04.08.2013 Views

Thermodynamics - Physics at Oregon State University

Thermodynamics - Physics at Oregon State University

Thermodynamics - Physics at Oregon State University

SHOW MORE
SHOW LESS

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

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

<strong>Thermodynamics</strong><br />

Henri J.F. Jansen<br />

Department of <strong>Physics</strong><br />

<strong>Oregon</strong> St<strong>at</strong>e <strong>University</strong><br />

August 19, 2010


Contents<br />

PART I. <strong>Thermodynamics</strong> Fundamentals 1<br />

1 Basic <strong>Thermodynamics</strong>. 3<br />

1.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.2 Some definitions. . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.3 Zeroth Law of <strong>Thermodynamics</strong>. . . . . . . . . . . . . . . . . . 12<br />

1.4 First law: Energy. . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

1.5 Second law: Entropy. . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

1.6 Third law of thermodynamics. . . . . . . . . . . . . . . . . . . . . 31<br />

1.7 Ideal gas and temper<strong>at</strong>ure. . . . . . . . . . . . . . . . . . . . . . 32<br />

1.8 Extra questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

1.9 Problems for chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2 Thermodynamic potentials and 49<br />

2.1 Internal energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

2.2 Free energies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

2.3 Euler and Gibbs-Duhem rel<strong>at</strong>ions. . . . . . . . . . . . . . . . . . 61<br />

2.4 Maxwell rel<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

2.5 Response functions. . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

2.6 Rel<strong>at</strong>ions between partial deriv<strong>at</strong>ives. . . . . . . . . . . . . . . . . 68<br />

2.7 Conditions for equilibrium. . . . . . . . . . . . . . . . . . . . . . 72<br />

2.8 Stability requirements on other free energies. . . . . . . . . . . . 78<br />

2.9 A magnetic puzzle. . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

2.10 Role of fluctu<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

2.11 Extra questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

2.12 Problems for chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

3 Phase transitions. 107<br />

3.1 Phase diagrams. . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

3.2 Clausius-Clapeyron rel<strong>at</strong>ion. . . . . . . . . . . . . . . . . . . . . . 116<br />

3.3 Multi-phase boundaries. . . . . . . . . . . . . . . . . . . . . . . . 121<br />

3.4 Binary phase diagram. . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

3.5 Van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e. . . . . . . . . . . . . . . . . . . 126<br />

3.6 Spinodal decomposition. . . . . . . . . . . . . . . . . . . . . . . . 138<br />

III


IV CONTENTS<br />

3.7 Generaliz<strong>at</strong>ions of the van der Waals equ<strong>at</strong>ion. . . . . . . . . . . 142<br />

3.8 Extra questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

3.9 Problems for chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . 144<br />

PART II. <strong>Thermodynamics</strong> Advanced Topics 152<br />

4 Landau-Ginzburg theory. 153<br />

4.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154<br />

4.2 Order parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . 156<br />

4.3 Landau theory of phase transitions. . . . . . . . . . . . . . . . . . 162<br />

4.4 Case one: a second order phase transition. . . . . . . . . . . . . . 164<br />

4.5 Order of a phase transition. . . . . . . . . . . . . . . . . . . . . . 169<br />

4.6 Second case: first order transition. . . . . . . . . . . . . . . . . . 170<br />

4.7 Fluctu<strong>at</strong>ions and Ginzburg-Landau theory. . . . . . . . . . . . . 176<br />

4.8 Extra questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 183<br />

4.9 Problems for chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . 184<br />

5 Critical exponents. 193<br />

5.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194<br />

5.2 Mean field theory. . . . . . . . . . . . . . . . . . . . . . . . . . . 200<br />

5.3 Model free energy near a critical point. . . . . . . . . . . . . . . . 208<br />

5.4 Consequences of scaling. . . . . . . . . . . . . . . . . . . . . . . . 211<br />

5.5 Scaling of the pair correl<strong>at</strong>ion function. . . . . . . . . . . . . . . 217<br />

5.6 Hyper-scaling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218<br />

5.7 Validity of Ginzburg-Landau theory. . . . . . . . . . . . . . . . . 219<br />

5.8 Scaling of transport properties. . . . . . . . . . . . . . . . . . . . 221<br />

5.9 Extra questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 225<br />

5.10 Problems for chapter 5 . . . . . . . . . . . . . . . . . . . . . . . . 226<br />

6 Transport in <strong>Thermodynamics</strong>. 229<br />

6.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230<br />

6.2 Some thermo-electric phenomena. . . . . . . . . . . . . . . . . . . 233<br />

6.3 Non-equilibrium <strong>Thermodynamics</strong>. . . . . . . . . . . . . . . . . . 237<br />

6.4 Transport equ<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . . 240<br />

6.5 Macroscopic versus microscopic. . . . . . . . . . . . . . . . . . . . 245<br />

6.6 Thermo-electric effects. . . . . . . . . . . . . . . . . . . . . . . . 252<br />

6.7 Extra questions. . . . . . . . . . . . . . . . . . . . . . . . . . . . 262<br />

6.8 Problems for chapter 6 . . . . . . . . . . . . . . . . . . . . . . . . 262<br />

7 Correl<strong>at</strong>ion Functions. 265<br />

7.1 Description of correl<strong>at</strong>ed fluctu<strong>at</strong>ions. . . . . . . . . . . . . . . . 265<br />

7.2 M<strong>at</strong>hem<strong>at</strong>ical functions for correl<strong>at</strong>ions. . . . . . . . . . . . . . . 270<br />

7.3 Energy consider<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . 275<br />

PART III. Additional M<strong>at</strong>erial 279


CONTENTS V<br />

A Questions submitted by students. 281<br />

A.1 Questions for chapter 1. . . . . . . . . . . . . . . . . . . . . . . 281<br />

A.2 Questions for chapter 2. . . . . . . . . . . . . . . . . . . . . . . 284<br />

A.3 Questions for chapter 3. . . . . . . . . . . . . . . . . . . . . . . 287<br />

A.4 Questions for chapter 4. . . . . . . . . . . . . . . . . . . . . . . 290<br />

A.5 Questions for chapter 5. . . . . . . . . . . . . . . . . . . . . . . 292<br />

B Summaries submitted by students. 295<br />

B.1 Summaries for chapter 1. . . . . . . . . . . . . . . . . . . . . . 295<br />

B.2 Summaries for chapter 2. . . . . . . . . . . . . . . . . . . . . . 297<br />

B.3 Summaries for chapter 3. . . . . . . . . . . . . . . . . . . . . . 299<br />

B.4 Summaries for chapter 4. . . . . . . . . . . . . . . . . . . . . . 300<br />

B.5 Summaries for chapter 5. . . . . . . . . . . . . . . . . . . . . . 302<br />

C Solutions to selected problems. 305<br />

C.1 Solutions for chapter 1. . . . . . . . . . . . . . . . . . . . . . . 305<br />

C.2 Solutions for chapter 2. . . . . . . . . . . . . . . . . . . . . . . 321<br />

C.3 Solutions for chapter 3. . . . . . . . . . . . . . . . . . . . . . . 343<br />

C.4 Solutions for chapter 4. . . . . . . . . . . . . . . . . . . . . . . 353


VI CONTENTS


List of Figures<br />

1.1 Carnot cycle in PV diagram . . . . . . . . . . . . . . . . . . . 20<br />

1.2 Schem<strong>at</strong>ics of a Carnot engine. . . . . . . . . . . . . . . . . . 21<br />

1.3 Two engines feeding eachother. . . . . . . . . . . . . . . . . . 22<br />

1.4 Two Carnot engines in series. . . . . . . . . . . . . . . . . . . 25<br />

2.1 Container with piston as internal divider. . . . . . . . . . . . 53<br />

2.2 Container where the internal degree of freedom becomes external<br />

and hence can do work. . . . . . . . . . . . . . . . . . . 54<br />

3.1 Model phase diagram for a simple model system. . . . . . . . 111<br />

3.2 Phase diagram for solid Ce. . . . . . . . . . . . . . . . . . . . . 112<br />

3.3 Gibbs energy across the phase boundary <strong>at</strong> constant temper<strong>at</strong>ure,<br />

wrong picture. . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

3.4 Gibbs energy across the phase boundary <strong>at</strong> constant temper<strong>at</strong>ure,<br />

correct picture. . . . . . . . . . . . . . . . . . . . . . . . 114<br />

3.5 Volume versus pressure <strong>at</strong> the phase transition. . . . . . . . . 114<br />

3.6 Model phase diagram for a simple model system in V-T space.115<br />

3.7 Model phase diagram for a simple model system in p-V space.116<br />

3.8 Gibbs energy across the phase boundary <strong>at</strong> constant temper<strong>at</strong>ure<br />

for both phases. . . . . . . . . . . . . . . . . . . . . . . . 118<br />

3.9 Typical binary phase diagram with regions L=liquid, A(B)=B<br />

dissolved in A, and B(A)=A dissolved in B. . . . . . . . . . . 124<br />

3.10 Solidific<strong>at</strong>ion in a reversible process. . . . . . . . . . . . . . . 125<br />

3.11 Typical binary phase diagram with intermedi<strong>at</strong>e compound<br />

AB, with regions L=liquid, A(B)=B dissolved in A, B(A)=A<br />

dissolved in B, and AB= AB with either A or B dissolved. . . 126<br />

3.12 Typical binary phase diagram with intermedi<strong>at</strong>e compound<br />

AB, where the intermedi<strong>at</strong>e region is too small to discern<br />

from a line. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

3.13 Impossible binary phase diagram with intermedi<strong>at</strong>e compound<br />

AB, where the intermedi<strong>at</strong>e region is too small to discern<br />

from a line. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

3.14 Graphical solution of equ<strong>at</strong>ion 3.24. . . . . . . . . . . . . . . . 131<br />

3.15 p-V curves in the van der Waals model. . . . . . . . . . . . . . 133<br />

VII


VIII LIST OF FIGURES<br />

3.16 p-V curves in the van der Waals model with neg<strong>at</strong>ive values<br />

of the pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

3.17 p-V curve in the van der Waals model with areas corresponding<br />

to energies. . . . . . . . . . . . . . . . . . . . . . . . . . . . 135<br />

3.18 Unstable and meta-stable regions in the van der Waals p-V<br />

diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138<br />

3.19 Energy versus volume showing th<strong>at</strong> decomposition lowers the<br />

energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

4.1 He<strong>at</strong> capacity across the phase transition in the van der Waals<br />

model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155<br />

4.2 He<strong>at</strong> capacity across the phase transition in an experiment. . 155<br />

4.3 Continuity of phase transition around critical point in p-T<br />

plane. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158<br />

4.4 Continuity of phase around singular point. . . . . . . . . . . . 159<br />

4.5 Continuity of phase transition around critical point in H-T<br />

plane. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

4.6 Magnetiz<strong>at</strong>ion versus temper<strong>at</strong>ure. . . . . . . . . . . . . . . . . 165<br />

4.7 Forms of the Helmholtz free energy. . . . . . . . . . . . . . . . 166<br />

4.8 Entropy near the critical temper<strong>at</strong>ure. . . . . . . . . . . . . . 167<br />

4.9 Specific he<strong>at</strong> near the critical temper<strong>at</strong>ure. . . . . . . . . . . . 168<br />

4.10 Magnetic susceptibility near the critical temper<strong>at</strong>ure. . . . . 169<br />

4.11 Three possible forms of the Helmholtz free energy in case 2. 171<br />

4.12 Values for m corresponding to a minimum in the free energy. 172<br />

4.13 Magnetiz<strong>at</strong>ion as a function of temper<strong>at</strong>ure. . . . . . . . . . . 173<br />

4.14 Hysteresis loop. . . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

4.15 Critical behavior in first order phase transition. . . . . . . . . 176<br />

6.1 Essential geometry of a thermocouple. . . . . . . . . . . . . . . 234<br />

C.1 m versus T-H . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363<br />

C.2 Figure 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366


INTRODUCTION IX<br />

Introduction.<br />

<strong>Thermodynamics</strong>??? Why? Wh<strong>at</strong>? How? When? Where? Many questions<br />

to ask, so we will start with the first one. A frequent argument against studying<br />

thermodynamics is th<strong>at</strong> we do not have to do this, since everything follows from<br />

st<strong>at</strong>istical mechanics. In principle, this is, of course, true. The argument, however,<br />

assumes th<strong>at</strong> we know the exact description of a system on the microscopic<br />

scale, and th<strong>at</strong> we can calcul<strong>at</strong>e the partition function. In practice, we can only<br />

calcul<strong>at</strong>e the partition function for a few simple cases, and in all other cases we<br />

need to make serious approxim<strong>at</strong>ions. This is where thermodynamics plays an<br />

invaluable role. In thermodynamics we derive basic equ<strong>at</strong>ions th<strong>at</strong> all systems<br />

have to obey, and we derive these equ<strong>at</strong>ions from a few basic principles. In<br />

this sense thermodynamics is a meta-theory, a theory of theories, very similar<br />

to wh<strong>at</strong> we see in a study of non-linear dynamics. <strong>Thermodynamics</strong> gives us<br />

a framework for the results derived in st<strong>at</strong>istical mechanics, and allows us to<br />

check if approxim<strong>at</strong>ions made in st<strong>at</strong>istical mechanical calcul<strong>at</strong>ions viol<strong>at</strong>e some<br />

of these basic results. For example, if the calcul<strong>at</strong>ed he<strong>at</strong> capacity in st<strong>at</strong>istical<br />

mechanics is neg<strong>at</strong>ive, we know we have a problem!<br />

There are some semantic issues with the words thermodynamics and st<strong>at</strong>istical<br />

mechanics. In the English speaking part of the world thermodynamics<br />

is often seen as a subset of the field of st<strong>at</strong>istical mechanics. In the German<br />

world it is often seen as an a different field from st<strong>at</strong>istical mechanics. I take<br />

the l<strong>at</strong>ter view. <strong>Thermodynamics</strong> is the field of physics describing thermal effects<br />

in m<strong>at</strong>ter in a manner which is independent of the microscopic details of<br />

the m<strong>at</strong>erial. St<strong>at</strong>istical mechanics starts <strong>at</strong> a microscopic model and derives<br />

conclusions for the macroscopic world, based on these microscopic details. In<br />

this course we discuss thermodynamics, we present equ<strong>at</strong>ions and conclusions<br />

which are independent of the microscopic details.<br />

<strong>Thermodynamics</strong> also gives us a language for the description of experimental<br />

results. It defines observable quantities, especially in the form of response<br />

functions. It gives the definitions of critical exponents and transport properties.<br />

It allows analyzing experimental d<strong>at</strong>a in the framework of simple models, like<br />

equ<strong>at</strong>ions of st<strong>at</strong>e. It provides a framework to organize experimental d<strong>at</strong>a. To<br />

say th<strong>at</strong> we do not need this is quite arrogant, and assumes th<strong>at</strong> if you cannot<br />

follow the (often very complic<strong>at</strong>ed) deriv<strong>at</strong>ions in st<strong>at</strong>istical mechanics you<br />

might as well give up. <strong>Thermodynamics</strong> is the meeting ground of experimenters<br />

and theorists. It gives the common language needed to connect experimental<br />

d<strong>at</strong>a and theoretical results.<br />

Classical mechanics has its limits of validity, and we need rel<strong>at</strong>ivity and/or<br />

quantum mechanics to extend the domain of this theory. <strong>Thermodynamics</strong> and<br />

st<strong>at</strong>istical mechanics do not have such a rel<strong>at</strong>ion, though, contrary to wh<strong>at</strong> people<br />

claim who believe th<strong>at</strong> we do not need thermodynamics. A prime example<br />

is the concept of entropy. Entropy is defined as a measurable quantity in ther-


X INTRODUCTION<br />

modynamics, and the definition relies both on the thermodynamic limit (a large<br />

system) and the existence of reservoirs (an even larger outside). We can also<br />

define entropy in st<strong>at</strong>istical mechanics, but purists will only call this an entropy<br />

analogue. It is a good one, though, and it reproduces many of the well known<br />

results. The st<strong>at</strong>istical mechanical definition of entropy can also be applied to<br />

very small systems, and to the whole universe. But problems arise if we now<br />

also want to apply the second law of thermodynamics in these cases. Small<br />

system obey equ<strong>at</strong>ions which are symmetric under time reversal, which contradicts<br />

the second law. W<strong>at</strong>ch out for Maxwell’s demons! On the large scale, the<br />

entropy of the universe is probably increasing (it is a very large system, and<br />

by definition isol<strong>at</strong>ed). But if the universe is in a well defined quantum st<strong>at</strong>e,<br />

the entropy is and remains zero! These are very interesting questions, but for<br />

a different course. Confusing paradoxes arise easily if one does not appreci<strong>at</strong>e<br />

th<strong>at</strong> thermodynamics is really a meta-theory, and when one applies concepts<br />

under wrong conditions.<br />

Another interesting question is the following. Do large systems obey the<br />

same equ<strong>at</strong>ions as small systems? Are there some new ingredients we need when<br />

we describe a large system? Can we simply scale up the microscopic models to<br />

arrive <strong>at</strong> the large scale, as is done in renormaliz<strong>at</strong>ion group theory? How does<br />

the arrow of time creep into the description when we go from the microscopic<br />

time reversible world to the macroscopic second law of thermodynamics? How<br />

do the large scale phenomena emerge from a microscopic description, and why<br />

do microscopic details become unimportant or remain observable? All good<br />

questions, but also for another course. Here we simply look <strong>at</strong> thermodynamics.<br />

And wh<strong>at</strong> if you disagree with wh<strong>at</strong> was said above? Keep reading nevertheless,<br />

because thermodynamics is also fun. Well, <strong>at</strong> least for me it is......<br />

The m<strong>at</strong>erial in front of you is not a textbook, nor is it an <strong>at</strong>tempt <strong>at</strong> a<br />

future textbook. There are many excellent textbooks on thermodynamics, and<br />

it is not very useful to add a new textbook of lower quality. Also, when you<br />

write a textbook you have to dot all the t-s and cross all the i-s, or something<br />

like th<strong>at</strong>. You get it, I am too lazy for th<strong>at</strong>. This set of notes is meant to be<br />

a tool to help you study the topic of thermodynamics. I have over the years<br />

collected the topics I found relevant, and working through these notes will give<br />

you a good basic understanding of wh<strong>at</strong> is needed in general. If any important<br />

topic is missing, I would like to know so I can add it. If you find a topic too far<br />

out, so be it. All mistakes in these notes are mine. If something is quite useful,<br />

it is stolen from somewhere else.<br />

You can simply take these notes and read them. After doing so, you will<br />

<strong>at</strong> least have seen the basic concepts, and be able to recognize them in the<br />

liter<strong>at</strong>ure. But a much better approach is to read these notes and use them as a<br />

start for further study. This could mean going to the library and looking up these<br />

topics in a number of books on thermodynamics. Nothing helps understanding<br />

more than seeing different descriptions of the same m<strong>at</strong>erial. If there is one skill<br />

th<strong>at</strong> is currently missing among many students, it is the capability of really<br />

using a library! Also, I do not want to give you examples of wh<strong>at</strong> I consider<br />

good textbooks. You should go find out. My opinion would only be a single


INTRODUCTION XI<br />

biased opinion anyhow.<br />

These notes started when I summarized discussions in class. In the current<br />

form, I have presented them as reading m<strong>at</strong>erial, to start class discussions.<br />

<strong>Thermodynamics</strong> can be taught easily in a non-lecture approach, and I am<br />

working on including more questions which could be discussed in class (they are<br />

especially lacking in the l<strong>at</strong>er parts). Although students feel uneasy with this<br />

approach, having a fear th<strong>at</strong> they miss something important, they should realize<br />

th<strong>at</strong> the purpose of these notes is to make sure th<strong>at</strong> all important m<strong>at</strong>erial is<br />

in front of them. Class discussions, of course, have to be guided. Sometimes a<br />

discussion goes in the wrong direction. This is fine for a while, but than the<br />

instructor should help bring it back to the correct p<strong>at</strong>h. Of course, the analysis<br />

of why the discussion took a wrong turn is extremely valuable, because one<br />

learns most often from one’s mistakes (<strong>at</strong> least, one should). To be honest,<br />

finding the right balance for each new group remains a challenge.<br />

The m<strong>at</strong>erial in these notes is sufficient for a quarter or a semester course.<br />

In a semester course one simply adds expansions to selected topics. Also, the<br />

m<strong>at</strong>erial should be accessible for seniors and first year gradu<strong>at</strong>e students. The<br />

m<strong>at</strong>hem<strong>at</strong>ics involved is not too hard, but calculus with many partial deriv<strong>at</strong>ives<br />

is always a bit confusing for everybody, and functional deriv<strong>at</strong>ives also need a bit<br />

of review. It is assumed th<strong>at</strong> basic m<strong>at</strong>erial covered in the introductory physics<br />

sequence is known, hence students should have some idea about temper<strong>at</strong>ure<br />

and entropy. Apart from th<strong>at</strong>, visit the library and discover some lower level<br />

texts on thermodynamics. Again, there are many good ones. And, if these<br />

textbooks are more than ten years old, do not discard them, because they are<br />

still as relevant as they were before. On the other hand, if you use the web<br />

as a source of inform<strong>at</strong>ion, be aware th<strong>at</strong> there are many web-sites posted by<br />

well-meaning individuals, which are full of wrong inform<strong>at</strong>ion. Nevertheless,<br />

browsing the web is a good exercise, since nothing is more important than to<br />

learn to recognize which inform<strong>at</strong>ion is incorrect!<br />

Problem solving is very important in physics, and in order to obtain a working<br />

knowledge of thermodynamics it is important to be able to do problems.<br />

Many problems are included, most of them with solutions. It is good to start<br />

problems in class, and to have a discussion of the general approach th<strong>at</strong> needs<br />

to be taken. When solving problems, for most people it is very beneficial to<br />

work in groups, and th<strong>at</strong> is encouraged. When you try to solve a problem and<br />

you get stuck, do not look <strong>at</strong> the solution! Go to other textbooks and try to<br />

find m<strong>at</strong>erial th<strong>at</strong> pertains to your problem. When you believe th<strong>at</strong> you have<br />

found the solution, then it is time to compare with the solution in the notes,<br />

and then you can check if the solution in the notes is correct.<br />

In many cases, solving problems in thermodynamics always follows the same<br />

general p<strong>at</strong>h. First you identify the independent st<strong>at</strong>e variables. If an experiment<br />

is performed <strong>at</strong> constant temper<strong>at</strong>ure, temper<strong>at</strong>ure is an independent<br />

st<strong>at</strong>e variable because it is controlled. Control means th<strong>at</strong> either we can set it<br />

<strong>at</strong> a certain value, or th<strong>at</strong> we can prevent changes in the variable. For example,<br />

if we discuss a gas in a closed container, the volume of the gas is an independent<br />

st<strong>at</strong>e variable, since the presence of the container makes it impossible for the


XII INTRODUCTION<br />

gas to expand or contract. Pressure, on the other hand, is not an independent<br />

st<strong>at</strong>e variable in this example, since we have no means of controlling it. Second,<br />

based on our determin<strong>at</strong>ion of independent st<strong>at</strong>e variables, we select the correct<br />

thermodynamic potential to work with. Finally, we calcul<strong>at</strong>e the response<br />

functions using this potential, and find rel<strong>at</strong>ions between these functions. Or<br />

we use these response functions to construct equ<strong>at</strong>ions of st<strong>at</strong>e using measured<br />

d<strong>at</strong>a. And so on.<br />

Problem solving is, however, only a part of learning. Another part is to ask<br />

questions. Why do I think this m<strong>at</strong>erial is introduced <strong>at</strong> this point? Wh<strong>at</strong> is<br />

the relevance? How does it build on the previous m<strong>at</strong>erial? Sometimes these<br />

questions are subjective, because wh<strong>at</strong> is obvious for one person can be obscure<br />

for another. The detailed order of topics might work well for one person but not<br />

for another. Consequently, it is also important to ask questions about your own<br />

learning. How did I understand the m<strong>at</strong>erial? Which steps did I make? Which<br />

steps were in the notes, and which were not? How did I fill in the gaps? In<br />

summary, one could say th<strong>at</strong> problem solving improves technical skills, which<br />

leads to a better prepar<strong>at</strong>ion to apply the knowledge. Asking questions improves<br />

conceptual knowledge, and leads to a better understanding how to approach new<br />

situ<strong>at</strong>ions. Asking questions about learning improves the learning process itself<br />

and will facilit<strong>at</strong>e future learning, and also to the limits of the current subject.<br />

Work in progress is adding more questions in the main text. There are<br />

more in the beginning than in the end, a common phenomenon. As part of<br />

their assignments, I asked students in the beginning which questions they would<br />

introduce. These questions are collected in an appendix. So, one should not<br />

take these questions as questions from the students (although quite a few are),<br />

but also as questions th<strong>at</strong> the students think are good to ask! In addition, I<br />

asked students to give a summary of each chapter. These responses are also<br />

given in an appendix.<br />

I provided this m<strong>at</strong>erial in the appendices, because I think it is useful in<br />

two different ways. If you are a student studying thermodynamics, it is good<br />

to know wh<strong>at</strong> others <strong>at</strong> your level in the educ<strong>at</strong>ional system think. If you are<br />

struggling with a concept, it is reassuring to see th<strong>at</strong> others are too, and to<br />

see with wh<strong>at</strong> kind of questions they came up to find a way out. In a similar<br />

manner, it it helpful to see wh<strong>at</strong> others picked out as the most important part<br />

of each chapter. By providing the summaries I do not say th<strong>at</strong> I agree with<br />

them (in fact, sometimes I do not)(Dutchmen rarely agree anyhow), but it gives<br />

a standard for wh<strong>at</strong> others picked out as important. And on the other hand,<br />

if you are teaching this course, seeing wh<strong>at</strong> students perceive to be the most<br />

important part of the content is extremely helpful.<br />

Finally, a thanks to all students who took my classes. Your input has been<br />

essential, your questions have lead to a better understanding of the m<strong>at</strong>erial,<br />

and your research interests made me include a number of topics in these notes<br />

which otherwise would have been left out.<br />

History of these notes:<br />

1991 Original notes for first three chapters written using the program EXP.


INTRODUCTION XIII<br />

1992 Extra notes developed for chapters four and five.<br />

2001 Notes on chapters four and five expanded.<br />

2002 Notes converted to L ATEX, significantly upd<strong>at</strong>ed, and chapter six added.<br />

2003 Notes corrected, minor additions to first five chapters, some additions to<br />

chapter six.<br />

2006 Correction of errors. Upd<strong>at</strong>ed section on fluctu<strong>at</strong>ions. Added m<strong>at</strong>erial on<br />

correl<strong>at</strong>ion functions in chapter seven, but this far from complete.<br />

2008 Upd<strong>at</strong>ed the m<strong>at</strong>erial on correl<strong>at</strong>ion functions and included the two equ<strong>at</strong>ions<br />

of st<strong>at</strong>e rel<strong>at</strong>ed to pair correl<strong>at</strong>ion functions.<br />

2010 Corrected errors and made small upd<strong>at</strong>es.


PART I<br />

<strong>Thermodynamics</strong> Fundamentals<br />

1


Chapter 1<br />

Basic <strong>Thermodynamics</strong>.<br />

The goal of this chapter is to introduce the basic concepts th<strong>at</strong> we use in thermodynamics.<br />

One might think th<strong>at</strong> science has only exact definitions, but th<strong>at</strong><br />

is certainly not true. In the history of any scientific topic one always starts<br />

with language. How do we describe things? Which words do we use and wh<strong>at</strong><br />

do they mean? We need to get to some common understanding of wh<strong>at</strong> terms<br />

mean, before we can make them equivalent to some m<strong>at</strong>hem<strong>at</strong>ical description.<br />

This seems r<strong>at</strong>her vague, but since our n<strong>at</strong>ural way of communic<strong>at</strong>ion is based<br />

on language, it is the only way we can proceed.<br />

We all have some idea wh<strong>at</strong> volume means. But it takes some discussion to<br />

discover th<strong>at</strong> our ideas about volume are all similar. We are able to arrive <strong>at</strong><br />

some definitions we can agree on. Similarly, we all have a good intuition wh<strong>at</strong><br />

temper<strong>at</strong>ure is. More importantly, we can agree how we measure volume and<br />

temper<strong>at</strong>ure. We can take an arbitrary object and put it in w<strong>at</strong>er. The rise of<br />

the w<strong>at</strong>er level will tell us wh<strong>at</strong> the volume of the object is. We can put an<br />

object in contact with a mercury thermometer, and read of the temper<strong>at</strong>ure.<br />

We have used these procedures very often, since they are reproducible and give<br />

us the same result for the same object. Actually, not the same, but in the same<br />

Gaussian distribution. We can do error analysis.<br />

In this chapter we describe the basic terms used in thermodynamics. All<br />

these terms are descriptions of wh<strong>at</strong> we can see. We also make connections with<br />

the ideas of energy and work. The words use to do so are all familiar, and we<br />

build on the vocabulary from a typical introductory physics course. We make<br />

m<strong>at</strong>hem<strong>at</strong>ical connections between our newly defined quantities, and postul<strong>at</strong>e<br />

four laws th<strong>at</strong> hold for all systems. These laws are independent of the n<strong>at</strong>ure of<br />

a system. The m<strong>at</strong>hem<strong>at</strong>ical formul<strong>at</strong>ion by necessity uses many variables, and<br />

we n<strong>at</strong>urally connect with partial deriv<strong>at</strong>ives and multi-variable calculus.<br />

And then there is this quantity called temper<strong>at</strong>ure. We all have a good<br />

”feeling” for it, and standard measurements use physical phenomena like thermal<br />

expansion to measure it. We need to be more precise, however, and define<br />

temper<strong>at</strong>ure in a complic<strong>at</strong>ed manner based on the efficiency of Carnot engines.<br />

At the end of the chapter we show th<strong>at</strong> our definition is equivalent to<br />

3


4 CHAPTER 1. BASIC THERMODYNAMICS.<br />

the definition of temper<strong>at</strong>ure measured by an ideal gas thermometer. Once we<br />

have made th<strong>at</strong> connection, we know th<strong>at</strong> our definition of temper<strong>at</strong>ure is the<br />

same as the common one, since all thermometers are calibr<strong>at</strong>ed against ideal<br />

gas thermometers.<br />

There are two reasons for us to define temper<strong>at</strong>ure in this complic<strong>at</strong>ed manner.<br />

First of all, it is a definition th<strong>at</strong> uses energy in stead of a thermal m<strong>at</strong>erials<br />

property as a basis. Second, it allows us to define an even more illustrious quantity,<br />

named entropy. This new quantity allows us to define thermal equilibrium<br />

in m<strong>at</strong>hem<strong>at</strong>ical terms as a maximum of a function. The principle of maximum<br />

entropy is the corner stone for all th<strong>at</strong> is discussed in the chapter th<strong>at</strong> follow.<br />

1.1 Introduction.<br />

Simple beginnings.<br />

Wh<strong>at</strong> st<strong>at</strong>e am I in?<br />

In the mechanical world of the 19th century, physics was very easy. All you<br />

needed to know were the forces acting on particles. After th<strong>at</strong> it was simply F =<br />

ma. Although this formalism is not hard, actual calcul<strong>at</strong>ions are only feasible<br />

when the system under consider<strong>at</strong>ion contains a few particles. For example,<br />

the motion of solid objects can be described this way if they are considered to<br />

be point particles. In this context, we have all played with Lagrangians and<br />

Hamiltonians. Liquids and gases, however, are harder to deal with, and are<br />

often described in a continuum approxim<strong>at</strong>ion. Everywhere in space one defines<br />

a mass density and a velocity field. The continuity and Newton’s equ<strong>at</strong>ions then<br />

lead to the time evolution of the flow in the liquid.<br />

Asking the right questions.<br />

The big difference between a solid and a liquid is complexity. In first approxim<strong>at</strong>ion<br />

a solid can be described by six coordin<strong>at</strong>es (center of mass, orient<strong>at</strong>ion),<br />

while a liquid needs a mass density field which is essentially an infinite set of<br />

coordin<strong>at</strong>es. The calcul<strong>at</strong>ion of the motion of a solid is rel<strong>at</strong>ively easy, especially<br />

if one uses conserv<strong>at</strong>ion laws for energy, momentum, and angular momentum.<br />

The calcul<strong>at</strong>ion of the flow of liquids is still hard, even today, and is often done<br />

on computers using finite difference or finite element methods. In the 19th century,<br />

only the simplest flow p<strong>at</strong>terns could be analyzed. In many cases these<br />

flow p<strong>at</strong>terns are only details of the overall behavior of a liquid. Very often one<br />

is interested in more general quantities describing liquids and gases. In the 19th<br />

century many important questions have been raised in connection with liquids


1.1. INTRODUCTION. 5<br />

and gases, in the context of steam engines. How efficient can a steam engine<br />

be? Which temper<strong>at</strong>ure should it oper<strong>at</strong>e <strong>at</strong>? Hence the problem is wh<strong>at</strong> we<br />

need to know about liquids and gases to be able to answer these fundamental<br />

questions.<br />

Divide and conquer.<br />

In thermodynamics we consider macroscopic systems, or systems with a large<br />

number of degrees of freedom. Liquids and gases certainly belong to this class of<br />

systems, even if one does not believe in an <strong>at</strong>omic model! The only requirement<br />

is th<strong>at</strong> the system needs a description in terms of density and velocity fields.<br />

Solids can be described this way. In this case the mass density field is given<br />

by a constant plus a small vari<strong>at</strong>ion. The time evolution of these devi<strong>at</strong>ions<br />

from equilibrium shows oscill<strong>at</strong>ory p<strong>at</strong>terns, as expected. The big difference<br />

between a solid and a fluid is th<strong>at</strong> the devi<strong>at</strong>ions from average in a solid are<br />

small and can be ignored in first approxim<strong>at</strong>ion.<br />

No details, please.<br />

In a thermodynamic theory we are never interested in the detailed functional<br />

form of the density as a function of position, but only in macroscopic or global<br />

averages. Typical quantities of interest are the volume, magnetic moment, and<br />

electric dipole moment of a system. These macroscopic quantities, which can<br />

be measured, are called thermodynamic or st<strong>at</strong>e variables. They uniquely determine<br />

the thermodynamic st<strong>at</strong>e of a system.<br />

St<strong>at</strong>e of a system.<br />

The definition of the st<strong>at</strong>e of a system is in terms of oper<strong>at</strong>ions. Wh<strong>at</strong><br />

are the possible quantities which we can measure? In other words, how do we<br />

assign numbers to the fields describing a m<strong>at</strong>erial? Are there any problems?<br />

For example, one might think th<strong>at</strong> it is easy to measure the volume of a liquid<br />

in a beaker. But how does this work close to the critical point where the index<br />

of refraction of the liquid and vapor become the same? How does this work if<br />

there is no gravity? In thermodynamics we simply assume th<strong>at</strong> we are able to<br />

measure some basic quantities like volume.<br />

Another question is how many st<strong>at</strong>e variables do we need to describe a<br />

system. For example we prepare a certain amount of w<strong>at</strong>er and pour it in a<br />

beaker, which we cap. The next morning we find th<strong>at</strong> the w<strong>at</strong>er is frozen. It<br />

is obvious th<strong>at</strong> the w<strong>at</strong>er is not in the same st<strong>at</strong>e, and th<strong>at</strong> the inform<strong>at</strong>ion we<br />

had about the system was insufficient to uniquely define the st<strong>at</strong>e of the system.<br />

In this case we omitted temper<strong>at</strong>ure.<br />

It can be more complic<strong>at</strong>ed, though. Suppose we have prepared many iron<br />

bars always <strong>at</strong> the same shape, volume, mass, and temper<strong>at</strong>ure. They all look


6 CHAPTER 1. BASIC THERMODYNAMICS.<br />

the same. But then we notice th<strong>at</strong> some bars affect a compass needle while others<br />

do not. Hence not all bars are the same and we need additional inform<strong>at</strong>ion to<br />

uniquely define the st<strong>at</strong>e of the bar. Using the compass needle we can measure<br />

the magnetic field and hence we find the magnetic moment of the bar.<br />

In a next step we use all bars with a given magnetic moment. We apply a<br />

force on these bars and see how much they expand, from which we calcul<strong>at</strong>e the<br />

elastic constants of the bar. We find th<strong>at</strong> different bars have different values of<br />

the elastic constants. Wh<strong>at</strong> are we missing in terms of st<strong>at</strong>e variables? This is<br />

certainly not obvious. We need inform<strong>at</strong>ion about defects in the structure, or<br />

inform<strong>at</strong>ion about the mass density field beyond the average value.<br />

Is it changing?<br />

Was I in a different st<strong>at</strong>e before?<br />

At this point it is important to realize th<strong>at</strong> a measured value of a st<strong>at</strong>e<br />

variable is always a time-average. The pressure in an ideal gas fluctu<strong>at</strong>es, but<br />

the time scale of the fluctu<strong>at</strong>ions is much smaller than the time scale of the<br />

externally applied pressure changes. Hence these short time fluctu<strong>at</strong>ions can be<br />

ignored and are averaged out in a measurement. This does imply a warning,<br />

though. If the fluctu<strong>at</strong>ions in the st<strong>at</strong>e variables are on a time scale comparable<br />

with the dur<strong>at</strong>ion of the experiment a standard thermodynamic description<br />

is useless. If they are on a very long time scale, however, we can use our<br />

thermodynamic description again. In th<strong>at</strong> case the motion is so slow and we<br />

can use linear approxim<strong>at</strong>ions for all changes.<br />

How do we change?<br />

The values of the st<strong>at</strong>e variables for a given system can be modified by applying<br />

forces. An increase in pressure will decrease the volume, a change in<br />

magnetic induction will alter the magnetic moment. The pressure in a gas in<br />

a container is in many cases equal to the pressure th<strong>at</strong> this container exerts on<br />

the gas in order to keep it within the volume of the container. It is possible to<br />

use this pressure to describe the st<strong>at</strong>e of the system and hence pressure (and<br />

magnetic induction) are also st<strong>at</strong>e variables. One basic question in thermodynamics<br />

is how these st<strong>at</strong>e variables change when external forces are applied. In<br />

a more general way, if a specific st<strong>at</strong>e variable is changed by external means,<br />

how do the other st<strong>at</strong>e variables respond?<br />

Number of variables, again.


1.2. SOME DEFINITIONS. 7<br />

The number of st<strong>at</strong>e variables we need to describe the st<strong>at</strong>e of a system depends<br />

on the n<strong>at</strong>ure of th<strong>at</strong> system. We expand somewh<strong>at</strong> more on the previous<br />

discussion. An ideal gas, for example, is in general completely characterized by<br />

its volume, pressure, and temper<strong>at</strong>ure. It is always possible to add more st<strong>at</strong>e<br />

variables to this list. Perhaps one decides to measure the magnetic moment of<br />

an ideal gas too. Obviously, th<strong>at</strong> changes our knowledge of the st<strong>at</strong>e of the ideal<br />

gas. If the value of this additional st<strong>at</strong>e variable is always the same, no m<strong>at</strong>ter<br />

wh<strong>at</strong> we do in our experiment, then this variable is not essential. But one can<br />

always design experiments in which this st<strong>at</strong>e variable becomes essential. The<br />

magnetic moment is usually measured by applying a very small magnetic induction<br />

to the system. This external field should be zero for all purposes. If it<br />

is not, then we have to add the magnetic moment to our list of st<strong>at</strong>e variables.<br />

It is also possible th<strong>at</strong> one is not aware of additional essential st<strong>at</strong>e variables.<br />

Experiments will often indic<strong>at</strong>e th<strong>at</strong> more variables are needed. An example is<br />

an experiment in which we measure the properties of a piece of iron as a function<br />

of volume, pressure, and temper<strong>at</strong>ure. At a temper<strong>at</strong>ure of about 770 ◦ C some<br />

abnormal behavior is observed. As it turns out, iron is magnetic below this<br />

temper<strong>at</strong>ure and in order to describe the st<strong>at</strong>e of an iron sample one has to<br />

include the magnetic moment in the list of essential st<strong>at</strong>e variables. An ideal<br />

gas in a closed container is a simple system, but if the gas is allowed to escape<br />

via a valve, the number of particles in this gas also becomes an essential st<strong>at</strong>e<br />

variable needed to describe the st<strong>at</strong>e of the system inside the container.<br />

Are measured values always sp<strong>at</strong>ial averages?<br />

Are there further classific<strong>at</strong>ions of st<strong>at</strong>es or processes?<br />

1.2 Some definitions.<br />

Two types of processes.<br />

If one takes a block of wood, and splits it into two pieces, one has performed<br />

a simple action. On the level of thermodynamic variables one writes something<br />

like V = V1 + V2 for the volumes and similar equ<strong>at</strong>ions for other st<strong>at</strong>e variables.<br />

The detailed n<strong>at</strong>ure of this process is, however, not accessible in this language.<br />

In addition, if we put the two pieces back together again, they do in general not<br />

stay together. The process was irreversible. In general, in thermodynamics one<br />

only studies the reversible behavior of macroscopic systems. An example would


8 CHAPTER 1. BASIC THERMODYNAMICS.<br />

be the study of the liquid to vapor transition. M<strong>at</strong>erial is slowly transported<br />

from one phase to another and can go back if the causes are reversed. The<br />

st<strong>at</strong>e variables one needs to consider in this case are the pressure, temper<strong>at</strong>ure,<br />

volume, interface area (because of surface tension), and perhaps others in more<br />

complic<strong>at</strong>ed situ<strong>at</strong>ions.<br />

When there is NO change.<br />

Obviously, all macroscopic systems change as a function of time. Most of<br />

these changes, however, are on a microscopic level and are not of interest. We<br />

are not able to measure them directly. Therefore, in thermodynamics one defines<br />

a steady st<strong>at</strong>e when all thermodynamic variables are independent of time. A<br />

resistor connected to a constant voltage is in a steady st<strong>at</strong>e. The current through<br />

the resistor is constant and although there is a flow of charge, there are no net<br />

changes in the resistor. The same amount of charge comes in as goes out.<br />

Thermodynamic equilibrium describes a more restricted situ<strong>at</strong>ion. A system<br />

is in thermodynamic equilibrium if it is in a steady st<strong>at</strong>e and if there are no<br />

net macroscopic currents (of energy, particles, etc) over macroscopic distances.<br />

There is some ambiguity in this definition, connected to the scale and magnitude<br />

of the currents. A vapor-liquid interface like the ocean, with large waves, is<br />

clearly not in equilibrium. But how small do the waves have to be in order th<strong>at</strong><br />

we can say th<strong>at</strong> the system is in equilibrium? If we discuss the thermal balance<br />

between oceans and <strong>at</strong>mosphere, are waves important? Also, the macroscopic<br />

currents might be very small. Glass, for example, is not in thermal equilibrium<br />

according to a strict definition, but the changes are very slow with a time scale<br />

of hundreds of years. Hence even if we cannot measure macroscopic currents,<br />

they might be there. We will in general ignore these situ<strong>at</strong>ions, since they tend<br />

not to be of interest on the time scale of the experiments!<br />

St<strong>at</strong>e functions.<br />

Wh<strong>at</strong> do you think about hysteresis loops in magnets?<br />

Once we understand the n<strong>at</strong>ure of thermal equilibrium, we can generalize<br />

the concept of st<strong>at</strong>e variables. A st<strong>at</strong>e function is any quantity which in thermodynamic<br />

equilibrium only depends on the values of the st<strong>at</strong>e variables and not<br />

on the history (or future?) of the sample. A simple st<strong>at</strong>e function would be the<br />

product of the pressure and volume. This product has a physical interpret<strong>at</strong>ion,<br />

but cannot be measured directly.<br />

Two types of variables.


1.2. SOME DEFINITIONS. 9<br />

Thermodynamic variables come in two varieties. If one takes a system in<br />

equilibrium the volume of the left half is only half the total volume (surprise)<br />

but the pressure in the left half is equal to the pressure of the total system.<br />

There are only two possibilities. Either a st<strong>at</strong>e variable scales linearly with the<br />

size of the system, or is independent of the size of the system. In other words,<br />

if we consider two systems in thermal equilibrium, made of identical m<strong>at</strong>erial,<br />

one with volume V1 and one with volume V2, a st<strong>at</strong>e variable X either obeys<br />

X1<br />

V1<br />

= X2<br />

V2 or X1 = X2. In the first case the st<strong>at</strong>e variable is called extensive and<br />

in the second case it is called intensive. Extensive st<strong>at</strong>e variables correspond to<br />

generalized displacements. For the volume this is easy to understand; increasing<br />

volume means displacing outside m<strong>at</strong>erial and doing work on it in the process.<br />

Intensive st<strong>at</strong>e variables correspond to generalized forces. The pressure is the<br />

force needed to change the volume. For each extensive st<strong>at</strong>e variable there is a<br />

corresponding intensive st<strong>at</strong>e variable and vice-versa.<br />

Extensive st<strong>at</strong>e variables correspond to quantities which can be determined,<br />

measured, or prescribed directly. The volume of a gas can be found by measuring<br />

the size of the container, the amount of m<strong>at</strong>erial can be measured using a<br />

balance. Intensive st<strong>at</strong>e variables are measured by making contact with something<br />

else. We measure temper<strong>at</strong>ure by using a thermometer, and pressure using<br />

a manometer. Such measurements require equilibrium between the sample and<br />

measuring device.<br />

Note th<strong>at</strong> this distinction limits where and how we can apply thermodynamics.<br />

The gravit<strong>at</strong>ional energy of a large system is not proportional to the amount<br />

of m<strong>at</strong>erial, but to the amount of m<strong>at</strong>erial to the five-thirds power. If the force<br />

of gravity is the dominant force internally in our system we need other theories.<br />

Electrical forces are also of a long range, but because we have both positive<br />

and neg<strong>at</strong>ive charges, they are screened. Hence in m<strong>at</strong>erials physics we normally<br />

have no fundamental problems with applying thermodynamics.<br />

Thermodynamic limit.<br />

At this point we are able to define wh<strong>at</strong> we mean by a large system. R<strong>at</strong>ios<br />

of an extensive st<strong>at</strong>e variable and the volume, like X<br />

V , are often called densities.<br />

It is customary to write these densities in lower case, x = X<br />

V . If the volume is too<br />

small, x will depend on the volume. Since X is extensive, this is not supposed to<br />

be the case. In order to get rid of the effects of a finite volume (surface effects!)<br />

one has to take the limit V → ∞. This is called the thermodynamic limit.<br />

All our m<strong>at</strong>hem<strong>at</strong>ical formulas are strictly speaking only correct in this limit.<br />

In practice, this means th<strong>at</strong> the volume has to be large enough in order th<strong>at</strong><br />

changes in the volume do not change the densities anymore. It is always possible<br />

to write x(V ) = x∞ +αV −1 +O(V −2 ). The magnitude of α decides which value<br />

of the volume is large enough.<br />

<strong>Physics</strong> determines the rel<strong>at</strong>ion between st<strong>at</strong>e variables.


10 CHAPTER 1. BASIC THERMODYNAMICS.<br />

Why are all these definitions important? So far we have not discussed any<br />

physics. If all the st<strong>at</strong>e variables would be independent we could stop right here.<br />

Fortun<strong>at</strong>ely, they are not. Some st<strong>at</strong>e variables are rel<strong>at</strong>ed by equ<strong>at</strong>ions of st<strong>at</strong>e<br />

and these equ<strong>at</strong>ions contain the physics of the system. It is important to note<br />

th<strong>at</strong> these equ<strong>at</strong>ions of st<strong>at</strong>e only rel<strong>at</strong>e the values of the st<strong>at</strong>e variables when<br />

the system is in thermal equilibrium, in the thermodynamic limit! If a system<br />

is not in equilibrium, any combin<strong>at</strong>ion of st<strong>at</strong>e variables is possible. It is even<br />

possible to construct non-equilibrium systems in which the actual definition or<br />

measurement of certain st<strong>at</strong>e variables is not useful or becomes ambiguous.<br />

Simple examples of equ<strong>at</strong>ions of st<strong>at</strong>e are the ideal gas law pV = NRT and<br />

Curie’s law M = CNH<br />

T . The first equ<strong>at</strong>ion rel<strong>at</strong>es the product of the pressure<br />

p and volume V of an ideal gas to the number of moles of gas N and the<br />

temper<strong>at</strong>ure T. The constant of proportionality, R, is the molar gas constant,<br />

which is the product of Boltzmann’s constant kB and Avogadro’s number NA.<br />

The second equ<strong>at</strong>ion rel<strong>at</strong>es the magnetic moment M to the number of moles<br />

of <strong>at</strong>oms N, the magnetic field H, and the temper<strong>at</strong>ure T. The constant of proportionality<br />

is Curie’s constant C. Note th<strong>at</strong> in thermodynamics the preferred<br />

way of measuring the amount of m<strong>at</strong>erial is in terms of moles, which again can<br />

be defined independent of a molecular model. Note, too, th<strong>at</strong> in electricity and<br />

magnetism we always use the magnetiz<strong>at</strong>ion density in the Maxwell equ<strong>at</strong>ions,<br />

but th<strong>at</strong> in thermodynamics we define the total magnetiz<strong>at</strong>ion as the relevant<br />

quantity. This makes M an extensive quantity.<br />

Equ<strong>at</strong>ions of st<strong>at</strong>e are rel<strong>at</strong>ed to st<strong>at</strong>e functions. For any system we can<br />

define the st<strong>at</strong>e function Fst<strong>at</strong>e = pV − NRT . It will take on all kinds of values.<br />

We then define the special class of systems for which Fst<strong>at</strong>e ≡ 0, identical to<br />

zero, as an ideal gas. The right hand side then leads to an equ<strong>at</strong>ion of st<strong>at</strong>e,<br />

which can be used to calcul<strong>at</strong>e one of the basic st<strong>at</strong>e variables if others are<br />

known. The practical applic<strong>at</strong>ion of this idea is to look for systems for which<br />

Fst<strong>at</strong>e is small, with small defined in an appropri<strong>at</strong>e context. In th<strong>at</strong> case we can<br />

use the class with zero st<strong>at</strong>e function as a first approxim<strong>at</strong>ion of the real system.<br />

In many cases the ideal gas approxim<strong>at</strong>ion is a good start for a description of a<br />

real gas, and is a start for system<strong>at</strong>ic improvements of the description!<br />

How do we get equ<strong>at</strong>ions of st<strong>at</strong>e?<br />

Equ<strong>at</strong>ions of st<strong>at</strong>e have two origins. One can completely ignore the microscopic<br />

n<strong>at</strong>ure of m<strong>at</strong>ter and simply postul<strong>at</strong>e some rel<strong>at</strong>ion. One then uses the<br />

laws of thermodynamics to derive functional forms for specific st<strong>at</strong>e variables as<br />

a function of the others, and compares the predicted results with experiment.<br />

The ideal gas law has this origin. This procedure is exactly wh<strong>at</strong> is done in thermodynamics.<br />

One does not need a model for the detailed n<strong>at</strong>ure of the systems,<br />

but derives general conclusions based on the average macroscopic behavior of a<br />

system in the thermodynamic limit.<br />

In order to derive equ<strong>at</strong>ions of st<strong>at</strong>e, however, one has to consider the microscopic<br />

aspects of a system. Our present belief is th<strong>at</strong> all systems consist of <strong>at</strong>oms.


1.2. SOME DEFINITIONS. 11<br />

If we know the forces between the <strong>at</strong>oms, the theory of st<strong>at</strong>istical mechanics will<br />

tell us how to derive equ<strong>at</strong>ions of st<strong>at</strong>e. There is again a choice here. It is possible<br />

to postul<strong>at</strong>e the forces. The equ<strong>at</strong>ions of st<strong>at</strong>e could then be derived from<br />

molecular dynamics calcul<strong>at</strong>ions, for example. The other route derives these<br />

effective forces from the laws of quantum mechanics and the structure of the<br />

<strong>at</strong>oms in terms of electrons and nuclei. The interactions between the particles<br />

in the <strong>at</strong>oms are simple Coulomb interactions in most cases. These Coulomb<br />

interactions follow from yet a deeper theory, quantum electro-dynamics, and are<br />

only a first approxim<strong>at</strong>ion. These corrections are almost always unimportant in<br />

the study of m<strong>at</strong>erials and only show up <strong>at</strong> higher energies in nuclear physics<br />

experiments.<br />

Why do we need equ<strong>at</strong>ions of st<strong>at</strong>e?<br />

Equ<strong>at</strong>ions of st<strong>at</strong>es can be used to classify m<strong>at</strong>erials. They can be used<br />

to derive <strong>at</strong>omic properties of m<strong>at</strong>erials. For example, <strong>at</strong> low densities a gas<br />

of helium <strong>at</strong>oms and a gas of methane <strong>at</strong>oms both follow the ideal gas law.<br />

This indic<strong>at</strong>es th<strong>at</strong> in this limit the internal structure of the molecules does not<br />

affect the motion of the molecules! In both cases they seem to behave like point<br />

particles. L<strong>at</strong>er we will see th<strong>at</strong> other quantities are different. For example, the<br />

internal energy certainly is larger for methane where rot<strong>at</strong>ions and transl<strong>at</strong>ions<br />

play a role.<br />

Classific<strong>at</strong>ion of changes of st<strong>at</strong>e.<br />

Since a st<strong>at</strong>ic universe is not very interesting, one has to consider changes<br />

in the st<strong>at</strong>e variables. In a thermodynamic transform<strong>at</strong>ion or process a system<br />

changes one or more of its st<strong>at</strong>e variables. A spontaneous process takes place<br />

without any change in the externally imposed constraints. The word constraint<br />

in this context means an external description of the st<strong>at</strong>e variables for the system.<br />

For example, we can keep the volume of a gas the same, as well as the<br />

temper<strong>at</strong>ure and the amount of gas. Or if the temper<strong>at</strong>ure of the gas is higher<br />

than the temper<strong>at</strong>ure of the surroundings, we allow the gas to cool down. In<br />

an adiab<strong>at</strong>ic process no he<strong>at</strong> is exchanged between the system and the environment.<br />

A process is called isothermal if the temper<strong>at</strong>ure of the system remains<br />

the same, isobaric if the pressure does not change, and isochoric if the mass<br />

density (the number of moles of particles divided by the volume) is constant. If<br />

the change in the system is infinitesimally slow, the process is quasist<strong>at</strong>ic .<br />

Reversible process.<br />

The most important class of processes are those in which the system starts<br />

in equilibrium, the process is quasist<strong>at</strong>ic, and all the intermedi<strong>at</strong>e st<strong>at</strong>es and the<br />

final st<strong>at</strong>e are in equilibrium. These processes are called reversible. The process


12 CHAPTER 1. BASIC THERMODYNAMICS.<br />

can be described by a continuous p<strong>at</strong>h in the space of the st<strong>at</strong>e variables, and<br />

this p<strong>at</strong>h is restricted to the surfaces determined by the equ<strong>at</strong>ions of st<strong>at</strong>e for<br />

the system. By inverting all external forces, the p<strong>at</strong>h in the space of the st<strong>at</strong>e<br />

functions will be reversed, which prompted the name for this type of process.<br />

Reversible processes are important because they can be described m<strong>at</strong>hem<strong>at</strong>ically<br />

via the equ<strong>at</strong>ions of st<strong>at</strong>e. This property is lost for an irreversible process<br />

between two equilibrium st<strong>at</strong>es, where we only have a useful m<strong>at</strong>hem<strong>at</strong>ical description<br />

of the initial and final st<strong>at</strong>e. As we will see l<strong>at</strong>er, the second law of<br />

thermodynamics makes another distinction between reversible and irreversible<br />

processes.<br />

How does a process become irreversible?<br />

An irreversible process is either a process which happens too fast or which is<br />

discontinuous. The sudden opening of a valve is an example of the last case. The<br />

system starts out in equilibrium with volume Vi and ends in equilibrium with<br />

a larger volume Vf . For the intermedi<strong>at</strong>e st<strong>at</strong>es the volume is not well defined,<br />

though. Such a process takes us outside of the space of st<strong>at</strong>e variables we<br />

consider. It can still be described in the phase space of all system variables, and<br />

m<strong>at</strong>hem<strong>at</strong>ically it is possible to define the volume, but details of this definition<br />

will play a role in the description of the process. Another type of irreversible<br />

process is the same expansion from Vi to Vf in a controlled way. The volume<br />

is well-defined everywhere in the process, but the system is not in equilibrium<br />

in the intermedi<strong>at</strong>e st<strong>at</strong>es. The process is going too fast. In an ideal gas this<br />

would mean, for example, pV ̸= NRT for the intermedi<strong>at</strong>e stages.<br />

Are there general principles connecting the values of st<strong>at</strong>e variables, valid for<br />

all systems?<br />

1.3 Zeroth Law of <strong>Thermodynamics</strong>.<br />

General rel<strong>at</strong>ions.<br />

An equ<strong>at</strong>ion of st<strong>at</strong>e specifies a rel<strong>at</strong>ion between st<strong>at</strong>e variables which holds<br />

for a certain class of systems. It represents the physics particular to th<strong>at</strong> system.<br />

There are, however, a few rel<strong>at</strong>ions th<strong>at</strong> hold for all systems, independent of the<br />

n<strong>at</strong>ure of the system. Following an old tradition, these rel<strong>at</strong>ions are called the<br />

laws of thermodynamics. There are four of them, numbered 0 through 3. The<br />

middle two are the most important, and they have been paraphrased in the<br />

following way. Law one tells you th<strong>at</strong> in the game of thermodynamics you<br />

cannot win. The second law makes it even worse, you cannot break even.


1.4. FIRST LAW: ENERGY. 13<br />

Law zero.<br />

The zeroth law is rel<strong>at</strong>ively trivial. It discusses systems in equilibrium.<br />

Two systems are in thermal equilibrium if they are in contact and the total<br />

system, encompassing the two systems as subsystems, is in equilibrium. In other<br />

words, two systems in contact are in equilibrium if the individual systems are in<br />

equilibrium and there are no net macroscopic currents between the systems. The<br />

zeroth law st<strong>at</strong>es th<strong>at</strong> if equilibrium system A is in contact and in equilibrium<br />

with systems B and C (not necessarily <strong>at</strong> the same time, but A does not change),<br />

then systems B and C are also in equilibrium with each other. If B and C are not<br />

in contact, it would mean th<strong>at</strong> if we bring them in contact no net macroscopic<br />

currents will flow.<br />

Significance of law zero.<br />

The importance of this law is th<strong>at</strong> it enables to define universal standards<br />

for temper<strong>at</strong>ure, pressure, etc. If two different systems cause the same reading<br />

on the same thermometer, they have the same temper<strong>at</strong>ure. A temper<strong>at</strong>ure<br />

scale on a new thermometer can be set by comparing it with systems of known<br />

temper<strong>at</strong>ure. Therefore, the first law is essential, without it we would not be<br />

able to give a meaningful analysis of any experiment. We postul<strong>at</strong>e it a a law,<br />

because we have not seen any exceptions. We postul<strong>at</strong>e it as a law, because we<br />

absolutely need it. We cannot prove it to be true. It cannot be derived from<br />

st<strong>at</strong>istical mechanics, because in th<strong>at</strong> theory it is also a basic or fundamental<br />

assumption. But if one rejects it completely, one throws away a few hundred<br />

years of successful science. But wh<strong>at</strong> if the situ<strong>at</strong>ion is similar to Newton’s<br />

F = ma, where Einstein showed the limits of validity? Th<strong>at</strong> scenario is certainly<br />

possible, but we have not yet needed it, or seen any reason for its need. Also,<br />

it is completely unclear wh<strong>at</strong> kind of theory should be used to replace all wh<strong>at</strong><br />

we will explore in these notes.<br />

Are there any consequences for the sizes of the systems?<br />

1.4 First law: Energy.<br />

He<strong>at</strong> is energy flow.<br />

The first law of thermodynamics st<strong>at</strong>es th<strong>at</strong> energy is conserved. The change<br />

in internal energy U of a system is equal to the amount of he<strong>at</strong> energy added to<br />

the system minus the amount of work done by the system. It implies th<strong>at</strong> he<strong>at</strong>


14 CHAPTER 1. BASIC THERMODYNAMICS.<br />

is a form of energy. Technically, he<strong>at</strong> describes the flow of energy, but we are<br />

very sloppy in our use of words here. The formal st<strong>at</strong>ement of the first law is<br />

dU = ¯ dQ − ¯ dW (1.1)<br />

The amount of he<strong>at</strong> added to the system is ¯ dQ and the amount of work done<br />

by the system is ¯ dW . The m<strong>at</strong>hem<strong>at</strong>ical formul<strong>at</strong>ion of the first law also shows<br />

an important characteristic of thermodynamics. It is often possible to define<br />

thermodynamic rel<strong>at</strong>ions only via changes in the thermodynamic quantities.<br />

Note th<strong>at</strong> we define the work term as work done on the outside world. It<br />

represents a loss of energy of the system. This is the standard definition, and<br />

represents the fact th<strong>at</strong> in the original analysis one was interested in supplying<br />

he<strong>at</strong> to an engine, which then did work. Some books, however, try to be consistent,<br />

and write the work term as work done on the system. In th<strong>at</strong> case there<br />

is no minus sign in equ<strong>at</strong>ion 1.1. Although th<strong>at</strong> is somewh<strong>at</strong> ne<strong>at</strong>er, it causes<br />

too much confusion with standard texts. It is always much too easy to lose a<br />

minus sign.<br />

The first law again is a st<strong>at</strong>ement th<strong>at</strong> has always been observed to be<br />

true. In addition, the first law does follow directly in a st<strong>at</strong>istical mechanical<br />

tre<strong>at</strong>ment. We have no reason to doubt the validity of the first law, and we<br />

discard any proposals of engines th<strong>at</strong> cre<strong>at</strong>e energy out of nothing. But again,<br />

there is no absolute proof of its validity. And, again as well, if one discards<br />

the first law, all the remainder of these notes will be invalid as well. A theory<br />

without the first law is very difficult to imagine.<br />

The internal energy is a st<strong>at</strong>e variable.<br />

The internal energy U is a st<strong>at</strong>e variable and an infinitesimal change in<br />

internal energy is an exact differential. Since U is a st<strong>at</strong>e variable, the value of<br />

any integral ∫ dU depends only on the values of U <strong>at</strong> the endpoints of the p<strong>at</strong>h<br />

in the space of st<strong>at</strong>e variables, and not on the specific p<strong>at</strong>h between the endpoints.<br />

The internal energy U has to be a st<strong>at</strong>e variable, or else we could devise<br />

a process in which a system goes through a cycle and returns to its original st<strong>at</strong>e<br />

while loosing or gaining energy. For example, this could mean th<strong>at</strong> a burning<br />

piece of coal today would produce less he<strong>at</strong> than tomorrow. If the internal<br />

energy would not be a st<strong>at</strong>e variable, we would have sources of free energy.<br />

Exact differentials.<br />

The concept of exact differentials is very important, and hence we will illustr<strong>at</strong>e<br />

it by using some examples. Assume the function f is a st<strong>at</strong>e function of<br />

the st<strong>at</strong>e variables x and y only, f(x,y). For small changes we can write<br />

df =<br />

( )<br />

∂f<br />

dx +<br />

∂x y<br />

( )<br />

∂f<br />

dy (1.2)<br />

∂y x


1.4. FIRST LAW: ENERGY. 15<br />

where in the not<strong>at</strong>ion for the partial deriv<strong>at</strong>ives the variable which is kept constant<br />

is also indic<strong>at</strong>ed. This is always very useful in thermodynamics, because<br />

one often changes variables. There would be no problems if quantities were defined<br />

directly like f(x, y) = x + y. In thermodynamics, however, most quantities<br />

are defined by or measured via small changes in a system. Hence, suppose the<br />

change in a quantity g is rel<strong>at</strong>ed to changes in the st<strong>at</strong>e variables x and y via<br />

¯dg = h(x, y)dx + k(x, y)dy (1.3)<br />

Is the quantity g a st<strong>at</strong>e function, in other words is g uniquely determined<br />

by the st<strong>at</strong>e of a system or does it depend on the history, on how the system<br />

got into th<strong>at</strong> st<strong>at</strong>e? It turns out th<strong>at</strong> a necessary and sufficient condition for g<br />

to be a st<strong>at</strong>e function is th<strong>at</strong><br />

( )<br />

∂h<br />

=<br />

∂y x<br />

( )<br />

∂k<br />

∂x y<br />

(1.4)<br />

everywhere in the x-y space. The necessity follows immedi<strong>at</strong>ely from 1.2, as<br />

long as we assume th<strong>at</strong> the partial deriv<strong>at</strong>ives in 1.4 exist and are continuous.<br />

This is because in second order deriv<strong>at</strong>ives we can interchange the order of the<br />

deriv<strong>at</strong>ives under such conditions. Th<strong>at</strong> it is sufficient can be shown as follows.<br />

Consider a p<strong>at</strong>h (x, y) = (ϕ(t), ψ(t)) from (x1, y1) <strong>at</strong> t1 to (x2, y2) <strong>at</strong> t2 and<br />

dt ,<br />

integr<strong>at</strong>e ¯ dg , using dx = dϕ<br />

dt<br />

Define<br />

∫ t2<br />

t1<br />

dt , dy = dψ<br />

dt<br />

(<br />

h(ϕ(t), ψ(t)) dϕ<br />

+ k(ϕ(t), ψ(t))dψ<br />

dt dt<br />

H(x, y) =<br />

)<br />

dt (1.5)<br />

∫ x<br />

dx<br />

0<br />

′ h(x ′ ∫ y<br />

, y) + dy<br />

0<br />

′ k(0, y ′ ) (1.6)<br />

and H(t) = H(ϕ(t), ψ(t)). It follows th<strong>at</strong><br />

dH<br />

dt =<br />

( )<br />

∂H dϕ<br />

∂x y dt +<br />

( )<br />

∂H dψ<br />

∂y x dt<br />

(1.7)<br />

The partial deriv<strong>at</strong>ives of H are easy to calcul<strong>at</strong>e:<br />

( )<br />

∂H<br />

(x, y) = h(x, y) (1.8)<br />

∂x y<br />

( )<br />

∂H<br />

(x, y) =<br />

∂y x<br />

This implies th<strong>at</strong><br />

∫ x<br />

0<br />

dx ′<br />

dx<br />

0<br />

′<br />

( )<br />

∂h<br />

∂y<br />

( )<br />

∂k<br />

∂x<br />

∫ x<br />

(x<br />

x<br />

′ , y) + k(0, y) =<br />

(x<br />

y<br />

′ , y) + k(0, y) = k(x, y) (1.9)


16 CHAPTER 1. BASIC THERMODYNAMICS.<br />

∫ t2<br />

t1<br />

¯dg =<br />

∫ t2<br />

t1<br />

dH<br />

dt<br />

(1.10)<br />

and hence the integral of ¯ dg is equal to H(t2)−H(t1) which does not depend<br />

on the p<strong>at</strong>h taken between the end-points of the integr<strong>at</strong>ion.<br />

Example.<br />

An example might illustr<strong>at</strong>e this better. Suppose x and y are two st<strong>at</strong>e variables,<br />

and they determine the internal energy completely. If we define changes<br />

in the internal energy via changes in the st<strong>at</strong>e variables x and y via<br />

dU = x 2 ydx + 1<br />

3 x3dy (1.11)<br />

we see immedi<strong>at</strong>ely th<strong>at</strong> this definition is correct, the energy U is a st<strong>at</strong>e function.<br />

The partial deriv<strong>at</strong>ives obey the symmetry rel<strong>at</strong>ion 1.4 and one can simply<br />

integr<strong>at</strong>e dU and check th<strong>at</strong> we get U(x, y) = 1<br />

3x3y + U0.<br />

The changes in he<strong>at</strong> and work are now assumed to be rel<strong>at</strong>ed in the following<br />

way<br />

¯dQ = 1<br />

2 x2 ydx + 1<br />

2 x3 dy (1.12)<br />

¯dW = − 1<br />

2 x2 ydx + 1<br />

6 x3 dy (1.13)<br />

These definitions do indeed obey the first law 1.1. It is also clear using the<br />

symmetry rel<strong>at</strong>ion 1.4 th<strong>at</strong> these two differentials are not exact.<br />

Suppose the system which is described above is originally in the st<strong>at</strong>e (x, y) =<br />

(0, 0). Now we change the st<strong>at</strong>e of the system by a continuous transform<strong>at</strong>ion<br />

from (0, 0) to (1, 1). We do this in two different ways, however. P<strong>at</strong>h one takes<br />

us from (0, 0) to (0, 1) to (1, 1) along two straight line segments, p<strong>at</strong>h two is<br />

similar from (0, 0) to (1, 0) to (1, 1). The integrals for dU, ¯ dQ , and ¯ dW are<br />

easy, since along each part of each p<strong>at</strong>h either dx or dy is zero.<br />

First take p<strong>at</strong>h one.<br />

U(1, 1) − U(0, 0) =<br />

∆Q =<br />

∆W =<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

dy 1<br />

2 (0)3 +<br />

dy 1<br />

6 (0)3 +<br />

dy 1<br />

3 (0)3 +<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

∫ 1<br />

0<br />

dx 1<br />

2 x21 = 1<br />

6<br />

dxx 2 1 = 1<br />

3<br />

dx(− 1<br />

2 )x21 = − 1<br />

6<br />

(1.14)<br />

(1.15)<br />

(1.16)<br />

First of all, the change in U is consistent with the st<strong>at</strong>e function we found,<br />

U(x, y) = 1<br />

3 x3 y + U0. Second, we have ∆U = ∆Q − ∆W indeed. It is easy to


1.4. FIRST LAW: ENERGY. 17<br />

calcul<strong>at</strong>e th<strong>at</strong> for the second p<strong>at</strong>h we have ∆U = 1<br />

1<br />

1<br />

3 , ∆Q = 2 , and ∆W = 6 .<br />

The change in internal energy is indeed the same, and the first law is again<br />

s<strong>at</strong>isfied.<br />

Importance of Q and W not being st<strong>at</strong>e functions.<br />

Life on earth would have been very different if Q and W would have been<br />

st<strong>at</strong>e variables. Steam engines would not exist, and you can imagine all consequences<br />

of th<strong>at</strong> fact.<br />

Expand on the consequences of Q and W being st<strong>at</strong>e functions<br />

Any engine repe<strong>at</strong>s a certain cycle over and over again. A complete cycle<br />

in our example above might be represented by a series of continuous changes in<br />

the st<strong>at</strong>e variables (x,y) like (0, 0) → (0, 1) → (1, 1) → (1, 0) → (0, 0). After the<br />

completion of one cycle, the energy U is the same as <strong>at</strong> the start of the cycle.<br />

The change in he<strong>at</strong> for this cycle is ∆Q = 1 1 1<br />

6 − 2 = − 3 and the work done on<br />

the environment is ∆W = − 1 1 1<br />

6 − 6 = − 3 . This cycle represents a he<strong>at</strong>er: since<br />

∆Q is neg<strong>at</strong>ive, he<strong>at</strong> is added to the environment and since ∆W is neg<strong>at</strong>ive<br />

the environment does work on the system. Running the cycle in the opposite<br />

direction yields an engine converting he<strong>at</strong> into work. If Q and W would be st<strong>at</strong>e<br />

variables, for each complete cycle we would have ∆Q = ∆W = 0, and no net<br />

change of work into he<strong>at</strong> and vice-versa would be possible!<br />

When was the first steam engine constructed?<br />

Work can be done in many different ways. A change in any of the extensive<br />

st<strong>at</strong>e variables of the system will cause a change in energy, or needs a force<br />

in order th<strong>at</strong> it happens. Consider a system with volume V, surface area A,<br />

polariz<strong>at</strong>ion ⃗ P , magnetic moment ⃗ M, and number of moles of m<strong>at</strong>erial N. The<br />

work done by the system on the environment is<br />

¯dW = pdV − σdA − ⃗ Ed ⃗ P − ⃗ Hd ⃗ M − µdN (1.17)<br />

where the forces are rel<strong>at</strong>ed to the intensive variables pressure p, surface tension<br />

σ, electric field ⃗ E, magnetic field ⃗ H, and chemical potential µ. Note th<strong>at</strong> some<br />

textbooks tre<strong>at</strong> the µdN term in a special way. There is, however, no formal<br />

need to do so. The general form is<br />

¯dW = − ∑<br />

xjdXj<br />

j<br />

(1.18)<br />

where the generalized force xj causes a generalized displacement dXj in the<br />

st<strong>at</strong>e variable Xj.


