
number, unless X is very small the last term will predominate over the entropic
contribution. More formally, if we now call X the number of monomers per
molecule of polymer 2 and assume for simplicity that the monomers of the two
polymers have equal volumes, we get the symmetric formula

g*( 4 "" X4> In 4> + Y(1 - 4 In( 1 - 4 + XYx4>(1 - 4 (12.5.14)

and
(12.5.15)

the latter showing that, since both X and Y are large numbers, miscibility requires
that X be essentially zero (or, of course, negative).
Returning to the solvent-polymer case, it should be stressed that equation
(12.5.8) does not generally apply to extremely dilute solutions. In fact, the
derivation of the Flory-Huggins equation is based on the tacit assumption that
concentration is homogeneous, an assumption which breaks down in dilute
solutions. When there are very few polymer molecules in the solution, there will
be regions of pure solvent as well as regions, roughly spherical in shape, where
isolated polymer molecules will be present and where the local solvent concentra-
tion is smaller than the average. The Flory-Huggins equation starts to be valid
when the polymer concentration has reached the point where the chains overlap 349
extensively, so that everywhere the local concentrations of polymeric segments
Polymers
and of solvent are uniform.
We can readily estimate the critical concentration at which the chains will
begin to overlap. By taking the end-to-end distance as a measure of the coil size
we obtain, from either equation (12.2.2) or (12.2.3),

(12.5.16)

where c* is the number of polymer molecules per unit volume of solution when
the spherical domains come into contact. As previously, the reported powers of
M refer to the case of a (J solvent or to a good one, respectively. In terms of
volume, or weight fraction, the dependence on M obviously becomes

(12.5.17)

which shows that the larger the molecular weight, the sooner a polymeric solution
ceases to be dilute. In most cases a concentration of the order of 1% by weight
is already sufficient.
In the range of dilute solutions, the thermodynamic calculations become
more involved. Indeed, it is required to estimate both the entropic and the
energetic contributions as a function of the mutual distance of the chains, i.e.,
for different degrees of overlapping. First considering a system of only two chains,
it is found that free enthalpy generally decreases with increasing distance, i.e.,
the chains repel one another. From this result, one calculates the excluded volume
which can be attributed to each chain, and therewith the thermodynamic proper-
ties of the solution. The result holds up to second order in the concentration,
since only binary interactions were considered. Multiple interactions must be
accounted for to extend the results to higher order in concentration. Leaving
aside this complex matter, we only consider the special case where, in the binary
interactions, the entropic and energetic terms exactly match so that the excluded
volume vanishes. This is the situation in a (J solvent where the chains, while
repelling each other for entropic reasons, also attract each other because energy
favors polymer-polymer contacts better than polymer-solvent contacts and the
two opposing effects cancel out.
In such a case, since the polymer coils can interpenetrate freely, the whole
free volume of the mixture is available to the center of mass of each molecule,
without exclusions. Now, remembering that uniform availability of the free
volume is the hypothesis required for the Aory-Huggins equation to hold (quite
apart from other "minor" assumptions), it is concluded that equation (12.5.8)
also holds for a dilute solution in a (J solvent, up to quadratic terms in the polymer
concentration.
Rather than work with equation 12.5.8 (itself), it is better to first calculate
the chemical potential ILl of the solvent, which is obtained from equation (12.5.8)
by taking the baric derivative with respect to {II' With cf>1 and cf>2 given by equation
(12.5.9), one obtains

(12.5.18)
350 which shows that the 8 solvent corresponds to X = !, for which value the effects
of the binary interactions, represented by the quadratic term, vanish.
Chapter Twelve
Although equation (12.5.18) does not hold in general, its first-order term,
being independent of polymer-polymer interactions of any order, is of general
validity. Indeed, the first-order term corresponds to the classical first-order theory
of dilute solutions and expresses the colligative properties in the classical sense;
see Section 8.8. To within first order for the solvent, the Aory-Huggins equation
is in agreement with the classical colligative theory. Using only the first term of
equation (12.5.18), little algebra is needed to obtain the classical result for the
osmotic pressure:

pas = RTc' (12.5.19)

where c' is the molar concentration of polymer. However, it should be taken into
account that, in a dilute polymer solution (dilute on a weight percent basis), the
molar concentration of polymer is very low and hence colligative properties such
as the boiling-point change are very small; osmotic pressure is, however, still
quite within measurable ranges, since it is very large for ordinary low molecular
weight solutes. Furthermore, membranes which are truly semipermeable do exist
in the case of polymers, which are easily rejected for steric reasons because of
their large molecular size. Thus equation (12.5.19) may be used to determine the
(number average) molecular weight of polymers.
Returning now to more concentrated solutions, it may be interesting to
calculate the solvent activity, which is of importance in vapor-liquid equilibria
of solutions, devolatilization of polymer melts, etc. By definition, the left-hand
side of equation (12.5.18) coincides with the logarithm of the solvent activity at.
Therefore, by using the full expression on the right hand side and by neglecting
11X with respect to 1, we obtain

(12.5.20)

This equation shows that the activity never coincides with the volume fraction,
not even when X = 0, as might perhaps have been surmised from a superficial
comparison of equations (12.5.6) and (12.5.1), i.e., by presuming that the solvent
volume fraction plays in this case an analogous role to the mole fraction in ideal
solutions.
In the limit of q,2 approaching unity, i.e., for a polymer containing a very
small amount of solvent, equation (12.5.20) becomes

(12.5.21)

which shows that the modified Henry constant of the solvent (modified in the
sense that it is based on volume fraction rather than mole fraction) is given by
pS exp(1 + X), with pS the vapor pressure of the pure solvent. [This result could
be considered as stating that exp(1 + x) is the modified infinite-dilution activity
coefficient of the solvent.]
A final comment on the Flory-Huggins equation concerns the temperature 351
dependence of the energy parameter X. Since Polymers
S = -8G/8T (12.5.22)

then in order that equations (12.5.6) and (12.5.8) be consistent with each other,
one should find that X varies inversely with absolute temperature. Although
indeed X generally decreases with increasing temperature, an exact inverse propor-
tionality is a rare event. This implies that the last term in the Flory-Huggins
equation, describing the polymer-solvent interactions, also cont.ains entropic, not
merely energetic, contributions. In other words, by maintaining the form of the
Flory-Huggins equation, equation (12.5.6) should assume the form
(12.5.23)

For this reason, it is preferable to refer to X with the more generic name of
interaction parameter rather than energy parameter, though there is no doubt
that the greatest part of the interaction term comes from energetic contributions.
It is perhaps appropriate to recall again that the Flory-Huggins equation is
obviously an approximation, based as it is on the assumption of constant free
volume fraction throughout the range of concentrations. It is also a very simple
equation, however, which is unquestionably convenient for quick, yet reasonably
accurate estimates. Refinements, obviously at the expense of simplicity, have
been, and still are being worked out, especially for the case of polymer blends
where, as previously shown, fine details may become crucial in deciding about
the miscibility (or compatibility) of the two polymers.
So far, we have only examined the thermodynamics of mixtures in which
the polymer molecules are not stretched. This is not the only possibility, however.
A classical example where stretched chains coexist with a solvent is that of swollen
rubbers. Let us consider a polymer and an ordinary liquid which are miscible in
all proportions, so that the polymer will readily dissolve in the liquid. If we now
first cross-link the polymer to form a network and then insert the resulting piece
of rubber into the liquid (or expose it to its vapor), since the cross-linking
operation has not significantly changed the "chemistry" of the polymer, there
remains a tendency to mix which, in view of the obvious impossibility for the
rubber to go into solution, will result in the fact that the solvent penetrates the
network, swelling it up to some larger volume. In the swelling process the chains
get stretched. It is this progressive stretching that will eventually stop the swelling
process. The problem then becomes one of determining the equilibrium volume
ratio between the swollen and dry rubber.
To this end, we write the free enthalpy change which occurs upon swelling
as the sum of a mixing and an elastic contribution:

(12.5.24)

The mixing term is given by the Flory-Huggins equation which, in view of the
fact that the rubber network is a single giant macromolecule, is here rewritten
without the n2 term:

(12.5.25)
352 For the elastic contribution, we can use equation (12.3.9) of tre section on
rubber elasticity where, since the swelling process is an isotropic expansion, we
Chapter Twelve
can put

(12.5.26)

a 3 being the volume expansion ratio. However, equation (12.3.9) needs to be


modified for the case at hand. We want to switch from free energy, which is the
correct potential for isochoric transformations, to free enthalpy, which applies
to isobaric expansions as considered here. Therefore, we must add the entropic
term which accounts for the expansion of the "gas" of cross-links (that is, of the
nodes of the network), which occurs upon swelling of the rubber.
If we now indicate by n2 the mole number of the network chains, assuming
that four chains depart from each node in the network (tetrafunctional junctions)
the mole number of cross-links is n2/2. Thus R (n2/ 2) In( a3 ) is the entropy increase
resulting from this effect.
Then, using equation (12.3.9) in its extensive form, i.e., as the total free
enthalpy of a network comprising n2 chains, we finally write the elastic contribu-
tion as

(125.27)

We are now ready to calculate the chemical potential of the solvent in the
swollen rubber. By regarding a as related to the polymer volume fraction in the
swollen polymer, q,2, namely

(12.5.28)

and that, as before [see equation (12.5.9)],

(125.29)

where X is now the ratio of the molar volumes of the network chain and solvent,
baric differentiation with respect to nl yields

(f.1 - f.~)/ RT = In(1- q,2) + q,2 + Xq,~ + (q,~/3 - q,J2)/ X (125.30)

where the rightmost term arises from the network elasticity.


Now, if the rubber is immersed in the liquid or exposed to its saturated
vapor, equilibrium will be reached when f.1 = f.~, i.e., when q,2 satisfies the
following equation:

0= In(l - q,2) + q,2 + Xq,~ + (q,~/3 - q,2/2)/ X (12.5.31)

It should be noted that, as long as X :5 !. i.e., for a solvent which mixes with the
parent polymer in all proportions, this equation would never be satisfied without
the elastic term; more precisely, it would be satisfied with q,2 = 0, i.e., for an
infinitely large swelling ratio. Conversely, with the elastic term included, a solution
is always found with 4>2 in the interval 0 < 4>2 < 1, which defines the equilibrium 353
swelling ratio 1'*. If 1'< 1'*, further solvent wants to penetrate the rubber. On Polymers
the other hand, if a> 1'*, the solvent is squeezed out as a result of rubber
elasticity. Similar considerations apply to the case of exposure to a nonsaturated
vapor, i.e., if the activity of the solvent is less than unity.
Although qualitatively the situation is as just described, quantitatively the
above equations should not be taken too literally and, in fact, they do not compare
too favorably with data. Now, quite apart from the minor imperfections of the
Flory-Huggins equation, which were discussed previously, there are major uncer-
tainties about the elastic term, and even about the additivity assumption embodied
in equation (12.5.24). It should be remembered that the rubber elasticity theory,
as presented here, refers to a highly idealized situation. To mention just one
source of possible deviations, we note that, when the polymer is cross-linked by
some chemical means, many entanglements that the chains formed get trapped
in the network. These topological constraints add further connectivity to the
network, which is ignored in the classical theory. The theory refers to a network
of so-called phantom chains, which can cross each other freely. Although there
have been many attempts to account for entanglements in rubber elasticity, there
is no general consensus as to the appropriate way of dealing with this problem.
As far as the additivity assumption is concerned, namely equation (12.5.24), there
have been speCUlations that the solvent does not behave in exactly the same way
when mixing with a relaxed polymer, as in an ordinary solution, or with stretched
chains, as in a swollen network. In the latter case, owing to local anisotropy
existing in the neighborhood of a stretched chain, the solvent close to the chain
would be induced by a polarization effect to orient somewhat, thus changing the
entropy as well as the interaction with the polymer.
Whatever the present limitations of the theory, the fact remains that, when
dealing with polymeric substances, one frequently encounters practical situations
where the molecules are stretched. Suffice it to mention processes where a
polymeric solution or a melt is subjected to a flow or deformation. The thermo-
dynamic effects arising from chain stretching can be significant. There exists
accumulating evidence showing that the polymer-solvent and polymer-polymer
compatibilities (or miscibilities) are changed during a flow process. Effects due
to chain stretching are also important in crystallization processes, to the point
that crystallization will take place at a temperature at which the relaxed chain
would not crystallize (see also Section 5.4). This effect had long been known for
certain rubbers, and was subsequently demonstrated for flowing polymers, both
in solution and in the melt.
The rationale for a change in crystallization temperature brought about by
chain stretching is readily found. We recall that the equilibrium temperature at
which a phase change occurs satisfies the equation

T* = DR/DS (12.5.32)

where DR and DS are the jumps of enthalpy and entropy across the phase
change. For the liquid-solid transition of ordinary substances, since both phases
are rather insensitive to pressure, there is no easy way of changing T*. For the
case of polymers, however, we enjoy an extra degree of freedom, Le., we can
354 stretch the molecules in the liquid (or rubbery) state. Both DH and DS will
Chapter Twelve
change in general but, prominently, there will be a change in DS upon stretching
the chains in the liquid. The change is in the direction of decreasing DS since
the stretched chains have less conformational entropy, i.e., they are in this respect
closer to the crystalline state. The transition temperature accordingly increases.

12.6. RIGID POLYMERS

Salia sapida, polyedra, diaphana, multiplicativa, solubilia in


particulas minimas injinitas, nihilominis semper con/ormes,
concrescentesque iterum iterumque in majores etiam conformes,
generant crystallisando, in et ex terris variis varios lapides.
Lynnaeus

In all previous sections, we have dealt with flexible polymers, particularly,


linear ones. As anticipated in the introductory section, however, this is not the
only possibility. Even within the realm of flexible polymers, there exist other
structures of the macromolecule which are not linear. Different types of branched
structures are a relatively common occurrence, low-density polyethylene being
one of the most important examples. Recently, also ring-shaped polymers have
attracted attention, especially in relation to possible mechanisms of relaxation
in the melt other than reptation. Although some of the results presented in
previous sections require to be modified when the structure is not linear, the
main features of the described behavior remain essentially unaltered.
Further complications may arise when, for example, the chain comprises
sequences of different monomers, i.e., when it is a block copolymer [see the
discussion following equation (9.2.3)], or when it contains electric charges, as in
most water-soluble polymers. In these cases, the effect of specific interactions
may become very important. For example, one might need to account for the
repulsions between different moieties in a copolymer or between like charges in
a polyelectrolyte, etc. Yet, even in these cases, the basic framework remains the
same. One is always dealing with long, chain-like objects which enjoy a large
number of internal degrees of freedom and are therefore capable of assuming
an even larger number of different conformations.
The situation changes drastically when rigid or semirigid polymers are
considered. By definition, a rigid polymer is one which exists in a single well-
defined conformation. There are essentially two types of rigid polymers, for which
the rigidity has a completely different origin. One of them is typified by the
globular proteins, which are linear macromolecules folded rigidly onto themselves
to form globules. In this case, the rigidity is brought about by secondary bonds,
which link together different segments of the folded chain. The chain per se is
not rigid and, indeed, under suitable conditions, such as by raising the tem-
perature, the protein may unfold to become a random coil. Here, we are not
interested in this kind of rigidity.
The case to be discussed IS that for which the rigidity is determined by the
nature of the primary bonds along the chain. In particular, we are interested in
the case where the single conformation available to the macromolecule is straight,
i.e., the polymer molecule is rod-like. The particular effects exhibited by these
polymers arise entirely from this geometrical feature, i.e., from the large aspect 355
ratio of their rod-like shape. Although the theory is well developed for the model
Polymers
of rigid rods only, some of its results are expected to apply, at least qualitatively,
to semirigid polymers as well. The chain of a semirigid polymer allows for some
flexibility and, consequently, a very long chain will be coiled up. At the same
time, however, the chain is rigid over shorter distances, i.e., the local configuration
is rod-like. It should be finally mentioned that long rod-like objects are not
necessarily single molecules. The tobacco mosaic virus is a classical example in
this regard. The theory applies in general to a suspension of rods as long as the
rods are Brownian particles, i.e., as long as they are small enough to be subjected
to Brownian motion.
The most spectacular effect of the rod-like shape is the spontaneous transition
from an isotropic distribution of the rod orientations to an anisotropic distribution,
brought about by an increase of concentration (in the terminology of Section
4.4, we are considering a lyotropic transition). In the oriented phase, the rod-like
molecules are roughly parallel to one another although their centers are distributed
randomly in space. The system remains a liquid but, since there is a degree of
order which extends over large distances, the behavior is also crystal-like. In
other words, the system belongs to the category of liquid crystals.
In order to understand how this isotropy-anisotropy transition may occur,
in the absence of any orienting force, one must consider again the concept of
excluded volume. With reference to binary interaction only, the excluded volume
of hard particles is conceptually very simple and also generally easy to calculate.
The simplest case is that of a system of hard spheres of equal diameter d. The
excluded volume, i.e., the volume where the center of the second sphere cannot
be found owing to the presence of the first, is clearly that of a sphere having
diameter twice as large, i.e., 2d.
In the case of hard rods with diameter d and length I, the excluded volume
depends on the relative orientation of the two rods. If they are parallel (see
Figure 12.6.1a), the excluded volume is of order d 2 /, i.e., as in the case of spheres,

a
b

FIGURE 12.6.1. (a) The excluded volume in the binary interactions of rigid rods for a
parallel configuration. (b) The same quantity for a perpendicular arrangement.
356 it is of the same order of magnitude as the particle volume. Conversely, if the
Chapter Twelve
two rods are perpendicular to each other (see Figure 12.6.1b), the excluded
volume becomes of the order of d1 2 , i.e., much larger. If the aspect ratio 1/ d is
a very large number, the "parallel" excluded volume can be neglected with respect
to the "perpendicular" excluded volume. Then, by calling 0 the angle formed
by the rod directions in binary interactions, one may write

(12.6.1)

where, as usual, the operator ( ) indicates an ensemble average which, this time,
is over the orientational distribution of the rod-like particles.
Now, if the number of particles per unit volume of suspension satisfies the
inequality

c 1/ dl 2 (12.6.2)

then the collective excluded volume of the particles is much smaller than the
suspension volume even if (sin 0) is of order unity. As a result, both the positional
and orientational distributions of the rods will be random so as to maximize the
entropy.
Conversely, let us assume that c is increased to values of order 1/ df or
larger. Then, with a (sin 0) value of order unity, as pertains to an isotropic
orientational distribution, the excluded volume would fill up the entire available
space, i.e., the particles would lose the translational degree of freedom, and the
corresponding positional entropy would drop to negligible values. If, however,
the particles choose to orient so as to become almost parallel to each other, then
(sin 0) approaches zero and the translational degrees of freedom are restored. In
other words, in this range of concentrations two opposing effects come into play.
By decreasing (sin 0), the orientational contribution to entropy decreases but,
simultaneously, the positional contribution increases because of the correspond-
ing reduction in the excluded volume. A compromise will be found with an
appropriate anisotropic orientational distribution which maximizes the total
entropy.
The detailed calculations are somewhat involved and will not be reported
here. The results show the following behavior. By increasing the concentration,
the suspension remains isotropic up to a critical concentration c1 If C1 < c < C2,
where C2 is a second critical concentration, two phases coexist, having concentra-
tions C1 and C2, respectively, the former isotropic and the latter liquid crystalline.
Finally, if c > C2, only the anisotropic phase exists, with a degree of orientational
order which progressively increases as c increases.
Both C1 and C2 are of order 1/ d1 2 We note that, if the aspect ratio of the
rods is large, such a concentration level can be fairly low. In terms of volumetric
fraction, it becomes

(12.6.3)

i.e., it is well below the close packing limit, which is approached when cP becomes
of order unity. Thus, a theory based on binary interactions alone is a good
approximation, and it becomes exact in the limit of an infinite aspect ratio.
We note further the peculiar character of the equilibrium between the two 357
phases with concentrations C 1 and C2' Since no energetic contribution is involved, Polymers
not only is the chemical potential the same in the two phases, but also the partial
molar entropy Sf is continuous across the phase transition (the latent heat is
zero: in this sense, this is a secondary phase transition; see Section 4.5). The two
phases differ only in the way the same Sf is obtained. This is possible because
we now have an internal state variable, which is the degree of orientational order.
If we call p this orientational parameter, defined in such a way as to become
zero in the isotropic solution, and if S~ is given by a constitutive function s~( C, p),
the phase equilibrium condition is

(12.6.4)

where P2 is the value of the orientational parameter in the anisotropic phase and
J = 1,2 indicates the solvent and the rod-like solute.
Of course, in actual systems energetic effects will also be generally present
and become superimposed on the purely entropic effects just discussed. Energetic
interactions can also induce anisotropy on their own account. Thus, in many
cases, the liquid crystalline state is brought about by a combination of factors.
Small molecules which give rise to liquid crystalline phases (such as discussed
in Section 4.4) are typically elongated in shape, Le., they might crudely be
compared with rods having a small value of the aspect ratio. This geometrical
factor, which by itself might be insufficient to induce anisotropy, is enhanced by
coupling with some form of anisotropic polarizability that reinforces the tendency
of the molecules to align parallel to each other.
Returning to long rods, i.e., to the polymeric case, we now briefly examine
the effects of this shape on transport properties, namely, on diffusivity and
viscosity. The friction coefficient of a rod varies somewhat depending on whether
the rod moves parallel or perpendicular to its own axis. The difference is only
by a factor of 2, however, and, to within a logarithmic factor containing the
aspect ratio, the friction coefficient is proportional to rod length. Therefore, in
dilute solutions, the diffusivity is obtained as

~=M; D= l/M (12.6.S)

In the case of rods, it is useful to introduce the concept of rotational dijfusivity,


which is linked to the rotational rather than the translational Brownian motion
of the particle. Since a diffusivity measures the square of the distance travelled
by the particle per unit time, but in this case the "distance" is an angle, the
dimensions of the rotational diffusivity DR are those of reciprocal time.
It will now be assumed that the orientational distribution of the rods has
been distorted from its equilibrium value as a consequence of, say, a flow process.
When the flow is arrested, the distribution relaxes back to equilibrium in a time
which is of the order of 1/ DR' Thus, 1/ DR measures the relaxation time of the
system. Hence, the question becomes: how does DR depend on rod length (or
molecular weight)?
To answer this question simply, let us consider the rigid dumbbell depicted
in Figure 12.6.2, which is substantially equivalent to the rod. The beads must be
358 endowed with a translational diffusivity which is of the order of that given by
equation (12.6.5), i.e., inversely proportional to M. When the dumbbell rotates
Chapter Twelve
through an angle 6, the bead displacement is of order 61. Hence, the time
required for such a displacement is of order (61)2/ D. It follows that

(12.6.6)

The difference between the rigid dumbbell and the rod-like particle is only
in a logarithmic term containing the aspect ratio, which results from considering
the hydrodynamic interactions between the various segments of the rod. Thus,
equation (12.6.6) is essentially correct, i.e., the relaxation time of a dilute solution
of rod-like polymers scales with about the cube of molecular weight.
It can further be shown that the contribution to viscosity due to rod-like
polymers in dilute solutions is related to the relaxation time in the same way as
for the flexible ones, i.e., that [compare equations (12.4.12) and (12.4.13)]

,",'= ckTT (12.6.7)

where c is the number of rods per unit volume. It follows that, in the present
case, the intrinsic viscosity scales with M according to

(12.6.8)

It should be noted how the rigid polymers give rise to larger power exponents
than encountered in the flexible case.
The difference in the dynamics between the rigid and flexible polymers
becomes even more pronounced in the concentrated case. The rod-lii5e solution
becomes concentrated in a dynamical sense when there is a sizeable amount
of interparticle collisions such that the particle rotations become hindered.
Since the volume spanned by a rod in its rotational motion is of order the e,
hindrance to rotation will start to occur above a critical concentration of the
order of

c* "" 1/13 (12.6.9)

This criterion does not involve the rod diameter d, but only its length. Since
1/ d ;> 1, c* is much below the concentration for the isotropy-anisotropy transition
discussed earlier.