18 CHAPTER 1. BASIC THERMODYNAMICS.<br />

The signs in work are normally neg<strong>at</strong>ive. If we increase the total magnetic<br />

moment of a sample in an external magnetic field, we have to add energy to<br />

the sample. In other words, an increase in the total magnetic moment increases<br />

the internal energy, and work has to be done on the sample. The work done<br />

by the sample is neg<strong>at</strong>ive. Note th<strong>at</strong> the pdV term has the opposite sign from<br />

all others. If we increase the volume of a sample, we push outside m<strong>at</strong>erial<br />

away, and do work on the outside. A positive pressure decreases the volume,<br />

while a positive magnetic field increases the magnetic magnetiz<strong>at</strong>ion in general.<br />

This difference in sign is on one historical, and is justified by the old, intuitive<br />

definitions of pressure and other quantities. But it also has a deeper meaning.<br />

Volume tells us how much space the sample occupies, while all other extensive<br />

quantities tell us how much of something is in th<strong>at</strong> space. In terms of densities,<br />

volume is in the denomin<strong>at</strong>or, while all other variables are in the numer<strong>at</strong>or.<br />

This gives a change in volume the opposite effect from all other changes.<br />

1.5 Second law: Entropy.<br />

Clausius and Kelvin.<br />

The second law of thermodynamics tells us th<strong>at</strong> life is not free. According<br />

to the first law we can change he<strong>at</strong> into work, apparently without limits. The<br />

second law, however, puts restrictions on this exchange. There are two versions<br />

of the second law, due to Kelvin and Clausius. Clausius st<strong>at</strong>ed th<strong>at</strong> there are<br />

no thermodynamic processes in which the only net change is a transfer of he<strong>at</strong><br />

from one reservoir to a second reservoir which has a higher temper<strong>at</strong>ure. Kelvin<br />

formul<strong>at</strong>ed it in a different way: there are no thermodynamic processes in which<br />

the only effect is to extract a certain amount of he<strong>at</strong> from a reservoir and convert<br />

it completely into work. The opposite is possible, though. These two st<strong>at</strong>ements<br />

are equivalent as we will see.<br />

He<strong>at</strong> is a special form of energy exchange.<br />

The second law singles out he<strong>at</strong> as compared to all other forms of energy.<br />

Since work is defined via a change in the extensive st<strong>at</strong>e variables, we can think<br />

of he<strong>at</strong> as a change of the internal degrees of freedom of the system. Hence<br />

he<strong>at</strong> represents all the degrees of freedom we have swept under the rug when<br />

we limited st<strong>at</strong>e variables to measurable, average macroscopic quantities. The<br />

only reason th<strong>at</strong> we can say anything <strong>at</strong> all about he<strong>at</strong> is th<strong>at</strong> it is connected to<br />

an extremely large number of variables (because of the thermodynamic limit).<br />

In th<strong>at</strong> case the m<strong>at</strong>hem<strong>at</strong>ical laws of large numbers apply, and the st<strong>at</strong>ements<br />

about he<strong>at</strong> become purely st<strong>at</strong>istical. In st<strong>at</strong>istical mechanics we will return<br />

to this point. Note th<strong>at</strong> the second law does not limit the exchange of energy<br />

switching from one form of work to another. In principal we could change


1.5. SECOND LAW: ENTROPY. 19<br />

mechanical work into electrical work without penalty! In practice, he<strong>at</strong> is always<br />

gener<strong>at</strong>ed.<br />

He<strong>at</strong> as a measurable quantity.<br />

One important implicit assumption in these st<strong>at</strong>ements is th<strong>at</strong> a large outside<br />

world does exist. In the first law we define the change of energy via an<br />

exchange of he<strong>at</strong> and work with the outside world. Hence we assume th<strong>at</strong> there<br />

is something outside our system. As a consequence, the second law does not<br />

apply to the universe as a whole. This is actually quite important. In thermodynamics<br />

we discuss samples, we observe sample, and we are on the outside.<br />

We have large reservoirs available to set pressure or temper<strong>at</strong>ure values. Hence<br />

when we take the thermodynamic limit for the sample, we first have to take the<br />

limit for the outside world and make it infinitely large. This point will come<br />

back in st<strong>at</strong>istical mechanics, and the first to draw <strong>at</strong>tention to it was Maxwell<br />

when he deployed his demon.<br />

The definitions of he<strong>at</strong> and entropy in thermodynamics are based on quantities<br />

th<strong>at</strong> we can measure. They are oper<strong>at</strong>ional definitions. The second law is<br />

an experimental observ<strong>at</strong>ion, which has never been falsified in macroscopic experiments.<br />

Maxwell started an important discussion trying to falsify the second<br />

law on a microscopic basis (his famous demon), but th<strong>at</strong> never worked either.<br />

It did lead to important st<strong>at</strong>ements about computing, though!<br />

If the second law is universally valid, it defines a preferred direction of time<br />

(by increasing entropy or energy stored in the unusable internal variables), and<br />

seems to imply th<strong>at</strong> every system will die a he<strong>at</strong> de<strong>at</strong>h. This is not true, however,<br />

because we always invoke an outside world, and <strong>at</strong> some point he<strong>at</strong> will have<br />

to flow from the system to the outside. This is another interesting point of<br />

discussion in the philosophy of science.<br />

In st<strong>at</strong>istical mechanics we can define entropy and energy by considering<br />

the system only, and is seems possible to define the entropy of the universe in<br />

th<strong>at</strong> way. Here one has to keep in mind th<strong>at</strong> the connection between st<strong>at</strong>istical<br />

mechanics and thermodynamics has to be made, and as soon as we make th<strong>at</strong><br />

connection we invoke an outside world. This is an interesting point of deb<strong>at</strong>e,<br />

too, which takes place on the same level as the deb<strong>at</strong>e in quantum mechanics<br />

about the interpret<strong>at</strong>ion of wave functions and changes in the wave functions.<br />

Carnot engine.<br />

We have seen before th<strong>at</strong> machines th<strong>at</strong> run in cycles are useful to do work.<br />

In the following we will consider such a machine. The important aspect of the<br />

machine is th<strong>at</strong> every step is reversible. The second law leads to the important<br />

conclusion th<strong>at</strong> all reversible machines using the same process have the same<br />

efficiency. We want to make a connection with temper<strong>at</strong>ure, and therefore we<br />

define an engine with a cycle in which two parts are <strong>at</strong> constant temper<strong>at</strong>ures, in<br />

order to be able to compare values of these temper<strong>at</strong>ures. The other two parts


20 CHAPTER 1. BASIC THERMODYNAMICS.<br />

Figure 1.1: Carnot cycle in PV diagram .<br />

are simplified by making them adiab<strong>at</strong>ic, so no he<strong>at</strong> is exchanged. Connecting<br />

the workings of such engines with the second law will then allow us to define a<br />

temper<strong>at</strong>ure scale, and also define entropy.<br />

An engine is a system which changes its thermodynamic st<strong>at</strong>e in cycles and<br />

converts he<strong>at</strong> into work by doing so. A Carnot engine is any system repe<strong>at</strong>ing<br />

the following reversible cycle: (1) an isothermal expansion <strong>at</strong> a high temper<strong>at</strong>ure<br />

T1, (2) an adiab<strong>at</strong>ic expansion in which the temper<strong>at</strong>ure is lowered to T2,<br />

(3) an isothermal contraction <strong>at</strong> temper<strong>at</strong>ure T2, and finally (4) an adiab<strong>at</strong>ic<br />

contraction back to the initial st<strong>at</strong>e. In this case work is done using a change<br />

in volume. Similar Carnot engines can be defined for all other types of work. It<br />

is easiest to talk about Carnot engines using the pressure p, the volume V, and<br />

the temper<strong>at</strong>ure T as variables. A diagram of a Carnot engine in the pV plane<br />

is shown in figure 1.1.<br />

The m<strong>at</strong>erial in a Carnot engine can be anything. For practical reasons it<br />

is often a gas. Also, because steps one and three are isothermal, contact with<br />

a he<strong>at</strong> reservoir is required, and the Carnot engine oper<strong>at</strong>es between these two<br />

he<strong>at</strong> reservoirs, by definition. Mechanical work is done in all four parts of the<br />

cycle. We can define Carnot engines for any type of work, but mechanical work is<br />

the easiest to visualize (and construction of Carnot engines based on mechanical<br />

work is also most common). One also recognizes the historical importance of<br />

steam engines; such engines ”drove” the development of thermodynamics!<br />

Carnot engines are the most efficient!<br />

The second law of thermodynamics has a very important consequence for<br />

Carnot engines. One can show th<strong>at</strong> a Carnot engine is the most efficient engine<br />

oper<strong>at</strong>ing between two reservoirs <strong>at</strong> temper<strong>at</strong>ure T1 and T2! This is a very strong<br />

st<strong>at</strong>ement, based on minimal inform<strong>at</strong>ion. The efficiency η is the r<strong>at</strong>io of the<br />

work W performed on the outside world and the he<strong>at</strong> Q1 absorbed by the system


Ì É<br />

1.5. SECOND LAW: ENTROPY. 21<br />

Ï<br />

Ì É<br />

Figure 1.2: Schem<strong>at</strong>ics of a Carnot engine.<br />

in the isothermal step one <strong>at</strong> high temper<strong>at</strong>ure. Remember th<strong>at</strong> in steps two<br />

and four no he<strong>at</strong> is exchanged. The he<strong>at</strong> absorbed from the reservoir <strong>at</strong> low<br />

temper<strong>at</strong>ure in step three is Q2 and the first law tells us th<strong>at</strong> W = Q1 + Q2.<br />

We define the flow of he<strong>at</strong> Qi to be positive when he<strong>at</strong> flows into the system.<br />

In most engines we will, of course, have Q1 > 0 and Q2 < 0. This gives us<br />

η = W<br />

= 1 +<br />

Q1<br />

Q2<br />

Q1<br />

(1.19)<br />

Work is positive when it represents a flow of energy to the outside world. A<br />

Carnot engine in reverse is a he<strong>at</strong>er (or refriger<strong>at</strong>or depending on which reservoir<br />

you look <strong>at</strong>).<br />

Can the efficiency be gre<strong>at</strong>er than one?<br />

A Carnot engine can be represented in as follows, see figure 1.2. In this<br />

figure the arrows point in the direction in which the energy flow is defined to<br />

be positive.<br />

Equivalency of Clausius and Kelvin.<br />

The two formul<strong>at</strong>ions of the second law of Clausius and Kelvin are equivalent.<br />

If a Kelvin engine existed which converts he<strong>at</strong> completely into work, this work<br />

can be transformed into he<strong>at</strong> dumped into a reservoir <strong>at</strong> higher temper<strong>at</strong>ure, in<br />

contradiction with Clausius. If a Clausius process would exist, we can use it to<br />

store energy <strong>at</strong> a higher temper<strong>at</strong>ure. A normal engine would take this amount<br />

of he<strong>at</strong>, dump he<strong>at</strong> <strong>at</strong> the low temper<strong>at</strong>ure again while performing work, and


Ì<br />

22 CHAPTER 1. BASIC THERMODYNAMICS.<br />

Ì É<br />

É ÏÏ É Ï<br />

<br />

É<br />

Figure 1.3: Two engines feeding eachother.<br />

there would be a contradiction with Kelvin’s formul<strong>at</strong>ion of the second law. The<br />

st<strong>at</strong>ement about Carnot engines is next shown to be true in a similar way.<br />

Contradictions if existence of more efficient engine.<br />

Assume th<strong>at</strong> we have an engine X which is more efficient than a Carnot<br />

engine C. We will use this engine X to drive a Carnot engine in reverse, see<br />

figure 1.3. The engine X takes an amount of he<strong>at</strong> QX > 0 from a reservoir <strong>at</strong><br />

high temper<strong>at</strong>ure. It produces an amount of work W = ηXQX > 0 and takes<br />

an amount of he<strong>at</strong> Q2X = (ηX − 1)QX from the reservoir <strong>at</strong> low temper<strong>at</strong>ure.<br />

Notice th<strong>at</strong> we need ηX < 1, (and hence Q2X < 0 ), otherwise we would<br />

viol<strong>at</strong>e Kelvin’s formul<strong>at</strong>ion of the second law. This means th<strong>at</strong> the net flow<br />

of he<strong>at</strong> is towards the reservoir of low temper<strong>at</strong>ure. Now take a Carnot engine<br />

oper<strong>at</strong>ing between the same two reservoirs. This Carnot engine is driven by<br />

the amount of work W, hence the amount of work performed by the Carnot<br />

engine is WC = −W . This Carnot engine takes an amount of he<strong>at</strong> Q1C =<br />

WC ηX = − ηC ηC QX from the reservoir <strong>at</strong> high temper<strong>at</strong>ure and an amount Q2C =<br />

WC − Q1C = ( 1<br />

ηC − 1)ηXQX from the reservoir <strong>at</strong> low temper<strong>at</strong>ure. Now<br />

consider the combin<strong>at</strong>ion of these two engines. This is a machine which takes<br />

an amount of he<strong>at</strong> (1 − ηx<br />

ηc )Qx from the reservoir <strong>at</strong> high temper<strong>at</strong>ure and the<br />

opposite amount from the reservoir <strong>at</strong> low temper<strong>at</strong>ure. Energy is conserved,<br />

but Clausius tells us th<strong>at</strong> the amount of he<strong>at</strong> taken from the high temper<strong>at</strong>ure<br />

reservoir should be positive, or ηx ≤ ηc. Hence a Carnot engine is the most<br />

efficient engine which one can construct!<br />

In a different proof we can combine an engine X and a Carnot engine, but<br />

require Q2X + Q2C = 0. Such an engine produces an amount of work Wnet<br />

which has to be neg<strong>at</strong>ive according to Kelvin.


1.5. SECOND LAW: ENTROPY. 23<br />

Show th<strong>at</strong> this implies ηX ≤ ηC.<br />

All Carnot engines are equally efficient.<br />

One can easily show th<strong>at</strong> all Carnot engines have the same efficiency. Suppose<br />

the efficiencies of Carnot engine one and two are η1 and η2, respectively.<br />

Use one Carnot engine to drive the other in reverse, and it follows th<strong>at</strong> we need<br />

η1 ≤ η2 and also η2 ≤ η1, or η1 = η2. Hence the efficiency of an arbitrary<br />

Carnot engine is ηC. This is independent of the details of the Carnot engine,<br />

except th<strong>at</strong> it should oper<strong>at</strong>e between a reservoir <strong>at</strong> T1 and a reservoir <strong>at</strong> T2.<br />

These are the only two variables which play a role, and the Carnot efficiency<br />

should depend on them only: ηC(T1, T2).<br />

Carnot efficiency can be measured experimentally.<br />

This efficiency function can be determined experimentally by measuring Q<br />

and W flowing in and out of a given Carnot engine. How? Th<strong>at</strong> is a problem.<br />

First, consider the work done. This is the easier part. For example, because of<br />

the work done a weight is lifted a certain distance. This gives us the change<br />

in energy, and hence the work done. In order to use this type of measurement,<br />

however, we need to know details about the type of work. This is essentially<br />

the same as saying th<strong>at</strong> we need to understand the measurements we are doing.<br />

How do we measure he<strong>at</strong>? We need a reference. For example, take a large<br />

closed glass container with w<strong>at</strong>er and ice, initially in a one to one r<strong>at</strong>io. Assume<br />

th<strong>at</strong> the amount of energy to melt a unit mass of ice is our basic energy value.<br />

We can measure the amount of he<strong>at</strong> th<strong>at</strong> went into this reference system by<br />

measuring the change in the volumes of w<strong>at</strong>er and ice. Also, if a sample of<br />

unknown temper<strong>at</strong>ure is brought into contact with the reference system, we can<br />

easily determine whether the temper<strong>at</strong>ure of the sample is higher or lower then<br />

the reference temper<strong>at</strong>ure of the w<strong>at</strong>er and ice system. If it is higher, ice will<br />

melt, if it is lower, w<strong>at</strong>er will freeze. Note th<strong>at</strong> we assume th<strong>at</strong> the temper<strong>at</strong>ure<br />

of the reference system is positive!<br />

Experimental definition of temper<strong>at</strong>ure.<br />

St<strong>at</strong>e variables are average, macroscopic quantities of a system which can be<br />

measured. This is certainly a good definition of variables like volume, pressure,<br />

and number of particles. They are rel<strong>at</strong>ed to basic concepts like length, mass,<br />

charge, and time. Temper<strong>at</strong>ure is a different quantity, however. A practical<br />

definition of the temper<strong>at</strong>ure of an object is via a thermometer. The active<br />

substance in the thermometer could be mercury or some ideal gas. But those<br />

are definitions which already incorpor<strong>at</strong>e some physics, like the linear expansion<br />

of solids for mercury or the ideal gas law for a gas. It is different from the


24 CHAPTER 1. BASIC THERMODYNAMICS.<br />

definitions of length and time in terms of the standard meter and clock. In a<br />

similar vein we would like to define temper<strong>at</strong>ure as the result of a measurement<br />

of a comparison with a standard. Hence we assume th<strong>at</strong> we have a known object<br />

of temper<strong>at</strong>ure T0, similar to the standard meter and clock. An example would<br />

be the container with the w<strong>at</strong>er and ice mixture mentioned above.<br />

Now how do we compare temper<strong>at</strong>ures on a quantit<strong>at</strong>ive level? If we want<br />

to find the temper<strong>at</strong>ure of an object of unknown temper<strong>at</strong>ure T, we take a<br />

Carnot engine and oper<strong>at</strong>e th<strong>at</strong> engine between the object and the standard. We<br />

measure the amount of he<strong>at</strong> Q flowing from the reference system to the Carnot<br />

engine or from the Carnot engine to the reference system. We also measure<br />

the amount of work W done by the Carnot engine. We use the first law to<br />

determine the amount of he<strong>at</strong> flowing out of the high temper<strong>at</strong>ure reservoir, if<br />

needed. The r<strong>at</strong>io of these two quantities is the efficiency of the Carnot engine,<br />

which only depends on the two temper<strong>at</strong>ures.<br />

We first determine if the object has a higher or lower temper<strong>at</strong>ure then<br />

the reference by bringing them in direct contact. If ice melts, the object was<br />

warmer, if ice forms, it was colder.<br />

If the temper<strong>at</strong>ure of the reference system is higher th<strong>at</strong> the temper<strong>at</strong>ure of<br />

the object, we use the reference system as the high temper<strong>at</strong>ure reservoir. We<br />

measure the amount of he<strong>at</strong> Q going out of the reference system and find:<br />

ηC = W<br />

Q<br />

(1.20)<br />

If the temper<strong>at</strong>ure of the reference system is lower th<strong>at</strong> the temper<strong>at</strong>ure of<br />

the object, we use the reference system as the low temper<strong>at</strong>ure reservoir. We<br />

measure the amount of he<strong>at</strong> Q going out of the reference system, which is now<br />

neg<strong>at</strong>ive, since he<strong>at</strong> is actually going in, and find:<br />

ηC = W<br />

W − Q<br />

In the first case we assign a temper<strong>at</strong>ure T to the object according to<br />

T<br />

T0<br />

and in the second case according to<br />

(1.21)<br />

= (1 − ηC) (1.22)<br />

T0<br />

T = (1 − ηC) (1.23)<br />

This is our definition of temper<strong>at</strong>ure on the Carnot scale. It is an important<br />

step forward, based on the unique efficiency of Carnot engines, which in<br />

itself is based on the second law. Theoretically, this is a good definition because<br />

Carnot engines are well-defined. Also, energy is well-defined. The important<br />

question, of course, is how this definition rel<strong>at</strong>es to known temper<strong>at</strong>ure scales.<br />

We will rel<strong>at</strong>e the Carnot temper<strong>at</strong>ure scale to the ideal gas temper<strong>at</strong>ure scale<br />

<strong>at</strong> the end of this chapter.


Ì<br />

1.5. SECOND LAW: ENTROPY.<br />

É<br />

25<br />

Ï<br />

Ì<br />

<br />

Ì É<br />

É<br />

Ï<br />

Figure 1.4: Two Carnot engines in series.<br />

Efficiency for arbitrary temper<strong>at</strong>ures.<br />

We can analyze the general situ<strong>at</strong>ion for a Carnot engine between arbitrary<br />

temper<strong>at</strong>ures as follows. Assume th<strong>at</strong> we have T > T0 > T ′ , all other cases<br />

work similarly. Consider the following couple of Carnot engines (see figure 1.4<br />

) and demand th<strong>at</strong> Q ′ 1 + Q2 = 0 (no he<strong>at</strong> going in or out the reference system).<br />

Argue th<strong>at</strong> this is equivalent to a single Carnot engine working between T and<br />

T ′ .<br />

′<br />

T For this system we have T0 = (1 − η′ T0<br />

C ) and T = (1 − ηC),<br />

′<br />

T or T = (1 −<br />

η ′ C )(1 − ηC). The efficiencies can be expressed in the energy exchanges and we<br />

′<br />

′<br />

T W have T = (1 − Q ′ )(1 −<br />

1<br />

W<br />

Q1 ). But we have Q′ 1 = −Q2 = Q1 − W and hence<br />

T ′<br />

′<br />

′<br />

W W<br />

W W<br />

T = (1 − Q1−W )(1 − ). The right hand side is equal to 1 − − Q1 Q1 Q1−W (1 −<br />

′<br />

W<br />

W W ) = 1 − − . In other words:<br />

Q1 Q1 Q1<br />

T ′<br />

T<br />

′ W + W<br />

= 1 − = 1 − ηC<br />

where the rel<strong>at</strong>ion now holds for arbitrary values of the temper<strong>at</strong>ure.<br />

Carnot cycle again.<br />

Q1<br />

Can we obtain neg<strong>at</strong>ive values of the temper<strong>at</strong>ure?<br />

(1.24)


26 CHAPTER 1. BASIC THERMODYNAMICS.<br />

Using the temper<strong>at</strong>ure scale defined by the Carnot engine, we can reanalyze<br />

the Carnot cycle. The efficiency is rel<strong>at</strong>ed to the he<strong>at</strong> ∆Q1 absorbed in the first<br />

step and ∆Q2 absorbed in the third step (which is neg<strong>at</strong>ive in an engine) by<br />

ηC = 1 + ∆Q2<br />

(1.25)<br />

∆Q1<br />

where we have used a not<strong>at</strong>ion with ∆Q to emphasize the fact th<strong>at</strong> we look<br />

<strong>at</strong> changes. But th<strong>at</strong> is not really essential. From the previous equ<strong>at</strong>ion we find:<br />

∆Q1<br />

+<br />

T1<br />

∆Q2<br />

= 0 (1.26)<br />

T2<br />

Since there is no he<strong>at</strong> exchanged in steps two and four of the Carnot cycle, this<br />

is equivalent to<br />

<br />

C<br />

¯dQ<br />

T<br />

= 0 (1.27)<br />

where the closed contour C specifies the p<strong>at</strong>h of integr<strong>at</strong>ion in the space of st<strong>at</strong>e<br />

variables.<br />

Integral for arbitrary cycles.<br />

Next we consider the combined effect of two Carnot engines, one working<br />

between T1 and T2, the other one between T2 and T3. Now compare this with a<br />

single system which follows the thermodynamic transform<strong>at</strong>ion defined by the<br />

outside of the sum of the two Carnot contours. One can think of the total process<br />

as the sum of the two Carnot steps, introducing an intermedi<strong>at</strong>e reservoir, in<br />

which no net he<strong>at</strong> is deposited. The contour integral of ¯ dQ<br />

T is also zero for the<br />

single process, since the two contributions over the common line are opposite<br />

and cancel. Any general closed p<strong>at</strong>h in the space of st<strong>at</strong>e variables, restricted to<br />

those surfaces which are allowed by the equ<strong>at</strong>ions of st<strong>at</strong>e, can be approxim<strong>at</strong>ed<br />

as the sum of a number of Carnot cycles with temper<strong>at</strong>ure difference ∆T . The<br />

error in this approxim<strong>at</strong>ion approaches zero for ∆T → 0. Hence:<br />

<br />

¯dQ<br />

= 0 (1.28)<br />

T<br />

where R is an arbitrary cyclic, reversible process.<br />

Definition of entropy.<br />

Formula 1.28 has the important consequence th<strong>at</strong><br />

We define a new variable S by<br />

R<br />

∫ 2<br />

S2 = S1 +<br />

1<br />

¯dQ<br />

T<br />

2∫<br />

1<br />

¯dQ<br />

T<br />

is p<strong>at</strong>h independent.<br />

(1.29)


1.5. SECOND LAW: ENTROPY. 27<br />

and because the integral is p<strong>at</strong>h independent S is a st<strong>at</strong>e function. When the<br />

integr<strong>at</strong>ion points are close together we get<br />

T dS = ¯ dQ (1.30)<br />

in which dS is an exact differential. The quantity S is called the entropy. In<br />

thermodynamics we define the entropy from a purely macroscopic point of view.<br />

It is rel<strong>at</strong>ed to infinitesimally small exchanges of thermal energy by requiring<br />

th<strong>at</strong> the differential ¯ dQ can be transformed into an exact differential by multiplying<br />

it with a function of the temper<strong>at</strong>ure alone. One can always transform a<br />

differential into an exact differential by multiplying it with a function of all st<strong>at</strong>e<br />

variables. In fact, there are an infinite number of ways to do this. The restriction<br />

th<strong>at</strong> the multiplying factor only depends on temper<strong>at</strong>ure uniquely defines<br />

this factor, apart from a constant factor. One could also define 5T dS = ¯ dQ ,<br />

which would simply re-scale all temper<strong>at</strong>ure values by a factor five.<br />

First law in exact differentials.<br />

The first law of thermodynamics in terms of changes in the entropy is<br />

dU = T dS − ¯ dW (1.31)<br />

For example, if we consider a system where the only interactions with the outside<br />

world are a possible exchange of he<strong>at</strong> and mechanical work, changes in the<br />

internal energy are rel<strong>at</strong>ed to changes in the entropy and volume through<br />

dU = T dS − pdV (1.32)<br />

An important note <strong>at</strong> this point is th<strong>at</strong> we often use the equ<strong>at</strong>ion 1.32 as a<br />

model. It does indeed exemplify some basic concepts, but for real applic<strong>at</strong>ions<br />

it is too simple. The simplest form of the first law th<strong>at</strong> has physical meaning is<br />

the following:<br />

Entropy is extensive.<br />

dU = T dS − pdV + µdN (1.33)<br />

In the definition of the Carnot temper<strong>at</strong>ure of an object the size of the object<br />

does not play a role, only the fact th<strong>at</strong> the object is in thermal equilibrium. As<br />

a consequence the temper<strong>at</strong>ure is an intensive quantity. On the other hand, if<br />

we compare the he<strong>at</strong> absorbed by a system during a thermodynamic process<br />

with the he<strong>at</strong> absorbed by a similar system which is α times larger, it is not<br />

hard to argue th<strong>at</strong> the amount of he<strong>at</strong> exchanged is α times larger as well. As<br />

a consequence, the entropy S is an extensive st<strong>at</strong>e variable.


28 CHAPTER 1. BASIC THERMODYNAMICS.<br />

N<strong>at</strong>ural variables for the internal energy are all extensive st<strong>at</strong>e<br />

variables.<br />

Changes in the internal energy U are rel<strong>at</strong>ed to changes in the extensive<br />

st<strong>at</strong>e variables only, since the amount of work done is determined by changes<br />

in extensive st<strong>at</strong>e variables only and S is extensive. In this sense, the n<strong>at</strong>ural<br />

set of variables for the st<strong>at</strong>e function U is the set of all extensive variables. By<br />

n<strong>at</strong>ural we mean the set of variables th<strong>at</strong> show as differentials in the first law,<br />

hence small changes are directly rel<strong>at</strong>ed.<br />

Importance of thermodynamic limit.<br />

Equ<strong>at</strong>ion 1.31 has an interesting consequence. Suppose th<strong>at</strong> we decide th<strong>at</strong><br />

another extensive st<strong>at</strong>e variable is needed to describe the st<strong>at</strong>e of a system.<br />

Hence we are adding a term xdX to ¯ dW . This means th<strong>at</strong> the number of internal<br />

degrees of freedom is reduced by one, since we are specifying one additional<br />

combin<strong>at</strong>ion of degrees of freedom via X. This in its turn indic<strong>at</strong>es th<strong>at</strong> the<br />

entropy should change, since it is a represent<strong>at</strong>ion of the internal degrees of<br />

freedom of the system. The definition of the entropy would therefore depend<br />

on the definition of work, which is an unacceptable situ<strong>at</strong>ion. Fortun<strong>at</strong>ely, the<br />

thermodynamic limit comes to rescue here. Only when the number of degrees of<br />

freedom is infinitely large, the change by one will not alter the entropy. Hence<br />

the entropy is only well-defined in the thermodynamic limit.<br />

Independent and dependent variables.<br />

and<br />

From equ<strong>at</strong>ion 1.32 we find immedi<strong>at</strong>ely th<strong>at</strong><br />

( )<br />

∂U<br />

T =<br />

∂S V<br />

p = −<br />

( )<br />

∂U<br />

∂V S<br />

(1.34)<br />

(1.35)<br />

which shows th<strong>at</strong> in the set of variables p,V,T,S only two are independent. If<br />

we know the basic physics of the system, we know the st<strong>at</strong>e function U(S, V ) and<br />

can derive the values for p and T according to the two st<strong>at</strong>e functions defined by<br />

the partial deriv<strong>at</strong>ives. Functions of the form T = f(S, V ) and p = g(S, V ) are<br />

called equ<strong>at</strong>ions of st<strong>at</strong>e. More useful forms elimin<strong>at</strong>e the entropy from these<br />

equ<strong>at</strong>ions and lead to equ<strong>at</strong>ions of st<strong>at</strong>e of the form p = h(T, V ). The rel<strong>at</strong>ion<br />

U = u(S, V ) is called an energy equ<strong>at</strong>ion and is not an equ<strong>at</strong>ion of st<strong>at</strong>e, since<br />

it defines an energy as a function of the independent st<strong>at</strong>e variables. Such<br />

rel<strong>at</strong>ions are the basis for equ<strong>at</strong>ions of st<strong>at</strong>e, but we use equ<strong>at</strong>ion of st<strong>at</strong>e only<br />

when we describe st<strong>at</strong>e variables th<strong>at</strong> occur in pairs, like T and S, or p and V .


1.5. SECOND LAW: ENTROPY. 29<br />

Equ<strong>at</strong>ions of st<strong>at</strong>e give dependent st<strong>at</strong>e variables of this n<strong>at</strong>ure as a function of<br />

independent ones.<br />

In the next chapter we will discuss how to change variables and make combin<strong>at</strong>ions<br />

like T and V independent and the others dependent.<br />

Entropy in terms of work.<br />

We now return to the definition of entropy according to equ<strong>at</strong>ion 1.29 above.<br />

If we apply the first law we get<br />

∫ 2<br />

dU +<br />

S2 = S1 +<br />

1<br />

¯ dW<br />

T<br />

and if we express work in its generalized form according to 1.18 we see<br />

∫ 2<br />

S2 = S1 +<br />

1<br />

dU + ∑<br />

j xjdXj<br />

T<br />

(1.36)<br />

(1.37)<br />

which shows th<strong>at</strong> entropy is defined based on basic properties of the system,<br />

which can be directly measured. There is no device th<strong>at</strong> measures entropy<br />

directly, and in th<strong>at</strong> sense it is different from all other st<strong>at</strong>e variables. But<br />

it is possible to perform processes in which the entropy does not change, and<br />

hence we do have control over changes in the entropy. Since entropy is based<br />

on changes in other st<strong>at</strong>e variables and the internal energy, it is well defined<br />

in reversible processes. The second law singles out he<strong>at</strong> as a more restricted<br />

change of energy, and this has consequences for entropy. We now discuss those<br />

consequences.<br />

Change in entropy in an irreversible process.<br />

Up to this point we only considered the entropy in connection with reversible<br />

processes. Different rules follow for irreversible processes. Consider a general<br />

process in which he<strong>at</strong> is transferred from a reservoir <strong>at</strong> high temper<strong>at</strong>ure to a<br />

reservoir <strong>at</strong> low temper<strong>at</strong>ure (and hence Q1 > 0). The efficiency of this process<br />

is <strong>at</strong> most equal to the Carnot efficiency, and hence<br />

W<br />

Q1<br />

For such a general process we have<br />

= 1 + Q2<br />

≤ ηC = 1 −<br />

Q1<br />

T2<br />

T1<br />

Q1<br />

T1<br />

+ Q2<br />

T2<br />

(1.38)<br />

≤ 0 (1.39)<br />

Suppose th<strong>at</strong> the he<strong>at</strong> is taken from the reservoir <strong>at</strong> high temper<strong>at</strong>ure T1<br />

in a general process, but dumped in the reservoir <strong>at</strong> low temper<strong>at</strong>ure T2 in a<br />

reversible way. Suppose we only consider small changes in energy. For comparison,<br />

consider a completely reversible way of transferring the he<strong>at</strong> from one


30 CHAPTER 1. BASIC THERMODYNAMICS.<br />

reservoir to another. In this process he<strong>at</strong> is transferred reversibly from the high<br />

temper<strong>at</strong>ure reservoir to the engine. This leads to<br />

∆Q g<br />

1<br />

T1<br />

≤ − ∆Qrev 2<br />

T2<br />

= ∆Qrev 1<br />

T1<br />

= ∆S1<br />

(1.40)<br />

For a general process <strong>at</strong> a temper<strong>at</strong>ure T, in which the amount of he<strong>at</strong> added<br />

to the system is ∆Q and the change in entropy of the system is ∆S, the changes<br />

are rel<strong>at</strong>ed by<br />

T ∆S ≥ ∆Q (1.41)<br />

The equal sign holds for all reversible processes only. Hence in an irreversible<br />

process the change in entropy accompanying the absorption of a certain amount<br />

of he<strong>at</strong> is larger than necessary. The change in entropy is minimal for a reversible<br />

process. An engine in which all processes are reversible will return to its initial<br />

st<strong>at</strong>e without a change in entropy. If, however, some processes are not reversible,<br />

the entropy of the engine will increase after a complete cycle. It does not return<br />

to the same initial st<strong>at</strong>e, but to a st<strong>at</strong>e of higher entropy. The internal energy<br />

of the system decreases more than necessary, and we lose energy, which can be<br />

seen from<br />

∆U = ∆Q − ∆W ≤ T ∆S − ∆W (1.42)<br />

The price in energy we pay to transform a system from one st<strong>at</strong>e to another<br />

is higher than needed if the transform<strong>at</strong>ion is irreversible.<br />

Maximum entropy principle.<br />

If we wait long enough, a completely isol<strong>at</strong>ed system will end up in equilibrium.<br />

If this system was not in equilibrium to start with, a spontaneous process<br />

will take it there. Since the system is completely isol<strong>at</strong>ed, there is no exchange<br />

of he<strong>at</strong> with the outside world, and <strong>at</strong> each time T ∆S ≥ 0. In reaching equilibrium<br />

by a spontaneous process, the entropy of the system has increased. This<br />

is true for any non-equilibrium initial st<strong>at</strong>e. Hence we find the following very<br />

important rule:<br />

The equilibrium st<strong>at</strong>e of a completely isol<strong>at</strong>ed system has maximal<br />

entropy.<br />

By completely isol<strong>at</strong>ed system we mean th<strong>at</strong> the internal energy U, the volume<br />

V, the amount of m<strong>at</strong>erial N, and all other extensive st<strong>at</strong>e variables are kept<br />

constant. The way to change the entropy is to change some internal variables,<br />

hence the entropy is a maximum as a function of the internal variables only.<br />

This is quite a strong st<strong>at</strong>ement based on some minimal input (the second law<br />

of thermodynamics).


1.6. THIRD LAW OF THERMODYNAMICS. 31<br />

1.6 Third law of thermodynamics.<br />

Using equ<strong>at</strong>ion 1.29 we can find the entropy difference between two arbitrary<br />

st<strong>at</strong>es. This still leaves us with one constant of integr<strong>at</strong>ion, which could be<br />

system dependent. The third law of thermodynamics sets this constant to zero<br />

<strong>at</strong> zero temper<strong>at</strong>ure for all systems. The short form of the third law is<br />

S(T = 0) = 0 (1.43)<br />

but this is not completely correct. First of all, we cannot reach zero temper<strong>at</strong>ure,<br />

and second we need to take the thermodynamic limit. Therefore, the correct<br />

formul<strong>at</strong>ion of the third law is:<br />

lim<br />

N→∞ lim<br />

S(T )<br />

= 0 (1.44)<br />

T →0 N<br />

Note th<strong>at</strong> we cannot interchange the order of the limits, we have to take the limit<br />

for the temper<strong>at</strong>ure first. This is a general characteristic of the thermodynamic<br />

limit, we have to take this limit last, after we have done all our calcul<strong>at</strong>ions!<br />

This is an important point to remember, and a source of errors if done wrong.<br />

One very important aspect of the third law is th<strong>at</strong> the value of the entropy is<br />

zero (in the sense discussed above) <strong>at</strong> zero temper<strong>at</strong>ure, no m<strong>at</strong>ter which values<br />

the other st<strong>at</strong>e variables have. Th<strong>at</strong> is quite restrictive. Another aspect of the<br />

third law is th<strong>at</strong> it does not tell us how close to zero the temper<strong>at</strong>ure has to<br />

be. There are systems in which the entropy appears to approach a non-zero<br />

value, until we reach a very low temper<strong>at</strong>ure, <strong>at</strong> which point the entropy does<br />

suddenly drop.<br />

Response <strong>at</strong> zero temper<strong>at</strong>ure.<br />

At zero temper<strong>at</strong>ure the entropy becomes infinitesimally small compared to<br />

the number of degrees of freedom in the thermodynamic limit. In the language of<br />

quantum mechanics, the degeneracy of the ground st<strong>at</strong>e is very small compared<br />

to the number of degrees of freedom. The third law seems to hold for all physical<br />

systems. It has an interesting consequence. Suppose we change the st<strong>at</strong>e of a<br />

system by changing the st<strong>at</strong>e variable X, but keep the temper<strong>at</strong>ure <strong>at</strong> zero.<br />

Since S(T = 0, X) = 0, we find th<strong>at</strong> the partial deriv<strong>at</strong>ive of S with respect to<br />

X has to be 0, or<br />

lim<br />

N→∞ lim<br />

T →0<br />

( )<br />

1 ∂S<br />

= 0 (1.45)<br />

N ∂X T<br />

This seems to hold for all experiments, and has important consequences for the<br />

values of response functions <strong>at</strong> low values of the temper<strong>at</strong>ure. It also allows us<br />

to check if we reached low enough temper<strong>at</strong>ures. If the entropy seems constant,<br />

but partial deriv<strong>at</strong>ives like the ones above are not zero, we are not <strong>at</strong> sufficiently<br />

low temper<strong>at</strong>ures yet.


32 CHAPTER 1. BASIC THERMODYNAMICS.<br />

He<strong>at</strong> capacity <strong>at</strong> zero temper<strong>at</strong>ure.<br />

An example of a simple system is a system where the only thermodynamic<br />

variables are S, T, p, and V. The internal energy follows from dU = T dS −pdV .<br />

If we work <strong>at</strong> constant volume, dV = 0, and the first law leads to<br />

S(V, T1) − S(V, T2) =<br />

∫ T1<br />

T2<br />

dU<br />

T =<br />

∫ T1<br />

dT<br />

T2 T<br />

( )<br />

∂U<br />

∂T V<br />

(1.46)<br />

Since according to the third law the left hand side exist for T2 → 0, the integral<br />

on the right hand side must exist, which implies<br />

lim<br />

T →0<br />

This is also observed experimentally.<br />

T=0 unreachable.<br />

( )<br />

∂U<br />

∂T V<br />

= 0 (1.47)<br />

Another consequence of the third law is the impossibility of reaching T = 0K<br />

in a finite number of steps using a reversible process. Problem 7 illustr<strong>at</strong>es this<br />

with an example.<br />

1.7 Ideal gas and temper<strong>at</strong>ure.<br />

There are several temper<strong>at</strong>ure scales used in daily life, like Celcius, Fahrenheit,<br />

Reamur, and Kelvin. In thermodynamics we define the temper<strong>at</strong>ure first, via<br />

an experiment. In st<strong>at</strong>istical mechanics one defines the entropy first (technically<br />

such a definition is called an entropy analogue), and derives the temper<strong>at</strong>ure<br />

from this. Since we cannot measure the entropy directly this is not a good<br />

oper<strong>at</strong>ional definition.<br />

Equivalency of temper<strong>at</strong>ure scales.<br />

There are many ways to measure temper<strong>at</strong>ure, and experimentally they have<br />

all been shown to be equivalent. Therefore, we need only to equ<strong>at</strong>e the formal<br />

definition of temper<strong>at</strong>ure via Carnot engines to one type of measurement in order<br />

to show th<strong>at</strong> the Carnot temper<strong>at</strong>ure is really the temper<strong>at</strong>ure we measure in<br />

real experiments.<br />

Ideal gas temper<strong>at</strong>ure.<br />

One very important standard of temper<strong>at</strong>ure is based on the ideal gas. The<br />

energy equ<strong>at</strong>ion and the equ<strong>at</strong>ion of st<strong>at</strong>e for a mono<strong>at</strong>omic ideal gas are


1.7. IDEAL GAS AND TEMPERATURE. 33<br />

and<br />

U = 3<br />

pV (1.48)<br />

2<br />

pV = NRT id<br />

(1.49)<br />

The second equ<strong>at</strong>ion of st<strong>at</strong>e is used to define the ideal gas temper<strong>at</strong>ure. If<br />

a known amount of an ideal gas is in equilibrium we measure its volume and<br />

pressure, and then evalu<strong>at</strong>e the temper<strong>at</strong>ure.<br />

Ideal gas cannot be <strong>at</strong> zero temper<strong>at</strong>ure.<br />

Note th<strong>at</strong> for the ideal gas we have U = 3<br />

2NRT , and hence ( )<br />

∂U<br />

∂T V<br />

= 3<br />

2 NR.<br />

This does not approach zero in the limit T → 0. The ideal gas law U = 3<br />

2 pV<br />

is therefore in contradiction with the third law. The ideal gas law cannot be<br />

applied <strong>at</strong> low temper<strong>at</strong>ures. This is not surprising, since all m<strong>at</strong>erials transform<br />

to a liquid <strong>at</strong> sufficiently low temper<strong>at</strong>ures.<br />

Carnot cycle for ideal gas.<br />

We now construct a Carnot engine based on an ideal gas. The equ<strong>at</strong>ions<br />

of st<strong>at</strong>e show th<strong>at</strong> the internal energy of the ideal gas is determined by the<br />

ideal gas temper<strong>at</strong>ure only, and hence does not change in an isothermal process.<br />

Therefore in steps one and three of the Carnot cycle all the he<strong>at</strong> absorbed by<br />

the system is changed into work. This is not in contradiction with the second<br />

law, because in each separ<strong>at</strong>e step of a Carnot cycle the system also changes.<br />

Note the word ”only” in the second law! The Carnot cycle is given by<br />

(p1, V1, T id<br />

1 ) → (p2, V2, T<br />

one id<br />

1 ) → (p3, V3, T<br />

two id<br />

2 ) → (p4, V4, T<br />

three id<br />

2 ) → (p1, V1, T<br />

four id<br />

1 )<br />

(1.50)<br />

During steps one and three the volume and pressure are rel<strong>at</strong>ed via the<br />

equ<strong>at</strong>ion of st<strong>at</strong>e, and the first law then st<strong>at</strong>es th<strong>at</strong> the work done is equal to<br />

the he<strong>at</strong> absorbed, since the energy is constant:<br />

∆W12 = ∆Q12 =<br />

∆W34 = ∆Q34 =<br />

∫ V2<br />

V1<br />

∫ V4<br />

V3<br />

pdV = NRT id<br />

∫ V2 dV<br />

1<br />

V1 V<br />

pdV = NRT id<br />

∫ V4<br />

dV<br />

2<br />

V3 V<br />

= NRT id<br />

= NRT id<br />

1 ln( V2<br />

V1<br />

2 ln( V4<br />

V3<br />

) (1.51)<br />

) (1.52)<br />

Steps two and four are adiab<strong>at</strong>ic, and there is no he<strong>at</strong> exchange. The change<br />

in energy is the work done by the system.


34 CHAPTER 1. BASIC THERMODYNAMICS.<br />

∆W23 = 3 id<br />

NR(T1 − T<br />

2 id<br />

2 ) = −∆W41<br />

(1.53)<br />

∆Q23 = ∆Q41 = 0 (1.54)<br />

We are now able to calcul<strong>at</strong>e the efficiency of this engine, which is wh<strong>at</strong> we<br />

need in order to compare the temper<strong>at</strong>ure of the ideal gas with the theoretical<br />

temper<strong>at</strong>ure. The efficiency is given by:<br />

which leads to<br />

Connecting the elements.<br />

η = ∆W12 + ∆W23 + ∆W34 + ∆W41<br />

∆Q12<br />

η = 1 +<br />

id T2 T id<br />

ln(<br />

1<br />

V4<br />

V3 )<br />

ln( V2<br />

V1 )<br />

(1.55)<br />

(1.56)<br />

In order to analyze this further, we need to calcul<strong>at</strong>e the entropy of an ideal<br />

gas, which is constant during the adiab<strong>at</strong>ic steps. This will then allow us to<br />

connect the values of the volume in st<strong>at</strong>es 2 and 3. Using the two equ<strong>at</strong>ions<br />

describing the ideal gas and the first law in terms of the entropy and volume<br />

only, we find th<strong>at</strong> small changes in the entropy are rel<strong>at</strong>ed to changes in the<br />

temper<strong>at</strong>ure and volume by<br />

T id dS = 3<br />

3<br />

d(pV ) + pdV =<br />

2 2 NRdT id +<br />

id NRT<br />

dV (1.57)<br />

V<br />

After dividing both sides by T id , it is not hard to find th<strong>at</strong> one can integr<strong>at</strong>e<br />

dS. The entropy for an ideal gas is<br />

S(T id , V, N) = NRln((T id ) 3<br />

2 V ) + S0(N) (1.58)<br />

Changes in the entropy are rel<strong>at</strong>ed to he<strong>at</strong> exchange. In an adiab<strong>at</strong>ic process<br />

there is no he<strong>at</strong> exchange, and hence an adiab<strong>at</strong>ic process is a process <strong>at</strong> constant<br />

entropy. In the adiab<strong>at</strong>ic steps two and four of the Carnot cycle changes in the<br />

volume are rel<strong>at</strong>ed to changes in the temper<strong>at</strong>ure according to<br />

( id<br />

V3 T1 =<br />

V2 T id<br />

)<br />

2<br />

3<br />

2<br />

= V4<br />

V1<br />

which shows th<strong>at</strong> the efficiency is given by<br />

η = 1 −<br />

T id<br />

2<br />

T id<br />

1<br />

The Carnot temper<strong>at</strong>ures were defined using this efficiency. We have<br />

(1.59)<br />

(1.60)


1.7. IDEAL GAS AND TEMPERATURE. 35<br />

T2<br />

T1<br />

= 1 − η =<br />

T id<br />

2<br />

T id<br />

1<br />

(1.61)<br />

and hence the two temper<strong>at</strong>ure scales are proportional to each other. This<br />

means th<strong>at</strong> they are really the same, since any temper<strong>at</strong>ure scale can always<br />

be modified by a constant scale factor. A similar analysis can be applied to all<br />

different kinds of thermometers. Standard thermometers are equivalent to the<br />

Carnot scale.<br />

Gibbs paradox.<br />

Suppose we have a system with N moles of m<strong>at</strong>erial in a volume V <strong>at</strong> temper<strong>at</strong>ure<br />

T and we add a similar system. If we use 1.58 the total entropy of the<br />

combined system is:<br />

which has to be equal to<br />

and hence we need<br />

which leads to<br />

Stotal(T id , 2V, 2N) = 2NR ln((T id ) 3<br />

2 V ) + 2S0(N) (1.62)<br />

S(T id , 2V, 2N) = 2NR ln((T id ) 3<br />

2 2V ) + S0(2N) (1.63)<br />

2S0(N) = S0(2N) + 2NR ln(2) (1.64)<br />

S0(N) = −NR ln(N) + NRc (1.65)<br />

where c is some constant. Therefore, the entropy of an ideal gas is<br />

S(T id (<br />

, V, N) = NR ln((T id ) 3<br />

)<br />

V<br />

2 ) + c<br />

N<br />

(1.66)<br />

Hence the dependence of the entropy on N is completely fixed by requiring<br />

th<strong>at</strong> the entropy is an extensive quantity. The constant c can be found from<br />

st<strong>at</strong>istical mechanical calcul<strong>at</strong>ions, and the expression for c contains the constant<br />

¯h. This is quite remarkable, since this is a direct manifest<strong>at</strong>ion of a quantum<br />

mechanical quantity in a macroscopic equ<strong>at</strong>ion!<br />

Without the constant of integr<strong>at</strong>ion S0(N) the entropy is not extensive, and<br />

using the first part of 1.58 only gives many problems. For example, consider a<br />

container of volume V with an amount NA of m<strong>at</strong>erial A and a similar container<br />

of volume V with an amount NB of m<strong>at</strong>erial B, all <strong>at</strong> temper<strong>at</strong>ure T. If we add<br />

these containers together, the new entropy is<br />

Stotal = S(T id , 2V, NA) + S(T id , 2V, NB) (1.67)<br />

since each gas now occupies twice the volume. The change in entropy is


36 CHAPTER 1. BASIC THERMODYNAMICS.<br />

∆S = Stotal − S(T id , V, NA) − S(T id , V, NB) =<br />

S(T id , 2V, NA) − S(T id , V, NA) + S(T id , 2V, NB) − S(T id , V, NB) (1.68)<br />

which gives<br />

∆S = (NA + NB)Rln(2) (1.69)<br />

which is called the entropy of mixing. Therefore, if we mix two different gases<br />

the entropy will increase. If we ignore the constant of integr<strong>at</strong>ion in equ<strong>at</strong>ion<br />

1.58 we get exactly the same answer if we make the m<strong>at</strong>erials A and B the same<br />

(Gibbs paradox), which cannot be true, since in this case there should not be a<br />

change in entropy. Including the constant of integr<strong>at</strong>ion resolves this paradox,<br />

as we demonstr<strong>at</strong>ed when we constructed the analytical form of this constant.<br />

The first steam engine<br />

One can argue about the exact definition of a steam engine, but it is fair to<br />

say th<strong>at</strong> Hero already constructed a steam engine more than 2000 years ago.<br />

The Romans did not have the analytical methods (m<strong>at</strong>hem<strong>at</strong>ics) to analyze<br />

this new toy, and never saw any use for it. Can you imagine wh<strong>at</strong> would have<br />

happened if the Romans, who were masters of engineering, would have had good<br />

smooth roads and steam driven cars?<br />

1.8 Extra questions.<br />

1. How do we measure temper<strong>at</strong>ure?<br />

2. How large does a thermometer need to be?<br />

3. How do we measure pressure?<br />

4. Wh<strong>at</strong> is the difference between direct and indirect measurements?<br />

5. Why is characteriz<strong>at</strong>ion in terms of st<strong>at</strong>e variables important?<br />

6. How realistic are reversible processes?<br />

7. Wh<strong>at</strong> happens when a process is not reversible?<br />

8. Comment on time-scales and equilibrium.<br />

9. Wh<strong>at</strong> is the law of large numbers, how does it play a role?<br />

10. Would pV = (NRT ) 2 be a possible equ<strong>at</strong>ion of st<strong>at</strong>e?<br />

11. Is pdV an exact differential?<br />

12. How is the internal energy defined in thermodynamics?


1.8. EXTRA QUESTIONS. 37<br />

13. Why is the internal energy a st<strong>at</strong>e function?<br />

14. Why would it be bad is Q and W were st<strong>at</strong>e functions?<br />

15. How many work terms can there be?<br />

16. Why would one tre<strong>at</strong> the µdN term differently?<br />

17. Why worry about measuring he<strong>at</strong> via Carnot engines?<br />

18. Why can the efficiency not be gre<strong>at</strong>er than one?<br />

19. Why is the maximum entropy principle important?<br />

20. Wh<strong>at</strong> is the thermodynamical basis for the maximum entropy principle?<br />

21. Wh<strong>at</strong> is the microscopic basis for the maximum entropy principle?<br />

22. Give a possible diagram for an irreversible engine.<br />

23. How do we measure entropy?<br />

24. How are the laws of thermodynamics obtained?<br />

25. Can we reach neg<strong>at</strong>ive temper<strong>at</strong>ures?<br />

26. Would a measure of temper<strong>at</strong>ure via T −1 be better?<br />

27. Give an example of a function where limits cannot be interchanged.<br />

28. Explain why we cannot reach T=0 (driving force to zero).<br />

29. Wh<strong>at</strong> is an ideal gas?<br />

30. Why is the calcul<strong>at</strong>ion in sec 1.7 important?<br />

31. Using dU = T dS − pdV find the equ<strong>at</strong>ions of st<strong>at</strong>e.<br />

32. How many of the variables T,S,p,V are independent?<br />

33. Suppose th<strong>at</strong> we have for an ideal gas pV = NRT and U = Nf(T ). Wh<strong>at</strong><br />

are the restrictions on f(T )?<br />

34. In a p-V-T-S system, suppose I know T=F(p,V), U=G(p,V). Can I find<br />

S? Do so.


38 CHAPTER 1. BASIC THERMODYNAMICS.<br />

1.9 Problems for chapter 1<br />

Problem 1.<br />

A system is characterized by the st<strong>at</strong>e variables x and y. The following<br />

functions are defined via differential changes in these st<strong>at</strong>e variables:<br />

(1) ¯ df = x 2 dx + y 2 dy<br />

(2) ¯ dg = y 2 dx + x 2 dy<br />

(3) ¯ dh = 3x 2 y 2 dx + (2x 3 y + y 4 )dy<br />

For each of these functions f,g, and h determine if they correspond to a st<strong>at</strong>e<br />

function. Express each st<strong>at</strong>e function in terms of x and y.<br />

Problem 2.<br />

The quantity q rel<strong>at</strong>ed to the st<strong>at</strong>e variables x and y is defined by ¯ dq =<br />

xydx + 1<br />

xdy. Show th<strong>at</strong> q is not a st<strong>at</strong>e function. The st<strong>at</strong>e function s is defined<br />

by ds = f(x) ¯ dq where f is a function of x only. Find f(x)<br />

Problem 3.<br />

The st<strong>at</strong>e of a system is determined by the st<strong>at</strong>e variables x and y. The<br />

actual st<strong>at</strong>e of the system can be changed experimentally by changing x and<br />

y. The energy U of the system is a function of x and y defined via dU =<br />

sin(x + y)dx + sin(x + y)dy. The work done on the outside world by the system<br />

when x and y are changed is given by ¯ dW = sin(x) cos(y)dx.<br />

(a) Show th<strong>at</strong> U is a st<strong>at</strong>e function and find U(x,y).<br />

(b) Show th<strong>at</strong> W is not a st<strong>at</strong>e function.<br />

(c) Give an expression for the differential change in he<strong>at</strong> ¯ dQ for this system.<br />

The system is cycled through st<strong>at</strong>es (x, y) in a reversible process. The cycle is<br />

given by (0, 0) → (απ, 0) → (απ, απ) → (0, απ) → (0, 0). α is a parameter with<br />

values restricted to [0, 2].<br />

(d) Calcul<strong>at</strong>e the work done on the environment in this cycle and the amount<br />

of he<strong>at</strong> taken up by the system as a function of α.<br />

(e) For which values of α is this process a he<strong>at</strong> engine?


1.9. PROBLEMS FOR CHAPTER 1 39<br />

Problem 4.<br />

An ideal gas is in equilibrium and has volume V0, temper<strong>at</strong>ure T0, and<br />

pressure p0. These are the only three st<strong>at</strong>e variables we can vary. The amount<br />

of m<strong>at</strong>erial is kept constant <strong>at</strong> N moles. This gas is cycled through the following<br />

process. First it expands <strong>at</strong> constant pressure to twice the volume. Next, it is<br />

cooled <strong>at</strong> constant volume to half the temper<strong>at</strong>ure it had <strong>at</strong> the beginning of<br />

this step. Then its volume is reduced to the original value <strong>at</strong> constant pressure.<br />

Finally, it is he<strong>at</strong>ed <strong>at</strong> constant volume to the original st<strong>at</strong>e. During this process<br />

an amount of he<strong>at</strong> Q1 > 0 flows into the system <strong>at</strong> steps one and four combined,<br />

and an amount of he<strong>at</strong> Q2 > 0 flows out of the system during steps two and<br />

three combined. The total amount of work performed on the environment is<br />

W. Calcul<strong>at</strong>e the efficiency for this process, defined by the r<strong>at</strong>io of the work W<br />

done on the environment and the amount of he<strong>at</strong> Q1 flowing into the system<br />

during of the process. Compare this value with the efficiency of a Carnot engine<br />

oper<strong>at</strong>ing between the highest and lowest temper<strong>at</strong>ure th<strong>at</strong> occur in the process<br />

described above.<br />

Problem 5.<br />

The internal energy of a system as a function of the entropy S and the volume<br />

V is given by U(S, V ) = S2<br />

V . The only other st<strong>at</strong>e variables of importance in<br />

this problem are the temper<strong>at</strong>ure T and the pressure p.<br />

(a) Using dU = T dS − pdV , calcul<strong>at</strong>e p(S, V ) and T (S, V ) for equilibrium<br />

st<strong>at</strong>es of the system.<br />

(b) Calcul<strong>at</strong>e the efficiency of a Carnot cycle for this system as a function of<br />

the temper<strong>at</strong>ures of the two reservoirs between which the Carnot engine is<br />

oper<strong>at</strong>ing. Calcul<strong>at</strong>e all contributions to the exchange of he<strong>at</strong> and work,<br />

and show th<strong>at</strong> the temper<strong>at</strong>ure defined in this problem is indeed equal to<br />

the Carnot temper<strong>at</strong>ure.<br />

Problem 6.<br />

A magnetic substance has temper<strong>at</strong>ure T and magnetiz<strong>at</strong>ion M. The first<br />

law in this case is dU = T dS +HdM, where H is the applied magnetic field. The<br />

internal energy U is a function of T only and does not depend on M, U = Nf(T ).<br />

The equ<strong>at</strong>ion of st<strong>at</strong>e is M = CH<br />

T , where C is a constant proportional to the<br />

size of the system. Calcul<strong>at</strong>e the efficiency of a Carnot cycle for an engine<br />

using a magnetic substance which is doing work on the outside world when<br />

reducing the magnetiz<strong>at</strong>ion for this substance, by calcul<strong>at</strong>ing all contributions<br />

to the exchange of he<strong>at</strong> and work, and show th<strong>at</strong> the temper<strong>at</strong>ure defined in<br />

this problem is indeed equal to the Carnot temper<strong>at</strong>ure.


40 CHAPTER 1. BASIC THERMODYNAMICS.<br />

Problem 7.<br />

This problem is an example of a consequence of the third law: it is impossible<br />

to reach absolute zero temper<strong>at</strong>ure in a finite number of steps if a reversible<br />

process is used. The entropy of a system is given by S = αV T , where α is a<br />

positive constant and V is the volume of the system. The system is originally<br />

<strong>at</strong> temper<strong>at</strong>ure T0 and volume V0. A reversible two-step process is used to cool<br />

this system. In the first part of the process the system expands adiab<strong>at</strong>ically<br />

to a volume 2V0. In the second part of the process the system is compressed<br />

isothermally back to V0. Calcul<strong>at</strong>e the temper<strong>at</strong>ure Tn of the system after this<br />

two-step process has been applied n times.<br />

Problem 8.<br />

The internal energy of systems 1 and 2 is given by Ui = CiT , where the<br />

constants C1 and C2 are both positive. Originally system 1 is <strong>at</strong> temper<strong>at</strong>ure<br />

T1 and system 2 <strong>at</strong> temper<strong>at</strong>ure T2. These systems are brought into thermal<br />

contact. This establishes an infinitesimally slow flow of energy between the<br />

systems, but the two systems are completely isol<strong>at</strong>ed from the rest of the world<br />

and energy cannot flow out of the combined system. The individual systems are<br />

always in equilibrium. The total system is not in equilibrium just after contact<br />

is established, but finally will reach equilibrium <strong>at</strong> a common temper<strong>at</strong>ure Tf .<br />

(a) Calcul<strong>at</strong>e Tf .<br />

(b) Calcul<strong>at</strong>e the change in entropy during this spontaneous process.<br />

(c) Show th<strong>at</strong> the entropy did increase.<br />

Problem 9.<br />

A removable partition divides a container in two parts with volumes V1 and<br />

V2. The container is filled with an ideal gas. The temper<strong>at</strong>ure and the pressure<br />

in each part are equal to T and p, respectively.<br />

(a) Calcul<strong>at</strong>e the amount of m<strong>at</strong>erial N1 and N2 in each part of the system.<br />

(b) Calcul<strong>at</strong>e the entropy of the combined system by adding the entropies of<br />

the two parts.<br />

The partition is now removed. Since the pressure and temper<strong>at</strong>ure of both parts<br />

were already equal, the combined system is already in equilibrium.<br />

(c) Calcul<strong>at</strong>e the total entropy of the system directly as a function of N =<br />

N1 + N2 and V = V1 + V2.


1.9. PROBLEMS FOR CHAPTER 1 41<br />

(d) Show th<strong>at</strong> the results of b and c are different.<br />

Since the removal of the partition does not cause any spontaneous process, the<br />

entropy before and after should be the same. The problem is rel<strong>at</strong>ed to our<br />

deriv<strong>at</strong>ion of the entropy for an ideal gas. The constant S0 does not depend on<br />

T or V, but could still depend on N.<br />

(e) Assuming th<strong>at</strong> S0 is a function of N, derive its functional form from the<br />

requirement th<strong>at</strong> the entropy calcul<strong>at</strong>ed in b and c is the same.<br />

Problem 10.<br />

An engine oper<strong>at</strong>es between a high temper<strong>at</strong>ure reservoir and a low temper<strong>at</strong>ure<br />

reservoir. Everything in this engine is ideal, except th<strong>at</strong> from each amount<br />

of he<strong>at</strong> Q taken by the engine from the high temper<strong>at</strong>ure reservoir an amount<br />

αQ with 0 < α < 1 flows <strong>at</strong> the same time directly into the low temper<strong>at</strong>ure<br />

reservoir through a leak.<br />

(a) How much entropy is cre<strong>at</strong>ed in this leak?<br />

(b) Calcul<strong>at</strong>e the efficiency of this engine.<br />

Problem 11.<br />

A Stirling engine uses the following cycle (all steps are reversible):<br />

(1) Isothermal expansion, absorbing an amount of he<strong>at</strong> Qh <strong>at</strong> Th.<br />

(2) Isochoric (constant volume) cooling to Tl, giving up an amount of he<strong>at</strong><br />

Qc.<br />

(3) Isothermal contraction, giving off an amount of he<strong>at</strong> Ql <strong>at</strong> Tl.<br />

(4) Isochoric warming to Th, absorbing an amount of he<strong>at</strong> Qa.<br />

Using these non-standard definitions all values of Qi are positive. Assume th<strong>at</strong><br />

Qa = Qc and th<strong>at</strong> this energy is recycled, th<strong>at</strong> is the he<strong>at</strong> given up in step 2<br />

returns to the system in step 4. Show th<strong>at</strong> for an ideal gas the efficiency of the<br />

Stirling cycle is less than the Carnot efficiency.<br />

Problem 12.<br />

The temper<strong>at</strong>ure is an intensive st<strong>at</strong>e variable since the definition is independent<br />

of the size of the reservoir. Show th<strong>at</strong> the entropy is extensive, in two<br />

different ways, using:


42 CHAPTER 1. BASIC THERMODYNAMICS.<br />

(a) Carnot engines which are working in parallel between the same reservoirs.<br />

(b) the knowledge th<strong>at</strong> the internal energy U is extensive.<br />

Problem 13.<br />

Car<strong>at</strong>heodory’s formul<strong>at</strong>ion of the second law is: There exists arbitrarily<br />

close to any given st<strong>at</strong>e of a system other st<strong>at</strong>es which cannot be reached from it<br />

by adiab<strong>at</strong>ic processes alone. In order to show th<strong>at</strong> this formul<strong>at</strong>ion is equivalent<br />

to Kelvin’s, we use a simple system where only mechanical work can be done.<br />

Hence an adiab<strong>at</strong>ic process is described by dU = −pdV .<br />

(a) Show th<strong>at</strong> in the U-V plane all points th<strong>at</strong> are connected by adiab<strong>at</strong>ic<br />

processes alone define a set of non-intersecting curves.<br />

These curves are labelled by a parameter S, with S(U,V) continuous and differentiable.<br />

(b) Show th<strong>at</strong> Car<strong>at</strong>heodory’s formul<strong>at</strong>ion of the second law implies th<strong>at</strong><br />

this is possible<br />

Define T −1 = ( )<br />

∂S<br />

∂U V .<br />

(c) Use the first law to show th<strong>at</strong> identifying S with the entropy and T with<br />

temper<strong>at</strong>ure is consistent with the standard definitions.<br />

(d) Is the identific<strong>at</strong>ion in part c the only possible one?<br />

Problem 14.<br />

An ideal gas is defined by the equ<strong>at</strong>ions of st<strong>at</strong>e pV = NRT , U = 3<br />

2NRT ,<br />

and S = NR ln(T √ T V<br />

N ). It is used in an engine following the Otto cycle:<br />

(1) Adiab<strong>at</strong>ic compression from volume V1 to a smaller volume V2, with temper<strong>at</strong>ure<br />

changing from Tc to T ′ .<br />

(2) He<strong>at</strong>ing <strong>at</strong> constant volume to a temper<strong>at</strong>ure Th.<br />

(3) Adiab<strong>at</strong>ic expansion V2 → V1 (power stroke). The temper<strong>at</strong>ure changes<br />

to T ′′ .<br />

(4) Cooling <strong>at</strong> constant volume to Tc.<br />

The amount of he<strong>at</strong> absorbed by the gas in step two is Qin, the amount of<br />

he<strong>at</strong> absorbed by the gas in step four is Qout (which is neg<strong>at</strong>ive). Calcul<strong>at</strong>e<br />

the efficiency of the Otto cycle and show th<strong>at</strong> it is less than the efficiency of a<br />

Carnot cycle oper<strong>at</strong>ing between Th and Tc.


1.9. PROBLEMS FOR CHAPTER 1 43<br />

Problem 15.<br />

In this problem we use dS = 1 p<br />

T dU + T dV . The internal energy is given by<br />

U(T, V ) = cV T 4 , with c a constant. Since the entropy is a st<strong>at</strong>e function, dS is<br />

exact. Use this to show th<strong>at</strong> p(T, V ) = 1<br />

3cT 4 + f(V )T , where f(V ) is a function<br />

of volume only. Calcul<strong>at</strong>e the entropy S(T, V ) and show th<strong>at</strong> f(V ) ≡ 0.<br />

Problem 16.<br />

Consider the following differential:<br />

¯df = g(x, y, z)dx + h(x, y, z)dy + k(x, y, z)dz<br />

Wh<strong>at</strong> are the necessary and sufficient conditions for ¯ df to be exact?<br />

Problem 17.<br />

Consider the following examples:<br />

¯df = 3x 2 y 2 zdx + 2x 3 yzdy + 6x 3 y 2 dz<br />

¯dg = 2xy 2 z 2 dx + 2x 2 yz 2 dy + 2x 2 y 2 zdz<br />

Are these differentials exact? If so, find the corresponding st<strong>at</strong>e function.<br />

Problem 18.<br />

A system is completely described by the four st<strong>at</strong>e variables T,S,p, and V.<br />

1. Show th<strong>at</strong> the general form of the energy equ<strong>at</strong>ion is U = V f( S<br />

V ).<br />

2. If U = S α V 1−α find the equ<strong>at</strong>ions of st<strong>at</strong>e p(S,V) and T (S,V).<br />

3. The entropy S is not always easy to deal with, and we often elimin<strong>at</strong>e this<br />

variable. In the previous case, solve for p(T,V). Wh<strong>at</strong> is special in this<br />

equ<strong>at</strong>ion, and why is th<strong>at</strong> the case?<br />

4. Calcul<strong>at</strong>e the efficiency of a Carnot engine using a m<strong>at</strong>erial with the energy<br />

equ<strong>at</strong>ion U = S α V 1−α .<br />

Problem 19.<br />

Consider the following cyclic process (Brayton), with an ideal mono-<strong>at</strong>omic<br />

gas as the working m<strong>at</strong>erial :


44 CHAPTER 1. BASIC THERMODYNAMICS.<br />

1. Adiab<strong>at</strong>ic expansion from pA, V1, T1 to pB, V2, T2.<br />

2. Isobaric expansion from pB, V2, T2 to pB, V3, T3.<br />

3. Adiab<strong>at</strong>ic contraction from pB, V3, T3 to pA, V4, T4.<br />

4. Isobaric contraction from pA, V4, T4 to pA, V1, T1.<br />

Calcul<strong>at</strong>e the efficiency in terms of pA and pB. Compare this with the<br />

Carnot efficiency (indic<strong>at</strong>e the temper<strong>at</strong>ures used, and why you choose these<br />

references).<br />

Problem 20.<br />

In an Ericsson cycle isothermal expansion and contraction is altern<strong>at</strong>ed with<br />

expansion and contraction <strong>at</strong> constant pressure. Use an ideal gas, pV = NRT<br />

and U = 3<br />

2pV , to show th<strong>at</strong> the r<strong>at</strong>io of the total work done to the he<strong>at</strong> in <strong>at</strong><br />

the isothermal expansion <strong>at</strong> high temper<strong>at</strong>ure is the same as th<strong>at</strong> for a Carnot<br />

engine oper<strong>at</strong>ing between the same two reservoirs.<br />

Problem 20.


1.9. PROBLEMS FOR CHAPTER 1 45<br />

The figure above shows the experimental results th<strong>at</strong> led to the most recent<br />

Nobel prize. Left are the d<strong>at</strong>a from Fert on multilayers, right the d<strong>at</strong>a from<br />

Grünberg for a double layer. At zero field the direction of the magnetic moments<br />

of the iron layers altern<strong>at</strong>es, <strong>at</strong> high field they all point in the same direction.<br />

They measured the resistance to currents parallel to the layers, and found th<strong>at</strong> it<br />

changed as a function of field. Based on the reading in the notes, sections 1.1-3,<br />

wh<strong>at</strong> can you say about the difference and similarities between these figures?<br />

Problem 21.<br />

Consider a system where we have<br />

and we are given th<strong>at</strong><br />

dU = pdV − HdM<br />

2 M<br />

p(V, M) = π<br />

V 2<br />

Find H(V, M), if we know th<strong>at</strong> H = 0 ⇒ M = 0. Find M(V, H).<br />

Problem 22.<br />

Derive an expression for ( )<br />

∂U<br />

∂V T,N th<strong>at</strong> contains ( )<br />

∂S<br />

∂V .<br />

T,N<br />

Calcul<strong>at</strong>e the<br />

l<strong>at</strong>ter deriv<strong>at</strong>ive in the case th<strong>at</strong> p = RT g( N<br />

V ), with an unknown function g(x).<br />

Use these results to show th<strong>at</strong> the internal energy has the form U = Nh(T ).<br />

Problem 23.<br />

You have measured a m<strong>at</strong>erials property, call it incognito, measured in<br />

ploughs, as a function of temper<strong>at</strong>ure, in Kelvin, and found the following d<strong>at</strong>a<br />

Temper<strong>at</strong>ure (Kelvin) Incognito (ploughs)<br />

0 0.000<br />

100 0.048<br />

200 0.091<br />

300 0.130<br />

400 0.167<br />

500 0.200<br />

600 0.231<br />

700 0.259<br />

800 0.286<br />

900 0.310<br />

1000 0.333<br />

Find a rel<strong>at</strong>ion th<strong>at</strong> gives incognito as a function of temper<strong>at</strong>ure reasonably<br />

well for this measurement.<br />

Next, you measure incognito for another m<strong>at</strong>erial, and you find


46 CHAPTER 1. BASIC THERMODYNAMICS.<br />

Temper<strong>at</strong>ure (Kelvin) Incognito (ploughs)<br />

0 0.000<br />

100 0.833<br />

200 0.909<br />

300 0.938<br />

400 0.952<br />

500 0.962<br />

600 0.968<br />

700 0.972<br />

800 0.976<br />

900 0.978<br />

1000 0.980<br />

Find a rel<strong>at</strong>ion th<strong>at</strong> gives incognito as a function of temper<strong>at</strong>ure reasonably<br />

well for this measurement.<br />

But now you realize th<strong>at</strong> these two m<strong>at</strong>erials are really similar in the kind<br />

of physics they display. They are different only because the temper<strong>at</strong>ure has<br />

different effects. Assume th<strong>at</strong> there is a typical temper<strong>at</strong>ure T0 associ<strong>at</strong>ed with<br />

each m<strong>at</strong>erial, a m<strong>at</strong>erials parameter. Find a rel<strong>at</strong>ion as a function of temper<strong>at</strong>ure<br />

which includes this typical temper<strong>at</strong>ure T0 as the only unknown parameter,<br />

which can explain both d<strong>at</strong>a sets. Find the value of T0 for each m<strong>at</strong>erial.<br />

Problem 24.<br />

A gas is contained in a cylinder with a piston th<strong>at</strong> can move. We can measure<br />

the pressure on the piston, the volume of the gas, and its temper<strong>at</strong>ure. We find<br />

the following set of d<strong>at</strong>a:<br />

Day Time p (<strong>at</strong>m) T (Cent) V (liters)<br />

Monday, March 16 11am 1.0 22.2 3.1<br />

Tuesday, March 17 11am 1.0 22.2 3.1<br />

Wednesday, March 18 11am 1.0 22.2 3.1<br />

Thursday, March 19 3pm 1.0 22.2 2.7<br />

Friday, March 20 11am 1.0 22.2 2.7<br />

S<strong>at</strong>urday, March 21 11am 1.0 22.2 2.7<br />

Sunday, March 22 11am 1.0 22.2 2.7<br />

Give an explan<strong>at</strong>ion for these observ<strong>at</strong>ions, which is consistent with the<br />

physics we have discussed, and does not assume irr<strong>at</strong>ional behavior by the observer.<br />

Problem 25.<br />

Consider a system which is electrically polarizable where we have<br />

dU = pdV − EdP


1.9. PROBLEMS FOR CHAPTER 1 47<br />

and we are given th<strong>at</strong><br />

p(V, P ) = ϵ0 P<br />

2<br />

2<br />

V 2<br />

Find U(V, P ), if we know th<strong>at</strong> E = 0 ⇒ P = 0.<br />

Problem 26.<br />

We all know th<strong>at</strong> it is more expensive to oper<strong>at</strong>e a refriger<strong>at</strong>or in summer<br />

than in winter. Give three reasons why this might be the case, and rank them<br />

according to how important they are. Give a clear justific<strong>at</strong>ion for your answer.


48 CHAPTER 1. BASIC THERMODYNAMICS.


Chapter 2<br />

Thermodynamic potentials<br />

and response functions.<br />

In the previous chapter we had a lot of discussion about the meaning of various<br />

quantities. Th<strong>at</strong> is important, because it does give us some idea wh<strong>at</strong> th<strong>at</strong><br />

mysterious quantity entropy is. The m<strong>at</strong>hem<strong>at</strong>ical deriv<strong>at</strong>ions were minimal,<br />

which is out of necessity. We first need to have definitions.<br />

One can also take a very practical approach, however, and simply st<strong>at</strong>e th<strong>at</strong><br />

the first law of thermodynamics is given by dU = T dS − pdV + µdN for the<br />

simple systems we considered. It is understood th<strong>at</strong> additional work terms can<br />

be added when needed. The second law transl<strong>at</strong>es into the st<strong>at</strong>ement th<strong>at</strong> for an<br />

isol<strong>at</strong>ed system the equilibrium st<strong>at</strong>e is th<strong>at</strong> st<strong>at</strong>e th<strong>at</strong> maximizes the entropy as<br />

a function of internal variables. These two st<strong>at</strong>ements are sufficient for starting<br />

the current chapter. Of course, there is some additional semantics th<strong>at</strong> needs<br />

to be referred to, and it will be important to go back to the first chapter when<br />

it is not quite clear wh<strong>at</strong> certain wording means.<br />

One other idea from the first chapter should not get lost, however. The<br />

fundamental definition of thermodynamical processes requires th<strong>at</strong> the system<br />

we consider is very large on some scale, but th<strong>at</strong> there always has to be an<br />

even larger outside world th<strong>at</strong> allows us to make contact with the system and<br />

measure properties of the system. <strong>Thermodynamics</strong> applies to systems looked<br />

<strong>at</strong> from the outside. The n<strong>at</strong>ure of the first law demands this. If we try to relax<br />

this idea, and use one part of the system to measure the other and vice-versa,<br />

we will run into some real problems. In such a case we have lost our common<br />

reference. So we are free to add the entropy of two blocks of iron, but we are<br />

not free to add the entropy of all separ<strong>at</strong>e blocks of m<strong>at</strong>erial in the universe<br />

to get the entropy of the universe. The st<strong>at</strong>ement th<strong>at</strong> based on the second<br />

law of thermodynamics the universe will keep getting warmer and warmer is<br />

an incorrect applic<strong>at</strong>ion of thermodynamics. We need to apply fundamental<br />

theories of gravity to answer the question if the universe is warming up or not.<br />

We start this chapter with a discussion of energy. At first, th<strong>at</strong> may seem<br />

49


50 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

redundant, because introductory physics has us well prepared to think about<br />

energy. The problem is th<strong>at</strong> in thermodynamical processes we have many different<br />

variables th<strong>at</strong> we can change of keep constant. So work done depends<br />

on the type of process. The mechanical work done on the outside world in a<br />

process with constant amounts of m<strong>at</strong>erial is not the same as the work done in<br />

a process where the chemical potential is kept constant. Since work is rel<strong>at</strong>ed<br />

to a change in available energy th<strong>at</strong> means th<strong>at</strong> in different processes different<br />

amounts of energy are available, or free, to do work.<br />

The introduction of different free energies allows us to change the maximum<br />

entropy st<strong>at</strong>ement in more useful terms. Maximum entropy turns out to be<br />

equivalent to minimum energy. Since we have several free energies we can now<br />

make st<strong>at</strong>ements about equilibrium for systems th<strong>at</strong> are not isol<strong>at</strong>ed.<br />

In the sections th<strong>at</strong> follow after th<strong>at</strong> we will explore the m<strong>at</strong>hem<strong>at</strong>ical consequences<br />

of the first law as st<strong>at</strong>ed above. The theory of multi-variable calculus<br />

leads to some important restrictions on partial deriv<strong>at</strong>ives, and the distinction<br />

between intensive and extensive variables gives us an Euler rel<strong>at</strong>ion. First order<br />

partial deriv<strong>at</strong>ives are fundamental connection between theory and experiment<br />

in thermodynamics, because they represent the linear response of systems. Experiment<br />

gives us values of partial deriv<strong>at</strong>ives, and theory tells us th<strong>at</strong> they are<br />

somehow rel<strong>at</strong>ed, and hence gives important checks on experimental d<strong>at</strong>a.<br />

The minimum energy principle gives us inform<strong>at</strong>ion about the signs of second<br />

order deriv<strong>at</strong>ives of the free energy. Those represent stability criteria, since a<br />

system th<strong>at</strong> is not in equilibrium will want to go back to equilibrium. As a<br />

consequence, we will find for example th<strong>at</strong> he<strong>at</strong> capacities have to be positive.<br />

It is truly remarkable th<strong>at</strong> the basic formul<strong>at</strong>ion of the first and second law,<br />

combined with pure m<strong>at</strong>hem<strong>at</strong>ics, gives us so many conclusions we can draw<br />

about thermodynamical processes. This chapter is general, and the conclusions<br />

hold for any system. We do not need to know the details of the system. The<br />

conclusions are independent of the equ<strong>at</strong>ions of st<strong>at</strong>e. In fact, they give us<br />

restrictions on wh<strong>at</strong> we can use for equ<strong>at</strong>ions of st<strong>at</strong>e! To me, this is the beauty<br />

of thermodynamics.<br />

In the last two sections we take two small excursions to show the applicability<br />

of wh<strong>at</strong> we found. In magnetic systems response functions th<strong>at</strong> should be<br />

positive can be neg<strong>at</strong>ive, which seems a contradiction. Th<strong>at</strong> contradiction is easily<br />

resolved, however, by correctly including the energy stored in the magnetic<br />

fields. Similar observ<strong>at</strong>ions hold for electric fields. Finally, we analyze in more<br />

detail the n<strong>at</strong>ure of fluctu<strong>at</strong>ions. Here we do introduce an additional element,<br />

based on ideas from st<strong>at</strong>istical mechanics. We propose th<strong>at</strong> the probability of a<br />

given fluctu<strong>at</strong>ion is rel<strong>at</strong>ed to the change in entropy for th<strong>at</strong> fluctu<strong>at</strong>ion, with<br />

the system tre<strong>at</strong>ed as being internally in equilibrium, only not with the outside<br />

world. We will come back to fluctu<strong>at</strong>ions when we discuss transport.


2.1. INTERNAL ENERGY. 51<br />

2.1 Internal energy.<br />

Potential energy.<br />

When a mass is dropped from a certain height, it will gain kinetic energy.<br />

If it hits the floor, this kinetic energy will be transformed into he<strong>at</strong> and/or do<br />

work. The amount of work this mass is able to do is completely determined<br />

by the difference in potential energy between the initial position and the floor.<br />

Hence the gravit<strong>at</strong>ional potential energy is a measure of the amount of energy<br />

stored in this mass initially, and this energy can be released to do work.<br />

Energy stored in a gas.<br />

Energy is also stored in a gas. If we allow a gas to expand, it will do work<br />

on the outside world. The amount of energy which can be released, however,<br />

depends on the n<strong>at</strong>ure of the process by which the gas is doing this work. In<br />

most cases some of the thermodynamic variables are kept constant. A standard<br />

example is the expansion of a gas <strong>at</strong> constant temper<strong>at</strong>ure or pressure. The<br />

energy which can be released depends on the values of these constant variables.<br />

For each situ<strong>at</strong>ion one defines an appropri<strong>at</strong>e thermodynamic potential or free<br />

energy. This makes thermodynamics slightly more complic<strong>at</strong>ed than ordinary<br />

mechanics. One has to remember which thermodynamic potential applies in<br />

which case.<br />

Why does the energy given up by a gas depend on the process?<br />

Internal energy is a free energy.<br />

The easiest thermodynamic potential pertains to a process which is adiab<strong>at</strong>ic.<br />

In an adiab<strong>at</strong>ic process there is no he<strong>at</strong> exchange and the entropy is<br />

constant. The first law is in this case ∆U = −∆W . In other words:<br />

The internal energy measures the amount of work a system is able<br />

to do in an reversible adiab<strong>at</strong>ic process.<br />

This tells us th<strong>at</strong> the internal energy is the appropri<strong>at</strong>e thermodynamic potential<br />

or free energy for an adiab<strong>at</strong>ic process. As usual, in reality the amount of work<br />

done is always less then the change in internal energy due to inefficiencies in the<br />

process. But those details are of a technical n<strong>at</strong>ure, and can be minimized by<br />

improving the design.


52 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

Internal energy and extensive st<strong>at</strong>e variables.<br />

Consider a completely isol<strong>at</strong>ed or closed system. All the extensive variables<br />

of the system have either a fixed value or interact with a generalized external<br />

force which is zero. In a closed system the amount of m<strong>at</strong>erial N is constant<br />

and typically the volume V is constant. We might, however, not be able to<br />

set all extensive st<strong>at</strong>e variables. For example, the magnetic moment ⃗ M might<br />

freely find its optimal value. This is only true if the external magnetic fields are<br />

small enough. If th<strong>at</strong> is the case, the magnetic moment is an unimportant st<strong>at</strong>e<br />

variable and we ignore it in the description of the system. One only includes<br />

st<strong>at</strong>e variables which can be measured, and if there are no external magnetic<br />

fields we cannot measure the magnetic moment. One has to include the entropy<br />

though. Since the system cannot exchange he<strong>at</strong> with the outside world, the<br />

entropy is constant, but its actual value will determine the value of the internal<br />

energy. In our closed system <strong>at</strong> constant volume and number of particles the<br />

internal energy is a function of the entropy, the volume, and the amount of<br />

m<strong>at</strong>erial only, when the system is in equilibrium. Hence we have U(S,V,N).<br />

Also, the temper<strong>at</strong>ure T and the pressure p, and the chemical potential µ are<br />

completely determined by the values of S, V, and N. The n<strong>at</strong>ural set of st<strong>at</strong>e<br />

variables to describe a completely isol<strong>at</strong>ed system are all the extensive st<strong>at</strong>e<br />

variables which cannot be changed without doing work on the outside world.<br />

The free energy for such a system is the internal energy.<br />

Minimum energy principle.<br />

The number of parameters to describe a macroscopic system is infinitely<br />

large. We could decide to measure some of those by using an appropri<strong>at</strong>e external<br />

probe. For example, we could divide the total system in two parts, separ<strong>at</strong>ed<br />

by a movable wall (no friction, but the wall has mass), which is an ideal thermal<br />

conductor.<br />

The pressure in the left part is pl and in the right part pr. Assume th<strong>at</strong><br />

the left and right part of the system are both in equilibrium, but not necessarily<br />

with each other. The parameter pl is an additional parameter needed to<br />

describe this system. This is a macroscopic parameter, since it pertains to a<br />

macroscopic part of the system. Once we know pl, we know Vl (equilibrium)<br />

and hence Vr and pr. In equilibrium we obviously need pl = pr (no net force<br />

on the partition), otherwise the wall would move. Suppose the system is not in<br />

equilibrium, however, and th<strong>at</strong> pl < pr. The opposite case can be tre<strong>at</strong>ed in a<br />

similar way.<br />

No friction.


2.1. INTERNAL ENERGY. 53<br />

ÔÐ ÔÖ<br />

Figure 2.1: Container with piston as internal divider.<br />

First assume th<strong>at</strong> this is the complete system, and th<strong>at</strong> there is no friction.<br />

It is easy to see wh<strong>at</strong> will happen. The wall will start to move to the left, and<br />

gain kinetic energy. At the point where pl = pr the wall is moving, and will keep<br />

moving to the left until the pressure on the left is wh<strong>at</strong> the original pressure was<br />

on the right. Then the process will continue in reverse. The wall will oscill<strong>at</strong>e!<br />

Since there is no exchange with the outside world and since there is no friction,<br />

for the complete system we have ∆V = ∆N = ∆S = 0 and indeed the internal<br />

energy does not change. The internal energy has the form:<br />

U = Ul + Ur + 1<br />

2 Mv2<br />

(2.1)<br />

where the first two terms represent the internal energy of the left and right side,<br />

and the last term the kinetic energy of the wall. When the pressure left and<br />

right are the same, the wall has the largest kinetic energy, and in th<strong>at</strong> case the<br />

sum of internal energies is smallest.<br />

The motion of the wall is an organized motion, and this kinetic energy could<br />

be extracted.<br />

With friction.<br />

Can you design an experiment th<strong>at</strong> will do this?<br />

In the second case suppose there is friction and the wall loses energy to the<br />

gas when it is moving. The wall will again oscill<strong>at</strong>e, but now this oscill<strong>at</strong>ion<br />

is damped, and the final st<strong>at</strong>e has the wall <strong>at</strong> a position with pl = pr. The<br />

mechanical energy of the partition has been transformed into thermal energy of


54 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

<br />

Å<br />

ÔÐ ÔÖ<br />

Figure 2.2: Container where the internal degree of freedom becomes external<br />

and hence can do work.<br />

the gases! Since the whole system is isol<strong>at</strong>ed we have ∆Q = 0 and hence for<br />

this process we find from 1.41 ∆S ≥ 0. This agrees with the maximum entropy<br />

principle, the equilibrium st<strong>at</strong>e corresponds to maximum entropy as far as the<br />

internal coordin<strong>at</strong>es are concerned. The maximum entropy principle tells us<br />

th<strong>at</strong> the organized energy of the wall has gone into the internal energy of the<br />

two parts of the system, and th<strong>at</strong> increase in energy is directly rel<strong>at</strong>ed to the<br />

increase in entropy.<br />

Wall corresponds to st<strong>at</strong>e variable.<br />

In the third case we connect the movable wall with a massless cord to some<br />

object with mass M (see figure). The value of this mass M is adjusted in such<br />

a way th<strong>at</strong> Mg = (pr − pl)A(1 − ϵ) is always true, where A is the area of the<br />

movable wall. We also take the limit ϵ → 0, which allows the wall to move in a<br />

quasi-st<strong>at</strong>ic manner. Therefore, in this system there is no net force acting on the<br />

wall, or the generalized force connected with the st<strong>at</strong>e variable corresponding<br />

to the wall position is always zero. We also assume th<strong>at</strong> there is no friction.<br />

At this point the total system is in equilibrium. Now we make the mass M<br />

smaller by very slowly taking very small pieces of th<strong>at</strong> mass away. At every point<br />

during this process the total system is in equilibrium, and the whole process is<br />

reversible. The wall will move slowly to the left, and the mass M will move<br />

upwards. Hence the system is doing work on this mass. At the end, the mass is<br />

zero and the wall does not interact with the outside any more. The total system<br />

is in equilibrium, as is the closed system itself.<br />

In this process there is always equilibrium and because the wall is a thermal<br />

conductor we always have Tl = Tr. Since there is no he<strong>at</strong> exchange with the<br />

outside world we have Tl∆Sl = ∆Ql = −∆Qr = −Tr∆Sr, and hence ∆Sl =<br />

−∆Sr. For the total system we have therefore ∆S = 0, as it should be for a


2.1. INTERNAL ENERGY. 55<br />

reversible process.<br />

Originally, the closed system was not in equilibrium. The internal energy of<br />

the original system was higher than in the final equilibrium situ<strong>at</strong>ion, since during<br />

the process of reaching equilibrium mass was lifted to a higher gravit<strong>at</strong>ional<br />

potential energy. The internal energy of the system is now lower. We arrive <strong>at</strong><br />

the important conclusion:<br />

The internal energy of a closed system is a minimum as a function<br />

of all parameters which are not fixed by external constraints.<br />

Explain carefully the difference between these three scenarios.<br />

Formal deriv<strong>at</strong>ion of minimum energy principle.<br />

This minimum energy st<strong>at</strong>ement also follows from the rel<strong>at</strong>ion between energy,<br />

entropy, and work, T ∆S ≥ ∆U + ∆W . In our example of case three<br />

we have a closed system, and the parameters S, V, and N are fixed, but the<br />

internal coordin<strong>at</strong>e pl is allowed to vary. No work is done on the outside world<br />

by changing the extensive parameters V and N, and neither is he<strong>at</strong> exchanged,<br />

which keeps S constant. Hence for a spontaneous process bringing us from<br />

the initial non-equilibrium situ<strong>at</strong>ion to the final equilibrium situ<strong>at</strong>ion, we have<br />

0 ≥ ∆U + 0. The internal energy decreases and is minimal <strong>at</strong> equilibrium. Note<br />

th<strong>at</strong> this minimum principle does not hold for those variables which are fixed<br />

by non-zero external constraints, since a change in those variables involves work<br />

done on the outside world and the n<strong>at</strong>ure of the outside world has to be taken<br />

into account.<br />

If a system is kept <strong>at</strong> constant entropy, volume, etc, it will still be able<br />

to exchange energy with the outside world. In the earlier thought experiment<br />

we had one of the internal degrees of freedom do work. In general, th<strong>at</strong> is<br />

not possible. But a system <strong>at</strong> constant entropy is still able to exchange he<strong>at</strong><br />

with the outside world! The loss in internal energy is due to he<strong>at</strong> dumped<br />

to the outside world. Remember, th<strong>at</strong> only for reversible processes we have<br />

T ∆S = ∆Q. If a system spontaneously goes to equilibrium, the process is by<br />

definition irreversible.<br />

Difference between minimum energy and maximum entropy.<br />

In the previous chapter we considered a system <strong>at</strong> constant energy, volume,<br />

amount of m<strong>at</strong>erial, etc. All the extensive variables th<strong>at</strong> would be able to do<br />

work are kept constant, and hence no work is done. The internal energy is also<br />

constant, and hence no he<strong>at</strong> is transferred either. We want to find the entropy.


56 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

Suppose we have a model th<strong>at</strong> allows us to calcul<strong>at</strong>e S(U, V, N, Xint) where Xint<br />

represents the internal degrees of freedom. The maximum entropy principle tells<br />

us th<strong>at</strong> we can find the values of these internal coordin<strong>at</strong>es (coordin<strong>at</strong>es th<strong>at</strong><br />

are not connected to terms in the work differential) by maximizing the function<br />

S. In the current chapter we consider a system <strong>at</strong> constant entropy, volume,<br />

amount of m<strong>at</strong>erial, etc. In this case there is again no work done, but he<strong>at</strong><br />

can be exchanged. Suppose we now have a model th<strong>at</strong> allows us to calcul<strong>at</strong>e<br />

U(S, V, N, Xint). The minimum energy principle tells us th<strong>at</strong> we can find the<br />

values of these internal coordin<strong>at</strong>es.<br />

Example.<br />

A typical applic<strong>at</strong>ion of this minimum principle would be the determin<strong>at</strong>ion<br />

of the magnetic moment of a gas. Consider the variable ⃗ M as an internal<br />

parameter. A typical form of the internal energy of a gas with magnetic moment<br />

⃗M would be U(S, V, N; ⃗ M) = U(S, V, N;⃗0) + 1<br />

2V M 2 . The last term represents<br />

the self-energy needed to cre<strong>at</strong>e a magnetic moment. The minimum principle<br />

tells us th<strong>at</strong> the magnetic moment has to be zero in equilibrium in this case.<br />

Additional interactions are needed to cre<strong>at</strong>e a magnetic moment, like applying<br />

a non-zero magnetic field.<br />

Internal energy and adiab<strong>at</strong>ic processes.<br />

The internal energy measures the amount of work a system can do in an<br />

adiab<strong>at</strong>ic process. Consider a thermally isol<strong>at</strong>ed system which can change its<br />

volume by expanding. An example is a gas in an completely insul<strong>at</strong>ed cylinder<br />

with a piston. The system starts in equilibrium with the piston fixed by an<br />

external force. In the next step, the piston is allowed to move and work is<br />

performed on the outside world. The piston stops again, and the volume of the<br />

system is increased. Is the amount of work done equal to the change in internal<br />

energy? The answer is no, since the final st<strong>at</strong>e might not be in equilibrium. If<br />

the final st<strong>at</strong>e is not in equilibrium, its value of the internal energy is too high,<br />

and hence the amount of work done is less than expected! The final st<strong>at</strong>e will go<br />

to equilibrium, but since the piston is now fixed, there is no mechanical contact<br />

with the outside world anymore. The only way to get rid of the excess energy is<br />

to dump he<strong>at</strong> into the outside world. The internal energy measures the amount<br />

of work a system can do in a reversible, adiab<strong>at</strong>ic process. The amount of work<br />

done in an irreversible process, starting from an initial st<strong>at</strong>e in equilibrium, is<br />

always less, because he<strong>at</strong> is gener<strong>at</strong>ed. If we assume th<strong>at</strong> the outside world is<br />

always in thermal equilibrium, this indic<strong>at</strong>es th<strong>at</strong> the entropy of the outside<br />

world must increase.<br />

The irreversible expansion is described in two steps. We start <strong>at</strong> volume Vi<br />

and entropy Si, with the system in equilibrium and Ui = U(Si, Vi, N). Next,<br />

the system expands to a volume Vf and entropy Sf . The amount of work done<br />

is ∆W . The system is not in equilibrium, however, and the internal energy


2.2. FREE ENERGIES. 57<br />

is NOT equal to U(Sf , Vf , N), but is equal to Ui − ∆W . The system now<br />

follows a spontaneous process and reaches equilibrium. This is a process <strong>at</strong><br />

constant S and V, and in the end the energy is Uf = U(Sf , Vf , N). The change<br />

in energy is ∆U = Uf − Ui. The value of ∆W depends on the n<strong>at</strong>ure of the<br />

irreversible process, and is in general hard to calcul<strong>at</strong>e. We only know for sure<br />

th<strong>at</strong> Ui −∆W ≥ Uf or ∆W ≤ −∆U. Note th<strong>at</strong> the change in entropy is hard to<br />

predict, since we have in general no formulas to describe non-equilibrium st<strong>at</strong>es.<br />

It will depend on the changes in the temper<strong>at</strong>ure in the second step. Our final<br />

conclusion is th<strong>at</strong>:<br />

The internal energy measures the amount of work a system is able<br />

to do in a reversible adiab<strong>at</strong>ic process. It gives an upper-bound for<br />

the amount of work done in irreversible processes.<br />

2.2 Free energies.<br />

The problem of having too many variables.<br />

The thermodynamic potential governing the adiab<strong>at</strong>ic expansion of a gas is<br />

the internal energy. Other measures of the free energy are needed in different<br />

types of processes. Wh<strong>at</strong> do we need when the system is able to exchange<br />

he<strong>at</strong> or particles with the outside world? In other words, which variables are<br />

controlled? Do we keep the volume or pressure constant?<br />

A construction of the appropri<strong>at</strong>e free energy is straightforward, and will be<br />

illustr<strong>at</strong>ed via an example system. Suppose the only extensive st<strong>at</strong>e variables<br />

which we can measure or change are the volume V and the amount of m<strong>at</strong>erial<br />

N. The work done in such changes depends on the pressure p and the chemical<br />

potential µ. We can write<br />

¯dW = pdV − µdN (2.2)<br />

Note again the difference in sign between the mechanical work term and all others.<br />

In this formula for the work done we always see a combin<strong>at</strong>ion of the differential<br />

of an extensive variable and the corresponding intensive variable. How<br />

can we change th<strong>at</strong> to the product of an extensive variable and the differential<br />

of the corresponding intensive st<strong>at</strong>e variable? The answer is found by considering<br />

the product of corresponding extensive and intensive st<strong>at</strong>e variables. The<br />

differential of such a product is easy to calcul<strong>at</strong>e, since d(xy) = xdy + ydx. This<br />

suggests th<strong>at</strong> we have to add to or subtract from the internal energy terms of the<br />

form pV and µN. Such transform<strong>at</strong>ions are called Legendre transform<strong>at</strong>ions.<br />

Helmholtz free energy.


58 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

As a first example we consider the Helmholtz free energy<br />

F = U − T S (2.3)<br />

By definition, F is a st<strong>at</strong>e function and its differential is exact. Note th<strong>at</strong> some<br />

books use the not<strong>at</strong>ion A for the Helmholtz free energy. The combin<strong>at</strong>ion T S is<br />

a mixture of an intensive variable, T, and the corresponding extensive variable,<br />

S. A small change in the Helmholtz energy is rel<strong>at</strong>ed to changes in he<strong>at</strong> and<br />

work via the first law:<br />

dF = dU − d(T S) = ( ¯ dQ − T dS ) − SdT − ¯ dW (2.4)<br />

In a reversible process ¯ dQ − T dS = 0, and now the interpret<strong>at</strong>ion of the<br />

formula for dF is easy. If we consider a thermodynamic process <strong>at</strong> constant<br />

temper<strong>at</strong>ure, the amount of work done on the outside world is equal to the<br />

decrease in Helmholtz free energy. Hence:<br />

The Helmholtz free energy measures the amount of work a system<br />

can do in an isothermal, reversible process!<br />

If a process is not reversible, the first term in 2.4 does not vanish. For an<br />

irreversible process we have ¯ dQ < T dS and hence for an irreversible isothermal<br />

process dF < − ¯ dW . Hence, like in the case of the internal energy, the<br />

Helmholtz free energy measures the maximal amount of work a system can do<br />

in an isothermal process. In order to do this maximal amount of work, the<br />

process has to be reversible. In addition, for a spontaneous process in which no<br />

work is done, dF < 0. Hence we arrive <strong>at</strong> the following minimum principle:<br />

The Helmholtz free energy of a system <strong>at</strong> constant temper<strong>at</strong>ure is<br />

a minimum as a function of all parameters which are not fixed by<br />

external constraints.<br />

Again, this is very similar to the minimum principle for the internal energy.<br />

The big difference is th<strong>at</strong> we are now considering a process in which we keep the<br />

temper<strong>at</strong>ure constant. This requires in general th<strong>at</strong> the entropy has to change,<br />

which means th<strong>at</strong> he<strong>at</strong> will flow. Therefore, the amount of work th<strong>at</strong> can be<br />

done is different than th<strong>at</strong> can be done <strong>at</strong> constant entropy!<br />

N<strong>at</strong>ural variables.<br />

The change in Helmholtz free energy in a reversible process is given by<br />

dF = −SdT − pdV + µdN (2.5)<br />

for a system which interacts with the outside world through he<strong>at</strong> exchange<br />

(dT), mechanical work (dV), and chemical work (dN). This form suggests th<strong>at</strong>


2.2. FREE ENERGIES. 59<br />

the n<strong>at</strong>ural variables for an expression of the Helmholtz free energy are T, V,<br />

and N. Similarly, for the internal energy U the n<strong>at</strong>ural variables are S, V, and<br />

N.<br />

Why are n<strong>at</strong>ural variables more useful than other combin<strong>at</strong>ions?<br />

Other free energies.<br />

Using the idea of Legendre transform<strong>at</strong>ions many more free energies can be<br />

constructed. Three very useful forms are:<br />

Gibbs potential G = U − T S + pV (2.6)<br />

Enthalpy H = U + pV (2.7)<br />

Grand potential Ω = U − T S − µN (2.8)<br />

All these potentials are st<strong>at</strong>e functions and their differentials are all exact.<br />

The Gibbs potential measures the amount of work a system can do <strong>at</strong> constant<br />

temper<strong>at</strong>ure and pressure. Since we work <strong>at</strong> constant pressure, the volume has<br />

to change and our system will do mechanical work on the outside world. This<br />

amount of energy cannot be used, because it is needed to keep the pressure<br />

constant. Similarly, the system is kept <strong>at</strong> constant temper<strong>at</strong>ure and he<strong>at</strong> will<br />

flow to the outside world. Again, this is energy which cannot be used. Hence<br />

the Gibbs potential measures the amount of non-mechanical (not-pdV ) work a<br />

system can do on the outside world <strong>at</strong> constant pressure and temper<strong>at</strong>ure. If<br />

µdN is the only other term in the expression for ¯ dW , then the Gibbs free energy<br />

is a measure of the amount of chemical work a system is able to do.<br />

The enthalpy measures the amount of work a system can do in an adiab<strong>at</strong>ic,<br />

isobaric process. Again, this is the amount of non-mechanical work, since the<br />

pressure has to be kept constant. The original interpret<strong>at</strong>ion of the enthalpy is<br />

th<strong>at</strong> it measures the amount of he<strong>at</strong> th<strong>at</strong> can be released in a process <strong>at</strong> constant<br />

pressure and constant amount of m<strong>at</strong>erial. The grand potential pertains to processes<br />

<strong>at</strong> constant temper<strong>at</strong>ure and chemical potential. Therefore, it measures<br />

non-chemical work, and in a simple system this is only mechanical, pdV type<br />

of work. The Gibbs potential and the enthalpy are typically used to describe<br />

solids, where one cannot change the volume directly but only through changes<br />

in the pressure. The grand potential is useful in chemistry, where the number of<br />

particles can change due to chemical reactions and where this number can only<br />

be determined by specifying the chemical potential. All these potentials obey a<br />

minimum principle similar to those we saw before.<br />

Summary of five free energies.


60 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

In summary, we introduced five free energies (others can be constructed in<br />

a similar manner):<br />

Symbol Differential (reversible) N<strong>at</strong>ural Name<br />

variables<br />

U dU = T dS − pdV + µdN S,V,N Internal energy<br />

F = U − T S dF = −SdT − pdV + µdN T,V,N Helmholtz free energy<br />

G = F + pV dG = −SdT + V dp + µdN T,p,N Gibbs free energy<br />

H = U + pV dH = T dS + V dp + µdN S,p,N Enthalpy<br />

Ω = F − µN dΩ = −SdT − pdV − Ndµ T,V,µ Grand potential<br />

The internal energy is the only free energy which is expressed in changes<br />

of extensive variables only. All the other energies involve <strong>at</strong> least one intensive<br />

variable. If we prescribe a given set of n<strong>at</strong>ural variables, the equilibrium st<strong>at</strong>e of<br />

a system can be found by minimizing the appropri<strong>at</strong>e free energy as a function<br />

of all internal variables. Which set of n<strong>at</strong>ural variables we use depends on the<br />

system we want to describe.<br />

In a given experiment, independent st<strong>at</strong>e variables are those variables whose<br />

values we can control. This can mean two things. First, it could be th<strong>at</strong> we are<br />

able to set the value of the variable <strong>at</strong> any value we want (often in a certain<br />

range only, though). For example, if we have a container with a piston we<br />

can set the volume of the container <strong>at</strong> arbitrary values by fixing the position<br />

of the piston. A more common way of controlling a st<strong>at</strong>e variable, however,<br />

is to prevent its value from changing! A gas in a closed container has a fixed<br />

volume. Any st<strong>at</strong>e variable which we can control by either setting its value or by<br />

preventing changes in its value is an independent variable. These independent<br />

variables are the variables we select as n<strong>at</strong>ural variables.<br />

General formalism.<br />

Is it possible to write H(T, V, N)?<br />

The general form for the differential of the internal energy is<br />

dU = T dS + ∑<br />

xjdXj<br />

j<br />

(2.9)<br />

which follows from 1.18. A general Legendre transform<strong>at</strong>ion pertaining to an<br />

adiab<strong>at</strong>ic environment would single out the values of j included in a set J and<br />

define<br />

YJ = U − ∑<br />

j∈J<br />

Xjxj<br />

(2.10)


2.3. EULER AND GIBBS-DUHEM RELATIONS. 61<br />

with a corresponding differential<br />

dYJ = T dS − ∑<br />

Xjdxj + ∑<br />

xjdXj<br />

j∈J<br />

j /∈J<br />

(2.11)<br />

In order to change to the temper<strong>at</strong>ure as a n<strong>at</strong>ural variable we also subtract a<br />

term T S. A typical example might be the discussion of magnetism in a solid. We<br />

have no way to fix the total magnetic moment ⃗ M other than using a magnetic<br />

field ⃗ H. Therefore, ⃗ H is the n<strong>at</strong>ural variable for this system, and typically we<br />

use a Gibbs-like energy G, defined by G = U − T S + pV − ⃗ M · ⃗ H, to describe<br />

this system. The differential for G is dG = −SdT + V dp + µdN − ⃗ M · d ⃗ H.<br />

2.3 Euler and Gibbs-Duhem rel<strong>at</strong>ions.<br />

Scaling of the internal energy.<br />

Changes in the internal energy are rel<strong>at</strong>ed to changes in the extensive variables,<br />

and the set of all extensive variables is the n<strong>at</strong>ural set of variables to<br />

describe U. In a general case we have U(S, V, N, X) where X represents all<br />

other extensive variables which can be measured and specified to determine<br />

the n<strong>at</strong>ure of the system. If we consider a homogeneous system, in which all<br />

extensive variables are scaled by a value λ, we can derive the following rel<strong>at</strong>ion:<br />

U(λS, λV, λN, λX) = λU(S, V, N, X) (2.12)<br />

This holds for all values of λ! This rel<strong>at</strong>ion can be differenti<strong>at</strong>ed with respect to<br />

λ. The result involves partial deriv<strong>at</strong>ives of U, and since we have so many choices<br />

for our independent variables, we will always specify the constant variables in<br />

the partial deriv<strong>at</strong>ives. We find<br />

( ) ( ) ( ) ( )<br />

∂U<br />

∂U<br />

∂U<br />

∂U<br />

S +<br />

V +<br />

N +<br />

X = U (2.13)<br />

∂S V,N,X ∂V S,N,X ∂N S,V,X ∂X S,V,N<br />

where we have set λ equal to one.<br />

Since the partial deriv<strong>at</strong>ives involve the n<strong>at</strong>ural st<strong>at</strong>e variables for U, they<br />

can be obtained from the differential for U:<br />

( )<br />

∂U<br />

= T (2.14)<br />

∂S V,N,X<br />

( )<br />

∂U<br />

= −p (2.15)<br />

∂V S,N,X<br />

( )<br />

∂U<br />

= µ (2.16)<br />

∂N<br />

S,V,X


62 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

( )<br />

∂U<br />

= x (2.17)<br />

∂X S,V,N<br />

where x is the generalized force corresponding to X. Therefore we find the important<br />

rel<strong>at</strong>ion (Euler equ<strong>at</strong>ion)<br />

U = T S − pV + µN + xX (2.18)<br />

The other free energies obey similar rel<strong>at</strong>ions, which can be derived immedi<strong>at</strong>ely<br />

from this expression. For example, the Gibbs energy for a system which<br />

can only exchange he<strong>at</strong>, do mechanical work (pdV ) and chemical work (µdN)<br />

is G = µN. This is a very simple expression indeed. In such a case it is often<br />

easier to work with the chemical potential then with the Gibbs energy itself.<br />

Note th<strong>at</strong> expression 2.18 implies th<strong>at</strong> we work in the thermodynamic limit.<br />

We need the freedom to choose x in 2.18 in an appropri<strong>at</strong>e way. This choice<br />

should not change the definition of S, and hence the number of internal degrees<br />

of freedom should be very large. In th<strong>at</strong> case a change in the number of internal<br />

degrees of freedom by a few does not m<strong>at</strong>ter.<br />

Independent variables.<br />

Equ<strong>at</strong>ions 2.14 through 2.17 show th<strong>at</strong> for a system in equilibrium the<br />

values of the temper<strong>at</strong>ure, pressure, etc are completely determined once we<br />

specify all the extensive st<strong>at</strong>e variables. In general, the set of st<strong>at</strong>e variables<br />

{S, T, p, V, µ, N, X, x} contains only four independent variables. The physics of<br />

the system is completely determined by the rel<strong>at</strong>ion U(S,V,N,X) and everything<br />

else follows. Obviously, we can change variables if we want. For example, we<br />

could use {T, V, N, X} as independent variables. The n<strong>at</strong>ural energy to determine<br />

in this case is the Helmholtz free energy F. The partial deriv<strong>at</strong>ives of F<br />

with respect to {T, V, N, X} are straightforward and easy to obtain from the<br />

expression for the differential of F. Since we switch variables so often, we always<br />

want to add the constants to the not<strong>at</strong>ion of the partial deriv<strong>at</strong>ives.<br />

Not all intensive variables are independent.<br />

The differential for dU is rel<strong>at</strong>ed to the changes in all extensive variables. If<br />

we take the differential of 2.18 we also find terms pertaining to changes in the<br />

intensive variables. These additional terms have to sum to zero, and hence<br />

SdT − V dp + Ndµ + Xdx = 0 (2.19)<br />

This is called the Gibbs-Duhem rel<strong>at</strong>ion. In words, not all changes in the intensive<br />

variables are independent. This means th<strong>at</strong> if we change basic variables in<br />

the set {S, T, p, V, µ, N, X, x} it is not possible to choose {T, p, µ, x} as our set of<br />

independent variables. Every other choice is allowed, though. It is not hard to<br />

understand this restriction. The intensive variables are independent of the size


2.3. EULER AND GIBBS-DUHEM RELATIONS. 63<br />

of the system, and cannot specify the size of the system. We need <strong>at</strong> least one<br />

measure of how large the system is. If all intensive variables were independent,<br />

we could, however, choose them as a basic set of st<strong>at</strong>e variables, which is in<br />

contradiction with the previous sentence.<br />

Formul<strong>at</strong>ion in terms of densities.<br />

Another way of looking <strong>at</strong> the dependency of the intensive variables is working<br />

with densities in stead of extensive variables. For example, we could specify<br />

the energy per unit volume or per mole. If we define all extensive variables per<br />

unit volume<br />

we find<br />

s = S<br />

V<br />

, u = U<br />

V<br />

N X<br />

, n = , X =<br />

V V<br />

(2.20)<br />

u = T s − p + µn + xX (2.21)<br />

and using V du + udV = d(uV ) = dU = T d(sV ) − pdV + µd(nV ) + xd(X V ),<br />

which is equal to (T s − p + µn + xX ) dV + (T ds + µdn + xdX ) V , we derive<br />

du = T ds + µdn + xdX (2.22)<br />

Similar formulas hold for densities per particle or other combin<strong>at</strong>ions. Hence<br />

one of the extensive variables can always be scaled out, and does not contribute<br />

to the basic physics of the system. It merely sets the size and is, of course,<br />

needed in th<strong>at</strong> respect.<br />

The obvious extensive variable to use for scale is volume. Volume is somewh<strong>at</strong><br />

special because it is a quantity directly rel<strong>at</strong>ed to the sp<strong>at</strong>ial geometry, and<br />

not to an amount of a physical quantity. The pdV work term has the opposite<br />

sign of all others. Hence there is something fundamentally different in volume<br />

compared to other extensive quantities. But one can also scale out amount of<br />

m<strong>at</strong>erial, and in many applic<strong>at</strong>ions we want to follow a given amount of m<strong>at</strong>erial<br />

and see wh<strong>at</strong> happens. We will return to this in the transport chapter.<br />

Using a Legendre transform<strong>at</strong>ion.<br />

Yet another way of looking <strong>at</strong> this problem is to define the following free<br />

energy via a Legendre transform<strong>at</strong>ion:<br />

Y = U − T S − µN − xX = −pV (2.23)<br />

The free energy Y is a st<strong>at</strong>e function and an extensive quantity. Its n<strong>at</strong>ural<br />

variables are {T, µ, x, V }. Hence we have<br />

Y (T, µ, x, V ) = −V p(T, µ, x, V ) (2.24)


64 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

If we scale the volume to αV and use the extensive n<strong>at</strong>ure of Y, we find<br />

p(T, µ, x, αV ) = p(T, µ, x, V ) (2.25)<br />

for all values of α. If we now take the limit α → 0, we see th<strong>at</strong> the pressure<br />

does not depend on the volume, and th<strong>at</strong> it is a function of the other intensive<br />

variables only.<br />

2.4 Maxwell rel<strong>at</strong>ions.<br />

Basic principle.<br />

One should not memorize Maxwell rel<strong>at</strong>ions, but in stead recognize the basic<br />

principle which determines all these rel<strong>at</strong>ions. If a function of many variables is<br />

twice differentiable, and if the second order deriv<strong>at</strong>ives are continuous, the order<br />

in which we take the second order deriv<strong>at</strong>ives is not important. We apply this<br />

rule to all possible free energies and all possible combin<strong>at</strong>ions of independent<br />

variables. This yields a large number of rel<strong>at</strong>ions, called Maxwell rel<strong>at</strong>ions.<br />

The most important occur when we take each free energy as a function of its<br />

n<strong>at</strong>ural variables, since in th<strong>at</strong> case the first order deriv<strong>at</strong>ives are simple st<strong>at</strong>e<br />

variables. These simple st<strong>at</strong>e variables can be found immedi<strong>at</strong>ely by inspecting<br />

the form of the differential. Since so many combin<strong>at</strong>ions of st<strong>at</strong>e variables are<br />

possible, we always write the constant st<strong>at</strong>e variables in the not<strong>at</strong>ion of the<br />

partial deriv<strong>at</strong>ives.<br />

Simple example.<br />

As an example consider F(T,V,N), with dF = −SdT − pdV + µdN. There<br />

are three choices for the mixed second order deriv<strong>at</strong>ives, leading to the three<br />

Maxwell equ<strong>at</strong>ions:<br />

( ) ( ) ( )<br />

2 ∂ F<br />

∂S<br />

∂p<br />

= −<br />

= −<br />

(2.26)<br />

∂T ∂V N ∂V T,N ∂T V,N<br />

( ) ( ) ( )<br />

2 ∂ F<br />

∂S ∂µ<br />

= −<br />

=<br />

(2.27)<br />

∂T ∂N ∂N ∂T<br />

( ∂ 2 F<br />

∂N∂V<br />

)<br />

V<br />

T<br />

=<br />

( ∂µ<br />

∂V<br />

)<br />

N,T<br />

T,V<br />

= −<br />

( ∂p<br />

∂N<br />

)<br />

V,N<br />

V,T<br />

(2.28)<br />

These rel<strong>at</strong>ions are very important, since they connect changes in one st<strong>at</strong>e<br />

variables to changes in another st<strong>at</strong>e variable in a different process. They limit<br />

the freedom of a system in equilibrium to respond to external causes. Maxwell<br />

rel<strong>at</strong>ions are easy to derive if you know which st<strong>at</strong>e variables are your basic set,<br />

which free energy belongs to th<strong>at</strong> set, and which form the differential has.


2.5. RESPONSE FUNCTIONS. 65<br />

Which rel<strong>at</strong>ions to use?<br />

It is not hard to find out if a Maxwell rel<strong>at</strong>ion can be found, as long as we use<br />

the convention of indic<strong>at</strong>ing) all variables th<strong>at</strong> are kept constant. For example,<br />

consider the deriv<strong>at</strong>ive . The set of variables used is {S, N, V } and this<br />

( ∂p<br />

∂S<br />

N,V<br />

is indeed the set of n<strong>at</strong>ural variables for a free energy, in this case the internal<br />

energy U. In general, we have a good set of variables if from each pair {T, S},<br />

{p, V }, {µ, N}, and {x, X} exactly one variable is used.<br />

The variable which is differenti<strong>at</strong>ed is one of the remaining variables (else<br />

the deriv<strong>at</strong>ive is trivial), and can be found by inspecting the differential of the<br />

free energy for which the set of variables is the set of n<strong>at</strong>ural variables. In our<br />

example we have p = − ( )<br />

∂U<br />

∂V and hence<br />

S,N<br />

( ) ( ) ( )<br />

2 ∂p<br />

∂ U<br />

∂T<br />

= −<br />

= −<br />

(2.29)<br />

∂S N,V ∂V ∂S N ∂V N,S<br />

2.5 Response functions.<br />

Linear response.<br />

How do we investig<strong>at</strong>e the physics of some system? How do we perform<br />

a basic measurement? We apply an external field, and measure the response<br />

of the system. For example, we change the temper<strong>at</strong>ure of a metal rod and<br />

measure the change in length. Life would be simple, and thermodynamics easy,<br />

if all responses were linear. This is not true, unfortun<strong>at</strong>ely. A linear rel<strong>at</strong>ion<br />

between cause and effect is only true in the limit of small changes of the causing<br />

agents, when we can write<br />

( )<br />

∂A<br />

∆A =<br />

∆B (2.30)<br />

∂B C,D,..<br />

for a change in B, keeping C,D,... constant. We do not know how small ∆B<br />

has to be in order for a linear rel<strong>at</strong>ion to be a good approxim<strong>at</strong>ion. Ohm’s law<br />

holds for metals for large changes in the applied voltage, but for semiconductors<br />

the linear range is very small. If we measure the magnetic susceptibility of iron,<br />

a change in temper<strong>at</strong>ure of 1K <strong>at</strong> room temper<strong>at</strong>ure is small, because the Curie<br />

temper<strong>at</strong>ure of iron is very large. Near the Curie point, however, changes are<br />

more rapid since some of the deriv<strong>at</strong>ives start to diverge.<br />

When is a linear approxim<strong>at</strong>ion to the response of a system good enough?<br />

The linear rel<strong>at</strong>ion 2.30 holds for small changes in B, and by measuring A we<br />

can find the partial deriv<strong>at</strong>ive. Since the partial deriv<strong>at</strong>ive measures the response


66 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

in A of the system to a small change in B, it is called a response function. The<br />

most important way we obtain inform<strong>at</strong>ion about thermodynamic systems is by<br />

measuring response functions. Hence measurements give us response functions,<br />

and the analysis of the behavior of response functions then will allow us to draw<br />

conclusions about the physics of the system!<br />

He<strong>at</strong> capacities.<br />

He<strong>at</strong> capacities are one of the most important thermodynamic response functions<br />

to measure. The tell us how much he<strong>at</strong> the system will take up if we increase<br />

its temper<strong>at</strong>ure. These measurements are always performed as a function<br />

of temper<strong>at</strong>ure, but can be done either <strong>at</strong> fixed pressure or volume. The other<br />

extensive thermodynamic variables kept constant are denoted by X. Hence we<br />

have<br />

( ) ( )<br />

∆Q<br />

∂S<br />

CV (T, V, X) =<br />

= T<br />

(2.31)<br />

∆T<br />

∂T<br />

Cp(T, p, X) =<br />

( ∆Q<br />

∆T<br />

)<br />

V,X<br />

p,X<br />

= T<br />

( ∂S<br />

∂T<br />

)<br />

V,X<br />

p,X<br />

(2.32)<br />

The he<strong>at</strong> capacities are rel<strong>at</strong>ed to the partial deriv<strong>at</strong>ives of the entropy, since<br />

the entropy governs he<strong>at</strong> exchange. The he<strong>at</strong> capacity <strong>at</strong> constant volume is<br />

directly rel<strong>at</strong>ed to the internal energy. Since we know th<strong>at</strong> for small changes<br />

∆U = T ∆S −p∆V +x∆X, we can write for a process in which the temper<strong>at</strong>ure<br />

changes by a small amount ∆T , but the volume V and the variables X are kept<br />

constant,<br />

∆U<br />

∆T ⎪V,X = T ∆S<br />

∆T ⎪V,X In the limit ∆T → 0 the r<strong>at</strong>ios become partial deriv<strong>at</strong>ives, and we have<br />

( )<br />

∂U<br />

CV (T, V, X) =<br />

∂T<br />

V,X<br />

(2.33)<br />

(2.34)<br />

Similarly, for a process <strong>at</strong> constant pressure and X, the differential of U<br />

yields<br />

and hence<br />

∆U<br />

∆T ⎪p,X Cp(T, p, X) =<br />

= T ∆S<br />

∆T ⎪<br />

− p<br />

p,X<br />

∆V<br />

∆T ⎪p,X ( ) ( )<br />

∂U<br />

∂V<br />

+ p<br />

∂T p,X ∂T p,X<br />

(2.35)<br />

(2.36)<br />

Since most m<strong>at</strong>erials expand when he<strong>at</strong>ed, the volume change in the last case<br />

will cause the system to do work and we have to supply this energy. We expect


2.5. RESPONSE FUNCTIONS. 67<br />

therefore Cp > CV . Also, in general we need he<strong>at</strong> to increase the temper<strong>at</strong>ure, so<br />

we also expect CV > 0. These st<strong>at</strong>ements will be proven l<strong>at</strong>er in the this chapter,<br />

where we talk about equilibrium conditions. He<strong>at</strong> capacities are extensive st<strong>at</strong>e<br />

functions. One often defines a specific he<strong>at</strong> by using the he<strong>at</strong> capacity per unit<br />

volume or per unit mass. Unfortun<strong>at</strong>ely, the word specific he<strong>at</strong> is used for both<br />

these r<strong>at</strong>ios, so make sure th<strong>at</strong> you understand from the context of the problem<br />

which one applies.<br />

Compressibilities.<br />

Another set of response functions which are often measured are rel<strong>at</strong>ed to<br />

volume changes, and called compressibilities. We apply pressure to a system<br />

and measure the change in volume. A measurement like this is done either<br />

adiab<strong>at</strong>ically or isothermally, and the two relevant compressibilities are<br />

κT (T, p, X) = − 1<br />

V<br />

κS(S, p, X) = − 1<br />

V<br />

( )<br />

∂V<br />

∂p T,X<br />

( )<br />

∂V<br />

∂p<br />

S,X<br />

(2.37)<br />

(2.38)<br />

We expect κT > κS > 0 because <strong>at</strong> constant entropy the internal variables<br />

of the system are essentially only scaled to a smaller volume, but <strong>at</strong> constant<br />

temper<strong>at</strong>ure they can also rearrange to allow for an even smaller volume. Again,<br />

these inequalities follow from the equilibrium conditions l<strong>at</strong>er in the this chapter.<br />

The minus sign is included to indic<strong>at</strong>e th<strong>at</strong> the volume decreases with increasing<br />

pressure. The size of the system is scaled out by dividing by the volume.<br />

Coefficient of thermal expansion.<br />

Finally, we often use the coefficient of thermal expansion<br />

α(T, p, X) = 1<br />

V<br />

( )<br />

∂V<br />

∂T p,X<br />

(2.39)<br />

Again, we divide by the volume to scale out the system size. We also expect<br />

α to be positive. This is, however, not always true. There are m<strong>at</strong>erials with<br />

a neg<strong>at</strong>ive coefficient of thermal expansion. These m<strong>at</strong>erials are exceptions,<br />

however.<br />

Response functions for ideal gas.<br />

It is useful to calcul<strong>at</strong>e the values of the response functions for a mono<strong>at</strong>omic<br />

ideal gas. This gives us a standard, a reference for comparison if we<br />

consider other systems. Using U = 3<br />

2pV and pV = NRT we find


68 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

CV = 3<br />

NR (2.40)<br />

2<br />

Cp = 5<br />

NR (2.41)<br />

2<br />

κT = 1<br />

p<br />

κS = 3<br />

5p<br />

α = 1<br />

T<br />

(2.42)<br />

(2.43)<br />

(2.44)<br />

The formula for κS follows from the formula for the entropy of an ideal gas<br />

3 − 1.58, which implies th<strong>at</strong> <strong>at</strong> constant entropy V ∝ T 2 . Together with the ideal<br />

3 − gas law this gives V ∝ p 5 . The ideal gas values indeed s<strong>at</strong>isfy κT > κS > 0<br />

and Cp > CV > 0.<br />

Compressibilities are often given in a different form, by st<strong>at</strong>ing the rel<strong>at</strong>ion<br />

between pressure and volume in the form pV γ = c, where c is a constant and<br />

the value of γ depends on the process we consider. In an isothermal process <strong>at</strong><br />

constant N we see from pV = NRT th<strong>at</strong> γ = 1, and in an adiab<strong>at</strong>ic process <strong>at</strong><br />

constant N we have γ = 5<br />

3 .<br />

There are some obvious problems with these numbers. We have seen before<br />

th<strong>at</strong> <strong>at</strong> low temper<strong>at</strong>ure the he<strong>at</strong> capacities should approach zero, clearly not<br />

the case for the ideal gas. Also, <strong>at</strong> low temper<strong>at</strong>ure the expansion coefficient<br />

becomes infinite. This is due to the fact th<strong>at</strong> the volume approaches zero.<br />

Also, the compressibilities diverge <strong>at</strong> zero pressure. All these facts point <strong>at</strong> the<br />

limit<strong>at</strong>ions of the model of an ideal gas. At low temper<strong>at</strong>ure this model is not<br />

valid.<br />

Why is the ideal gas not a good model <strong>at</strong> low temper<strong>at</strong>ure?<br />

2.6 Rel<strong>at</strong>ions between partial deriv<strong>at</strong>ives.<br />

Two deriv<strong>at</strong>ives.<br />

<strong>Thermodynamics</strong> makes extensive use of the theory of calculus of many<br />

variables. This is a very powerful m<strong>at</strong>hem<strong>at</strong>ical theory! It limits the changes<br />

in dependent variables. For example, response functions are partial deriv<strong>at</strong>ives,<br />

and they are rel<strong>at</strong>ed via Maxwell rel<strong>at</strong>ions if they are also the second order


2.6. RELATIONS BETWEEN PARTIAL DERIVATIVES. 69<br />

deriv<strong>at</strong>ives of a free energy. M<strong>at</strong>hem<strong>at</strong>ics proves other rel<strong>at</strong>ions between partial<br />

deriv<strong>at</strong>ives too. For example, assume we have a st<strong>at</strong>e function z(x, y). It is<br />

possible to define a function f(x, y, z) in such a way th<strong>at</strong> f(x, y, z) = 0 leads<br />

exactly to this function. Hence f(x, y, z(x, y)) ≡ 0 for all values of x and y. The<br />

partial deriv<strong>at</strong>ive of this identity with respect to x is<br />

0 =<br />

( )<br />

∂f<br />

(x, y, z(x, y)) +<br />

∂x y,z<br />

( )<br />

∂f<br />

(x, y, z(x, y))<br />

∂z x,y<br />

( )<br />

∂z<br />

(x, y) (2.45)<br />

∂x y<br />

In a similar way, we can use f(x, y, z) = 0 to define a function x(y, z), with<br />

f(x(y, z), y, z) ≡ 0. The partial deriv<strong>at</strong>ive of this equ<strong>at</strong>ion with respect to z is:<br />

0 =<br />

( )<br />

∂f<br />

(x(y, z), y, z)<br />

∂x y,z<br />

( )<br />

∂x<br />

(y, z) +<br />

∂z y<br />

( )<br />

∂f<br />

(x(y, z), y, z) (2.46)<br />

∂z x,y<br />

Comparing these two rel<strong>at</strong>ions shows th<strong>at</strong> for any point (x0, y0, z0) obeying<br />

f(x0, y0, z0) = 0 we have<br />

( ) ( )<br />

∂z<br />

∂x<br />

(x0, y0) (y0, z0) = 1 (2.47)<br />

∂x y ∂z y<br />

when the partial deriv<strong>at</strong>ives of f(x, y, z) are zero in isol<strong>at</strong>ed points only, and<br />

the partial deriv<strong>at</strong>ives of x and z are continuous. Hence given a st<strong>at</strong>e function<br />

z(x, y) we can use z as a st<strong>at</strong>e variable and change variables to {y, z} , which<br />

defines a new st<strong>at</strong>e function x(y, z). Partial deriv<strong>at</strong>ives of these two functions<br />

are rel<strong>at</strong>ed.<br />

Three deriv<strong>at</strong>ives.<br />

A more surprising result is obtained when all three combin<strong>at</strong>ions of independent<br />

variables are used. Using {x, y} as independent variables we have as<br />

before<br />

−<br />

( )<br />

∂f<br />

(x, y, z(x, y)) =<br />

∂x y,z<br />

In a similar way we obtain<br />

−<br />

−<br />

( )<br />

∂f<br />

(x(y, z), y, z) =<br />

∂y x,z<br />

( )<br />

∂f<br />

(x, y(x, z), z) =<br />

∂z x,y<br />

( )<br />

∂f<br />

(x, y, z(x, y))<br />

∂z x,y<br />

( )<br />

∂f<br />

(x(y, z), y, z)<br />

∂x y,z<br />

( )<br />

∂f<br />

(x, y(x, z), z)<br />

∂y x,z<br />

( )<br />

∂z<br />

(x, y) (2.48)<br />

∂x y<br />

( )<br />

∂x<br />

(y, z) (2.49)<br />

∂y z<br />

( )<br />

∂y<br />

(x, z) (2.50)<br />

∂z x


70 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

Multiplying these equ<strong>at</strong>ions leads to<br />

( ) ( ) ( )<br />

∂x<br />

∂y<br />

∂z<br />

(y0, z0) (x0, z0) (x0, y0) = −1 (2.51)<br />

∂y z ∂z x ∂x y<br />

for all points (x0, y0, z0) obeying f(x0, y0, z0) = 0, with similar assumptions one<br />

the partial deriv<strong>at</strong>ives as before. This rel<strong>at</strong>ion is surprising because of the minus<br />

sign. It is often found in the form<br />

( ∂x<br />

∂y<br />

)<br />

(y0, z0) = −<br />

z<br />

( )<br />

∂x<br />

(y0, z0)<br />

∂z y<br />

( )<br />

∂z<br />

(x0, y0) (2.52)<br />

∂y x<br />

This rule can be generalized to arbitrary numbers of variables. It is not hard<br />

to show th<strong>at</strong> a minus sign always occurs when the total number of variables is<br />

odd.<br />

Difference of he<strong>at</strong> capacities.<br />

There are subtle differences in semantics between questions posed by physicists<br />

and m<strong>at</strong>hem<strong>at</strong>icians, and it is important to realize them. If a physicist<br />

asks the question wh<strong>at</strong> is U(T, p, N) if U(T, V, N) = N( 3 N<br />

2RT + V (T − T0)) it is<br />

understood th<strong>at</strong> we try to find a way th<strong>at</strong> allows us to find first V from T, p, N<br />

and then use the formula above. We are looking for two different mappings, and<br />

need to find (T, p, N) → U if we know (T, V, N) → U. We denote the function<br />

by the outcome, which is the same in both cases. M<strong>at</strong>hem<strong>at</strong>icians would simply<br />

respond U(T, p, N) = N( 3 N<br />

2RT + p (T − T0)), since it is a simple substitution of<br />

variables. Like always, we ask questions but assume a specific context. In the<br />

following we will be clear, though, and use the m<strong>at</strong>hem<strong>at</strong>ical distinction, which<br />

says find U2(T, p, N) if we know U1(T, V, N). It is useful, believe me.<br />

As an example of changing variables, consider the st<strong>at</strong>e variables temper<strong>at</strong>ure<br />

T, pressure p, volume V, and the extensive variable(s) X. Considering<br />

the entropy as a function of {T, V, X} gives S1(T, V, X) and considering the<br />

entropy as a function of {T, p, X} gives a different function S2(T, p, X). These<br />

two functions are clearly rel<strong>at</strong>ed by<br />

S2(T, p, X) = S1(T, V (T, p, X), X) (2.53)<br />

where V(T,p,X) is the function for the equilibrium volume with {T, p, X} as independent<br />

variables. Remember th<strong>at</strong> all these rel<strong>at</strong>ions only hold for equilibrium<br />

systems! Therefore the partial deriv<strong>at</strong>ives are rel<strong>at</strong>ed by<br />

( ∂S2<br />

∂T<br />

)<br />

p,X<br />

=<br />

Using the Maxwell rel<strong>at</strong>ion<br />

( )<br />

∂S<br />

∂V<br />

T,X<br />

( ∂S1<br />

∂T<br />

)<br />

V,X<br />

+<br />

( ∂S1<br />

∂V<br />

)<br />

( ) 2 ∂ F<br />

= −<br />

=<br />

∂V ∂T X<br />

T,X<br />

( ∂V<br />

∂T<br />

( )<br />

∂p<br />

∂T V,X<br />

)<br />

p,X<br />

(2.54)<br />

(2.55)


2.6. RELATIONS BETWEEN PARTIAL DERIVATIVES. 71<br />

and the chain rule<br />

we then derive<br />

( ) ( ) ( )<br />

∂p<br />

∂p ∂V<br />

= −<br />

∂T V,X ∂V T,X ∂T p,X<br />

V T α2<br />

Cp − CV =<br />

which tells us th<strong>at</strong> Cp > CV indeed if κT > 0.<br />

Important points.<br />

κT<br />

(2.56)<br />

(2.57)<br />

In this deriv<strong>at</strong>ion we did two things. First, we write these equ<strong>at</strong>ions in<br />

terms of functions, leaving out the variables, since they are all contained in<br />

the partial deriv<strong>at</strong>ives. For example, ( )<br />

∂S<br />

∂V T,X clearly is rel<strong>at</strong>ed to S1(T, V, X),<br />

)<br />

while would imply S2(T, p, X). Even though in the partial deriv<strong>at</strong>ives<br />

( ∂S<br />

∂p<br />

T,X<br />

we write the same symbol S in the top part, which is correct since in both cases<br />

the result of the function is the entropy, the functional form is understood to be<br />

different since the variables are different. This is probably the most confusing<br />

aspect of the not<strong>at</strong>ion used in calculus of many variables. Second, there are<br />

several ways to get from the beginning to the end. We started by using 2.55<br />

since often it is a good idea to get rid of the entropy. In order to do th<strong>at</strong> we<br />

either use the definition of he<strong>at</strong> capacity, or a Maxwell rel<strong>at</strong>ion.<br />

R<strong>at</strong>ios of response functions.<br />

In a similar way we find<br />

V T α2<br />

κT − κS =<br />

Cp<br />

and from the r<strong>at</strong>io of these expressions we find<br />

Cp<br />

CV<br />

= κT<br />

κS<br />

(2.58)<br />

(2.59)<br />

which is a remarkably simple rel<strong>at</strong>ion between very different response functions.<br />

Note th<strong>at</strong> the last three formulas hold for all systems, temper<strong>at</strong>ures, pressures,<br />

etc. Of course, on has to measure Cp and CV <strong>at</strong> the same temper<strong>at</strong>ure and pressure<br />

in order to be able to compare them. These rel<strong>at</strong>ions are experimentally<br />

confirmed. The most important message here is th<strong>at</strong> many response functions<br />

are dependent and such dependencies give a good consistency check for experimental<br />

results. The ideal gas values do obey these rel<strong>at</strong>ions, as can be easily<br />

checked. The fact th<strong>at</strong> such rel<strong>at</strong>ions exist is based on the fact th<strong>at</strong> the function


72 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

U(S, V, N, X) gives a complete description of the system and th<strong>at</strong> all other st<strong>at</strong>e<br />

variables follow from this function.<br />

Check th<strong>at</strong> 2.59 holds for the ideal gas.<br />

Equ<strong>at</strong>ion 2.59 has an interesting consequence. Suppose we have an ideal gas,<br />

but this gas is not mono-<strong>at</strong>omic. In th<strong>at</strong> case we still have pV = NRT , but the<br />

energy versus temper<strong>at</strong>ure rel<strong>at</strong>ion is more complic<strong>at</strong>ed. The isothermal compressibility<br />

κT ia still equal to 1<br />

p , which means th<strong>at</strong> the adiab<strong>at</strong>ic compressibility<br />

is given by:<br />

or<br />

− 1<br />

V<br />

( )<br />

∂V<br />

∂p S,N<br />

= κS = 1 CV<br />

p Cp<br />

( )<br />

∂V<br />

= −<br />

∂p S,N<br />

V CV<br />

p Cp<br />

(2.60)<br />

(2.61)<br />

Suppose th<strong>at</strong> we have a range of volumes for which the temper<strong>at</strong>ure changes<br />

while compressing are such th<strong>at</strong> the r<strong>at</strong>io CV remains constant. Such a range<br />

Cp<br />

is often found for poly<strong>at</strong>omic gases. For example, <strong>at</strong> some intermedi<strong>at</strong>e temper<strong>at</strong>ures<br />

all rot<strong>at</strong>ional degrees of freedom might be active, but the vibr<strong>at</strong>ional<br />

are not. In th<strong>at</strong> case we find for adiab<strong>at</strong>ic compression th<strong>at</strong> p ∝ V −γ with<br />

γ = Cp<br />

. The coefficient of thermal expansion for this poly-<strong>at</strong>omic gas is still<br />

CV<br />

determined by pV = NRT , and hence α = 1<br />

T . Equ<strong>at</strong>ion 2.57 then tells us th<strong>at</strong><br />

Cp − CV = NR, and hence γ = 1 + NR . Therefore, measuring the compress-<br />

CV<br />

ibility gives us the he<strong>at</strong> capacity, from which we can derive inform<strong>at</strong>ion about<br />

the internal degrees of freedom of the molecules!<br />

2.7 Conditions for equilibrium.<br />

P<strong>at</strong>h to equilibrium.<br />

Wh<strong>at</strong> happens when two equilibrium systems are brought into contact? Obviously,<br />

the combined system will try to reach thermal equilibrium. Wh<strong>at</strong> are<br />

the conditions for this total equilibrium? These conditions are easy to derive<br />

if the combined system is completely isol<strong>at</strong>ed or closed, and cannot exchange<br />

energy in the form of work with the rest of the world. He<strong>at</strong> exchange is possible,<br />

though. This means th<strong>at</strong> the total volume, amount of m<strong>at</strong>erial, etc. is constant.<br />

We also assume th<strong>at</strong> we can keep the entropy of the system constant. This does<br />

not conflict with the st<strong>at</strong>ement th<strong>at</strong> he<strong>at</strong> transfer is possible, since we consider<br />

non-equilibrium systems!