FIGURE 12.6.2. The rigid dumbbell: a simplification of the rod-like molecule.


When the rods become concentrated in the sense discussed above, the 359
calculation of their rotational diffusivity becomes more complex. In Figure 12.6.3,
Polymers
a test rod is depicted together with neighboring rods. Owing to collisions, the
test rod is encaged in a sort of tube the diameter of which can be estimated
to be

(12.6.10)

If c c*, then a I. Within the tube, the rotations are constrained to small
angles of order a/ I, i.e., the rod is entangled in much the same way as for the
con.centrated flexible case.
In order to perform sizeable rotations, i.e., much larger than a/ I, the rod
must diffuse longitudinally. Each time it diffuses longitudinally by a distance ~
the orientation will change by an amount of order a/ I. Now, since the time
interval t to perform such a step is of order

(12.6.11)

where D is the translational diffusivity, it follows that in this case we find

(12.6.12)

Making use of equation (12.6.6) for the M dependence of D/12, and of equation
(12.6.10) for the value of a, equation (12.6.12) gives

(12.6.13)

where c' is, as before, the polymer concentration by weight.


Finally, because equation (12.6.7) continues to remain valid also in the
concentrated case, the viscosity is obtained in the form

(12.6.14)

where we have written JL instead of JL' because the solvent contribution is


negligible in the concentrated case.

FIGURE 12.6.3. Rigid rods in a concentrated


solution. The test rod is encaged in a tubular
region of diameter a.
360 We note how much larger are the power dependencies for rigid polymers
Chapter Twelve
as compared to those found in the flexible case. The experiments fully conffrm
these predictions, at least qualitatively. The experimental power laws are actually
slightly weaker than predicted by the equations above, but this discrepancY is
probably due to the fact that most so-called rigid polymers are in fact semirigid
to some extent.
Equation (12.6.14) shows that the viscosity of a rigid polymer solution grows
rapidly with the concentration. However, this result only applies as long as the
orientational distribution remains isotropic. After the isotropy-anisotropy transi-
tion, the situation changes drastically. In the liquid crystalline phase, the viscosity
suddenly drops to a much smaller value and, furthermore, by further increasing
c above C2, the viscosity continues to decrease, at least for a range of concentra-
tions.
The explanation of this effect is again to be found in the behavior of the
rotational diffusivity. When the solution becomes anisotropic, i.e., when the rods
choose to align almost parallel to one another, the tube of constraints depicted
in Figure 12.6.3 suddenly increases in diameter and, as a consequence, the rod
can change its orientation more rapidly. A drop in viscosity is the immediate
consequence of the increase in rotational diffusivity. The further decrease of
viscosity observed for c increasing above C2 has the same explanation, since the
degree of order also increases with increasing c. It is noteworthy that there is no
contradiction between the fact that the rods become almost aligned and that, at
the same time, they can change their orientation more easily. Thermodynamics
imposes on them that the angle explored as a consequence of their rotational
Brownian motion be smaller. At the same time, they can explore this limited
angular space more rapidly, because of reduced hindrances.
The drop in viscosity due to the isotropy-anisotropy transition is so typicl
that it is often used as a viable experimental method of determining the occurrence
of the transition. The lower viscosity of the liquid crystalline phase is also
technologically important with regard to the processability of these substances.
We conclude by recalling the various ranges of concentrations where the
different behaviors are encountered. In terms of volume fractions, the characteris-
tic polymer concentrations are ordered as follows:

(12.6.15)

where cp* = (d/I)2, while CPt and CP2 are both of order d/ L
If 0 < cP < cP *, the solution is dilute and, of course, isotropic. The relaxation
time and the viscosity increment scale approximately with M3 and c' M2, respec-
tively. If cp* < cP < CPt, the solution is concentrated but still isotropic. The relaxa-
tion time and the viscosity scale with C'2 M7 and C'3 M 6, respectively. In the small
range CPt < cP < CP2 two phases exist, one isotropic and the other liquid crystalline.
For cP > CP2, the solution is anisotropic and has a smaller viscosity and relaxation
time than the limiting isotropic solution at CPt. However, when the close packing
condition is approached, the polymer-polymer friction dominates over the poly-
mer-solvent friction and the viscosity will rise again. Indeed, there will eventually
be a transition to a solid state, be it glassy or crystalline. Finally, the inclusion
of energetic effects in the above picture may result in the fact that transition to 361
a solid state occurs much earlier, thus canceling some of the above regimes.
Polymers

ACKNOWLEDGMENT. G. Astarita should be credited for editing the original


manuscript of this chapter by introducing appropriate references to concepts and
equations encountered in the rest of the book. The beautiful quotations at the
beginning of each section are also his choice.

EXAMPLES AND PROBLEMS

Examples

1. We wish to obtain the result in equation (12.2.2) directly, i.e., without using the
distribution function. We refer to Figure 1.2.2 which shows the chain of Kuhn segments,
or random walk. If we indicate by bI the vectors corresponding to the chain segments,
the end-to-end vector is obviously obtained in the form

(12.E.l)

where the sum is over the n segments. Therefore

r2 = r r = (I b I ) (I bJ )
= nb 2 +I I bI bJ (12.E.2)

where the last double sum is over J different from 1. When taking the ensemble average
of this equation, the last sum vanishes because, when I differs from J, there will be as
many positive terms as there are negative terms, and hence equation (12.2.2) is obtained.

2. For a shear deformation of small amplitude, two principal directions of deformation


are, with reference to Figure 12.3.1, 45 in the plane of the figure, while the third is
orthogonal to the plane. Using simple geometrical results we obtain

a~ = [1 + (1 - 'd]/2; a~ = [1 + (1 - y)2]/2; a~ = 1 (12.E.3)


Therefore

U12 = dA/ dy = ckTy (12.E.5)

3. We wish to find the disorientation entropy of the aliphatic chain by assigning equal
probabilities to trans and gauche conformers. Since to each next carbon atom in the chain
there are one trans and two gauche alternatives, the number of conformations is of order
3 n , where n is the number of carbon atoms along the chain. All of these conformations
have equal probability. Therefore

8 D1SOR = k In(3 n ) = kn In 3"" kn (12.E.5)

4. The mean-square end-to-end distance of an isolated chain in a good solvent is


roughly (i.e., ignoring numerical factors) given by [compare equation (12.2.3)]

(12.E.6)
362 We wish to calculate the concentration dependence of the end-to-end distance in the
concentrated case.
Chapter Twelve The polymer volume fraction </>* at which the isolated chains will just "touch" each
other (nb 3 is taken as the volume occupied by the segments of the chain) is given by
(12.E.7)
By assuming power-law dependencies in the concentrated case as well, we write
(r2)CONC in the form

(12.E.8)
which is dimensionally correct, and automatically satisfies the crossover between the two
regimes. The unknown value of m is now found by imposing the condition that (r2)CONC
be proportional to n since, in the concentrated case, the excluded volume effects are
screened out at the scale of the chain dimensions. Thus, from the condition
6/5+4m/5=1 (12.E.9)
we get m = -1/4. The result is therefore
(r 2>CONC = nb2c/J -1/4 (l2.E.IO)
This result, first obtained by Daoud, is in good agreement with neutron scattering
experiments of deuterated chains in concentrated solutions.

Problems

12.1. Find the general relationship between F and r which replaces equation (12.3.3) for arbitrary
chain extensions. [Hint: The distribution of segment orientations in the chain subjected to force F
is the Boltzmannian C exp(Fb cos 6/ kT), where 6 is the angle formed by the segment with the
direction of the force and C is a normalization constant. Next calculate r as nb(cos 6).]
12.2. Find the stresses generated by deforming a rubber by a traction in direction 1 accompanied
by a contraction in direction 2, while in direction 3 the rubber dimension stays constant (planar
extension).
12.3. Explain the justification for using results which apply to free chains, such as the Gaussian
distribution of equation (12.2.1) [from which equations (12.3.2), and hence (12.3.9), are obtained],
to predict the behavior of deformed rubbers. In a deformed rubber, the chains are not free nor are
they randomly distributed.
12.4. By assuming that, in a concentrated solution of entangled polymers in good solvents, the
concentration dependence of the diameter of the tube of constraints is of the form a = </> -3/4, find
the concentration dependence of the relaxation time by further assuming that the friction coefficient
stays constant.
12.S. Find the limiting value of the interaction parameter for complete miscibility of two
polymers having X = Y = 100.

LITERATURE

To Thales . .. the primary question was not "What do we


know, " but "How do we know it. "
Aristotle

The fundamental reference for the equilibrium thermodynamics of polymers is the


classic book by P. J. Flory, Principles of Polymer Chemistry, Cornell University Press,
Ithaca, NY (1975), first published in 1953. Questions of chain conformation, rubber
elasticity, and equilibrium properties of mixtures are treated there, at least in their basic 363
aspects.
The book by P. G. De Gennes, Scaling Concepts in Polymer Physics, Cornell University Polymers
Press, Ithaca, NY (1979), is an excellent modern monograph on solutions in good solvents.
Both equilibrium and transport properties are considered, although the latter also refer,
as in this chapter, to the equilibrium strncture of the material, i.e., when the gradients are
small. .
The fundamental reference for transport properties of polymers, both close to and
far from equilibrium, is the recent book by M. Doi and S. F. Edwards, The Theory of
Polymer Dynamics, Clarendon Press, Oxford (1986), where also the case of rigid polymers
is considered. These authors are responsible for the most important advancements in the
theory of dynamics of entangled systems.
A detailed treatment of various kinds of dumbbell models or of multiple bead-spring
and bead-rod chains, and a prediction of the corresponding rheological properties, can
be found in the book by R. B. Bird, C. F. Curtiss, R. C. Armstrong, and O. Hassager,
Dynamics of Polymeric Liquids, Vol. 2, Wiley, New York (1987).
Chapter Thirteen

THERMODYNAMICS OF
ELECTROMAGNETISM

R. E. ROSENSWEIG
Exxon Research and Engineering Company
Annandale. New Jersey

Be sure to ask your teacher


his reasons and sources.
Rashi

NOTATION

A Helmholtz free energy per unit mass Uoule/kg) m2 S-2


.sa Helmholtz free energy per mole m2 kg S-2 mol-I
A Augmented Helmholtz free energy per m2 S-2
unit mass Uoule/kg)
B Magnetic induction (tesla) kg S-2 A-I
Ci Mol fraction of species i
Cn Surface couple density kg S-2
C Surface couple stress tensor kgs- 2
D Displacement field m- 2 sA
e Electric potential (volt) m2 kgs- 3 A-I
E Electric field (volt/ meter) mkgs- 3 A- '
g Acceleration of gravity ms- 2
G Gibbs free energy density Uoule/kg) m2 S-2
G Body couple per unit volume m S-2
H Magnetic field (ampere/meter) m-IA

I". Electric dipole moment PV,I eo of a total system m4 kgs- 3 A-I


It,m Magnetic dipole moment MY, of a total system m2 A 365
366 .1 Magnetic moment per mol m 2 A mol-l
Chapter Thirteen Jr Free current density m-2 A
ko Boltzmann constant (ko = R/ N), m 2 kgs-2 K- l
1.38062 x 10-23 joule/K
Kp Equilibrium constant expressed in
atmospheric partial pressures
m Mass m
M Magnetization m-1A
.J(, Molecular weight kg mol-l
n Number of moles mol
N Avogadro number, mol- l
6.02217 x 1023 particles mol- l
p Polarization vector m-2 sA
p Pressure in the presence of field m- l kg S-2
Q Heat added per unit mass m 2 s- 2
q Heat flux vector kgs-3
R Gas constant per mol m 2 kg S-2 K- l mor l
S Entropy per unit mass Goule/kilogram K); m 2 s- 2 K- l
also, area of closed surface m2
Y' Entropy per mole Goule/mol K) m 2 kg S-2 K- l mol- l
tn Stress vector m- l kgs- 2
T Temperature K
T Surface stress tensor m- l kgs- 2
U Internal energy density m 2 s- 2
v Mass velocity vector ms- l
V Volume m3
W Work done on a system m 2 kg S-2
Xi Mass fraction of species i
E Permittivity (general) m- 3 kg- l S4 A2
Eo Permittivity of free space, m- 3 kg- l S4 A2
8.854 X 10-12 farad/meter
l Magnetomotive force A
(J Entropy production rate per unit volume m- l kg S-3 K- l
K Thermal conductivity kgs- 3

JL Chemical potential per unit mass; m 2 s- 2


also, magnetic permeability
(units of JLo), JL = B/ H
JLo Standard value of chemical potential m 2 s- 2
I-'s Bohr magneton, 9.27410 x 10- 24 joule/telsa 367
1-'0 Permeability of free space, Thermodynamics of
41T x 10-7 henry/meter Electromagnetism

IIi Stoichiometric coefficient mols


g Extent of reaction
p Mass density m- 3 kg
(T Surface charge density (coulomb/m2) m- 2 sA
conductivity, siemen
Electril~al m- 3 kg-I S3 A2
\) Specific volume kg-I m3
Magnetic flux m2 kg S-2 A-I
Dissipation function m- I kg S-2
Susceptibility (X = M / H)

Subscripts

Total
Species i; also, internal
m Magnetic; also, per unit mass
e Electric; also, external
n Normal
1/ Viscous
v Per unit volume
o Value in absence of a field (except in 1-'0 and Eo)

Superscripts

* Modified
(over) Denotes partial quantity
o Reference value
Field averaged quantity
Denotes evaluation in the material frame of reference

Notes

(i) The mole is the amount of substance of a system which contains as many
elementary entities as there are atoms in 0.012 kilogram of carbon-12.
(ii) iYl denotes a closed region with a piecewise smooth boundary aiYl.
13.1. AN OVERVIEW

All the phenomena of magnetism offer /illie difficulty in their


explanation. It is no longer considered to be the action of an
incomprehensible allraction completely similar to the occult
faculties of Aristotle. Each body has its poles and its surfaces;
the universal fluid, composed of a twofold stream, penetrates
this body by means of each pole. This fluid always keeps the
same direction, as long as that direction is not altered by
another current which is stronger than the first. This is
what constitutes the reinforcement of mineral magnetism
as well as that of animal magnetism.
Translation from Discours de M. Mesmer
sur Ie Magntisme (1784)

In the past chemical engineers have paid scant attention to the thermodynamics
of electromagnetic systems. With the advent of new processing concepts, novel
systems, and new materials including higher temperature superconductors, the
topic is assuming a greater importance. This treatment introduces the student or
graduate to the subject, building on fundamental principles in sequential
fashion with the goal of encompassing both physicochemical and transport
principles. A number of diverse but related topics are developed in a unified
fashion in comparison to the existing literature in the area which is incom-
plete, scattered, and often deficient. Certain of the topics developed are
important in current engineering practice and others represent areas for future
innovation.
Our treatment begins with a review of electromagnetic units and the statement
of Maxwell's equations, of which use will be made of every term at some point
along the way. However, most of the subject matter can be understood with
concepts of the uncoupled magnetostatic and electrostatic fields, and the corre-
spondence between these analogous systems is developed initially. Obtaining
correct expressions for electromagnetic work is of paramount concern. Because
of its importance this task is approached from several directions: via one-
dimensional systems wherein the physics is easily understood although the gener-
ality of results is uncertain; from the complete Maxwell's equations which yield
the generally accepted but limited result applicable to fixed volumes only; and
finally with a treatment based entirely on macroscopic classical concepts that
broadens the Maxwell result to apply for a substance that changes its mass
density in the process of becoming magnetically or electrically polarized. This
fundamental result is subsequently used in developing all the remaining topics
in the chapter. 369
370 The culminating topics include applications to problems of phase change,
chemical reaction, and other equilibrium processes as well as the formulation of
Chapter Thirteen
rate processes in dissipational systems of momentum and mass transport with
polarizable species.

13.2. ELECTROMAGNETIC UNITS

Nowhere is there a greater need for meticulous attention to the dimenSional


homogeneity of equations than in electromagnetism, especially when mingled
with thermodynamics parameters. We use SI units (Systeme International
d'Unites), the base units for which are taken from the rationalized mksa system
of units: Distances are measured in meters (m), mass in kilograms (kg), time in
seconds (s), and electric current in amperes (A). The qualifier "rationalized" is
used because a factor of 411" is introduced to simplify the appearance of Maxwell's
equations at the cost of complicating the appearance of Coulomb's law if it were
to appear; the latter, however, is not used in this chapter.
The notion of a magnetic field H (or electric field E) simplifies the detailed
description of external conditions. Thus, instead of specifying for a given test
that the magnet (electrode) was located at a particular distance and orientation
and was constructed with certain specifications, it may be said that the apparatus
was placed at a given location in a field H (or E).
In SI, magnetic field H has units of amperes per meter. An induction field
B (in tesla) is defined such that in vacuum B = #oH while in the presence of matter

B = #o(H+M) (2.1)

The parameter #0 is called the permeability of free space and has the value
#0 = 411" X 10-7 H m-t, where H (not to be confused with the magnetic field)
stands for the henry. Other numerically equivalent units of #0 are tesla-meter
per ampere (T m A-I), and newtons per square ampere (N A-2) or m kg S-2 A-2,
and m2r N- I The quantity M denotes magnetization, a material property denot-
ing the state of magnetic polarization of magnetized matter. It can be seen from
equation (2.1) that only in the presence of magnetized matter does B differ from
#oH.
Materials scientists frequently work in the cgs system of units in which
distance is measured in centimeters, mass in grams, time in seconds, and electric
current in amperes. In a vacuum in the cgs system and B field (in gauss) is
numerically equal to the H field (in oersteds) and B = H + 411"1 = H + M where
1 is the intensity of magnetization and M is known as the ferric induction. The
SI unit of magnetic induction, the tesla, is equal to 104 gauss.
The Earth's magnetic induction is about 7 x 10-5 T (0.7 gauss) in magnitude
with a horizontal component at the latitude of New York City of about 5 x 10-5 T;
the field of a ceramic permanent magnet is typically 10-1 T, and an iron yoke
laboratory electromagnetic 2 T. Liquid-helium-cooled (4.2 K) superconducting
magnets made of metal alloys routinely produce magnetic induction of 5 T and
have been operated at 20 to 50 T. The limiting high temperature above which a
superconductor reverts to a normal conducting state was dramatically increased 371
in 1986 with the discovery of superconductivity in a class of rare earth copper
Thermodynamics of
oxide compositions for which J. G. Bednorz and K. A. Muller won the Nobel Electromagnetism
Prize in physics in 1987. Operation above the liquid nitrogen temperature (77.3 K)
has been demonstrated and there is general optimism for achieving higher
temperature operation, possibly above room temperature in suitably modified
materials. The cost of replacing liquid nitrogen coolant is less by a factor of
about 500 than liquid helium due to the 10 times larger volumetric heat capacity
and 50 times cheaper cost of the nitrogen.
Electrostatic fields are conveniently produced by applying a potential
~ifference to an electrode pair. Because electrodes can be configured as thin
conducting surfaces, they are often quite compact (think of the transparent
electrode in the liquid crystal alphanumeric display of a wristwatch). However,
the achievable energy density of magnetostatic fields exceeds that of electrostatic
fields as free magnetic charges (monopoles) are unknown and breakdown does
not occur as field intensity increases.
It will be recalled that the strength and direction of an electric field is
described at each point by a vector E such that the force acting on a small
stationary test charge q placed at this point is q E. The vector E is called the
electricfield. The conceptually analogous magnetic vector is H. Magnetized matter
behaves as a collection of an equal number of oppositely charged poles, Le., as
dipolar matter. Polarized dielectrics and ferroelectric materials provide the electric
dipolar analog. The electrically polarized media are characterized by the polariz-
ation vector P and a vector D, the displacement field, defined as

D=eoE+P (2.2)

The time derivative of D is termed the displacement current. Here eo is a constant


called the permittivity of free space [eo = 8.854 x 10-12 = 10-9 / (36'7T) farad m-1].
Quantity (eoIL0)-1/2 has the dimensions of a velocity; it is known from a solution
to Maxwell's equations that this quantity equals the propagation speed c of light
or any other electromagnetic wave in a vacuum (c = 2.9979 X 108 m S-1).
To spare the reader the inconvenience of searching through the general list
of notation, Table 13.1 provides a summary ofthe purely electromagnetic variables
and units used in this work.

13.3. ELECfROMAGNETIC THEORY

Maxwell had been reading Faraday's "Experimental


Researches "; and, gifted as he was with a physical imagination
akin to Faraday's, he had been profoundly impressed by the
theory of lines of force. At the same time, he was a trained
mathematician; and the distinguishing feature of almost all his
researches was the union of the imagination and the analytical
faculties to produce results partaking of both natures.
Sir Edmund Whitaker, A History of the Theories of Aether
and Electricity (1910)
372 TABLE 13.1
Summary of Electromagnetic Field Variables and Units
Chapter Thirteen
Symbol Name SI unit SI base units

So Permittivity of free space farad/ meter m- 3 kg- I s' A2


E Electric field volt/meter mkgs- 3 A-I
0 Electric displacement coulomb/(meter)2 m- 2 sA
P Polarization coulomb/(meter)2 m- 2 sA
Jr Free current density ampere/(meter)2 m- 2 A
Pr Free charge density coulomb/(meter)3 m- 3 sA
JLo Permeability of free space henry/meter mkgs- 2 A-2
H Magnetic field ampere/meter m-IA
B Magnetic induction tesla kg S-2 A-I
M Magnetization ampere/meter m-IA
D=soE+P 1 tesla (SI) is 10' gauss (cgs)
B= JLo(H+M) H of one oersted (cgs) is B of
So = 8.854 X 10- 12 farad/meter one gauss (cgs) in vacuum
JLo = 411" X 10-7 henry/meter

13.3.1. Maxwell's Equations

The development of electromagnetic theory followed two separate paths


until the nineteenth century. One of these was the study of electric charges and
their fields, and the other concerned electric currents and the magnetic fields they
produce. This was the state of affairs until Faraday showed that a time-varying
magnetic field can generate an electric field and Maxwell, by introducing the
displacement current, showed that a time-varying electric field produces a mag-
netic field. The mathematical relationships governing electromagnetic phenomena
are the celebrated Maxwell's equations. In differential form and written for an
observer fixed in the laboratory frame they appear as follows, in the so-called
macroscopic form suitable for the calculation of fields both inside and outside
of matter:

Faraday's law

v x E = -aD/at (3.1.1)

Ampere's law with Maxwell's "correction"

VxH=Jr+aD/at (3.1.2)

Gauss's law (I)

VD=Pr (3.1.3)

Gauss's law (II)

VD=O (3.1.4)
Faraday's law relates the circuital voltage that appears when the flux linkages 373
vary with time, as in an electrical generator, or for that matter in an electromagnetic
Thermodynamics of
wave traveling through space; there need be no electrical current flow nor Electromagnetism
conductor present. Ampere's law relates the magnetic field that curls round a
current flux, corrected for unsteady values of electric field; wound magnetic field
sources rely on the former while the latter is important in wave propagation.
Gauss's law (I) tallies the field lines emanating from a distribution of charge,
and Gauss's law (II) reflects the circumstance that isolated magnetic poles are
unknown; a line of magnetic induction closes on itself.
Constitutive relations connect the fields D and H to the fields E and B. In
matter that is moving relative to the laboratory frame the field variables must be
evaluated in the frame of the moving matter. For quasi-stationary processes no
distinction need be made and one may write

D = D(E,p, T) (3.1.5)

and
H = H(B,p, T) (3.1.6)

where it is also assumed that the magnetic and dielectric parameters are uncoupled
from each other. The defining equations, discussed previously, are

D = EoE+P (3.1.7)

and
B = J.'o(H+M) (3.1.8)

P and J.'oM have physical significance as electric and magnetic dipolar density
in substances.
In equation (3.1.2) J r is the free current density. The term "free" distinguishes
such charge from the "bound charge" associated with the electric dipolar matter.
If there is more than one type of charge carrier, the net charge density Pr is equal
to the algebraic sum of all the charge densities, while the net current density'Jr
equals the vector sum of the current densities due to each carrier. Then, because
net charge is conserved, an equation of conservation can be written a priori as

(3.1.9)

It should be noted that in taking the divergence of Ampere's law in equation


(3.1.2) the term containing H vanishes because the divergence of the curl of any
vector is zero. Then using Gauss's law (I) it is seen that the equation of net
charge conservation (3.1.9) is again obtained. It was in fact this circumstance
that lead Maxwell to add the displacement current "correction" aD/at tu
Ampere's law.
Additionally, boundary conditions are imposed on fields at interfaces
between media having differing properties of M and P; in a magnetic system
these relationships assert the continuity of normal B and tangential H:

[B 0] = 0 (3.1.10)
374 and
Chapter Thirteen [H x nJ =0 (3.1.11)

where brackets indicate difference across the interface.