2.7. CONDITIONS FOR EQUILIBRIUM. 73<br />

The two subsystems are able to exchange energy by changing their individual<br />

volumes (⇒ mechanical work), amount of m<strong>at</strong>erial (⇒ chemical work), or<br />

entropies (⇒ he<strong>at</strong> exchange). We will label the two subsystems l for left and r<br />

for right. The total internal energy is<br />

with<br />

Ut = Ur + Ul<br />

(2.62)<br />

Ut = U(S, V, N, X) (2.63)<br />

Ul = U(Sl, Vl, Nl, Xl) (2.64)<br />

Ur = U(S − Sl, V − Vl, N − Nl, X − Xl) (2.65)<br />

where we use the symbol U for the function th<strong>at</strong> gives the values Ul, Ur, Ut for<br />

the (sub)systems, and where X stands for all other extensive variables. Since<br />

the total system is closed, the total internal energy in equilibrium will be a<br />

minimum. Also the total sum of each extensive variable cannot change, and<br />

we have S = Sl + Sr, etc. The equilibrium st<strong>at</strong>e follows by minimizing the<br />

internal energy as a function of the four internal (for the total system) variables<br />

{Sl, Vl, Nl, Xl}. This leads to four equ<strong>at</strong>ions of the form<br />

( )<br />

∂U<br />

(Sl, Vl, Nl, Xl) −<br />

∂S V,N,X<br />

( )<br />

∂U<br />

(S − Sl, V − Vl, N − Nl, X − Xl) = 0<br />

∂S V,N,X<br />

(2.66)<br />

Using the expression for the differential dU we then see immedi<strong>at</strong>ely th<strong>at</strong><br />

the conditions for equilibrium are:<br />

Tl = Tr, pl = pr, µl = µr, xl = xr<br />

(2.67)<br />

All intensive variables have to be the same. This is of course exactly wh<strong>at</strong> we<br />

expected.<br />

Using the minimum principle.<br />

At this point we only used the fact th<strong>at</strong> the internal energy is an extremum<br />

<strong>at</strong> equilibrium. We also know th<strong>at</strong> is should be a minimum, and this gives a<br />

number of very useful inequalities for the second order deriv<strong>at</strong>ives. Assume th<strong>at</strong><br />

both subsystems are identical. Assume th<strong>at</strong> the total entropy is 2S, the total<br />

volume 2V, etc. Using identical systems is just for convenience, but does not<br />

affect the general conclusion. The entropy of the system on the left is S + ∆S<br />

and for the system on the right S − ∆S. The volume and all other extensive<br />

quantities are defined in a similar way. The difference in internal energy of the<br />

total system between equilibrium (∆S = 0, etc) and non-equilibrium is:


74 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

∆U = U(S + ∆S, V + ∆V, N + ∆N, X + ∆X)+<br />

U(S − ∆S, V − ∆V, N − ∆N, X − ∆X) − 2U(S, V, N, X) (2.68)<br />

The changes ∆S, ∆V , ∆N, and ∆X in this formula are independent and<br />

can be chosen arbitrarily. The total energy is a minimum as a function of the<br />

internal variables {∆S, ∆V, ∆N, ∆X}, and hence ∆U ≥ 0. Since the values of<br />

{S, V, N, X} are arbitrary, this means th<strong>at</strong> U(S, V, N, X) is a convex function.<br />

Consequences for second order deriv<strong>at</strong>ives.<br />

The previous discussion was true for any value of the parameters ∆S, etc.<br />

When these changes are small, however, we can use a Taylor series expansion for<br />

the energy of the subsystems. The zeroth order term vanishes by construction.<br />

The first order term gives us the equilibrium conditions for the intensive st<strong>at</strong>e<br />

variables we derived before. Hence we can write T for the temper<strong>at</strong>ure of the<br />

whole system and do not have to specify left or right. The same holds for all<br />

other intensive variables. The second order terms form an expression of the<br />

form<br />

( 2 ∂ U<br />

∆U =<br />

∂S2 )<br />

(∆S)<br />

V,N,X<br />

2 ( ) 2 ∂ U<br />

+<br />

∆S∆V + · · · (2.69)<br />

∂S∂V N,X<br />

This expression has to be convex. By changing only one extensive variable <strong>at</strong> a<br />

time we find the necessary inequalities on the st<strong>at</strong>e function U(S, V, N, X) of a<br />

single system:<br />

( ) 2 ∂ U<br />

≥ 0 (2.70)<br />

∂S 2<br />

( ) 2 ∂ U<br />

∂V 2<br />

( ) 2 ∂ U<br />

∂N 2<br />

( 2 ∂ U<br />

∂X 2<br />

)<br />

V,N,X<br />

S,N,X<br />

S,V,X<br />

S,V,N<br />

≥ 0 (2.71)<br />

≥ 0 (2.72)<br />

≥ 0 (2.73)<br />

This is not all, however. Suppose we only change S and V, in which case we<br />

have<br />

∆U =<br />

( 2 ∂ U<br />

∂S2 )<br />

(∆S)<br />

V,N,X<br />

2 ( ) 2 ∂ U<br />

+2<br />

∆S∆V +<br />

∂S∂V N,X<br />

( 2 ∂ U<br />

∂V 2<br />

)<br />

(∆V )<br />

S,N,X<br />

2 ≥ 0<br />

(2.74)<br />

If we divide by (∆V ) 2 we have an inequality of the form ax 2 + bx + c ≥ 0, which<br />

has to hold for all values of x. We already determined th<strong>at</strong> a ≥ 0 and c ≥ 0.


2.7. CONDITIONS FOR EQUILIBRIUM. 75<br />

The form ax 2 + bx + c has two distinct roots when b 2 > 4ac. In our case we<br />

therefore need b 2 ≤ 4ac, or<br />

( ( ) ) 2<br />

2 ∂ U<br />

≤<br />

∂S∂V N,X<br />

He<strong>at</strong> capacities are positive.<br />

( 2 ∂ U<br />

∂S2 )<br />

V,N,X<br />

( 2 ∂ U<br />

∂V 2<br />

)<br />

S,N,X<br />

(2.75)<br />

These inequalities have ( a number of consequences for the signs of response<br />

2<br />

∂ U<br />

functions. For example, ∂S2 )<br />

=<br />

V,N,X<br />

( )<br />

∂T<br />

T<br />

∂S = . Since the tempera-<br />

V,N,X CV<br />

ture is positive (the existence of neg<strong>at</strong>ive temper<strong>at</strong>ure values viol<strong>at</strong>es the second<br />

law), we find<br />

CV ≥ 0 (2.76)<br />

or the specific he<strong>at</strong> <strong>at</strong> constant volume must be positive. This can be rel<strong>at</strong>ed<br />

to a condition for local stability. Consider a small volume in a large system in<br />

equilibrium. Due to some fluctu<strong>at</strong>ion the energy inside this volume is larger<br />

than average. If we would have CV < 0, the temper<strong>at</strong>ure of this volume would<br />

be less than average! Due to the second law, energy would flow into this volume,<br />

making this fluctu<strong>at</strong>ion in energy even larger. Therefore a system with CV < 0 is<br />

unstable against fluctu<strong>at</strong>ions in the energy. Stability therefore requires positive<br />

he<strong>at</strong> capacities <strong>at</strong> constant volume.<br />

Compressibilities are positive.<br />

Similarly, we find from the condition for<br />

(<br />

2<br />

∂ U<br />

∂V 2<br />

)<br />

S,N,X th<strong>at</strong><br />

κS ≥ 0 (2.77)<br />

Compressibilities also have to be positive because of a local stability argument.<br />

Suppose a small volume element has a pressure less than average. The outside<br />

environment has a larger pressure and hence will the imbalance of forces will<br />

reduce the volume of the small element. If the compressibility would be neg<strong>at</strong>ive,<br />

the inside pressure would reduce even more and the process would run<br />

away. Therefore, a neg<strong>at</strong>ive compressibility implies an instability against small<br />

fluctu<strong>at</strong>ions in the pressure.<br />

The more m<strong>at</strong>erial we have, the harder is it to add some.<br />

The stability requirement for the chemical potential is<br />

( )<br />

∂µ<br />

∂N S,V<br />

≥ 0 (2.78)


76 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

which implies th<strong>at</strong> the energy µ it takes to add one unit mass of m<strong>at</strong>erial to<br />

a system increases when the amount of m<strong>at</strong>erial increases. It is harder to add<br />

m<strong>at</strong>erial to a system which already contains a lot! This is again connected with<br />

a local stability requirement for fluctu<strong>at</strong>ions in the number of particles.<br />

Finally we can use equ<strong>at</strong>ion 2.75 to get<br />

T<br />

V κSCV<br />

≥<br />

( ( ) ) 2 ( ( ) ) 2<br />

2 ∂ U<br />

∂T<br />

=<br />

∂S∂V N,X ∂V S,N,X<br />

The chain rule gives<br />

( ) ( )<br />

∂T<br />

∂T<br />

= −<br />

∂V S,N,X ∂S V,N,X<br />

( )<br />

∂S<br />

∂V T,N,X<br />

(2.79)<br />

(2.80)<br />

The first factor on the right hand side is T , while the second is transformed<br />

CV<br />

according to the Maxwell equ<strong>at</strong>ion<br />

( ∂S<br />

∂V<br />

)<br />

T,N,X<br />

Another applic<strong>at</strong>ion of the chain rule gives<br />

( ) ( )<br />

∂p<br />

∂p<br />

= −<br />

∂T<br />

∂V<br />

or<br />

V,N,X<br />

Combining everything gives<br />

=<br />

( )<br />

∂p<br />

∂T V,N,X<br />

T,N,X<br />

( )<br />

∂V<br />

∂T p,N,X<br />

(2.81)<br />

(2.82)<br />

( )<br />

∂p<br />

= −<br />

∂T V,N,X<br />

−1<br />

αV (2.83)<br />

V κT<br />

T<br />

V κSCV<br />

(<br />

αT<br />

≥<br />

CV κT<br />

) 2<br />

(2.84)<br />

Multiplying these inequality by C2 V V<br />

T , which is positive, and using 2.57 leads to<br />

CV<br />

≥<br />

κS<br />

Cp − CV<br />

κT<br />

With the help of 2.76 and 2.59 we then find<br />

and hence due to 2.57<br />

and using 2.58<br />

(2.85)<br />

κT ≥ 0 (2.86)<br />

Cp ≥ CV ≥ 0 (2.87)<br />

κT ≥ κS ≥ 0 (2.88)


2.7. CONDITIONS FOR EQUILIBRIUM. 77<br />

The inequalities for the he<strong>at</strong> capacities and compressibilities hold for all systems.<br />

They have been derived without any knowledge of the function U(S,V,N,X),<br />

apart from the requirements of continuity of the second order deriv<strong>at</strong>ives and<br />

the restrictions due to the laws of thermodynamics. These are remarkable consequences<br />

indeed, based on minimal input.<br />

Would it have been possible to derive these inequalities using the maximum<br />

entropy principle?<br />

Note th<strong>at</strong> we can write the difference between the he<strong>at</strong> capacities in the<br />

following form:<br />

( ) ( )<br />

∂U ∂U<br />

Cp − CV =<br />

−<br />

+ αpV (2.89)<br />

∂T p,X ∂T V,X<br />

and this expression is always positive, even if the coefficient of thermal expansion<br />

α is neg<strong>at</strong>ive! In such a case the internal energy <strong>at</strong> constant pressure has to<br />

increase rapidly as a function of temper<strong>at</strong>ure. This has consequences for the<br />

microscopic description of neg<strong>at</strong>ive thermal expansion m<strong>at</strong>erials!<br />

On paradoxes.<br />

Consider a system where a thermally insul<strong>at</strong>ed piston can move back and<br />

forth in a closed cylinder. We have identical amounts of m<strong>at</strong>erial in both parts<br />

of the cylinder, but they start out <strong>at</strong> different temper<strong>at</strong>ures. At the start, both<br />

parts of the system are in thermal equilibrium, but they are not in equilibrium<br />

with each other. Wh<strong>at</strong> do you expect th<strong>at</strong> the equilibrium st<strong>at</strong>e is? The common<br />

answer is in the middle, with the temper<strong>at</strong>ures and pressures on both sides equal<br />

to each other. Well, people have even done simul<strong>at</strong>ions showing th<strong>at</strong> this is the<br />

case, so it must be true! But if you read the system description very carefully<br />

you will see th<strong>at</strong> this is not true. Both parts of the system are thermally<br />

isol<strong>at</strong>ed. Assume th<strong>at</strong> we take a p<strong>at</strong>h from the initial st<strong>at</strong>e to the final st<strong>at</strong>e<br />

along which each part of the system is always in equilibrium. The initial values<br />

of the entropy of the left and right sides are Sl and Sr, and these could be<br />

different. Equilibrium is obtained when<br />

pl = −<br />

( )<br />

( )<br />

∂U<br />

∂U<br />

(Vl, Sl, N) = pr = − (Vr, Sr, N) (2.90)<br />

∂V S,N<br />

∂V S,N<br />

with V = Vl + Vr. Only when Sl = Sr does this lead to a stable situ<strong>at</strong>ion with<br />

the piston in the middle.<br />

A paradox is a result which seems to contradict common experience. They<br />

often arise from either not giving all inform<strong>at</strong>ion (like in our example), or by<br />

making a mistake in the analysis. Careful analysis will always show where we


78 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

went wrong. For example, in the Gibbs paradox we obtain wrong results if<br />

we omit the N dependence of the constant of integr<strong>at</strong>ion. Unfortun<strong>at</strong>ely, in<br />

thermodynamics we always deal with many variables, and the opportunity to<br />

cre<strong>at</strong>e paradoxes is large. But as long as we always first determine the set of<br />

independent st<strong>at</strong>e variables for a given problem, we will avoid many of these<br />

paradoxes.<br />

2.8 Stability requirements on other free energies.<br />

The conditions for equilibrium discussed in the previous section can be generalized<br />

to all thermodynamic potentials as a function of extensive variables. It<br />

does not hold for intensive variables, though. As an example, consider the Gibbs<br />

free energy G(T,p,N). It is easy to show th<strong>at</strong><br />

( 2 ∂ G<br />

∂N 2<br />

)<br />

≥ 0 (2.91)<br />

using the same thought experiment as before. But these experiments do not<br />

work for intensive variables. It is not possible, for example, to exchange an<br />

amount of temper<strong>at</strong>ure. On the other hand, one can calcul<strong>at</strong>e the second order<br />

deriv<strong>at</strong>ives in this case directly:<br />

( ) 2 ∂ G<br />

∂T 2<br />

( ∂ 2 G<br />

∂p 2<br />

)<br />

p,N<br />

T,N<br />

p,T<br />

( )<br />

∂S<br />

= −<br />

∂T<br />

( )<br />

∂V<br />

=<br />

∂p<br />

p,N<br />

T,N<br />

= − Cp<br />

T<br />

≤ 0 (2.92)<br />

= −V κT ≤ 0 (2.93)<br />

and the signs are opposite from wh<strong>at</strong> we find for the extensive variables. On<br />

the other hand<br />

( ) ( )<br />

2 ∂ G ∂V<br />

=<br />

= αV (2.94)<br />

∂p∂T N ∂T p,N<br />

and this leads to<br />

( 2 ∂ G<br />

∂T 2<br />

)<br />

p,N<br />

V κT Cp<br />

T<br />

( 2 ∂ G<br />

∂p2 ) (( ) )2<br />

2 ∂ G<br />

−<br />

=<br />

T,N ∂p∂T N<br />

− α 2 V 2 = V κT CV<br />

T<br />

≥ 0 (2.95)<br />

where we used 2.57. As a result we find th<strong>at</strong> G(p,T,N) is concave as a function<br />

of p and T, but convex as a function of N. Note th<strong>at</strong> these conclusions are<br />

independent of the minimum property of the Gibbs free energy. In th<strong>at</strong> case we<br />

have the st<strong>at</strong>ement th<strong>at</strong> <strong>at</strong> given values of p,T, and N the Gibbs free energy in<br />

equilibrium is a minimum as a function of the internal degrees of freedom!


2.8. STABILITY REQUIREMENTS ON OTHER FREE ENERGIES. 79<br />

Entropy rel<strong>at</strong>ions.<br />

We can also perform the thought experiment for a completely closed system,<br />

in which no he<strong>at</strong> is allowed to go to the external world. In th<strong>at</strong> case we specify<br />

U,V,N,X for the total system. The equilibrium st<strong>at</strong>e is found by maximizing the<br />

entropy. Because the n<strong>at</strong>ure of the extremum has now changed to a maximum,<br />

this implies th<strong>at</strong> S(U,V,N,X) is a concave function. As a consequence we have<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

≤ 0 (2.96)<br />

V,N,X<br />

( ∂ 2<br />

∂V 2<br />

)<br />

U,N,X<br />

≤ 0 (2.97)<br />

( 2 ∂ S<br />

∂N 2<br />

)<br />

≤ 0 (2.98)<br />

V,U,X<br />

( 2 ∂ S<br />

∂X 2<br />

)<br />

≤ 0 (2.99)<br />

U,V,N<br />

with opposite signs from wh<strong>at</strong> we found for the second order deriv<strong>at</strong>ives of the<br />

internal energy. But we still have the additional requirements of the form<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

V,N,X<br />

( 2 ∂ S<br />

∂V 2<br />

) ( ( ) ) 2<br />

2 ∂ S<br />

≥<br />

U,N,X ∂U∂V N,X<br />

(2.100)<br />

as before. Here the sign does not change because these requirements are rel<strong>at</strong>ed<br />

to a no-root condition.<br />

No inversion for second order deriv<strong>at</strong>ives.<br />

It is interesting to compare the deriv<strong>at</strong>ives containing S and U. We derived<br />

in general th<strong>at</strong><br />

( )<br />

∂S<br />

∂U V,N,X<br />

( )<br />

∂U<br />

= 1 (2.101)<br />

∂S V,N,X<br />

Such a simple rel<strong>at</strong>ion does not hold for the second order deriv<strong>at</strong>ives. From<br />

the inequalities we find<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

V,N,X<br />

( 2 ∂ U<br />

∂S2 )<br />

≤ 0 (2.102)<br />

V,N,X


80 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

2.9 A magnetic puzzle.<br />

Wh<strong>at</strong> is wrong?<br />

Consider a magnetic system, in which the basic extensive variables are the<br />

entropy S, volume V, amount of m<strong>at</strong>erial N, and total magnetic moment M.<br />

The magnetic work term is given by<br />

¯dWmag = −HdM (2.103)<br />

where we assume th<strong>at</strong> the direction of the applied field and the direction of the<br />

resulting moment are always the same. We define the magnetic susceptibilities<br />

by<br />

( )<br />

∂M<br />

χT =<br />

(2.104)<br />

∂H T<br />

and<br />

χS =<br />

( )<br />

∂M<br />

∂H S<br />

(2.105)<br />

In problem 7 we analyze this, and we find th<strong>at</strong> the normal stability requirements<br />

demand th<strong>at</strong><br />

χT ≥ χS ≥ 0 (2.106)<br />

for all m<strong>at</strong>erials, all values of the temper<strong>at</strong>ure, pressure, etc.<br />

This is very similar to the case of compressibilities, and in experiments compressibilities<br />

are always positive. Wh<strong>at</strong> could go wrong here? Well, there are<br />

many diamagnetic m<strong>at</strong>erials, for which the susceptibilities are neg<strong>at</strong>ive!<br />

Possible solutions.<br />

One possible answer is th<strong>at</strong> thermodynamics is wrong. Before we take this<br />

drastic step, however, it is advisable to consider other possibilities. Could the<br />

experiments be wrong? No, they are not. Is there some physics missing? Is<br />

there anything th<strong>at</strong> takes up energy? Yes! In order to apply a field we cre<strong>at</strong>e a<br />

field in the m<strong>at</strong>erial, and this magnetic field also needs energy! Therefore, the<br />

work term is incomplete. We need to include an extra term th<strong>at</strong> accounts for<br />

the energy stored in the field.<br />

Some basic E&M.<br />

In order to understand the effects of electric and magnetic fields on a sample,<br />

we divide space into two parts, separ<strong>at</strong>ed by a surface A. This surface is just


2.9. A MAGNETIC PUZZLE. 81<br />

outside the sample, so th<strong>at</strong> the enclosed volume is equal to the volume V of the<br />

sample. The only interactions through this surface are via electric and magnetic<br />

fields. Outside the sample we have applied electric and magnetic fields ⃗ E and<br />

⃗H, gener<strong>at</strong>ed by charges and currents. Outside the sample these fields are<br />

equivalent to the ⃗ D and ⃗ B fields, since there is no m<strong>at</strong>ter outside the sample.<br />

In real life one might have to correct for the gas in the sample chamber, but<br />

th<strong>at</strong> is not important here. The energy flow through a surface is given by the<br />

Poynting vector, ⃗ S, and the increase in internal energy of the sample is<br />

<br />

∆U = −∆t ⃗S · d 2 A ⃗ (2.107)<br />

where we have a minus sign because the surface element points away from the<br />

sample. Using the divergence theorem:<br />

∫<br />

∆U = −∆t ⃗∇ · ⃗ Sd 3 V (2.108)<br />

In MKSA units we have<br />

and hence<br />

∫<br />

∆U = −∆t<br />

V<br />

V<br />

A<br />

⃗S = ⃗ E × ⃗ H (2.109)<br />

⃗∇ · ⃗ E × ⃗ Hd 3 V (2.110)<br />

We now apply a product rule, of the form ⃗ ∇ · ⃗ A × ⃗ B = ⃗ B · ⃗ ∇ × ⃗ A − ⃗ A · ⃗ ∇ × ⃗ B<br />

to get<br />

∫ [ ]<br />

∆U = −∆t ⃗H · ∇ ⃗ × E⃗ − E ⃗ · ∇ ⃗ × H⃗<br />

d<br />

V<br />

3 V (2.111)<br />

At this point we can use Maxwell’s equ<strong>at</strong>ions to simplify m<strong>at</strong>ters. We have<br />

and<br />

⃗∇ × ⃗ H = ⃗j + ∂ ⃗ D<br />

∂t<br />

⃗∇ × ⃗ E = − ∂ ⃗ B<br />

∂t<br />

(2.112)<br />

(2.113)<br />

where the current density represents the free currents in the sample. Because<br />

we have magnetic fields th<strong>at</strong> can change, Eddy currents can be induced. We<br />

now have an expression for the energy change in terms of changing fields and<br />

currents:<br />

∫ [<br />

∆U = −∆t ⃗H · (−<br />

V<br />

∂ ⃗ B<br />

∂t ) − ⃗ E · (⃗j + ∂ ⃗ D<br />

∂t )<br />

]<br />

(2.114)


82 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

The middle term we recognize as the Joule he<strong>at</strong>ing taking place in the sample.<br />

This is he<strong>at</strong> th<strong>at</strong> is gener<strong>at</strong>ed inside the system, and hence does indeed<br />

add to the internal energy. We will get back to this term l<strong>at</strong>er in Chapter 6. In<br />

equilibrium thermodynamics we have no net currents, though, and so for now<br />

we ignore this term.<br />

To simplify m<strong>at</strong>ters we assume th<strong>at</strong> the fields are all uniform in the sample,<br />

so the volume integrals can be done. We also integr<strong>at</strong>e over a small time and<br />

find the changes in the fields. Th<strong>at</strong> leads to<br />

∆U = V<br />

[ ]<br />

⃗H · ∆B ⃗ + E ⃗ · ∆D⃗ (2.115)<br />

The B field and the H field are rel<strong>at</strong>ed via the magnetic moment density<br />

according to<br />

and similarly for the electric quantities<br />

⃗B = µ0( ⃗ H + 1<br />

V ⃗ M) (2.116)<br />

⃗D = ϵ0 ⃗ E + 1<br />

V ⃗ P (2.117)<br />

Note th<strong>at</strong> in Electricity and Magnetism the quantities in the previous equ<strong>at</strong>ions<br />

contain magnetiz<strong>at</strong>ion and polariz<strong>at</strong>ion density. In <strong>Thermodynamics</strong>, on<br />

the other hand, we use extensive quantities for these properties, and hence ⃗ M<br />

is the total magnetiz<strong>at</strong>ion and ⃗ P the total polariz<strong>at</strong>ion.<br />

Energy stored in the fields.<br />

This leaves us with four terms in the energy expression. Two depend on the<br />

fields only, they are:<br />

[<br />

∆Uf = V µ0 ⃗ H · ∆ ⃗ H + ϵ0 ⃗ E · ∆ ⃗ ]<br />

E<br />

(2.118)<br />

These two terms we recognize as the energy stored in the fields. If we change<br />

fields inside a sample it changes the m<strong>at</strong>erials properties, but we also need to<br />

cre<strong>at</strong>e the fields. Since the fields are not part of the sample, we could ignore<br />

them, but th<strong>at</strong> is part of the problem.<br />

Energy rel<strong>at</strong>ed to m<strong>at</strong>ter.<br />

The remaining two terms give the energy of the sample in a magnetic field. If<br />

we change the sample properties in a fixed field, we need more energy, according<br />

to<br />

[<br />

∆Um = µ0 ⃗ H · ∆ ⃗ M + ⃗ E · ∆ ⃗ ]<br />

P<br />

(2.119)


2.9. A MAGNETIC PUZZLE. 83<br />

Units are not always obvious in E&M.<br />

In thermodynamics we are used to seeing the formulas for magnetic work<br />

without the constants µ0. This is not hard to accomplish, as long as we remember<br />

to use the correct units. To obtain Gaussian units we make the following<br />

replacements:<br />

H →<br />

1<br />

√ H (2.120)<br />

4πµ0<br />

B → µ0<br />

√4π B (2.121)<br />

√<br />

4π<br />

M → √ M (2.122)<br />

µ0<br />

E →<br />

1<br />

√ E (2.123)<br />

4πϵ0<br />

D → ϵ0<br />

√4π D (2.124)<br />

P → √ 4πϵ0P (2.125)<br />

and in Gaussian units we have, assuming th<strong>at</strong> the polariz<strong>at</strong>ions are parallel to<br />

the fields,<br />

Interpret<strong>at</strong>ion.<br />

dU = V V<br />

HdH + EdE + HdM + EdP (2.126)<br />

4π 4π<br />

In the remainder of this section we focus on the magnetic part of the energy,<br />

and assume th<strong>at</strong> we apply external magnetic fields only. Th<strong>at</strong> is not essential<br />

for the discussion, though. The work term in 2.126 includes a term th<strong>at</strong> does<br />

not depend on the magnetiz<strong>at</strong>ion of the m<strong>at</strong>erial. If we do an experiment to<br />

measure the magnetic susceptibility and leave out the sample, we would still<br />

measure an effect, since we are storing energy in the magnetic field in the region<br />

where the sample would have been. The first term in 2.126 is really not a term<br />

connected with a sample, and should be subtracted. This is, however, never<br />

done in practice, and one always measure the total response of sample and field<br />

together, and defines the susceptibility from th<strong>at</strong> total response. If we would<br />

subtract the energy stored in the field, susceptibilities would be positive indeed.<br />

Experimentally measured susceptibilities.<br />

A simple rel<strong>at</strong>ion between the B and H fields is given by B = µH, which<br />

combines with H = B − 4π<br />

V M (Gaussian units!) gives


84 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

(µ − 1)H = 4π<br />

M (2.127)<br />

V<br />

from which we derive for the work done, which is the opposite from the change<br />

in energy,<br />

¯dW = − µ<br />

HdM (2.128)<br />

µ − 1<br />

or by defining M ′ = µ<br />

µ−1M we have in this new variable<br />

¯dW = −HdM ′<br />

(2.129)<br />

exactly as we had before we introduced the energy in the field. In terms of this<br />

new variable M ′ we have therefore:<br />

or<br />

or<br />

But we also have using 2.127<br />

which leads to<br />

or<br />

χ =<br />

( 2 ∂ U<br />

∂M ′2<br />

)<br />

≥ 0 (2.130)<br />

(<br />

∂H<br />

∂M ′<br />

)<br />

≥ 0 (2.131)<br />

µ − 1<br />

µ<br />

1<br />

≥ 0 (2.132)<br />

χ<br />

( )<br />

∂M µ − 1<br />

= V<br />

∂H 4π<br />

4π<br />

µV<br />

(2.133)<br />

≥ 0 (2.134)<br />

µ ≥ 0 (2.135)<br />

In other words, the permeability is a positive quantity. The consequence for the<br />

susceptibility is then from 2.133 th<strong>at</strong><br />

χ ≥ − V<br />

4π<br />

which is indeed observed experimentally.<br />

(2.136)


2.10. ROLE OF FLUCTUATIONS. 85<br />

2.10 Role of fluctu<strong>at</strong>ions.<br />

In this section we take a first look <strong>at</strong> the role of fluctu<strong>at</strong>ions in m<strong>at</strong>erials. A<br />

n<strong>at</strong>ural explan<strong>at</strong>ion of phenomena rel<strong>at</strong>ed to fluctu<strong>at</strong>ions is given in st<strong>at</strong>istical<br />

mechanics, where we consider the <strong>at</strong>omic n<strong>at</strong>ure of m<strong>at</strong>erials on a microscopic<br />

scale. In this section we look <strong>at</strong> fluctu<strong>at</strong>ions on a larger scale, and derive some<br />

conclusions which should always be valid, independent of the <strong>at</strong>omic n<strong>at</strong>ure of<br />

the m<strong>at</strong>erial. This again gives us some basic rel<strong>at</strong>ions which a true microscopic<br />

theory should obey. This is another example of the limits set by thermodynamics<br />

on microscopic theories.<br />

It is important to realize th<strong>at</strong> the way we will be describing fluctu<strong>at</strong>ions<br />

here is on a macroscopic level. We assume th<strong>at</strong> the laws of thermodynamics<br />

are obeyed, and th<strong>at</strong> st<strong>at</strong>e variables can be defined. In st<strong>at</strong>istical mechanics we<br />

will look <strong>at</strong> microscopic st<strong>at</strong>es, and see how the average of those microscopic<br />

st<strong>at</strong>es lead to the thermodynamic averages. Because of the fact th<strong>at</strong> in this section<br />

we consider thermodynamic fluctu<strong>at</strong>ions, we will always average over many<br />

microscopic st<strong>at</strong>es. In st<strong>at</strong>istical mechanics we will discuss how to perform such<br />

averages, which need to be consistent with the choice of independent variables<br />

for our thermodynamic system.<br />

A small subsystem.<br />

Consider a system with volume VL, amount of m<strong>at</strong>erial NL (in moles!), and<br />

energy UL. These quantities are fixed and they completely define the system.<br />

In other words, the temper<strong>at</strong>ure TL, the pressure pL, and the chemical potential<br />

µL of the system are completely defined in equilibrium. Also, the entropy SL<br />

of the system in equilibrium is determined.<br />

Consider a small part of this system. In order to define a small part, we<br />

need <strong>at</strong> least on extensive variable. The n<strong>at</strong>ural variable is to use the volume,<br />

although one could also use the amount of m<strong>at</strong>erial. Hence we consider a small<br />

volume V inside the system. We know th<strong>at</strong> on average we have for the energy<br />

U and the amount of m<strong>at</strong>erial N in this subsystem:<br />

and<br />

V<br />

U = UL<br />

VL<br />

V<br />

N = NL<br />

VL<br />

(2.137)<br />

(2.138)<br />

The energy and the amount of m<strong>at</strong>erial in the subsystem can fluctu<strong>at</strong>e by<br />

amounts ∆U and ∆N, however, and as derived before we have for the entropy<br />

of the subsystem:<br />

S(U + ∆U, V, N + ∆N) = S(U, V, N) + 1<br />

TL<br />

∆U − µL<br />

∆N+<br />

TL


86 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

1<br />

2<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

(∆U)<br />

V,N<br />

2 ( ) 2 ∂ S<br />

+<br />

∆U∆N +<br />

∂U∂N V<br />

1<br />

( 2 ∂ S<br />

2 ∂N 2<br />

)<br />

(∆N)<br />

U,V<br />

2 + · · ·<br />

(2.139)<br />

which is exactly the rel<strong>at</strong>ion we used to determine the equilibrium conditions<br />

for a stable m<strong>at</strong>erial. Note th<strong>at</strong> we assume th<strong>at</strong> the m<strong>at</strong>erial inside V is always<br />

in thermal equilibrium with itself, but not necessarily with the environment.<br />

Th<strong>at</strong> is why we can use the entropy formula. A discussion of more complic<strong>at</strong>ed<br />

fluctu<strong>at</strong>ions is beyond whet we can do here, because we need a definition of<br />

thermal equilibrium. One can make more progress in st<strong>at</strong>istical mechanics,<br />

since the microscopic details will play a role. But even there one has to be<br />

careful, and not go against basic theoretical assumptions.<br />

Note th<strong>at</strong> in our discussion we start with extensive variables, since these<br />

quantities can be measured directly. This is the logical first step. Quantities like<br />

pressure and temper<strong>at</strong>ure will also fluctu<strong>at</strong>e, but in order to find the fluctu<strong>at</strong>ions<br />

in intensive quantities we will need to invoke thermal equilibrium.<br />

The discussion we had before used fluctu<strong>at</strong>ions to derive criteria for stability,<br />

by requiring th<strong>at</strong> the entropy in equilibrium is maximal. This discussion does<br />

not answer the question how likely these fluctu<strong>at</strong>ions are to happen. In other<br />

words, wh<strong>at</strong> is the probability W (∆U, ∆N) of such a fluctu<strong>at</strong>ion? Which terms<br />

in the entropy should play a role?<br />

Subtracting the reference values.<br />

The first term in 2.139 simply sets the value of the entropy as expected from<br />

equilibrium, when the total system is completely homogeneous. This is just a<br />

reference value. The linear terms determine the conditions for equilibrium, and<br />

show th<strong>at</strong> T = TL and µ = µL are needed. Th<strong>at</strong> is why we use TL and µL in<br />

equ<strong>at</strong>ion 2.139 and not the actual values, since the difference has to show up in<br />

the second order terms. In equilibrium both the temper<strong>at</strong>ure and the chemical<br />

potential are uniform. Therefore we consider the function:<br />

R(U, V, N, ∆U, ∆N) = S(U + ∆U, V, N + ∆N) − S(U, V, N)−<br />

( ) ( )<br />

∂S<br />

∂S<br />

∆U −<br />

∆N (2.140)<br />

∂U V,N ∂N U,V<br />

where by definition we do not subtract terms with fluctu<strong>at</strong>ions in the volume,<br />

since the volume is considered to be a constant. Another motiv<strong>at</strong>ion for subtracting<br />

the linear terms is th<strong>at</strong> these terms are equal to the change in entropy<br />

of the remainder of the system. Since the remainder is extremely large, th<strong>at</strong><br />

change in entropy only contains the first order terms. Hence the function R<br />

really represents the change in entropy for the whole system! We have to add<br />

the changes in entropy of the rest of the system, but since the changes in U and<br />

N are opposite, we have to subtract the two terms in the equ<strong>at</strong>ion above.


2.10. ROLE OF FLUCTUATIONS. 87<br />

This function R contains all higher order terms in the fluctu<strong>at</strong>ions. Also,<br />

this function is neg<strong>at</strong>ive, since in equilibrium the entropy of the complete system<br />

is a maximum. Therefore we have th<strong>at</strong> 0 ≤ e R ≤ 1 and this could perhaps be<br />

used for a probability. Another reason is th<strong>at</strong> the law of large numbers tells<br />

us th<strong>at</strong> the probability as a function of the fluctu<strong>at</strong>ions should be a Gaussian<br />

distribution for small fluctu<strong>at</strong>ions, and e R has th<strong>at</strong> form. It also incorpor<strong>at</strong>es<br />

the fact th<strong>at</strong> large fluctu<strong>at</strong>ions are less likely, in an exponential manner.<br />

We have introduced a new element here. Based on wh<strong>at</strong> we know from<br />

st<strong>at</strong>istical mechanics we can say th<strong>at</strong> probabilities are rel<strong>at</strong>ed to the exponent<br />

of the entropy. We note here th<strong>at</strong> a function of the n<strong>at</strong>ure e R shows the right<br />

behavior, so we will simply postul<strong>at</strong>e th<strong>at</strong> it indeed represents the probabilities<br />

we need. It is also the n<strong>at</strong>ural extension in a st<strong>at</strong>istical mechanical theory.<br />

Finally, the consequences are supported by experimental evidence.<br />

Two new variables.<br />

and<br />

In order to make the m<strong>at</strong>h easier, we introduce two new variables:<br />

XU = 1<br />

TL<br />

XN = − µL<br />

TL<br />

(2.141)<br />

(2.142)<br />

If we assume th<strong>at</strong> the m<strong>at</strong>erial in the volume V is in equilibrium with the total<br />

sample when there are no fluctu<strong>at</strong>ions, we can then rewrite the function R in<br />

the form:<br />

R(U, V, N, ∆U, ∆N) = S(U + ∆U, V, N + ∆N) − S(U, V, N) − XU∆U − XN ∆N<br />

(2.143)<br />

The entropy is the extensive st<strong>at</strong>e variable describing the internal structure<br />

of a m<strong>at</strong>erial. It is therefore reasonable to use the entropy as the st<strong>at</strong>e function<br />

th<strong>at</strong> governs the probability of the fluctu<strong>at</strong>ions. Using the fact th<strong>at</strong> for small<br />

fluctu<strong>at</strong>ions in the thermodynamic limit (V large, but always small compared<br />

to the total volume VL) the probabilities need to be Gaussian we define for the<br />

probability W of a fluctu<strong>at</strong>ion in U and V:<br />

W (∆U, ∆N) = Ae ωR(U,V,N,∆U,∆N)<br />

(2.144)<br />

This definition contains two parameters, but the value of A is easily fixed by<br />

normaliz<strong>at</strong>ion:<br />

∫<br />

Ae ωR(U,V,N,∆U,∆N) d∆Ud∆N = 1 (2.145)<br />

The remaining question is how to find the value of ω. This requires microscopic<br />

theory or experiment. It turns out th<strong>at</strong> ωkB = 1, where kB is the


88 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

Boltzmann constant, rel<strong>at</strong>ed to the molar gas constant R by R = NAkB, with<br />

NA Avogadro’s number. This result easily explained in st<strong>at</strong>istical mechanics.<br />

This observ<strong>at</strong>ion also allows us to connect macroscopic experiments with microscopic<br />

theory, since we can measure probability distributions of macroscopic<br />

fluctu<strong>at</strong>ions!<br />

In equ<strong>at</strong>ion 2.144, wh<strong>at</strong> is really new and unexpected? The basic form can be<br />

derived from the law of large numbers, which st<strong>at</strong>es th<strong>at</strong> in the thermodynamic<br />

limit the probability distribution has to be Gaussian. This requires th<strong>at</strong> the<br />

leading term in the exponent should be of the form<br />

−A (∆U) 2 − B∆U∆N − C (∆N) 2<br />

(2.146)<br />

and th<strong>at</strong> A > 0, C > 0 , and B 2 ≤ 4AC. It does not require any further rel<strong>at</strong>ion<br />

between the values of A, B, and C, though. It also does not say anything<br />

about the higher order terms. Equ<strong>at</strong>ion 2.144 limits how these coefficients<br />

are rel<strong>at</strong>ed. This has consequences for experimental observ<strong>at</strong>ions, and rel<strong>at</strong>ion<br />

based on 2.144 have been confirmed both theoretically (st<strong>at</strong>istical mechanics)<br />

and experimentally.<br />

Averages.<br />

The use of equ<strong>at</strong>ion 2.144 follows from the calcul<strong>at</strong>ion of average values.<br />

Suppose we have a st<strong>at</strong>e function f th<strong>at</strong> depends on the amount of m<strong>at</strong>erial and<br />

the energy in the volume V. We assume th<strong>at</strong> the fluctu<strong>at</strong>ions are slow, and th<strong>at</strong><br />

the m<strong>at</strong>erial in V is always in equilibrium internally (but not with the outside<br />

world, of course). We can then define an average of this function by<br />

∫<br />

⟨f(∆U, ∆N)⟩ = Ae ωR(U,V,N,∆U,∆N) f(∆U, ∆N)d∆Ud∆N (2.147)<br />

Obviously we need to have ⟨∆U⟩ = 0 and ⟨∆N⟩ = 0.<br />

Ignore higher order terms.<br />

Why is this obvious?<br />

The definition 2.144 includes third order terms. Because of these terms, the<br />

probability distribution is not symmetric around the maximum value, and as a<br />

result a calcul<strong>at</strong>ion of ⟨∆U⟩ using 2.144 might not produce zero. This seems<br />

to be a problem, until we realize th<strong>at</strong> we have to take the thermodynamic<br />

limit (this thing is hidden everywhere!) and as a consequence the effects of the<br />

higher order terms become much smaller than those of the second order terms.<br />

Therefore we approxim<strong>at</strong>e the probability by


2.10. ROLE OF FLUCTUATIONS. 89<br />

W (∆U, ∆N) = Ae ω<br />

( (<br />

1 ∂<br />

2<br />

S<br />

2 ∂U 2<br />

)<br />

V,N (∆U)2 ( ) (<br />

∂<br />

2<br />

S<br />

1 ∂<br />

2<br />

S<br />

+ ∂U∂N ∆U∆N+<br />

V<br />

2 ∂N 2<br />

)<br />

U,V (∆N)2<br />

)<br />

(2.148)<br />

Using this form in the thermodynamic limit will give the same results as using<br />

the original definition.<br />

Consequences.<br />

As an example we study the case where only energy can fluctu<strong>at</strong>e, and hence<br />

the amount of m<strong>at</strong>erial in V is constant, or ∆N = 0. Because ( )<br />

∂S<br />

1<br />

∂U = V,N T we<br />

have<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

V,N<br />

= − 1<br />

T 2<br />

( )<br />

∂T<br />

∂U V,N<br />

= − 1<br />

T 2CV (2.149)<br />

and the probability for a fluctu<strong>at</strong>ion in the energy is<br />

ω(∆U)2<br />

−<br />

2T W (∆U) = Ae 2CV (2.150)<br />

The calcul<strong>at</strong>ion of the average mean square fluctu<strong>at</strong>ion is easy, and we find<br />

⟨(∆U) 2 2 T<br />

⟩ =<br />

ω CV<br />

(2.151)<br />

This formula has some interesting consequences. Since the right hand side<br />

is extensive, the left hand side is proportional to N, and we have<br />

√<br />

⟨(∆U) 2 ⟩<br />

U<br />

∝ 1<br />

√ N<br />

(2.152)<br />

In other words, in the thermodynamic limit the rel<strong>at</strong>ive magnitude of the fluctu<strong>at</strong>ions<br />

goes to zero!<br />

A second interesting observ<strong>at</strong>ion is th<strong>at</strong> fluctu<strong>at</strong>ions are rel<strong>at</strong>ed to the corresponding<br />

response function. M<strong>at</strong>erials with a large value of the he<strong>at</strong> capacity<br />

show large fluctu<strong>at</strong>ions in the energy. In other words, if we have a m<strong>at</strong>erial with<br />

a large he<strong>at</strong> capacity a small fluctu<strong>at</strong>ion in the temper<strong>at</strong>ure will give a large<br />

fluctu<strong>at</strong>ion in the energy. Such m<strong>at</strong>erials apparently respond more dram<strong>at</strong>ically<br />

to temper<strong>at</strong>ure fluctu<strong>at</strong>ions.<br />

Both previous observ<strong>at</strong>ions hold in general, and are useful to remember.<br />

Fluctu<strong>at</strong>ions are correl<strong>at</strong>ed.<br />

Wh<strong>at</strong> is the value of ⟨∆U∆N⟩? Naively one might say zero, but th<strong>at</strong> is not<br />

true since we assume th<strong>at</strong> the m<strong>at</strong>erial in V is always in equilibrium. A change<br />

in U will change the chemical potential, and the effect of changes in N are now


90 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

dependent on how U is changing. This is rel<strong>at</strong>ed to the presence of the mixed<br />

term in the exponent in the definition of the probability. The form is<br />

( ) 2 ∂ S<br />

= −<br />

∂U∂N<br />

1<br />

T 2<br />

( )<br />

∂T<br />

(2.153)<br />

∂N<br />

V<br />

which indeed rel<strong>at</strong>es changed in N to changes in T and hence changes in U.<br />

A calcul<strong>at</strong>ional trick.<br />

We use 2.144 in connection with 2.143, and consider XU and XN to be<br />

independent variables, which in equilibrium will take the appropri<strong>at</strong>e values<br />

rel<strong>at</strong>ed to the big system. This leads to<br />

( )<br />

∂W<br />

= −ω∆UW (2.154)<br />

∂XU<br />

and<br />

This leads to<br />

U,V<br />

( )<br />

∂W<br />

= −ω∆NW (2.155)<br />

∂XN<br />

⟨∆U∆N⟩ = − 1<br />

ω ⟨∆U<br />

( )<br />

∂W 1<br />

⟩ (2.156)<br />

∂XN W<br />

Using the definition of an average we find<br />

⟨∆U∆N⟩ = − 1<br />

∫ (<br />

∂W<br />

∆U<br />

ω ∂XN<br />

)<br />

d∆Ud∆N (2.157)<br />

where the 1<br />

W cancels the probability distribution W needed in the integral.<br />

Using integr<strong>at</strong>ion by parts<br />

⟨∆U∆N⟩ = − 1<br />

ω<br />

∂<br />

∂XN<br />

∫<br />

∆UW d∆Ud∆N + 1<br />

ω<br />

∫ ( ∂∆U<br />

∂XN<br />

)<br />

W d∆Ud∆N<br />

(2.158)<br />

The first term is zero because ∫ ∆UW d∆Ud∆N = 0. The second term<br />

contains<br />

( ) ( )<br />

∂∆U ∂(Uinstantaneous − U)<br />

=<br />

∂XN<br />

∂XN<br />

( )<br />

∂U<br />

= −<br />

∂XN<br />

(2.159)<br />

since the instantaneous energy is an independent variable. This expression is<br />

also independent of the fluctu<strong>at</strong>ion, and we obtain<br />

⟨∆U∆N⟩ = − 1<br />

( )<br />

∂U<br />

(2.160)<br />

ω ∂XN<br />

V,XU


2.10. ROLE OF FLUCTUATIONS. 91<br />

In a similar manner, elimin<strong>at</strong>ing ∆U, we could have obtained<br />

⟨∆U∆N⟩ = − 1<br />

( )<br />

∂N<br />

ω ∂XU<br />

V,XN<br />

(2.161)<br />

which shows a type of rel<strong>at</strong>ion between partial deriv<strong>at</strong>ives which is similar to<br />

the Maxwell rel<strong>at</strong>ions discussed in this chapter.<br />

We now use the last equ<strong>at</strong>ion and use th<strong>at</strong> XU is in equilibrium determined<br />

by TL via XU = 1 and obtain:<br />

TL<br />

( )<br />

2 T ∂N<br />

⟨∆U∆N⟩ = (2.162)<br />

ω ∂T V,XN<br />

where the partial deriv<strong>at</strong>ive is calcul<strong>at</strong>ed <strong>at</strong> equilibrium values corresponding<br />

to V, T = TL, µ = µL. The only point of concern is th<strong>at</strong> the partial deriv<strong>at</strong>ive<br />

is calcul<strong>at</strong>ed <strong>at</strong> constant XN, or <strong>at</strong> constant µ<br />

T . Using th<strong>at</strong> in this case<br />

N(T, V, µ) = N(T, V, XN T ) we derive<br />

( ) ( ) ( )<br />

∂N<br />

∂N<br />

∂N<br />

= + XN<br />

(2.163)<br />

∂T<br />

∂T V,µ ∂µ T,V<br />

V,XN<br />

and we have for our final result<br />

⟨∆U∆N⟩ = T<br />

[<br />

T<br />

ω<br />

( ) ( ) ]<br />

∂N<br />

∂N<br />

+ µ<br />

∂T V,µ ∂µ T,V<br />

(2.164)<br />

which indic<strong>at</strong>es th<strong>at</strong> the fluctu<strong>at</strong>ions in U and N are correl<strong>at</strong>ed, unless the<br />

partial deriv<strong>at</strong>ives cancel in a special way.<br />

If we would have used the first formula for this correl<strong>at</strong>ion, we would have<br />

found<br />

⟨∆U∆N⟩ = − 1<br />

(<br />

∂U<br />

ω ∂(− µ<br />

T )<br />

)<br />

V, 1<br />

T<br />

= T<br />

ω<br />

( )<br />

∂U<br />

∂µ V,T<br />

(2.165)<br />

and using the energy rel<strong>at</strong>ions dU = T dS + µdN, where there is no dV term<br />

since the volume is constant, we find<br />

⟨∆U∆N⟩ = T<br />

( ( ) ( ) )<br />

∂S<br />

∂N<br />

T + µ<br />

(2.166)<br />

ω ∂µ V,T ∂µ V,T<br />

and Mr. Maxwell tells us th<strong>at</strong> this is indeed the same as our first equ<strong>at</strong>ion for<br />

the correl<strong>at</strong>ion between fluctu<strong>at</strong>ions.<br />

Fluctu<strong>at</strong>ions in U again.<br />

In a similar way we derive th<strong>at</strong><br />

⟨(∆U) 2 ⟩ =<br />

T 2<br />

ω<br />

( )<br />

∂U<br />

∂T V,XN<br />

(2.167)


92 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

and this differs from the form we derived before, because in the partial deriv<strong>at</strong>ive<br />

XN is held constant. This difference can be explained very easily. In the current<br />

calcul<strong>at</strong>ion we allow both U and N to fluctu<strong>at</strong>e, and we know th<strong>at</strong> these are correl<strong>at</strong>ed.<br />

In the case we considered before we demanded th<strong>at</strong> ∆N = 0. The space<br />

over which the average value is calcul<strong>at</strong>ed is therefore different, and the result<br />

will be different! As usual, this shows how important it is in thermodynamic<br />

processes to specify the conditions under which we do an experiment!<br />

Fluctu<strong>at</strong>ions in other variables.<br />

In our approach to fluctu<strong>at</strong>ions we assume th<strong>at</strong> the m<strong>at</strong>erial inside the volume<br />

V is always in thermal equilibrium with itself, and th<strong>at</strong> the fluctu<strong>at</strong>ions are<br />

therefore very slow. In th<strong>at</strong> case it is easy to calcul<strong>at</strong>e the average root mean<br />

square fluctu<strong>at</strong>ions in other variables. For example, for small fluctu<strong>at</strong>ions we<br />

have<br />

( ) ( )<br />

∂T<br />

∂T<br />

∆T =<br />

∆U +<br />

∆N (2.168)<br />

∂U V,N ∂N U,V<br />

and this gives<br />

⟨(∆T ) 2 ⟩ =<br />

( ( ) ) 2<br />

∂T<br />

⟨(∆U)<br />

∂U V,N<br />

2 ⟩+<br />

( ) ( )<br />

( ( ) ) 2<br />

∂T ∂T<br />

∂T<br />

2<br />

⟨∆U∆N⟩ +<br />

⟨(∆N)<br />

∂U V,N ∂N U,N<br />

∂N U,N<br />

2 ⟩ (2.169)<br />

An interesting cross correl<strong>at</strong>ion.<br />

We now evalu<strong>at</strong>e the following correl<strong>at</strong>ion<br />

( )<br />

( )<br />

∂T<br />

∂T<br />

⟨∆T ∆N⟩ =<br />

⟨∆U∆N⟩ +<br />

∂U<br />

∂N<br />

V,N<br />

U,V<br />

⟨(∆N) 2 ⟩ (2.170)<br />

In order to evalu<strong>at</strong>e this expression we need ⟨(∆N) 2 ⟩. Similar to the deriv<strong>at</strong>ion<br />

before, we find<br />

⟨∆N∆N⟩ = − 1<br />

∫ ( )<br />

∂W<br />

∆N d∆Ud∆N (2.171)<br />

ω ∂XN<br />

and<br />

⟨∆N∆N⟩ = − 1<br />

ω<br />

∂<br />

∂XN<br />

∫<br />

∆NW d∆Ud∆N + 1<br />

ω<br />

∫ ( ∂∆N<br />

∂XN<br />

)<br />

W d∆Ud∆N<br />

(2.172)


2.10. ROLE OF FLUCTUATIONS. 93<br />

which leads to<br />

or<br />

Hence we find<br />

⟨∆T ∆N⟩ =<br />

( )<br />

∂T T<br />

∂U V,N ω<br />

⟨∆N∆N⟩ = − 1<br />

( )<br />

∂N<br />

ω ∂XN V,XU<br />

⟨∆N∆N⟩ = + T<br />

ω<br />

( )<br />

∂N<br />

∂µ V,T<br />

( ) ( )<br />

∂U ∂T<br />

+<br />

∂µ V,T ∂N U,V<br />

T<br />

ω<br />

( )<br />

∂N<br />

∂µ V,T<br />

(2.173)<br />

(2.174)<br />

(2.175)<br />

which can be calcul<strong>at</strong>ed using the following observ<strong>at</strong>ion. Suppose we have the<br />

temper<strong>at</strong>ure T as a function of N, V, U, and want to calcul<strong>at</strong>e the change in T<br />

as a function of µ while we keep the variables V and X constant, where X is to<br />

be specified. We have<br />

( ) ( )<br />

∂T ∂T<br />

=<br />

∂µ V,X ∂U V,N<br />

( ) ( )<br />

∂U ∂T<br />

+<br />

∂µ V,X ∂N V,U<br />

( )<br />

∂N<br />

∂µ V,X<br />

(2.176)<br />

which is exactly the form of the right hand side of the previous equ<strong>at</strong>ion, with<br />

X = ) T . But if we keep T constant, it does not change, and the partial deriv<strong>at</strong>ive<br />

is zero! Hence we find an important result<br />

( ∂T<br />

∂µ<br />

V,T<br />

⟨∆T ∆N⟩ = 0 (2.177)<br />

which means th<strong>at</strong> fluctu<strong>at</strong>ions in temper<strong>at</strong>ure and amount of m<strong>at</strong>erial are uncorrel<strong>at</strong>ed!<br />

It can be shown th<strong>at</strong> this is true in general for the correl<strong>at</strong>ion<br />

between fluctu<strong>at</strong>ions in an extensive st<strong>at</strong>e variable and an intensive st<strong>at</strong>e variable,<br />

as long as they are not a conjug<strong>at</strong>e pair. Hence also ⟨∆p∆N⟩ = 0. For<br />

correl<strong>at</strong>ions between the members of a conjug<strong>at</strong>e pair we find the following<br />

⟨∆µ∆N⟩ =<br />

( )<br />

∂µ T<br />

∂U V,N ω<br />

( ) ( )<br />

∂U ∂µ<br />

+<br />

∂µ V,T ∂N U,V<br />

and now the right hand side involves the partial deriv<strong>at</strong>ive<br />

hence<br />

T<br />

ω<br />

( )<br />

∂N<br />

∂µ V,T<br />

( ∂µ<br />

∂µ<br />

)<br />

V,T<br />

(2.178)<br />

= 1, and<br />

⟨∆µ∆N⟩ = T<br />

ω = kBT (2.179)<br />

and again one can show th<strong>at</strong> this is true for all correl<strong>at</strong>ions between members<br />

of conjug<strong>at</strong>e pairs. With the exception of the pressure and volume pair, since


94 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

we have defined our system to be <strong>at</strong> constant volume cells. If we would have<br />

worked with constant amount of m<strong>at</strong>erial cells, then we would have had this<br />

rel<strong>at</strong>ion to be valid for the pressure volume combin<strong>at</strong>ion, too.<br />

Non-equilibrium fluctu<strong>at</strong>ions.<br />

In the previous discussion we have always assumed th<strong>at</strong> the volume we are<br />

considering is in local equilibrium. Wh<strong>at</strong> if th<strong>at</strong> is not the case. We will first<br />

look <strong>at</strong> a simple example. Suppose the volume is divided into two parts, both in<br />

local equilibrium, but not in equilibrium with each other. For example, assume<br />

th<strong>at</strong> the pressure in the two parts is different, but the temper<strong>at</strong>ure is the same.<br />

We can define the entropy of this system, it is the sum of the entropies of the<br />

two parts. The entropy is lower than the entropy of the same system after<br />

it came into equilibrium, and hence the probability of such a non-equilibrium<br />

fluctu<strong>at</strong>ion is smaller (exponentially!) than the local equilibrium fluctu<strong>at</strong>ion. In<br />

this example we could define the temper<strong>at</strong>ure but not the pressure of the nonequilibrium<br />

fluctu<strong>at</strong>ion, but the temper<strong>at</strong>ure is not the same as the temper<strong>at</strong>ure<br />

of the equilibrium fluctu<strong>at</strong>ion!<br />

We can generalize this to the following case. Suppose we have a nonequilibrium<br />

fluctu<strong>at</strong>ion in which we can divide the volume into smaller cells,<br />

all in local equilibrium, but not in equilibrium with each other. The entropy<br />

of such a fluctu<strong>at</strong>ion can be measured, and is lower than the entropy of the<br />

corresponding equilibrium fluctu<strong>at</strong>ion. Hence the probability of such a nonequilibrium<br />

fluctu<strong>at</strong>ion is much smaller. Also, the values of intensive st<strong>at</strong>e<br />

variables are generally not defined for such a fluctu<strong>at</strong>ion.<br />

It is important to realize th<strong>at</strong> in order to measure the entropy we need thermal<br />

equilibrium and the measuring device. So there will be many fluctu<strong>at</strong>ions<br />

for which we cannot measure the entropy, and for which we hence cannot say<br />

anything quantit<strong>at</strong>ively about the probability of it happening. But if we can<br />

describe such a fluctu<strong>at</strong>ion, this means th<strong>at</strong> we can define some set of internal<br />

parameters characterizing this fluctu<strong>at</strong>ion, and we could include these in our<br />

thermodynamic description to some approxim<strong>at</strong>ion. Hence we can approxim<strong>at</strong>e<br />

the entropy and again we know th<strong>at</strong> it will be even lower than before. Hence<br />

the probability of such a non-equilibrium fluctu<strong>at</strong>ion will be even smaller!<br />

2.11 Extra questions.<br />

1. If the only st<strong>at</strong>e variables are T, S, p,and V,wh<strong>at</strong> is the meaning of U −<br />

T S + pV ?<br />

2. Why is there a difference in useful work done between a constant N and a<br />

constant /mu process?<br />

3. Explain the three cylinder scenarios.<br />

4. Give examples for the use of each free energy.


2.11. EXTRA QUESTIONS. 95<br />

5. The n<strong>at</strong>ural variables for U are S, V, and N. Why does th<strong>at</strong> make sense?<br />

6. So we have U = T S − pV + µN and also U = T S − pV + µN + MH. How<br />

can th<strong>at</strong> be?<br />

7. Wh<strong>at</strong> is the importance of Euler and Gibbs-Duhem?<br />

8. How many of the Maxwell rel<strong>at</strong>ions should you memorize?<br />

9. Should α always be positive?<br />

10. For a neg<strong>at</strong>ive thermal expansion m<strong>at</strong>erial, would Cp < CV ?<br />

11. Which additional response functions could one define in a magnetic system?<br />

How many he<strong>at</strong> capacities?<br />

12. Which Maxwell to use?<br />

13. Wh<strong>at</strong> are the signs of an ideal gas model failing <strong>at</strong> low T?<br />

14. Find a Maxwell rel<strong>at</strong>ion to change ( )<br />

∂S<br />

∂T V,N<br />

15. Show th<strong>at</strong> α = 1<br />

( ) ( )<br />

2<br />

∂ G<br />

1 ∂S<br />

V ∂p∂T = − V ∂p<br />

N N<br />

16. Minus sign in three deriv<strong>at</strong>ive rel<strong>at</strong>ion Why also in five, seven?<br />

17. Why is the r<strong>at</strong>io rel<strong>at</strong>ion for he<strong>at</strong> capacities, compressibilities important?<br />

18. Explain why a neg<strong>at</strong>ive compressibility leads to an instability<br />

19. f(x, y, z) = x2 + y2 + z2 − R2 . x = √ R2 − y2 − z2 ( )<br />

∂x y<br />

. ∂y = − x<br />

z<br />

chain rule OK?<br />

20. Why is there a minimum only in the extensive variables?<br />

21. Are there more rel<strong>at</strong>ions to make sure th<strong>at</strong> the entropy is a maximum?<br />

22. Wh<strong>at</strong> is the difference between H and T?<br />

Is the<br />

23. Which m<strong>at</strong>erials have χ = − V<br />

4π ?<br />

)<br />

24. Is always positive, no m<strong>at</strong>er which variables are held constant?<br />

( ∂µ<br />

∂N<br />

??<br />

25. Wh<strong>at</strong> kind of fluctu<strong>at</strong>ions can we describe?<br />

26. Why use the entropy to characterize fluctu<strong>at</strong>ions?<br />

27. Which R is in ωR = 1?<br />

28. Why do higher order terms drop out in the TD limit?<br />

29. Wh<strong>at</strong> are uses of the rel<strong>at</strong>ion between < (∆U) 2 > and CV ?<br />

30. If the correl<strong>at</strong>ion between fluctu<strong>at</strong>ions cancel, wh<strong>at</strong> is the condition? (<br />

µ ∝ T !!)


96 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

2.12 Problems for chapter 2<br />

Problem 1.<br />

A container with volume V contains N moles of an ideal gas. The container is<br />

divided into two parts by a movable partition. This partition can move without<br />

any friction. The volume of the left part is V1. The number of moles in the left<br />

part is N1. The partition always moves infinitesimally slow, and the ideal gas<br />

in each part is always in equilibrium. The temper<strong>at</strong>ure is T everywhere in the<br />

system. T is kept constant. M<strong>at</strong>erial cannot go through the partition, hence<br />

N1 remains constant. V1 can vary, though, and hence we consider V1 to be an<br />

internal parameter.<br />

(a) Which thermodynamic potential does one have to minimize as a function<br />

of V1 in order to find the equilibrium st<strong>at</strong>e of the total system?<br />

(b) Show th<strong>at</strong> the minimum occurs when the pressure left and right is the<br />

same. Note: do not only show th<strong>at</strong> it is an extremum, but show th<strong>at</strong> it is<br />

not a maximum!<br />

Problem 2.<br />

Show th<strong>at</strong> κT − κS =<br />

Problem 3.<br />

V T α2<br />

Cp<br />

(equ<strong>at</strong>ion 2.58 ).<br />

In first approxim<strong>at</strong>ion, the electrons in a conductor can be modelled by a<br />

gas of particles <strong>at</strong> a given temper<strong>at</strong>ure T in a volume V <strong>at</strong> a chemical potential<br />

µ. The chemical potential is set by connecting the conductor to a b<strong>at</strong>tery, the<br />

other terminal of the b<strong>at</strong>tery is grounded.<br />

(a) Which thermodynamic potential is needed to describe this system?<br />

The number of electrons in the conductor is measured as a function of the<br />

temper<strong>at</strong>ure. Experimentally one finds the rel<strong>at</strong>ion N(T, V, µ) = n0V (1 − RT<br />

µ )<br />

for a range of temper<strong>at</strong>ures with RT ≪ µ.<br />

(b) Using a Maxwell rel<strong>at</strong>ion, calcul<strong>at</strong>e how the entropy S depends on µ.<br />

(c) Show th<strong>at</strong> lim<br />

µ→0 S(T, V, µ) = ∞.<br />

The last result seems to indic<strong>at</strong>e th<strong>at</strong> lim<br />

T →0 S(T, V, 0) = ∞, in contradiction<br />

with the third law.


2.12. PROBLEMS FOR CHAPTER 2 97<br />

(d) Argue th<strong>at</strong> this experiment does not invalid<strong>at</strong>e the third law.<br />

Problem 4.<br />

Show th<strong>at</strong><br />

to show th<strong>at</strong><br />

Problem 5.<br />

( )<br />

∂µ<br />

∂N<br />

( ) S,V<br />

∂µ<br />

∂N<br />

T,V<br />

≤ CV<br />

T<br />

≥ 0.<br />

( ( ) ) 2<br />

∂T<br />

∂N . Use this expression for µ(S, V, N)<br />

S,V<br />

Show th<strong>at</strong> the energy function U(S, V ) = S2<br />

V s<strong>at</strong>isfies the criteria for stability.<br />

Calcul<strong>at</strong>e CV , Cp, κS, κT , and α for this system as a function of the temper<strong>at</strong>ure<br />

T.<br />

Problem 6.<br />

Calcul<strong>at</strong>e<br />

(<br />

2<br />

∂ S<br />

∂U 2<br />

)<br />

V,N,X<br />

and<br />

(<br />

2<br />

∂ U<br />

∂S2 )<br />

V,N,X<br />

in terms of response functions dis-<br />

cussed in this chapter. Use these results to show th<strong>at</strong> the product of these two<br />

deriv<strong>at</strong>ives is neg<strong>at</strong>ive.<br />

Problem 7.<br />

A magnetic system is characterized by the st<strong>at</strong>e variables T, S, M, H, µ, and<br />

N. N is always constant in this problem. All partial deriv<strong>at</strong>ives are <strong>at</strong> constant<br />

N. The magnetic susceptibilities are defined by χT = ( )<br />

∂M<br />

∂H T and χS = ( )<br />

∂M<br />

∂H S<br />

.<br />

(a) Show th<strong>at</strong> the he<strong>at</strong> capacities <strong>at</strong> constant field and <strong>at</strong> constant moment<br />

are rel<strong>at</strong>ed to the susceptibility χT by CH − CM = T<br />

(( ) )<br />

∂M 2.<br />

χT ∂T H<br />

(b) Show th<strong>at</strong> the susceptibilities <strong>at</strong> constant temper<strong>at</strong>ure and entropy obey<br />

χT − χS = T<br />

(( ) )<br />

∂M 2.<br />

CH ∂T H<br />

(c) Show th<strong>at</strong> χT ≥ χS ≥ 0 and th<strong>at</strong> CH ≥ CM ≥ 0.<br />

(d) Calcul<strong>at</strong>e CH, CM, χT , and χS for a paramagnetic gas containing N moles<br />

of m<strong>at</strong>erial with U = 3<br />

2<br />

Problem 8.<br />

NRT and M = CNH<br />

T .<br />

Show th<strong>at</strong> ⟨(∆M) 2 ⟩ = T<br />

ω χV when ∆U = ∆N = 0.