Finally, in the treatment of dissipational processes of moving, polarized
matter, a transformation is required to related field quantities in the moving frame
of the material to field quantities evaluated in the laboratory reference frame;
these requisite couplings are introduced into the description where needed.

13.3.2. Correspondence between Magnetic and Electric Systems

In the absence of current and free charge (Jr = 0 and Pr = 0) Maxwell's


equations reduce to
VxE=-aD/at (3.2.1)

VxH= aD/at (3.2.2)

VD=O (3.2.3)

VD=O (3.2.4)

D = EoE+P (3.2.5)

B = lLo(H+ M) (3.2.6)

Further, when the fields change slowly, as in a reversible process carried out at
an infinitesimal rate,
aD/at = 0 (3.2.7)

and
aD/at = 0 (3.2.8)

In this case the equations break into decoupled sets of magnetostatic and
electrostatic relationships having analogous forms.
Magnetostatic Electrostatic

VxH = 0 (3.2.9) VxE=O (3.2.12)


V B = 0 (3.2.10) VD=O (3.2.13)
B = lLo(H + M) (3.2.1I) D = EoE+P (3.2.14)

The following substitutions then transform a magnetostatic relationship into an


electrostatic one, and vice versa:
(3.2.15)

(3.2.16)
(3.2.17) 375
Thermodynamics of
(3.2.18)
Electromagnetism

This correspondence between such magnetic and electric systems will spare the
necessity of separately discussing both topics in every possible instance.
As mentioned previously, electric field systems differ from magnetic systems
in that free charge exists in the former with no phenomenological analog in the
latter. To preserve the analogy space-charge effects must be eliminated in the
electric field systems. This can be accomplished if alternating electric field is
applied at a frequency w greater than the reciprocal dielectric relaxation time
CT / Ii of the polarizable substance (CT is the electrical conductivity).

13.4. ELECfROMAGNETIC WORK WITH CONSTANT MASS DENSITY

In speaking of the Energy of the field 1 wish to be understood


literally. All energy is the same as mechanical energy, whether
it exists in the form of motion or in that of elasticity, or in any
other form. The energy in electromagnetic phenomena is
mechanical energy. The only question is, Where does it reside?
On our theory it resides in the electromagnetic field, in the
space surrounding the electrified and magnetic bodies, as well
as in those bodies themselves ....
James Clerk Maxwell, "On Physical Lines of Force,"
Philosophical Magazine, 1861/62

The modern view coincides with Maxwell's-the energy resides in the field.
As convincing evidence, the electromagnetic energy constituting a light ray
continues to propagate after the source is extinguished.
This section develops expressions for the electromagnetic energy in simple
electric and magnetic field systems. The systems are chosen deliberately simple
to aid physical understanding. However, the results have a more general
significance as will be established.

13.4.1. One-Dimensional Dielectric

Let us consider parallel-plate electrodes subject to potential difference e


enclosing a dielectric medium with permittivity e (see Figure 13.1). Dielectric
is the name for materials that are electrically insulating. Another name for
permittivity is dielectric constant. In isotropic media

D=eE (4.1.1)

which is the definition of e. In this simple geometry lEI = cp / I. Despite its name,
e is not usually a constant but depends on temperature, mass density, frequency,
and magnitude of the field. For so-called linear dielectrics Ii is independent of
E. Typical values of relative permittivity lir are listed in Table 13.2 for various
376
Chapter Thirteen
! Gauss, Surface / +q +

FIGURE 13.1. Two thin parallelplate electrodes of area A at potential difference e. The
dashed line represents a Gaussian surface to derive the relationship between displacement
D and surface charge density q (the free charge).

common substances; relative permittivity is defined as

e r ,= el eo (4.1.2)

The values cited in Table 13.2 pertain to a steady value of the applied electric
field E. Materials having a large relative permittivity are usually composed of
highly polar molecules. When the potential difference across the electrodes is
increased from e to e + de, a differential of electrical work dW. is performed on
the slab,
dW. = edq (4.1.3)

where Iql is the charge on a plate. Assuming a negligible effect of fringe fields
so that the charge is equally distributed over the surface area A of a plate, charge
is given by
q = uA (4.1.4)

where u is charge density, coulomb/m 2 Thus, Gauss's law (I) can be integrated
over a volume with the shape of a small pillbox straddling the interface between

TABLE 13.2
The Relative Permittivity for Various Substances at Room Temperature
and Zero Frequency

Dielectric strength
Material er (kV'm- 1 )

Vacuum 1.00000 00
Air 1.00054 800
Teflon (polytetrafluorethylene) 2.1 60,000
Transformer oil 4.5 12,000
Pyrex glass 4.5 13,000
Ruby mica 5.4 160,000
Pure water 80
A ferroelectric titanate >2100 2000-12,000
the positively charged electrode and the dielectric substance, 377

LV, L
Thermodynamics of
Electromagnetism
D dV = Pr dV = uA (4.1.5)

Application of Gauss's divergence theorem to the left-hand side of equation


(4.1.5) gives

r V. DdV = Ja9t
J~
rD' n dS = DA (4.1.6)

where n is a unit outward-facing normal to the pillbox surface. The right-hand


side of equation (4.1.6) is simply DA, because the field is uniform in the dielectric
and absent within the conductor. Combining these results the boundary conditivn
on the displacement is
D=u (4.1.7)

Thus, the displacement evaluated at the interface equals the surface density of
electric charge and q = DA. Because E = lEI = ell, equation (4.1.3) now becomes

dWe = EI d(DA) (4.1.8)

or, because A is constant,

1
-dW
V e =EdD (4.1.9)

where V = Al is the volume of the dielectric slab. Hence the work done on the
slab by the electrical circuit has been expressed as a volumetric density in terms
of electric field quantities evaluated within the dielectric substance. It should be
noted that the dielectric substance was tacitly assumed constant in volume in
this derivation, a restriction that is removed in Section 13.5.
Thermodynamics of a closed system at its roots is concerned with heat flow
or its absence. Because the conditions of heat flow during the charging process
have not yet been specified, the right-hand side of equation (4.1.9) lacks precise
meaning. To give the result a definite meaning the first law of thermodynamics
is applied to a unit volume of the dielectric slab,

(4.1.10)

where dQv is the differential of heat added to unit volume of the slab, dWy =
d (Wei V) = (dWe)1 V is the differential electric work done on unit volume of the
slab material, and U y is the internal energy of a unit volume of the slab. For a
reversible process dQy = TdSy, so from equations (4.1.10) and (4.1.9) we obtain

E dD = dUy - T dS y (4.1.11)
378 Supposing the process to be accomplished isothermally through efficient heat
exchange with the surroundings permits writing dUv - T dSv = d( U v - TSJ, so
Chapter Thirteen
that

EdD = dAv (T = constant) (4.1.12)

where Av is the per unit volume Helmholtz free energy,

(4.1.13)

As an alternative, if the capacitor is charged reversibly and adiabatically (dSv = 0),


equations (4.1.11) and (4.1.9) give

EdD = dUv (Sv = constant) (4.1.14)

E dD has path-dependent integrated values because E is dependent on


temperature T as well as the displacement D.

13.4.2. Simple Magnetic Case

Referring to Figure 13.2 we consider a cylindrical sample of a homogeneous


magnetizable material wound uniformly with n turns of an insulated conductive
wire. The length of the sample is assumed large compared to its diameter, so
that demagnetizing effects due to poles on the ends of the sample produce
negligible corrections to the applied field. The terminals of the wire are connected
to a source of emf e. The voltage at the terminals is given by

d4> dB
e=n-=nA- (4.2.1)
dt dt

the first part of which can be deduced from integration of Faraday's law (3.1.1)
though not so simply if rigor is desired. The flux in the system is 4> = BA, where
A is the constant cross-sectional area of the sample and B is the uniform induction
field. Ampere's law (3.1.2) in the absence of displacement current integrates

Magnetizable Matter
Cross Section A

"- ............ ... ....

---- e---
FIGURE 13.2. Sketch to develop expression for electrical work performed in magnetizing
a material.
readily as shown in standard texts to yield the relationship between magnetic 379
field magnitude H and current i,
Thermodynamics of
Electromagnetism
H = nill (4.2.2)

The electrical work dW. done on the system in time dt is

dW. = eidt (4.2.3)

and substituting for e and i from the above and cancelling common factors gives

dWe= AIHdB (4.204)

The volume V = Al is constant by assumption and thus the differential of work


per unit volume is dWv = d( Wei V) = (dWe)1 V. The work done on the material
per unit volume is thus given by

dWv = HdB (4.2.5)

Similar to the electrical case, the work again has been expressed purely in terms
of field variables. Incorporating this work expression into the combined first and
second laws applied to a unit volume of the magnetizable material proceeds in
the analogous manner as for the electrical case and gives

H dB = dUv - TdSv (general) (4.2.6)

HdB = dAy where Av = Uv - TSv (T = constant) (4.2.7)

HdB = dUv (Sv = constant) (4.2.8)

EXAMPLE 13.4.1. Density of Energy Storage in Electric and Magnetic Fields


The expressions developed above permit the numerical computation of the energy
stored in fields.
For the electric field, from equation (4.1.14) dUv = E dD and D = EoE + P = eE
or E = Eo(1 + X.) where X. = P / EoE is the electric susceptibility. Assuming isothermal
conditions and constant Xe, e is constant and the integrated expression for the free
energy density Av = Av . is

E2 D2
A v.. = E2 - =2E- (E.4.1.1)

Similarly, for isothermal magnetization, from equation (4.2.7) dAy = H dB and B =


10(H + M) = IH or 1 = 10(1 + Xm) where Xm = M / H is the magnetic susceptibility. With
constant Xm the integrated expression for magnetic field energy density Av = Av.m is given
by

(EA.I.2)
380 Table 13.3 lists computed values of field energy density for various systems. All the energy
stored in fields is potentially available to perform work. Hydrocarbon fuel such as gasoline
Chapter Thirteen
has a heat of combustion of about 3.2 x 10 10 J/m 3 Assuming a 30% thermal efficiency
for conversion of the energy to work in an internal combustion engine the available energy
density is 9.6 x 109 J/m 3 Thus, the calculated energy storage in superconductors
approaches that of chemical fueL The difficulty of containing the energy is considered in
a later example.

13.4.3. General Case from Maxwell's Equations

When a magnetic or an electric field is established in a region of space, an


expenditure of energy is required in addition to any energy consumed in irrevers-
ible processes such as ohmic heating, and this energy can be related to just the
spatial distribution of the electromagnetic field vectors. While the field expressions
for the energy were derived in reference to systems of simple geometries in the
preceding section, the purpose of the following is to illustrate the general nature
of these relationships.
This more general derivation is based on use of the full set of Maxwell's
equations. It will be assumed that fields are established in an arbitrary system
starting from a field-free state. The fields are assumed to be generated by a spatial
distribution of electric current J r flowing in response to a spatial distribution of
electric field E over a region of volume V. Thus, in any differential volume of
the system the differential work performed by the external sources in time interval
dt is given by

dWe = E . JrdVdt (4.3.1)

The term E J r can be manipulated into a form in which only electromagnetic


field variables appear. Thus

E J r = E (V x H - aD/at) - H (V x E + aB/at) (4.3.2)

Quantity J r was eliminated using Ampere's law (3.1.2), in the first term on the
right-hand side, and the second term is identically zero on account of Faraday's
law (3.1.1). By introducing the vector identity

V . (E x H) = E . (V x H) - H . (V x E) (4.3.3)

equation (4.3.2) can be rewritten as an integral expression giving the work done
on the system by the external sources in establishing the field:

We=f f'(E.aD+H.aB)dtdV+f f'V'(EXH)dtdV (4.3.4)


9'1 0 at at 9'1 0
TABLE 13.3
Field Energy Storage

Field System e/ EO Eb a (kYmm-') 1'-/1'-0 B(T) Av (Jm-3 )

Electric Air gap capacitor 1.0 0.8 2.83


Electric Ruby mica capacitor 5.4 160 2.10 X 10'
Magnetic Air gap of iron yoke magnet 2 1.59 X 10
Magnetic Present superconductive magnet 6 1.43 X 10'
Magnetic Speculated future superconductor 100b 3.98 X 109

tl Eb is the breakdown electric field (dielectric strength).


h As the critical transition temperature Tc of a superconductor increases, the transition field at lower temperatures increases; assuming the usual relationship
for a type II superconductor it is computed that a steady magnetic field as large as 300 tesla (3 x 106 gauss) could be generated at liquid helium temperature
with a superconductor for which Tc exceeds room temperature.

hl~
......
"-
a cI'>.~
~'S
""'~" '~~." w
~. f'J 00
~..sa, ......
382 The volume integral in the last term is transformed to a surface integral using
the divergence theorem:
Chapter Thirteen

r V. (E x H) dV =!J.~ (E x H) . n dS
J~
(4.3.5)

Here E x H, called the Poynting vector, having units of watts per square meter,
is interpreted as the flux density of electromagnetic energy through a spatial or
material surface; only its integral over a closed surface has measurable sig-
nificance. On the assumption that Hand E go to zero on the system boundaries,
expression (4.3.4) becomes

W. =

i f, t(
~ 0
aD
E-+H-
at dtdV aB)
at (4.3.6)

The volume integrations are carried out over stationary volume elements, so
expression (4.3.6) can equivalently be written as

(4.3.7)

W. represents work done on the system by the external sources.


Based on the form given in equation (4.3.7), E dD and H dB are considered
to represent differential work densities on the unit volume basis. When Hand
B, and E and D are collinear, which is the case for isotropic material in a
time-steady field, these scalar products are expressible in terms of the field
magnitudes as H dB and E dD, respectively. Thus, the expression for differential
electromagnetic work per unit volume done on a system, denoted dW is y ,

dWy = H dB + E dD (4.3.8)

in agreement with the results obtained in the uniform field systems and given as
equations (4.1.9) and (4.2.5).
Expression (4.3.8) requires further transformation to treat the thermo-
dynamics of mass having variable volume, a step accomplished in the next section.

13.5. ELECI'ROMAGNETIC WORK WITH VARIABLE MASS DENSITY

In order to apply thermodynamics to magnetic systems we have


merely to extend our previous formulae by including extra
terms for the magnetic work. In principle, the procedure is
straighiforward and should cause no difficulty. There is however,
a serious incidental difficulty, namely that of finding the correct
general expression for magnetic work.
E. A. Guggenheim in Thermodynamics,
fifth revised edition (1967)
Guggenheim masterfully circumvented the problem by analyzing selected 383
aspects of a particular, variable density system; the technique falls short of
Thermodynamics of
providing a general expression that can be used in subsequent analyses. Other Electromagnetism
authors appear not to notice there is a problem.
In this section, we obtain electromagnetic work for a system with variable
mass density by employing a straightforward transformation (in retrospect) of
the known expression for electromagnetic volumetric work density given in
Section 13.4.
Because at this stage we know electromagnetic work only on the per unit
volume basis given by equation (4.3.8), this treatment begins with a statement of
the first law written on the same basis, Le., per unit volume. For simplicity, terms
of magnetic polarization alone are included because the terms for electric polariz-
ation are analogous and will be written by inspection at the end. The first-law
statement is
dQv + H dB + IL dp = dUv (5.1)

where Qv is heat added per unit volume, p is the mass per unit volume, Uv is
internal energy per unit volume, and H dB is the expression for magnetic work
performed per unit volume, derived previously. It is noted that equation (5.1)
contains no term for work of expansion and this might at first appear an oversight.
However, for a system defined as the variable content of a constant volume in
space, the boundaries do not move and thus it is strictly true that no work of
expansion is performed. Nonetheless, material is free to move across the boundary
should expansion or contraction occur, Le., the mass content of the system is
variable. The system is an open system of variable mass with the thermodynamic
consequence embodied in the term IL dp wherein IL is the chemical potential per
unit mass of the substance present in the system. The reader can note that in the
analogous first-law statement written in Section 13.4 as equation (4.2.6), density
was assumed constant, so there dp = 0 and the term IL dp did not appear~
To begin transforming equation (5.1) we assert the compatability of the first
law written for a homogeneous substance in any of the following three forms:

dQy + H dB + IL dp = dUy (Unit-volume open-system) (5.1)

dQ + dWm - pd(l/p) = dU (Unit-mass closed-system) (5.2)

dQ,+ dW, - pdV, + ILdm, = dU, (Variable volume and


mass open-system) (5.3)

Subscript t refers to total mass and subscript v to unit volume. Total mass is
denoted m,. The quantity dWm denotes differential magnetic work performed on
a unit mass while dW. denotes the differential magnetic work performed on the
total mass. The unprimed quantities Q and U pertain to the unit mass basis. In
equations (5.2) and (5.3) it is assumed that a scalar pressure p exists whose
product with volume change yields mechanical work of expansion, so that a
different treatment would generally be needed for solids. The pressure p must
be assumed field-dependent, reducing to ordinary pressure only in the absence
of the electromagnetic field.
384 Having defined dWm and dW, the next objective is to determine definite
expressions for both. In a reversible process dQ = T dS, dQv = T dSv , and dQ. =
Chapter Thirteen
T dS. where S, S., and S. are the corresponding entropy variables. Using these
variables equations (5.1)-(5.3) can be written as the respective Gibbs equations,

TdS,,+ HdB + JLdp = dUv (5.4)

TdS + dWm - pd(l/p) = dU (5.5)

TdS. + dW, - p dY, + JL dm. = dU. (5.6)

By definition the total quantities are related to the volumetric quantities as follows:

(5.7)

(5.8)

m. = Y,p, dm. = Y, dp + pdY, (5.9)

Substituting these expressions into equation (5.6) and rearranging gives

Y,( T dSv + JL dp - dU.) + dY,( TSv - p + JLP - Uv) + dW, = 0 (5.10)

Next we attempt to express the total work in the form

dW, = XdY, (5.11)

and determine whether this form is compatible within the posited framework. It is
desired that Y, be an extensive variable as are the other differential variables in
equation (5.6) and, correspondingly, that its coefficient X be intensive.
Accordingly, Y, is sought in the form

Y,=Y,y, dY. = Y, dY + Y dY, (5.12)

Substitution of the assumed form of Y. in equation (5.10) then yields

Y,(TdSv + JL dp + X dY - dUv)
+ dY,(TSv - p + JLP + XY - Uv) = 0 (5.13)

for which it will next be shown that the coefficient of dY, vanishes.
To proceed, dW, from relation (5.11) is substituted into equation (5.6) giving

T dS. + X dY, - p dY, + JL dm. = dU. (5.14)

Because this equation is the sum of products of intensive variables with extensive
variables (still a hypothesis for X dY,) the equation can be integrated along a
path in which mass is added to build up the system from nothing while holding
constant the intensive variables T, X, p, and JL. The result of the integration is
an Euler equation in the form 385
Thermodynamics of
TS. + Xy. - Pv. + p.m. = U. (!1.l5)
Electromagnetism

Neither equation (5.4) nor (5.5) can be integrated in this manner because in them
the volume and the mass, respectively, are fixed at constant values. Substituting
into relation (5.15) from equations (5.7)-(5.9) and (5.12) to eliminate total
variables in favor of the volumetric definitions gives

(5.16)

Comparison of equation (5.16) with (5.13) now confirms that the final term of
relation (5.13) disappears, and what remains can be arranged to read

(5.17)

Comparison of equations (5.17) and (5.4) permits the following important


identification to be made:
X=H, Y=B (5.18)

which supports the hypothesis (5.11). From the definition of y. in equation (5.12),
y. = BV.. Thus, the total work dW. can be expressed as

dW. = X dy' = H d(BV.) (5.19)

The combined law for the variable volume and mass system from equation (5.6)
therefore becomes
TdS. + H d(BV.) - p dV. + p. dm. = dU. (5.10)

We are now in a position to complete the determination of the expression


for electromagnetic work in (5.2). The total quantities are related to the unit mass
quantities by the following definitions wherein m. is the total mass in the
homogeneous system:

S. = Sm" dS. = Sdm.+ m.dS (!l.U)

1
v.=-m"
p
(!l.ll)

U.= Um dU. = Udm.+ m.dU (5.l3)

Substituting the differential relationships into equation (5.20) gives upon


rearrangement

m.[TdS+Hd(;) -Pd(;) -dU]

+ dm.[ TS + ~B - ; + p. - U] = 0 (5.24)
386 Equations (5.18) and (5.12) enable the Euler relationship (5.15) to be written as
Chapter Thirteen
TS,+ HBy' - py.+ ,.,.m, - U, = 0 (Sol5)

or, in terms of mass variables, using system (5.21)-(5.23),

HB p .
TS+---+,.,.- U=O (5.26)
p p

Comparison of equations (5.26) and (5.24) demonstrates that the last term in
,equation (5.24) vanishes and what remains can be expressed as

(5.27)

Comparison of equations (5.27) and (5.5) now identifies the work term in relation
(5.2) as

(5.28)

This is the central result of our treatment giving dWm as the appropriate work
expression to utilize in diverse problems, such as: compressible flow dynamics,
magnetostriction and other problems with density change; and magnetochemistry
wherein density change and molar transformation occurs. For problems in which
density is constant equation (5.28) reduces to (H / p) dB as previously obtained.
Employing the analogy for an electrically polarizable system developed
previously permits writing the result corresponding to equation (5.28) as

dW. = Ed(D/p) (5.29)

One small but critical point remains to be commented upon. In equation


(5.11) it was assumed that 1'; is extensive which permitted integration of equation
(5.14) to give the Euler form (5.15). Then it was found that 1'; is given by the
expression 1'; = By'. Accordingly, we must now examine By' to determine its
status as an extensive variable. Now in a homogeneous region B is spatially
uniform, and it is certainly true that By' is extensive as y. increases in proportion
to the total amount of mass. Thus, the variable 1'; = By' is a legitimate extensive
variable. Concomitantly, it is necessary to retain B under the differential in
equation (5.20) as B can be time dependent.

* More generally, from expression (4.3.7) the magnetic work term in equation (5.1) is H dB where
Hand B need not be collinear. Retracing the derivation then yields dWm = Hd(B/p) in place of
equation (5.28). By the same token the generalization of equation (5.29) gives dW. = Ed(D/p).
These more general expressions are useful in formulating relationships for the ftow of ftuids having
internal structure, such as magnetic colloidal dispersions (see Section 13.9.1},
As an overall check on the work we should find that equation (5.20), which 387
is the most general statement of the combined laws, reduces to the result for
Thermodynamics of
constant mass given by relation (5.27), and to the result for constant volume Electromagnetism
given by equation (5.1). The reductions are readily confirmed in both cases.
Having found no contraditions it may be concluded that our attempt
to determine expressions for electromagnetic work in the variable density,
polarizable systems has been satisfactorily achieved.

Gibbs Equations
Combining the results of equations (5.28) and (5.29) for a constant composi-
tion system of fixed mass containing matter that is magnetically and electrically
polarizable yields a Gibbs equation in the form

(5.30)

Equation (5.20) derived previously expresses the combined first and second
laws for a variable-mass system of one component. That equation is readily
generalized to apply to a system of variable composition by replacing the term
II- dm. by a sum of terms of the form II-i dm" where subscript i denotes a species
of an s-component mixture. Including work of electric polarization along with
the work of magnetic polarization, the broadened relationship is written as
follows:

T dS. - p dV. + H d (BY,) + Ed (DY,) + I


. II-i dmi = dU. (5.31)
i

An equation with chemical potentials of this form is sometimes referred to as


the fundamental equation, to distinguish it from the Gibbs equation, but we will
not make this distinction, particularly in view of the circumstance that Gibbs
originated the notation of chemical potential.

EXAMPLE 13.5.1. Work of Magnetizing an Expansible Cylinder


This example illustrates the meaning of the work expression (5.28) in relation to a
system of simple geometry.
Figure 13.3 illustrates the system, which differs from that of Figure 13.2 in that a
space is present between the windings of the field source and the surface of the cylindrical
sample. Considering the conductor to be free of electrical resistance, the electrical work
done by the external circuit in time dt is

dW. = eidt (E.5.1.1)

where from Faraday's law

dtb
e=n- (E.5.1.2)
dt
388 For this geometry the magnetic flux is given by
Chapter Thirteen
cf> = BA+ B'A' (E.5.1.3)

where B is the induction in the inner region, i.e., in the space occupied by the material,
and B' is the induction in the space between the material and the coil. Quantities A and
A' are the inner and outer cross-sectional areas, respectively. In accord with Ampere's
law the magnetic field H is the same in both regions, i.e., H = H', and

H = nill (E.5.1.4)

Thus, the work in equation (E.5.1.1) can be expressed as

dW. = Hld(BA + B'A') (E.5.1.5)

The mass ... of the cylinder is constant,

.,. = Alp (E.5.1.6)

Eliminating area A and expressing the result as the work done per unit mass gives

1
-dW.
.,.
= Hd - (B) +-d(BoAo)
p".
HI (E.5.1.7)

The first term on the right-hand side is the work done on the inne r region per unit mass,
in agreement with relation (5.28). The second term is the work done in increasing the field
in the outer region, normalized to unit mass of the inner region. Any reversible work of
expansion performed by the cylinder equals the work done on surrounding nonmagnetic
material, and the sum of the mechanical work done by the whole system vanishes.
Returning to equation (E.5.1.5) and substituting B' = /LoH' and B = /Lo(H + M),
where M is magnetization of the material in the inner region, gives

dW. = HI d[foLoMA + foLoH(A + A')]


= HI d(/LoMA) + V, d(foLoH2/2) (E.5.1.8)

Final Boundary

FIGURE 13.3. Sketch to derive


magnetic work in a system contain-
ing an expansible cylindrical
sample.
where V. = I(A + A') is the constant total volume. Eliminating A with expression (E.5.1.6) 389
and rearranging allows the work per unit mass to be expressed as
Thermodynamics of
Electromagnetism
(E.5.1.9)

In this representation the system is divided into two subsystems, the matter comprising
the cylinder less the space it occupies, and the total space of the inner and outer regions
located within the coil. The first term on the right-hand side is the density of work done
on the mass; the second term is the work done on the total space, normalized to the inner
mass.

PROBLEM 13.5.1. Practice in Separating a Mass From the Space It Occupies


Choosing only the mass", of the cylindrical sample as the system write the first law using for
work done on the system the sum of mechanical work -p dV and magnetic work /LoH d(M/ p) found
above. V is the volume occupied by the mass. Reconcile the associated Gibbs equation with the
result for a closed system listed in the second row of Table 6.1.
PROBLEM 13.5.2. Work of Polarizing a Dielectric Slab of Variable Density
For a system containing a slab of expansible dielectric material located between spaced-apart
flat electrodes having a fixed separation distance, derive from equation (E.5.1.1) the expression for
electrical work corresponding to equation (E.5.1.9), i.e.,

(P.5.2.1)

13.6. THE GIBBS EQUATION AND THERMODYNAMIC RELATIONS

In the following, the Gibbs equation derived previously in the form of


equation (5.30) is used to relate pressure and entropy in the presence of elec-
tromagnetic fields to their expressions in the absence of a field. In addition, the
Gibbs equation is recast into forms useful in the subsequent development of
hydrodynamic relationships and relationships of physicochemical interest.
To begin we rewrite the Gibbs equation,

TdS + H d(B/ p) + E d(D/ p) - pd(1/p) = dU (6.1)

The Helmholtz free energy per unit mass A is defined conventionally as

A=U-TS (6.2)

An additional thermodynamic function A can be defined incorporating the


electromagnetic field variables,

A=A-BH/p-DE/p (6.3)

The motivation for this definition is that A possesses utility. As usual in thermody-
namics we are free to define any new function. The electromagnetic terms in A
are chosen to yield simple relationships when A is differentiated and combined
390 with the Gibbs equation. We shall see subsequently that A serves as a potential
function, namely, various desired thermodynamic parameters result from partial
Chapter Thirteen
differentiation of A.
Equations (6.3) and (6.2) yield

-
dA = dU - T dS - S dT - H d (B)p - pBdH - Ed (D) DdE
P -P (6.4)

by eliminating dU from equations (6.4) and (6.1) and expanding d (1/ p ) we obtain

- P B D
dA = -SdT+2dp --dH --dE (6.5)
p p p

from which it can be seen that A is a function of the independent variables of


temperature, density, magnetic field, and the electric field:

A = A(T,p, H,E) (6.6)

Given that A is a point function we have

dA = (OA) dT+ (PA) dp + (OA) dH + (OA) dE (6.7)


oT p,H,E op T,H,E oH T,p,E oE T,p,H

Comparison of equations (6.7) and (6.5) leads to the following prescriptions for
obtaining entropy S, pressure p, magnetic induction B, and the electric displace-
ment D from function A:

p=p - 2(OA) B= _p(OA) ,


op T,H,E' oH T,p,E

D--p - OA)
(oE (6.8)
T,p,H

It will be assumed that the polarizable fluid is free of the anisotropy that results
from micropolar effects, such as the body couple exerted on molecules or particles
in suspension (see Section 13.9.1). Hence we can assume that M is collinear with
H, and E with D. Thus, we can write

and P=X.E (6.9)

where Xm and Xc denote the electric and magnetic susceptibility, respectively.


The susceptibilities in general are dependent on density, temperature, and field.
Hence
Xm = Xm(P, T, H) and X. = X.(p, T, E) (6.10)
Function A may now be found by integrating equation (6.5) along the following 391
path: initially, at zero values of the fields Hand E integrate over temperature
Thermodynamics of
T at constant density p, then over density p at constant temperature T. The effect Electromagnetism
of this procedure is to integrate the first two terms on the right-hand side of
equation (6.5) to obtain A, the Helmholtz function in the absence of a field.
Subsequently, we integrate over the magnetic field at constant temperature,
density, and electric field intensity. Finally, integration is carried out over the
electric field in an analogous manner. The final result of integration at this stage
is

A = Ao--IfH BdH - -If" DdE (6.11)


PoP 0

where Ao = A(p, T, H = 0).


Introduction of the defining relations previously stated as equations (3.1.7)
and (3.1.8) gives

H fH H2 fH
f B dH = P-o( H + M) dH = P-o - + P-oM dH (6.12)
o 0 2 0

and

f"
o DdE
r"
= Jo (60E + P) dE
E2
= 60 2+ r"
Jo PdE (6.13)

Thus, the equation for A can be expressed in the form

(6.14)

Expressions (6.8) with this expression for A yield the following relationships for
pressure and entropy:

P-OH2 60E2)
P=Po+ ( - - + - -
2 2

- u)
+P-o f H(aM dH+ f"(apu)
- dE (6.15)
o au T,H.E 0 au T.H.E
and

S=So+-1 fH(a)
- dH+-1 fE(ap)
- dE (6.16)
p 0 aT p.H.E p 0 aT p,H,E

Here Po = p2(aAo/ap)T is pressure at density p and temperature T in the absence


of fields; similarly So = -(aAo/aT)p is entropy in the absence of fields, and
U = P-) is the specific volume. A pressure-like variable P* is defined as

P* = Po + P-o J, H(aMU)
-- dH + J,E(aPv)
- dE (6.17)
o au T,H,E 0 au T,H,E
392 Subtracting this relationship from equation (6.15) gives a simple expression
relating p* to pressure p:
Chapter Thirteen

(6.18)

Recasting the Gibbs Equation


Substituting expression (6.16) into the Gibbs equation (6.1) and expanding
the field-containing terms gives, with slight rearrangement, the following
representation of the Gibbs equation:

fJ-oH2 60E2 ) dp
pdU=pTdS+ ( p*-------fJ-oMH-PE -+HdB+EdD (6.19)
2 2 p

For an incompressible substance (dp = 0) equation (6.19) reduces to

pdU = pTdS + HdB + EdD (6.20)

which also is obtainable directly from equation (6.1). The form of the Gibbs
equation (6.19) is found to be convenient later in deriving the surface stress
tensor, from which one obtains the equation of motion for a polarizable
fluid.
Another form of the Gibbs equation is useful in treating problems of phase
equilibrium and chemical change in which the thermodynamic system is chosen
as the matter exclusive of the space that the matter occupies.
It is assumed for simplicity that a magnetic field is present in the absence
of an electric field. The objective is to transform the open system form of the
Gibbs equation (5.31) applicable to a multicomponent system. By substituting
for pressure p using expression (6.18) and employing the defining equation
B = fJ-o(H + M), equation (5.31) can be put into the desired form representing
the Gibbs relation for an open system excluding space occupied by the mass.
This relationship is

TdS, + fJ-oH dI, - p* dy' + L fJ-. dm. = dUt (6.21)

where ut is the excess of internal energy over the background field energy
associated with the volume y. of space,

(6.22)

and I, = My' is the total magnetic moment.


A form applicable to a system of unit mass and variable composition is
useful in treating problems with diffusing species. This form is obtained by
substituting definitions (5.21) for S" V" and U, into relationship (5.31) and
eliminating the Euler equation corresponding to (5.31). The result is 393
Thermodynamics of
TdS - p d(!) + H d(!!') + E d(!!') + .f f.J.i dxi = dU
Electromagnetism
(6.23)
p p P ,-I

where Xi = md m, is mass fraction of species i with m, = L mi'


Table 13.4 classifies and lists various forms of the combined first and second
laws that are useful in this work corresponding to various choices of the thermo-
dynamic system.

Chemical Potential in Presence of Fields


Application of relationship (6.21) requires expressions for the chemical
potential functions appearing in it. For the purpose of obtaining such expressions
it is convenient to define a modified Gibbs function ot for the total system:

ot = ut - TS, + p*V; - f.J.o H1, (6.24)

Differentiating this expression and eliminating ut with the aid of equation (6.21)
yields

i
dOt = -S, dT + V; dp* - J.toI, dH + L J.ti dmi (6.25)
i=l

The form of equation (6.25) indicates that ot is a function of T, p*, H, and mio
i.e.,

ot = ot(T, p*, H, mJ (6.26)

Thus, by expanding ot we obtain

dO*, = (ao t) dT + (aot)


aT. p ,H,mj
aP* T,H.m; dP*

+ ( -aot) dH+ (aot)


- dm (6.27)
aH T,p.,mt ani T,p H.m) I

Comparison of like coefficients in expressions (6.27) and (6.25) establishes the


following relationships:

( ao t)
aT = -S" ( aot)
ap* = V;,

( aot) = f.J.i (6.28)


ami
W
Q 10
~
~
~
:.l
I'

TABLE 13.4
Combined First and Second Laws for Various Electromagnetic Systems

System Type Gibbs equation Reference

Unit mass including space Closed TdS-Pd(;) + Hd(;) +Ed(7) = dU (5.27)

Unit mass excluding space Closed TdS - P*d(;) + JLoH d( ~) + E d(;) = dU* Problem 13.5.1

Total mass including space Closed TdS, - pdV, + Hd(BV,) + Ed(DV,) = dU,
Total mass excluding space Closed T dS, - p* dV, + JLoH dI,.m + E dI,.c = dU~

Unit volume of variable mass Open TdSy + HdB + EdD+ JLdp = dUv (5.4)

Total mass excluding space Open TdS, - p* dV, + JLoH dI,.m + E dI'.e + L JL,dm, = dU~ (6.21)
i-I

Unit mass of variable composition Open TdS - Pd(;) + Hd(;) + Ed(7) + t JL,dx, = dU (6.23)

,
Total mass including space Open TdS,-pdV,+ Hd(BV,) + Ed(DV,) + L JL,dm,=dU, (5.31)"
i=l

u* = u - JLoH2/2p - EoE2/2p I t m = MVt B=l'o(H+M)


p*= P - ~lLoH2 - 4OE2 It.e= pVt E = EoE + P
u~ = Ut - ~lLoH2 Vt - !eoE 2 Vt
a Every equation in the table is obtainable as a reduction or a rearrangement of equation (5.31).
Then, because order of differentiation is commutative, we obtain analogs to 395
Maxwell's thermodynamic relations, Thermodynamics of
Electromagnetism
( af.Li) = (ast ) = -B. (6.29)
aT p',H,m, ami p',T,H

( af.Li) = (av.) = V, (6.30)


ap* T,H,m, ami p',T,H

( -af.Li) = -f.Lo (al


- t) = -f.Loli- (6.31)
aH T,p',m, ami p',T,H

where Bi is the partial mass entropy of species i, V; the partial mass volume, and
L the partial mass dipole moment.
The chemical potential is a function of temperature T, pressure p*, magnetic
field H, and the masses mi of the s species present. Using the definition of mass
fraction Xi = m;/ m" only (s - 1) mass fractions are independent and the
functional dependence of f.Li can be expressed as

f.Li = f.Li(T, p*, H, xJ (6.32)

Because T, p*, H, and Xj are state parameters, f.Li is a function of state and its
total differential can be written as

d .=(af.L i) dT+(af.L i ) d *
f.L. aT. p ,H,Xj
aP* T.H,xj
P

(6.33)

Substitution for the coefficients in equation (6.33) using relations (6.29)-(6.31)


produces the desired relationship for the chemical potential exhibiting the
dependence on the magnetic field parameters:

df.Li = -Bi dT + V; dp* - f.LoL dH + sf (~)


j=l aXj p*,H,T,xj
dXj (6.34)

A generalized form of the Gibbs-Duhen equation is obtainable from equation


(6.21) by integrating to its Euler form, differentiating the results, subtracting
equation (6.21), and dividing each term by the total mass. The result is

1 f.L M s
S dT - - dp* + _ 0- dH + L Xi df.Li = 0 (6.35)
p P i~l

This relationship could be useful in solution thermodynamics for relating the


chemical potential of a species to known values of the chemical potential of the
other species in a mixture, for example, when T, p*, and H are constant.
396 13.7. EQUILIBRIUM IN MULTIPHASE AND MULTICOMPONENT
POLARIZABLE SYSTEMS
Chapter Thirteen

Criteria of equilibrium are bedrock on which multiphase and multi com-


ponent thermodynamic analyses are based. Prior authors dealing with electromag-
netic thermodynamics tacitly assume that the criteria developed for pressure-
volume-temperature (P- V - T) systems apply unaltered for pressure-volume-
temperature-field (P- V - T -H) systems. We return to the fundamental postulate
for an isolated system and from it find that, while the criteria for eqUilibrium of
heat transfer and species transfer are unchanged in the presence of a field, the
criterion of mechanical eqUilibrium is altered from the usual equality of pressure
in the phases. Together with the first- and second-law expressions developed
previously, the results of this section permit the analysis of very diverse
electromagnetic thermodynamic problems.

13.7.1. The Fundamental Postulate

A broad statement of the second law for a system having entropy Ss can be
formulated in terms of the entropy variation dSs written as the sum of two terms,

(7.1.1)

where deS is entropy transferred to the system by its surroundings while diS is
entropy generated within the system.(8) The second law assertion is that diS must
be zero for reversible transformations and positive for irreversible transformations,

(7.1.2)

The entropy supplied, deS, may be positive, zero, or negative, depending on the
interaction of the system with its surroundings. For a system which can exchange
neither heat nor mass with its surroundings, i.e., an adiabatic closed system,
deS = O. For systems which may exchange mass and heat with their surroundings,
deS contains additional terms that can be derived. This is done in Section 13.9,
where diS is normalized to unit volume and time and denoted by the symbol 8.
A system restrained from exchanging heat, mass, or work with its surround-
ings is referred to as isolated. Because an isolated system is a special case of an
adiabatic closed system, it follows that its entropy change in any process must
be nonnegative. In the present section the second law is used in simple form in
reference to isolated systems.
Let us consider a multicomponent, heterogeneous system ~ made up of a
number a of homogeneous, magnetizable phases. Figure 13.4 depicts an isolated
system ~ of the type described located between permeable pole pieces of an
electromagnet. The system is contrained to a constant volume so that no work
of expansion can be performed. The walls enclosing the system are adiabatic and
impermeable so that no heat or mass is exchanged with the surroundings. An
electrical winding of N turns carrying current I establishes a magnetomotiveforce
~ = NI driving a magnetic flux ch. that links the N turns and permeates the
system~ . Associated with an increase offlux d4>~ is electrical work dWe = -~ d4>~ 397
whose value is zero when 4>~ is constant. Thus, a further constraint of constant Thermodynamics of
4>~ is applied to ensure that the system is truly isolated. The constraint can be Electromagnetism
enforced by adjusting the voltage at the terminals of the winding during any
process that takes place. Alternatively, if the winding is a closed circuit super-
conductor, constancy of magnetic flux is guaranteed by Faraday's law and the
properties of the superconductor.
Initially it may be imagined that the a phases are separated by partitions
that are permeable to the magnetic flux 4>~ but impermeable to the flow of heat
or diffusion of species. The material in each of the a regions can be considered
homogeneous and at equilibrium with itself initially. However, the phases may
possibly be at sI1ghtly different temperatures, pressures, magnetic field intensities,
etc., such that, if the partitions are "dissolved," transfer of heat and mass,
displacements of phase boundaries, and redistributions of magnetic field energy
may take place. Applying the fundamental postulate to this process, it is asserted
that the entropy of the isolated system either increases or remains constant,

(7.1.3)

Labeling the phases with the index a , we have

(7.1.4)

The equality in relationship (7.1.3) applies when the system is at equilibrium


while the inequality applies to a natural process. As a consequence S~ is a
maximum at equilibrium subject to the constraints of constant total internal
energy U~ , volume V~, and magnetic flux 4>~ .
The criterion (7.1.3) can be transformed to an equivalent but more convenient
form by assuming that a natural process has taken place in which S~ has increased
differentially. It is then assumed that heat is extracted from the system while
holding V~ and 4>~ constant in order to restore S~ to its original value. From the

(IPermeabl e Yoke (fJ _00

h I

r l
, ------0,
::> \
,
I
I
I
I
I N'c ::>
\
e
I I I
____ L.. I ;:> /
...J <: A/
~

FIGURE 13.4. Sketch to derive


equilibrium criteria of magnet-
ized heterogeneous systems. The
isolated system l is subjected to
constant magnetic flux cP ... .
398 first law, dU~ = dfb. + dW~ where dQ~ is heat added to the system and dW~ is
work done on the system. Because dW~ = 0 and dfb. :s; 0 we have dU~ :s; O. Thus,
Chapter Thirteen
the well-known equivalent criterion is

(7.1.5)

This indicates that total internal energy decreases in a natural process and is
unchanged in a reversible process for the constraints stated. It follows that with
these constraints of constant total entropy, volume, and magnetic flux, the total
internal energy is a minimum at equilibrium. We shall use the statement (7.1.5)
in the next section to derive criteria of equilibrium in terms of intensive variables
characterizing the phases present within the system.
Another form of the equilibrium criterion is useful in operations in which
temperature T, volume v., and magnetic induction B are held constant. Suppose
the apparatus of Figure 13.4 is modified with a' constant volume thermostatic
reservoir brought into thermal contact with the sample of matter contained
between the pole pieces. An isolated system L can be defined comprising the
sample and the thermostatic reservoir. Writing S~ = St + Sr where St is the entropy
of the total sample and Sr the entropy of the reservoir, criterion (7.1.3) can be
written as
d(St + Sr) 2: 0 (7.1.6)

When a process occurs in which an amount of heat dQt flows into the sample,
the reservoir change of entropy is dSr = -dQtl T. Substituting for dQt from the
first law (5.3) with the mass m, constant gives

dSr = -dU./T (7.1.7)

The volume V. is constant so that no mechanical work p dV. is done and, because
B also is constant, no magnetic field work H d(BV.) is done. Combining (7.1.7)
with (7.1.6) and rearranging yields the criterion

dA,:s;O (T, v., B constant) (7.1.8)

where At = Ut - TS, is the Helmholtz free energy for the total sample.

13.7.2. Criteria of Equilibrium in Terms of Intensive Variables

Conditions in the isolated system when the partitions are removed will be
analyzed using criterion (7.1.5). The phases become open systems with respect
to each other and we can express dul a ) for the a phase in terms of the Gibbs
equation previously presented in the form (5.31),

dul a ) = T(a) dsl a ) - p(a) dv,a) + H(a) d(B(a)v,a + L p.~a) dm~a) (7.2.1)
i
Thus, from relationships (7.1.4) and (7.2.1), 399
Thermodynamics of
Electromagnetism

+ L H(a) d(B(a) vla + L L lL~a) dm~a) (7.2.2)


a i

The phase index a takes values 1 to 8, and i is the species index having values
1 to s. Expanding dU~ gives

dU~ = T(t) dslt) - p(1) dVP) + H(l) d(B(1)VP


+ lLit) dmit) + IL&I) dm&1) + ... + IL~I) dm~1)
+ T(2) dsl 2) _ p(2) dvl 2) + H(2) d(B(2)vl 2
+ lLi2 ) dm\2) + 1L&2) dm~2) + ... + 1L~2) dm~2)

+ T(6) dsi 6) _ p(6) dvi O) + H(6) d(B(O)vi O


+ lLiO) dmiO) + IL~O) dm~O) + ... + IL~O) dm~O) (7.2.3)

The individual variations dS;I), etc., are subject to the constraints of constant
total entropy, constant total volume, constant total mass of each species (chemical
reaction excluded), and constant total magnetic flux. From the entropy, volume,
and mass constraints we have

dS~ = dSP) + ... + dS;6) = 0


dV~ = dV;1) + ... + dvl 6) = 0

(dm,)~ = dmp) + ... + dmiO) = 0

(dm2h = dm~1) + ... + dm&6) = 0

(dmsh = dm~1) + ... + dm~6) = 0 (7.2.4)

The magnetic constraint is absent from the above list. This constraint can be
specified in alternate ways, and we shall see that the orientation of the field and
geometric arrangement of the phases influence the result obtained.