98 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

Problem 9.<br />

A closed system (no exchange of he<strong>at</strong> and work with the outside world)<br />

contains an ideal gas. The volume V of the total system is constant. An internal<br />

partition divides the system in two parts. There is no flow of m<strong>at</strong>erial through<br />

this partition. The volumes of the two parts are V1 and V − V1. The parameter<br />

V1 is an internal parameter. He<strong>at</strong> can flow back and forth through the partition.<br />

As a consequence, the temper<strong>at</strong>ure T is always the same left and right, and is<br />

equal to T (V1). The partition always moves in a quasi-st<strong>at</strong>ic manner and each<br />

part of the system is always in thermal equilibrium. The internal energy of<br />

the system is the sum of the internal energy of the parts plus the mechanical<br />

potential energy of the partition (the pressures in both parts are not the same!).<br />

The total entropy of the system is constant.<br />

(a) Use the entropy equ<strong>at</strong>ion for an ideal gas to find<br />

V1.<br />

( )<br />

∂T as a function of<br />

∂V1<br />

(b) Use this result to show th<strong>at</strong> in equilibrium the temper<strong>at</strong>ure of a system<br />

with given S, V, and N is a minimum.<br />

(c) The minimum energy principle in connection with U = 3<br />

2NRT seems to<br />

lead to the same conclusion. Wh<strong>at</strong> is wrong in th<strong>at</strong> argument?<br />

Problem 10.<br />

A gas has the equ<strong>at</strong>ion of st<strong>at</strong>e pV = NRT , but does not have to be ideal,<br />

hence U does not have to be pV . The volume expands by a small amount ∆V<br />

<strong>at</strong> constant temper<strong>at</strong>ure and constant number of particles.<br />

(A) Use a Maxwell rel<strong>at</strong>ion to calcul<strong>at</strong>e ( )<br />

∂S<br />

∂V T N .<br />

(B) Calcul<strong>at</strong>e the amount of work ∆W done during this process, in terms of<br />

st<strong>at</strong>e variables and ∆V .<br />

(C) Calcul<strong>at</strong>e the amount of he<strong>at</strong> ∆Q entering the system during this process,<br />

in terms of st<strong>at</strong>e variables and ∆V .<br />

(D) Show th<strong>at</strong> ∆U = 0 during this process.<br />

(E) If (D) is true, show th<strong>at</strong> U is a function of T and N only.<br />

(F) If (E) is true, show th<strong>at</strong> U has the form U = Nf(T ).<br />

Problem 11.


2.12. PROBLEMS FOR CHAPTER 2 99<br />

Two containers are connected by a valve. The first container, volume V,<br />

contains a gas <strong>at</strong> temper<strong>at</strong>ure T, pressure p. The second container is empty<br />

(vacuum). The total volume of both containers is V’. Both containers are completely<br />

isol<strong>at</strong>ed from the environment. The valve is opened, and the gas in the<br />

first container expands freely, but irreversibly, and finally fills both containers.<br />

In the end we have again thermodynamic equilibrium, with st<strong>at</strong>e variables V’,<br />

T’, and p’. Calcul<strong>at</strong>e T’ in the following two cases: (a) the gas is ideal (b) the<br />

2<br />

N gas has equ<strong>at</strong>ions of st<strong>at</strong>e (p + a V 2 )(V − Nb) = NRT , CV = 3<br />

2NR. Problem 12.<br />

Show th<strong>at</strong> the osmotic pressure p exerted by N moles of solute in a very<br />

dilute solution of temper<strong>at</strong>ure T and volume V is given by p = NRT<br />

V . Use<br />

entropy arguments.<br />

Problem 13.<br />

The function U is given by U(S, V, N) = 1<br />

N (S2 +S −V 2 ). Give three reasons<br />

why this is not a correct function describing the internal energy of a m<strong>at</strong>erial.<br />

Problem 14.<br />

The internal energy of system is given by U(S, V, N) = S α V β N γ .<br />

(a) Using the fact th<strong>at</strong> U is extensive, find a rel<strong>at</strong>ion between α, β, and γ.<br />

(b) Calcul<strong>at</strong>e the pressure p, temper<strong>at</strong>ure T, and chemical potential µ as a<br />

function of S, V, and N.<br />

(c) Calcul<strong>at</strong>e the he<strong>at</strong> capacity <strong>at</strong> constant volume V (and constant N).<br />

(d) Calcul<strong>at</strong>e the adiab<strong>at</strong>ic compressibility (<strong>at</strong> constant N).<br />

(e) Based on the sign of these response functions, find inequalities for α and<br />

β.<br />

Problem 15.<br />

Use a Maxwell rel<strong>at</strong>ion to rel<strong>at</strong>e the coefficient of thermal expansion to a<br />

partial deriv<strong>at</strong>ive of the entropy and use th<strong>at</strong> to show th<strong>at</strong> the coefficient of<br />

thermal expansion approaches zero when the temper<strong>at</strong>ure goes to zero.<br />

Problem 16.


100 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

A system is characterized by four st<strong>at</strong>e variables, S, T, p, and V. The internal<br />

energy is given by<br />

with c being a constant.<br />

(A) Calcul<strong>at</strong>e ( )<br />

∂S<br />

∂T V .<br />

(B) Calcul<strong>at</strong>e S(T, V ).<br />

(C) Calcul<strong>at</strong>e p(T, V ).<br />

Problem 17.<br />

U(T, V ) = cV T 2<br />

The equ<strong>at</strong>ions of st<strong>at</strong>e for an ideal gas are pV = NRT and U = 3<br />

2NRT .<br />

(A) Show th<strong>at</strong> CV = 3<br />

2 NR.<br />

(B) Calcul<strong>at</strong>e the entropy difference S(T, V, N) − S(T0, V, N).<br />

(C) Wh<strong>at</strong> happens in the limit T → 0? Which conclusion can you draw from<br />

this?<br />

(D) Calcul<strong>at</strong>e the enthalpy H.<br />

(E) Use the enthalpy to show th<strong>at</strong> Cp = 5<br />

2 NR.<br />

Problem 18.<br />

The chemical potential of an ideal gas with density n = N<br />

V<br />

µ = RT ln( n<br />

nG<br />

)<br />

is given by<br />

with nG = cT 3<br />

2 , where c is a constant with the appropri<strong>at</strong>e dimensions. This<br />

gas is put in a uniform gravit<strong>at</strong>ional filed, acceler<strong>at</strong>ion g. Calcul<strong>at</strong>e the pressure<br />

in this isothermal gas as a function of height.<br />

Problem 19.<br />

A steam engine works changing w<strong>at</strong>er to steam, and using the steam’s pressure<br />

to do mechanical work. Consider a steam engine pulling a train of cars (no<br />

friction), holding w<strong>at</strong>er and fuel. At the beginning the mass of the empty train<br />

is M, the mass of the w<strong>at</strong>er is Mw<strong>at</strong>er, and the mass of the fuel is Mfuel. A<br />

unit mass of fuel releases J Joules upon burning. This energy evapor<strong>at</strong>es the<br />

w<strong>at</strong>er and brings the steam to 200C. The he<strong>at</strong> per mole needed to do this is H.<br />

Mechanical energy is gained from the steam by adiab<strong>at</strong>ically cooling it to 100C.<br />

The adiab<strong>at</strong>ic index γ of w<strong>at</strong>er is 4<br />

3 .


2.12. PROBLEMS FOR CHAPTER 2 101<br />

(a) Calcul<strong>at</strong>e the fraction f of the fuel’s energy th<strong>at</strong> is available to do mechanical<br />

work.<br />

(b) The train moves up a hill on a 10km stretch with a 1% slope. Calcul<strong>at</strong>e<br />

the amount of fuel used in this trip, assuming it is much less than the<br />

total mass of the train.<br />

(c) Calcul<strong>at</strong>e the amount of w<strong>at</strong>er used during this trip, assuming th<strong>at</strong> this,<br />

too, is much less than the mass of the train.<br />

Problem 20.<br />

Each member of a family of 4 opens the refriger<strong>at</strong>or 8 times a day. Every<br />

time it is opened, 75% of the 1 m 3 of air in the refriger<strong>at</strong>or is replaced by air<br />

from the room. In addition to the air, the refriger<strong>at</strong>or contains different things,<br />

equivalent to 200 liters of w<strong>at</strong>er. The temper<strong>at</strong>ure inside the refriger<strong>at</strong>or is 2C,<br />

the room is 20C. No he<strong>at</strong> leaks out when the door is closed. The refriger<strong>at</strong>or is<br />

cooled by an ideal Carnot machine. The vapor pressure of w<strong>at</strong>er <strong>at</strong> 20C is 0.1<br />

<strong>at</strong>m, <strong>at</strong> 2C it can be ignored. The he<strong>at</strong> of vaporiz<strong>at</strong>ion of w<strong>at</strong>er is 2.26 MJ/kg.<br />

R = 8.31 J/Kmol. A mole of gas occupies 22.4 liters <strong>at</strong> 0C.<br />

(a) Wh<strong>at</strong> are the molar he<strong>at</strong> capacities cV and cp for air?<br />

(b) Wh<strong>at</strong> is the average power needed to oper<strong>at</strong>e the refriger<strong>at</strong>or when the<br />

air in the house is dry?<br />

(c) Wh<strong>at</strong> is the average power needed to oper<strong>at</strong>e the refriger<strong>at</strong>or when the<br />

air in the house is s<strong>at</strong>ur<strong>at</strong>ed with w<strong>at</strong>er vapor?<br />

Problem 21.<br />

The he<strong>at</strong> capacity of a solid is C = AT 3 , where A is a constant. This solid is<br />

the low temper<strong>at</strong>ure reservoir of a reversible refriger<strong>at</strong>or. The high temper<strong>at</strong>ure<br />

reservoir is <strong>at</strong> room temper<strong>at</strong>ure. The solid is cooled from room temper<strong>at</strong>ure<br />

to almost absolute zero.<br />

(a) Find an expression for the amount of work required to cool this solid.<br />

(b) Wh<strong>at</strong> is the decrease in entropy of the solid?<br />

(c) Wh<strong>at</strong> is the decrease in internal energy of the solid?<br />

(d) Wh<strong>at</strong> are the increases in entropy and internal energy of the high temper<strong>at</strong>ure<br />

reservoir?


102 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

Problem 22.<br />

The equ<strong>at</strong>ion of st<strong>at</strong>e of an imperfect gas is given by p = N<br />

V RT ( 1 + B(T ) N<br />

)<br />

V<br />

where B(T ) is some function of the temper<strong>at</strong>ure. This gas undergoes a small<br />

isothermal expansion ∆V <strong>at</strong> constant N. The amount of he<strong>at</strong> going into the<br />

system is ∆Q.<br />

(a) Rel<strong>at</strong>e ∆Q to ∆V , keeping only linear terms.<br />

(b) Evalu<strong>at</strong>e the partial deriv<strong>at</strong>ives in (a).<br />

(c) Interpret the result.<br />

Problem 23.<br />

A well insul<strong>at</strong>ed steel tank has volume V . The tank contains air <strong>at</strong> pressure<br />

p1 and temper<strong>at</strong>ure T1. It is <strong>at</strong>tached to a compressed air line which supplies<br />

air <strong>at</strong> steady pressure p ′ and temper<strong>at</strong>ure T ′ . Initially the valve between the<br />

tank and the line is closed. The valve is then opened long enough to change the<br />

pressure in the tank to p2, and it is then closed again.<br />

Show th<strong>at</strong> in transport of this kind the energy balance is given by N2(u2 −<br />

h ′ ) − N1(u1 − h ′ ) = Q, where u is the molar internal energy, h’ is the molar<br />

enthalpy h’=u’+p’v’ where v’ is the molar volume occupied by air in the line,<br />

and Q is the he<strong>at</strong> added to the air.<br />

How many moles N2 are in the tank <strong>at</strong> th<strong>at</strong> moment, and wh<strong>at</strong> is the temper<strong>at</strong>ure<br />

T2?<br />

Assume th<strong>at</strong> (1) the tank and the air are always <strong>at</strong> the same temper<strong>at</strong>ure,<br />

(2) Air is a perfect gas with molar specific he<strong>at</strong>s cp = 7<br />

2 R and cV = 5<br />

For numbers use V1 = 20m 3 , p1 = 1bar , T1 = 22C, p ′ = 5bar, T ′ = 30C,<br />

p2 = 3bar. The mass of the steel tank is mt 1000 kg, the he<strong>at</strong> capacity of steel<br />

cs is 100 cal/kg/K and the gas constant is R=8.3 x 10 −5 m 3 bar/mol/K.<br />

Problem 24.<br />

A thermodynamic model for gravit<strong>at</strong>ional collapse can be constructed by<br />

adding the gravit<strong>at</strong>ional self energy to the ideal gas. A spherical mass M of a<br />

cloud of dust with radius R has gravit<strong>at</strong>ional energy Ug = − 3M 2 G<br />

5R .<br />

(a) Find the equ<strong>at</strong>ion of st<strong>at</strong>e of the cloud.<br />

(b) Find an expression for the isothermal compressibility.<br />

(c) Discuss the conditions under which the cloud is no longer thermally stable.<br />

Problem 25.<br />

2 R.


2.12. PROBLEMS FOR CHAPTER 2 103<br />

The he<strong>at</strong> capacity <strong>at</strong> constant pressure for a certain solid is γT with γ =<br />

5mcal/degree 2 /mol. At T=300K the solid changes into a classical ideal gas<br />

which is kept <strong>at</strong> constant pressure. The he<strong>at</strong> of this transition is 150 cal/mol.<br />

The entropy of this gas <strong>at</strong> T=600K is 5.5 cal/degree/mol. Is this gas mono<strong>at</strong>omic<br />

or di<strong>at</strong>omic? Use R=2cal/degree/mol.<br />

Problem 26.<br />

A substance has the following properties:<br />

(I) The amount of m<strong>at</strong>erial is fixed <strong>at</strong> N moles.<br />

(II) At constant temper<strong>at</strong>ure T0 the work W done by it on expansion from a<br />

volume V0 to a volume V is<br />

(III) The entropy is given by<br />

W = NRT0 ln( V<br />

S = NR V0<br />

V<br />

( T<br />

T0<br />

V0<br />

)<br />

) x<br />

In these expressions T0, V0, and x are fixed constants.<br />

(A) Calcul<strong>at</strong>e the Helmholtz free energy of the system.<br />

(B) Find the equ<strong>at</strong>ion of st<strong>at</strong>e for the pressure p in terms of V , T , and N.<br />

(C) Find the work done on expansion <strong>at</strong> constant T, where T is arbitrary.<br />

Problem 27.<br />

The he<strong>at</strong> capacity <strong>at</strong> constant pressure is given by<br />

( )<br />

∂S<br />

Cp = T<br />

∂T<br />

and the coefficient of thermal expansion is given by<br />

α = 1<br />

( )<br />

∂V<br />

V ∂T<br />

The amount of m<strong>at</strong>erial N is always kept constant.<br />

If we consider S(T, p) we have<br />

( ) ( )<br />

∂S ∂S<br />

dS = dT + dp<br />

∂T p ∂p T<br />

p<br />

p


104 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

(A) Show th<strong>at</strong> this leads to<br />

T dS = CpdT − αT V dp<br />

One can show th<strong>at</strong> (but you do NOT have to!)<br />

where X is a finite constant.<br />

αV<br />

lim<br />

T →0 Cp<br />

= X<br />

(B) Show th<strong>at</strong> this has the consequence th<strong>at</strong> in an adiab<strong>at</strong>ic process one can<br />

never reach T = 0 by a finite change in pressure.<br />

Problem 28.<br />

A container of volume V0 is divided into two equal parts by a removable<br />

wall. There are N0 moles of gas in equilibrium <strong>at</strong> temper<strong>at</strong>ure Ti on the left,<br />

the right side is empty. The container is isol<strong>at</strong>ed, no he<strong>at</strong> goes through. The<br />

internal wall is suddenly removed and the gas can expand freely until it finally<br />

fills up the whole container.<br />

(A) Is the gas in equilibrium just after the wall has been removed?<br />

(B) Is the free expansion a reversible process?<br />

(C) Does the entropy of the gas increase, decrease, or stay the same?<br />

(D) Does the energy of the gas increase, decrease, or stay the same?<br />

Assume th<strong>at</strong> we have a different, reversible process th<strong>at</strong> brings a gas <strong>at</strong><br />

constant internal energy from a temper<strong>at</strong>ure Ti and volume 1<br />

2V0 to a volume V0<br />

and temper<strong>at</strong>ure Tf . At each stage of this reversible process the gas is described<br />

by an internal energy function of the form U = u(T, V, N).<br />

(E) Is this reversible process adiab<strong>at</strong>ic?<br />

(F) Express the response function ( )<br />

∂T<br />

∂V<br />

the energy function u(T, V, N).<br />

U,N<br />

in terms of partial deriv<strong>at</strong>ives of<br />

(G) A mono-<strong>at</strong>omic ideal gas has u(T, V, N) = 3<br />

2NRT . How much does the<br />

temper<strong>at</strong>ure change in a free expansion from volume 1<br />

2V0 to V0?<br />

Problem 29.<br />

Under adiab<strong>at</strong>ic compression a gas changes its temper<strong>at</strong>ure when the pressure<br />

is changed, in a reversible manner. The response function describing this<br />

situ<strong>at</strong>ion is Ω = ( )<br />

∂T<br />

∂P S,N .


2.12. PROBLEMS FOR CHAPTER 2 105<br />

T V α<br />

Show th<strong>at</strong> Ω = Cp<br />

( )<br />

1 ∂V<br />

with α = − V ∂T p,N and Cp<br />

constant pressure.<br />

the he<strong>at</strong> capacity <strong>at</strong><br />

For an ideal mono-<strong>at</strong>omic gas αT = 1 and Cp = 5<br />

2NR. Find the function f<br />

defined by T = f(p) in this adiab<strong>at</strong>ic process.<br />

Problem 30.<br />

The Jacobian J is defined by<br />

J(x, y; u, v) =<br />

∂(x, y)<br />

∂(u, v) =<br />

Show th<strong>at</strong> J(T, S; p, V ) = 1.<br />

Problem 31.<br />

( )<br />

∂x<br />

∂u v<br />

( )<br />

∂y<br />

−<br />

∂v u<br />

( )<br />

∂x<br />

∂v u<br />

( )<br />

∂y<br />

∂u v<br />

A fluid undergoes an adiab<strong>at</strong>ic throttling process. This means th<strong>at</strong> the fluid<br />

is pressed through a porous plug, with initial pressure p1 and final pressure p2.<br />

(A) Show th<strong>at</strong> the enthalpy of the fluid is constant in this process.<br />

As a result of this process, the temper<strong>at</strong>ure changes (remember when you<br />

pump air in your bicycle tire, or when you let the air escape?). The amount of<br />

change is measured by the Joule-Thomson coefficient ηJT , defined by<br />

( )<br />

∂T<br />

ηJT =<br />

∂p<br />

(B) Show th<strong>at</strong> ηJT = V (T α − 1)<br />

Cp<br />

(C) Wh<strong>at</strong> is the value for an ideal gas?<br />

Problem 32.<br />

2NRTf<br />

5V<br />

The pressure of a gas of fermions <strong>at</strong> low temper<strong>at</strong>ure is given by p =<br />

[<br />

1 + 5π2<br />

12<br />

( T<br />

Tf<br />

H<br />

) 2 ]<br />

, where Tf is a m<strong>at</strong>erial dependent parameter called the<br />

Fermi temper<strong>at</strong>ure. One finds experimentally th<strong>at</strong> Tf = c ( N<br />

V<br />

constant. One also finds th<strong>at</strong> CV = 1<br />

2 π2 NR T<br />

Tf<br />

Helmholtz free energy of this gas.<br />

Problem 33.<br />

Assume the he<strong>at</strong> capacity <strong>at</strong> constant volume is given by<br />

) 2<br />

3 where c is a<br />

. Calcul<strong>at</strong>e the entropy and the


106 CHAPTER 2. THERMODYNAMIC POTENTIALS AND<br />

T 2<br />

CV = NR<br />

T 2 + T 2 0<br />

Show th<strong>at</strong> this implies th<strong>at</strong> the coefficient of thermal expansion is equal to<br />

zero.<br />

Problem 34.<br />

The internal energy of a system is given by U = cV T 4 , where c is a constant.<br />

Find the entropy, pressure, and chemical potential for this system.<br />

Problem 35.<br />

Suppose we know th<strong>at</strong><br />

F (T, V, N) = N<br />

∞∑<br />

( ) n<br />

N<br />

fn(T )<br />

V<br />

n=0<br />

Find the corresponding expansions for S(T, V, N), p(T, V, N), and U(T, V, N)<br />

Problem 36.<br />

Define a st<strong>at</strong>e function A by A = U − cT N, where c is a constant with the<br />

appropri<strong>at</strong>e dimensions to make the equ<strong>at</strong>ion dimensionally correct. Assume<br />

th<strong>at</strong> we have A(T, p, N), hence A as a function of the variables T , p , and N.<br />

Show th<strong>at</strong><br />

( )<br />

∂A<br />

= Cp − pV α − cN<br />

∂T p,N<br />

Problem 37.<br />

In cases where chemical reactions occur quickly, we can discuss the change<br />

in volume as a function of pressure in an adiab<strong>at</strong>ic process where we change the<br />

amount of m<strong>at</strong>erial or the chemical potential. The ) chemical volume expansion<br />

. The adiab<strong>at</strong>ic r<strong>at</strong>e of<br />

coefficient αchem is defined by αchem = 1<br />

V<br />

( ∂V<br />

∂µ<br />

change of amount ) of m<strong>at</strong>erial as a function of chemical potential is defined by<br />

. Show th<strong>at</strong><br />

RS = 1<br />

N<br />

( ∂N<br />

∂µ<br />

S,p<br />

κS,µ − κS,N = V<br />

N<br />

α 2 chem<br />

RS<br />

S,p


Chapter 3<br />

Phase transitions.<br />

In the first two chapters the philosophical and m<strong>at</strong>hem<strong>at</strong>ical basis for thermodynamics<br />

was developed. The notion of thermal equilibrium plays a very<br />

important role, and helps us to define the stable st<strong>at</strong>e of the system. Thermal<br />

equilibrium is defined between the system and the outside world, in cases where<br />

work can be done or he<strong>at</strong> can be exchanged. For example, if we say th<strong>at</strong> a<br />

system is <strong>at</strong> a given temper<strong>at</strong>ure we mean th<strong>at</strong> the average temper<strong>at</strong>ure of the<br />

system is equal to the temper<strong>at</strong>ure of some outside reservoir. In order for this<br />

st<strong>at</strong>ement to be meaningful, the system has to be very large, so rel<strong>at</strong>ive fluctu<strong>at</strong>ions<br />

are small. In addition, the fluctu<strong>at</strong>ions in temper<strong>at</strong>ure of the reservoir<br />

have to be much smaller, and hence the reservoir has to be much larger than<br />

the system.<br />

Thermodynamical equilibrium also implies th<strong>at</strong> different parts of the system<br />

are in equilibrium with each other. On average, different parts have the same<br />

density, and even though m<strong>at</strong>erial can flow back and forth, there are no net<br />

macroscopic currents.<br />

N<strong>at</strong>ure does cre<strong>at</strong>e extra complic<strong>at</strong>ions, though, and they are by far the most<br />

interesting. Wh<strong>at</strong> happens if a system can change its st<strong>at</strong>e abruptly? This is<br />

a combin<strong>at</strong>ion of the two ideas described above. In phase equilibrium we have<br />

two parts of the same system th<strong>at</strong> look quite different. M<strong>at</strong>erial can go from<br />

one phase to another. So on one hand we have two separ<strong>at</strong>e systems, but on<br />

the other hand they are just two parts of the same system.<br />

We can apply the ideas developed in the first two chapters to describe phase<br />

transitions. Th<strong>at</strong> is the topic of this chapter. In most cases we can change the<br />

phase of a system by applying a temper<strong>at</strong>ure or pressure change. Therefore,<br />

the free energy th<strong>at</strong> we need is the Gibbs free energy. The minimum energy<br />

principle for the Gibbs free energy then allows us to predict which phase is<br />

stable. We start this chapter by describing the prototype phase diagram for a<br />

system with variables p, V, T, S, N, µ. As a function of intensive st<strong>at</strong>e variables<br />

we can move continuously, but as a function of extensive st<strong>at</strong>e variables we have<br />

gaps. All these properties can be derived from the minimal energy principle.<br />

We can also follow phase equilibrium along phase boundaries, and this leads to<br />

107


108 CHAPTER 3. PHASE TRANSITIONS.<br />

some restrictions as well. Next, we discuss briefly the n<strong>at</strong>ure of binary phase<br />

diagrams, and how they can be discussed in the same manner. This is a very<br />

important and practical applic<strong>at</strong>ion of thermodynamics.<br />

So far we have assumed th<strong>at</strong> we have two different forms of the free energy,<br />

and we compare them to get the stable st<strong>at</strong>e. One way to connect the free<br />

energies of the two different phases is via the use of equ<strong>at</strong>ions of st<strong>at</strong>e. The<br />

most important one is the van der Waals equ<strong>at</strong>ion. We will use this equ<strong>at</strong>ion to<br />

show how a liquid to gas phase transition can be modelled. Integr<strong>at</strong>ions along<br />

isotherms allow us to connect the free energy of the two phases, and analyze<br />

phase equilibrium.<br />

The n<strong>at</strong>ure of this chapter is quite basic. We focus on a description of phase<br />

equilibrium and some simple consequences based on the assumption th<strong>at</strong> one<br />

can analyze the system by dividing it into two parts, each of which is internally<br />

in thermal equilibrium. We do allow net macroscopic flows of m<strong>at</strong>ter from one<br />

phase to another, though, during the transition. But <strong>at</strong> a given set of external<br />

conditions the thermodynamics will describe exactly how much m<strong>at</strong>erial is in<br />

phase one as compared to phase two. The r<strong>at</strong>io of these two amounts can be<br />

changed by changing temper<strong>at</strong>ure or pressure. We will analyze these r<strong>at</strong>ios <strong>at</strong><br />

the end of the chapter.<br />

3.1 Phase diagrams.<br />

Wh<strong>at</strong> happens when I suddenly change my st<strong>at</strong>e dram<strong>at</strong>ically?<br />

Study of rel<strong>at</strong>ions.<br />

In the previous two chapters we have studied the m<strong>at</strong>hem<strong>at</strong>ical description<br />

of thermal systems, i.e. systems which have a large number of internal degrees of<br />

freedom. We found the existence of many rel<strong>at</strong>ions between reversible changes<br />

in in st<strong>at</strong>e variables. Note th<strong>at</strong> the study of irreversible changes is much harder,<br />

but certainly not less important. Th<strong>at</strong> is for a different course, however.<br />

Entropy.<br />

The fundamental question in thermodynamics is: Wh<strong>at</strong> is entropy? Is it<br />

simply a quantity th<strong>at</strong> represents the energy stored in the internal degrees of<br />

freedom, or is it more? This question can not be answered in thermodynamics,<br />

and we have to apply st<strong>at</strong>istical mechanics to obtain one possible answer. In<br />

thermodynamics we simply use entropy as a st<strong>at</strong>e variable, and find ways to<br />

measure it.


3.1. PHASE DIAGRAMS. 109<br />

Consequences of rel<strong>at</strong>ions.<br />

Based on the laws of thermodynamics and the m<strong>at</strong>hem<strong>at</strong>ics of functions<br />

of many variables we were able to draw some very strong conclusions about<br />

response functions and stability. The rel<strong>at</strong>ions between response functions have<br />

important technical consequences. For example, if we try to design a m<strong>at</strong>erial<br />

with a large coefficient of thermal expansion α we know th<strong>at</strong> we always have<br />

α 2 T V = κT (Cp − CV ) and hence the right hand side should be large. But in<br />

a solid Cp − CV is always small and hence we need a m<strong>at</strong>erial with a large<br />

isothermal compressibility! And then the next question is: does th<strong>at</strong> have<br />

neg<strong>at</strong>ive consequences for the applic<strong>at</strong>ions of our new m<strong>at</strong>erial?<br />

Discontinuities.<br />

In this chapter we will focus on another important general aspect of the<br />

equ<strong>at</strong>ions of thermodynamics. We will investig<strong>at</strong>e the effects of discontinuities in<br />

the equ<strong>at</strong>ions of st<strong>at</strong>e. These discontinuities manifest themselves in the physical<br />

world as phase transitions. Wh<strong>at</strong> happens when a m<strong>at</strong>erial abruptly changes<br />

its properties?<br />

Wh<strong>at</strong> is the most important phase transition in the world?<br />

For example, zirconia is a very important ceramic m<strong>at</strong>erial. It has a high<br />

melting point, is very hard, and is inert. In short, it is very useful in many<br />

applic<strong>at</strong>ions. The problem, however, is th<strong>at</strong> <strong>at</strong> around 1100 centigrade zirconia<br />

transforms from a tetragonal high temper<strong>at</strong>ure phase into a monoclinic low temper<strong>at</strong>ure<br />

phase. This transition is very disruptive, and fractures the m<strong>at</strong>erial.<br />

The more we know about wh<strong>at</strong> happens <strong>at</strong> this phase transition, the more we<br />

can control it and make the consequences more benign.<br />

Practical importance.<br />

The study of phase transitions is of enormous practical importance. When<br />

does a m<strong>at</strong>erial become superconductive? How does magnetism change? In this<br />

and the following chapters we will study phase transitions from a macroscopic<br />

point of view. This is important, because this rel<strong>at</strong>es directly to the way phase<br />

transitions are observed. In order to explain why phase transitions happen, however,<br />

we need to look <strong>at</strong> the microscopic n<strong>at</strong>ure of the m<strong>at</strong>erial, and hence those<br />

answers are only found in st<strong>at</strong>istical mechanics. Nevertheless, thermodynamics<br />

will provide us again with important boundary conditions th<strong>at</strong> all m<strong>at</strong>erials will<br />

have to obey.<br />

Model system.


110 CHAPTER 3. PHASE TRANSITIONS.<br />

In order to make the discussion less cumbersome, we will consider a model<br />

system in which the only relevant st<strong>at</strong>e variables are {p, V, µ, N, S, T } . There<br />

is only one m<strong>at</strong>erial present. These restrictions are, however, not essential, and<br />

other systems can be discussed in a similar way. If we ignore chemical reactions<br />

and their influence on phase transitions, we can assume th<strong>at</strong> the amount of<br />

m<strong>at</strong>erial, N, is fixed. The three st<strong>at</strong>e variables which are easiest to change<br />

and control are the pressure p, the volume V, and the temper<strong>at</strong>ure T. Since we<br />

already prescribed the value of N, only two of the variables in the set {p, V, T } are<br />

independent. Since N is an extensive variable, we can choose any combin<strong>at</strong>ion<br />

of two variables from {p, V, T } . Systems like this are called pVT systems. Note<br />

th<strong>at</strong> we assume th<strong>at</strong> the system is in equilibrium, else equ<strong>at</strong>ions of st<strong>at</strong>e like<br />

p = f(V, T ) would not exist.<br />

Phase diagram.<br />

The variables p and T are intensive variables, and their values are chosen<br />

by bringing the system in thermal and mechanical contact with appropri<strong>at</strong>e<br />

reservoirs of pressure p and temper<strong>at</strong>ure T. From an experimental point of view,<br />

these are good handles on the system, and often easy to use. These reservoirs<br />

are infinitely large systems in thermodynamic equilibrium. Any choice for the<br />

values of p and T is possible, and hence we can give our model system arbitrary<br />

values of the pressure and temper<strong>at</strong>ure. Of course, we assume positive values<br />

for both p and T. Hence for any combin<strong>at</strong>ion of p and T we are able to specify<br />

the volume and all other aspects of the system. The system is represented by a<br />

plot in the p-T diagram, called a phase diagram. A very general phase diagram<br />

looks like figure 3.1. The solid lines indic<strong>at</strong>e phase boundaries. Crossing such<br />

a line in the p-T diagram will cause a jump in the value of the volume, and<br />

in the values of other extensive st<strong>at</strong>e variables. Sometimes the jump is only in<br />

the value of the deriv<strong>at</strong>ives, though. Most simple pVT systems have a phase<br />

diagram of the form shown in figure 3.1.<br />

Give an example of a solid to gas transition.<br />

An<strong>at</strong>omy of a phase diagram.<br />

The solid lines separ<strong>at</strong>e the well-known phases of most m<strong>at</strong>erials. At low<br />

temper<strong>at</strong>ure and high pressure the system is in a solid st<strong>at</strong>e. At higher temper<strong>at</strong>ure<br />

it either becomes a gas or a liquid. At the phase boundary the two phases<br />

can coexist. The points t and c are special. At the triple point t all three phases<br />

can coexist. At the critical point c the phase boundary between liquid and gas<br />

vanishes. The phase boundary between solid and liquid never vanishes, <strong>at</strong> least<br />

as far as we know. This implies th<strong>at</strong> gas and liquid are more alike than gas and<br />

solid, which makes sense intuitively.


3.1. PHASE DIAGRAMS. 111<br />

Figure 3.1: Model phase diagram for a simple model system.<br />

Critical point.<br />

The interpret<strong>at</strong>ion of a critical point is as follows. If the temper<strong>at</strong>ure is in<br />

the range (Tt, Tc), an increase in pressure will cause a phase transition between<br />

two st<strong>at</strong>es. At low pressure the st<strong>at</strong>e of the system resembles a gas, <strong>at</strong> high<br />

pressure a liquid. Since these two st<strong>at</strong>es are separ<strong>at</strong>ed by a phase transition,<br />

we can clearly identify them. If the temper<strong>at</strong>ure is above Tc, however, this<br />

clear separ<strong>at</strong>ion vanishes. At low pressure the system still behaves like a gas,<br />

and <strong>at</strong> high temper<strong>at</strong>ure like a liquid. The change from one to the other is<br />

continuous and it is not possible to uniquely identify where we change from one<br />

type of behavior to another. This continuous change is not possible from liquid<br />

to solid, as far as we know from experiments.<br />

There is some semantics which is useful to follow. Technically we make a<br />

distinction between liquid, gas, and vapor. All three phases are rel<strong>at</strong>ed, but can<br />

be distinguished experimentally by the following. If we are working <strong>at</strong> a fixed<br />

temper<strong>at</strong>ure and increase the pressure, we either see a transition or we do not.<br />

If we see a transition, we call those phases liquid (<strong>at</strong> high pressure) and vapor<br />

(<strong>at</strong> low pressure). If we do not see a transition, we call th<strong>at</strong> phase the gas phase.<br />

Hence we have a liquid if T < Tc and p > pt, a vapor if T < Tc and p < pt, and<br />

a gas if T > Tc.<br />

critical points in a solid.<br />

A famous example of a critical point in a solid is the α to γ phase transition in<br />

solid Ce. The phase diagram looks like figure 3.2. If we cool Ce <strong>at</strong> <strong>at</strong>mospheric<br />

pressure we have to go through the phase boundary, and the volume changes in<br />

a discontinuous manner. If the whole m<strong>at</strong>erial were uniform, this would not give<br />

any problems. But any imperfections in the m<strong>at</strong>erial cause some parts of the


112 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.2: Phase diagram for solid Ce.<br />

m<strong>at</strong>erial to transform before other parts, causing large strains in the m<strong>at</strong>erial.<br />

As a result, large single crystals of α-Ce are hard to make. The solution to this<br />

problem is to go around the critical point. First increase the pressure, than<br />

decrease the temper<strong>at</strong>ure, and finally decrease the pressure again.<br />

Energy consider<strong>at</strong>ions.<br />

At a phase boundary two distinct phases of the m<strong>at</strong>erial can coexist. This is<br />

only possible if these two phases have the same energy <strong>at</strong> the phase boundary.<br />

If not, one phase would be preferred. The question is, which energy plays a<br />

role. Since the phase boundary is determined by specific values of p and T, and<br />

since N is kept constant, the energy of interest is the Gibbs energy G(T, p, N).<br />

Therefore, for a p<strong>at</strong>h in p-T space crossing a phase boundary, the Gibbs free<br />

energy is continuous, but its deriv<strong>at</strong>ives are not. The reason for this is th<strong>at</strong> the<br />

Gibbs free energy is minimal as a function of all internal parameters <strong>at</strong> a given<br />

value of p and T. The internal parameters define the phase we are in. Therefore<br />

phase transitions are found by studying singular behavior in the deriv<strong>at</strong>ives of<br />

the Gibbs free energy.<br />

Discontinuity in the volume.<br />

The Gibbs energy is continuous <strong>at</strong> a phase boundary. If the Gibbs free<br />

energy were not continuous <strong>at</strong> a phase boundary, a picture of the Gibbs en-<br />

ergy as ) a function of pressure would look like figure 3.3 for example (note th<strong>at</strong><br />

≤ 0 because of stability). One could then follow the lower branch<br />

( ∂ 2 G<br />

∂p 2<br />

T,N<br />

across the phase boundary, and have a st<strong>at</strong>e of lower energy, which by definition<br />

would be the stable st<strong>at</strong>e. This would simply indic<strong>at</strong>e th<strong>at</strong> we marked the<br />

phase boundary <strong>at</strong> the wrong place. Hence the Gibbs free energy as a function


3.1. PHASE DIAGRAMS. 113<br />

Figure 3.3: Gibbs energy across the phase boundary <strong>at</strong> constant temper<strong>at</strong>ure,<br />

wrong picture.<br />

of pressure <strong>at</strong> constant temper<strong>at</strong>ure and N looks like 3.4.<br />

Once we have the Gibbs energy) as a function of pressure, the volume follows<br />

autom<strong>at</strong>ically through V =<br />

and this is shown in figure 3.5. First<br />

( ∂G<br />

∂p<br />

T,N<br />

of all, because of the curv<strong>at</strong>ure of the Gibbs free energy the volume decreases<br />

with pressure. Second, <strong>at</strong> the phase transition the volume jumps down in a<br />

discontinuous manner, it never jumps up.<br />

Is it possible for the volume to be continuous as a function of pressure <strong>at</strong> the<br />

phase transition?<br />

Is it possible for the volume to decrease as a function of temper<strong>at</strong>ure <strong>at</strong> a phase<br />

transition?<br />

Other pictures of the phase transition, VT diagram.<br />

In the V-T plane this phase diagram looks like figure 3.6. In this diagram<br />

we see regions which are not accessible, because the volume of the system jumps<br />

<strong>at</strong> the phase boundary. There is no continuous p<strong>at</strong>h from the solid phase to<br />

the others, but there is a continuous p<strong>at</strong>h from liquid to gas, around the critical<br />

point. Inaccessible regions only show up in diagrams when extensive st<strong>at</strong>e<br />

variables are involved. Wh<strong>at</strong> happens when the system is prepared <strong>at</strong> a given<br />

volume V0 <strong>at</strong> a high temper<strong>at</strong>ure and then cooled down to a temper<strong>at</strong>ure which


114 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.4: Gibbs energy across the phase boundary <strong>at</strong> constant temper<strong>at</strong>ure,<br />

correct picture.<br />

Figure 3.5: Volume versus pressure <strong>at</strong> the phase transition.


3.1. PHASE DIAGRAMS. 115<br />

would force the st<strong>at</strong>e of the system to be in an inaccessible region? The answer<br />

is simple, the system will phase separ<strong>at</strong>e. Part of it will have a small volume,<br />

part of it a large volume, and both phases are in equilibrium with each other.<br />

The amount of each phase is determined by where in the inaccessible region the<br />

point (T, V0) is situ<strong>at</strong>ed.<br />

Figure 3.6: Model phase diagram for a simple model system in V-T space.<br />

Phase separ<strong>at</strong>ion analyzed.<br />

As an example, consider th<strong>at</strong> the system has volume V0 and is cooled down<br />

<strong>at</strong> constant volume to a temper<strong>at</strong>ure T below Tc, but above Tt. The volume<br />

per mole (specific volume) of the gas and the liquid <strong>at</strong> this temper<strong>at</strong>ure and<br />

<strong>at</strong> a pressure p are vg(p, T ) and vl(p, T ). Because gas and liquid will be in<br />

equilibrium, they will be <strong>at</strong> the same pressure. We need to determine the<br />

amounts of m<strong>at</strong>erial Ng and Nl in the gas and liquid phase. We obviously have:<br />

and<br />

N = Ng + Nl<br />

(3.1)<br />

V0 = Ngvg(p, T ) + Nlvl(p, T ) (3.2)<br />

This seems to be incomplete, but we have to keep in mind th<strong>at</strong> the pressure is<br />

determined from the phase diagram 3.1. Hence we know exactly how the system<br />

phase separ<strong>at</strong>es.<br />

We can combine these equ<strong>at</strong>ions by using the volume per mole v0 and write<br />

which is equivalent to<br />

(Ng + Nl)v0 = Ngvg(p, T ) + Nlvl(p, T ) (3.3)


116 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.7: Model phase diagram for a simple model system in p-V space.<br />

or<br />

Ng(vg(p, T ) − v0) = Nl(v0 − vl(p, T )) (3.4)<br />

Ng<br />

Nl<br />

= v0 − vl(p, T )<br />

vg(p, T ) − v0<br />

(3.5)<br />

which is called the lever rule. The phase with a molar volume most different<br />

from the actual volume has the smallest amount of m<strong>at</strong>erial in it.<br />

If we have our initial volume larger than the critical volume, V0 > Vc and<br />

we cool the system, we see th<strong>at</strong> <strong>at</strong> the temper<strong>at</strong>ure when we hit the forbidden<br />

region droplets of m<strong>at</strong>erial with a small volume (liquid) will form. The more<br />

we lower the temper<strong>at</strong>ure, the more liquid we get. On the other hand, if the<br />

original volume is less than the critical volume, V0 < Vc, and we cool down, we<br />

will see droplets of the high volume phase, gas, form in the liquid. The more<br />

we cool, the more gas we get! Th<strong>at</strong> seems counterintuitive, but we have to<br />

remember th<strong>at</strong> when we cool the mixture, the pressure decreases, and <strong>at</strong> lower<br />

pressure we will see more gas phase!<br />

The same system in a p-V diagram looks like figure 3.7. Notice again the<br />

forbidden regions. It is also possible to draw the three dimensional surface<br />

V (p, T, N) for fixed N, but th<strong>at</strong> is beyond my capabilities and artistic talents.<br />

3.2 Clausius-Clapeyron rel<strong>at</strong>ion.<br />

So far we made some general st<strong>at</strong>ements about wh<strong>at</strong> is happening <strong>at</strong> a phase<br />

boundary. Our next task is to give a more quantit<strong>at</strong>ive description of the phase<br />

boundaries. <strong>Thermodynamics</strong> limits the freedom we have in drawing phase<br />

diagrams!


3.2. CLAUSIUS-CLAPEYRON RELATION. 117<br />

basic energy.<br />

Wh<strong>at</strong> is the best way to describe m<strong>at</strong>hem<strong>at</strong>ically wh<strong>at</strong> is happening? Since<br />

in terms of p and T all values of these st<strong>at</strong>e variables are possible, we will use<br />

the pressure and temper<strong>at</strong>ure as independent variables. The free energy in this<br />

case is the Gibbs potential G = U − T S + pV . The Gibbs free energy is a<br />

minimum as a function of all internal parameters when p, T, and N are fixed.<br />

Hence the st<strong>at</strong>e with the lowest Gibbs free energy is stable in a p-T diagram.<br />

In our simple system the Euler rel<strong>at</strong>ion tells us th<strong>at</strong> G = µN. Since N is fixed,<br />

the chemical potential in this case is equal to the Gibbs free energy per mole,<br />

and is often used in stead of the Gibbs free energy in this case. To be precise,<br />

however, we will use the Gibbs energy as our basic energy.<br />

Phase boundary.<br />

In order to determine the phase boundary between to phases we need an<br />

analytical form of the Gibbs free energy for both phases. Assume th<strong>at</strong> we have<br />

expressions for this energy in each phase, which we can extend to all values in<br />

p-T space by analytic continu<strong>at</strong>ion. The phase boundary then follows from<br />

G1(p, T, N) = G2(p, T, N) (3.6)<br />

and since N is fixed, this yields an equ<strong>at</strong>ion p = pco(T ) for the pressure as a<br />

function of temper<strong>at</strong>ure. This rel<strong>at</strong>ion describes the coexistence curve of the<br />

two phases.<br />

The slope) of the Gibbs free energy as a function op pressure is equal to the<br />

volume, = V , and hence as a function of the pressure the Gibbs free<br />

( ∂G<br />

∂p<br />

T,N<br />

energy is always increasing (volumes are positive by definition). The second )<br />

order deriv<strong>at</strong>ive of the Gibbs free energy with pressure is equal to ,<br />

( ∂V<br />

∂p<br />

and is always neg<strong>at</strong>ive because of stability requirements. Therefore the slope of<br />

the Gibbs free energy as a function of pressure is decreasing. A picture of the<br />

Gibbs free energy of the two phases which we study might look like the figure<br />

3.8. The lowest curve gives the stable st<strong>at</strong>e for each value of the pressure. At<br />

the phase transition the slope of the Gibbs free energy jumps to a smaller value,<br />

and hence the volume jumps to a smaller value.<br />

In a picture of volume versus pressure, why can the volume not have the same<br />

value for two different values of the pressure.<br />

Following the coexistence curve <strong>at</strong> both sides.<br />

The differential for the Gibbs free energy is given by<br />

T,N


118 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.8: Gibbs energy across the phase boundary <strong>at</strong> constant temper<strong>at</strong>ure<br />

for both phases.<br />

dG = V dp − SdT + µdN (3.7)<br />

but since N is constant, the last term is zero. The partial deriv<strong>at</strong>ives of the<br />

Gibbs free energy give us V(p,T,N) and S(p,T,N) for both phases. If we are <strong>at</strong><br />

a point on the coexistence curve, and make small changes in the temper<strong>at</strong>ure<br />

and pressure in such a manner th<strong>at</strong> we remain on the coexistence curve, the<br />

Gibbs free energy of the two phases remains the same (by definition), and hence<br />

along the coexistence curve we find:<br />

V1dp − S1dT = V2dp − S2dT (3.8)<br />

or<br />

( )<br />

dp<br />

=<br />

dT co<br />

S1(p, T ) − S2(p, T )<br />

(3.9)<br />

V1(p, T ) − V2(p, T )<br />

which is the famous Clausius-Clapeyron rel<strong>at</strong>ion. The pressure and temper<strong>at</strong>ure<br />

on the right hand side are rel<strong>at</strong>ed by p = pco(T ). Both the numer<strong>at</strong>or and the<br />

denomin<strong>at</strong>or on the right hand side are often divided by N, yielding a r<strong>at</strong>io of<br />

differences in molar volume and molar entropy.<br />

L<strong>at</strong>ent he<strong>at</strong>.<br />

The l<strong>at</strong>ent he<strong>at</strong> of a phase transform<strong>at</strong>ion is the amount of he<strong>at</strong> needed to<br />

transform the system <strong>at</strong> a given pressure and temper<strong>at</strong>ure along the coexistence<br />

curve completely from one phase to the other. The l<strong>at</strong>ent he<strong>at</strong> is positive if we<br />

transform m<strong>at</strong>erial from a low temper<strong>at</strong>ure phase to a high temper<strong>at</strong>ure phase.<br />

We indeed need energy to melt ice and boil w<strong>at</strong>er! The fact th<strong>at</strong> the l<strong>at</strong>ent<br />

he<strong>at</strong> is positive is easy to prove, similar to how we showed th<strong>at</strong> the volume


3.2. CLAUSIUS-CLAPEYRON RELATION. 119<br />

decreases ( when we go to higher pressure. In the previous chapter we saw th<strong>at</strong><br />

2<br />

∂ G<br />

∂T 2<br />

)<br />

≤ 0 and hence the picture of G versus T looks qualit<strong>at</strong>ively like figure<br />

p,N<br />

3.8. Since S = − ( )<br />

∂G<br />

∂T a decrease in the slope of G versus T is a jump down<br />

p,N<br />

in the value of -S and hence an increase in S.<br />

The l<strong>at</strong>ent he<strong>at</strong> is denoted by L1→2(T ) for the process going from phase 1<br />

to phase 2 <strong>at</strong> a temper<strong>at</strong>ure T and a corresponding value of the pressure along<br />

the coexistence curve. Since he<strong>at</strong> is rel<strong>at</strong>ed to changes in the entropy, we have<br />

L1→2(T ) = T (S2(pco(T ), T, N) − S1(pco(T ), T, N)) (3.10)<br />

and the Clausius-Clapeyron rel<strong>at</strong>ion takes the form<br />

( )<br />

dp<br />

l1→2(T )<br />

=<br />

dT co T (v2(p, T ) − v1(p, T ))<br />

(3.11)<br />

where we use the lower case characters to indic<strong>at</strong>e the molar volume and molar<br />

l<strong>at</strong>ent he<strong>at</strong> (i.e. the quantities are per mole).<br />

Consequence for phase boundary.<br />

The Clausius-Clapeyron rel<strong>at</strong>ion 3.9 is still in its general form. Note th<strong>at</strong> if 2<br />

denotes the high temper<strong>at</strong>ure phase, and if the volume of the high temper<strong>at</strong>ure<br />

phase is largest, both the numer<strong>at</strong>or and the denomin<strong>at</strong>or are positive. Hence<br />

the coexistence curve pco(T ) always has a positive slope in such a case. On the<br />

other hand, if we consider the phase boundary between w<strong>at</strong>er and ice, we know<br />

th<strong>at</strong> ice has a larger volume per mole, and hence the denomin<strong>at</strong>or is neg<strong>at</strong>ive.<br />

In this case the slope of the coexistence curve is neg<strong>at</strong>ive!<br />

Example.<br />

As a specific example take the transition from liquid to gas. In this case<br />

the volume of the liquid is much smaller than the volume of the gas and can be<br />

ignored. Using the ideal gas law gives:<br />

( )<br />

dp<br />

≈<br />

dT co<br />

pco(T )l1→2(T )<br />

RT 2<br />

(3.12)<br />

In many cases the l<strong>at</strong>ent he<strong>at</strong> is only very weakly temper<strong>at</strong>ure dependent. In<br />

th<strong>at</strong> case we can do the integr<strong>at</strong>ion and obtain<br />

pco(T ) ≈ p0e l (<br />

1→2 1<br />

R T −<br />

0 1<br />

)<br />

T<br />

(3.13)<br />

which indeed has a positive slope. This is a useful formula for a number of<br />

applic<strong>at</strong>ions. The limit T→ ∞ of pco(T ) is finite. Consequently, this rel<strong>at</strong>ion<br />

does not hold for the complete solid-liquid coexistence curve (but might be close<br />

for part of it). For such a phase boundary we need to take the values of the


120 CHAPTER 3. PHASE TRANSITIONS.<br />

volumes of both phases into account. Also, this rel<strong>at</strong>ion does not hold near a<br />

critical point, where the l<strong>at</strong>ent he<strong>at</strong> goes to zero, and the molar volumes become<br />

equal.<br />

To continue the study of this example, consider a closed volume V containing<br />

N moles of w<strong>at</strong>er. The temper<strong>at</strong>ure T is the control variable. If we are <strong>at</strong> room<br />

temper<strong>at</strong>ure, and the volume is chosen in such a manner th<strong>at</strong> the pressure is<br />

around <strong>at</strong>mospheric pressure, we will see w<strong>at</strong>er <strong>at</strong> the bottom, an interface, and<br />

vapor <strong>at</strong> the top.<br />

Which implicit assumption did we make?<br />

This system lends itself to an easy experiment. By measuring the level of the<br />

interface, we can determine the volume of the liquid and gas, Vl and Vg = V −Vl.<br />

By doing some optical absorption measurement we can determine the density<br />

of the vapor and hence we find the amount of vapor Ng and hence we know the<br />

amount of w<strong>at</strong>er Nl = N − Ng. If the gas can be described by the ideal gas law,<br />

we find pco(T ) = NgRT<br />

. Vg<br />

Up or down or wh<strong>at</strong>??<br />

Wh<strong>at</strong> happens if we increase the temper<strong>at</strong>ure? Will the interface go up or<br />

down? The answer is, th<strong>at</strong> depends on the volume V. The rel<strong>at</strong>ive volumes of<br />

the liquid and gas phase follow from<br />

V = Nlvl(T, pco(T )) + Ngvg(T, pco(T )) (3.14)<br />

for all temper<strong>at</strong>ures as long as we are below the critical temper<strong>at</strong>ure Tc. If<br />

the volume is larger than the critical volume Vc the phase diagram 3.6 shows<br />

th<strong>at</strong> <strong>at</strong> some temper<strong>at</strong>ure below Tc all m<strong>at</strong>erial will be vapor, and hence we see<br />

the interface drop to the bottom of the container. If the volume V is less than<br />

the critical volume Vc, on the other hand, all m<strong>at</strong>erial will transform to liquid,<br />

and the interface will actually rise! This seems counter-intuitive. How can we<br />

increase the temper<strong>at</strong>ure and see more liquid form? The answer is found in<br />

the fact th<strong>at</strong> the total volume is constant, and th<strong>at</strong> the increase in temper<strong>at</strong>ure<br />

comes with an increase in pressure which is sufficient to condens<strong>at</strong>e the vapor<br />

into liquid if the volume is small enough (i.e. smaller than Vc).<br />

Wh<strong>at</strong> happens if V = Vc? In this case the interface does not go to the top<br />

or the bottom of the container, but ends up somewhere in the middle. So wh<strong>at</strong><br />

happens when T reaches Tc. The interface is supposed to disappear, and it does<br />

indeed. If we reach Tc in this experiment, the properties of the liquid phase and<br />

the vapor phase become more and more similar, and <strong>at</strong> the critical interface<br />

they are identical. We only see the interface between w<strong>at</strong>er and vapor because<br />

the index of refraction is different in the two phases. At the critical point the


3.3. MULTI-PHASE BOUNDARIES. 121<br />

indices of refraction become identical, and the interface disappears from our<br />

view. The same holds for any type of measurement.<br />

3.3 Multi-phase boundaries.<br />

At the triple point in the phase diagram three phases coexist, and hence the<br />

equ<strong>at</strong>ions determining the triple point are<br />

G1(p, T, N) = G2(p, T, N) = G3(p, T, N) (3.15)<br />

Since the total amount of m<strong>at</strong>erial is fixed, there are two equ<strong>at</strong>ions for two<br />

unknowns. In general, this set of equ<strong>at</strong>ions has one unique solution. Therefore,<br />

we only find one triple point.<br />

Critical point.<br />

It is also easy to see th<strong>at</strong> in general we have only one critical point. The<br />

coexistence line is determined by one equ<strong>at</strong>ion for the free energy, and the critical<br />

point also st<strong>at</strong>es th<strong>at</strong> a parameter like the volume has to be the same.<br />

In this case the simultaneous equ<strong>at</strong>ions are G1(p, T, N) = G2(p, T, N) and<br />

V1(p, T, N) = V2(p, T, N).<br />

Quadruple point.<br />

Suppose in our model system we would have four different phases. The solid<br />

could, for example, change its structure <strong>at</strong> a certain temper<strong>at</strong>ure, or become<br />

magnetic. Would it be possible to have all four phases coexist? The answer is<br />

clearly no for our simple case. Since the only two independent variables are p and<br />

T, the set of three equ<strong>at</strong>ions determining this special point is over-complete and<br />

in general no solution exists. If we want to solve three simultaneous equ<strong>at</strong>ions,<br />

we need three independent variables. One could allow N to change by bringing<br />

the system into contact with a reservoir of constant chemical potential. This is<br />

not possible, however, since in th<strong>at</strong> case the system would be described by the<br />

set of st<strong>at</strong>e variables {p, T, µ} and these are all intensive. Therefore, we have<br />

to introduce another st<strong>at</strong>e variable. For example, take a (possibly) magnetic<br />

system with total magnetic moment M. For simplicity, we assume th<strong>at</strong> the<br />

direction of this moment is not important. The corresponding intensive st<strong>at</strong>e<br />

variable is the magnetic field H. In this case our system is described by the four<br />

variables {p, T, H, N} and we now use the modified, magnetic Gibbs free energy<br />

G(p, T, H, N) = U − T S + pV − HM (3.16)<br />

which is still equal to µN, since we now have to add one term to the Euler<br />

rel<strong>at</strong>ion. A point where all four phases coexist is now described by certain<br />

values {pq, Tq, Hq} and if such a quadruple point exists, it needs in general a


122 CHAPTER 3. PHASE TRANSITIONS.<br />

non-zero applied magnetic field. Note th<strong>at</strong> we ignored the energy stored in the<br />

field (see previous chapter). This energy is independent of the particular phase,<br />

however, and cancels in the equ<strong>at</strong>ions.<br />

Multi-component systems.<br />

If M+1 phases coexist for a certain system, we need to solve M equ<strong>at</strong>ions.<br />

The number of unknowns cannot be smaller than M. Hence we arrive <strong>at</strong> the<br />

general conclusion th<strong>at</strong><br />

The number of phases which can coexist for a system described by<br />

M independent intensive st<strong>at</strong>e variables is <strong>at</strong> most equal to M+1.<br />

For example, we have a system with r different types of m<strong>at</strong>erial, with number<br />

of moles Ni, i = 1 . . . r, for each type. Assume th<strong>at</strong> the total amount of<br />

m<strong>at</strong>erial is fixed, hence we have something like ∑ wiNi = N is constant, where<br />

the weights wi depend on the chemical reactions th<strong>at</strong> are described. In th<strong>at</strong><br />

case, using also p and T as variables, we have r+1 independent variables th<strong>at</strong><br />

can change, and hence we have <strong>at</strong> most r+2 phases th<strong>at</strong> can coexist <strong>at</strong> one time.<br />

Applic<strong>at</strong>ion to chemical reactions.<br />

In a chemical experiment one might look <strong>at</strong> reactions between two species<br />

A and B, according to A + B ↔ AB. The total amount of moles of A and B,<br />

either free or combined, are fixed. The rel<strong>at</strong>ive number can change. Therefore,<br />

this only leaves one st<strong>at</strong>e variable to describe the st<strong>at</strong>e of the reaction, and<br />

we could take this variable NAB, the number of molecules AB. It is better to<br />

use the corresponding chemical potential µAB, however, since th<strong>at</strong> is in general<br />

the st<strong>at</strong>e variable which is controlled in chemical experiments. The other st<strong>at</strong>e<br />

variable we can easily vary is the temper<strong>at</strong>ure T. Since there are two independent<br />

st<strong>at</strong>e variables, <strong>at</strong> most three phases can coexist. One phase is easily identified<br />

as A and B free, the other as A and B bound in AB. Requiring these two<br />

phases to coexist elimin<strong>at</strong>es one independent st<strong>at</strong>e variable, and leads to an<br />

equ<strong>at</strong>ion for the coexistence curve of the form NAB(T ). Hence the st<strong>at</strong>e of the<br />

reaction is completely determined by the temper<strong>at</strong>ure. If we also allow the AB<br />

m<strong>at</strong>erial to solidify, we introduce a third phase. The AB m<strong>at</strong>erial is either in<br />

solution or precipit<strong>at</strong>ed out. There is only one temper<strong>at</strong>ure Tt <strong>at</strong> which all three<br />

phases coexist. Interesting problems arise when one assumes th<strong>at</strong> only the AB<br />

molecules in solution are able to dissoci<strong>at</strong>e within the time of the experiment.<br />

An efficient reaction p<strong>at</strong>h then often requires manipul<strong>at</strong>ing both µAB and T to<br />

obtain a complete reaction.


3.4. BINARY PHASE DIAGRAM. 123<br />

3.4 Binary phase diagram.<br />

Basic variables.<br />

A very important applic<strong>at</strong>ion is found in binary phase diagrams. Suppose<br />

we have two m<strong>at</strong>erials, A and B, and each can dissolve in the other. Suppose we<br />

have NA moles of m<strong>at</strong>erial A with B dissolved in it, and NB moles of m<strong>at</strong>erial B<br />

with A dissolved in it. Since m<strong>at</strong>erial cannot disappear, we have NA + NB = N<br />

is constant. Therefore, <strong>at</strong> most four phases can coexist <strong>at</strong> a given time. In these<br />

solid solutions we typically work <strong>at</strong> <strong>at</strong>mospheric pressure, however, which for<br />

practical purposes is almost equal to zero (for solids!!) and which therefore can<br />

be ignored. The variables th<strong>at</strong> can be changed easily are the temper<strong>at</strong>ure T<br />

and the fraction of m<strong>at</strong>erial B in the total m<strong>at</strong>erial, cB, which is the r<strong>at</strong>io of<br />

the amount of pure m<strong>at</strong>erial B (without any A dissolved in it) to N. In such a<br />

phase diagram <strong>at</strong> most three phases can coexist.<br />

N<strong>at</strong>ure of cB.<br />

Although the variable cB is intensive, it is only so because we divided by<br />

N to get a density. The variable cB really measures an amount of m<strong>at</strong>erial,<br />

and really plays the role of an extensive st<strong>at</strong>e variable. The true intensive st<strong>at</strong>e<br />

variable would be the corresponding chemical potential, but th<strong>at</strong> one is hard<br />

to control directly in alloys. Just as we saw in the VT diagram, we will have<br />

regions in the cBT diagram which are not allowed, which do not correspond to<br />

st<strong>at</strong>es of the system.<br />

Typical phase diagram.<br />

A typical phase diagram for a binary alloy looks like figure 3.9. The curv<strong>at</strong>ure<br />

of the phase boundaries is always the same as in this figure, and can be<br />

derived from stability criteria. At low temper<strong>at</strong>ures, for very low or very high<br />

concentr<strong>at</strong>ions cB of B we have one type of m<strong>at</strong>erial. If the concentr<strong>at</strong>ion of B<br />

is in between, the m<strong>at</strong>erial will phase separ<strong>at</strong>e into regions rich in B, consisting<br />

of B(A) m<strong>at</strong>erial where A is dissolved in B, and regions poor in B, consisting<br />

of A(B) m<strong>at</strong>erial. The division between these can easily be calcul<strong>at</strong>ed if we<br />

know where the phase boundaries are. We use a similar lever rule as we used<br />

for the volumes earlier in this chapter. At higher temper<strong>at</strong>ures the system is<br />

either liquid, pure m<strong>at</strong>erial, or a mixture of liquid and pure m<strong>at</strong>erial. There is<br />

exactly one temper<strong>at</strong>ure and concentr<strong>at</strong>ion where all three phases can coexist.<br />

This point is called the eutectic point. Of course, when we have more phases or<br />

ordered structures, we can have more eutectic points.<br />

Crossing the line.


124 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.9: Typical binary phase diagram with regions L=liquid, A(B)=B<br />

dissolved in A, and B(A)=A dissolved in B.<br />

Suppose we prepare a liquid <strong>at</strong> high temper<strong>at</strong>ure containing both elements<br />

A and B and the concentr<strong>at</strong>ion of B is C0 B is low. We now cool the mixture and<br />

end in a stable region of m<strong>at</strong>erial A(B). Wh<strong>at</strong> happens in between and wh<strong>at</strong> is<br />

the result of our experiment? In order to describe this, see figure 3.10.<br />

When we reach the upper boundary of the forbidden region, we have an<br />

equilibrium of solid m<strong>at</strong>erial with a concentr<strong>at</strong>ion of B of CL and liquid with<br />

the original concentr<strong>at</strong>ion. When we lower the temper<strong>at</strong>ure, more solid will<br />

form, and the concentr<strong>at</strong>ion of the solid phase will increase. But since we form<br />

more solid with too small an amount of B, the concentr<strong>at</strong>ion of B in the liquid<br />

will also increase. Near the lower boundary of the forbidden region we have<br />

mainly solid m<strong>at</strong>erial with almost the original concentr<strong>at</strong>ion of B and some<br />

liquid with a high concentr<strong>at</strong>ion of B m<strong>at</strong>erial. Since no m<strong>at</strong>erial is lost, we can<br />

easily find the amounts of m<strong>at</strong>erial Nl in the liquid phase and Ns in the solid<br />

phase:<br />

Ns(c 0 B − c1) = Nl(c2 − c 0 B) (3.17)<br />

This rule is called the lever rule.<br />

In the scenario described above we have made one basic assumption, th<strong>at</strong><br />

the system is always in thermal equilibrium! In other words, we assume th<strong>at</strong><br />

the process is infinitely slow. Wh<strong>at</strong> are the time scales involved? First of all,<br />

the cooling time. The second time scale is rel<strong>at</strong>ed to the diffusion of m<strong>at</strong>erial B<br />

in the liquid and in the solid. The last one is the limiting factor, and this time<br />

can be long. In addition, in order for the solid to start forming we need initial<br />

seeds (nucle<strong>at</strong>ion of m<strong>at</strong>erial). In many cases, this is not a limiting factor.<br />

Suppose we cool the m<strong>at</strong>erial and grains of solid form. Suppose th<strong>at</strong> the<br />

diffusion of B m<strong>at</strong>erial is fast, and we end up with only grains of solid with<br />

the correct composition. Since we are now dealing with a solid, most likely we<br />

will have empty space in between the grains! If the original melt had some


3.4. BINARY PHASE DIAGRAM. 125<br />

Figure 3.10: Solidific<strong>at</strong>ion in a reversible process.<br />

impurities, it is possible for these impurities to collect in these empty spaces.<br />

In general, the m<strong>at</strong>erial <strong>at</strong> the grain boundaries will have a lower density and<br />

the volume of the resulting m<strong>at</strong>erial is too large. Wh<strong>at</strong> can we do about th<strong>at</strong>?<br />

If the m<strong>at</strong>erial is a metal, we can roll it and press it to get rid of these empty<br />

spaces. For ceramics, th<strong>at</strong> is much harder.<br />

But th<strong>at</strong> is not all folks. Typically, the diffusion of B m<strong>at</strong>erial in the solid<br />

is incomplete. Therefore, the grains will have a concentr<strong>at</strong>ion gradient. The<br />

centers are poor in B m<strong>at</strong>erial. Near the grain boundaries we have more or<br />

less the correct concentr<strong>at</strong>ion. The excess of B m<strong>at</strong>erial will be in between the<br />

grains! Wh<strong>at</strong> to do now? Since there is a concentr<strong>at</strong>ion gradient, there is a<br />

difference in chemical potential, and there is a driving force which wants to<br />

make the concentr<strong>at</strong>ion uniform. Since the diffusion is always much better <strong>at</strong><br />

higher temper<strong>at</strong>ure, we want to keep the solid <strong>at</strong> a temper<strong>at</strong>ure just below the<br />

lower bound for a while. This annealing process is very important when making<br />

m<strong>at</strong>erials. Sometimes the times involved are quite long. Sometimes other tricks<br />

have to be applied during the anneal.<br />

The previous discussion shows th<strong>at</strong> phase transitions in solids are quite difficult<br />

to deal with. Th<strong>at</strong> is why we deal with gas to liquid transitions in most<br />

of this chapter. In the next chapter we will discuss magnetic phase transitions,<br />

which also are much easier to deal with. Nevertheless, a study of phase transitions<br />

in solids is of gre<strong>at</strong> practical importance! Just talk with your mechanical<br />

engineering friends.<br />

Eutectic point.<br />

The eutectic point is of gre<strong>at</strong> practical importance. Opposed to the gas-liquid<br />

coexistence, phase separ<strong>at</strong>ion <strong>at</strong> low temper<strong>at</strong>ures is a very slow process, since<br />

it involves <strong>at</strong>omic motion in a solid. It is therefore possible to cre<strong>at</strong>e meta-stable<br />

m<strong>at</strong>erials in the forbidden regions by cooling a sample very rapidly. Typically


126 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.11: Typical binary phase diagram with intermedi<strong>at</strong>e compound<br />

AB, with regions L=liquid, A(B)=B dissolved in A, B(A)=A dissolved in<br />

B, and AB= AB with either A or B dissolved.<br />

this is done by dropping a drop of liquid into cold w<strong>at</strong>er. If we are not <strong>at</strong> the<br />

eutectic point, it is almost certain th<strong>at</strong> small amounts of m<strong>at</strong>erial A(B) or B(A)<br />

will form during cooling (when we go through the top forbidden regions) and<br />

the final m<strong>at</strong>erial will not be homogeneous. If we are <strong>at</strong> the eutectic point, this<br />

region where m<strong>at</strong>erial separ<strong>at</strong>es is reached <strong>at</strong> the lowest possible temper<strong>at</strong>ure<br />

(making the process the slowest), furthest from the stable phases (increasing<br />

the energy needed for the process). Meta-stable m<strong>at</strong>erials <strong>at</strong> a concentr<strong>at</strong>ion<br />

corresponding to the eutectic point are therefore easiest to prepare.<br />

Intermedi<strong>at</strong>e compounds.<br />

In many cases there are stable compounds of the m<strong>at</strong>erials A and B present,<br />

and the phase diagram then looks different. A typical example is figure 3.11<br />

where we included a stable m<strong>at</strong>erial AB . The overall structure and interpret<strong>at</strong>ion<br />

is the same, except th<strong>at</strong> we now have a region of AB m<strong>at</strong>erial in which<br />

either A or B is dissolved. Quite often the range of stability of the intermedi<strong>at</strong>e<br />

compound is very small, so small th<strong>at</strong> it collapses to a single line in the diagram,<br />

see figure 3.12. An example of an impossible phase diagram is figure 3.13, where<br />

we have four phases coexisting <strong>at</strong> one point.<br />

3.5 Van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e.<br />

More quantit<strong>at</strong>ive approaches based on energy.<br />

The previous discussion of phase transitions was very general. In order to


3.5. VAN DER WAALS EQUATION OF STATE. 127<br />

Figure 3.12: Typical binary phase diagram with intermedi<strong>at</strong>e compound<br />

AB, where the intermedi<strong>at</strong>e region is too small to discern from a line.<br />

Figure 3.13: Impossible binary phase diagram with intermedi<strong>at</strong>e compound<br />

AB, where the intermedi<strong>at</strong>e region is too small to discern from a line.