Normal Field
It will first be assumed that the phases are arranged in layers, each of a
uniform thickness, with the magnetic flux oriented normal to the layers, as
sketched in Figure 13.5. Using the condition that magnetic induction is continuous
across an interface, we can write

(7.2.5)
400 where B is the uniform value of magnetic induction in all the phases. Thus, the
magnetic induction can be factored out of each differential term in which it
Chapter Thirteen
appears in equation (7.2.3):

(7.2.6)

This summation of internal magnetic energy changes in general does not sum to
zero, even though the electrical work done on the system by the external circuit
is nil. A portion of the magnetic field energy is converted to the other energy
forms, or vice versa.
Collecting terms having the common factor V: a ), the expression for dU~
can be written as

dU~ = T(I) dSP) - (p(1) - H(1) B(I) dVP)

+ f.L \1) dm\1) + f.L~1) dm~1) + ... + f.L~I) dm~1)


+ T(2) dS?) - (p(2) - H(2) B(2 dvf)

+ f.L\2) dm\2) + f.L~2) dm~2) + ... + f.L~2) dm~2)

+ T(8) dsl 8) _ (p(8) _ H(8) B(8 dvl 8)

+ f.L \8) dm\8) + f.L~8) dm~8) + ... + f.L~8) dm~8) (7.2.7)

Equation (7.2.7) has 8(s + 2) independent variables while the number of con-
straints from system (7.2.4) is s + 2. The expression for dU~ may be expressed
in terms of s + 2 fewer independent variables by using the constraining equations
to eliminate, for example, dSP), dvl 1 ), and the s quantities dn~l). The result yields
an expression for dU~ in terms of (8 - l)(s + 2) truly independent variables. The

FIGURE 13.S. Isolated system of (J


homogeneous, layered phases with
orientation of magnetic flux ch normal
to the layers.
expression that results is 401
Thermodynamics of
dU~ = [T(2) - T(1)] dsl 2) + [(p(2) _ H(2)B(2) - (p(l) - H(1)B(l))] d0,2) Electromagnetism
+[1L~2) -1L~1)] dm~2) + ... + [1L~2) -1L~1)] dm~2)
+ [T(3) _ T(1)] dSP) - [(p(3) - H(3) B(3) _ (p(l) - H(1) B(l))] d0,3)

+[ILP) -1L~1)] dm~3) + ... + [1L~3) _1L~1)] dm~3)

Here, the variations dsl2), dV?>' dm~2), dm&2), etc. are truly independent. There-
fore, at equilibrium in the closed system, by using the second-law result that U~
is minimum, and employing the necessary mathematical criteria characterizing
an expansion of a function of multivariables(4) it follows that

au~ au~ aU~ aU~


as?) = 0, av~2) = 0, am?) = 0, am(2) = 0, etc. (7.2.9)

Upon computing the$e various derivatives from expression (7.2.8) for dU~, we
have at internal equilibrium with respect to heat transfer and mass transfer, and
in the presence of a field, the following results:

(7.2.10)

These criteria have the same appearance as the standard results for unpolarized
systems, i.e., temperature is uniform in the system, and the chemical potential
of any given species has the same value in every phase. But, in addition, at
internal equilibrium with respect to boundary displacement, we also obtain the
following result on pressure:

(p(1) _ H(1) B(1)


= (p(2) _ H(2) B(2) = ... = (p(8) _ H(8) B(8) (normal field) (7.2.U)

This result is well known in ferrohydrodynamics(5) but apparently not in


thermodynamics. Unlike in ordinary systems, pressure is not uniform throughout
the system but varies from layer to layer in such a manner that the difference
pea) _ H(a) B(a) is the same in each layer. Because the magnetic induction is the
402 same in each phase, the result (7.2.11) can be expressed alternatively as
Chapter Thirteen
p*(l) + 1-'0 M(l)2 = p*(2) + 1-'0 M(2)2 = ... = p*(8) + 1-'0 M(8)2 (7.2.12)
2 2 2

where definition (6.18) for p* has been introduced. Thus, the pressure can vary
across a flat phase boundary only when there is a difference of the normal
component of magnetization.
From the analogy between electrostatic and magnetostatic systems, the result
(7.2.11) translates for electric systems to

(7.2.13)

and the pressure can vary across a flat phase boundary only when there is a
difference of the normal component of polarization P.

Tangential Field
It is now assumed that the phases are arranged in layers that are tangential
to the direction of the field; see Figure 13.6. From Ampere's law the tangential
component of magnetic field is continuous across a boundary between phases:

(7.2.14)

The general relationship (7.2.3) for dU~ can be particularized to this special case,
recognizing that while the overall size and shape of the vessel containing the (}
phases is unchanged, the boundaries between phases shift and so the cross-
sectional area A;a) of the phases varies with an exchange of species from one
phase to another. We represent the volume of a phase a as vl a) = Ala)/;a) where
l?) = W) = ... = W) = I is the fixed distance between the pole faces of the
permeable yoke. Adding together the magnetic terms of expression (7.2.8) and

--:-1
I
I
V-system ~
I
I
I H~I=H")= ... =H")

I
I
I
I
I FIGURE 13.6. Isolated system of fI
- --+-"'_""':.JI homogeneous, layered phases with
orientation of magnetic flux tPI. tangen-
tial to the layers.
substituting these relations yields 403
HI d(B(1)A(1) + B(2) A(2) + ... + B(8)A(8 Thermodynamics of
Electromagnetism
= HI d(c/J(I) + c/J(2) + ... + c/J(8
=Hldc/J=O (7.2.15)

where the constraint c/J = constant initially imposed on the system was employed.
Thus, the sum of the magnetic terms vanish, indicating that for this tangential
field system there is no net conversion of magnetic field energy to the other forms.
The remainder of the argument proceeds as in the case of normal field
orientation developed previously and results again in the equilibrium criteria
(7.2.10). In addition, with tangential magnetic field the pressure as well as the
temperature and chemical potential of each species are uniform throughout the
system, similar to conditions in unpolarized matter. Then p(J) = p(2) = ... = p(9),
or equivalently, because of the continuity of the tangential magnetic field intensity,
p*(I) = p*(2) = ... = p*(8) (tangential field) (7.2.16)

For a given orientation and intensity of the field, fixed mass ratios of the
species, and given temperature and external pressure, the equilibria remain the
same, regardless of the constraints, for we assume that local conditions govern
the distribution of species among the phases and the variation of pressure from
phase to phase. The constraints chosen for the derivation were selected for
convenience and not as a necessity. For example, if the magnetomotive force ~e
rather than the magnetic flux c/J~ were held invariant, the criterion that U~ is
minimum would be altered, but the final results (7.2.10) and (7.2.11) or (7.2.16)
are unchanged.
When the species in the system react chemically with each other, more
elaborate derivations are required to establish the equilibrium criteria.

An Experimental Verification
The relationships (7.2.12) and (7.2.16) are manifested in measuring pressure
in magnetizable fluids using a thin flexible liquid-filled sensor connected to a
manometer.(6) If the sensor surface is oriented perpendicular to the field, the
manometer registers a pressure that is larger by the magnitude JLoM2/2 than
when the sensor surface is oriented tangential to the field. Denoting the pressure
difference by t:.p, the magnetization of the fluid is given as M = (2t:.p/ JLO)I/2.
Figure 13.7 shows curves of the magnetization of a ferrofluid comprising collodial
magnetite in kerosene, calculated from pressure measurements parallel and per-
pendicular to the field. Values of magnetization determined with a ballistic
galvanometer are plotted on the curves for comparison. It can be seen that the
agreement between the two methods of measurement is close, aside from what
appears to be a small systematic difference.
The same pressure jump mechanism accounts for the behavior of a magnetic
fluid drop surrounded by nonmagnetic fluid and subjected to a uniform applied
magnetic field. An equilibrium is established between the jump and capillary
404 pressure due to curvature of the drop surface leading to elongation of the drop
in the direction along the magnetic field.
Chapter Thirteen
The steps to obtaining the relationship /:l.p = /LoM2/2 are as follows. Let (2)
denote a position outside the magnetic fluid, hence within the nonmagnetic
manometer fluid, and (1) a position inside the magnetic fluid; while subscripts
nand t denote normal and tangential orientation of the sensor interface relative
to the field. Then relationships (7.2.12) and (7.2.16) can be stated as

P*(1)
n
+ /Lo
2
M(1)2 = p*(2) + /Lo M(2)2 = P
n 2 e.n (7.2.17)

and
(7.2.18)

where it is recognized that M~2) = 0, p!(2) = Pe.n, and pt(2) = Pe., in the nonmag-
netic manometer fluid. Manometer fluid pressure with normal orientation of the
field has been denoted by Pe.n, and Pe., denotes the manometer fluid pressure
with tangential field orientation. From its definition (6.17) p* is a scalar quantity,
hence for a constant value H of the field p!(l) = pt(l) and subtraction of equation
(7.2.17) from (7.2.18) gives

(7.2.19)

where /:l.p = Pe.n - Pe., and M = M(l).


The condition of interfacial mechanical equilibrium plays an important role
in problems of meniscus shape and the interfacial stability of magnetically(5) and
electrically(36) polarized fluids. An interface subject to oblique orientation of the
field experiences a normal stress difference J.LO(M~2)2 - M~1)2)/2 where subscript
n denotes normal component, and there is no shear stress of magnetic origin;
both these conclusions may be established from application of the magnetic stress
tensor developed as equation (9.1.25) below.

FIGURE 13.7. Magnetization curves of a ferroftuid


determined by measurements of pressure difference
(open circles) and magnetometry (closed circles);
the close agreement furnishes a critical test of
relationship (7.2.11). M, denotes saturation value
Magnetic Field, H (kA/m) of magnetization. (Data after Chekanov.(6)
13.8. APPLICATIONS TO PROBLEMS OF EQUILIBRIUM IN A FIELD 405
Thermodynamics of
13.8.1. Phase Change Electromagnetism

Let us consider two phases denoted by (1) and (2) in equilibrium with each
other in the presence of a magnetic field. The criteria of equilibrium with respect
to transfer of species, from system (7.2.10), requires equality of the chemical
potential of each ith species in the separate phases:

(8.1.1)

For a small change in temperature, which may be caused by a change of pressure


or of the intensity of the magnetic field, equilibrium is maintained, so that

(8.1.2)

.1,
For a pure species .4lS =g, .4lV = 'Y, and .4llm = where a script font denotes
molar, subscript m denotes magnetic, and .4l denotes molecular weight. Thus,
from relations (8.1.2) and (6.34) with dxj = 0,

- g(1) dT(1) + '0 1) dp*(1) - 11-0.1(1) dH(I)


= _g(2) dT(2) + 'Y(2) dp*(2) - 11-0.1(2) dH(2) (8.1.3)

This is a general expression requiring separate consideration for the different


possible orientations of the applied field.

Tangential Field
When the magnetic field is oriented tangentially to the interface between flat
layers of two phases, the criterion (7.2.16) for mechanical equilibrium requires
p(1) = p(2). From expression (6.18), recognizing that T(1) = T(2) = Te and H(1) =
H(2) = H. = Bel 11-0 , where Be is the applied induction field, for a system in
equilibrium at the environmental pressure Pe

(p*)(I) = (p*)(2) = Pe (8.1.4)

Because Pe is constant, (dp*)(1) = (dp*)(2) = 0 and equation (8.1.3) rearranges to

aT) = _(.1(2)
(aBe P.
- .1(1)
g(2) _ g(1) (8.1.5)

At constant temperature, equations (8.1.3) and (8.1.4) yield

(8.1.6)
406 Normal Field
Chapter Thirteen
The mechanical equilibrium condition (7.2.11) is i l ) - H(I) B(I) =

p(2) _ H(2) B(2), from which it follows [see equation (7.2.19)] that an equivalent
statement is
(8.1.7)

where

(8.1.8)

Equation (8.1.3) can be rewritten by adding and subtracting r dPn on both


sides. The result can be expressed in the form

[-y dT + r d(p* + Pn).1 dB](I)


= [-y dT + r d(p* + Pn).1 dB] (2) (8.1.9)

If it is assumed that the phase change equilibrium occurs at the environmental


pressure Pe, then P* + Pn = Pe so that d(p* + Pn) = o. Using the condition of
thermal equilibrium T(I) = T(2) = T, and continuity of the field of magnetic
induction B(I) = B(2) = Be, equation (8.1.9) yields

(
aT) = _(.1(2) - .1(1)
(8.1.10)
aBe P. y(2) - y(l)

which is identical to equation (8.1.5). Quantity B. is the magnetic induction field


in the absence of the material, and Pe is the environmental pressure.
Similarly, it can be shown that equation (8.1.6) applies when the field
orientation is normal to the interface between the phases.
In summary, although orientation of the magnetic induction affects the
pressure and the magnetic field within separate phases, when the pressure is
referred to the suroundings and the field to the external field in the absence of
matter, the relationships for phase change assume a common form.

A note on the magnetism of materials. Classical atomic models describe an


atom as electrons orbiting about a positively charged nucleus. Particles of the
nucleus and individual electrons can also be imagined to be spinning. The orbital
and the spinning motions are thus sources of a magnetic field. The electron orbital
motions produce diamagnetism, in which the induced dipoles oppose the applied
field, usually a small effect.
Paramagnetism is due to electronic spin and, in quantum theory, the spin
is not a classical quantity but an inherent property of the electron. The spin
moment is quantized with the smallest unit of magnetic moment known as the
Bohr magneton II-B' 407
Thermodynamics of
Electromagnetism

where e = 1.602 x 10- 19 C is the electron charge, h = 6.626 X 10-34 J s is Planck's


constant, and me = 9.11 x 10-3 kg is the electron rest mass. The Bohr magneton
closely equals the spin magnetic moment of a free electron. An atom such as H
or a free radical such as CH 3 possesses an unpaired electron (denoted by the
dot) and a magnetic moment of about one Bohr magneton. The magnetic moment
of oxygen is equivalent to about two Bohr magnetons per molecule.

EXAMPLE 13.8.1. Influence of Magnetic Field on Boiling Point of Liquid Oxygen


Unlike most common gases, oxygen gas is paramagnetic as is liquid oxygen. The
susceptibility of oxygen vapor is somewhat greater than that of the liquid. One can think
of a Le Chatelier principle in which liquid converts to vapor, attracted by the field, and
accordingly the boiling temperature should be reduced in the field.
To compute this reduction in temperature the molar entropy difference is expressed
as
y<v) _ g(l) = AI T (E.8.1.1)

where A is the molar latent heat. The molar moments are approximately given by

(E.8.l.l)

and
(E.8.t.3)

where XCv) and x(l) are mass susceptibilities defined as volumetric susceptibility divided
by density and, as previously defined, .4l is the molecular weight. Thus equation (8.1.8)
can be written in the form

(E.8.1.4)

The susceptibilities and A decrease with an increase in temperature. Assuming that suitable
average values can be chosen, the integration of this equation gives

T(H)
- - = exp [B;.4l
- - - (x (v) - x(l) ] (E.8.1.5)
T(O) 2P,oA

Data
A = 75.5 J mor l K- I
T(O) = 90 K (atmospheric boiling point in the absence of a field)
XCv) = 4.345 X 10-6 m3 kg- 1
X(I) = 2.920 X 10-6 m3 kg- 1

.4l = 0.032 kg mol- I


p = 1180 kg m-3
On substituting the data into equation (E.8.1.5) it is found, when B. = 10 T, that AT =
T(H) - T(O) = -2.1 kelvins, and for Be = 50 T the value AT = -40.6 kelvins. This effect
408 will alter the normal vapor equilibrium curve of liquefied air and might result in an easier
separation of the gases. Prediction of the behavior of the two-phase equilibrium of mixtures,
Chapter Thirteen such as nitrogen, oxygen, and argon, is complicated by the possibility ofnonideaI behavior.
In computations authors(I.2t) err in replacing the difference in the molar moment by
the product of a susceptibility (assumed the same in both phases) and the difference in
the molar volume of the phases. Meachin and Biddulph(2) report measurements at 77.4 K
in which the vapor pressure of oxygen increases in the field.

PROBLEM 13.8.1. Influence of an Electric Field on the Boiling Point of Oxygen


(a) Using the analogy developed in Section 13.3.2, write the electric field relationship correspond-
ing to equation (E.8.1.5).
(b) The dielectric constant of O2 vapor in equilibrium with the liquid is 1.000523, and the liquid
has a dielectric constant of 1.507. Compute the shift in the boiling temperature at an applied electric
field of 10 volts/meter. Is the boiling point elevated or depressed?
(c) Give an explanation why the maximum electric field influence on boiling temperature is
much smaller than the magnetic field effect.
The relationship between vapor and liquid dielectric constants is an interesting problem on
which much work has been done:(24)

13.S.2. Chemical Reactions in a Field

A general homogeneous reaction can be written as

aA + bB + ... = cC + dD + ... (8.2.1)

Denoting the ith chemical reactant by Ri the reaction can be rewritten compactly
as
s
0= L ViRi (8.2.2)
i=1

where Vi are whole number stoichiometric coefficients and s is the total number
of species. The values of Vi for species that appear on the left-hand side of
equation (8.2.1) are thereby assigned negative values. Let us consider a reaction
that proceeds a differential amount and denote the change in mols of the general
species i as dni' From relation (8.2.1) the ratio of mole change to stoichiometric
coefficient is the same for all the species:

dnA dnB dni


-=-==-=d~ (8.2.3)
VA VB Vi

where ~ is known as the extent of reaction.


Relationships governing the equilibrium of the reaction can be obtained
from criterion (7.1.5) with subscript t (for total) in place of I. (for system) as the
reactants are present in a single homogeneous phase:

dU.:s; 0 (S., V;, CPt constant) (8.2.4)


From the fundamental equation (5.31) the total internal energy for the 409
mixture present in the system is Thermodynamics of
Electromagnetism
,
dUt = TdSt - pdVt + Hd(BVt) + I iLidni (8.2.5)
i

and, assuming the magnetic induction B is uniform over area At of volume


Vt = Atlt , the magnetic term can be rewritten H deB\/;) = Hit dcPt. Then, for the
constraints of constant entropy St, volume Vt , and flux cPt> equation (8.2.5)
reduces to

dUt = I iLidmi (St, Vt , cPt constant) (8.2.6)

Using relationship (8.2.3), the internal energy is dUt = I; lI;iL;.M; dt, and applying
criterion (7.1.5) with the equal sign in the form a Uti at = 0, we obtain the general
criterion for the chemical equilibrium,

aUt
ag = ~ lIiiL;.M; = 0
S
(8.2.7)

The relationship of the second equality displays the usual form for reactions
unaffected by a field. However, in the present case the chemical potentials iL; are
magnetic-field dependent and so integral expressions are next developed for the
chemical potential by integrating the differential form (6.34);

_ _
diL; = -S; dT + V; dp* - iLol; dH
_ +I
S-l(a~iL ;) dXj (8.2.8)
j uXj p*,H,T,nj

When applying equation (8.2.8) to one mole of a pure species in the absence of
a magnetic field, .MS; = g, .MY, = 'Y, .Ml, = off = 0, p* = p, dXj = 0 for all j, and
diL; = diL where .M is the molecular weight. Hence

.MdiL = -gdT+ 'Ydp (8.2.9)

For one mole of an ideal gas

'Y = RTlp (8.2.10)

Thus, from equation (8.2.9) at constant temperature,

(8.2.11)
410 where Po is a reference level of pressure at which /. = /.o(T). Integration between
the indicated limits yields
Chapter Thirteen

(
RT p / po
/. T,p,H=O)=/. (T)+-:ijln (8.2.12)

Reference pressure Po is conventionally selected as one atmosphere. The conven-


tion is adopted here, so the final term of equation (8.2.12) becomes RT In P with
p expressed in atmospheres. An expression for the chemical potential qf a species
in a gas mixture can now be developed. We consider a gas mixture under pressure
P having mol fraction Xi of species i. The gas mixture is equilibrated through a
semipermeable membrane with pure species i contained in a vessel at total
pressure Pi' It is seen from system (7.2.10) that the chemical potential of the
species in mixture equals that of the pure species, so the chemical potential can
immediately be written as

(8.2.13)

At constant temperature T, pressure p*, and composition Xj, equation (8.2.8)


yields

(8.2.14)

On integrating this latter equation between field limits 0 to H, we obtain

/'i(T, p*, H) = /'i(T, p, H = 0) - /.0 ~H


I
(8.2.15)

where the circumflex denotes the field average quantity defined by

.1. = -1
A

I HO'
H
i
.1dH
- (8.2.16)

where ji denotes the partial molar magnetic moment of the ith species.
The function of integration /.i( T, p, H = 0) in equation (8.2.15) is given by
expression (8.2.13). Thus, the chemical potential for a component of the gas
mixture can finally be expressed in the form

(8.2.17)

This relationship is applicable to a magnetic species having a linear or nonlinear


magnetization curve. An example of the latter is a paramagnetic species that is
magnetized beyond the constant initial susceptibility range.
The application of these relationships to the equilibrium of a homogeneous, 411
reactive gas mixture is now examined. Substitution of the chemical potential Thermodynamics of
(8.2.17) into the general criterion (8.2.7) yields Electromagnetism

(8.2.18)

The relationship can alternatively be written as

(8.2.19)

where
s
Kp IT p;'
= i=1 (8.2.20)

where IT;~l denotes the product of the factors; Kp is well known as the definition
of the equilibrium constant. Because JL?( T) is synonymous with free energy per
unit mass of the pure species at the reference pressure [set G~ = G t for H = 0
in the last equation in system (6.28)], the quantity L;~l ViAtiJL?( T) in equation
(8.2.19) is conventionally denoted as the standard change of free energy .6.G~ of
the reaction at temperature T. Quantity .6.G~ is the standard free energy change
accompanying total conversion of a stoichiometric amount of reactants to
products in the absence of a field. With this nomenclature relationship (8.2.19)
can be written as

.6.G(B) = .6.G~ + .6.G~ (8.2.21)

where
.6.G(B) = -RTIn Kp(B) (8.2.22)

.6.G~ = -RTIn Kp(O) (8.2.23)

s
.6.G~ = - L VijiB (8.2.24)
i=1

and B = JLoH is a reference induction field corresponding to the magnetic field


H. As can be seen, the influence of magnetism is to shift the free energy change
by an amount .6.G~, which we will term the standard change of magnetic energy.
When the products of the reaction are more magnetizable than the reactants, the
effect of the magnetic field is to drive the reaction to the right. These are the
desired relationships with which to determine the equilibrium distribution of
species partial pressure in a reactive, magnetizable, gas mixture. The relationships
reduce for linearly magnetizable media to the result stated by Kaneto. (37)
As a concluding remark, we note it does not matter that the reaction
conditions of interest in a particular circumstance are different from the con-
straints (8.2.4). For example, the pressure may be constant and the volume v..
412 variable, while the magnetic field rather than the flux 4Jt is held constant. The
reason for this is that equilibrium at a given pressure, volume, field, and number
Chapter Thirteen
of moles of each species must be the same regardless of the path followed to
arrive at that equilibrium state. Thus, we were able to establish the equilibrium
relationships based on using constraints {8.2.4} which were chosen for con-
venience only.

EXAMPLE 13.8.2. Magnetization of a Free Radical Species


The unpaired electron of a free radical, such as atomic hydrogen or a methyl radical,
is a permanent magnetic dipole. In the absence of an applied field the orientation of the
dipole is random and a collection of the free radicals exhibits no net magnetic moment.
Quantum mechanics teaches that in the presence of an applied magnetic field the dipolar
spin (projection along the field direction) aligns either parallel or antiparallel to the field,
but not in intermediate orientations. Denoting the parallel orientation of the radical by
Mt and the antiparallel orientation by M~ the process of magnetization can be represented
as a reaction equilibrium
M~'=: Mt (E.8.2.1)

Assuming the radicals are present as a gas phase, an equilibrium constant can be written
as
(E.8.2.2)

From equation (8.2.21), recognizing that for this system .Ii takes only two values,
(E.8.2.3)

where .9'f is the molar moment of an up spin and .9'~ the moment of a down spin. The
value of Kp(O) is 1, because the. number of up and down spins in a given direction is
equal in the absence of an applied magnetic field. Rearranging the equation and using
relation (E.8.2.2) thus gives

(E.8.2.4)

If it is assumed that the magnetic moment of a spin is equivalent to one Bohr magneton
ILB' we obtain (J!f - J!~) = NILB - (- NILB) = 2NILB where N is the Avogadro number.
Thus,

(E~.2.5)

where kB is the Boltzmann constant equal to 1.38062 x 10-23 J K- I and ILB =


9.27410 X 10-24 J r l . Substitution into equation (E.8.2.4) gives

PM~ = ex p{2ILBB} (E.8.2.6)


PMf kBT

Magnetization is conventionally expressed in terms of fractional saturation defined as


.1 / .1. where .1 is the molar magnetic moment (net moment of the collection of up and
down spins), and !f, is the saturation moment of a mole; 413
Thermodynamics of
fractional saturation = .!... = PMt - PM, (E.8.2.7) Electromagnetism
.1, PMt + PM,
where an ideal gas mixture is assumed with the number density proportional to the partial
pressure. Then, by combining relationships (E.8.2.7) and (E.8.2.6) to eliminate the ratio
of partial pressures, we obtain the magnetization law in the form

!fl!f, = tanh a (E.8.2.8)

where
(E.8.2.9)

The product /LoB measures the energy of one spin in the field, and ko T the thermal energy
per particle. This relationship shows that the dimensionless ratio of these energy terms
determines the fractional saturation.
The molar absolute value of the radical's magnetization at saturation is 45.9% that
of iron, for which the Bohr magneton number is 2.18 at room temperature. Calculated
values of the fractional saturation vs. applied magnetic field for several values of absolute
temperature T are shown in Figure 13.8.

EXAMPLE 13.8.3. Dissociation of Methane in a Magnetic Field


Consider the chemical equilibrium for dissociation of methane into a methyl free
radical CH 3 and a hydrogen free radical H (hydrogen atom),

The dot indicates the presence of an unpaired electron in the molecular fragment or atom.
From the prior discussion the unpaired electron carries a magnetic moment, and so the
free radicals are magnetizable. In the presence of an applied magnetic field a fraction a
of the free radicals has magnetic moment aligned with the field while the remaining
fraction (1 - a) is aligned against the field. The net moment per mole !fi =

FIGURE 13.8. Magnetization of a


one-spin species, such as a methyl 6 10
radical or hydrogen atom. 8(hT or Megagauss)
414 a.1'f + (I - a).1,. where .1'f and 9>,. are as defined in Example 13.8.2. From equation
(8.2.24) we have
Chapter Thirteen
m
aG~ = - L ",j,B
;=1

(E.8.3.1)

aG~ = -2jB

= -2/Lo IH .1 dH (E.8.3.2)

where definition (8.2.16) was introduced with j assumed equal to 9>. Equation (E.8.2.8)
yields

B)
/LB-
.1 = .1, tanh ( - (E.8.3.3)
kBT
Thus

aG~ = -2/Lo.1, IH tanh(/LBB/kBT) dH


= -2RT La tanh a da (E.8.3.4)

where .1, = N/LB, and a = /LBB/kBT as defined previously. The integral above is listed
giving
aG~ = - RT In cosh2 a (E.8.3.5)

Thus, from the relationships of (8.2.21) to (8.2.23),

K(B)/ K(O) = cosh2 a (E.8.3.6)

where

(E.8.3.7)

(E.8.3.8)

Under conditions where the extent of dissociation is small, PCH 4 (B) = PCH 4 (O). Then the
increase in free radical partial pressure in the field is given by

PCH,.(B) = PH(B) = cosh a (E.8.3.9)


PCH,.(O) PH(O)

Table 13.5 lists results of computation for the dissociation of methane using this formula.
As the field increases, the magnetization of the product free radicals approaches saturation
far in advance of substantial conversion of the methane to free radical products.
TABLE 13.5 415
Methane Dissociation in a Magnetic Field (298 K)
Thermodynamics of
Electromagnetism
B (tesla) a Fractional saturation Free radical ratio

0 0 0 1.000
200 0.4508 0.4225 1.l03
400 0.9016 0.7171 1.435
600 1.352 0.8745 2.062
800 1.803 0.9471 3.116
1000 2.254 0.9782 4.815
2000 4.508 0.9998 45.38
00 00 1.0000 very large

Explosively compressed magnetic fields have generated pulse field intensities in excess
of 11r tesla. Fields of Hf tesla are found in magnetic white dwarf stars while pulsars give
indication of field strength equal to about 108 tesla. (31)
Water is too stable to be easily decomposed at moderate temperatures into hydrogen
and oxygen according to

However, because oxygen is paramagnetic while water and hydrogen are not, application
of a magnetic field to water tends to enhance the reaction, thus extracting oxygen. Cur1(30)
studied the thermomagnetochemistry and concluded that a field in excess of 10 tesla is
required. While this field level is daunting the study holds interest for the engineering
problems that it addresses. For example, the oxygen produced is forcefully attracted by
the field so that its removal requires expenditure of energy.

PROBLEM 13.8.2. Free Energy of a Magnetized Liquid Mixture


(a) From the definition of the Helmholtz free energy for a total system

At=Ut-TSt (P.8.2.1)

and the Gibbs equation (5.31), derive the relationship applicable to a binary mixture,

dA t = -St dT + H d(BVt ) + JAo, dm, + JAo, dm, (P.8.2.2)

(b) For conditions of constant temperature T, magnetic induction B, and total volume V" derive
the following relationship for free energy per mol from equation (P8.2.2),

(P.8.2.3)

(c) An expression relating chemical potential to magnetic field intensity and the partial pressure
of a species in a perfect gas mixture was developed as equation (8.2.17). Assuming Raoult's law is
obeyed in the form p, = p~c" where p~ is the vapor pressure of pure liquid species i and c, is the
mol fraction in the liquid mixture, show that the chemical potential for unit mass of species 1 in the
liquid mixture is given by

(P.8.2.4)

where 9, is defined by expression (8.2.16), and that a similar relationship applies for species 2.
416 (d) Using the results of (c) and (b) above, show that the/ree energy o/mixing per mole of liquid
mixture is given by the expression
Chapter Thirteen
(P.8.2.5)

where silO = (c,.4l'/L~ + C2.4l2/L~) + RT(c,lnp~ + c2 1n p~) is the free energy of the amount of the pure
components making up the mixture.
Show that the magnetic terms on the righthand side of expression (P.8.2.5) reduce to
-(/LoH 2/2)(c,xl l ) + C2X~I) when the molar susceptibilities XII) and X~I) can be assumed constant.
It is seen from expression (P.8.2.5) that the free energy of mixing is negative and decreases with
increasing intensity of the applied magnetic field.

13.8.3. Magnetocaloric Processes

There are terrible problems said the doctor who performed the
first human heart transplant in 1968. "The main one is to
develop an inexhaustible, implantable, non-heat-producing
power source of energy." If so he speculated, "It would be in
autos first, then refrigerators and finally in biological devices. "
The New York Times, January 13, 1988

Magnetic materials having a temperature-dependent magnetization heat up


when they are placed in a magnetic field, and cool down when they are removed .
. The phenomenon is known as the magnetocaloric effect. Magnetic refrigeration
using paramagnetic salts has been used since 1933 in this way as a batch process
to reduce temperatures of research samples to a small fraction of a kelvin. At
elevated temperatures the effect is most pronounced near the Curie point of a
ferromagnetic sample which is the temperature above which spontaneous mag-
netization disappears. Engineers have studied the magnetocaloric effect in cyclic
processes as a nonmechanical pump of magnetizable liquid in heat pipes for
cooling electronic equipment; solar collectors for circulating heated fluid; and,
more ambitiously, generation of power in direct conversion cycles. Ferromagnetic
elemental gadolinium with a Curie temperature near room temperature can be
mechanically cycled in and out of a magnetic field to achieve refrigeration, and
materials with optimized Curie temperature can be produced for a broad range
of operating temperatures.
An analysis of the relationships involved is conveniently begun with the
first-law statement (5.4) assuming constant mass density p and absence of electric
field E. The first law for a unit volume of magnetic substance then reads

dQv+ HdB = dUv (8.3.1)

Suppose we wish to determine the temperature change of a magnetizable


material put into a magnetic field. Assuming the process is adiabatic, dQv = 0
and equation (8.3.1) simplifies to H dB = dUv' Experience in P- V - T thermody-
namics suggests that the internal energy be expressed as the product of a specific
heat and a temperature increment. However, if matter is absent so that B = JLoH,
then dUv = d(JLoH2/2) and the right-hand side of equation (8.3.1) bears no
relationship to temperature. Let us examine whether it will help if this amount
is subtracted from both sides of the equation leaving JLoR dM = dU~ where
U~ = Uv - JLoH2/2. Then, because magnetization M depends on temperature 417
and field, the modified internal energy U~ must also. A methodology for handling
Thermodynamics of
this more complex problem is presented in what follows. Electromagnetism
Thus, if magnetic field H and temperature T are chosen as independent
variables, then U v = Uv(H, T) so that the differential can be written as

dUv = (a Uv) dH + (a Uv) dT (8.3.2)


aH T aT H

With M = M(H, T) the differential of the magnetization has the form

dM = (ilM\ dH + (aM\ dT (8.3.3)


aH)T aT)H

and with the defining equation B = JLo(H + M) the differential of the magnetic
induction can be expressed as

dB = JLo[ 1 + e~ T] dH + JLOe~ H dT (8.3.4)

Substitution of expressions (8.3.4) and (8.3.2) into equation (8.3.1) and collecting
common terms yields the differential of heat added dQv as

dQv = e(H, T) dT + g(H, T) dH (8.3.5)

where the functions e(H, T) and g(H, T) satisfy the equations

( auv) = e(H, T) + JLOH(ilM\ (8.3.6)


aT H aT)H
and

(8.3.7)

In equation (8.3.5), e(H, T) is a field-dependent specific heat while g(H, T)


accounts for transfer of heat when magnetic field changes. Because Uv is a
function of state its differential is exact in the mathematical sense, so
a(aUv/aT)/aH = a(aUv/aH)/aT giving from equations (8.3.6) and (8.3.7)

which simplifies to

(8.3.8)
418 For a reversible process dSv = dQvl T, so from equation (8.3.5) dSv is expressed
as
Chapter Thirteen

dS = c(H, T) dT + g(H, T) dH (8.3.9)


v T T

Because entropy is a function of state, its differential is also exact. Hence equation
(8.3.9) yields

ac) = (aaTg ) -:rg


(aH T H
(8.3.10)

On comparing equations (8.3.10) and (8.3.8) it is seen that

(8.3.11)

Substitution of this expression for g into equation (8.3.10) gives

(8.3.12)

Near the Curie temperature Be the saturation magnetization of a ferro-


magnetic body can be expressed as M = K(Be - T), where K has a positive
value known as the pyromagnetic coefficient,

K=-e~L (8.3.13)

From equation (8.3.12) with K taken as constant, it can be seen that the specific
heat becomes independent of the field intensity,

c(H, T) = c(T) (8.3.14)

where c( T) is the ordinary specific heat.


From expression (8.3.11) the function g is determined as

(8.3.15)

Integration of equation (8.3.9) using the above and further assuming c(T) = Co
is constant gives an explicit expression for the entropy of the ferromagnetic
substance:
Sv = Co In T - lLoKH + constant (8.3.16)

The first term on the right-hand side displays the familiar logarithmic increase
of entropy with temperature.
EXAMPLE 13.4. Cooling with Adiabatic Demagnetization of Elemental Gadolinium 419
Consider the process of temperature reduction that occurs when a magnetic substance
is rapidly removed from a region of magnetic field (adiabatic demagnetization). Denoting Thermodynamics of
Electromagnetism
initial conditions by subscript i and final conditions by f, expression (8.3.16) with (Sv)i =
(Sv)' yields

(E.8A.I)

If the material is the ferromagnetic element gadolinium (Gd) at its near room temperature
Curie point, we have 1i = 8 = 295 K, Co '" 4.2 x lIf J m-3 K- J , and K = 29,500 A m- J K- J
(370 gauss K- J ). If the magnetic induction B = P-oH = 7 T (70,000 gauss), such as is
routinely produced with superconducting magnets, the adiabatic temperature decrease is
(Ti - T,) = 14 K.
Figure 13.9 illustrates experimental values of the isentropic temperature change
produced by applying a 7 T field to a Gd sample at various initial temperatures. The effect
persists above the Curie temperature 8e because any material that is ferromagnetic below
8e is paramagnetic above 8e The peaked shape of the curve is related to the changing
value of the pyromagnetic coefficient. Brown(17) computes a thermodynamically determined
curve showing a reasonable match to the data of Figure 13.9, using the known spin state
of 7/2 for gadolinium to estimate M(T,H).

PROBLEM 13.8.3. Cyclic Work in a Magnetocaloric Power System


Consider a power cycle consisting of two isotherms joined by two constant field paths. By
referring to expression (8.3.16) for entropy, sketch the cycle on temperature T (ordinate) vs. entropy
S. (abscissa) coordinates. Describe in words the physical processes occurring along each leg around
the cycle. Which direction around the cycle corresponds to the conversion of heat to work, and vice
versa? Derive the following expression for cyclic work assuming magnetic field intensity is negligible
in the low field path:

(P.8.3.I)

where subscript c denotes cold and h hot.

PROBLEM 13.8.4. Science Fair Motor


A magnetic motor sometimes demonstrated at science fairs consists of the rotatable, steel wheel
of a bicycle mounted with the wheel axis horizontal. A source of heat warms a short sector on the
rim of the wheel located between the poles of a permanent magnet, and it is observed that the wheel

g 18

~ 14

e~
,,;
f 12 00
!() 0
\
I
10
I!
i"
00
8 0
l 0
0
E
. 8 ~
f! of} 00
00
FIGURE 13.9. The isentropic temperature A. 4 0 0
change produced by applying a 7 tesla mag- ~ 2
netic field to elemental gadolinium (after J 200 240 280 320 380
Brown(I7). Initial Tempenltu.. (I<)
420 rotates continuously. Give a mechanistic explanation of this phenomenon and a thermodynamic
explanation in terms of the construction (Mollier diagram) of Problem 13.8.3.
Chapter Thirteen Analysis(S) of the conversion of heat to work for a particular recuperative cycle shows that the
efficiency can approach that of a Carnot engine which no heat engine can exceed.

PROBLEM 13.8.5. Magnetocaloric Refrigeration


In a magnetocaloric cycle for refrigeration discussed by Brown, (17) a recuperative cycle is
described employing a regenerator composed of a liquid column in a vertical tube with magnetic
field impressed over the middle and upper zones of the column. The rare earth Gd fabricated into
a movable element with a large heat transfer area passes cyclically up and down through the entire
column of liquid. In steady operation a temperature gradient is established with the hotter fluid at
the top, where heat is rejected to the surroundings. Describe the process of heat recovery in this cycle
with reference to the Mollier diagram of Problem 13.8.3.
Sketch an apparatus using the principle of this device to liquefy natural gas on a continuous
basis. What modifications would be desirable if the device were used to separate air?

13.9. APPLICATIONS TO PROBLEMS OF TRANSPORT IN A FIELD

In describing momentum transport as well as transport of a scalar property


such as mass or heat when the material is polarizable, the fundamental thermody-
namic assertion is the nonnegative character of the internal entropy generation
rate. To this end, an equation of entropy balance is introduced below and is
applied to problems of momentum and mass transport in magnetic fields.

13.9.1. Momentum Transfer

Until recently the engineering applications of fluid mechanics were largely


restricted to systems in which magnetic and electric fields play no role. However,
the interaction of electromagnetic fields and fluids has been attracting increasing
attention, with the study of the interactions conveniently divided into three main
categories: magnetohydrodynamics or MHO, which treats magnetic fields and
electric currents interacting in electrically conductive fluids such as liquid metals
and ionized gases; electrohydrodynamics or EHD, in which force arises due to
free charge acted upon by an electric field; and flows in which no electric current
need flow while the forces and torques that arise are due to interaction of the
field with the magnetic or electric polarization of matter. The connection with
thermodynamics is intimate in the latter category and forms the subject of this
discussion.
When the fluid is strongly polarized in a ferromagnetic or possibly ferro-
electric sense, the interactions that develop are intense compared to usual para-
magnetic or dielectric interactions in modest intensity fields and the resultant
topic of study has been termed je"ohydrodynamicS<S) (FHD).
The ultrastable colloidal dispersion of subdomain magnetic particles in a
carrier fluid (je"ojluid) furnishes the most common example of a magnetizable
liquid. A typical ferrofluid contains magnetite (Fe304) particles of approximately
100 A size as the magnetic constituent. Each particle is coated with a molecular
adsorbed layer, such as oleic acid when the carrier liquid is a hydrocarbon liquid.
The adsorbed layer prevents particles from clustering to each other, and thermal 421
motion prevents the particles from settling. Aqueous solutions of paramagnetic
Thermodynamics of
salts furnish additional examples of magnetizable fluids. (15) Electromagnetism
Colloidal magnetic fluids have found widespread application in: zero-leakage
rotary shaft seals; precision fluid film bearings; motion dampers; motion sensors;
and as a coolant in acoustic transducers. Diverse other applications have been
studied in power conversion, oil production, printing, and other fields. In
sufficiently high fields ordinary paramagnetic fluids become highly magnetized,
and in low gravity operations even modest field intensities can provide significant
forces driving fluid motions.
Basic to understanding all flows of polarized media is the formulation of
the equations of motion and constitutive relationships that determine the
dynamics. Thus, one is concerned with formulating the electromagnetic stress
tensor from which can be found body and surface forces acting on the fluid.
Body and surface couples can also arise due to rotational motion of the particles
relative to the fluid. Then, in addition to usual viscous stresses resisting the
motion, a number of other phenomena need to be considered including transfer
of angular momentum from particles to fluid, diffusion of angular momentum,
and relaxation of magnetization. A systematic methodology for tre.ating these
problems attempts to combine second-law thermodynamics with the familiar
balance laws of mass and momentum. (7) From an entropy balance based on
expressions of physical rate processes, guidance is obtained in formulating
acceptable constitutive equations. This method of treating nonequilibrium
phenomena is part of what is usually termed irreversible thermodynamics,
although coupling need not exist in the Onsager sense. (8)
The following utilizes the nomenclature of stress tensors. Thus, at any
arbitrary point P in a continuum, the stress principle of Cauchy associates a stress
vector t with each unit normal n representing the orientation of an infinitesimal
surface element having P as an interior point. The totality of all possible pairs
of the vectors t. and n at P defines the state of stress at that point. Fortunately,
it is not necessary to specify every pair of stress and normal vectors to describe
completely the state of stress at a given point. This may be accomplished instead
by merely specifying the stress vector on each of three mutually orthogonal planes
at P. Vector component relations then serve to relate the stress vector on any
other plane at the point to the original three. Thus, from a force balance on a
small tetrahedral volume having its vertex at P, a second-order stress tensor T
can be defined such that

(9.1.1)

or, in indicial notation, representing t. by its components ti ,

(9.1.2)

The components Til' T22 , and T33 perpendicular to the planes are the normal
stresses; those acting in the planes are shear stresses. The latter include all the
remaining stresses: T\2, T13 , T210 T23 , T310 and T32
422 A balance of energy for a fixed mass of variable shape, velocity, position,
and temperature can be written as the following integral relationship, equating
Chapter Thirteen
the rate of increase of energy of the mass to the rate at which work is done on
the mass by surface and body stresses less the rate at which heat is removed plus
the rate at which electromagnetic energy is transmitted in:

~J
dt 9'1
p(V+V2)dV=J tn.VdS+J pgvdV
2 9'1 9'1

-r Ja9'1
q. 0 dS - r
Ja9'1
o (E' x H') dS (9.1.3)

Here d j dt = aj t + V V is the convective derivative. The last term in equation


(9.1.3) expresses the influx of field-transmitted energy via the Poynting vector
(see Section 13.4.3). The surface is moving with the local velocity of the fluid,
so the electric and magnetic field vectors must be evaluated in the reference frame
of the fluid; that is the meaning of the primed variables. It is shown in electromag-
netism texts that motion induces fields so that E' = E + v x Band H' = H - v x D
at ordinary velocities. In the absence of an applied electric field, E == D = 0 and
so E' x H' = (v x B) x H. Also, V is internal energy per unit mass as introduced
previously, v is the vector velocity having magnitude v, g is gravitational acceler-
ation, and q is the heat flux. The stress vector t represents the sum of pressure,
viscous, and electromagnetic contributions.
The objective in this section is to illustrate the methodology for a case that
is simple enough to discuss in a short space, yet sufficiently complex to yield an
interesting result.
It can be seen that equation (9.1.3) does not have the electromagnetic work
terms appearing in the Gibbs equation (6.19); instead, the electromagnetic work
is formulated in terms of the displacement of a surface on which stress acts.
Accordingly, one objective of this treatment is to determine the form of the
electromagnetic stress tensor that makes these formulations compatible with one
another.
It is desired to transform equation (9.1.3) to a differential equation, but first
each surface integral in it must be expressed as a volume integral. Application
of equation (9.1.1) and Gauss's theorem gives

fa9'1
(tn v) dS = f
a9'1
(0 T v) dS = f a9'1
(T v) 0 dS

= L V (Tv) dV (9.1.4)

With the identity


V(Tv) = (VT)v+T:Vv (9.1.5)

the result (9.1.4) can be expressed as

fa9'1 (tn.V)dS=f9'1 [(VT)v+T:Vv]dV (9.1.6)


For the heat removal term it follows from Gauss's theorem that 423
Thermodynamics of
f q'O dS = f (Vq) dV (9.1.7) Electromagnetism
Jam Jm
Likewise, the term in the Poynting flux of energy transforms to

-L Vx[(vxB)xH]dV= L H[Vx(vxB)]dV

because V x H = 0 in a nonconductive fluid. From Reynolds's transport


theorem(3S) the terms on the left-hand side of equation (9.1.3) may be transformed
as follows:

-d
dt m
ipUdV= dU
p-dV
m dt
f (9.1.8)

-d
dt
im 2
v .
p-dV=
2
fm 2
d(v /2)
p---dV=
dt
fm dv
pv-dV
dt
(9.1.9)

When the results (9.1.6)-(9.1.9) are substituted into equation (9.1.3) and the
arbitrariness of the volume of integration is invoked, the following equation is
obtained after minor rearrangement:

dU ( p--VT-pg
p-+ dv ) v=T:Vv-Vq-H[Vx(vxB)] (9.1.10)
dT dt

This differential equation for the balance of energy is not yet expressed in its
simplest form. To proceed further the momentum balance for the chosen mass
of fluid will be formulated. Applying Newton's law of motion for a continuum,
the integral momentum balance reads

f pv dV =
ddtJm f t dS + f pg dV (9.1.11)
Jam Jm
which transforms to
dv
p-=VT+pg (9.1.11)
dt

Inspection of this equation shows that the bracketed terms on the left-hand side
of equation (9.1.10) disappear, leaving the following differential equation
expressing the balance of internal energy:

dU .
Pdi= T:Vv-Vq-H[Vx(vxB)] (9.1.13)

It is noteworthy that the kinetic energy and body force terms have dropped out
of the equation, a consequence of subtracting the mechanical energy terms.
424 Dividing each term of the Gibbs equation (6.19) by dt and omitting field
terms other than the magnetic ones gives, with slight rearrangement,
Chapter Thirteen

dS dU (JLOH2
pT-=p-- p*----JL MH ) -dp- HdB
- (9.1.14)
dt dt 2 0 p dt dt

This Gibbs equation provides a means for computing entropy rate of change
when all other terms including dU / dT are known. The relationship (9.1.13),
which also contains dU / dT, is derived with reference to an arbitrary irreversible
process. Assuming that both equations are valid, it is possible to eliminate the
term p dU / dt between them and obtain an equation for the entropy production
rate. As an adjunct to carrying out this step the equation of mass conservation
is needed. In differential form, valid for a compressible medium, it reads

ap
-+V(pv)=O (9.1.15)
at
Because V(pv) = vVp + pVv, equation (9.1.15) can be rewritten as

dp
- = -Vv= -1:Vv (9.1.16)
pdt

where, as previously, d/ dt is the convective derivative and I = ii + jj + kk is the


unit dyadic. .