128 CHAPTER 3. PHASE TRANSITIONS.<br />

give details we need to assume a certain form for the energy as a function of<br />

the st<strong>at</strong>e variables. Energy is the most basic thermodynamic st<strong>at</strong>e variable for<br />

which we can develop models th<strong>at</strong> can easily be interpreted. This will be done<br />

in the next chapter.<br />

More quantit<strong>at</strong>ive approaches based on equ<strong>at</strong>ions of st<strong>at</strong>e.<br />

In a p-V-T system the three st<strong>at</strong>e variables which are easily controlled or<br />

measured are the pressure p, the volume V, and the temper<strong>at</strong>ure T. We assume<br />

th<strong>at</strong> the number of particles N is again fixed, and hence of the three variables<br />

{p, V, T } only two are independent. A m<strong>at</strong>hem<strong>at</strong>ical rel<strong>at</strong>ion describing this<br />

dependency, an equ<strong>at</strong>ion of st<strong>at</strong>e, is needed to show how they are rel<strong>at</strong>ed. The<br />

ideal gas law, pV = NRT , is the easiest example of such an equ<strong>at</strong>ion of st<strong>at</strong>e,<br />

but unfortun<strong>at</strong>ely a system obeying the ideal gas law does not show any phase<br />

transitions. An ideal gas is an approxim<strong>at</strong>ion of a real gas, and describes a real<br />

gas in the limit of very low densities. Therefore, an <strong>at</strong>tempt to improve upon<br />

the ideal gas law would be to assume an equ<strong>at</strong>ion of st<strong>at</strong>e of the form<br />

p<br />

RT =<br />

∞∑<br />

( ) j<br />

N<br />

Bj(T )<br />

V<br />

j=1<br />

(3.18)<br />

with B1(T ) = 1. The value of this first term is determined by the requirement<br />

th<strong>at</strong> in the low density limit the ideal gas law should be obtained. This expansion<br />

is called a virial expansion and is the basis for many <strong>at</strong>tempts to describe a real<br />

gas by some kind of perturb<strong>at</strong>ion theory.<br />

Validity of equ<strong>at</strong>ions of st<strong>at</strong>e.<br />

In our discussion of phase transitions we mentioned th<strong>at</strong> every point in a p-T<br />

diagram corresponds to a st<strong>at</strong>e of the system, but th<strong>at</strong> this is not necessarily<br />

the case for every point in a p-V or T-V diagram. Equ<strong>at</strong>ion 3.18 yields a value<br />

of the pressure for all values of V and T. Hence equ<strong>at</strong>ion 3.18 in a technical<br />

sense only represents an equ<strong>at</strong>ion of st<strong>at</strong>e for those points in a T-V diagram<br />

which are equilibrium or stable st<strong>at</strong>es! The advantage of using an equ<strong>at</strong>ion like<br />

3.18 for all points is th<strong>at</strong> it gives an analytic function connecting all possible<br />

regions in phase space. It has to be supplemented by equilibrium conditions.<br />

The most obvious requirements ) are those rel<strong>at</strong>ed to local stability. If we have<br />

p(T,V) we need ≤ 0. This inequality holds for the ideal gas law for all<br />

( ∂p<br />

∂V<br />

T<br />

temper<strong>at</strong>ures and volumes. All points in a T-V diagram are stable st<strong>at</strong>es for an<br />

ideal gas, and there are no phase transitions. This is not the case, however, for<br />

more general equ<strong>at</strong>ions of st<strong>at</strong>e.<br />

Models of the equ<strong>at</strong>ion of st<strong>at</strong>e.


3.5. VAN DER WAALS EQUATION OF STATE. 129<br />

It is also possible to assume the existence of an equ<strong>at</strong>ion of st<strong>at</strong>e, rel<strong>at</strong>ing the<br />

st<strong>at</strong>e variables describing a system. Van der Waals analyzed such a situ<strong>at</strong>ion.<br />

He was interested in the phase transition from a gas to a liquid and derived a<br />

modified form of the ideal gas law which is valid for both the gas and the liquid<br />

phase. The fe<strong>at</strong>ures of his model equ<strong>at</strong>ion of st<strong>at</strong>e are remarkable general. The<br />

explan<strong>at</strong>ion for this fact is found in a microscopic deriv<strong>at</strong>ion of his equ<strong>at</strong>ion of<br />

st<strong>at</strong>e, which we will ignore <strong>at</strong> this point.<br />

Van der Waals followed a route different from the virial approach when he<br />

investig<strong>at</strong>ed liquid to gas phase transitions. He postul<strong>at</strong>ed th<strong>at</strong> for a not-so-ideal<br />

gas the equ<strong>at</strong>ion of st<strong>at</strong>e should have the form<br />

( ) 2 N<br />

p + a (V − Nb) = NRT (3.19)<br />

V 2<br />

where a and b are two parameters. It is possible to give a physical interpret<strong>at</strong>ion<br />

to these parameters a and b if the Van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is derived from<br />

a microscopic model. The parameter b represents the volume of a molecule. The<br />

total volume V has to be larger than the volume of all molecules tightly packed<br />

together. The internal energy of an ideal gas only depends on the temper<strong>at</strong>ure,<br />

and a correction term is needed for higher densities. The parameter a represents<br />

the average interaction energy between two molecules and is positive for an<br />

<strong>at</strong>tractive interaction.<br />

Unique solutions or not.<br />

is<br />

The pressure as a function of volume for the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e<br />

p = NRT<br />

− aN2<br />

V − Nb V 2<br />

(3.20)<br />

Therefore, the pressure is a unique function of the volume and temper<strong>at</strong>ure.<br />

The volume, on the other hand, is not a unique function of the pressure and<br />

temper<strong>at</strong>ure, and the van der Waals equ<strong>at</strong>ion allows for multiple solutions in<br />

certain ranges of pressure and temper<strong>at</strong>ure. Switching from one solution to<br />

another causes a jump in the value of the volume, exactly wh<strong>at</strong> we expect for a<br />

phase transition.<br />

It is also useful to discuss the difference between the pressure in a van der<br />

Waals gas and in an ideal gas. For this difference we find:<br />

2 N<br />

p − pideal =<br />

V 2<br />

(<br />

)<br />

V<br />

bRT − a<br />

(3.21)<br />

V − Nb<br />

which shows th<strong>at</strong> the difference is of second order in the density. This difference<br />

is useful if we want to calcul<strong>at</strong>e the energy difference between a van der Waals<br />

gas and an ideal gas, for example.<br />

Unstable regions.


130 CHAPTER 3. PHASE TRANSITIONS.<br />

( ∂p<br />

∂V<br />

Mechanical ) stability requires th<strong>at</strong> compressibilities are positive, and hence<br />

< 0. This deriv<strong>at</strong>ive for the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is<br />

T<br />

( )<br />

∂p<br />

= −<br />

∂V T,N<br />

NRT<br />

V 2<br />

(<br />

)<br />

1 n<br />

− 2a<br />

(1 − nb) 2 RT<br />

(3.22)<br />

where n = N<br />

V is the molar density of the m<strong>at</strong>erial in the gas or liquid. Instabilities<br />

occur when this deriv<strong>at</strong>ive is positive. This can only happen when a<br />

is positive. In a microscopic interpret<strong>at</strong>ion this means th<strong>at</strong> the interaction between<br />

the molecules has to be <strong>at</strong>tractive. This makes sense, for if the interaction<br />

between the molecules were repulsive, they would not want to condense into a<br />

liquid. The stability condition can be rewritten in the form:<br />

Phase space.<br />

RT<br />

2a ≥ (1 − nb)2 n (3.23)<br />

Phase space for a p-V-T system is given by the octant of three dimensional<br />

{p, V, T } space in which all three variables are positive. The van der Waals<br />

equ<strong>at</strong>ion of st<strong>at</strong>e limits the possible st<strong>at</strong>es of the system to a two-dimensional<br />

surface ) in this space. The lines on this surface where the partial deriv<strong>at</strong>ive<br />

is zero separ<strong>at</strong>es stable and unstable regions. Therefore it is of interest<br />

( ∂p<br />

∂V<br />

T<br />

to find these lines by solving<br />

n(1 − nb) 2 = RT<br />

2a<br />

(3.24)<br />

In order th<strong>at</strong> V > Nb we need n < 1<br />

b , hence we do not include the high density<br />

region in which the <strong>at</strong>oms would have to overlap in order to fit into the volume<br />

V.<br />

Case a=0.<br />

There are no solutions to this equ<strong>at</strong>ion when a = 0. In th<strong>at</strong> case the equ<strong>at</strong>ion<br />

of st<strong>at</strong>e is simply<br />

p = NRT<br />

V − Nb<br />

(3.25)<br />

which is well defined for all volumes V > Nb and has a neg<strong>at</strong>ive deriv<strong>at</strong>ive.<br />

Hence the effect of a finite volume of the molecules in itself is not sufficient to<br />

give rise to phase transitions.<br />

case b=0.


3.5. VAN DER WAALS EQUATION OF STATE. 131<br />

is<br />

Figure 3.14: Graphical solution of equ<strong>at</strong>ion 3.24.<br />

There is only one solution to 3.24 if b = 0. In th<strong>at</strong> case the equ<strong>at</strong>ion of st<strong>at</strong>e<br />

p = n(RT − an) (3.26)<br />

and stability requires n < RT<br />

2a . This system shows a sign of a phase transition.<br />

If we start with a gas <strong>at</strong> a high temper<strong>at</strong>ure, and cool it down, nothing special<br />

happens as long as T > 2an<br />

2an<br />

R . As soon as we reach the temper<strong>at</strong>ure R ,<br />

however, the system becomes mechanically unstable, and the system collapses<br />

to a singular point with V = 0. The <strong>at</strong>tractive interaction <strong>at</strong> low temper<strong>at</strong>ure<br />

is strong enough to force condens<strong>at</strong>ion into a liquid, but the liquid is unstable.<br />

We need the finite volume of the molecules to stabilize the liquid phase.<br />

General case.<br />

The general equ<strong>at</strong>ion 3.24 with both a and b non-zero has either one or three<br />

solutions. In the region where the molecules do not overlap we have n < 1<br />

b . For<br />

very large values of the temper<strong>at</strong>ure there are no instabilities for n < 1<br />

b and<br />

there are no phase transitions as a function of volume or density or pressure.<br />

If the temper<strong>at</strong>ure is very small, the solutions are approxim<strong>at</strong>ely n = RT<br />

2a and<br />

√<br />

RT<br />

nb = 1 ± 2a . Two solutions are in the allowed range nb < 1. If we increase<br />

the temper<strong>at</strong>ure, these two solutions will approach each other and <strong>at</strong> a certain<br />

temper<strong>at</strong>ure Tc they will merge together and disappear. The third solution will<br />

always be in the un-physical region n > 1<br />

b . This is represented in figure 3.14.<br />

If the temper<strong>at</strong>ure T is larger than Tc, all values of n in the range [0, 1<br />

b ]<br />

correspond to mechanically stable st<strong>at</strong>es. On the other hand, if T is below<br />

Tc the values of N in between the two crossings correspond to mechanically<br />

unstable st<strong>at</strong>es, and only low and high density phases are possible.


132 CHAPTER 3. PHASE TRANSITIONS.<br />

critical temper<strong>at</strong>ure.<br />

The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e shows no phase transitions for T > Tc<br />

and a region of instability for T < Tc. Therefore Tc is identified as the critical<br />

temper<strong>at</strong>ure, belonging to the critical point for this liquid-gas phase transition.<br />

The deriv<strong>at</strong>ive of the function<br />

is easy to find:<br />

f(n) = n(1 − nb) 2<br />

(3.27)<br />

f ′ (n) = (1 − nb) 2 − 2nb(1 − nb) = (1 − 3nb)(1 − nb) (3.28)<br />

and is zero for nb = 1 (as expected) and 3nb = 1. Therefore 3ncb = 1. This<br />

requires b > 0 as expected. The value of Tc follows from RTc<br />

2a = f(nc) = 4<br />

27b .<br />

The pressure follows directly from the van der Waals equ<strong>at</strong>ion.<br />

Loc<strong>at</strong>ion of critical point.<br />

The critical point is therefore specified by<br />

RTc = 8a<br />

27b<br />

(3.29)<br />

Vc = 3Nb (3.30)<br />

pc = a<br />

27b2 (3.31)<br />

The critical volume is indeed larger than Nb, as required. Its value, however,<br />

is close enough to the high density limit and a strict quantit<strong>at</strong>ive interpret<strong>at</strong>ion<br />

of the parameters a and b in terms of microscopic quantities in invalid. In such<br />

an interpret<strong>at</strong>ion higher order effects will certainly play a role. Therefore, it is<br />

better to consider a and b as adjustable parameters rel<strong>at</strong>ed to the interaction<br />

energy and molecular volume, and find their values from experiment. It is also<br />

clear th<strong>at</strong> the equ<strong>at</strong>ions for the critical point only make sense when a and b are<br />

positive.<br />

pV-curves.<br />

The general shape of the p-V curves is shown in figure 3.15. Note th<strong>at</strong> <strong>at</strong> the<br />

critical point not only the slope ) of the p(V) curve is zero, but also the curv<strong>at</strong>ure!<br />

In addition, we have<br />

> 0 and with increasing temper<strong>at</strong>ure<br />

( ∂p<br />

∂T<br />

V,N<br />

= NR<br />

V −Nb<br />

the curves move always up. This means th<strong>at</strong> the p-V-curves will never cross!<br />

Another interesting question pertains to the situ<strong>at</strong>ion depicted in figure 3.16.<br />

Is it possible to have neg<strong>at</strong>ive values of the pressure? Mechanical instability


3.5. VAN DER WAALS EQUATION OF STATE. 133<br />

Figure 3.15: p-V curves in the van der Waals model.<br />

Figure 3.16: p-V curves in the van der Waals model with neg<strong>at</strong>ive values of<br />

the pressure.


134 CHAPTER 3. PHASE TRANSITIONS.<br />

only requires th<strong>at</strong> the slope of the p(V) curve is neg<strong>at</strong>ive. It does not preclude<br />

a neg<strong>at</strong>ive value of the pressure. Neg<strong>at</strong>ive pressure means th<strong>at</strong> the system gains<br />

energy by expanding. One could simul<strong>at</strong>e this by connecting springs to the<br />

outside of a m<strong>at</strong>erial, which try to stretch the m<strong>at</strong>erial. Also, if we replace p-V<br />

by H-M there is no problem <strong>at</strong> all, neg<strong>at</strong>ive magnetic fields are possible!<br />

Do you expect neg<strong>at</strong>ive pressures in real liquids?<br />

Neg<strong>at</strong>ive pressures occur in the van der Waals model as soon as T < 27<br />

32 Tc,<br />

which means for a large range of values of the temper<strong>at</strong>ure. The equ<strong>at</strong>ion p = 0<br />

leads to<br />

an(1 − nb) = RT (3.32)<br />

with solutions<br />

√<br />

1 ± 1 − 4RT<br />

n = b<br />

a<br />

(3.33)<br />

2b<br />

which are always less than 1<br />

b and hence always occur in the physically allowed<br />

region. The square root has to be real, though, which leads to 4RT b < a, which<br />

gives the inequality for the temper<strong>at</strong>ure cited above.<br />

Finding the minimal energy st<strong>at</strong>e.<br />

The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e for a temper<strong>at</strong>ure below the critical<br />

value leads to a p(V) curve of the form shown in figure 3.17. If the value of<br />

the pressure is between pmin and pmax, there are three values of the volume<br />

possible. Since the volume is a st<strong>at</strong>e variable, only one of these values can be<br />

correct. Obviously we have to discard the middle value of the volume, since the<br />

deriv<strong>at</strong>ive of the p(V) curve has the wrong sign in this case. The Gibbs free<br />

energy is needed in order to decide which of the other two solutions is stable.<br />

How to calcul<strong>at</strong>e the energy.<br />

The difference in Gibbs free energy between the point 1 and 2 is found the<br />

change in energy along the p<strong>at</strong>h following the equ<strong>at</strong>ion of st<strong>at</strong>e p(V) <strong>at</strong> this<br />

temper<strong>at</strong>ure. This is not a reversible process, since some of the intermedi<strong>at</strong>e<br />

st<strong>at</strong>es are unstable. These intermedi<strong>at</strong>e st<strong>at</strong>es are special, however, since they<br />

are m<strong>at</strong>hem<strong>at</strong>ically well described. The equ<strong>at</strong>ion of st<strong>at</strong>e is still valid for these<br />

points. At constant temper<strong>at</strong>ure and particle number the differential of the<br />

Gibbs free energy is dG = V dp and therefore<br />

G2 − G1 =<br />

∫ 2<br />

1<br />

V dp = A1 − A2<br />

(3.34)


3.5. VAN DER WAALS EQUATION OF STATE. 135<br />

Figure 3.17: p-V curve in the van der Waals model with areas corresponding<br />

to energies.<br />

where A1 and A2 are the indic<strong>at</strong>ed areas on the left and right. If A1 > A2 the<br />

Gibbs free energy of the system in st<strong>at</strong>e 2 is the highest and the system will be<br />

in st<strong>at</strong>e 1, with a low volume. The opposite is true for A2 > A1. If A1 = A2<br />

the Gibbs free energies of both phases are equal and the liquid and gas coexist.<br />

This condition therefore determines pco(T ).<br />

Is this procedure allowed?<br />

One question th<strong>at</strong> might be asked is if the previous procedure is allowed.<br />

We have followed the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e through unstable regions,<br />

which does not seem to be correct. Fortun<strong>at</strong>ely, as expected, the answer is<br />

reassuring, we can do this. At each point along the van der Waals equ<strong>at</strong>ion<br />

of st<strong>at</strong>e the st<strong>at</strong>e of the system is an extremal st<strong>at</strong>e for the energy, either a<br />

maximum or a local or global minimum. If we follow the van der Waals equ<strong>at</strong>ion<br />

of st<strong>at</strong>e in a reversible process, in theory we can always go back and forth and<br />

the energies calcul<strong>at</strong>ed before correspond to the actual energy differences. Of<br />

course, in practice there would always be small fluctu<strong>at</strong>ions th<strong>at</strong> throw us out<br />

of the unstable st<strong>at</strong>e when the energy is a maximum. Hence the procedure we<br />

followed is not a practical experimental procedure to measure energy differences.<br />

Neg<strong>at</strong>ive pressure never correspond to physical st<strong>at</strong>es.<br />

The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e allows for neg<strong>at</strong>ive pressures. The equ<strong>at</strong>ion<br />

p = 0 has two solutions when T < 27<br />

32Tc. The pressure <strong>at</strong> very large volumes<br />

always approaches zero inversely proportional to V. The stable st<strong>at</strong>e for a small,<br />

positive pressure is determined by the equality of the areas A1 and A2. In the<br />

limit for the pressure approaching zero, the area on the right A2 becomes in-


136 CHAPTER 3. PHASE TRANSITIONS.<br />

finitely large. Therefore, we always have A2 > A1 in this limit, and the system<br />

will be found <strong>at</strong> the high volume st<strong>at</strong>e. Hence, the st<strong>at</strong>es with neg<strong>at</strong>ive pressure<br />

are never stable, even though one of them is mechanically stable. Mechanical<br />

stability only involves the deriv<strong>at</strong>ive of pressure versus volume. If a system is<br />

stable with a neg<strong>at</strong>ive pressure, it behaves like an expanded rubber band. We<br />

need to stretch it (neg<strong>at</strong>ive pressure) in order to keep it in th<strong>at</strong> st<strong>at</strong>e. Stable<br />

systems with a neg<strong>at</strong>ive pressure are possible on general thermodynamical<br />

grounds. The specific physics of the system under consider<strong>at</strong>ion will decide if<br />

th<strong>at</strong> system can be stable under neg<strong>at</strong>ive pressures. A liquid-gas system should<br />

not be stable under neg<strong>at</strong>ive pressure, since cooling through forced evapor<strong>at</strong>ion<br />

is possible. The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is consistent with the experimental<br />

facts for a liquid-gas system and does not allow for stable st<strong>at</strong>es with<br />

a neg<strong>at</strong>ive pressure. All in all we conclude th<strong>at</strong> the van der Waals equ<strong>at</strong>ion of<br />

st<strong>at</strong>e is a good model for a liquid to gas phase transition, ending in a critical<br />

point.<br />

Joule-Thomson effect.<br />

If a fluid is pushed through a porous plug, this throttling process gives a<br />

change in temper<strong>at</strong>ure due to the change in pressure <strong>at</strong> constant enthalpy. This<br />

was a problem assigned in the previous chapter. The Joule-Thomson coefficient<br />

describing this process is therefore<br />

( )<br />

∂T<br />

ηJT =<br />

(3.35)<br />

∂p H,N<br />

and can be shown to be equal to<br />

ηJT = V<br />

Cp<br />

(αT − 1) (3.36)<br />

For m<strong>at</strong>erials with αT > 1 the fluid will cool down on the expansion through<br />

the plug, and for m<strong>at</strong>erials with αT < 1 the fluid will he<strong>at</strong> up on this expansion.<br />

For an ideal gas we have αT = 1, so no change in temper<strong>at</strong>ure occurs. But wh<strong>at</strong><br />

is happening in a real gas?<br />

The van der Waals equ<strong>at</strong>ion is a good approxim<strong>at</strong>ion for an ideal gas, so<br />

we will calcul<strong>at</strong>e the coefficient of thermal expansion for the van der Waals gas.<br />

Th<strong>at</strong> is a bit tedious, so we will look <strong>at</strong> two extremes. First of all, wh<strong>at</strong> happens<br />

<strong>at</strong> very high temper<strong>at</strong>ures? In th<strong>at</strong> case the pressure will be very high for a<br />

given volume, and hence we can ignore the a ( )<br />

N 2<br />

V term and use:<br />

which gives<br />

p(V − Nb) ≈ NRT (3.37)<br />

( )<br />

∂V<br />

p ≈ NR (3.38)<br />

∂T p,N


3.5. VAN DER WAALS EQUATION OF STATE. 137<br />

and hence<br />

αT − 1 ≈ NRT<br />

pV<br />

pα ≈ NR<br />

V<br />

− 1 ≈ p(V − Nb)<br />

pV<br />

− 1 = −b N<br />

V<br />

(3.39)<br />

(3.40)<br />

So <strong>at</strong> high temper<strong>at</strong>ures the Joule-Thomson coefficient for a real gas is neg<strong>at</strong>ive,<br />

which means th<strong>at</strong> the gas cools down in a throttling process. But wh<strong>at</strong><br />

happens a a lower temper<strong>at</strong>ure? We will study this in the low density limit,<br />

where we have<br />

p = NRT<br />

V − Nb −a<br />

This gives<br />

0 = NR<br />

V<br />

− NRT<br />

V 2<br />

( ) 2<br />

N<br />

V<br />

0 = NR<br />

V<br />

≈ N<br />

RT (1+bN<br />

V V )−a<br />

( )<br />

∂V<br />

+ bR<br />

∂T p,N<br />

− NRT<br />

V<br />

α + bR<br />

( ) 2<br />

N<br />

V<br />

( ) 2<br />

N<br />

V<br />

( ) 2<br />

N<br />

=<br />

V<br />

NRT<br />

+(bRT −a)<br />

V<br />

2 N<br />

− 2(bRT − a)<br />

V 3<br />

( )<br />

∂V<br />

∂T p,N<br />

− 2(bRT − a)<br />

RT α + 2(bRT − a) N<br />

N<br />

α = R + bR<br />

V V<br />

α =<br />

αT − 1 =<br />

αT − 1 = N<br />

V<br />

R + bR N<br />

V<br />

RT + 2(bRT − a) N<br />

V<br />

bRT N<br />

V<br />

− 2(bRT − a) N<br />

V<br />

RT + 2(bRT − a) N<br />

V<br />

2a − bRT<br />

RT + 2(bRT − a) N<br />

V<br />

N 2<br />

( ) 2<br />

N<br />

V<br />

(3.41)<br />

(3.42)<br />

α (3.43)<br />

V 2<br />

(3.44)<br />

(3.45)<br />

(3.46)<br />

(3.47)<br />

This shows th<strong>at</strong> the Joule-Thomson effect changes sign when we lower the<br />

temper<strong>at</strong>ure. The effect is zero <strong>at</strong> the so-called inversion temper<strong>at</strong>ure Ti and<br />

for temper<strong>at</strong>ures below Ti a gas he<strong>at</strong>s up in a throttling process. We see th<strong>at</strong><br />

the inversion temper<strong>at</strong>ure is given by<br />

Ti = 2a 27<br />

=<br />

bR 4 Tc<br />

(3.48)<br />

which is a temper<strong>at</strong>ure well above the critical temper<strong>at</strong>ure! Hence for almost<br />

all gases <strong>at</strong> room temper<strong>at</strong>ure the Joule-Thomson effect is positive.


138 CHAPTER 3. PHASE TRANSITIONS.<br />

Figure 3.18: Unstable and meta-stable regions in the van der Waals p-V<br />

diagram.<br />

3.6 Spinodal decomposition.<br />

Names of separ<strong>at</strong>ion lines.<br />

The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e determines the temper<strong>at</strong>ure as a function<br />

of pressure and volume for every point (st<strong>at</strong>e) in the p-V plane. As we<br />

have seen, ) however, not all st<strong>at</strong>es are stable. The region in the p-V plane<br />

where<br />

> 0 corresponds to mechanically unstable st<strong>at</strong>es. The line<br />

( ∂p<br />

∂V<br />

T,N<br />

which separ<strong>at</strong>es this region from the region of mechanically stable (but thermodynamically<br />

meta-stable) st<strong>at</strong>es is called the spinodal and is determined by<br />

)<br />

(p, V, N) = 0. One has to keep in mind th<strong>at</strong> N is fixed. The line sepa-<br />

( ∂p<br />

∂V<br />

T,N<br />

r<strong>at</strong>ing the thermodynamically stable and meta-stable st<strong>at</strong>es is called a binodal<br />

and is determined by the condition th<strong>at</strong> the Gibbs free energy be the same in<br />

both phases.<br />

Meta-stable regions.<br />

Wh<strong>at</strong> happens when we prepare a system in the mechanically unstable region?<br />

This could be done, for example, by cooling a system with volume Vc<br />

rapidly to a temper<strong>at</strong>ure below Tc. Assume th<strong>at</strong> the pressure for this un-physical<br />

system is p0, and th<strong>at</strong> the system is kept <strong>at</strong> a constant temper<strong>at</strong>ure T0. At first,<br />

the system will be homogeneous. Since it is mechanically unstable, any small<br />

fluctu<strong>at</strong>ion in the density will be amplified. A region which momentarily has a<br />

density which is too high, will increase its density following the van der Waals<br />

curve until it reaches the spinodal, where it becomes mechanically stable. A


3.6. SPINODAL DECOMPOSITION. 139<br />

low density region will follow the van der Waals curve in the opposite direction.<br />

The system will separ<strong>at</strong>e in small regions of high density, N , and small regions<br />

Va<br />

with low density N<br />

Vb . The values of the volumes Va and Vb are found from the<br />

curve in figure 3.18. These are the values of the volume for the whole system, so<br />

they have to be transformed to densities. This regions are all mixed together.<br />

The total volume has to be constant, and hence<br />

Vtotal = Na<br />

N Va + Nb<br />

N Vb<br />

(3.49)<br />

where Vtotal is the initial volume. Also we have N = Na + Nb. These two<br />

rel<strong>at</strong>ions together determine how much m<strong>at</strong>erial will go into each phase.<br />

Spinodal decomposition.<br />

The process described in the previous section is an important one and is<br />

called spinodal decomposition. The end st<strong>at</strong>e is mechanically stable. It is thermodynamically<br />

unstable, though, since pa < pb. Hence the process will continue<br />

by a further expansion of the low density regions and a further contraction of<br />

the high density regions until the pressure in both regions is equal again (and<br />

equal to p0). During this process we again follow a van der Waals curve. But<br />

now time scales become important.<br />

Spinodal decomposition in a liquid-gas system.<br />

If the spinodal decomposition occurred in a gas-liquid mixture, the road to<br />

thermal equilibrium is without obstacles and the final system will be a mixture<br />

of gas and liquid <strong>at</strong> the same pressure p0 and temper<strong>at</strong>ure T0 (corresponding to<br />

points 1 and 2 in figure 3.18). The amount of each m<strong>at</strong>erial is determined by the<br />

original volume of the gas. In free space, this is exactly wh<strong>at</strong> happens. Clouds<br />

and fog are such mixtures of gas and liquid. One way to do an experiment along<br />

these lines is to take a liquid in a cylinder with a piston and suddenly pull the<br />

piston out.<br />

Complic<strong>at</strong>ing factors.<br />

In many experiments gravity plays a role and will try to collect all the liquid<br />

in the bottom of the container. Surface tension is another factor, and has to be<br />

taken into account to find the distribution of the sizes of the liquid droplets in a<br />

fog. Note th<strong>at</strong> it is possible for a system to be in the meta-stable region. W<strong>at</strong>er<br />

can be under-cooled well below 0 C. Any small perturb<strong>at</strong>ion (like an abrupt<br />

movement of the container) will cause it to freeze, however, demonstr<strong>at</strong>ing the<br />

meta-stability.<br />

Example in binary phase diagram.


140 CHAPTER 3. PHASE TRANSITIONS.<br />

In a solid compound AxB1−x the composition x is a st<strong>at</strong>e variable and a<br />

spinodal curve might show up in the T-x plane. Certain compositions are mechanically<br />

unstable in th<strong>at</strong> case, and if we quench a liquid of th<strong>at</strong> combin<strong>at</strong>ion<br />

to a low temper<strong>at</strong>ure, the quenched solid is mechanically unstable. It will phase<br />

separ<strong>at</strong>e following a spinodal decomposition and become a mixture of grains of<br />

the compound with low x and high x. The intensive st<strong>at</strong>e variable corresponding<br />

to the concentr<strong>at</strong>ion x is the chemical potential. Therefore, we have cre<strong>at</strong>ed differences<br />

in chemical potential between the grains, and this will force m<strong>at</strong>erial to<br />

move, so we can follow the equivalent of a p-V curve to arrive <strong>at</strong> the stable st<strong>at</strong>es<br />

for this temper<strong>at</strong>ure. In order to do so, however, m<strong>at</strong>erial needs to diffuse in a<br />

solid, which is a very slow process. When the system has become mechanically<br />

stable, it might try to find other ways to rearrange and become thermodynamically<br />

stable. In a solid, however, the grains of each m<strong>at</strong>erial cannot be deformed<br />

randomly and the elastic properties of the m<strong>at</strong>erial will determine how close the<br />

system is able to approach thermodynamical equilibrium. In most cases, large<br />

internal stresses will persist, which has lots of implic<strong>at</strong>ions for the strength of<br />

the m<strong>at</strong>erial.<br />

Wh<strong>at</strong> are the consequences for the measurement of phase diagrams?<br />

If we cool a binary alloy, with a concentr<strong>at</strong>ion corresponding to a stable st<strong>at</strong>e,<br />

from the liquid, do we get a homogeneous solid?<br />

An energy view of the decomposition<br />

If we follow the Helmholtz free energy as a function of volume <strong>at</strong> a given<br />

temper<strong>at</strong>ure below the critical temper<strong>at</strong>ure, we find a behavior as sketched in<br />

figure 3.19. The slope of the energy curve <strong>at</strong> each point gives us the pressure<br />

<strong>at</strong> th<strong>at</strong> st<strong>at</strong>e (with a minus sign). The Helmholtz energy is calcul<strong>at</strong>ed by the<br />

following integr<strong>at</strong>ion along the equ<strong>at</strong>ion of st<strong>at</strong>e:<br />

∫ V<br />

F (V, T, N) = F0(T, N) −<br />

V0<br />

pdV (3.50)<br />

where the integr<strong>at</strong>ion variable is now V (because we switched to the Helmholtz<br />

free energy).<br />

Suppose we are <strong>at</strong> a volume Vi in between points 1 and 2. In th<strong>at</strong> case the<br />

Helmholtz free energy obtained by following the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e<br />

is larger th<strong>at</strong> the value obtained by drawing a straight line between 1 and 2.<br />

Wh<strong>at</strong> does the l<strong>at</strong>ter mean? It corresponds to cre<strong>at</strong>ing a mixture of m<strong>at</strong>erials


3.6. SPINODAL DECOMPOSITION. 141<br />

F<br />

1<br />

V i<br />

Figure 3.19: Energy versus volume showing th<strong>at</strong> decomposition lowers the<br />

energy.<br />

corresponding to the st<strong>at</strong>es <strong>at</strong> points 1 and 2. The equ<strong>at</strong>ion of the straight line<br />

connecting 1 and 2 is<br />

Fs(V ) =<br />

1<br />

V2 − V1<br />

2<br />

V<br />

((V2 − V )F1 + (V − V1)F2) (3.51)<br />

which corresponds to a fraction x1 = V2−V<br />

V2−V1 of m<strong>at</strong>erial in st<strong>at</strong>e 1, and x2 =<br />

V −V1<br />

V2−V1<br />

in st<strong>at</strong>e 2. It is easy to check th<strong>at</strong> this indeed adds up to the correct<br />

volume, x1V1 + x2V2 = V . Therefore, in any point between 1 and 2 we can<br />

lower the energy of the m<strong>at</strong>erial by phase separ<strong>at</strong>ion.<br />

How do we find the loc<strong>at</strong>ion of points 1 and 2? They need to have the same<br />

slope, hence<br />

( ) ( )<br />

∂F<br />

∂F<br />

(V1) =<br />

(V2) = −pcoex<br />

∂V T,N ∂V T,N<br />

and this slope should be the slope of the connecting line:<br />

which leads to<br />

( )<br />

∂F<br />

(V1) =<br />

∂V T,N<br />

F2 − F1<br />

V2 − V1<br />

pcoex(V2 − V1) = −(F2 − F1) =<br />

∫ 2<br />

1<br />

(3.52)<br />

(3.53)<br />

pdV (3.54)<br />

which is an altern<strong>at</strong>ive formul<strong>at</strong>ion of the st<strong>at</strong>ement th<strong>at</strong> the areas are equal.<br />

This approach is due to Maxwell, and the graphical approach in figure 3.19 is<br />

called the Maxwell construction.


142 CHAPTER 3. PHASE TRANSITIONS.<br />

3.7 Generaliz<strong>at</strong>ions of the van der Waals equ<strong>at</strong>ion.<br />

The van der Waals equ<strong>at</strong>ion takes a very simple form when expressed in values<br />

of the st<strong>at</strong>e variables rel<strong>at</strong>ive to the critical values. Define<br />

P = p<br />

, V = V<br />

, T = T<br />

The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e in these variables is<br />

(<br />

P + 3<br />

V2 ) (<br />

V − 1<br />

)<br />

=<br />

3<br />

8T<br />

3<br />

pc<br />

Vc<br />

Tc<br />

(3.55)<br />

(3.56)<br />

which is independent of pc, Vc, and Tc! In rel<strong>at</strong>ive st<strong>at</strong>e variables the van der<br />

Waals equ<strong>at</strong>ion does not contain any inform<strong>at</strong>ion about the specific system and<br />

in these units the binodal coexistence curve T (V) should be the same for all<br />

m<strong>at</strong>erials. This has indeed been observed for a number of different substances<br />

like inert gases. The binodal curves for these substances in rel<strong>at</strong>ive units are<br />

almost identical. This common binodal curve differs somewh<strong>at</strong> from th<strong>at</strong> predicted<br />

by the van der Waals equ<strong>at</strong>ion, however, and corrections to the higher<br />

order virial coefficients are needed. Nevertheless, ignoring these rel<strong>at</strong>ively small<br />

differences, it is amazing how well the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e describes<br />

a liquid to gas phase transition, and predicts the universality of the spinodal<br />

curve. In order to understand the reasons for this success, one has to look <strong>at</strong><br />

the microscopic found<strong>at</strong>ion of the van der Waals equ<strong>at</strong>ion. Also, the notion<br />

th<strong>at</strong> the equ<strong>at</strong>ions of st<strong>at</strong>e in scaled units are the same for similar m<strong>at</strong>erials<br />

introduces for the first time the idea of universality. If the same physics is <strong>at</strong><br />

work in different m<strong>at</strong>erials, it is possible to scale experimental results from one<br />

to the other.<br />

Mixtures.<br />

The van der Waals equ<strong>at</strong>ion can also be extended to mixtures. In this case<br />

it loses many of its simple characteristics. For a gas mixture of M species, the<br />

excluded volumes are bi and the average interactions are aij. The number of<br />

<strong>at</strong>oms of type i is Ni. The simplest generaliz<strong>at</strong>ion of the van der Waals equ<strong>at</strong>ion<br />

for this case is<br />

⎛<br />

⎝p + 1<br />

V 2<br />

⎞ (<br />

∑<br />

M∑<br />

)<br />

aijNiNj ⎠ V − Nibi = NRT (3.57)<br />

i


3.8. EXTRA QUESTIONS. 143<br />

where the functions a and b only depend on the composition. As is turns out,<br />

this generaliz<strong>at</strong>ion is not correct from a microscopic point of view. The change<br />

in the first factor can be explained in terms of particle-particle interactions.<br />

Although the modific<strong>at</strong>ion of the second factor intuitively seems correct, a st<strong>at</strong>istical<br />

mechanical approach yields a more complic<strong>at</strong>ed expression.<br />

3.8 Extra questions.<br />

1. Wh<strong>at</strong> is the most important phase transition?<br />

2. Can we have a solid to gas transition?<br />

3. Why is G continuous?<br />

4. Why does V drop with p?<br />

5. Can V be continuous?<br />

6. Should V drop with T?<br />

7. Consider a container with w<strong>at</strong>er and vapor. Increase T. Wh<strong>at</strong> is formed?<br />

8. Wh<strong>at</strong> happens <strong>at</strong> the critical point in an experiment?<br />

9. Suppose we have two phases coexisting with p1 = p2, T1 = T2, but µ1 ̸= µ2.<br />

Wh<strong>at</strong> happens?<br />

10. Why is it easiest to discuss phase boundaries using the Gibbs free energy?<br />

11. Wh<strong>at</strong> can you tell about the slope of the coexistence curve?<br />

12. Could ice 1 and ice 2 meet by accident <strong>at</strong> the triple point?<br />

13. Apply Clausius-Clapeyron to a superconducting phase transition in B field.<br />

14. Wh<strong>at</strong> are the consequences for the slope of pco <strong>at</strong> the w<strong>at</strong>er-ice boundary?<br />

15. How do you measure phase diagrams?<br />

16. Construct phase diagram with a compound forming <strong>at</strong> low T.<br />

17. Wh<strong>at</strong> determines the curv<strong>at</strong>ure of the lines in a binary phase diagram?<br />

18. Do perfect compounds exist <strong>at</strong> low temper<strong>at</strong>ure?<br />

19. Wh<strong>at</strong> does universality of the van der Waals equ<strong>at</strong>ion mean?<br />

20. Discuss the meaning of the sign of the second virial coefficient.<br />

21. Show th<strong>at</strong> ηJT = − N<br />

Cp B2 for constant B2<br />

22. Give an example of a system which has a neg<strong>at</strong>ive pressure.


144 CHAPTER 3. PHASE TRANSITIONS.<br />

23. How did we get the unstable region in the forbidden region?<br />

24. Why can we measure the spinodal curve?<br />

25. Can the curves p(V, T1) and p(V, T2) cross?<br />

26. Explain equal areas in terms of work done.<br />

27. Why do we need the Gibbs free energy to decide in the vdW equ<strong>at</strong>ion, but<br />

use the Helmholz free energy in the Maxwell construction?<br />

3.9 Problems for chapter 3<br />

Problem 1.<br />

The equ<strong>at</strong>ion of st<strong>at</strong>e for some m<strong>at</strong>erial is given by the first two terms of a<br />

virial expansion,<br />

p<br />

RT<br />

= N<br />

V<br />

(<br />

1 + B2(T ) N<br />

)<br />

V<br />

(3.59)<br />

The second virial coefficient B2(T ) has a very simple form, B2(T ) = R (T −T0).<br />

p0<br />

(a) Make a picture of the isotherms in a p-V diagram.<br />

(b) Sketch the region of points (p,V) which correspond to unstable st<strong>at</strong>es.<br />

(c) Sketch the region of unstable st<strong>at</strong>es in the V-T diagram.<br />

(d) Are there any meta-stable st<strong>at</strong>es?<br />

Problem 2.<br />

The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is not the only form proposed to study<br />

phase transitions. Dietrici investig<strong>at</strong>ed the following expression:<br />

aN −<br />

p(V − Nb) = NRT e RT V (3.60)<br />

(a) Show th<strong>at</strong> in the limit V ≫ N this equ<strong>at</strong>ion of st<strong>at</strong>e becomes equivalent<br />

to the ideal gas law.<br />

(b) Calcul<strong>at</strong>e Vc, Tc, and pc for this equ<strong>at</strong>ion of st<strong>at</strong>e.<br />

(c) Sketch the isotherms in a p-V plane for T > Tc, T = Tc, and T < Tc.


3.9. PROBLEMS FOR CHAPTER 3 145<br />

Problem 3.<br />

Drops of a liquid are in equilibrium with a gas of the same m<strong>at</strong>erial. This<br />

experiment is done in an environment where there is no gravity, so the drops<br />

flo<strong>at</strong> freely in the vapor. The drops are spherical because of surface tension.<br />

For a given drop, the change in energy dU is given by<br />

where A is the surface area of the drop.<br />

dU = T dS − pdV + µdN + σdA (3.61)<br />

(a) Show th<strong>at</strong> for spherical drops this rel<strong>at</strong>ion can be written in the form<br />

dU = T dS − peff dV + µdN (3.62)<br />

(b) How is the pressure inside this spherical drop rel<strong>at</strong>ed to the pressure in<br />

the vapor?<br />

The chemical potential of the liquid is µl(p, T ) and for the gas µg(p, T ).<br />

Assume th<strong>at</strong> both gas and liquid have the same temper<strong>at</strong>ure T.<br />

(c) Wh<strong>at</strong> is the equilibrium condition which determines th<strong>at</strong> in equilibrium<br />

there is no net flow of particles between the liquid and the vapor?<br />

(d) Generalize the Clausius-Clapeyron rel<strong>at</strong>ion for a system of vapor and<br />

spherical droplets, including surface tension.<br />

(e) Show th<strong>at</strong> the coexistence pressure of the vapor <strong>at</strong> a given temper<strong>at</strong>ure<br />

increases when the average radius of the droplets decreases.<br />

Problem 4.<br />

A closed container contains w<strong>at</strong>er and helium gas. Assume th<strong>at</strong> helium does<br />

not dissolve in w<strong>at</strong>er.<br />

(a) Show th<strong>at</strong> <strong>at</strong> a given pressure there is <strong>at</strong> most one temper<strong>at</strong>ure <strong>at</strong> which<br />

w<strong>at</strong>er, ice, and w<strong>at</strong>er vapor coexist.<br />

(b) The triple point for w<strong>at</strong>er is pt, Tt. Calcul<strong>at</strong>e the change in these values<br />

as a function of the Helium pressure. Assume th<strong>at</strong> the chemical potential<br />

of w<strong>at</strong>er is independent of pressure<br />

Problem 5.<br />

Show th<strong>at</strong> the l<strong>at</strong>ent he<strong>at</strong> per mole of the liquid to vapor phase transition<br />

using the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is given by


146 CHAPTER 3. PHASE TRANSITIONS.<br />

Problem 6.<br />

l1→2(T ) = RT ln<br />

( )<br />

V2 − Nb<br />

V1 − Nb<br />

(3.63)<br />

(A) Calcul<strong>at</strong>e the difference in Helmholtz free energy between a van der Waals<br />

gas and an ideal gas using F = −pdV <strong>at</strong> constant temper<strong>at</strong>ure and number<br />

of moles of m<strong>at</strong>erial. Assume th<strong>at</strong> this difference is zero in the limit<br />

V → ∞.<br />

(B) Calcul<strong>at</strong>e the difference in entropy between a van der Waals gas and an<br />

ideal gas <strong>at</strong> constant temper<strong>at</strong>ure and number of moles of m<strong>at</strong>erial. Assume<br />

th<strong>at</strong> this difference is zero in the limit V → ∞.<br />

(C) Show th<strong>at</strong> the constant ”a” in the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e leads<br />

2<br />

N to a term −a V 2 in the internal energy of a van der Waals gas. Interpret<br />

this term in a molecular picture.<br />

Problem 7.<br />

(A) For values of the temper<strong>at</strong>ure below the critical temper<strong>at</strong>ure find an upper<br />

and a lower limit for the transition pressure in a van der Waals gas.<br />

(B) For values of the temper<strong>at</strong>ure below the critical temper<strong>at</strong>ure and values<br />

of the pressure in between these two limits, the van der Waals equ<strong>at</strong>ion<br />

of st<strong>at</strong>e has two mechanically stable solutions with volumes Vl and Vh.<br />

Calcul<strong>at</strong>e the difference in Gibbs free energy between these two st<strong>at</strong>es.<br />

(C) Outline a procedure for calcul<strong>at</strong>ing the coexistence curve pco(T ) numerically.<br />

Problem 8.<br />

The boiling point of Hg is 357C <strong>at</strong> 1 <strong>at</strong>m. and 399C <strong>at</strong> 2 <strong>at</strong>m. The density<br />

of liquid mercury is 13.6 g/cm 3 . The he<strong>at</strong> of vaporiz<strong>at</strong>ion is 270 J/g. Calcul<strong>at</strong>e<br />

the order of magnitude of the volume of 200g Hg vapor <strong>at</strong> 357C and 1 <strong>at</strong>m.<br />

Note, 1 <strong>at</strong>m. is about 10 5 N/m 2 , 0C=273K, and G=U-TS+pV.<br />

Problem 9.<br />

For a certain m<strong>at</strong>erial the gas and liquid phase coexist even <strong>at</strong> 0K. Show<br />

th<strong>at</strong> the third law implies th<strong>at</strong> dpco<br />

dT = 0 <strong>at</strong> T=0.


3.9. PROBLEMS FOR CHAPTER 3 147<br />

Problem 10.<br />

A closed container of volume V0 contains N0 moles of m<strong>at</strong>erial, partly liquid,<br />

partly gas, in equilibrium. The energy is constant, the container is thermally<br />

isol<strong>at</strong>ed. The gas is ideal, with pV = NRT . The Gibbs free energy of the liquid<br />

is Gl(p, T, N) = Nµl(p, T ). The unknown variables are (1) the number of moles<br />

Ng in the gas phase, (2) the number of moles Nl in the liquid phase, (3) the<br />

volume Vg of the gas phase, and (4) the volume Vl of the liquid phase.<br />

(A) Why are the pressure and temper<strong>at</strong>ure in the gas phase the same as in<br />

the liquid phase?<br />

(B) Specify four equ<strong>at</strong>ions which we need to solve to find the four unknown<br />

quantities, given th<strong>at</strong> the pressure in the container is p and the temper<strong>at</strong>ure<br />

is T.<br />

(C) Solve this system of equ<strong>at</strong>ions for Nl<br />

(D) Is there a solution for all values of p and T?<br />

Problem 11.<br />

In the vicinity of the triple point of a m<strong>at</strong>erial the vapor pressure of the<br />

liquid is given by log( p<br />

20T0<br />

) = 10 − p0 T . The vapor pressure of the solid is given<br />

by log( p<br />

25T0<br />

) = 15 − p0 T .<br />

(A) Wh<strong>at</strong> are the temper<strong>at</strong>ure and the pressure of the triple point in terms of<br />

T0 and p0?<br />

(B) Using the Clausius-Clapeyron equ<strong>at</strong>ion, show th<strong>at</strong><br />

d log(pco)<br />

dT = l<br />

pcoT ∆v .<br />

(C) Is the volume per mole of the gas larger or smaller than th<strong>at</strong> of the liquid,<br />

assuming th<strong>at</strong> T0 and p0 are both positive?<br />

(D) Assume th<strong>at</strong> the values of the l<strong>at</strong>ent he<strong>at</strong> are independent of temper<strong>at</strong>ure<br />

near the triple point, and th<strong>at</strong> the molar volume of the gas is much larger<br />

than the molar volume of the liquid and of the solid. Calcul<strong>at</strong>e the r<strong>at</strong>ios<br />

lSL : lLG : lSG, where S,L, and G stand for solid, liquid, and gas.<br />

Problem 12.<br />

Grüneisen parameter.


148 CHAPTER 3. PHASE TRANSITIONS.<br />

In many solids we can approxim<strong>at</strong>e the entropy by S = Nf(T/TD) where<br />

the Debye temper<strong>at</strong>ure TD is a function of the molar volume only, TD = g( V<br />

N ).<br />

The Grüneisen parameter γg is defined by<br />

Show th<strong>at</strong><br />

γg = − V<br />

NTD<br />

γg =<br />

V α<br />

CV κT<br />

dTD<br />

d V<br />

N<br />

For many solids the Grüneisen parameter is almost constant. If we assume<br />

th<strong>at</strong> it is a constant, show th<strong>at</strong> we have the equ<strong>at</strong>ion of st<strong>at</strong>e<br />

P V = γgU + Nh( V<br />

N )<br />

This is the Mie-Grüneisen equ<strong>at</strong>ion of st<strong>at</strong>e.<br />

Problem 13.<br />

Murnaghan equ<strong>at</strong>ion.<br />

The isothermal bulk modulus BT is given by BT = −V<br />

( )<br />

∂p<br />

∂V<br />

T<br />

= 1 . The<br />

κT<br />

bulk modulus can often be approxim<strong>at</strong>ed by BT (p, T ) = BT (0, T ) + βT p where<br />

βT is independent of pressure.<br />

Show th<strong>at</strong> this leads to the equ<strong>at</strong>ion of st<strong>at</strong>e<br />

p(V, T ) = BT<br />

( [V0<br />

(0, T )<br />

V<br />

βT<br />

] β<br />

T<br />

− 1<br />

where V0 is the volume <strong>at</strong> p = 0 and temper<strong>at</strong>ure T.<br />

Problem 14.<br />

Derive the virial coefficients Bj(T ) for the Van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e.<br />

Comment on the sign of the second virial coefficient.<br />

Problem 15.<br />

)


3.9. PROBLEMS FOR CHAPTER 3 149<br />

The Gibbs free energy of pure m<strong>at</strong>erials A and B is given by NAµA(p, T ) and<br />

NBµB(p, T ). The entropy of mixing for m<strong>at</strong>erials A and B is −NAR log( NA<br />

NA+NB )−<br />

NBR log( NB<br />

NA+NB<br />

), while the interaction energy between A and B is given by<br />

A(T, p) NANB<br />

NA+NB . Show th<strong>at</strong> a system where we mix NA moles of A and NB<br />

moles of B is stable for 2RT > A and th<strong>at</strong> there are forbidden regions when<br />

2RT < A.<br />

Problem 16.<br />

The Clausius-Clapeyron rel<strong>at</strong>ion reads dpco<br />

dT = l<br />

T (v2−v1) . The l<strong>at</strong>ent he<strong>at</strong> of<br />

fusion for ice is l = 333 kJ kg−1 , the density of w<strong>at</strong>er is 1000 kg m−3 and the<br />

density of ice is 900 kg m−3 . At normal pressure, 105 N m−2 , the melting point<br />

of ice is 273 K. A sk<strong>at</strong>er of weight 1000 N using sk<strong>at</strong>es with a contact area<br />

of 10−5 m2 sk<strong>at</strong>es well if the ice under the sk<strong>at</strong>es melts. Wh<strong>at</strong> is the lowest<br />

temper<strong>at</strong>ure in degrees centigrade <strong>at</strong> which this sk<strong>at</strong>er will be able to sk<strong>at</strong>e? Do<br />

not use a calcul<strong>at</strong>or, I have changed the numbers slightly so you can do without.<br />

Problem 17.<br />

The boiling point of Hg is 357C <strong>at</strong> 1 <strong>at</strong>m. and 399C <strong>at</strong> 2 <strong>at</strong>m. The density<br />

of liquid mercury is 13.6 g/cm 3 . The he<strong>at</strong> of vaporiz<strong>at</strong>ion is 270 J/g. Calcul<strong>at</strong>e<br />

the order of magnitude of the volume of 200g Hg vapor <strong>at</strong> 357C and 1 <strong>at</strong>m.<br />

Note, 1 <strong>at</strong>m. is about 10 5 N/m 2 , 0C=273K, and G=U-TS+pV.<br />

Problem 18.<br />

The pressure of a gas is measured and it is found th<strong>at</strong> the following equ<strong>at</strong>ion<br />

is a good model for the pressure <strong>at</strong> low densities:<br />

p = NRT<br />

V<br />

3 N<br />

− aN2 + bRT<br />

V 2 V 3<br />

with a and b positive numbers. The model is now used <strong>at</strong> all densities. This<br />

gas shows a phase transition to a liquid st<strong>at</strong>e. Find the critical temper<strong>at</strong>ure of<br />

the transition, in terms of the constants a and b and the gas constant R.<br />

Problem 19.<br />

Near a critical point in the liquid to vapor phase transition we have the following<br />

rel<strong>at</strong>ion between volume and temper<strong>at</strong>ure on the border of the forbidden<br />

region:<br />

Tc − T = A(V − Vc) 2


150 CHAPTER 3. PHASE TRANSITIONS.<br />

A gas is <strong>at</strong> a temper<strong>at</strong>ure T above Tc and a volume V close to but not<br />

equal to V0. It is cooled down. Find the temper<strong>at</strong>ure <strong>at</strong> which we see phase<br />

separ<strong>at</strong>ion, and find the fraction of liquid and vapor below this temper<strong>at</strong>ure as<br />

a function of temper<strong>at</strong>ure.<br />

Problem 20.<br />

According to the Goff-Gr<strong>at</strong>ch equ<strong>at</strong>ion (Smithsonian Tables, 1984) we have<br />

for the vapor pressure pi of w<strong>at</strong>er vapor over ice:<br />

10 log( pi<br />

p0<br />

) = −9.09718 ( Tt<br />

T − 1) − 3.56654 10 log( Tt<br />

T<br />

) + 0.876793 (1 − )<br />

T<br />

with T in [K] and pi in [hPa]. p0 is the pressure <strong>at</strong> the triple point, p0 =<br />

6.1173 hP a. Tt is the temper<strong>at</strong>ure <strong>at</strong> the triple point, Tt = 273.16 K.<br />

Calcul<strong>at</strong>e the he<strong>at</strong> of sublim<strong>at</strong>ion of ice as a function of temper<strong>at</strong>ure. Wh<strong>at</strong><br />

is the r<strong>at</strong>io of the he<strong>at</strong> of sublim<strong>at</strong>ion of ice <strong>at</strong> 0 ◦ C and −10 ◦ C?<br />

PS: 1 <strong>at</strong>m = 1013.25 hPa.<br />

Problem 21.<br />

Consider the phase change of a m<strong>at</strong>erial AcB1−c from liquid to a solid solution,<br />

th<strong>at</strong> is in the solid the m<strong>at</strong>erials can always be mixed <strong>at</strong> any concentr<strong>at</strong>ion.<br />

When <strong>at</strong> a given temper<strong>at</strong>ure and pressure there is equilibrium between liquid<br />

and solid, we need to have th<strong>at</strong> the chemical potential for each <strong>at</strong>om type is the<br />

same in the liquid as in the solid, µl = µs. The chemical potential of <strong>at</strong>om type<br />

A and <strong>at</strong>om type B do not have to be the same. The total number of A <strong>at</strong>oms,<br />

in solid and liquid combined, is given by cN, where N is the total number of<br />

<strong>at</strong>oms. Similar for B <strong>at</strong>oms. The chemical potential for <strong>at</strong>oms A in the liquid<br />

is given by µAl = gAl(p, T ) + RT log(cAl), where cAl is the concentr<strong>at</strong>ion of A<br />

<strong>at</strong>oms in the liquid, the r<strong>at</strong>io of the number of A <strong>at</strong>oms in the liquid divided<br />

by the total number of <strong>at</strong>oms in the liquid. We have a similar expression for all<br />

other three cases.<br />

The solid and the liquid can have different concentr<strong>at</strong>ions of A <strong>at</strong>oms, call<br />

these cs and cl. Find formulas for these concentr<strong>at</strong>ions, rel<strong>at</strong>ing them directly<br />

to the quantities e gAs (p,T )−gAl (p,T )<br />

RT and e gBs (p,T )−gBl (p,T )<br />

RT .<br />

Find the r<strong>at</strong>io of the number of <strong>at</strong>oms in the liquid, Nl, and the solid, Ns.<br />

Problem 22.<br />

We can model a peak in the specific he<strong>at</strong> by the following formula:<br />

CV (T, V, N) = C0(T, V, N) + NRf( n ∆<br />

)T<br />

∆2 + (T − T0) 2<br />

n0<br />

Tt


3.9. PROBLEMS FOR CHAPTER 3 151<br />

where n is the density n = N<br />

V , f(n) is a smooth function of n, n0 is a<br />

reference density, and ∆ is a small parameter, measured in Kelvin. R is the gas<br />

constant.<br />

Is there a peak in the coefficient of thermal expansion? In the compressibility?<br />

Evalu<strong>at</strong>e these quantities!


152<br />

PART II<br />

<strong>Thermodynamics</strong> Advanced<br />

Topics


Chapter 4<br />

Landau-Ginzburg theory.<br />

The discussion of phase transitions in the previous chapter was motiv<strong>at</strong>ed by experimental<br />

observ<strong>at</strong>ions. If we have an equ<strong>at</strong>ion of st<strong>at</strong>e, wh<strong>at</strong> can we conclude<br />

from its behavior when it is singular? How does it lead to phase transitions?<br />

All those observ<strong>at</strong>ions are very important, they set the stage for a discussion of<br />

phase transitions from a more fundamental perspective.<br />

The ultim<strong>at</strong>e basis for a description of phase transitions is found in st<strong>at</strong>istical<br />

mechanics, and we will need a microscopic description for th<strong>at</strong>. From such a<br />

microscopic discussion we obtain free energies. But free energies are the domain<br />

of thermodynamics, and hence we can discuss the consequences of different forms<br />

of the free energy without knowing the microscopic found<strong>at</strong>ion for them. Th<strong>at</strong><br />

is exactly wh<strong>at</strong> we will do in this chapter.<br />

The most important control variable we have for phase transitions is the<br />

temper<strong>at</strong>ure, and therefore we need the free energy as a function of temper<strong>at</strong>ure.<br />

In other words, we need the Helmholtz free energy. Landau gave a very<br />

ingenious discussion of phase transitions in terms of expansions of the Helmholtz<br />

free energy in terms of some important extensive parameter. This analysis forms<br />

the basis for a classific<strong>at</strong>ion of all possible phase transitions, and any form of<br />

the Helmholtz free energy obtained from microscopic theory can be analyzed in<br />

the terms explained in this chapter. Again, this is an example of how thermodynamics<br />

dict<strong>at</strong>es the limit<strong>at</strong>ions of wh<strong>at</strong> can happen in a general theory.<br />

In this chapter we first discuss the very important concept of an order parameter,<br />

the way we can measure the occurrence of a phase transition. Next<br />

we set up Landau’s approach to the analysis of phase transitions. We apply the<br />

theory to the two most common phase transitions, first and second order.<br />

In the final section we step away from uniform behavior and ask the question<br />

wh<strong>at</strong> happens when we make a local change in the system. How far does it<br />

extends? This leads to the introduction of a new concept, the correl<strong>at</strong>ion length.<br />

This turns out to be an extremely important parameter for a description of phase<br />

transitions. It is the macroscopic description of the effectiveness of microscopic<br />

interactions between particles and the range over which they extend. It is the<br />

parameter th<strong>at</strong> describes the effects of the microscopic many body interactions.<br />

153


154 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

It constitutes the key connection between microscopic and macroscopic theory.<br />

4.1 Introduction.<br />

Phase transition in equilibrium thermodynamics.<br />

Equilibrium thermodynamics as presented in the previous chapters gives us<br />

st<strong>at</strong>e functions and free energies. In particular, we should be able to find the<br />

Gibbs free energy G(p,T,N). In Chapter 3 we used these ideas to describe phase<br />

transitions. Suppose a system can exist in two different phases, 1 and 2, and<br />

we have a model for the Gibbs free energy in each phase. For each value of<br />

the temper<strong>at</strong>ure and pressure we can therefore determine the Gibbs free energy<br />

of each phase, and by comparing the values we can determine which phase is<br />

stable and which is meta-stable. Sometimes we do not have the explicit forms<br />

of the free energies, but have a model equ<strong>at</strong>ion of st<strong>at</strong>e (like the van der Waals<br />

equ<strong>at</strong>ion of st<strong>at</strong>e) which allows us to rel<strong>at</strong>e the free energies of the two phases to<br />

each other. Since only the difference in free energy is important, this approach<br />

is equivalent to the one sketched first.<br />

Wh<strong>at</strong> happens <strong>at</strong> the transition?<br />

The approach outlined above is very useful, since it gives us equ<strong>at</strong>ions like the<br />

Clausius-Clapeyron equ<strong>at</strong>ion which tell us how phase boundaries change. This<br />

approach is an either/or approach, however. It allows us to determine which<br />

phase is stable, but it does not say anything about the transition. For example,<br />

the he<strong>at</strong> capacity as a function of temper<strong>at</strong>ure in the van der Waals model looks<br />

like figure 4.1, while experimentally it looks much more like figure 4.2. There<br />

is a critical region in which the he<strong>at</strong> capacity is very large (and perhaps even<br />

divergent). The width of this region could be 10K or 1mK, depending on the<br />

m<strong>at</strong>erial.<br />

Fluctu<strong>at</strong>ions.<br />

The next question is wh<strong>at</strong> is missing from our simple equilibrium thermodynamics<br />

model? Anyone who has seen w<strong>at</strong>er boil knows the answer. There are<br />

fluctu<strong>at</strong>ions in the system. Even below the boiling point vapor bubbles appear<br />

in w<strong>at</strong>er. These instabilities live only a short time, but when we get closer to<br />

the boiling point, the vapor bubbles become larger and live longer. To describe<br />

wh<strong>at</strong> is happening is very difficult, and we need a non-equilibrium theory. It<br />

is very important in practical applic<strong>at</strong>ions, though. Any prepar<strong>at</strong>ion technique<br />

of a m<strong>at</strong>erial th<strong>at</strong> involves a phase transition has to deal with this question,<br />

and understanding and being able to change the mechanism could be worth<br />

millions of dollars for industry. In this chapter we describe a theory developed


4.1. INTRODUCTION. 155<br />

Figure 4.1: He<strong>at</strong> capacity across the phase transition in the van der Waals<br />

model.<br />

Figure 4.2: He<strong>at</strong> capacity across the phase transition in an experiment.


156 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

by Landau and Ginzburg, which is a first approach <strong>at</strong> incorpor<strong>at</strong>ing the effects<br />

of fluctu<strong>at</strong>ions.<br />

4.2 Order parameters.<br />

How do we bring order or organiz<strong>at</strong>ion into chaos?<br />

How do we describe a phase transition?<br />

Phase transitions are observed by doing measurements. The question is,<br />

wh<strong>at</strong> do we measure? The phase transition from w<strong>at</strong>er to ice is easy to observe<br />

experimentally. The properties of w<strong>at</strong>er and ice are very different and we rely<br />

on the fact th<strong>at</strong> these properties have a range of values to distinguish between<br />

ice and w<strong>at</strong>er. Since snow is also frozen w<strong>at</strong>er, but different from ice, we also<br />

know th<strong>at</strong> not all properties of ice are characteristic for frozen w<strong>at</strong>er. We need<br />

to single out one or more parameters which will really tell us which phase we are<br />

dealing with. In a solid the rel<strong>at</strong>ive positions of the molecules are fixed (within<br />

small vibr<strong>at</strong>ions) but in a liquid the molecules are able to move freely. In a liquid<br />

the motion of the molecules is locally correl<strong>at</strong>ed, and the local neighborhood of<br />

a molecule in a liquid is similar to th<strong>at</strong> of a molecule in a solid. There is no<br />

long range order, though. In a gas, even the local correl<strong>at</strong>ion has disappeared,<br />

and the molecules are almost independent.<br />

Requirements for characterizing parameter.<br />

If we want to describe a phase transition in m<strong>at</strong>hem<strong>at</strong>ical terms, we need to<br />

find a parameter which uniquely defines which phase we are in. Therefore, this<br />

parameter has to be a st<strong>at</strong>e variable. In a gas to liquid transition we used the<br />

volume to distinguish between the low and high density phases, but this choice<br />

has some problems. If we use a st<strong>at</strong>e variable to distinguish between different<br />

phases of a system, it has to be measurable. Hence, it must be possible to use<br />

this st<strong>at</strong>e variable to do work. But we also require th<strong>at</strong> a measurement of this<br />

st<strong>at</strong>e variable does not change the phase of the system, hence in principle is<br />

should be measured with a corresponding field which approaches zero. In th<strong>at</strong><br />

case the st<strong>at</strong>e variable we use to describe the phase of the system is considered<br />

an internal parameter. It also has to be an extensive parameter. The volume in<br />

a gas to liquid transform<strong>at</strong>ion could play this role, but the ideal gas law in the<br />

limit p → 0 becomes singular. To describe a gas one needs a non-zero pressure,<br />

and the work done by changing the volume becomes an essential part of the<br />

physics of a gas to liquid phase transition.


4.2. ORDER PARAMETERS. 157<br />

Order parameter.<br />

A st<strong>at</strong>e variable which is used to describe the phase of a system is called<br />

an order parameter. The difference in molar volume between the phase under<br />

consider<strong>at</strong>ion and the gas phase is an example of an order parameter. This order<br />

parameter is zero in the gas phase, but non-zero in the liquid phase. Another<br />

standard example is the magnetic moment ⃗ M of a solid. We often use the molar<br />

density ⃗ M<br />

N in stead. This order parameter is zero in the non-magnetic phase,<br />

but non-zero in the ferro-magnetic phase. There is always a generalized force<br />

associ<strong>at</strong>ed with the order parameter, the corresponding intensive st<strong>at</strong>e variable.<br />

For the liquid to gas transition this is the pressure p and for the magnetic system<br />

it is the magnetic field ⃗ H. This generalized force is wh<strong>at</strong> we need to apply in<br />

order to be able to measure the value of the order parameter.<br />

Conserved or non-conserved order parameters.<br />

Some people make a distinction between two types of order parameters. If<br />

we study a ferromagnetic m<strong>at</strong>erial as a function of temper<strong>at</strong>ure <strong>at</strong> zero applied<br />

magnetic field, we find th<strong>at</strong> below the critical temper<strong>at</strong>ure the magnetiz<strong>at</strong>ion<br />

has a non-zero value. Hence we have an extensive st<strong>at</strong>e variable changing its<br />

value, its ”amount” is not conserved. On the other hand, if we study a fixed<br />

volume of gas and cool it down, we see liquid appear. We have two phases with<br />

different densities, but the total density remains constant, by construction. In<br />

this case the order parameter is called conserved. But it is clear th<strong>at</strong> this is only<br />

a m<strong>at</strong>ter of choosing st<strong>at</strong>e variables. If we cool a gas <strong>at</strong> constant pressure we<br />

are in the same situ<strong>at</strong>ion as the ferromagnet, and the volume is not conserved.<br />

Hence the distinction between two types of order parameters is not fundamental,<br />

but has, of course, very important practical consequences!<br />

Measuring the order parameter.<br />

The external parameters used to determine the st<strong>at</strong>e a solid are often the<br />

temper<strong>at</strong>ure T and the pressure p. For the following discussion, let us assume<br />

th<strong>at</strong> we work <strong>at</strong> <strong>at</strong>mospheric pressure, which is essentially zero for a solid. The<br />

appropri<strong>at</strong>e free energy for a system <strong>at</strong> constant temper<strong>at</strong>ure is the Helmholtz<br />

energy F, which in our case only depends on the st<strong>at</strong>e variables T (determined<br />

by the outside world) and ⃗ M (used to determine the st<strong>at</strong>e of the system). The<br />

Helmholtz free energy of the thermodynamic equilibrium st<strong>at</strong>e is found by minimizing<br />

F as a function of ⃗ M. This is consistent with the following observ<strong>at</strong>ion.<br />

In the process of measuring the order parameter we need to consider it as an<br />

external st<strong>at</strong>e variable. In our case the measurement consists of applying a field<br />

⃗H and the appropri<strong>at</strong>e free energy in th<strong>at</strong> case is the Gibbs like free energy<br />

G(T, ⃗ H) = U − T S − ⃗ M · ⃗ H (4.1)


158 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.3: Continuity of phase transition around critical point in p-T plane.<br />

Measuring the work done on the sample <strong>at</strong> constant temper<strong>at</strong>ure as a function<br />

of field gives this energy, and the partial deriv<strong>at</strong>ive with respect to ⃗ H in the<br />

limit ⃗ H → ⃗0 gives us ⃗ M as required. When we consider the magnetic moment<br />

as an external parameter, it is rel<strong>at</strong>ed to the external field via<br />

( )<br />

∂F<br />

∂ ⃗ M<br />

T<br />

= ⃗ H (4.2)<br />

and this partial deriv<strong>at</strong>ive is indeed zero in the limit ⃗ H → ⃗0. In conclusion, the<br />

free energy used to describe the system and the free energy used to describe the<br />

measurement of the order parameter differ by a simple Legendre transform<strong>at</strong>ion.<br />

Phase transitions with critical point.<br />

In the phase diagram for a PVT-system we find two different types of behavior.<br />

The transition from solid to gas or liquid is described by a coexistence curve<br />

which probably extends to infinity. There is no critical point. The liquid to gas<br />

coexistence curve, on the other hand, ends in a critical point. If we start <strong>at</strong> a<br />

st<strong>at</strong>e of the system <strong>at</strong> a high temper<strong>at</strong>ure, and cool the system <strong>at</strong> low pressure,<br />

we will end up in the gas phase. The same procedure <strong>at</strong> high pressure gives a<br />

liquid. Hence from any given initial st<strong>at</strong>e one can reach a st<strong>at</strong>e just above or<br />

below the coexistence curve in a continuous manner. This is shown in figure 4.3.<br />

One way of defining the order of a phase transition is using this continuity. A<br />

phase transition is called second order if it is possible to find a continuous p<strong>at</strong>h<br />

around a critical point from one phase to the other, and it is called first order<br />

if this is not the case. Note, however, th<strong>at</strong> other definitions of the order of a<br />

phase transition are in use, too.<br />

Branch cuts.


4.2. ORDER PARAMETERS. 159<br />

Figure 4.4: Continuity of phase around singular point.<br />

The coexistence curve plays a role which is very similar to th<strong>at</strong> of a branch<br />

cut in a complex function. The function √ z, for example, shows a similar<br />

picture which a line of discontinuities, see figure 4.4. From any initial value<br />

of z it is possible to go to a point just below or just above the branch cut<br />

in a continuous manner. The branch cut represents a line of discontinuities.<br />

The critical point in the p-T diagram is comparable to the essential singularity<br />

which ends a branch cut. Just like √ z cannot be described by a power series <strong>at</strong><br />

z = 0, the function V(p,T) has an essential singularity <strong>at</strong> pc, Tc. This approach<br />

has been generalized. All we need to know is the exact behavior of V(p,T) <strong>at</strong><br />

the singularity in order to predict wh<strong>at</strong> is happening near this point. This is<br />

very much like the complex analysis of a singular complex function. A magnetic<br />

phase transition is characterized in the same way by a real singular point Tc, Hc.<br />

If the system is symmetric for the change M ⇔ −M, it follows th<strong>at</strong> Hc = 0 and<br />

the singular behavior in the H-T plane is like in figure 4.5.<br />

Wh<strong>at</strong> conclusions can you draw from the similarity between the m<strong>at</strong>hem<strong>at</strong>ical<br />

forms of the free energy and a branch cut?<br />

Symmetry in system.<br />

When the external field coupling to the order parameter is zero, one often<br />

expects the free energy to be symmetric as a function of the order parameter.<br />

This is not true in general, since a vanishing external field only requires th<strong>at</strong> the<br />

first order deriv<strong>at</strong>ive is zero. But if we consider this simple symmetric case the<br />

free energy of the system without field is the same for a given value of the order<br />

parameter and for the opposite value. The most symmetric situ<strong>at</strong>ion therefore


160 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.5: Continuity of phase transition around critical point in H-T<br />

plane.<br />

will have a value of zero for the order parameter. This does often, but not<br />

necessarily, correspond to the st<strong>at</strong>e of the system <strong>at</strong> high temper<strong>at</strong>ures. For<br />

example, ⃗ M = ⃗0 <strong>at</strong> high temper<strong>at</strong>ures. There is no preferred direction of the<br />

magnetiz<strong>at</strong>ion. When we apply a magnetic field we will have in general th<strong>at</strong><br />

⃗M∥ ⃗ H.<br />

No singularities in finite system.<br />

If the equilibrium st<strong>at</strong>e of the system has a non-zero value of the order<br />

parameter, there are <strong>at</strong> least two st<strong>at</strong>es for the system with the same energy,<br />

corresponding to a certain value of the order parameter and to the opposite<br />

value. Therefore in th<strong>at</strong> case the ground st<strong>at</strong>e is degener<strong>at</strong>e. Although in<br />

principle the st<strong>at</strong>e of the system can fluctu<strong>at</strong>e between these degener<strong>at</strong>e ground<br />

st<strong>at</strong>es, the time involved is in practice infinitely long since macroscopic variables<br />

are involved. Quantum mechanical tunnelling is highly improbable in th<strong>at</strong> case.<br />

But we also need to remember th<strong>at</strong> we have to take the thermodynamic limit<br />

N → ∞. In th<strong>at</strong> case the tunnelling time becomes infinitely long, and the<br />

system is found in one st<strong>at</strong>e only. There is a m<strong>at</strong>hem<strong>at</strong>ical st<strong>at</strong>ement th<strong>at</strong> says<br />

th<strong>at</strong> for a finite system there are no singularities in st<strong>at</strong>e functions! Hence sharp<br />

transitions and critical points only occur in infinite systems!<br />

Wh<strong>at</strong> do phase transitions look like in finite systems?<br />

When is a finite system infinite for all practical purposes?


4.2. ORDER PARAMETERS. 161<br />

Broken symmetry.<br />

At the coexistence curve the system chooses one of the possible ground st<strong>at</strong>es,<br />

and breaks the symmetry between the possible values of the order parameter by<br />

choosing a particular one. In a scenario like we sketched here a phase transition<br />

is rel<strong>at</strong>ed to spontaneously broken symmetry in a degener<strong>at</strong>e ground st<strong>at</strong>e. Of<br />

course, the external field used to measure the order parameter, will always<br />

break the symmetry. In th<strong>at</strong> case the broken symmetry is forced by external<br />

constraints.<br />

Supercooling and superhe<strong>at</strong>ing.<br />

If we follow a complex function like √ z in a continuous manner through a<br />

branch cut, we end up on the wrong Riemann sheet. In the case of √ z this<br />

simply means th<strong>at</strong> we end up with the wrong sign, since there are only two<br />

Riemann sheets. If we go continuously through a coexistence curve, we end<br />

up in a meta-stable st<strong>at</strong>e. W<strong>at</strong>er vapor, for example, can be super-s<strong>at</strong>ur<strong>at</strong>ed.<br />

Introducing so-called seeds will then allow droplets of liquid to form around<br />

these seeds, and will force the transition to the stable liquid st<strong>at</strong>e.<br />

Could a free energy have more than two Riemann sheets, and if so, can they be<br />

observed?<br />

Essential singularity.<br />

A branch cut of a complex function end in an essential singularity, like the<br />

point z=0 for the function √ z. Second order phase transitions end in an essential<br />

singularity <strong>at</strong> the critical point. M<strong>at</strong>hem<strong>at</strong>ically, the idea of a critical point and<br />

an essential singularity are closely rel<strong>at</strong>ed.<br />

Model systems.<br />

The gas to liquid phase transition is not easy to describe. The problem is<br />

th<strong>at</strong> a gas has a non-zero volume, and th<strong>at</strong> a non-zero pressure exists, needed<br />

for thermal equilibrium. The driving force is non-zero! Therefore we need to<br />

consider quantities like the difference in volume and the difference in pressure.<br />

A symmetric magnetic phase transition is easier to describe. In one phase we<br />

have M=0, which occurs <strong>at</strong> H=0, and in the other phase we have a finite value<br />

of the magnitude M. Choosing M as the order parameter is therefore equivalent<br />

to choosing half the difference between the volumes in the liquid to gas system<br />

as the order parameter. Since magnetic systems are easier to model, we will use<br />

these systems as examples in the rest of this chapter. A model of a liquid to gas


162 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

transition, however, can be done in a similar manner, and only the m<strong>at</strong>hem<strong>at</strong>ical<br />

description becomes more elabor<strong>at</strong>e.<br />

4.3 Landau theory of phase transitions.<br />

Energy is the basic quantity.<br />

Landau devised a very simple method to describe phase transitions. It is<br />

similar in spirit to the approach of van der Waals, but on a more fundamental<br />

level. Landau recognized th<strong>at</strong> the free energy is the important st<strong>at</strong>e function to<br />

parameterize, since all other rel<strong>at</strong>ions follow from a knowledge of the free energy.<br />

The route van der Waals followed is more empirical. The van der Waals equ<strong>at</strong>ion<br />

of st<strong>at</strong>e works very well for a liquid-gas transition, but is hard to generalize to<br />

other phase transitions. Landau suggested to parameterize the free energy as a<br />

function of the order parameter.<br />

Magnetic systems are good working example.<br />

We will use a magnetic system as an example. We will assume th<strong>at</strong> the<br />

system has rot<strong>at</strong>ional symmetry, and th<strong>at</strong> the free energy does not depend on<br />

the direction of the magnetiz<strong>at</strong>ion ⃗ M, but only on its value M, apart from the<br />

obvious minus sign th<strong>at</strong> reflects up or down. In this discussion we therefore<br />

allow M to be neg<strong>at</strong>ive, which is necessary in order to show the degeneracy of<br />

the ground st<strong>at</strong>e. We assume th<strong>at</strong> we use the temper<strong>at</strong>ure T as an independent<br />

variable, and th<strong>at</strong> the volume V and the amount of m<strong>at</strong>erial N are constant.<br />

Therefore, we need to work with the Helmholtz free energy.<br />

The applied magnetic field follows from ( )<br />

∂F<br />

∂M = H and since we plan<br />

T,V,N<br />

to discuss the stable st<strong>at</strong>e <strong>at</strong> zero applied field, this partial deriv<strong>at</strong>ive is zero. We<br />

now switch our point of view, and consider M to be an internal parameter. The<br />

condition mentioned before is now consistent with the fact th<strong>at</strong> the Helmholtz<br />

free is a minimum as a function of the internal parameters. Since V and N<br />

are constant, we will drop these variables from the discussion, unless explicitly<br />

needed.<br />

Model Helmholtz energy.<br />

Because of the symmetry we imposed we have F (T, M) = F (T, −M) and an<br />

expansion of the free energy F only contains even powers of M. Because of this<br />

symmetry the st<strong>at</strong>e with M=0 always corresponds to zero applied external field.<br />

This st<strong>at</strong>e is therefore an extremum, but could be a maximum. If the minimum<br />

of the free energy occurs <strong>at</strong> a non-zero value of M, there will always be two<br />

minima, +M and -M. These are st<strong>at</strong>es with a spontaneous magnetiz<strong>at</strong>ion and<br />

give rise to the broken symmetry discussed before.


4.3. LANDAU THEORY OF PHASE TRANSITIONS. 163<br />

Landau studied the following form for the free energy:<br />

f(T, m) = a(T ) + 1<br />

2 b(T )m2 + 1<br />

4 c(T )m4 + 1<br />

6 d(T )m6 + · · · (4.3)<br />

, as a function<br />

of the average magnetiz<strong>at</strong>ion density, m = M<br />

V . The advantage of a formul<strong>at</strong>ion<br />

in terms of energy and magnetiz<strong>at</strong>ion densities is th<strong>at</strong> it does not depend on<br />

the size of the system. The coefficients a(T), etc. are intensive quantities. They<br />

take the form as expected in the thermodynamic limit, and are not dependent<br />

on N and V (but could depend on the density n = N<br />

V or equivalently on the<br />

pressure p, which dependence we will ignore here for simplicity).<br />

This specifies the Helmholtz free energy per unit volume, f = F<br />

V<br />

Validity of the expansion.<br />

One problem comes to mind <strong>at</strong> once. Landau assumed th<strong>at</strong> the order parameter<br />

is the same for the whole system. When the system changes phase,<br />

however, there is a coexistence of both phases and the order parameter depends<br />

strongly on position and time. This short-coming of the theory is not essential,<br />

and the improved theory of Ginzburg and Landau takes fluctu<strong>at</strong>ions into<br />

account, as we will discuss l<strong>at</strong>er. But, as we discussed before, <strong>at</strong> Tc there is an<br />

essential singularity, and a power series is not appropri<strong>at</strong>e, because it fails to<br />

converge! Hence Landau’s theory in not able to describe exactly wh<strong>at</strong> is<br />

happening for the phase transition <strong>at</strong> a critical point! In m<strong>at</strong>hem<strong>at</strong>ical<br />

terms, <strong>at</strong> the critical point the power series has a radius of convergence equal<br />

to zero.<br />

Value of the theory.<br />

Landau’s theory is very vers<strong>at</strong>ile for classifying all kinds of phase transitions<br />

by comparing the st<strong>at</strong>es <strong>at</strong> both sides of the transition. Keep in mind<br />

th<strong>at</strong> Landau’s theory was developed in 1936, before many models were solved<br />

analytically. By assuming very general functional forms for the coefficients<br />

a(T),b(T),· · ·, only a few basic types of phase transitions appear. Experimentalists<br />

use Landau theory very often to interpret their d<strong>at</strong>a, because it is a<br />

good macroscopic, first approach. It is therefore a good meeting ground for theory<br />

and experiment. Complex theoretical calcul<strong>at</strong>ions should try to derive the<br />

functional forms of the coefficients in Landau’s expansion, while experimentalists<br />

can measure them. This gives an excellent framework to discuss agreements<br />

and conflicts.<br />

The use of this classific<strong>at</strong>ion of phase transitions has been very helpful in<br />

understanding phase transitions, and is done l<strong>at</strong>er more rigorously in renormaliz<strong>at</strong>ion<br />

group theory. It also is another hint <strong>at</strong> universality, phase transitions<br />

with a similar analytical form of the coefficients behave essentially in the same<br />

way. This forms a basis for scaling theory, which is discussed in the next chapter.


164 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

4.4 Case one: a second order phase transition.<br />

One of the basic phase transitions is found when we assume th<strong>at</strong> b(T ) = b0(T −<br />

Tc) with b0 > 0, and c(T ) = c > 0, d(T ) = d > 0, with all higher order<br />

coefficients vanishing. Hence we study wh<strong>at</strong> happens when the first coefficient<br />

changes sign. The last coefficient, d, has to be positive or else the minimum free<br />

energy would occur <strong>at</strong> infinite magnetiz<strong>at</strong>ion. The partial deriv<strong>at</strong>ives are:<br />

( )<br />

∂f<br />

∂m T<br />

= b0(T − Tc)m + cm 3 + dm 5<br />

(4.4)<br />

( 2 ∂ f<br />

∂m2 )<br />

= b0(T − Tc) + 3cm 2 + 5dm 4<br />

(4.5)<br />

The slope is zero if either<br />

T<br />

m = 0 (4.6)<br />

b0(T − Tc) + cm 2 + dm 4 = 0 (4.7)<br />

In the first case this only corresponds to a minimum in the free energy if T > Tc,<br />

otherwise the curv<strong>at</strong>ure of the free energy 4.5 would be neg<strong>at</strong>ive. If a solution<br />

for m exists in the second case, it always corresponds to a local minimum since<br />

the curv<strong>at</strong>ure 4.5 in th<strong>at</strong> case is 2cm 2 +4dm 4 , which is always positive. In order<br />

to see if this st<strong>at</strong>e is the equilibrium st<strong>at</strong>e we have to compare the free energies.<br />

We have<br />

f(T, m) − f(T, 0) = 1<br />

2 b0(T − Tc)m 2 + 1<br />

4 cm4 + 1<br />

6 dm6 =<br />

1<br />

2 ⟨b0(T − Tc) + cm 2 + dm 4 ⟩m 2 − 1<br />

4 cm4 − 1<br />

3 dm6<br />

(4.8)<br />

and this is always neg<strong>at</strong>ive in the second case. Hence if 4.7 has a solution, this<br />

solution corresponds to the equilibrium st<strong>at</strong>e.<br />

General non-zero solution.<br />

The general solution of 4.7 is<br />

m 2 = 1<br />

(<br />

−c ±<br />

2d<br />

√ c2 )<br />

− 4db0(T − Tc)<br />

(4.9)<br />

If T > Tc both roots are neg<strong>at</strong>ive (remember c > 0) and m is not real. This is<br />

not a physical solution and hence for T > Tc the equilibrium st<strong>at</strong>e of the system<br />

has m = 0. For T < Tc, only one of the roots is positive and we find<br />

m 2 = 1<br />

(√ )<br />

c2 + 4db0(Tc − T ) − c<br />

2d<br />

(4.10)


4.4. CASE ONE: A SECOND ORDER PHASE TRANSITION. 165<br />

or<br />

Figure 4.6: Magnetiz<strong>at</strong>ion versus temper<strong>at</strong>ure.<br />

m 2 = c<br />

(√<br />

2d<br />

1 + 4db0<br />

c2 (Tc<br />

)<br />

− T ) − 1<br />

(4.11)<br />

In the limit T ↑ Tc, we can use √ 1 + x−1 ≈ 1<br />

2x, and the values for the magnetic<br />

moment are<br />

m = ±<br />

√ b0<br />

c<br />

√ Tc − T (4.12)<br />

which is shown in figure 4.6. Experiments show a very similar behavior, except<br />

th<strong>at</strong> the general form is<br />

m ∝ (Tc − T ) β<br />

(4.13)<br />

where typical values of β are around 0.33 in stead of 0.5 found in our Landau<br />

model. Therefore the magnetiz<strong>at</strong>ion is qualit<strong>at</strong>ively correct in the Landau model<br />

(for example, it goes to zero without a jump), but the exponent is wrong. For<br />

such quantit<strong>at</strong>ive details it is clearly important to know wh<strong>at</strong> happens <strong>at</strong> the<br />

phase transition in terms of fluctu<strong>at</strong>ions, and Landau’s theory is too simplistic,<br />

as mentioned before, because it ignores fluctu<strong>at</strong>ions .<br />

Free energy pictures.<br />

A picture of the Helmholtz free energy as a function of m looks like figure<br />

4.7. The analysis of these curves is easy. In the limit of infinite magnetiz<strong>at</strong>ion<br />

the curves go to positive infinity. At m=0 the curves have either a maximum<br />

or a minimum. There are two cases. For T > Tc there is only one extremum,<br />

and this has to be the point m=0, which therefore has to be a minimum. For


166 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.7: Forms of the Helmholtz free energy.<br />

T < Tc there are three extrema, one <strong>at</strong> zero, one positive, and one neg<strong>at</strong>ive.<br />

Therefore, the point m=0 has to be a maximum, and the minima are <strong>at</strong> the<br />

symmetric non-zero points.<br />

Since we are dealing with a fifth order equ<strong>at</strong>ion, could we have the sequence<br />

minimum, maximum, minimum, maximum, minimum for the extrema?<br />

Response functions.<br />

One of the important response function for our simple T system is the he<strong>at</strong><br />

capacity. The he<strong>at</strong> capacity is rel<strong>at</strong>ed to the partial deriv<strong>at</strong>ive of the entropy.<br />

Since volume in our model is not a st<strong>at</strong>e variable which can be varied, there is<br />

only one he<strong>at</strong> capacity and it is rel<strong>at</strong>ed to the free energy via:<br />

C = −V T d2 f<br />

dT 2<br />

(4.14)<br />

Since temper<strong>at</strong>ure is the only external st<strong>at</strong>e variable, there are no partial deriv<strong>at</strong>ives.<br />

The free energy is<br />

T > Tc : f(T ) = a(T ) (4.15)<br />

T < Tc : f(T ) = a(T ) + 1<br />

2 b0(T − Tc)m 2 + 1<br />

4 cm4 + 1<br />

6 dm6<br />

(4.16)<br />

Near Tc both the second and the third term in 4.16 are proportional to (T −Tc) 2 .<br />

Near Tc the higher order terms do not play a role, since they are proportional<br />

to larger powers of T − Tc. Hence f is approxim<strong>at</strong>ed by


4.4. CASE ONE: A SECOND ORDER PHASE TRANSITION. 167<br />

Figure 4.8: Entropy near the critical temper<strong>at</strong>ure.<br />

T ↑ Tc : f(T ) ≈ a(T ) − 1<br />

4<br />

The entropy per unit volume, s = S<br />

V<br />

= − df<br />

dT<br />

b2 0<br />

2<br />

(T − Tc)<br />

c<br />

and follows from<br />

(4.17)<br />

T > Tc : s(T ) = − da<br />

(T ) (4.18)<br />

dT<br />

T < Tc : s(T ) = − da<br />

dT (T ) + b2 0<br />

2c (T − Tc) (4.19)<br />

which has the behavior as shown in figure 4.8. The specific he<strong>at</strong> per unit volume<br />

and is given by<br />

follows from c = C<br />

V<br />

= T ds<br />

dT<br />

T > Tc : c(T ) = −T d2a (T ) (4.20)<br />

dT 2<br />

T < Tc : c(T ) = −T d2a dT 2 (T ) + T b20 (4.21)<br />

2c<br />

which is shown in figure 4.9. This is the same type of behavior as found in the<br />

van der Waals model, where the he<strong>at</strong> capacity also jumps. There is no divergent<br />

behavior, however, as seen in many experiments. This is again rel<strong>at</strong>ed to the<br />

fact th<strong>at</strong> we did not include fluctu<strong>at</strong>ions. We have a homogeneous system with<br />

m independent of time and position.<br />

Rel<strong>at</strong>ions across Tc.<br />

The Landau model rel<strong>at</strong>es the free energy below and above the transition<br />

temper<strong>at</strong>ure, and leads to two distinct analytical forms in these regions, with<br />

discontinuities. This is similar to the van der Waals model of a phase transition.


168 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.9: Specific he<strong>at</strong> near the critical temper<strong>at</strong>ure.<br />

The difference, however, is in the origin of the energy rel<strong>at</strong>ion. In stead of<br />

modelling an equ<strong>at</strong>ion of st<strong>at</strong>e, we have a model for the Helmholtz free energy<br />

th<strong>at</strong> is valid everywhere, but selects a different minimum in the different regions.<br />

This approach can be connected to more detailed models in an easier way, since<br />

it is more fundamental.<br />

The term rel<strong>at</strong>ed to a(T) is always present and gives a contribution to s<br />

and C which is continuous <strong>at</strong> the phase transition. The second term in the free<br />

energy for the ordered phase leads to a discontinuity in the slope of s, and the<br />

he<strong>at</strong> capacity jumps by an amount V Tc b2<br />

0<br />

2c . The entropy has an additional term<br />

b 2<br />

0<br />

2c (T − Tc) for temper<strong>at</strong>ures below Tc. Hence the entropy in the ordered st<strong>at</strong>e<br />

is less than the entropy in the st<strong>at</strong>e with m = 0. This is in agreement with our<br />

intuitive notion of entropy as a measure of the internal randomness of a system.<br />

The ordered st<strong>at</strong>e has lower entropy.<br />

Since the entropy is lower in the ordered st<strong>at</strong>e, why does the maximum entropy<br />

principle not imply th<strong>at</strong> the disordered st<strong>at</strong>e should be the stable st<strong>at</strong>e?<br />

Susceptibility.<br />

Very often it is easier to measure the magnetic susceptibility in stead of the<br />

magnetiz<strong>at</strong>ion density itself. This is due to the fact th<strong>at</strong> it is often easier to do a<br />

differential measurement. By definition χT (T, M) = ( )<br />

∂M<br />

∂H and χ−1<br />

T T (T, M) =<br />

)<br />

. Hence we find th<strong>at</strong><br />

( ∂ 2 F<br />

∂M 2<br />

T<br />

χ −1<br />

T (T, M) = V −1 ( b0(T − Tc) + 3cm 2 + 5dm 4)<br />

(4.22)


4.5. ORDER OF A PHASE TRANSITION. 169<br />

Figure 4.10: Magnetic susceptibility near the critical temper<strong>at</strong>ure.<br />

If we now use the solutions for m(T) we find<br />

T > Tc : χT (T ) =<br />

V<br />

b0(T − Tc)<br />

V<br />

T < Tc : χT (T ) =<br />

2b0(Tc − T )<br />

(4.23)<br />

(4.24)<br />

This response function is sketched in the figure 4.10. Note th<strong>at</strong> the susceptibility<br />

diverges <strong>at</strong> the critical point. The n<strong>at</strong>ure of the divergence is the same <strong>at</strong> each<br />

side of the transition, here a power of minus one in both cases. The exponent<br />

is the same. This is a general characteristic which we will discuss more in the<br />

next chapter. The pre-factors are different, though, and in this case the low<br />

temper<strong>at</strong>ure form has an extra factor of two.<br />

4.5 Order of a phase transition.<br />

One of the basic ways to distinguish phase transitions is by looking <strong>at</strong> the<br />

change in order parameter <strong>at</strong> the critical temper<strong>at</strong>ure. If the value of the order<br />

parameter jumps from one value to another, and hence is discontinuous, the<br />

phase transition is called first order. If this is not the case, the phase transition<br />

is called of second order. It is actually possible to even further distinguish the<br />

order of phase transitions, although perhaps not very useful. One way is to<br />

consider the functions ( ∂ n F<br />

∂T n<br />

)<br />

as a function of temper<strong>at</strong>ure near Tc. The order<br />

of the phase transition is then given by the lowest value of n for which this<br />

partial deriv<strong>at</strong>ive is discontinuous. Our previous example would be of second<br />

order in th<strong>at</strong> case.


170 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

4.6 Second case: first order transition.<br />

General observ<strong>at</strong>ions.<br />

In the general form of the Landau expansion, equ<strong>at</strong>ion 4.3, for a symmetric<br />

model, we can typically assume th<strong>at</strong> the lower order terms are the most important<br />

and have the most important temper<strong>at</strong>ure dependence. The term a(T) is<br />

always present and gives a continuous background contribution. If all the other<br />

terms are positive, then the st<strong>at</strong>e with m=0 is the only st<strong>at</strong>e corresponding to<br />

a minimum of the free energy. Therefore, in order to see some interesting behavior,<br />

<strong>at</strong> least on term should change sign as a function of temper<strong>at</strong>ure, or be<br />

neg<strong>at</strong>ive <strong>at</strong> all temper<strong>at</strong>ures. The easiest case is when the term b(T) changes<br />

sign. Here we study the next important case. Other fun cases are tre<strong>at</strong>ed in the<br />

problems.<br />

Neg<strong>at</strong>ive sign in fourth order term.<br />

This very interesting case is obtained when we assume th<strong>at</strong> b(T ) > 0, c(T ) =<br />

c < 0, d(T ) = d > 0, and all other coefficients are zero. We also assume th<strong>at</strong><br />

the dominant temper<strong>at</strong>ure dependence is in the b(T) term, and have indeed set<br />

the higher order terms to be constant. In order to see interesting behavior like<br />

normally observed, we assume th<strong>at</strong> db<br />

dT<br />

> 0. Since b(T ) > 0 the point m = 0 will<br />

always correspond to a local minimum of the Helmholtz free energy. It might,<br />

however, not be the global minimum. Because c < 0 it is possible th<strong>at</strong> the free<br />

energy curve bends down and goes through a maximum as a function of m. In<br />

th<strong>at</strong> case there has to be another minimum in the free energy as a function of<br />

m, since d > 0 implies th<strong>at</strong> for very large values of m the free energy approaches<br />

plus infinity. Three possible forms of the Helmholtz free energy as a function of<br />

m are shown in figure 4.11.<br />

Why don’t we have the sequence minimum, maximum, minimum in the free<br />

energy as a function of the order parameter?<br />

We again construct the partial deriv<strong>at</strong>ives of the Helmholtz free energy:<br />

( )<br />

∂f<br />

= b(T )m + cm<br />

∂m T<br />

3 + dm 5<br />

(4.25)<br />

( 2 ∂ f<br />

∂m2 )<br />

= b(T ) + 3cm 2 + 5dm 4<br />

(4.26)<br />

The solutions for m are<br />

T<br />

m = 0 (4.27)


4.6. SECOND CASE: FIRST ORDER TRANSITION. 171<br />

Figure 4.11: Three possible forms of the Helmholtz free energy in case 2.<br />

m 2 = −c ± √ c 2 − 4b(T )d<br />

2d<br />

(4.28)<br />

Hence if 4b(T )d > c2 only m = 0 is a solution. In th<strong>at</strong> case the fourth order<br />

term in the expansion of the Helmholtz free energy is not strong enough to pull<br />

the function down to a second minimum. In the other case we find two extrema.<br />

Note th<strong>at</strong> in this case the two possible values for m2 are always real, and hence<br />

the values of m are always real. The argument in the previous paragraph shows<br />

th<strong>at</strong> the extremum closest to m2 = 0 corresponds to a maximum, and the other<br />

one to a minimum. In order to obtain a value for m pertaining to th<strong>at</strong> second<br />

minimum we need to take the plus sign in 4.28. A picture of the solutions for m2 is given in figure 4.12, where the temper<strong>at</strong>ure T0 follows from 4b(T0)d = c2 . The<br />

values for m2 <strong>at</strong> this temper<strong>at</strong>ure are m2 (T0) = −c<br />

2d . Since b(T) is increasing<br />

with temper<strong>at</strong>ure, the non-zero solution only plays a role for T < T0.<br />

Energy difference.<br />

The important question, when there are more minima, is to determine which<br />

st<strong>at</strong>e corresponds to a global minimum. The difference in free energy is again<br />

f(T, m) − f(T, 0) = 1<br />

2 b0(T − Tc)m 2 + 1<br />

4 cm4 + 1<br />

6 dm6 =<br />

1<br />

2 ⟨b0(T − Tc) + cm 2 + dm 4 ⟩m 2 − 1<br />

4 cm4 − 1<br />

3 dm6 (4.29)<br />

and the term in brackets is zero <strong>at</strong> a minimum. Since c < 0 the difference in<br />

free energy can either be positive or neg<strong>at</strong>ive and the sign of the difference is<br />

determined by the sign of c + 4<br />

3dm2 . If the value of m2 for the minimum is<br />

small, the term with c will domin<strong>at</strong>e and the free energy for the m = 0 solution


172 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.12: Values for m corresponding to a minimum in the free energy.<br />

is the lowest. Substituting the solution for m in the difference of the free energy<br />

values we find<br />

∆F = − 1<br />

(<br />

m4 c + 2<br />

12 √ c2 )<br />

− 4b(T )d<br />

(4.30)<br />

The function b(T) is an increasing function of temper<strong>at</strong>ure, and for large values<br />

of the temper<strong>at</strong>ure (but always keeping 4b(T )d < c 2 ) the term in brackets is<br />

determined mainly by c and the difference in the free energy is positive (remember<br />

th<strong>at</strong> c is neg<strong>at</strong>ive). In th<strong>at</strong> case m = 0 is the stable st<strong>at</strong>e. For small<br />

values of the temper<strong>at</strong>ure, if b(T) becomes small enough, the term in brackets<br />

will change sign and the st<strong>at</strong>e with non-zero magnetic moment is stable.<br />

Critical temper<strong>at</strong>ure.<br />

Since we assume th<strong>at</strong> b(T) is monotonous, there will only be one value of<br />

the temper<strong>at</strong>ure for which the difference in free energy is zero. This critical<br />

temper<strong>at</strong>ure is determined by<br />

or<br />

c + 2 √ c 2 − 4b(T )d = 0 (4.31)<br />

b(Tc) = 3c2<br />

16d<br />

(4.32)<br />

At the phase transition the value of the order parameter jumps from zero to mc<br />

which is given by<br />

mc =<br />

√<br />

− 3c<br />

4d<br />

(4.33)


4.6. SECOND CASE: FIRST ORDER TRANSITION. 173<br />

Figure 4.13: Magnetiz<strong>at</strong>ion as a function of temper<strong>at</strong>ure.<br />

This transition is therefore of first order.<br />

Analysis of the transition.<br />

The complete plot of the value of the magnetiz<strong>at</strong>ion density as a function of<br />

temper<strong>at</strong>ure is given in figure 4.13. There are some important conclusions to<br />

be drawn. If we start <strong>at</strong> low temper<strong>at</strong>ure in a magnetized st<strong>at</strong>e and he<strong>at</strong> the<br />

system slowly, we will follow the meta-stable magnetic st<strong>at</strong>e even above Tc if<br />

we do things very carefully. This is because in order to go to the stable nonmagnetic<br />

st<strong>at</strong>e the system first has to increase its energy. There is an energy<br />

barrier. This is very typical for a first order phase transition. A system in this<br />

case can be superhe<strong>at</strong>ed or under-cooled into a meta-stable st<strong>at</strong>e. The further<br />

we are in the meta-stable region, the easier it is to perturb the system and force<br />

it to go to the stable st<strong>at</strong>e. The detailed mechanism of such phase transitions is<br />

often described in the framework of nucle<strong>at</strong>ion and growth. One needs to know<br />

how small seeds of the stable phase can form through fluctu<strong>at</strong>ions and how they<br />

can expand by growing. Clearly, in a solid this is a much slower process than<br />

in a liquid, since the motion of <strong>at</strong>oms in a solid is very slow.<br />

Hysteresis.<br />

When we perform a real experiment in which we bring a system through a<br />

first order phase transition we always do the experiment <strong>at</strong> a measurable speed,<br />

never reversibly slow. Therefore, we will always bring the system into the metastable<br />

st<strong>at</strong>e before it transforms to the stable st<strong>at</strong>e. This is called hysteresis,<br />

see figure 4.14. The amount by which we can bring the system into the metastable<br />

st<strong>at</strong>e depends on the m<strong>at</strong>erial we study. If nucle<strong>at</strong>ion and growth is<br />

slow, this can be quite far. Typically, nucle<strong>at</strong>ion is a thermal process with


174 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.14: Hysteresis loop.<br />

probability inversely proportional to a Boltzmann factor, or exponential in the<br />

energy difference between the st<strong>at</strong>es. Therefore, the probability of transforming<br />

becomes exponentially small near the critical temper<strong>at</strong>ure in such a model. As<br />

a consequence, reducing the speed of the process by a factor of two does not<br />

reduce the width of the hysteresis loop by a factor of two, but only by a small<br />

amount. This is of gre<strong>at</strong> importance in many applic<strong>at</strong>ions rel<strong>at</strong>ed to magnetism<br />

or phase transitions in crystal structures.<br />

Wh<strong>at</strong> do you expect to happen for hysteresis loops for very small samples?<br />

Discontinuities.<br />

In the first order phase transition discussed here the order parameter m<br />

jumps in value <strong>at</strong> the critical point. Since we switch from one minimum to another<br />

in the free energy curve, the change in free energy with temper<strong>at</strong>ure will<br />

in general be different in the two st<strong>at</strong>es. Hence df<br />

dT<br />

will also be discontinuous<br />

<strong>at</strong> the critical points, and hence the entropy jumps in value. This is completely<br />

different in a second order phase transition, where both m and S are continuous.<br />

Since m is continuous, there is also no hysteresis in a second order phase<br />

transition.<br />

Susceptibility.<br />

The susceptibility is obtained from χ −1<br />

(<br />

2<br />

∂ F<br />

T (T, M) = ∂M 2<br />

)<br />

T<br />

and is:


4.6. SECOND CASE: FIRST ORDER TRANSITION. 175<br />

T < Tc : χT (T ) =<br />

T > Tc : χT (T ) = V<br />

b(T )<br />

V<br />

b(T ) + 3cm2 V<br />

=<br />

+ 5dm4 2m2 (T ) (c + 2dm2 (T ))<br />

(4.34)<br />

(4.35)<br />

Since b(T) is positive and increasing, the susceptibility above Tc is decreasing.<br />

Also 4.28 shows th<strong>at</strong> m(T) is a decreasing function of T, and hence the susceptibility<br />

is an increasing function of temper<strong>at</strong>ure below Tc. Note th<strong>at</strong> c + 2dm 2<br />

is always positive.<br />

Entropy.<br />

( )<br />

∂f<br />

∂m<br />

( ) T<br />

∂f<br />

∂m<br />

T<br />

(<br />

∂f<br />

The entropy as a function of temper<strong>at</strong>ure follows from s = −<br />

dm<br />

dT<br />

∂T<br />

)<br />

m −<br />

. The deriv<strong>at</strong>ives of m versus temper<strong>at</strong>ure do not play a role, since<br />

= 0. Hence we find<br />

T > Tc : S(T ) = − da<br />

dT<br />

T < Tc : S(T ) = − da<br />

dT<br />

(4.36)<br />

1 db<br />

−<br />

2 dT m2 (T ) (4.37)<br />

Since b(T) is an increasing function of T, the entropy in the ordered st<strong>at</strong>e is<br />

again lower than the entropy in the disordered st<strong>at</strong>e, as expected. This is not<br />

in disagreement with the st<strong>at</strong>ement in chapter one, where we concluded th<strong>at</strong><br />

the entropy has to be a maximum, since th<strong>at</strong> st<strong>at</strong>ement pertains to a situ<strong>at</strong>ion<br />

of constant internal energy, and here the internal energy is different in the two<br />

st<strong>at</strong>es. If we assume th<strong>at</strong> b is linear in T, the he<strong>at</strong> capacity in the ordered<br />

phase differs from the he<strong>at</strong> capacity in the m = 0 phase by a term − db dm<br />

dT m(T ) dT<br />

and this is a positive term which does not diverge <strong>at</strong> Tc. The behavior of these<br />

quantities near Tc is sketched in figure 4.15.<br />

Basic difference in behavior.<br />

We have studied the description of a first order phase transition and of a<br />

second order phase transition. The stable st<strong>at</strong>e of the system is associ<strong>at</strong>ed with<br />

the global minimum in the free energy. The position of such a minimum depends<br />

on an external control variable, like the temper<strong>at</strong>ure. If as a function of<br />

the temper<strong>at</strong>ure the global minimum changes position by an interchange in a<br />

global-local minimum pair, the transition is first order. On the other hand, in<br />

a second order phase transition the global minimum becomes a maximum and<br />

in this process two minima are cre<strong>at</strong>ed on either side. This gives an important<br />

difference, which we have to keep in mind when we construct models for


176 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Figure 4.15: Critical behavior in first order phase transition.<br />

phase transitions. In a first order transition both st<strong>at</strong>es are available <strong>at</strong> either<br />

side of the transition, but in a second order transition new st<strong>at</strong>es are cre<strong>at</strong>ed.<br />

Therefore, hysteresis never occurs in a second order phase transition!<br />

Combined behavior.<br />

In real m<strong>at</strong>erials we see often combined types of behavior. For example,<br />

a transition could be second order in temper<strong>at</strong>ure like in our magnetic case.<br />

If we now work <strong>at</strong> a temper<strong>at</strong>ure below the critical temper<strong>at</strong>ure and apply a<br />

magnetic field, we see a first order transition in which the magnetiz<strong>at</strong>ion flips<br />

from one direction to the opposite direction. But this first order transition is<br />

clearly rel<strong>at</strong>ed to wh<strong>at</strong> happens <strong>at</strong> the critical point. In fact, the Landau model<br />

for a range around the critical point is given by the model in our first case. It<br />

makes therefore more sense to distinguish continuous phase transitions (anding<br />

in a critical point) from discontinuous transitions, and model the behavior<br />

accordingly.<br />

4.7 Fluctu<strong>at</strong>ions and Ginzburg-Landau theory.<br />

Fluctu<strong>at</strong>ions near a phase transition.<br />

The phenomenological theory of Landau has been very useful in classifying<br />

the order and n<strong>at</strong>ure of phase transitions. But the form 4.3 of the free energy<br />

is incomplete near a phase transition. We know th<strong>at</strong> near a phase transition<br />

fluctu<strong>at</strong>ions are important. For example, if we w<strong>at</strong>ch w<strong>at</strong>er boil, we see th<strong>at</strong><br />

vapor bubbles form near the bottom of a pan. Gravity will send these bubbles<br />

up. Since inside these bubbles the pressure is too low, they will collapse. Closer


4.7. FLUCTUATIONS AND GINZBURG-LANDAU THEORY. 177<br />

to the boiling point we will see both more bubbles forming, and larger bubbles<br />

forming. Especially near the critical point, one finds very large bubbles!<br />

How to deal with these?<br />

In order to study fluctu<strong>at</strong>ions we need non-equilibrium thermodynamics.<br />

We need to study a system th<strong>at</strong> is not uniform in space, and also changes over<br />

time. This is a difficult task. Ginzburg, however, found a very good first step<br />

to include sp<strong>at</strong>ial non-uniformity. He started from Landau’s model. He still<br />

assumed th<strong>at</strong> the problem was st<strong>at</strong>ic, but allowed for sp<strong>at</strong>ial vari<strong>at</strong>ions in the<br />

order parameter. In the approach studied here we do not discuss the origin of<br />

the driving forces towards non-uniformity, we simply assume th<strong>at</strong> they exist. We<br />

also ignore surface tension, which certainly plays a role, and simply assume th<strong>at</strong><br />

the effects of surface tension are incorpor<strong>at</strong>ed in the values of the parameters in<br />

our model.<br />

Divide and conquer.<br />

To understand the origin of our problem, we divide the volume of the total<br />

system in a large number of much smaller systems with a much smaller volume.<br />

On one hand, the number of small sub-systems is very large, but these subsystems<br />

themselves are still large enough to be considered to be macroscopic,<br />

e.g. they contain many <strong>at</strong>oms. All these small sub-systems are in thermal<br />

contact. All sub-systems are in thermal equilibrium. The total system is in<br />

thermal equilibrium and the average of the internal energy of each subsystem<br />

is a fixed value and the same everywhere. But the internal energy for a small<br />

sub-system may vary. The larger the devi<strong>at</strong>ion from the average value, the<br />

more unlikely the fluctu<strong>at</strong>ion. In this sense we consider the total system to be<br />

an ensemble of small systems in thermal contact.<br />

Wh<strong>at</strong> is a vapor bubble containing one <strong>at</strong>om?<br />

Position dependent order parameter.<br />

If the system is far away from a phase transition, all sub-systems will be in<br />

the same phase. If the total system is very close to a phase transition, about<br />

half of the sub-systems will have st<strong>at</strong>e variables which correspond to a metastable<br />

thermodynamical st<strong>at</strong>e. These sub-systems try to change phase until<br />

they have exchanged enough energy with the neighboring sub-systems, in which<br />

case they correspond to a stable st<strong>at</strong>e again. In terms of the order parameter<br />

the following description arises. A subsystem <strong>at</strong> position ⃗r has its own order<br />

parameter m(⃗r). The average value of the order parameter, ∫ m(⃗r)d 3 r, is equal


178 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

to the order parameter M for the total system and can be calcul<strong>at</strong>ed from the<br />

values of the st<strong>at</strong>e variables for the total system. Near a phase transition, there<br />

are large fluctu<strong>at</strong>ions in m(⃗r), however. Because our sub-systems contain many<br />

<strong>at</strong>oms, this implies th<strong>at</strong> vari<strong>at</strong>ions in m(⃗r) are on a length scale much larger<br />

than the inter-<strong>at</strong>omic distance, but possibly small compared to the sample size.<br />

Note th<strong>at</strong> for a homogeneous system we have m(⃗r) = m and M=mV indeed.<br />

How does this description of the order parameter compare with the definition<br />

of fields like ⃗ E(⃗r) in electromagnetism?<br />

Energy consider<strong>at</strong>ions.<br />

Ginzburg and Landau improved the phenomenological theory discussed before<br />

by including these fluctu<strong>at</strong>ions. They assumed th<strong>at</strong> the st<strong>at</strong>es of neighboring<br />

sub-systems are close together and th<strong>at</strong> m(⃗r) varies slowly. In th<strong>at</strong> case the<br />

fluctu<strong>at</strong>ions are taken into account by adding terms in the free energy which<br />

contain deriv<strong>at</strong>ives of m(⃗r). Precisely <strong>at</strong> the phase transition this will still be incorrect,<br />

since the order parameter or its slope varies in a discontinuous manner,<br />

and m(⃗r) is not analytical. An expansion of the free energy in terms of gradients<br />

of the order parameter is not valid near Tc and the improved Ginzburg-Landau<br />

theory is still not strictly valid near Tc. Nevertheless, it is a very useful theory,<br />

because it will be a good description of the st<strong>at</strong>e of a system much closer to the<br />

critical temper<strong>at</strong>ure than the previous Landau theory.<br />

Formal theory.<br />

The Helmholtz free energy is in Ginzburg-Landau theory given by an integr<strong>at</strong>ion<br />

over all space. As an example, we will extend our model of a second<br />

order phase transition to include fluctu<strong>at</strong>ions. In th<strong>at</strong> case we write<br />

∫<br />

F (T, m(⃗r)) =<br />

d 3 (<br />

r a(T ) + 1<br />

2 b0(T − Tc)m 2 (⃗r) + 1<br />

4 cm4 (⃗r) + 1<br />

6 dm6 )<br />

(⃗r)<br />

+ 1<br />

2 f<br />

∫<br />

d 3 <br />

<br />

r ⃗ <br />

<br />

∇m(⃗r)<br />

2<br />

(4.38)<br />

with b0, c, d, and f all positive. When the magnetic moment is homogeneous,<br />

m(⃗r) = M<br />

V , this expression is the same as we studied before, except for a redefinition<br />

of the coefficients a, b, c, and d by factors of powers of the volume. The<br />

last term in 4.38 represents fluctu<strong>at</strong>ions in the order parameter. The coefficient<br />

f has to be positive, else the system would be unstable against fluctu<strong>at</strong>ions.<br />

With a positive value of f we see th<strong>at</strong> it costs energy to cre<strong>at</strong>e a fluctu<strong>at</strong>ion, as


4.7. FLUCTUATIONS AND GINZBURG-LANDAU THEORY. 179<br />

expected. Assuming th<strong>at</strong> we have an isotropic system, the free energy cannot<br />

depend on the direction of ⃗ ∇m, and the form 4.38 is the simplest choice we can<br />

make.<br />

Functional deriv<strong>at</strong>ives derived.<br />

How does the Helmholtz free energy 4.38 change when the order parameter<br />

changes by a small amount δm(⃗r)? We need to calcul<strong>at</strong>e δF (T, m(⃗r)) =<br />

F (T, m(⃗r) + δm(⃗r)) − F (T, m(⃗r)). If the change in the order parameter is small,<br />

we only need to keep first order terms, and obtain<br />

∫<br />

δF (T, m(⃗r)) = d 3 rδm(⃗r) ( b0(T − Tc)m(⃗r) + cm 3 (⃗r) + dm 5 (⃗r) )<br />

∫<br />

+f d 3 r⃗ ∇m(⃗r) · ⃗ ∇δm(⃗r) (4.39)<br />

The last term is equal to<br />

∫<br />

−f d 3 rδm(⃗r) ⃗ ∇ 2 <br />

m(⃗r) + f d 2 σδm(⃗r)ˆn(⃗r) · ⃗ ∇m(⃗r) (4.40)<br />

The surface integral vanishes for a finite sample, since in th<strong>at</strong> case m(⃗r) = 0<br />

outside the sample. This integral also vanishes if we employ periodic boundary<br />

conditions. The fluctu<strong>at</strong>ions in the free energy are therefore:<br />

∫<br />

δF (T, m(⃗r)) =<br />

d 3 (<br />

rδm(⃗r) b0(T − Tc)m(⃗r) + cm 3 (⃗r) + dm 5 (⃗r) − f ⃗ ∇ 2 )<br />

m(⃗r)<br />

Position dependent magnetic field.<br />

(4.41)<br />

If we consider the magnetiz<strong>at</strong>ion as a external st<strong>at</strong>e variable, the magnetic<br />

field needed to cause a magnetiz<strong>at</strong>ion m(⃗r) is found by generalizing H = ( )<br />

∂F<br />

∂M T<br />

to a functional deriv<strong>at</strong>ive<br />

( )<br />

δF<br />

h(⃗r) =<br />

(4.42)<br />

δm T<br />

which is equivalent to<br />

and which leads to<br />

∫<br />

δF = d 3 rδm(⃗r)h(⃗r) (4.43)<br />

h(⃗r) = b0(T − Tc)m(⃗r) + cm 3 (⃗r) + dm 5 (⃗r) − f ⃗ ∇ 2 m(⃗r) (4.44)


180 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

by construction from equ<strong>at</strong>ion 4.41. This equ<strong>at</strong>ion is a typical differential equ<strong>at</strong>ion.<br />

The term with the Laplace oper<strong>at</strong>or occurs in many different places. The<br />

difference with other often studied second order differential equ<strong>at</strong>ions is the fact<br />

th<strong>at</strong> equ<strong>at</strong>ion 4.41 is non-linear.<br />

Spontaneous solutions.<br />

A spontaneous magnetiz<strong>at</strong>ion is possible when this equ<strong>at</strong>ion has solutions<br />

for h(⃗r) = 0, and this gives us exactly the situ<strong>at</strong>ion we discussed before. When<br />

h(⃗r) = 0 fluctu<strong>at</strong>ions will increase the free energy (this follows directly from<br />

equ<strong>at</strong>ion 4.38) and the ground st<strong>at</strong>e corresponds to a homogeneous magnetiz<strong>at</strong>ion<br />

m(⃗r) = M<br />

V . Equ<strong>at</strong>ion 4.44 then reduces to the form we solved before,<br />

leading to the same second order phase transition.<br />

Response to a driving force.<br />

The easiest way to study fluctu<strong>at</strong>ions is to consider m(⃗r) as an external<br />

st<strong>at</strong>e variable and ask the question how the system responds to certain kinds<br />

of external fields h(⃗r). This is like in resonance situ<strong>at</strong>ions, where we ask which<br />

driving force gives the largest response. For example, an interesting situ<strong>at</strong>ion<br />

arises when we use h(⃗r) = h0δ(⃗r). The system is locally perturbed and one<br />

would like to know how far away the effects of this perturb<strong>at</strong>ion can be seen. In<br />

terms of the sub-systems we introduced before, this means th<strong>at</strong> we change the<br />

st<strong>at</strong>e of one sub-system away from the average by applying a external magnetic<br />

field, and we want to know how the other sub-systems respond. The use of a<br />

Dirac delta-function is m<strong>at</strong>hem<strong>at</strong>ically justified if we assume th<strong>at</strong> it is possible<br />

to divide the large total system in a very large number of sub-systems. One<br />

can think of this by considering a delta function in the limit of a very narrow<br />

Gaussian, for example.<br />

Wh<strong>at</strong> could be the cause of such a small external force?<br />

Small external force.<br />

We will assume th<strong>at</strong> h0 is small, and th<strong>at</strong> the resulting magnetiz<strong>at</strong>ion m(⃗r)<br />

is close to the value without an external field, m(⃗r, T ) = m0(T ) + ϕ(⃗r, T ). The<br />

function ϕ will also be small. We are interested in changes in the magnetiz<strong>at</strong>ion<br />

near the critical point, when the magnetiz<strong>at</strong>ion itself is small. To simplify the<br />

m<strong>at</strong>hem<strong>at</strong>ics we will therefore ignore the sixth order term, which is small in this<br />

region. In other words, we set d = 0. We then know th<strong>at</strong><br />

0 = b0(T − Tc)m0(T ) + cm 3 0(T ) (4.45)


4.7. FLUCTUATIONS AND GINZBURG-LANDAU THEORY. 181<br />

√ b0(Tc−T )<br />

which has the solutions m0(T ) = 0 for T > Tc and m0(T ) = c for<br />

T < Tc. This equ<strong>at</strong>ion is subtracted from the equ<strong>at</strong>ion 4.44 determining m(⃗r, T )<br />

and only terms linear in ϕ are retained. Note th<strong>at</strong> m<strong>at</strong>hem<strong>at</strong>ically it is possible<br />

to have solutions with large values of ϕ, essentially given by cϕ3 − f ⃗ ∇2ϕ = 0,<br />

but such solutions also imply fast vari<strong>at</strong>ions of ϕ which is physically not possible<br />

because of the <strong>at</strong>omic structure of m<strong>at</strong>ter. As a result we get<br />

h0δ(⃗r) = b0(T − Tc)ϕ(⃗r, T ) + c3m 2 0(T )ϕ(⃗r, T ) − f ⃗ ∇ 2 ϕ(⃗r, T ) (4.46)<br />

The second term on the right is different above and below Tc. Using the solution<br />

for m0(T ) gives us<br />

T > Tc : ⃗ ∇ 2 ϕ(⃗r, T ) − b0(T − Tc)<br />

f<br />

T < Tc : ⃗ ∇ 2 ϕ(⃗r, T ) + 2 b0(T − Tc)<br />

f<br />

Solution away from origin.<br />

ϕ(⃗r, T ) = − h0<br />

δ(⃗r) (4.47)<br />

f<br />

ϕ(⃗r, T ) = − h0<br />

δ(⃗r) (4.48)<br />

f<br />

First we consider the differential equ<strong>at</strong>ion <strong>at</strong> all points away from the origin.<br />

In th<strong>at</strong> case the equ<strong>at</strong>ion has the form:<br />

⃗∇ 2 ϕ + αϕ = − h0<br />

δ(⃗r) = 0 (4.49)<br />

f<br />

with α < 0, which can be solved easily. The solutions are of the form<br />

1<br />

r ul(r)Ylm(Ω) (4.50)<br />

because the system has rot<strong>at</strong>ional symmetry. Since fluctu<strong>at</strong>ions increase the<br />

energy, the energy will be the lowest when ϕ has spherical symmetry, and hence<br />

we need to exclude all terms but the l=0 term. Therefore, ϕ = u(r)<br />

r (the factor<br />

√<br />

4π from the spherical harmonic is now included in the radial function u) and<br />

we need to solve<br />

u ′′ + αu = 0 (4.51)<br />

which has the solution u = e ±ı√αr . Since the total response has to be calcul<strong>at</strong>ed<br />

from ∫ ϕd3r we discard the increasing solution and we get ϕ = A<br />

r<br />

r<br />

e− ξ , where<br />

ξ √ α = ı.<br />

Effects of the origin.<br />

The effects of the delta function singularity <strong>at</strong> the origin are taken into<br />

account by integr<strong>at</strong>ing the equ<strong>at</strong>ion


182 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

⃗∇ 2 ϕ + αϕ = − h0<br />

δ(⃗r) (4.52)<br />

f<br />

over a sphere of radius R around the origin, and then take the limit R → 0.<br />

Th<strong>at</strong> leaves us with<br />

The first integral is<br />

∫<br />

r


4.8. EXTRA QUESTIONS. 183<br />

The perturb<strong>at</strong>ion is also proportional to the local magnetic field, as expected<br />

in our simple calcul<strong>at</strong>ion where we only included linear terms. For small fluctu<strong>at</strong>ions<br />

this is certainly justified. If the external field is zero, the response is<br />

zero, as expected.<br />

The perturb<strong>at</strong>ion is inversely proportional to f, the factor scaling the energy<br />

rel<strong>at</strong>ed to fluctu<strong>at</strong>ions. When f is large, fluctu<strong>at</strong>ions in the magnetiz<strong>at</strong>ion cost a<br />

lot of energy and are more unlikely. ϕ is indeed small in th<strong>at</strong> case. Amazingly,<br />

the correl<strong>at</strong>ion length is proportional to √ f. If in a system f is large, fluctu<strong>at</strong>ions<br />

are small, but they have a very large range! On the other hand, if f approaches<br />

zero, all fluctu<strong>at</strong>ions are essentially localized since ξ approaches zero, but the<br />

fluctu<strong>at</strong>ions themselves can be very large. In other words, when it is easy to<br />

cause a perturb<strong>at</strong>ion, this perturb<strong>at</strong>ion is also screened rapidly because the<br />

system can also gener<strong>at</strong>e a response easily.<br />

Note th<strong>at</strong> in the Ginzburg-Landau theory we merely explain wh<strong>at</strong> happens to<br />

a system when a fluctu<strong>at</strong>ion occurs. We do not explain why fluctu<strong>at</strong>ions occur,<br />

and how instabilities might arise. We only discovered a theory th<strong>at</strong> shows why<br />

fluctu<strong>at</strong>ions are screened and how we can measure th<strong>at</strong> screening.<br />

Divergence <strong>at</strong> Tc.<br />

The most important fe<strong>at</strong>ure of the equ<strong>at</strong>ions for the correl<strong>at</strong>ion length is<br />

the fact th<strong>at</strong> in the limit T → Tc the correl<strong>at</strong>ion length diverges. Near a phase<br />

transition any small change in the magnetiz<strong>at</strong>ion <strong>at</strong> one point will be noticeable<br />

all over the sample. This is an example of the inherent instability of a system<br />

<strong>at</strong> a phase transition. Note th<strong>at</strong> the n<strong>at</strong>ure of the divergence is the same when<br />

we approach the critical temper<strong>at</strong>ure from below or from above. Both go like<br />

the square root of the temper<strong>at</strong>ure difference. The pre-factors are not the same,<br />

however, and in our case they differ by a factor of √ 2.<br />

4.8 Extra questions.<br />

1. Why do we work with the Helmholtz free energy and not Gibbs?<br />

2. Why is there a peak in the he<strong>at</strong> capacity?<br />

3. Give examples of order parameters.<br />

4. How do you measure the value of the order parameter?<br />

5. Why are there no phase transitions in a finite system?<br />

6. Discuss the rel<strong>at</strong>ion between correl<strong>at</strong>ion length versus system size.<br />

7. Wh<strong>at</strong> is broken symmetry?<br />

8. To get f(t,m) from F(T,M,N), wh<strong>at</strong> did we use?<br />

9. When do power series converge?


184 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

10. Why does Landau’s theory imply some kind of universality?<br />

11. Why is Landau more fundamental than using an EOS?<br />

12. Work out the results for a second order transition with d = 0<br />

13. Would a theory with b(T ) > 0 and c(T ) changing sign give a phase transition?<br />

14. Work out results for b = 0, c changing sign, d positive.<br />

15. Why do we go from something of the form (T,m) to (T), and wh<strong>at</strong> happened<br />

with the additional degree of freedom?<br />

16. Using the free energy pictures, analyze the n<strong>at</strong>ure of phase transitions.<br />

17. Is hysteresis equilibrium?<br />

18. Can we under-cool a gas?<br />

19. If magnetic phase transitions are second order, why do we see hysteresis<br />

loops?<br />

20. Why is the gradient term the lowest order and the only second order term?<br />

21. Wh<strong>at</strong> are functional deriv<strong>at</strong>ives?<br />

22. Wh<strong>at</strong> does it mean th<strong>at</strong> the correl<strong>at</strong>ion length diverges <strong>at</strong> the critical temper<strong>at</strong>ure?<br />

23. Interpret the meaning of formula 4.56<br />

24. Do correl<strong>at</strong>ions decay <strong>at</strong> the critical temper<strong>at</strong>ure?<br />

25. Is it significant th<strong>at</strong> the scaling is the same <strong>at</strong> both sides of the critical<br />

temper<strong>at</strong>ure.<br />

4.9 Problems for chapter 4<br />

Problem 1.<br />

Suppose th<strong>at</strong> in the expansion for the Helmholtz free energy per unit volume<br />

f = F<br />

M<br />

V as a function of the magnetiz<strong>at</strong>ion density m = V a third order term<br />

is present. This is possible in systems consisting of large anisotropic molecules.<br />

The free energy in this case is<br />

f(m, T ) = a(T ) + 1<br />

2 b(T )m2 + 1<br />

3 c(T )m3 + 1<br />

d(T )m4<br />

4<br />

Assume th<strong>at</strong> b(T ) = b0 ∗ (T − T0) and b0 > 0, c(T ) = c < 0, and d(T ) = d > 0.


4.9. PROBLEMS FOR CHAPTER 4 185<br />

(a) Show th<strong>at</strong> this system exhibits a first-order phase transition <strong>at</strong> a temper<strong>at</strong>ure<br />

Tc.<br />

(b) Calcul<strong>at</strong>e Tc in terms of b0, T0, c, and d.<br />

(c) Calcul<strong>at</strong>e lim m(T ) .<br />

T ↑Tc<br />

Problem 2.<br />

In this chapter we have only considered the temper<strong>at</strong>ure as a st<strong>at</strong>e variable<br />

driving a phase transition. Suppose the pressure is also important. The<br />

Helmholtz free energy per unit volume is then given by<br />

f(m, p, T ) = a(p, T ) + 1<br />

2 b(p, T )m2 + 1<br />

4 c(p, T )m4 + 1<br />

d(P, T )m6<br />

6<br />

where we ignored all higher order terms. Assume th<strong>at</strong> d(p, T ) = d > 0.<br />

(a) If c(p, T ) = c > 0, give an expression which determines Tc as a function<br />

of pressure for a second order phase transition.<br />

(b) If c(p, T ) = c < 0, b(p, T ) > 0, and b(p,T) increases as a function of temper<strong>at</strong>ure,<br />

give an expression which determines Tc as a function of pressure<br />

for a first order phase transition.<br />

(c) Suppose b(p, T ) = b0p(T − T0) and c(p, T ) = c0T (p − p0), with b0 > 0 and<br />

c0 > 0. Sketch in a p-T diagram the function Tc(p) for both the first order<br />

and the second order phase transition.<br />

(d) Show th<strong>at</strong> the two curves in part (c) connect <strong>at</strong> one point (called a<br />

tri-critical point) and th<strong>at</strong> the deriv<strong>at</strong>ives of the curves are the same in<br />

this tri-critical point.<br />

Problem 3.<br />

The Helmholtz free energy per unit volume for a two dimensional system is<br />

given by<br />

f(mx, my, T ) = a(T ) + 1<br />

2 b(T )(m2 x + m 2 y) + 1<br />

4 c(T )m2 xm 2 y + 1<br />

4 d(T )(m4 x + m 4 y)<br />

Assume th<strong>at</strong> b(T ) = b0(T − Tc) with b0 > 0, c(T ) = c > 0 and d(T ) = d > 0.<br />

(a) Show th<strong>at</strong> this system has a second order phase transition <strong>at</strong> Tc.


186 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

(b) If c = 2d, show th<strong>at</strong> m(T) is like we derived in the text for a second order<br />

phase transition, but th<strong>at</strong> mx and my cannot be determined apart from<br />

the value for m =<br />

√<br />

m 2 x + m 2 y.<br />

(c) If c ̸= 2d, show th<strong>at</strong> mx = my.<br />

Problem 4.<br />

Discuss wh<strong>at</strong> happens in a Landau theory with f(T ; m) = a(T ) + 1<br />

2b0(T −<br />

T0)m2 + 1<br />

4cm4 + 1<br />

6m6 and b0 > 0 , c < 0, and d > 0.<br />

Problem 5.<br />

Discuss wh<strong>at</strong> happens in a Landau theory with f(T ; m) = a(T ) + 1<br />

4 c0(T −<br />

T0)m 4 + 1<br />

6 m6 and c0 > 0 and d > 0. Calcul<strong>at</strong>e the response functions.<br />

Problem 6.<br />

Consider Ginzburg-Landau in d dimensions. Assume again a small perturb<strong>at</strong>ion<br />

of the form h(⃗r) = h0δ(⃗r) and look for a spherical solution. Show th<strong>at</strong><br />

the solutions for ϕ decay exponentially and th<strong>at</strong> for small r ϕ diverges like r 2−d .<br />

Problem 7.<br />

Calcul<strong>at</strong>e the l<strong>at</strong>ent he<strong>at</strong> for the first order phase transition discussed in the<br />

text.<br />

Problem 8.<br />

A system is characterized by a three dimensional order parameter ⃗ S. The<br />

Landau expansion is<br />

f(T ; ⃗ S) = a(T ) + 1<br />

2 b0(T − T0)S 2 + 1<br />

2 c0(T − 1<br />

2 T0)(ˆn · ⃗ S) 2 + 1<br />

4 dS4<br />

with b0 > 0, c0 > 0, d > 0 and ˆn a constant unit vector. Discuss the phase<br />

transitions in this model.<br />

Problem 9.<br />

The Helmholtz free energy per unit volume is given by


4.9. PROBLEMS FOR CHAPTER 4 187<br />

f(T ; m) = R(T + 1<br />

8 T0) cosh( 2m<br />

) − RT0 cosh( m<br />

)<br />

This models a phase transition. Calcul<strong>at</strong>e Tc. Is the transition first or second<br />

order?<br />

Problem 10.<br />

m0<br />

The Helmholtz free energy per unit volume has the form<br />

f(T ; m) = a(T ) + b0(T − T0)m √ m + 1<br />

4 cm4<br />

with b0 > 0 and c > 0. We have introduced an unusual square root term.<br />

Show th<strong>at</strong> this gives a second order phase transition and calcul<strong>at</strong>e the temper<strong>at</strong>ure<br />

dependence of the order parameter near the critical temper<strong>at</strong>ure.<br />

Problem 11.<br />

Consider a model for a second order phase transition in a magnetic field.<br />

The Helmholtz free energy per unit volume is given by<br />

f(T ; m) = a(T ) + 1<br />

2 b(T )m2 + 1<br />

c(T )m4<br />

4<br />

Assume c(T ) = c > 0, a positive constant, and b(T ) = b0(T −Tc) with b0 > 0.<br />

The magnetiz<strong>at</strong>ion density is still considered to be an internal parameter, which<br />

should be chosen to minimize the Helmholtz free energy. The condition is now<br />

( )<br />

∂f<br />

= H<br />

∂m<br />

Calcul<strong>at</strong>e m(T,H) (hint: use Maple or equivalent) and plot m(H) for various<br />

temper<strong>at</strong>ures and m(T) for various fields.<br />

Problem 12.<br />

Consider the following model of a tri-critical point:<br />

f(m, T, p) = 1<br />

2 a(T, p)m2 + 1<br />

4 b(T, p)m4 + 1<br />

6 cm6<br />

where c is a positive constant. The functions a and b are given by<br />

T − Tc p − pc<br />

a(T, p) = A + B<br />

Tc pc<br />

T − Tc p − pc<br />

b(T, p) = C + D<br />

Tc pc<br />

T<br />

m0


188 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Depending on the p<strong>at</strong>h in p-T space, along p = f(T ) with pc = f(Tc), the<br />

behavior of the response functions and the order parameter takes on a different<br />

form. For example, the behavior of the magnetiz<strong>at</strong>ion m(p, T ) always varies<br />

like (Tc − T ) β , but the value of β depends on the p<strong>at</strong>h taken. The two distinct<br />

regions with different values of β are called critical and tri-critical.<br />

(A) Calcul<strong>at</strong>e the β this system.<br />

(B) Draw the phase diagram in p-T space.<br />

(C) Draw the critical and tri-critical regions in this phase diagram.<br />

(D) Give an explan<strong>at</strong>ion of these results, write it in such a way th<strong>at</strong> a new,<br />

incoming gradu<strong>at</strong>e student could understand it.<br />

Problem 13.<br />

Consider the following model Helmholtz free energy per unit volume:<br />

f(T ; m) = a(T ) + e T 0 −T<br />

T 0 e −m2<br />

+ ce m2<br />

where a(T ) is an arbitrary function of temper<strong>at</strong>ure, and T0 and c are positive<br />

constants. Show th<strong>at</strong> this system undergoes a second order phase transition,<br />

calcul<strong>at</strong>e the critical temper<strong>at</strong>ure Tc, and find the critical exponent β.<br />

Problem 14.<br />

A magnetic system has a spontaneous magnetiz<strong>at</strong>ion below a critical temper<strong>at</strong>ure<br />

Tc, given by m2 = m2 0 Tc−T<br />

. The system is in a st<strong>at</strong>e <strong>at</strong> temper<strong>at</strong>ure<br />

Tc<br />

Ti = 2Tc and magnetiz<strong>at</strong>ion mi = 1<br />

2m0. This system is cooled down to very low<br />

temper<strong>at</strong>ure, but <strong>at</strong> constant magnetiz<strong>at</strong>ion. Describe the st<strong>at</strong>e of the system<br />

<strong>at</strong> this very low temper<strong>at</strong>ure. Calcul<strong>at</strong>e relevant quantities as a function of<br />

temper<strong>at</strong>ure.<br />

Problem 15.<br />

Consider the following Landau model, f(T, m) = a(T ) + b0<br />

2 (T − T0)m2 +<br />

1<br />

2N m2N , with b0 > 0. Find the magnetiz<strong>at</strong>ion as a function of temper<strong>at</strong>ure.<br />

Wh<strong>at</strong> happens to the critical exponent β when N → ∞? Wh<strong>at</strong> is the critical<br />

exponent if we change the model to f(T, m) = a(T ) + b0<br />

2 (T − T0)m2 + 1<br />

2em2 ?<br />

Problem 16.<br />

Consider the equ<strong>at</strong>ion for the perturb<strong>at</strong>ion of the magnetiz<strong>at</strong>ion density<br />

when there is a localized field, but write down the equ<strong>at</strong>ion <strong>at</strong> T = Tc. This


4.9. PROBLEMS FOR CHAPTER 4 189<br />

equ<strong>at</strong>ion is simpler than the one discussed in the notes. Find the behavior<br />

for large values of r. Compare this with the behavior away from the critical<br />

temper<strong>at</strong>ure.<br />

Problem 17.<br />

Consider the Landau model for magnetism in an AB alloy,<br />

f(T, mA, mB) = a(T )+ 1<br />

2 b0(T −Tc)(m 2 A+m 2 B)+bm(T −Tm)mAmB+ 1<br />

4 c0(m 4 A+m 4 B)<br />

where mA is the magnetiz<strong>at</strong>ion density for <strong>at</strong>oms A and mB for <strong>at</strong>oms B.<br />

The constants b0, c0, and bm are positive. Find mA(T ) and mB(T ).<br />

Problem 18.<br />

Consider a Landau model, for a second order phase transition,<br />

f(T, m) = a(T ) + b(T )m 2 + c0m 4<br />

Design coefficient a(T ), b(T ), and c0 th<strong>at</strong> give a second order phase transition<br />

where the magnetiz<strong>at</strong>ion is m(T ) = m0(Tc − T ) β for T < Tc and zero above.<br />

Discuss why you think such a model would be realistic or not.<br />

Problem 19.<br />

Using the general Landau formalism, equ<strong>at</strong>ion (4.3), show th<strong>at</strong> there is never<br />

a peak in the he<strong>at</strong> capacity <strong>at</strong> a phase transition.<br />

Problem 20.<br />

Use a Landau model of the form:<br />

and assume th<strong>at</strong><br />

f(T, m) =<br />

∞∑<br />

cn(T )m 2n<br />

n=0<br />

cn(T ) ≡ 0 , n = 0, · · · , N − 1<br />

cN (T ) = CN (T − Tc)<br />

cn(T ) ≡ Cn , n = N + 1, · · ·<br />

where CN , CN+1, etc are constants independent of the temper<strong>at</strong>ure.