EXAMPLE 13.9.1. Verification of the Tensor Identity V v = I: Vv


The repeated product is defined according to the nesting convention (a b): (c d) =
a(bc)d. Thus

ii:Vv
=j.(iV)v

= avx
ax

Likewise (j j): Vv = av.,/ ax and (k k): (Vv) = avz / az. Summing the separate terms yields

(QED)

As another preliminary, the last term H dB/ dt in equation (9.1.14) will be


transformed. It is helpful in doing this to rewrite H dB/ dt as the scalar vector
product H dB/ dt which is permissible when Hand B are collinear. Now dB/ dt
refers to the time derivative of induction within a mass of fixed identity.
Accordingly, 425
dB ilB Thermodynamics of
-=-+(v'V)B (9.1.17) Electromagnetism
dt ilt

where v, as previously, is the velocity of the mass at the point in question. Using
the vector identity

v x (X x Y) = X(V . Y) - Y(V X) + Y VX - X' VY (9.1.18)

and substituting v for X and B for Y gives, with rearrangement,

V VB = -B(V 'v) + (BV)v - V x (v x B) (9.1.19)

where it was recognized that V B = 0 from Maxwell's equation. Combining


equations (9.1.17) and (9.1.19) then yields

dB ilB
- = - - B(V 'v) + (B V)v - V x (v x B) (9.1.20)
dt ilt

If it is assumed that B is temporally constant, then the first term on the right-hand
side of this equation disppears. Thus

dB
H- = -BH(V 'v) + H(BV)v - HV x (v x B)
dt
= -BHI:Vv+ HB:Vv- HV x (v x B)
= (BH - BHI):Vv- HV x (v x B) (9.1.21)

In the last line the substitution HB = H(JLH) = BH was introduced on the


assumption that Hand B are collinear. Substituting from equation (9.1.21) for
H dBI dt in the Gibbs equation (9.1.14), from expression (9.1.16) for dpi pdt,
and eliminating p dU I dt with the aid of the internal energy balance (9.1.13)
yields a compact equation expressing the rate of entropy production, namely

dS
pT- = cI> - Vq (9.1.22)
dt

where cI> is the dissipation function per unit volume given by

cI> = Tv:Vv (9.1.23)


and
Tv = T + Pol - Tm (9.1.24)

with Tm given by

Tm = -{JLO LH[il(~V)L.T dH +!JLOH2}1+ BH (9.1.25)


426 where substitution was made for p* from its definition (6.17). Because Tm alone
encompasses magnetic parameters, it evidently represents the magnetic stress
Chapter Thirteen
tensor in the presence of matter. Quantity Tm is the generalization to nonlinearly
magnetizable media of the force density of Korteweg and Helmholtz, previously
derived(1o,lI) and utilized extensively(5) in the study of magnetic fluid response
to applied magnetic field. t
The work done by pressure stress is nondissipative. This is illustrated by
rewriting the entropy production term in equation (9.1.23) in the form pol: Vv =
PoVv = -(Pol p) dpi dt, where equation (9.1.16) is employed to introduce the
mass density. When the process is run backward the sign of dpi dt reverses, and
accordingly the entropy change dSI dt reverses. For the same reason, the term
with I in expression (9.1.25) for Tm also yields reversible change. The remaining
part ofTmcontributes the term DU: Vv, which can be indicated in indicial notation
as HjBj avdaxj ; when the direction of flow is reversed, avdaxj changes sign while
HjBj is unchanged in sign. Thus, reversibility is displayed also for this term and
therefore for the total working of the magnetic stress.
This reversible nature of the magnetic stress stems from the model assumption
that magnetization M is collinear with magnetic field U at all times, or, stated
more broadly, relaxation is instantaneous, i.e., hysteresis is absent. The form of
the magnetic work density H d (B I p) employed in deriving Tm was developed
on the collinearity assumption. It was noted in Section 13.5 that retracing the
development of the work density, it can easily be shown that the more general
form is U d (DI p) when U and D are oblique. Thus, a tensor permeability II may
exist such that D = II" U. Relaxation processes in colloidal magnetic fluids produce
values of II that are time dependent; aspects of the irreversibilities accompanying
relaxation are discussed further below.
Quantity Tv given by expression (9.1.24) is the total stress T less the pressure
and the magnetic stress. The only stress acting in this system other than pressure
or magnetic stress by assumption is viscous stress, and so the tensor Tv must
represent the viscous stress. From equation (9.1.22) and the ensuing discussion
it is seen that entropy is generated solely through the action of the viscous stress
and heat flow. Thus, magnetic stress in this model with instantaneous relaxation
cannot directly generate entropy.
As a comment on notation, the ordinary pressure Po(p, T) appearing in
expression (9.1.24) can be eliminated using equation (6.17) or (6.18) giving
Pol - Tm = (p* + lLoH2/2)1 - BU = pi - BU as alternative forms. No one
pressure Po, p*, or p is more fundamental than the other and the choice of one
or other in a particular circumstance can be made for convenience.

Constitutive Relationships
As a general statement, the rate of entropy increase in a mass offixed identity
equals the rate of internal production of entropy plus the rate at which entropy is

t The stress balance at a planar interface between media using expression (9.1.25) is found to be
given by continuity of p* + !LoM~/2, where Mn is the normal component of magnetization evaluated
adjacent to the interface, in agreement with the mechanical equilibrium criteria of Section 13.7 (see
Reference 5, p. 129).
transported into the mass. This balance statement is expressed mathematically as 427
(see notation of Figure 13.10).
Thermodynamics of
Electromagnetism
~
dt
r pSdV= J9'r1 6dV+ Ja9'r 1 j,'DdS
J9'1
(9.1.26)

in which 6 is the internal rate of entropy production per unit volume and j, is
the entropy flux vector. Applying the Reynolds transport theorem, transforming
the last term using Gauss's divergence theorem, and noting the arbitrariness of
the volume chosen for integration, we obtain the differential equation equivalent
to equation (9.1.26),
dS
6=p--V'J (9.1.27)
dt '

When, as in the present circumstances, the only mechanism for transporting


entropy into the volume is via heat transfer, the entropy flux is given by j, = -q/ T
where q is the heat flux vector and T the local temperature. Thus

dS
6 = p-+ V(q/T) (9.1.28)
dt

If each term is multiplied by temperature T, the quantity pTdS/dt substituted


from equation (9.1.22), and the terms simplified, we obtain

T6 = cI> + q' V In(1/ T) (9.1.29)

When the fundamental postulate of the second law is applied to a unit mass of
the moving fluid, it follows from the discussion in Section 13.7 that 6 is
nonnegative:

6~0 (9.1.30)

This statement of the second law is known as the Clausius-Duhem inequality.


Assuming that heat flow is dissipational in the presence of mechanical dissipation
permits equation (9.1.29) to be split into two inequalities:

q. V In(1/ T) ~ 0 (9.1.3la)

FIGURE 13.10. Integral control volume for deriving the entropy balance equation.
428 and
Chapter Thirteen (9.I.3Ib)

These relationships provide guidance in the selection of constitutive relationships


for q and Tv that close the equation set while ensuring that the second law remains
unviolated. The simplest choice in inequality (9.1.3la) is to assume a linear
relation between the flux q and the quantity V ln(ll T) = -(II T2)V T. This
produces the familiar Fourier law of heat conduction,

(9.1.32)

where the factor T- 2 is absorbed into K, the thermal conductivity. Thus,


substituting expression (9.1.32) back in relationship (9.1.3la) gives

K
T2 (VT) (VT) 2:: 0 (9.1.33)

Because (V T) (V T) = (V T)2 is inherently positive, as is temperature T, it can


be concluded that the Fourier coefficient K must be a positive quantity.
Assumption of a linear relationship between each of the nine components
of the viscous stress tensor Tv and the nine components of the velocity gradient
tensor Vv implies 81 constants of proportionality. Through arguments of
kinematics and isotropy these are reduced to two constants giving(34)

(9.1.34)

in which (VV)T is the transpose of Vv, TJ is the ordinary or first coefficient of


viscosity, and A the second coefficient of viscosity.
Substitution of expression (9.1.34) for Tv into dissipation function (9.1.23)
and assuming the fluid incompressible (V v = 0) yields the inherently positive-
valued expression for the dissipation function cP,

(9.1.35)

In indicial notation this is

(9.1.36)

A colloidal subdomain particle carrying the magnetic property in a ferrofluid


tends to align with shifting orientation of magnetic field in the manner of a
compass needle. Any resultant rotation of the particle is resisted by a couple due
to the surrounding viscous fluid and, accordingly, an additional mechanism exists
for energy dissipation. Concomitantly, Hand B are not collinear and magnetic
work density is given by the term H d (BI p). The thermodynamic methodology
has been extended to treat this case, providing a framework for describing the
dynamics and formulating the associated additional constitutive equations in a
series of papers by Shizawa and Tanahashi. (26)
For example, denoting the body couple per unit mass of fluid as G and the 429
angular rate of particle rotation as .., a rate of work is performed per unit volume
Thermodynamics of
equal to pG..,; over the volume within a fluid region :Pl, the total rate of work is Electromagnetism

L pG..,dV (9.1.37)

Similarly, it may be assumed that in addition to the body couple G a surface


couple c" acts upon unit area of the surface enclosing the volume; over the surface
the associated rate of work is

(9.1.38)

Again invoking the concept of Cauchy a couple stress tensor C is defined such
that c" = n' C. Then, following the same mathematical steps given in equation
(9.1.4), it can be seen that the surface integral transforms to a volume integral
given by

r
J'fJ1!
Cn ' ' ' ' dS = r V.(C",) dV
JfJ1!
(9.1.39)

These expressions of couple work give additional terms that must be added to
the right-hand side of equation (9.1.3) in treating the problem of fluid with
intrinsic angular momentum. A term modeling the rotational kinetic energy of
spinning particles is added to the accumulation term on the left-hand side.
The foregoing indicates an approach to this topic that yields a formulation
of the governing 'relationships complete with constitutive relationships for mag-
netic couple stress and intrinsic angular momentum conversion rate. (26) However,
issues concerning the role of magnetic relaxation remain to be clarified. Experi-
mental studies of related flow fields using ferrofluids are emerging.(27) Analyses
of a number of flow fields with couple stress are reviewed in the literature. (5.32)

EXAMPLE 13.9.2. Bottles Containing an Electromagnetic Field


Generating 'an electric or magnetic field invariably produces stresses on the sources
of the field. A particularly convenient tool for evaluating these stresses is furnished by
the field stress tensor. For a magnetic field the stress tensor Tm is given by expression
(9.1.25). In the absence of magnetic material M = 0, and it will be assumed that the
integral term on the right-hand side of equation (9.1.25) vanishes. Also, 8 = JLoH so that
Tm reduces to

(E.9.2.1)

This is the Maxwell vacuum stress tensor.


Referring to the wound cylindrical coil depicted in Figure 13.2 choose a thin control
volume that straddles the windings. One surface of the control volume is located within
the field and is oriented tangential to the field. The opposite surface lies immediately
outside the coil where the field is negligibly small. The field forces exerted over the surfaces
430 of the control volume are experienced as a net force that must be withstood by the coil.
Denoting 0 the unit normal on the surface, expression (E.9.2.1) gives for the stress vector
Chapter Thirteen
tn, recognizing that oH = 0,

{E.9.2.2}

The magnitude of tn is given as


J.o 2 B2
t =Ot =--H = - - {E.9.2.3}
n n 2 2J.o

Because tn is negative it is directed opposite to 0, hence it represents a pressure stress on


the containing wall, i.e., on the windings. There is no field and hence no stress acting on
the outer surface of the control volume. Thus expression (E.9.2.3) represents the total
stress on the wall. The total force is the integrated product of stress and area over the
entire wall.
In the case of an electric field in free space,

{E.9.2.4}

Let us consider one plate of the electrode pair in Figure 13.1. A control volume that
straddles the upper electrode is free of field on its top surface, while a uniform field is
present over the lower surface and is oriented normal to the surface. We can imagine that
the electrode is slightly lifted away from the dielectric, with the lower surface of the control
volume located within that space; in this case the free space form of To given as expression
(E.9.2.4) is applicable. Because the unit normal 0 on the lower surface is parallel to the
field, the stress vector tn is obtained in the form

{E.9.2.5}

and its magnitude tn = o tn is given by

(E.9.2.6)

Here tn is positive, giving a tensile stress acting to pull the electrode toward the dielectric
slab. The slab in tum is under compression.
In both these cases the field source serves as a "bottle" which must support a stress
to "contain" the field within it. The magnitude of these stresses corresponds identically
to the expressions for field energy density developed in Section 13.4. Because a stress level
of one atmosphere equals 1.013 N m-2 , the values of Av in Table 13.3 can immediately
be converted (except in one case) to equivalent stress in atmospheres or psi; see Table 13.6.
The exceptional case is that of the ruby mica capacitor. The energy density Av listed
applies for the dielectric and not the thin air gap. Because the displacement .Q is contino
uous, BoEg = BEd where subscript g denotes gap, d dielectric, and thus (Av)gap =
(Av)dioloeuie X B/ Bo.
The very large magnitude of the wall stress for the future superconductor poses a
formidable challenge to the designer. One investigator employs a series of discrete, coaxial
TABLE 13.6 431
Computed Values of the Electromagnetic Wall Stress (Systems of Table 13.3)
Thermodynamics of
Wall stress Electromagnetism

Field System A v (Jm- 3 ) atm psi

Electric Air gap capacitor 2.83 2.79 x 10-' 4.1 X to-'


Electric Ruby mica capacitor 2.10 x 10' 1.12 1.65 x 10"
Magnetic Air gap of iron yoke magnet 1.59 x 10 1.57 x 10 2.30 x 10"
Magnetic Present superconductive magnet 1.43 x 107 1.41 X 102 2.08 x 10'
Magnetic Future superconductor 3.98 x 10 3.93 x 10" 5.78 x 10'

cylindrical windings said to limit the bursting stress in any cylindrical layer; using ordinary
conductors and pulse operation to limit the temperature rise, a magnetic field of 100 tesla
is generated nondestructively.(29) In principle, a high-temperature superconducting magnet
of this design could operate continuously.

13.9.2. Mass Transfer

In Section 13.9.1 the Gibbs equation for a homogeneous fluid was manipu-
lated to obtain an equation for the rate of entropy production, yielding along
the way the magnetic stress tensor and, with assertion of the second law, allowable
forms of constitutive relationships for viscous stress and heat flow. This section
introduces the Gibbs equation for a fluid mixture and employs the corresponding
strategy to derive relationships for diffusion rates of magnetizable species in a
mixture subjected to a magnetic field. Standard treatments(S) of diffusion in
external force fields are inappropriate to this situation in which the component
electromagnetic force densities are unknown a priori. '
From the Gibbs equation (6.23) applicable to mixtures, the rate of change
of entropy per unit mass can be expressed as

(9.2.1)

Introducing the definition of p* from equation (6.18) and the defining equation
B = ILo(O + M), the above equation can be expressed as

dS dU ILoH2(*) dp dB s dx;
pT-=p-- p ----ILoHM - - H - - L I L J J - (9.2.2)
dt dt 2 pdt dt ;=1 dt

This relationship for entropy production rate extends that of equation (9.1.14)
with the addition of the last term in equation (9.2.2) giving the contribution of
composition rate of change.
An equation for conservation of species i can be developed in reference to
an integral balance on a control volume whose bounding surface moves at the
local mass average velocity v. The mass rate of accumulation of species j within
432 the volume is (d/ dt) 1PXi dV and equals the transport rate of the species into
the volume by the process of diftusion across the bounding surface. The latter is
Chapter Thirteen
given by -1 JiB dS where J i is the mass flux vector of species i relative to the
plane of no net mass flux,

Ji = (Vi - V)PXi (9.2.3)

The mass average velocity V is defined in terms of the species velocities Vi and
mass fractions Xi,
s
v= L
;=1
ViXi (9.2.4)

Applying the Reynolds transport theorem to the volume integral and Gauss's
divergence theorem to the surface integral, the arbitrariness of the volume gives
the species conservation equation as

(9.2.5)

Expressions for dU/dt, dp/dt, and dB/dt were developed as equations


(9.1.13), (9.1.16), and (9.1.21), respectively. For mixtures, the heat flux appearing
in equation (9.1.13) is augmented with a diffusive flux ofintemal energy. Denoting
the combined flux q', substitution of these relationships and the expression for
dxd dt of equation (9.2.5) into relationship (9.2.2) yields the production rate of
entropy in the form
dS s
pT- = 111- V .q' + L lI-iV Ji (9.2.6)
dt i=1

where III is the dissipation function defined by expression (9.1.23). Again it is


desired to manipulate the entropy production rate into the form of the entropy
balance. From the vector identity V (yX) = yV-' X + X' Vy the flux terms in
equation (9.2.6) are expanded as

(9.2.7)

and

11--
-~VJ- =J-'V ~ (11--) -V, (11--)
~J- (9.2.8)
T ' 'T T'

With these expressions the entropy increase rate in equation (9.2.6) can be written
in desired form as

dS
p-= -V
dt T
1 , (-1) -
[q, - L JilLi] +-Tv:Vv+q'V
T T
Ls
;=1
J;'V (11-1)
-
T
(9.2.9)
This equation can be compared to the form of the entropy balance equation 433
(9.1.27)
Thermodynamics of
Electromagnetism
dS V
dt= 9+ 'J s
P- (9.2.10)

from which we identify the entropy flux is and internal entropy production rate
8 as

(9.2.11)

and

1 , ( -1 ) - ~
9 =-Tv:Vv+q'V L.. J,'V (IL')
- (9.2.12)
T T '=1 T

Alternatively, the internal entropy production rate can be expressed in the form
s
T9 =Tv:Vv-isVT- 1:
i=l
J"VILi (9.2.13)

This expression displays the characteristic form with sums of the products of
fluxes and their driving forces on the right-hand side. The fluxes are respectively
the momentum flux Tv, the entropy flux is, and the mass fluxes of the s species
J,.

Binary Diffusion in Magnetic Field


Let us consider isothermal diffusion of species with negligible viscous dissipa-
tion in the presence of a nonuniform magnetic field. From equation (9.2.13) the
expression for entropy production reduces to

T8 = - f
i=1
Ji,VIL, (9.2.14)

For the binary mixture,

(9.2.15)

By their definition the mass fluxes J 1 and J 2 are not independent of each other.
From expressions (9.2.3) and (9.2.4), J 1 + J 1 = 0 and so

(9.2.16)

This expression for entropy production rate has the same appearance whether
or not a field is applied. However, because the chemical potentials are field
dependent, the influence of a field is indeed inherent in the expression.
434 By referring to equation (P.8.2.4) it can be seen that chemical potential ILl
depends on temperature T, mol fraction c;, magnetic field H, and somewhat on
Chapter Thirteen
total pressure, as the latter affects the vapor pressure of a pure species. Neglecting
the pressure effect, the gradient V(ILI - IL2) can be expanded for an isothermal
system to give

(9.2.17)

From the cited expression for chemical potential of a liquid mixture, the
coefficients in the above expansion are calculated to be

(9.2.18)

and

(9.2.19)

where ""-j is the molecular weight of species i and J is the partial molar moment.
j

From equation (9.2.16) the form of the linear constitutive law guaranteeing that
(1 2: 0 is

(9.2.20)

where L is a factor of proportionality. Thus

(9.2.21)

Mol fraction is related to mass fraction by

(9.2.22)

(9.2.23)

and

( 1 + 1)
""-ICI ""-2 CZ =
J1
XIX2
(9.2.24)

Substitution into equation (9.2.21) to eliminate the mol fraction variables gives

(9.2.25)
In the absence of a magnetic field J I is conventionally represented as J I = 435
- Dp VXl> where D is defined as the binary diffusion coefficient. Thus, from
Thermodynamics of
relationship (9.2.25), with H = 0, L = Dp.J,{I.J,{2 ill RTxIX2, and elimination of L Electromagnetism
from equation (9.2.25) yields

(9.2.26)

which expresses the mass flux in terms of mass fraction concentration.


As a special case when the mixture is dilute (XI ~ 0, X 2 ~ 1) and only one
species is magnetizable (i 2 = 0), (9.2.25) reduces to

(9.2.27)

Setting the partial molar magnetic moment equal to that of the pure species
il =.11 yields a form employed by Blums et al.(9,33) in the study of colloidal
magnetophoresis.

EXAMPLE 13.9.3. Magnetic Separation of Molecules


A nondilute binary liquid mixture of paramagnetic molecular species having suscep-
tibilities XI = M,/ H, X2 = M2/ H, and nearly equal molecular weights Atl = At2 = At and
mass densities PI = P2 = p, is contained in a tube positioned in a gradient magnetic field
H having maximum intensity Hm at one end and negligible intensity at the other. It is
desired to determine the equilibrium composition distribution.
At equilibrium the mass fiux disappears, so from (9.2.26) with J I = 0 we find

(E.9.3.1)

Assuming that the partial molar moment equals the molar moment of the pure species,
i, = .i, where .i, = M, V = X,HV and 'V = At/ P is the molar volume. Substitution into
equation (E.9.3.1) and rearranging gives

(E.9.3.2)

Integration of the left-hand side between positions at the two ends of the tube gives

(E.9.3.3l

where a is the separation factor defined by

(E.9.3.4)
436 Thus, the separation factor is found from equation (E.9.3.2) in the form
Chapter Thirteen
(E.9.3.5)

Separation can occur only when there is a difference in the magnetic susceptibilities ..
Separation is favored by a large ratio of magnetic energy density (XI - X2)lL oH 2/2 to
thermal energy density RTI 'Y.

Retrospective
The chemical engineer may be called upon to design a process containing
polarizable matter which must operate in a region of strong magnetic or electric
fields. Although some of the thermodynamic effects are normally so small as to
be negligible, careful design practice requires that anticipated effects be evaluated
and not merely assumed to be small. This chapter presents analyses by which
several magnetic and electric effects may be evaluated. Shifts of equilibria in the
largest attainable magnetic fields exceed the sh~fts attainable in the largest electric
fields.
Present day engineering concepts in which fields do play an important role
include field energy storage systems, systems of heat and work interconversion,
fluid mechanics with ferro- and paramagnetic species, and processes of mass
transfer with polarizable species.
Recent advances in superconductivity enhance the interest in the phenomena
utilizing magnetic fields. Creative use of the basic principles and mechanisms
discussed can be expected to yield future innovations.

LITERATURE

The reader desiring a clearly written, brief introduction to electromagnetism is referred


to the monograph of Thomas and Meadows(12) which "explores the physical meaning of
the equations and brings out the relationship between this physical context and its
mathematical representation." The introductory text of Reitz et ai.(14) is sound and well
written. An introductory text recommended for its modem format and detailed problem
solving approach is that of ZahnY3) At a more advanced level the book of Stratton(15)
deservedly continues to be held in high regard. The text of Jackson(16) is widely used,
although the mixture of microscopic and macroscopic concepts may prove a distraction
to some.
The expressions for variable-volume electromagnetic work E d (Dip) and H d (DI p )
developed in the spirit of classical thermodynamics in Section 13.5 substantiate the results
of Chu(19) obtained on the basis of a relativistic representation for the stress tensor in a
time-varying electromagnetic field and other premises including an assumption for the
form of the electromagnetic momentum density. The form of the latter remains to be
established as discussed recently by Eu and Oppenheim.(20) In my view the expressions
E d (Dip) and H d (D/ p) are a litmus for testing the adequacy of the electromagnetic
momentum expression, and not vice versa.
Standard texts invariably express polarization work as (EI p). dD and (HI p). dD, or
an equivalent, and hence are limited in a strict sense to the treatment of constant density
processes (see, for example, the book by Modell and Reid(21)). The recent monograph of 437
Blums et al (9) is based on the form employed in this chapter which is presented there
with a literature citation to Chu and no discussion. Thermodynamics of
Ericksen(22) comments on the thermodynamical criteria recommended by Gibbs for Electromagnetism
analyzing stability of equilibrium explaining that the basic criterion for stability of such
equilibria is to be taken as the condition that the entropy should be a maximum, subject
to the constraint that the energy remain fixed; or equivalently for Gibbs that, for fixed
entropy, the energy is a minimum. Ericksen then comments: "As to the alleged equivalence,
he gives a plausibility argument; I cannot call it a proof." Then later: "Conceptually, we
must be comparing those stable eqUilibria with- something not so stable, and it is less than
clear what they ought to be compared with can be described so simply. Landau shook
the apple cart, a bit, convincing physicists, at least, that it was important to account for
other variables, involved in a significant way in bifurcations, what are called, after him,
order parameters."
An instance can be mentioned in which bifurcation of a magnetic system yields a
distinctly different static equilibrium state replacing another as magnetic field, the order
parameter, increases. Thus, linear stability theory predicts that uniform magnetic field
oriented perpendicular to a flat interface between magnetizable and nonmagnetizable fluid
destabilizes the interface when a critical value of magnetization is exceeded. (10) Experiments
validated the theory and showed that a static state is reached in which elevated peaks are
arranged on the surface, usually in a hexagonal periodic pattern. Boudouvis et al (23) using
nonlinear numerical computations predicted and experimentally observed first-order
excitation to finite amplitude peaks accompanied by hysteresis in the peak height.
Analyses of conditions for phase equilibria in fields could apparently be established
in a systematic manner using Legendre transformation as exemplified in Ref. 21.
Current research(18) is probing the influence of magnetic fields 0Ii. biological systems.
It is only in recent years that validated phenomena have been advanced.

REFERENCES

1. F. W. Camp and E. F. Johnson, Magnetic effects in certain systems of chemical engineering


interest, I&:EC Fundamentals, 4(2), 145-150 (1965).
2. A. J. Meachin and M. W. Biddulph, The effect of high magnetic fields on the vapour pressure
of nitrogen, oxygen and argon, Cryogenics, 18, 29-32 (1978).
3. K. Denbigh, The Principles of Chemical Equilibrium, Third Edition, Cambridge University Press,
London (1971).
4. F. B. Hildebrand, Methods of Applied Mathematics, Prentice-Hall, Englewood Cliffs, NJ (1952).
5. R. E. Rosensweig, Ferrohydrodynamics, Cambridge University Press, New York (1985).
6. V. V. Chekanov, On measuring pressure in ferrofluid, Magnetohydrodynamics, 13(4), 394-398
(1977).
7. K. Shizawa and T. Tanahashi, Thermodynamic discussions on the basic equations of conducting
magnetic fluids, BulL J5ME, 29(250), 1171-1176 (1986).
8. S. R. de Groot and P. Mazur, Non-EqUilibrium Thermodynamics, Dover Publications, New York
(1984). Originally published by North-Holland, Amsterdam (1962).
9. E. Blums, Yu. A. Mikhailov, and R. Ozols, Heat and Mass Transfer in MHD Flows, World
Scientific Pub!. Co., Singapore (1987). Distributed by Taylor and Francis Inc. (in USA).
10. M. D. Cowley and R. E. Rosensweig, The interfacial stability of a ferromagnetic fluid, J. Fluid
Mech., 30(4), 671-688 (1967).
11. P. Penfield and H. A. Haus, Electrodynamics of Moving Media, MIT Press, Cambridge,
Massachusetts (1967).
438 12. E. G. Thomas and A. J. Meadows, Maxwell's Equations and their Applications, Adam Hilger Ltd.,
Bristol and Boston (1985).
Chapter Thirteen 13. M. Zahn, Electromagnetic Field Theory, W'uey, New York (1979).
14. J. R. Reitz, F. J. Milford, and R. W. Christy, Foundations of Electromagnetic Theory, Addison-
Wesley, Reading, Massachusetts (1979).
15. J. W. Stratton, Electromagnetic Theory, McGraw-Hill, New York (1941).
16. J. D. Jackson, Classical Electrodynamics, Wiley, New York (1'975).
17. G. V. Brown, Magnetic stirling cycles-a new application for magnetic materials, IEEE Trans.
Magn. MAG-13(5), 1146-1148 (1977).
18. G. Maret, J. Kiepenheuer, and N. Boccara (eds.), Biophysical Effects of Steady Magnetic Fields,
Springer-Verlag, New York (1986).
19. B. T. Chu, Thermodynamics of electrically conducting ftuids, Phys. Fluids, 1(5), 473-484 (1959).
20. B. C. Eu and I. Oppenheim, On the Minkowski tensor and thermodynamics of media in an
electromagnetic field, Ph"ica, I36A, 233-254 (1986).
21. M. Modell and R. C. Reid, ThermodynamiCS and its Applications, 2nd edn., Prentice-Hall,
Englewood Cliffs, NJ (1983).
22. J. L. Ericksen, Thermodynamics and stability of equilibrium, Appendix G3, pp. 503-509, in:
C. Truesdell, Rational Thermodynamics, second edition, Springer-Verlag, New York (1984).
23. A. G. Boudouvis, J. L. Puchalla, L. E. Scriven, and R. E. Rosensweig, Normal field instability
and patterns in pools of ferroftuid, J. Magn. & Magn. Mater., 65, 307-310 (1987).
24. R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures on Physics, Addison-Wesley,
Reading, Massachusetts, sixth printing (1977). See Volume II.
25. E. A. Guggenheim, ThermodynamiCS, An Advanced Treatment for Chemists and Physicists, fifth
revised edition, North-Holland, Amsterdam (1967).
26. K. Shizawa and T. Tanahashi, A new complete set of basic equations for magnetic ftuids with
internal rotation, Bull. JSME, 28(243), 1942-1948 (1985); 19(255), 2878-2884 (1986).
27. S. Kamiyama and R. E. Rosensweig, Introduction to the magnetic ftuids bibliography, J. Magn.
& Magn. Mater., 65(2 & 3), 401-402; Magnetic ftuids bibliography, 403-439 (1987).
28. R. K. Lyon, Effect of strong electrical fields on the boiling points of some alcohols, Nature, 191,
1285-1286 (1971).
29. M. Date and A. Yamagishi, Generation and applications of non-destructive pulsed magnetic
field, IEEE Trans. Magn., MAG-13(5), 3257-3262 (1987).
30. R. L. Curl, Direct thermomagnetic splitting of water, Int. J. Hydrogen Energy, 4, 13-20 (1979).
31. J. Landstreet and J. R. P. Angel, Astrophys. J., 196, 918 (1975); M. A. Rudderman and P. G.
Sutherland, Possible origin of magnetic fields in neutron stars and magnetic white dwarfs, Nat.,
Phys. Sci., 146, 93 (December 10, 1973).
32. B. M. Berkovsky, A. N. Vislovich, and B. E. Kashevsky, Magnetic ftuid as a continuum with
internal degrees of freedom, IEEE Trans. Magn., MAG-16(2), 329-342 (1980).
33. E. Blums, J. Plavins, and A. Chukhrov, High-gradient magnetic separation of magnetic colloids
and suspensions, J. Magn. & Magn. Mater., 39,147-151 (1983).
34. Sir H. Jeffreys, Cartesian Tensors, Cambridge University Press (1957).
35. R. Aris, Vector, Tensors, and the Basic Equations of Fluid Mechanics, Prentice-Hall, Englewood
Cliffs, NJ (1962).
36. J. R. Melcher, Continuum Electromechanics, MIT Press, Cambridge, Massachusetts (1981).
37. S. Kaneto, The inftuence of magnetic field on chemical reactions, J. Chem. Soc. Ind. Jpn., 34,
Suppl. Binding, 133B-134B (1931).
INDEX

Absolute entropy, 214 Birefringence, 338


Absolute temperature, 58, 65 Block copolymer, 354
Acceleration wave, 152 Body, 13
Acoustics, 150 couples, 2, 18, 42, 390
Activity coefficients, 226 rigid, 13, 18
Admissibility, 65, 277 Brownian motion, 110, 334, 355
Adsorption, 260 Buffer, 314
Affinity, 66, 99, 105, 115, 121
distribution, 91 Canonical ensemble, 201
vector, 74 Capillarity, 165
Air conditioner, 27 Capillary pressure, 158,248
Aliphatic chain, 345 Carnot cycle, 58, 66, 188
Alternate current, 302 Chain conformation, 324
Amorphous state, 213 Chafge density, 376
Ampere's law, 372 Chemical potential, 80, 149, 174
Angular momentum, 1,42 distribution, 90
Anisotropy, 338 at interface, 163
Atmosphere, 177 Cholesteric crystal, 117
Atomic weight, 74 Clausius-Clapeyron equation, 107, 120
Avogadro number, 6 Clausius-Duhem inequality, 32, 65, 427
Axiom of cbemistry, 71 Clausius-Planck inequality, 32
Azeotrope, 250 Coion,299
Colligative properties, 235, 350
Baric derivatives, 47, 52 Composition space, 72, 109, 113
Benzene, 325 Compressibility, 124
BET equation, 261 Compressor, 189
Biological systems, 4, 76, 139, 190, 198, 239 Conductive flux, 176
Biophysics, 199 Conformational entropy, 147, 154,213,339 439
440 Confonnational isomers, 325 Electrical conductivity, 302
Contact angle, 159, 261 Electrical current, 193
Index Continuous azeotrope, 264 Electrical double layer, 300
Continuous mixtures, 87, 112, 128, 142 Electrical potential, 85, 301
Convective flux, 176 Electrical work, 16, 85
of energy, 179 Electrochemical cell, 61, 84
Corresponding states, 231 Electrochemical reactions, 314
Couple stress tensor, 429 Electrohydrodynamics, 420
Coupling, 5, 172, 187, 301 Electromagnetic phenomena, 4, 375
Cracking, 89 Electromagnetic units, 370
Critical point, 108, 231 Electromotive force, 85, 298
Crosslink density, 331 Electronic industry, 4
Crystal symmetry, 191 Electronic spin, 406
Crystallinity, 117 Electrostatic field, 369
classes, 173 Embryo, 251
flow induced, 149, 339 Enantiotropic system, 118
Cubic expansion coefficient, 215, 231 Energy dissipation, 24, 51, 144, 146
Curie theorem, 190, 193 Energy parameter, 348
Cyclic processes, 76, 199 End-to-end distance, 141, 326
Engineering science, 2
Damping, ISO Enlanglements, 333
Debye-Huckel model, 297 Entropic elasticity, 147, 153, 212, 329
Defonnation, 18, 38, 185 Entropy
Dew point, 248 density, 22
Diamagnetism, 406 flow, 180
Dielectric constant, 300, 375 of mixing, 224
Diffusional instability, 110 production, 192
Diffusion, 176, 190, 215 supply, 22
coefficient, 333 Enzyme catalysts, 199
velocities, 178 Epigenetic system, 199
Diffusivity tensor, 182 Equilibrium history, 145
Dilute solutions, 227 Equimolar reaction, 139
Dilute suspensions, 335 Equimolar counterdiffusion, 178
Dimensional analysis, 5 Ergodic description, 328
Discontinuity, 137, 151, 157 Erythrite, 173
Disorientation, 345 Euler equations, 150
Displacement field, 371 Eulerian mechanics, 18
Dissociation of electrolytes, 240, 293 Eutectic, 259
Distribution function, 87, 337 Excess entropy, 148
bicontinuous, 89 Excluded volume, 355
Doi-Edwards theory, 344 Extensive properties, 80
Dolomite, 173 Extent of reaction, 66, 72, 89, 135
Drag, 150, 181 External state variable, 38, 46
Dufour effect, 190 Extra stress, 141

Ehrenfest relations, 121 Fading memory, 136, 142, 338


Einstein relationship, 334 Faraday
Elastic dumbbell, 336 constant, 6, 84, 300
Elastic energy, 147, 184, 212 law, 372
Elastic moduli, 151, 213, 329 Fennentation, 199
Elastic spring, 60 Ferrohydrodynamics, 420
Elastic theory, 38, 58, 137, 145, 150 Fick's law, 182
Electric field, 371 Flame fronts, 151
Electrical capacity, 80 Flexible chain, 325
Electrical charge density, 300 Flexible string, 151
Flory-Huggins equation, 347 Hydrostatic distribution, 173 441
Flory temperature, 327 Hydrostatic pressure, 103
Flow birefringence, 173 Hyperbolic equations, 150, 182 Index
Flue gas desulfurization, 281 Hysteresis, 261
Fluid mechanics, 1
Fluxes, 192 Ideal gas, 211
Forces, 192 Ideal mixture, 221
Formation reaction, 82 Incompressibility, 2, 128, 183
Fourier's law, 172, 194 Infmite dilution, 228
Frame Intensity, 84, 301
invariance, 141 Interaction parameters, 225, 235, 351
of reference, 179 Interatomic potential, 329
Frechet differential, 94, 143 Interface, 103
Free convection, 187 Internal coordinate, 337
Free current density, 373 Internal state variable, 38, 49, 121
Free draining coil, 334 Inverse Langevin function, 329
Free volume, 346 Invertibility, 53, 65, 76
Freezing, 122 Isochoric deformation, 330
Friction coefficient, 333 Isoenthalpic process, 188
Fugacity, 217 Isoentropic compressibility, 150, 217
coefficient, 247 Isoentropic process, 188
Functional, 93 Isomerization, 99, 115, 128
Functional analysis, 90, 100, 135 Isothermal compressibility, 215
Functional derivative, 90, 92, 145
Jump, 140
Galvani potential, 316
Gas dynamics, 128, 241 Kernel,95
Gauss Kidney, 190
distribution, 327 Kinetic energy, 14
law, 372 Kinetics, 50, 75, 93, 124, 136, 139, 147, 190
theorem, 20 mass action, 99, 152
Generalized pressure, 42 tensor, 191
Generalized volume, 42 Kuhn segment, 326, 361
Genetic locus, 200
Gibbs-Duhem equation, 80, 91, 175, 395 Label of reaction, 92
Gibbs-Helmholtz equation, 242 Lagrangian form, 176
Glass, 108, 123 Lambda transition, 119
Globular proteins, 354 Laminar flow, 195
Good solvents, 327 Langmuir isotherm, 261
Gravity Laplace transform, 136
acceleration, 175 Latent heat, 57, 120
effects, 181 Lattice model, 345
forces, 173 Length scale, 175
Gypsum, 173,282 Lever rule, 105, 112
Liapounov function, 76, 100
Hamiltonian, 182, 201 Lift force, 150
Heat Limestone, 281
conduction, 16, 24, 122 Linear friction, 151
flux, 171, 179, 185,215 Linear momentum, 1
pump, 30 Liquid crystal, 4, 117, 355, 371
transfer, 126, 171 Liquid helium, 123, 370
Helmholtz-Korteweg theorem, 195 Liquid nitrogen, 371
Henry's law, 228, 350 Liquid oxygen, 407
Heterogeneous eqUilibria, 276 Local action, 21, 37
History, 142, 338 Local minima, 83, 116
442 Long-range interactions, 298 Oil wells, 176
Lyotropic system, 119, 355 Olive oil, 159
Index Orientation, 141
Magnetic energy, 411 Orientational correlation, 325
Magnetic field, 120, 370 Osmotic pressure, 238
Magnetic induction, 401
Magnetic moment, 392 Parabolic equations, 182
Magnetic separation, 435 Paramagnetism, 406
Magnetization, 370 Partial mass properties, 179
Magnetocaloric processes, 416 Partial molar properties, 79, 95
Magnetohydrodynamics, 420 Partition coefficient, 255
Magnetostatic field, 369 Peltier coulers, 316
Marangoni effect, 161 Penetration theory, 182
Margnles equation, 232 Perfect crystals, 218
Mass Permanence of atoms, 71, 95
conservation, 1 Permeability, 370
, transfer, 109, 113, 117, 173, 178 Permittivity, 371
Material point, 176 pH, 5, 306
Maxwell Phase, 103
equation, 140, 185, 369 eqUilibria, 109
expansion, 64 Phenomenological coefficients, 192
kinetic theory, 3, 5, 46, 140, 151, 179, 182, Phosphoric acid, 309
215, 236, 323, 333 pK,306
relations, 55 Poisson ratio, 332
Mechanical stability, 211 Polarizability, 152, 357, 396
Messenger RNA, 200 Polybutadiene, 325
Metabolic products, 76, 199 Polyethylene, 324
Metastable points, 11 1 Polyfunctional acids, 309
Metric, 94 Polymerization, 117
Microcrystalline solid, 173 Polymers, 3, 119, 139, 146, 173, 186,213
Microscopic reversibility, 99, 203 blends, 351
Minimum entropy production, 193 crosslinked, 81
Miscibility gap, 258 mixing, 345
Mobility, 181 rigid, 354
tensor, 182 Polystyrene, 240, 324
Molar flux, 177 PopUlation genetics, 199
Molecular weight vector, 174 Positional entropy, 356
Momentum transfer, 183, 333 Potassium
Monodisperse polymers, 324 citrate, 318
Monolayer, 159, 261 nitrate, 299
Monotropic system, 11 8 Potential energy, 17
Morphology, 139, 186 Power law, 195
Multiphase system, 103 Predator-prey, 97, 100
Pressure transducer, 127
Navier-Stokes equations, 2, 151, 178 Propagation of discontinuities, 141, 149, 220, 241
Nematic crystal, 117
Nernst-Einstein equation, 302 Quartz, 123, 220
Network, 329 Quasistatic process, 62, 153
Neutron scattering, 327 Quasi-wave, 151
Newtonian fluid, 45, 141, 176, 215 Quenching, 108
Nitrogen oxide, 153
Norm, 94, 142 Radiant heat, 16, 172, 180
Nozzle valve, 188 Radius of curvature, 157
NRTL model, 242 Random walk, 324
Nucleation, 108, Ill, 117 selfavoiding, 327
Rank, 73, 89, 99 Stochastic process, 339
443
Raoult's law, 224 Stoichiometric coefficients, 114
Reaction Stoichiometry, 71 Index
pathways, 99 Stokes
subspace, 73, 77, 83, 99, 1l3, 271 hypothesis, 45, 161, 181
Redlich-Kwong equation, 231 law, 334
Redox reactions, 315 Strain, 42
Refrigeration cycle, 187 Stress
Refrigerator, 29 relaxation, 149
Relative adsorption, 163 tangential, 161, 183, 196
Relaxation, 4, 43, 135, 333 tensor, 1,41, 141, 186, 194
time, 138, 147, 185, 339, 358 Strong electrolytes, 293
Reptation, 341 Styrene, 240
Residual, 94 Subharmonic phenomena, 199
Retardation, 146 Sublimation, 107
Rigid dumbbell, 357 Superconductivity, 120
Rodlike polymers, 324 Supercooled vapor, 53, 108, 121
Rotational diffusivity, 357 Supercritical state, 108, 123
Rouse chain, 336 Supercritical theories, 119
Rouse-Zimm chain, 337 Superheated liquid, 53, 108, 121
Rubber, 63, 120, 125, 148, 153, 213, 327 Surface
crosslinked, 329 density, 163
engines, 33 layer, 158
swollen, 351 phenomena, 4
rate of strain, 160
Salt bridge, 317 tension, 103, 128, 157, 187
Semipermeable membrane, 239, 315, 350 tension tensor, 160
Sensible heat, 57 velocity, 160
Seveso accident, 127 viscosity, 128, 160
Shear Swelling, 261
deformation, 332 Symmetry relations, 191
motion, 161, 183
rate, 194 o Solvent, 335, 349
viscosity, 45, 141, 339 Talandic thermodynamics, 201
Shock wave, 151, 221, 241 Tangential stresses, 215
Short-range interactions, 298 Temperature gradient, 171
Silver iodide, 298 Tensile strength, 104
Single crystal, 173 Tensor analysis, 135
Site, 37, 60, 138 Tetrafunctional junctions, 352
Smectic crystal, 117 Tetrahedral structure, 325
Solid-solid transitions, 107, 149 Thermal engines, 15, 26
Solubility product, 299 Thermal conductivity, 172, 197
Sorel effect, 190 tensor, 172
Speed of sound, 65, 127, 149,217 Thermocapillary motion, 161, 187
Spinodal region, 104, 106, 110, 112, 121, 233, 253 Thermocouples, 316
Standard free enthalpy, 117 Thermoelectric effect, 203
Standard hydrogen electrode, 317 Thermostatics, 49
Standard state, 81, 296, 272 Thermotropic systems, 118
Static continuation, 145 Third law, 218
Statistical mechanics, 3, 141, 153, 199 Tie line, 116, 254
Steady state, 177, 195 Time scale, 136
Steam reforming, 282 Tobacco mosaic virus, 355
Stefan Topology, 94, 142, 341
hypothesis, 181 Transport phenomena, 202
problem, 122 Tree line, 202
Viscous theory, 137, 140, 147,
444 Triangular rule, 142
Triple point, 106, 119. 126 ISO
Index Turbulent flow. 195
Two-phase enve/ope, 116 Water hammer, 217
Water shift reaction, 96
UNIQUAC, 242 Weak electrolytes, 303
Weight ftaction. 177
Van der Waals equation, 128, 230 Wetting liquid, 159
Van Laar equation. 234 Wilson equation, 242
Vapor pressure, 107
Velocity gradient, I, 45, 184 X-ray diffraction, 117
Virial coefficients, 230
Viscoelasticity, 328. 336 Young modulus, 332
Viscosity, 2. 43, 119, 194, 333
intrinsic, 340, 358 Zinc, 284

Das könnte Ihnen auch gefallen