190 CHAPTER 4. LANDAU-GINZBURG THEORY.<br />

Find the critical exponent β for this system<br />

Problem 21.<br />

Consider the following Landau model, f(T, m) = a(T ) + b0<br />

2 (T − T1)(T −<br />

T2)m2 + d<br />

4 m4 , with b0 > 0, T1 > 0, T2 > 0, and d > 0. Find the magnetiz<strong>at</strong>ion<br />

as a function of temper<strong>at</strong>ure. Find the entropy as a function of temper<strong>at</strong>ure.<br />

Problem 22.<br />

δ.<br />

Suppose we have in a Landau model<br />

T − Tc<br />

f(T, m) = a(T ) + b0m<br />

4<br />

4 + 1<br />

6 cm6<br />

with b0 > 0 and c > 0. Find the values of the critical exponents α, β, γ, and<br />

Problem 23.<br />

Suppose we have in a Landau model<br />

f[T, m(⃗r)] = a(T )+ 1<br />

∫ ∫<br />

b(T )<br />

2<br />

m(⃗r)J(⃗r−⃗r ′ )m(⃗r ′ )d 3 rd 3 r ′ + 1<br />

[∫<br />

c(T )<br />

4<br />

m 2 (⃗r)d 3 ] 2<br />

r<br />

Find the differential equ<strong>at</strong>ion for m(⃗r) th<strong>at</strong> yields the minimum for the<br />

functional.<br />

Problem 24.<br />

Suppose we have in a Landau model<br />

f[T, m(⃗r)] =<br />

∫ (<br />

a(T ) + 1<br />

2 b(T )m2 (⃗r) + 1<br />

4 c(T )m4 )<br />

(⃗r) d 3 ∫<br />

r+g(T )<br />

|⃗r× ⃗ ∇m(⃗r)| 2 d 3 r<br />

Find the differential equ<strong>at</strong>ion for m(⃗r) th<strong>at</strong> yields the minimum for the<br />

functional.<br />

Problem 25.<br />

Consider Landau’s model for a first order phase transition, but now including<br />

pressure:<br />

f(T, p; m) = a(T ) + 1<br />

2 b(T )m2 + 1<br />

4 c(p)m4 + 1<br />

6 dm6


4.9. PROBLEMS FOR CHAPTER 4 191<br />

> 0, c(p) < 0, d > 0.<br />

Find an expression for dTc in terms of the coefficients b, c, and d. If the<br />

with b(T ) > 0, db<br />

dT<br />

dp<br />

critical temper<strong>at</strong>ure decreases as a function of pressure, wh<strong>at</strong> can you say about<br />

dc<br />

dp ?<br />

Problem 26.<br />

Consider the Landau model for a second order phase transition. We set<br />

the temper<strong>at</strong>ure equal to the critical temper<strong>at</strong>ure, T = Tc. Of course, we then<br />

find m = 0. Now we apply a small magnetic field H and measure the induced<br />

magnetic moment. We find th<strong>at</strong> m ∝ H 1<br />

δ . Find the value of δ.<br />

Problem 27.<br />

Give a Landau type of expansion for the liquid to vapor phase transition.<br />

Note: you need to define an order parameter first!


192 CHAPTER 4. LANDAU-GINZBURG THEORY.


Chapter 5<br />

Critical exponents.<br />

At first glance the analysis of a phase transition is easy. We have two possible<br />

st<strong>at</strong>es for the system, and the st<strong>at</strong>e with lowest free energy wins. But the observ<strong>at</strong>ions<br />

in the last part of the previous chapter open a new way of discussing<br />

wh<strong>at</strong> happens <strong>at</strong> a phase transition. Suppose we are close to the critical temper<strong>at</strong>ure<br />

of a phase transition. Locally, it is possible to have a thermal fluctu<strong>at</strong>ion<br />

which transforms a small part of the m<strong>at</strong>erial to the unstable phase. This region<br />

will grow until it reaches a size dict<strong>at</strong>ed by the correl<strong>at</strong>ion length, but then will<br />

disappear again because its internal energy is too high. We observe this directly<br />

when we boil w<strong>at</strong>er. Also, <strong>at</strong> very low temper<strong>at</strong>ures thermal fluctu<strong>at</strong>ions are<br />

possibly too small, and quantum fluctu<strong>at</strong>ions near a critical point become dominant.<br />

Transitions where the instabilities are driven by quantum fluctu<strong>at</strong>ions<br />

are called quantum phase transitions.<br />

The observ<strong>at</strong>ions made above lead to another description of the st<strong>at</strong>e of the<br />

system. If we are close in temper<strong>at</strong>ure to the phase transition, it is possible to<br />

consider the st<strong>at</strong>e of the system as a dynamic mixture of the stable phase embedded<br />

in which we have regions of the unstable phase. Such a st<strong>at</strong>e clearly has<br />

a higher internal energy, but since there are many possibilities for the unstable<br />

regions, the entropy in this st<strong>at</strong>e is also higher, and the Helmholtz free energy<br />

will be lower! Therefore, the correct description of the st<strong>at</strong>e of the system close<br />

to the critical point is such a mixed st<strong>at</strong>e!<br />

A microscopic description of such a dynamic mixed st<strong>at</strong>e is extremely difficult,<br />

but in thermodynamics we can derive some general st<strong>at</strong>ements th<strong>at</strong> the<br />

free energy of this st<strong>at</strong>e has to obey. On an absolute scale the free energy of the<br />

mixed st<strong>at</strong>e is always close to the free energy of the stable st<strong>at</strong>e, and the difference<br />

between these two free energies is small. Close to the transition, however,<br />

this difference in free energy is much larger than the difference between the free<br />

energies of the two homogeneous st<strong>at</strong>es, and for such temper<strong>at</strong>ure values this is<br />

the dominant free energy. Since observable variables are determined by slopes of<br />

the free energy, they can be quite different from the values for the homogeneous<br />

st<strong>at</strong>es.<br />

In this chapter we start with a discussion of the phase transition in terms<br />

193


194 CHAPTER 5. CRITICAL EXPONENTS.<br />

of critical exponents, a measurement how response functions diverge near the<br />

critical point. We then calcul<strong>at</strong>e these exponents in a mean field description.<br />

The values we obtain in this way are close to the observed values, but differences<br />

between mean field theory and experiment are significant. Next, we discuss<br />

the free energy of the mixed st<strong>at</strong>e, and describe how we can derive some of<br />

its properties based on the idea of scaling. Here we use the physics th<strong>at</strong> is<br />

important near a phase transition to predict how the general form of the free<br />

energy should be. This gives us several rel<strong>at</strong>ions between critical exponents,<br />

which are all observed experimentally. The scaling idea is then extended to<br />

another form called hyper-scaling. We combine the ideas of scaling with the<br />

theory of Landau and Ginzburg and show the validity of the l<strong>at</strong>ter when we<br />

change the number of dimensions in the system. Finally, we apply the scaling<br />

idea to time as well, and are then able to discuss the critical exponents for<br />

transport parameters.<br />

5.1 Introduction.<br />

Critical point is essential singularity.<br />

In this chapter we will discuss the n<strong>at</strong>ure of second order phase transitions<br />

in more detail. Our language is a bit sloppy here, and by second order we<br />

really mean all phase transitions th<strong>at</strong> end in a critical point. We have argued<br />

before th<strong>at</strong> the behavior <strong>at</strong> the critical point determines wh<strong>at</strong> happens away<br />

from the critical point. Therefore, in this chapter we focus on properties <strong>at</strong><br />

the critical point. Response functions will diverge, and near the critical point<br />

this divergent behavior overshadows any background behavior. It is not clear a<br />

priori, however, wh<strong>at</strong> near is. In terms on temper<strong>at</strong>ure, this could be 100K or<br />

1µK.<br />

Wh<strong>at</strong> causes this divergent behavior?<br />

Wh<strong>at</strong> is meant by background?<br />

Descriptions of behavior <strong>at</strong> the critical point.<br />

In chapter 3 we saw th<strong>at</strong> we can model a phase transition if we know the<br />

Gibbs free energy for the two distinct phases. The coexistence curve was found<br />

by comparing these energies. In this scenario the Gibbs free energies of the two


5.1. INTRODUCTION. 195<br />

phases were quite uncorrel<strong>at</strong>ed, there was no hint th<strong>at</strong> they actually came from<br />

the same system. Near a phase transition, in many cases nothing diverged. We<br />

either were in one type of analytical behavior or in another. Landau theory<br />

improved on this by giving these two free energies a common origin. We defined<br />

fequilibrium(T ) = min<br />

m f(T ; m) (5.1)<br />

which rel<strong>at</strong>ed the free energy in each phase to a common model, which includes<br />

an internal parameter. This procedure again gave two distinct analytical forms<br />

of the free energy, one above and one below the transition. The n<strong>at</strong>ure of the<br />

construction of these model energies, however, now often introduces divergent<br />

behavior <strong>at</strong> the critical point.<br />

Background.<br />

We are now able to define wh<strong>at</strong> we mean by a background. Expressions<br />

like Landau theory give us a uniform value of the order parameter, constant in<br />

time. This would be the correct average behavior of the system far away from<br />

a phase transition. We can use a theoretical expression like Landau theory to<br />

extrapol<strong>at</strong>e this average behavior to the point of the phase transition. This<br />

extrapol<strong>at</strong>ion uses the continuous behavior of thermodynamic functions and<br />

incorpor<strong>at</strong>es all uniform behavior of the system.<br />

At non zero temper<strong>at</strong>ures we will have thermal fluctu<strong>at</strong>ions. Far away from<br />

a phase transition these thermal fluctu<strong>at</strong>ions are small. They average out to<br />

zero, both in space and time. They are, however, time dependent phenomena.<br />

There is energy stored in these fluctu<strong>at</strong>ions, but far from a phase transition this<br />

energy is small. Near the transition, however, this energy becomes large and<br />

eventually domin<strong>at</strong>es the system.<br />

Characteriz<strong>at</strong>ion of the singularities.<br />

Any realistic theory of a second order phase transition has to include the<br />

effects of fluctu<strong>at</strong>ions near the critical point. These fluctu<strong>at</strong>ions give rise to divergent<br />

behavior in response functions, different from mean field theory, and we<br />

therefore need to develop a m<strong>at</strong>hem<strong>at</strong>ical description how to quantify differences<br />

in divergent behavior. The main characteristic of a divergence is the r<strong>at</strong>e with<br />

which certain quantities go to infinity. This idea leads us then autom<strong>at</strong>ically to<br />

the idea of critical exponents, which are the parameters th<strong>at</strong> show how fast the<br />

divergence is.<br />

The four common exponents.<br />

We will define the four common critical exponents in terms of the model of<br />

a magnetic phase transition used before. This is just an easy way to describe<br />

wh<strong>at</strong> is happening, and similar procedures are valid for any second order phase


196 CHAPTER 5. CRITICAL EXPONENTS.<br />

transition. The order parameter for a second order phase transition approaches<br />

zero when the temper<strong>at</strong>ure approaches the critical temper<strong>at</strong>ure. Near the phase<br />

transition we have<br />

T < Tc : m(T, H = 0) ∝ (Tc − T ) β<br />

(5.2)<br />

The power β in this rel<strong>at</strong>ion is called a critical exponent. Note th<strong>at</strong> the rel<strong>at</strong>ion<br />

above only holds when there is no applied external magnetic field.<br />

If we apply a small magnetic field for a temper<strong>at</strong>ure equal to the critical<br />

temper<strong>at</strong>ure, we find another critical exponent δ:<br />

H → 0 : m(T = Tc, H) ∝ H 1<br />

δ (5.3)<br />

The magnetic susceptibility χ = ( )<br />

∂M<br />

∂H T also often diverges <strong>at</strong> Tc, due to the<br />

instability of a system near a phase transition.<br />

Critical exponents γ and γ ′ are defined by:<br />

T > Tc : χ(T, H = 0) ∝ (T − Tc) −γ<br />

T < Tc : χ(T, H = 0) ∝ (Tc − T ) −γ′<br />

(5.4)<br />

(5.5)<br />

Experiments suggest th<strong>at</strong> in general γ = γ ′ . Our example is a magnetic system,<br />

but can be generalized to an arbitrary phase transition by replacing M with the<br />

order parameter for th<strong>at</strong> transition and H by the corresponding force.<br />

The specific he<strong>at</strong> ( the he<strong>at</strong> capacity per unit volume) behaves in a special<br />

way:<br />

T > Tc : C(T, H = 0) ∝ (T − Tc) −α<br />

T < Tc : C(T, H = 0) ∝ (Tc − T ) −α′<br />

Experiment shows again th<strong>at</strong> in general α = α ′ .<br />

Correl<strong>at</strong>ion length.<br />

Are these critical exponents independent variables?<br />

(5.6)<br />

(5.7)<br />

In the previous chapter we have seen th<strong>at</strong> including fluctu<strong>at</strong>ions in a description<br />

of a m<strong>at</strong>erial leads to the idea of a correl<strong>at</strong>ion length ξ. This correl<strong>at</strong>ion<br />

length ξ is associ<strong>at</strong>ed with a parameter ν:<br />

T > Tc : ξ(T ) ∝ (T − Tc) −ν<br />

T < Tc : ξ(T ) ∝ (Tc − T ) −ν′<br />

(5.8)<br />

(5.9)


5.1. INTRODUCTION. 197<br />

Again, we will usually find ν = ν ′ .<br />

Universality.<br />

The powers in all these rel<strong>at</strong>ions are called critical exponents. It turns out<br />

th<strong>at</strong> critical exponents are a unique characteriz<strong>at</strong>ion of phase transitions. If two<br />

seemingly different phase transitions have exactly the same critical exponents,<br />

the underlying physics must be similar. This is an example of universality. We<br />

use the values of the critical exponents to classify second order phase transitions,<br />

and we find th<strong>at</strong> there are only a few groups. For example, the critical exponents<br />

in the magnetic phase transitions for iron and nickel are the same, which tells us<br />

th<strong>at</strong> the same physics is <strong>at</strong> work. Of course, the critical temper<strong>at</strong>ure has a very<br />

different value, which simply reflects the fact th<strong>at</strong> we are dealing with different<br />

m<strong>at</strong>erials. But if we find an other magnetic phase transition with different<br />

values of the critical exponents, we know th<strong>at</strong> something different is going on,<br />

and th<strong>at</strong> we will need different models to describe this new m<strong>at</strong>erial.<br />

Critical exponents for the liquid to gas transition.<br />

Critical exponents can also be defined for a liquid to gas phase transition.<br />

The intensive parameters controlling the phase transition are pressure and temper<strong>at</strong>ure<br />

and Tc in the formulas above must be set equal to the temper<strong>at</strong>ure <strong>at</strong><br />

the critical point. The coexistence curve pco(T ) has the same role as the line<br />

H(T ) = 0 in the magnetic phase transition, it represents a discontinuity. If we<br />

expose a ferromagnetic m<strong>at</strong>erial to a magnetic field, cool it below the critical<br />

temper<strong>at</strong>ure, and then reduce the field to zero, the final result will de different<br />

for positive or neg<strong>at</strong>ive fields. The line (T, H = 0) from T = 0 to T = Tc represents<br />

all points in phase space where this discontinuity occurs. In the same way,<br />

if we apply pressure to a gas/liquid, cool it below the critical temper<strong>at</strong>ure, and<br />

then have the pressure approach pco(T ), we observe a discontinuity depending<br />

on the p<strong>at</strong>h in (p,T) space.<br />

Superconductivity.<br />

A more complex example of a second order phase transition is found in<br />

superconductivity. In this case the order parameter is the pseudo wave function<br />

ψ(⃗r) representing the super-conducting electrons. The density |ψ(⃗r)| 2 represents<br />

the density of super-conducting electrons <strong>at</strong> point ⃗r. There is a bit of a problem,<br />

since we need to know wh<strong>at</strong> the force corresponding to the order parameter is.<br />

If we use the square of the pseudo wave function, the generalized force obviously<br />

is some sort of chemical potential. It is not easy to find out which generalized<br />

force corresponds to the pseudo wave functions itself.<br />

Landau model for superconductivity.


198 CHAPTER 5. CRITICAL EXPONENTS.<br />

Deep inside a superconductor we assume th<strong>at</strong> we can describe the free energy<br />

by<br />

f(T ; ψ) = fnormal(T ) + 1<br />

2 α(T )|ψ(⃗r)|2 + 1<br />

β(T )|ψ(⃗r)|4<br />

4<br />

(5.10)<br />

and in order to model a second order phase transition we assume th<strong>at</strong> α(T )<br />

changes sign <strong>at</strong> Tc and th<strong>at</strong> β(T ) > 0. Therefore we find below Tc th<strong>at</strong><br />

|ψ(⃗r)| 2 α(T )<br />

= −<br />

β(T )<br />

(5.11)<br />

The difference in free energy between the two st<strong>at</strong>es, the normal st<strong>at</strong>e and<br />

the super-conducting st<strong>at</strong>e, is given by<br />

fsc − fnormal = − 1 α<br />

4<br />

2<br />

β<br />

(5.12)<br />

and this energy can be rel<strong>at</strong>ed to the critical field Hc(T ) for a type I superconductor<br />

by setting this equal to the energy density, because a type I superconductor<br />

expels all magnetic fields. This gives<br />

− H2 c (T )<br />

8π<br />

α<br />

= −1<br />

4<br />

2<br />

β<br />

Why is this a good assumption?<br />

(5.13)<br />

Next we assume th<strong>at</strong> in the description of α near Tc we only need to include<br />

the linear terms, and hence th<strong>at</strong> α(T ) ∝ (T − Tc). This gives<br />

which is indeed observed, and<br />

which agrees with the older theory due to London.<br />

Fluctu<strong>at</strong>ions.<br />

Hc(T ) ∝ Tc − T (5.14)<br />

|ψ(⃗r)| 2 ∝ Tc − T (5.15)<br />

Our next task is to include fluctu<strong>at</strong>ions in this model. The pseudo wave<br />

function is described by a Schrödinger type of equ<strong>at</strong>ion, and the corresponding<br />

energy, including a magnetic field (using a vector potential ⃗ A) is<br />

1<br />

2m∗ (<br />

<br />

<br />

¯h<br />

ı ⃗ ∇ − e∗<br />

c ⃗ ) <br />

<br />

A ψ<br />

<br />

2<br />

(5.16)


5.1. INTRODUCTION. 199<br />

where m ∗ is the mass of the super-conducting species (i.e. a Cooper pair) and<br />

e ∗ its charge. Even though deep inside the superconductor there is no magnetic<br />

field, we need to include the vector potential in order to describe the interface<br />

between the super-conducting and normal regions. This form of the extra free<br />

energy term is very similar to the Ginzburg term discussed before.<br />

Next, we separ<strong>at</strong>e the pseudo wave function into a magnitude and phase by<br />

ψ = |ψ|e ıϕ which leads to the following form of the energy term:<br />

1<br />

2m∗ (<br />

¯h 2 ( ) (<br />

2<br />

∇|ψ| ⃗ + ¯h ⃗ ∇ϕ − e∗<br />

c ⃗ ) 2<br />

A |ψ| 2<br />

)<br />

(5.17)<br />

where all cross terms disappear since they are purely imaginary. The first term<br />

depends only on the magnitude of the order parameter and is the same as the<br />

term introduced in the last chapter. The second term reflects a gauge invariance,<br />

and is a pure quantum mechanical effect.<br />

Phase of the pseudo wave function.<br />

If there are no external fields and ⃗ A = 0 the minimum in the free energy<br />

occurs when ⃗ ∇ϕ = 0 or ϕ = constant, which means th<strong>at</strong> ϕ is not important in<br />

this case. In general we have<br />

rel<strong>at</strong>ing the phase and the vector potential.<br />

Correl<strong>at</strong>ion length.<br />

¯h ⃗ ∇ϕ = e∗<br />

c ⃗ A (5.18)<br />

The results from the previous chapter allow us to say th<strong>at</strong> the correl<strong>at</strong>ion<br />

length in this case is given by<br />

√<br />

¯h<br />

ξ(T ) =<br />

2<br />

2m∗α(T )<br />

(5.19)<br />

which diverges in the usual way near the critical temper<strong>at</strong>ure. The interpret<strong>at</strong>ion<br />

of this correl<strong>at</strong>ion length is simple. If we have an interface between a<br />

normal st<strong>at</strong>e and a super-conducting st<strong>at</strong>e, the domain wall separ<strong>at</strong>ing these<br />

two st<strong>at</strong>es is of thickness ξ.<br />

Another important length scale is the length λL(T ) over which a magnetic<br />

field dies out inside the superconductor due to the flux expulsion. A comparison<br />

of the values of λL(T ) and ξ allows us to decide if a superconductor is of type<br />

I or type II.


200 CHAPTER 5. CRITICAL EXPONENTS.<br />

5.2 Mean field theory.<br />

Meaning of different parameters.<br />

In the study of phase transitions we distinguish two types of parameters<br />

th<strong>at</strong> characterize a transition. When we consider the terms in the work done<br />

on a sample, we have products like HM and pV. We also have he<strong>at</strong> exchange,<br />

corresponding to a term TS. A phase transition is first characterized by a critical<br />

value in the temper<strong>at</strong>ure, Tc, below which temper<strong>at</strong>ure the sample is ordered<br />

when other fields are not present, i.e. H=0, p=0, etc. But we need also consider<br />

the roles of the other external controls we have on the sample. For example,<br />

in a magnetic phase transition we also have a critical value of the field, Hc,<br />

which is often equal to zero. The critical point in the (T, H) plane is (Tc, Hc).<br />

If we reduce the temper<strong>at</strong>ure through the critical value with a field slightly<br />

higher than Hc we end up with one type of magnetiz<strong>at</strong>ion, and if we lower the<br />

temper<strong>at</strong>ure with a field slightly below Hc we obtain the opposite magnetiz<strong>at</strong>ion.<br />

If we also allow pressure to play a role, we find th<strong>at</strong> the critical point is given<br />

by three parameters, Tc, Hc, and pc.<br />

The position of the critical point depends on the strength of the interactions,<br />

and on the type of interactions. Even though iron and nickel are both<br />

ferromagnetic m<strong>at</strong>erials, the critical temper<strong>at</strong>ures are very different. In terms<br />

of models, they have different exchange interactions.<br />

Phase transitions are also characterized by critical exponents, as defined<br />

earlier. As it turns out, these critical exponents do not depend on the strength<br />

of the interactions, they only depend on the form of the Hamiltonian, or<br />

on the n<strong>at</strong>ure of the interactions. Therefore, systems with similar values of<br />

the critical exponents are similar in n<strong>at</strong>ure, or have similar physics. Critical<br />

exponents are a tool to classify different types of phase transitions, or to sort<br />

them into different universality classes. In the following we discuss an important<br />

example.<br />

Landau and mean field.<br />

Wh<strong>at</strong> does ”form of the Hamiltonian” mean?<br />

The second order phase transition we have studied as an example in Landau<br />

theory has a special meaning. If we ignore fluctu<strong>at</strong>ions, and only include the<br />

lowest order terms, the external magnetic field is given by (with T < Tc)<br />

H =<br />

( )<br />

∂f<br />

= b(T )m + cm<br />

∂m T<br />

3<br />

(5.20)


5.2. MEAN FIELD THEORY. 201<br />

which can be written as<br />

b0T m = H + m(b0Tc − cm 2 ) (5.21)<br />

and near the critical temper<strong>at</strong>ure this is approxim<strong>at</strong>ely b0T m = H + mb0Tc.<br />

In free space the magnetiz<strong>at</strong>ion and field are proportional. In a paramagnetic<br />

substance the rel<strong>at</strong>ion between the magnetiz<strong>at</strong>ion and the magnetic field is given<br />

by Curie’s law m = α H<br />

T . The equ<strong>at</strong>ion 5.21 looks like Curie’s law, except th<strong>at</strong><br />

the magnetic field is modified. Because a magnetiz<strong>at</strong>ion is present there is an<br />

additional effective magnetic field H ′ = m(b0Tc −cm2 ). This can be interpreted<br />

as follows. Because a magnetiz<strong>at</strong>ion is already present, each sub-system in our<br />

large system experiences an additional magnetic field which is simply rel<strong>at</strong>ed<br />

to the average magnetiz<strong>at</strong>ion in the system as a whole. One can think of the<br />

surroundings of a sub-system providing such a magnetic field, based on the<br />

average value of the magnetiz<strong>at</strong>ion in the system. This is called a mean field<br />

approach. It ignores the effect th<strong>at</strong> there are fluctu<strong>at</strong>ions in the magnetiz<strong>at</strong>ion.<br />

Critical exponents in mean field.<br />

The critical exponents for Landau’s theory in this mean field approach follow<br />

immedi<strong>at</strong>ely from the equ<strong>at</strong>ions we derived before:<br />

χ −1 (T, H) =<br />

m ∝ (Tc − T ) 1<br />

2 ⇒ β = 1<br />

2<br />

(5.22)<br />

( 2 ∂ f<br />

∂m2 )<br />

∝ |T − Tc| ⇒ γ = γ<br />

T<br />

′ = 1 (5.23)<br />

T = Tc, f ∝ m 4 ⇒ δ = 3 (5.24)<br />

The value of α cannot be determined, since we did not specify a(T). Since<br />

a(T) is actually independent of the phase transition, it is reasonable to assume<br />

th<strong>at</strong> it does not diverge <strong>at</strong> Tc, and in th<strong>at</strong> case we find<br />

α = α ′ = 0 (5.25)<br />

Any phase transition with the same critical exponents as we listed here has the<br />

same physics incorpor<strong>at</strong>ed. Any model with these critical exponents is therefore<br />

essentially a mean field model!<br />

Example values.<br />

The following table gives some examples of measured values of critical exponents<br />

and theoretical results:


202 CHAPTER 5. CRITICAL EXPONENTS.<br />

type system Tc α β γ δ<br />

ferromagnetic Ni 632 -0.10 0.33 1.32 4.2<br />

F e 1044 -0.12 0.34 1.33<br />

anti-ferromagnetic RbMnF3 83 -0.14 0.32 1.40<br />

liquid-gas Co2 304 0.12 0.34 1.20 4.2<br />

Xe 290 0.08 0.34 1.20 4.4<br />

ordered alloy CoZn 739 0.31 1.25<br />

mean field 0 0.5 1.0 3.0<br />

Ising, d=2 0 0.125 1.75 15<br />

Ising, d=3 0.12 0.31 1.25 5.0<br />

Heisenberg -0.14 0.38 1.38<br />

If mean field theory would be valid for all systems, all systems would have<br />

the same values for the critical exponents. In some sense this is obeyed. They<br />

are indeed quite close, but distinct from mean field. The fact th<strong>at</strong> the critical<br />

exponents do not vary wildly is due to the fact th<strong>at</strong> in many systems mean field<br />

theory is a good first approxim<strong>at</strong>ion.<br />

A very different set of values is obtained in the 2 dimensional Ising model.<br />

This shows the importance of the dimensionality of the system, and hence we<br />

expect big differences between the properties of thin films and bulk solids, which<br />

is indeed observed. The difference between the 3 dimensional Ising model and<br />

the Heisenberg model is not very large, but large enough th<strong>at</strong> experiments can<br />

distinguish between the validity of the two.<br />

In experimental situ<strong>at</strong>ions we often study thin films, and the question always<br />

arises if these films are two or three dimensional. The answer to this question<br />

is easy. One length scale of importance is the thickness of the film, Tfilm. The<br />

other quantity of interest is the correl<strong>at</strong>ion length ξ(T ). If we have ξ(T ) < Tfilm,<br />

even in the direction perpendicular to the film we have many domains, or regions<br />

of different organized behavior, and the film is three dimensional. On the other<br />

hand, if ξ(T ) > Tfilm the film is always coherent over the whole thickness, and<br />

we have a two dimensional system. In the thermodynamic limit we assume th<strong>at</strong><br />

the extend of the film in the plane is infinite, and in th<strong>at</strong> direction we always have<br />

a correl<strong>at</strong>ion length smaller than the dimensions of the film. This argument<br />

therefore shows th<strong>at</strong> a thin film near the critical temper<strong>at</strong>ure always<br />

behaves as a two dimensional system. This situ<strong>at</strong>ion is more complic<strong>at</strong>ed<br />

when we have two different types of correl<strong>at</strong>ion length to consider, in th<strong>at</strong> case<br />

we have to ask the question large or small for each of those lengths.<br />

Example van der Waals.<br />

Wh<strong>at</strong> is near?<br />

In Chapter 3 we studied a model of the liquid to gas phase transition devel-


5.2. MEAN FIELD THEORY. 203<br />

oped by van der Waals. This model was very instructive, and explained many<br />

characteristics of the phase transition. It essentially captured the behavior using<br />

two parameters, a and b. Based on the values for a and b we found the critical<br />

temper<strong>at</strong>ure and pressure. Hence, it is also possible to use the critical temper<strong>at</strong>ure<br />

and pressure as parameters describing the phase transition. There are no<br />

parameters left, and therefore the critical exponents should all be determined<br />

by the form of the equ<strong>at</strong>ion!<br />

Parameter free van der Waals equ<strong>at</strong>ion.<br />

Once the critical temper<strong>at</strong>ure and pressure are known, the van der Waals<br />

equ<strong>at</strong>ion should be unique. This leads us to the introduction of the following<br />

rel<strong>at</strong>ive variables:<br />

T − Tc<br />

θ =<br />

Tc<br />

p − pc<br />

π =<br />

pc<br />

V − Vc<br />

υ =<br />

and in terms of these variables the van der Waals equ<strong>at</strong>ion is<br />

[<br />

(1 + π) +<br />

3<br />

(1 + υ) 2<br />

Vc<br />

(5.26)<br />

(5.27)<br />

(5.28)<br />

]<br />

[3(1 + υ) − 1] = 8(1 + θ) (5.29)<br />

which is indeed parameter-free. Since we are interested in the behavior near<br />

the critical point, where all scaled parameter values are small, we rewrite this<br />

equ<strong>at</strong>ion as follows by multiplying by (1 + υ) 2 :<br />

or<br />

or<br />

[ (1 + π)(1 + υ) 2 + 3 ] [2 + 3υ] = 8(1 + θ)(1 + υ) 2<br />

(5.30)<br />

[ 4 + 2υ + υ 2 + π + 2πυ + πυ 2 ] [2 + 3υ] = 8(1 + θ)(1 + 2υ + υ 2 ) (5.31)<br />

8+16υ +8υ 2 +2π +7πυ +8πυ 2 +3υ 3 +3πυ 3 = 8+16υ +8υ 2 +8θ +16θυ +8θυ 2<br />

or<br />

π ( 2 + 7υ + 8υ 2 + 3υ 3) + 3υ 3 = θ ( 8 + 16υ + 8υ 2)<br />

(5.32)<br />

(5.33)<br />

which is a good start for the deriv<strong>at</strong>ion of the values of the critical exponents.


204 CHAPTER 5. CRITICAL EXPONENTS.<br />

Reduced pressure expansion.<br />

In order to analyze the behavior near the critical point, we expand the<br />

reduced pressure as a function of reduced temper<strong>at</strong>ure and reduced volume<br />

near the critical point and find th<strong>at</strong>:<br />

π ≈ 4θ − 6θυ + 9θυ 2 − 3<br />

2 υ3<br />

(5.34)<br />

where in the third order volume term the small θ term is ignored, and the<br />

higher order volume terms are ignored. This equ<strong>at</strong>ion gives a good and valid<br />

description of the behavior near the critical point.<br />

The partial deriv<strong>at</strong>ive of the pressure versus volume is:<br />

( )<br />

∂π<br />

≈ −6θ + 18θυ −<br />

∂υ θ<br />

9<br />

2 υ2<br />

(5.35)<br />

and equ<strong>at</strong>ing this to zero has solutions if 18 2 θ 2 ≥ 108θ or θ ≤ 0 for the solution<br />

close to the critical point, as expected.<br />

Order parameter.<br />

For θ ≤ 0 and close to zero we can find the volume of the liquid and the gas<br />

phase from the equ<strong>at</strong>ions:<br />

and<br />

The first equ<strong>at</strong>ion is equivalent to:<br />

π(υl) = π(υg) (5.36)<br />

0 =<br />

∫ υg<br />

υl<br />

υdπ (5.37)<br />

−6θυl + 9θυ 2 l − 3<br />

2 υ3 l = −6θυg + 9θυ 2 g − 3<br />

2 υ3 g<br />

while the last equ<strong>at</strong>ion is equivalent to:<br />

or<br />

or<br />

0 =<br />

0 =<br />

∫ υg<br />

υl<br />

∫ υg<br />

υl<br />

υ<br />

(5.38)<br />

( )<br />

∂π<br />

dυ (5.39)<br />

∂υ θ<br />

υ(−6θ + 18θυ − 9<br />

2 υ2 )dυ (5.40)<br />

−3θυ 2 l + 6θυ 3 l − 9<br />

8 υ4 l = −3θυ 2 g + 6θυ 3 g − 9<br />

8 υ4 g<br />

(5.41)


5.2. MEAN FIELD THEORY. 205<br />

Close to the critical point the middle term in each equ<strong>at</strong>ion can be ignored,<br />

because it is small compared to the first term. We cannot ignore the third term,<br />

since it does not contain the temper<strong>at</strong>ure, and we do not yet know the rel<strong>at</strong>ive<br />

scaling of all quantities. As a consequence, we find th<strong>at</strong> we need to solve:<br />

and<br />

4θυl + υ 3 l = 4θυg + υ 3 g<br />

8θυ 2 l + 3υ 4 l = 8θυ 2 g + 3υ 4 g<br />

(5.42)<br />

(5.43)<br />

How to solve these equ<strong>at</strong>ions? First of all, note th<strong>at</strong> we expect υl < 0 < υg.<br />

The two previous equ<strong>at</strong>ions can be rewritten in the form:<br />

and<br />

4θ(υl − υg) = υ 3 g − υ 3 l<br />

(5.44)<br />

8θ(υ 2 l − υ 2 g) = 3(υ 4 g − υ 4 l ) (5.45)<br />

The first has solutions υl = υg, which is not physical because υl < 0 < υg.<br />

Hence we are left with<br />

The second equ<strong>at</strong>ion has solutions<br />

−4θ = υ 2 g + υlυgυ 2 l<br />

υ 2 l = υ 2 g<br />

(5.46)<br />

(5.47)<br />

or υl = −υg. Together with the first equ<strong>at</strong>ion this gives the solutions according<br />

to<br />

υg = −υl = 2 √ |θ| (5.48)<br />

which are only valid for θ < 0, as expected. The other solution of the second<br />

equ<strong>at</strong>ion is<br />

8θ = 3(υ 2 g + υ 2 l ) (5.49)<br />

which cannot be right since it would imply θ > 0.<br />

The order parameter ∆υ is defined as the difference between the volumes,<br />

and we have<br />

along the coexistence curve given by<br />

He<strong>at</strong> capacity.<br />

∆υ = 4 √ |θ| (5.50)<br />

πcoex = 4θ (5.51)


206 CHAPTER 5. CRITICAL EXPONENTS.<br />

We have seen in chapter 3 th<strong>at</strong> the he<strong>at</strong> capacity as a function of temper<strong>at</strong>ure<br />

always shows a finite jump. There is no divergence, but the values of the he<strong>at</strong><br />

capacity are not the same coming from below or above the critical temper<strong>at</strong>ure.<br />

This is simply rel<strong>at</strong>ed to the fact th<strong>at</strong> we have a continuous p<strong>at</strong>h from the gas<br />

to the liquid st<strong>at</strong>e following a van der Waals curve. The Gibbs free energy of<br />

the gas and liquid are the same <strong>at</strong> the coexistence curve, but the slope as a<br />

function of temper<strong>at</strong>ure changes.<br />

The Gibbs free energy follows from<br />

g = g0(θ) +<br />

which leads to the entropy<br />

−s =<br />

∫ υ<br />

0<br />

υdπ = g0(θ) − 3θυ 2 + 6θυ 3 − 9<br />

8 υ4<br />

( )<br />

∂g<br />

= g<br />

∂θ π<br />

′ 0(θ) − 3υ 2 + 6υ 3 (<br />

+ −6θυ + 18θυ 2 − 9<br />

2 υ3<br />

) ( )<br />

∂υ<br />

∂θ π<br />

and a specific he<strong>at</strong><br />

−c = θ<br />

(<br />

θ −6θ + 36θυ − 27<br />

2 υ2<br />

( )<br />

∂s<br />

= θg<br />

∂θ π<br />

′′<br />

0 (θ) + 2θ ( −6υ + 18υ 2) ( )<br />

∂υ<br />

+<br />

∂θ π<br />

) ( ) 2 (<br />

∂υ<br />

+θ −6θυ + 18θυ<br />

∂θ π<br />

2 − 9<br />

2 υ3<br />

) ( 2 ∂ υ<br />

∂θ2 )<br />

π<br />

(5.52)<br />

(5.53)<br />

(5.54)<br />

which does not diverge, but is discontinuous (i.e. does not have the same value<br />

for υl and υg. Near the transition we can therefore model the he<strong>at</strong> capacity by<br />

θ > 0 : C(θ, π = πcoex) ∝ θ −α<br />

θ < 0 : C(θ, π = πcoex) ∝ |θ| −α′<br />

and the only way to get a discontinuity is to have<br />

Order parameter.<br />

From equ<strong>at</strong>ion 5.50 gives immedi<strong>at</strong>ely th<strong>at</strong><br />

Order parameter versus field.<br />

(5.55)<br />

(5.56)<br />

α = α ′ = 0 (5.57)<br />

β = 1<br />

2<br />

(5.58)


5.2. MEAN FIELD THEORY. 207<br />

This response function follows from (leaving out a factor<br />

not change the value of the critical exponent):<br />

−κ −1 =<br />

which we need to follow along πcoex = 4θ or<br />

( )<br />

∂π<br />

= −6θ + 18θυ −<br />

∂υ θ<br />

9<br />

2 υ2<br />

0 = −6θυ + 9θυ 2 − 3<br />

2 υ3<br />

V pc , which does<br />

Vc<br />

(5.59)<br />

(5.60)<br />

which can be used (after dividing by υ ) to elimin<strong>at</strong>e the last term, and we get<br />

θ < 0 : κ −1 = −12θ + 9θυ (5.61)<br />

in which only the first term is important near the critical point. Similarly, above<br />

the critical temper<strong>at</strong>ure we follow the constant volume curve and find:<br />

which leads immedi<strong>at</strong>ely to<br />

θ > 0 : κ −1 = +6θ (5.62)<br />

γ = γ ′ = 1 (5.63)<br />

Volume versus pressure <strong>at</strong> the critical temper<strong>at</strong>ure.<br />

The equ<strong>at</strong>ion here is simple, we have<br />

which shows immedi<strong>at</strong>ely th<strong>at</strong><br />

Van der Waals is mean field.<br />

π = − 3<br />

2 υ3<br />

(5.64)<br />

δ = 3 (5.65)<br />

We have found th<strong>at</strong> the critical exponents in a van der Waals model of<br />

a phase transition are the same as the mean field exponents. Therefore, the<br />

van der Waals model is a mean field model. This can be explained easily in<br />

a microscopic model, where the parameters a and b only describe the average<br />

interaction between molecules, or the interaction between one molecule and<br />

its average (mean) environment. The van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is no<br />

exception. In fact, any analytic equ<strong>at</strong>ion of st<strong>at</strong>e model of a phase transition<br />

gives the same mean field exponents, since in rel<strong>at</strong>ive units the form of the<br />

equ<strong>at</strong>ion near the critical point is always like equ<strong>at</strong>ion 5.34!


208 CHAPTER 5. CRITICAL EXPONENTS.<br />

5.3 Model free energy near a critical point.<br />

Fluctu<strong>at</strong>ions.<br />

By now it will be clear th<strong>at</strong> any realistic description of a phase transition<br />

near a critical point will have to include the effects of fluctu<strong>at</strong>ions. Without<br />

th<strong>at</strong>, we only have analytical theories with discontinuities <strong>at</strong> the transition,<br />

which lead to mean field values of the critical exponents. In order to illustr<strong>at</strong>e<br />

the approach to the problem of including fluctu<strong>at</strong>ions, we will again study a<br />

magnetic phase transition, but the procedure can be generalized to arbitrary<br />

systems.<br />

Scaled variables.<br />

The important variables which we can use to change the st<strong>at</strong>e of our model<br />

system are the temper<strong>at</strong>ure T and the magnetic field H. The critical point is<br />

given by the values Tc and Hc = 0. In order to focus on the behavior near the<br />

critical point we used scaled values of these variables:<br />

and<br />

T − Tc<br />

θ =<br />

Tc<br />

(5.66)<br />

h = H (5.67)<br />

where the scaling in the last case is simple, since we cannot divide by zero.<br />

Model energy.<br />

The important free energy in this case is a analogue to the Gibbs free energy<br />

(the magnetic Gibbs free energy) defined by<br />

g(θ, h) = 1<br />

(U − T S − MH) (5.68)<br />

N<br />

where we also divided by the total amount of m<strong>at</strong>erial to get an energy density.<br />

One important contribution to this free energy is the normal equilibrium thermodynamical<br />

form we have discussed before. This term only leads to continuous<br />

behavior, without any divergent behavior in the energy. The previous section<br />

showed th<strong>at</strong> the free energy for a van der Waals model is smooth, for example.<br />

This term is the background energy. The free energy, however, also contains a<br />

term describing the fluctu<strong>at</strong>ions in the system. This is the term we do not know<br />

how to write down formally, but we can separ<strong>at</strong>e the effects:<br />

g(θ, h) = gbackground(θ, h) + gfluctu<strong>at</strong>ions(θ, h) (5.69)


5.3. MODEL FREE ENERGY NEAR A CRITICAL POINT. 209<br />

Analysis of fluctu<strong>at</strong>ion term.<br />

When we get closer to the critical temper<strong>at</strong>ure, the effect of fluctu<strong>at</strong>ions<br />

will domin<strong>at</strong>e the behavior of the system. These fluctu<strong>at</strong>ions will cause all<br />

kinds of divergencies in the response functions. We therefore assume th<strong>at</strong> near<br />

the critical point only the second term in the last equ<strong>at</strong>ion is important. The<br />

analysis of critical exponents is therefore determined by the second term, and<br />

from now on we ignore the first term and omit the subscript fluctu<strong>at</strong>ions. The<br />

main question is, wh<strong>at</strong> kind of general st<strong>at</strong>ements can we make about this energy<br />

term!<br />

Scaling.<br />

Near a critical point the correl<strong>at</strong>ion length becomes very large. Any other<br />

length scale we have in the system is not important anymore, and the correl<strong>at</strong>ion<br />

length determines the behavior of the system. Suppose we have a picture of a<br />

system close to a critical point. We see bubbles of one phase in another, and the<br />

average size of these bubbles is rel<strong>at</strong>ed to the correl<strong>at</strong>ion length. If we change<br />

the values of the temper<strong>at</strong>ure or field slightly, we obtain a different picture with<br />

a different average size of the bubbles. If we have an absolute measure of the<br />

length scale we can identify these pictures. But imagine th<strong>at</strong> both pictures are<br />

scaled and th<strong>at</strong> one unit of length in each picture corresponds to the appropri<strong>at</strong>e<br />

correl<strong>at</strong>ion length. In this case the average size of the bubbles will be the same<br />

in both pictures. Can we still identify the pictures? Can we tell them apart?<br />

The experimental answer seems to be no. Therefore, the free energy due to<br />

the fluctu<strong>at</strong>ions in both cases must be rel<strong>at</strong>ed, and the density is changed by<br />

simply accounting for the extra amount of m<strong>at</strong>erial in a scaled unit volume.<br />

M<strong>at</strong>hem<strong>at</strong>ically this is expressed by saying th<strong>at</strong> the free energy g (due to the<br />

fluctu<strong>at</strong>ions) near the critical point should be a homogeneous function of its<br />

variables.<br />

Homogeneous functions.<br />

Homogeneous functions are easiest to discuss when only one variable is involved.<br />

By definition, a function f(x) is called homogeneous if<br />

f(λx) = S(λ)f(x) (5.70)<br />

where S is called a scaling function. This rel<strong>at</strong>ion has to hold for all values of<br />

λ and x. In this simple case it is easy to derive wh<strong>at</strong> the possible homogeneous<br />

functions are. When we apply to successive transform<strong>at</strong>ions we get<br />

but also<br />

f((µλ)x) = S(µλ)f(x) (5.71)


210 CHAPTER 5. CRITICAL EXPONENTS.<br />

or<br />

f((µλ)x) = f(µ(λx)) = S(µ)S(λ)f(x) (5.72)<br />

S(µλ) = S(µ)S(λ) (5.73)<br />

for all values of λ and µ. If we now take the deriv<strong>at</strong>ive with respect to λ we<br />

obtain<br />

and similarly for µ:<br />

from which follows:<br />

or<br />

which means th<strong>at</strong><br />

S ′ (µλ)µ = S(µ)S ′ (λ) (5.74)<br />

S ′ (µλ)λ = S ′ (µ)S(λ) (5.75)<br />

S(µ)S ′ (λ)λ = µS ′ (µ)S(λ) (5.76)<br />

S ′ (λ)λ<br />

S(λ) = µS′ (µ)<br />

S(µ)<br />

S ′ (λ)λ<br />

S(λ)<br />

(5.77)<br />

= p (5.78)<br />

where p is a constant. Since S(1)=1 we can solve this immedi<strong>at</strong>ely and we get<br />

Form of the scaling functions.<br />

S(λ) = λ p<br />

(5.79)<br />

We arrive <strong>at</strong> the important conclusion th<strong>at</strong> scaling functions are simple powers<br />

of the variable, where the constant p can take any real value. In equ<strong>at</strong>ion<br />

5.70 we can take x=1 and find<br />

f(λ) = λ p f(1) (5.80)<br />

which shows th<strong>at</strong> the homogeneous functions of one variable are simple powers.<br />

For example, for the volume V of a cube with side a we find th<strong>at</strong> V (λa) =<br />

λ 3 V (a).<br />

Homogeneous function in many variables.<br />

Homogeneous functions of many variables are defined in a similar manner,<br />

where we already include the fact th<strong>at</strong> the scaling functions are powers. The


5.4. CONSEQUENCES OF SCALING. 211<br />

only change we make is where we put the scaling function, and in this case it is<br />

easier to have the simple multiplier variable λ in front of the function:<br />

f(λ p x, λ q y) = λf(x, y) (5.81)<br />

Since the powers p and q can take any real value, we need to restrict ourselves<br />

to positive values of lambda only.<br />

Scaling applied to the free energy.<br />

The basic idea th<strong>at</strong> the correl<strong>at</strong>ion length is the only important length scale<br />

near the critical point transl<strong>at</strong>es in the following st<strong>at</strong>ement for the energy density.<br />

Near the critical point θ = 0 , h = 0 we have<br />

g(λ p θ, λ q h) = λg(θ, h) (5.82)<br />

The motiv<strong>at</strong>ion is more or less as follows. Nearer to the critical point the energy<br />

density becomes smaller, since the energy stored in a fluctu<strong>at</strong>ion is spread<br />

out over a larger volume. So if we make θ and h smaller, the energy density<br />

becomes smaller. Scaling theory simple assumes th<strong>at</strong> this reduction in temper<strong>at</strong>ure<br />

and field is homogeneous, like in equ<strong>at</strong>ion 5.82.<br />

5.4 Consequences of scaling.<br />

The scaling rel<strong>at</strong>ion in equ<strong>at</strong>ion 5.82 has some very important ) consequences.<br />

For example, if we calcul<strong>at</strong>e the magnetiz<strong>at</strong>ion m = we find th<strong>at</strong><br />

and therefore if we set θ = 1 and h = 0 we obtain<br />

or with t = λ p<br />

( ∂g<br />

∂h<br />

λ q m(λ p θ, λ q h) = λm(θ, h) (5.83)<br />

m(λ p , 0) = λ 1−q m(1, 0) (5.84)<br />

m(t, 0) = t 1−q<br />

p m(1, 0) (5.85)<br />

Since m(1, 0) = 0 we find therefore th<strong>at</strong> m(t, 0) = 0 for all values of t and th<strong>at</strong><br />

there is no magnetiz<strong>at</strong>ion. But here we have to be careful. Since λ > 0 this<br />

only holds for t > 0, and now our conclusion is indeed correct. Scaling tells us<br />

th<strong>at</strong> if the magnetiz<strong>at</strong>ion is zero <strong>at</strong> one point above the critical temper<strong>at</strong>ure, it<br />

is zero everywhere above the critical temper<strong>at</strong>ure.<br />

Rel<strong>at</strong>ion to critical exponent β.<br />

In order to discuss the behavior below the critical temper<strong>at</strong>ure we need to<br />

start with θ = −1. This gives:<br />

θ


212 CHAPTER 5. CRITICAL EXPONENTS.<br />

or with t = −λ p<br />

m(−λ p , 0) = λ 1−q m(−1, 0) (5.86)<br />

m(t, 0) = (−t) 1−q<br />

p m(−1, 0) (5.87)<br />

In this case m(−1, 0) is not zero and we find th<strong>at</strong> the magnetiz<strong>at</strong>ion near the<br />

critical temper<strong>at</strong>ure follows a simple power law as a function of the reduced temper<strong>at</strong>ure,<br />

as expected. A comparison with the definition of the critical exponent<br />

β gives<br />

β =<br />

1 − q<br />

p<br />

(5.88)<br />

and hence the value of the critical exponent follows immedi<strong>at</strong>ely from the scaling<br />

parameters p and q in the energy! This is true for all critical exponents! Since<br />

we only have two independent parameters in the energy scaling, we see th<strong>at</strong><br />

only two values of the critical exponents are independent, and th<strong>at</strong> the others<br />

should de dependent!<br />

Critical exponent δ.<br />

From equ<strong>at</strong>ion 5.83 we see immedi<strong>at</strong>ely th<strong>at</strong><br />

and with h = 1 we obtain<br />

or by defining |H| = λ q we find<br />

or<br />

Critical exponent γ.<br />

to<br />

λ q m(0, λ q h) = λm(0, h) (5.89)<br />

m(0, λ q ) = λ 1−q m(0, 1) (5.90)<br />

m(0, |H|) = |H| 1−q<br />

q m(0, 1) (5.91)<br />

δ = q<br />

1 − q<br />

(5.92)<br />

The susceptibility follows from χ = ( )<br />

∂m<br />

∂h and the scaling rel<strong>at</strong>ion 5.83 leads<br />

θ<br />

λ 2q χ(λ p θ, λ q h) = λχ(θ, h) (5.93)<br />

In this equ<strong>at</strong>ion we use again h = 0, and again distinguish between temper<strong>at</strong>ures<br />

above and below the critical temper<strong>at</strong>ure. We find


5.4. CONSEQUENCES OF SCALING. 213<br />

and<br />

t > 0 : χ(t, 0) = t 1−2q<br />

p χ(1, 0) (5.94)<br />

t < 0 : χ(t, 0) = (−t) 1−2q<br />

p χ(−1, 0) (5.95)<br />

which shows th<strong>at</strong> the critical exponents are indeed the same above and below<br />

the critical point:<br />

γ = γ ′ =<br />

but th<strong>at</strong> the pre-factors are different.<br />

Specific he<strong>at</strong>.<br />

2q − 1<br />

p<br />

(5.96)<br />

( 2<br />

∂ g<br />

The specific he<strong>at</strong> follows from ch = −T ∂θ2 )<br />

where the temper<strong>at</strong>ure in<br />

h<br />

front is the real temper<strong>at</strong>ure, not the scaled one. This gives<br />

which then leads to<br />

and<br />

λ 2p ch(λ p θ, λ q h) = λch(θ, h) (5.97)<br />

t > 0 : ch(t, 0) = t 1−2p<br />

p ch(1, 0) (5.98)<br />

t < 0 : ch(t, 0) = (−t) 1−2p<br />

p ch(−1, 0) (5.99)<br />

which shows again th<strong>at</strong> the critical exponents are indeed the same above and<br />

below the critical point:<br />

α = α ′ =<br />

and again th<strong>at</strong> the pre-factors are different.<br />

Scaling rel<strong>at</strong>ions.<br />

2p − 1<br />

p<br />

(5.100)<br />

Using the expressions for the critical exponents it is possible to derive rel<strong>at</strong>ions<br />

between these exponents by elimin<strong>at</strong>ing the variables p and q. For<br />

example, we have<br />

q<br />

(δ + 1)β = (<br />

1 − q<br />

+ 1)1 − q<br />

p<br />

1<br />

= = 2 − α (5.101)<br />

p<br />

and if we check this for the mean field exponents we see th<strong>at</strong> there indeed<br />

(δ + 1)β = 4 × 1<br />

2 = 2 − α indeed.<br />

Similarly we find th<strong>at</strong>


214 CHAPTER 5. CRITICAL EXPONENTS.<br />

q<br />

(δ − 1)β = (<br />

1 − q<br />

− 1)1 − q<br />

p<br />

= 2q − 1<br />

p<br />

= γ (5.102)<br />

which was already discovered by Widom before the formul<strong>at</strong>ion of scaling theory.<br />

It is easy to check th<strong>at</strong> the mean field exponents s<strong>at</strong>isfy this equality. Another<br />

equ<strong>at</strong>ion, named after Rushbrooke, is also easily derived:<br />

and<br />

α + 2β + γ = 2 (5.103)<br />

Note th<strong>at</strong> we can work backwards and find th<strong>at</strong> in mean field theory we have<br />

pMF = 1<br />

2<br />

qMF = 3<br />

4<br />

(5.104)<br />

(5.105)<br />

A final note, in this type of scaling we always find th<strong>at</strong> α = α ′ and γ = γ ′ ,<br />

because we assume th<strong>at</strong> the scaling is the same below and above the critical<br />

temper<strong>at</strong>ure.<br />

Comparison with experiment.<br />

We can look <strong>at</strong> the experimental d<strong>at</strong>a presented before, and construct the<br />

following table:<br />

type 1 − α+γ<br />

2 β 1 + γ<br />

β δ<br />

ferromagnetic Ni 0.39 0.33 5.0 4.2<br />

ferromagnetic F e 0.39 0.34<br />

anti-ferromagnetic RbMnF3 0.37 0.32<br />

liquid-gas Co2 0.34 0.34 4.5 4.2<br />

liquid-gas Xe 0.36 0.34 4.5 4.4<br />

mean field 0.5 0.5 3.0 3.0<br />

Ising, d=2 0.125 0.125 15 15<br />

Ising, d=3 0.31 0.31 5.0 5.0<br />

Heisenberg 0.38 0.38<br />

which is reasonable. One has to keep in mind th<strong>at</strong> critical exponents are very<br />

difficult to measure, and th<strong>at</strong> the error bars are large. If we take th<strong>at</strong> into<br />

account, we conclude th<strong>at</strong> the experimental results do not contradict scaling.<br />

Equalities or inequalities?<br />

The basis for the equalities can be discussed in a completely different manner.<br />

In chapter 2 we derived inequalities rel<strong>at</strong>ed to thermodynamic stability. For<br />

example, we found th<strong>at</strong>


5.4. CONSEQUENCES OF SCALING. 215<br />

CH − CM = T<br />

χT<br />

( )<br />

∂M<br />

2<br />

∂T H<br />

and since CM > 0 we can drop the term CM in the left hand side and find:<br />

CH ≥ T<br />

χT<br />

( )<br />

∂M<br />

2<br />

∂T H<br />

Near the critical point we have CH = c1|θ| −α , χT = c2|θ| −γ , and ( ∂M<br />

∂T<br />

c3|θ| β−1 , which gives<br />

c1|θ| −α ≥ T c23 |θ|<br />

c2<br />

2β+γ−2<br />

and this should be true even in the limit θ → 0, which leads to<br />

or<br />

as originally discovered by Rushbrooke.<br />

In a similar manner one can derive<br />

(5.106)<br />

(5.107)<br />

)<br />

H =<br />

(5.108)<br />

−α ≤ 2β + γ − 2 (5.109)<br />

α + 2β + γ ≥ 2 (5.110)<br />

γ ≥ β(δ − 1) (5.111)<br />

which was originally shown by Widom.<br />

The deriv<strong>at</strong>ion above is based on equilibrium thermodynamics. In scaling<br />

theory we describe a non-equilibrium st<strong>at</strong>e with fluctu<strong>at</strong>ions. But if we think<br />

of a given st<strong>at</strong>e with two phases present, where one phase is meta-stable, we<br />

can still apply the same kind of analysis to the meta-stable st<strong>at</strong>e, since it still<br />

corresponds to a minimum in the free energy.<br />

Explain this carefully.<br />

One can therefore argue th<strong>at</strong> the inequality 5.110 should also hold for the singular<br />

part of the free energy. Scaling theory confirms th<strong>at</strong>, but is also much<br />

more restrictive, because it forces the inequality to be an equality.<br />

Pair correl<strong>at</strong>ion function.<br />

In the previous chapter we discussed how the effects of a small perturb<strong>at</strong>ion<br />

are felt further away in the system. In Ginzburg-Landau theory it was<br />

possible to calcul<strong>at</strong>e the change in magnetiz<strong>at</strong>ion density caused by a perturb<strong>at</strong>ion<br />

<strong>at</strong> the origin. The function in equ<strong>at</strong>ion 4.56 is an example of a


216 CHAPTER 5. CRITICAL EXPONENTS.<br />

pair correl<strong>at</strong>ion function, i.e. a function describing how the magnetiz<strong>at</strong>ion <strong>at</strong><br />

two different sites is rel<strong>at</strong>ed. The formal definition of the pair correl<strong>at</strong>ion function<br />

Γ is<br />

Γ(T, ⃗r, ⃗r ′ ) = ⟨(m(⃗r) − m0)(m(⃗r ′ ) − m0)⟩ (5.112)<br />

where m0 is the average value of the magnetiz<strong>at</strong>ion. The pair correl<strong>at</strong>ion function<br />

will also depend on the temper<strong>at</strong>ure, of course, and on other thermodynamic<br />

variables. In a homogeneous system the pair correl<strong>at</strong>ion function will<br />

only depend on the distance |⃗r − ⃗r ′ | between the points. In Ginzburg-Landau<br />

theory we found th<strong>at</strong> the pair correl<strong>at</strong>ion is proportional to the function calcul<strong>at</strong>ed<br />

in equ<strong>at</strong>ion 4.56, since this function gives us the extra magnetiz<strong>at</strong>ion<br />

when the origin has a large magnetiz<strong>at</strong>ion (driven by a delta function applied<br />

field). Therefore, in th<strong>at</strong> case:<br />

Two critical exponents.<br />

Γ(T, r) ∝ 1 r<br />

e− ξ (5.113)<br />

r<br />

The critical exponents ν and ν ′ describe the behavior of the correl<strong>at</strong>ion<br />

length ξ close to the critical temper<strong>at</strong>ure:<br />

T > Tc : ξ(T ) ∝ (T − Tc) −ν<br />

T < Tc : ξ(T ) ∝ (Tc − T ) −ν′<br />

(5.114)<br />

(5.115)<br />

and again we expect th<strong>at</strong> scaling theory will give ν = ν ′ . Ginzburg-Landau<br />

theory gives ν = ν ′ = 1<br />

2 . At the critical temper<strong>at</strong>ure the correl<strong>at</strong>ion length is<br />

infinite, and the decay of the pair correl<strong>at</strong>ion function is given by the r dependence<br />

in front. Another critical exponent η is defined by:<br />

Γ(Tc, r) ∝ r 2−d−η<br />

(5.116)<br />

where d is the number of dimensions in the problem. Using the results of one<br />

of the problems in the previous chapter we see th<strong>at</strong> η = 0 in all dimensions for<br />

mean field (Ginzburg-Landau) theory.<br />

Momentum space.<br />

In many cases we need the Fourier transform of the pair correl<strong>at</strong>ion function,<br />

and calcul<strong>at</strong>e<br />

∫<br />

d d re ı⃗q·⃗r Γ(T, r) (5.117)


5.5. SCALING OF THE PAIR CORRELATION FUNCTION. 217<br />

which for small q is proportional to q η−2 . This is often used as the definition<br />

for η, but is of course completely equivalent to our definition. L<strong>at</strong>er we will<br />

discuss the exact definition of the structure factor, which is almost the same<br />

as the Fourier transform above. The difference is not essential for the current<br />

discussion.<br />

Rel<strong>at</strong>ion to the susceptibility.<br />

Since the pair correl<strong>at</strong>ion function describes the average of fluctu<strong>at</strong>ions, we<br />

can use the techniques from the end of Chapter 2 to rel<strong>at</strong>e this to a response<br />

function, which in this case is the susceptibility. In order to find the susceptibility<br />

we need to take an average over all space, and hence we find th<strong>at</strong><br />

∫<br />

χ(T ) ∝ d d rΓ(T, r) (5.118)<br />

or th<strong>at</strong> the susceptibility is proportional to the Fourier transform of the pair<br />

correl<strong>at</strong>ion function <strong>at</strong> q=0.<br />

Make sure th<strong>at</strong> you understand the previous argument based on chapter 2.<br />

5.5 Scaling of the pair correl<strong>at</strong>ion function.<br />

We assume th<strong>at</strong> near the critical point the correl<strong>at</strong>ion length is the only important<br />

length scale. As a result, the pair correl<strong>at</strong>ion function has to be a function<br />

and we can write<br />

of the r<strong>at</strong>io r<br />

ξ<br />

Γ(T, r) = f( r<br />

ξ(T ) )ξu (T ) = ˜ f( r<br />

ξ(T ) )ru<br />

(5.119)<br />

At the critical temper<strong>at</strong>ure the correl<strong>at</strong>ion length is infinite and therefore the<br />

second part of the equ<strong>at</strong>ion above gives<br />

Γ(Tc, r) = ˜ f(0)r u<br />

(5.120)<br />

which shows th<strong>at</strong> u = 2 − d − η. The Fourier transform of the pair distribution<br />

function is<br />

∫<br />

d d re ı⃗q·⃗r f( r<br />

ξ(T ) )ξu (T ) (5.121)<br />

which can be transformed using ⃗r = ξ(T )⃗x and ⃗s = ξ(T )⃗q into<br />

∫<br />

d d xe ı⃗s·⃗x f(x)ξ u+d (T ) (5.122)


218 CHAPTER 5. CRITICAL EXPONENTS.<br />

Rel<strong>at</strong>ion between the susceptibility and the correl<strong>at</strong>ion length.<br />

Since the susceptibility is proportional to the q=0 component of the Fourier<br />

transform of the pair correl<strong>at</strong>ion function, we see th<strong>at</strong> near the critical point<br />

the susceptibility is proportional to<br />

χ(T ) ∝ ξ u+d = ξ 2−η ∝ |T − Tc| −ν(2−η)<br />

which shows the important rel<strong>at</strong>ion between the critical exponents:<br />

5.6 Hyper-scaling.<br />

(5.123)<br />

γ = ν(2 − η) (5.124)<br />

Near the critical point the correl<strong>at</strong>ion length is the only important length scale,<br />

and this has one final consequence. The energy needed to cre<strong>at</strong>e a fluctu<strong>at</strong>ion,<br />

independent of its size, will be a constant proportional to Tc. L<strong>at</strong>er we can<br />

derive in a microscopic picture th<strong>at</strong> it will be on the order of kBTc, but for our<br />

present discussion it is sufficient to assume th<strong>at</strong> it is a constant per fluctu<strong>at</strong>ion.<br />

The volume of a fluctu<strong>at</strong>ion is proportional to ξ d , and therefore we find th<strong>at</strong><br />

near the critical point the energy density scales like<br />

g(θ, h = 0) ∝ ξ −d (T ) (5.125)<br />

This assumption is called hyper-scaling. Near the critical point we therefore<br />

have<br />

g(θ, h = 0) ∝ θ νd<br />

From the energy scaling rel<strong>at</strong>ion 5.82 we find:<br />

and with t = λ p we get<br />

which shows th<strong>at</strong> νd = 1<br />

p , or<br />

(5.126)<br />

g(λ p , 0) = λg(1, 0) (5.127)<br />

g(t, 0) = t 1<br />

p g(1, 0) (5.128)<br />

dν = 2 − α (5.129)<br />

which is NOT true for mean field theory combined with Ginzburg-Landau in<br />

three dimensions! The equ<strong>at</strong>ion above would give ν = 2<br />

3 , while Ginzburg-<br />

Landau gives ν = 1<br />

2 . Therefore, Ginzburg-Landau theory does not obey hyperscaling,<br />

and cannot be a complete theory <strong>at</strong> the transition. Note, however, th<strong>at</strong><br />

in four dimensions the equ<strong>at</strong>ion above is obeyed. Perhaps something is special<br />

in four dimensions!


5.7. VALIDITY OF GINZBURG-LANDAU THEORY. 219<br />

5.7 Validity of Ginzburg-Landau theory.<br />

Criterium for validity.<br />

The theory put forward by Ginzburg and Landau is only a start for a description<br />

of the role of fluctu<strong>at</strong>ions. The critical exponents in this theory have<br />

the mean field values, and in real experiments those values are different. But<br />

it is still a fair question to ask when this theory is a good approxim<strong>at</strong>ion. One<br />

criterium is the following. If we have fluctu<strong>at</strong>ions in a parameter, a theory th<strong>at</strong><br />

tre<strong>at</strong>s these as a perturb<strong>at</strong>ion needs to require th<strong>at</strong> the fluctu<strong>at</strong>ions are small,<br />

or th<strong>at</strong><br />

∆M ≪ M (5.130)<br />

which, obviously, can only be applied for temper<strong>at</strong>ures below the critical temper<strong>at</strong>ure.<br />

Note th<strong>at</strong> we use the total magnetiz<strong>at</strong>ion in this formula. In Ginzburg-<br />

Landau we have<br />

(∆M) 2 ∫<br />

= d 3 rd 3 r ′ ⟨(m(⃗r)−m0)(m(⃗r ′ ∫<br />

)−m0)⟩ = V d 3 ∫<br />

rΓ(T, r) ∝ V d 3 r 1<br />

r<br />

(5.131)<br />

which leads to something like<br />

(∆M) 2 = C1V ξ 2 (T ) (5.132)<br />

where C1 is a constant independent of the temper<strong>at</strong>ure.<br />

Inequality for validity.<br />

The total magnetiz<strong>at</strong>ion is simple, and we have M = V m0. Therefore we<br />

have a condition of the form<br />

C1V ξ 2 (T ) ≪ m 2 0V 2<br />

e− r<br />

ξ<br />

(5.133)<br />

which is always obeyed in the thermodynamic limit V → ∞. But th<strong>at</strong> is not<br />

correct. The thermodynamic limit should always be taken in the end,<br />

and we first need to consider the behavior as a function of temper<strong>at</strong>ure. If we<br />

approach the critical points the correl<strong>at</strong>ion length becomes infinite, signalling<br />

th<strong>at</strong> interchanging limits can be dangerous! Our criterium for validity should<br />

be th<strong>at</strong> the fluctu<strong>at</strong>ions measured in a relevant volume must be small. Since<br />

the range of fluctu<strong>at</strong>ions are given by the correl<strong>at</strong>ion length, the obvious choice<br />

is to consider a volume determined by this correl<strong>at</strong>ion length! Therefore, we<br />

take V = ξ 3 (T ).<br />

Since the magnetiz<strong>at</strong>ion itself is determined by the critical exponent β we<br />

have the following inequality th<strong>at</strong> gives a necessary criterium for the validity of<br />

Ginzburg-Landau theory:


220 CHAPTER 5. CRITICAL EXPONENTS.<br />

or<br />

C1 ≪ m 2 0ξ (5.134)<br />

C2 ≪ θ 2β−ν<br />

(5.135)<br />

where all constants are lumped together in C2. Using the appropri<strong>at</strong>e values<br />

β = ν = 1<br />

2 we get<br />

|θ| ≫ C 2 2<br />

(5.136)<br />

which shows th<strong>at</strong> close to the critical point Ginzburg-Landau theory is not<br />

self-consistent, i.e. it predicts fluctu<strong>at</strong>ions which are so large th<strong>at</strong> the basic<br />

assumptions of the theory are not valid anymore. Higher order gradients become<br />

important, and the series as an expansion in terms of gradients does not converge<br />

anymore.<br />

Necessary or sufficient.<br />

Strictly speaking the criterium above is only a necessary criterium for the<br />

validity of the theory. In general, however, it will also be sufficient. If |θ| ≫ C 2 2<br />

higher order gradients will be less and less important, and an expansion in terms<br />

of gradients is appropri<strong>at</strong>e. But this is only a st<strong>at</strong>ement about very high orders.<br />

Ginzburg-Landau theory assumes th<strong>at</strong> the lowest order gradient is the dominant<br />

term. This is not guaranteed, it could be possible th<strong>at</strong> all the action is in the<br />

next term. Therefore, the condition we have derives is not sufficient in general.<br />

In practice, however, one will often find th<strong>at</strong> the lowest order term is dominant,<br />

and th<strong>at</strong> the inequality derived above is also sufficient. An important exception<br />

does occur in all those cases where symmetry arguments make this term exactly<br />

zero! One should always keep th<strong>at</strong> in mind.<br />

Use of Ginzburg-Landau theory.<br />

Nevertheless, the theory of Ginzburg and Landau is an improvement, and<br />

will give a better description of the system as long as the inequality above<br />

is obeyed. Hence we have the following picture. Far away from the critical<br />

point mean field theory is a good description of the system under consider<strong>at</strong>ion.<br />

When we get closer to the critical point, we can improve by including gradient<br />

terms in the expansion of the free energy. Very close to the critical point we<br />

need a different theory, however. In this chapter we have given a model of the<br />

free energy th<strong>at</strong> describes the behavior <strong>at</strong> the critical point, but the critical<br />

exponents are parameters in th<strong>at</strong> model (following from the values of p and q<br />

in equ<strong>at</strong>ion 5.82). In order to derive the values of these critical exponents we<br />

need a different theory, however. This is usually done in st<strong>at</strong>istical mechanics,<br />

where we include the microscopic n<strong>at</strong>ure of the m<strong>at</strong>erial.


5.8. SCALING OF TRANSPORT PROPERTIES. 221<br />

Range of validity.<br />

The inequality |θ| ≫ C 2 2 gives us also an estim<strong>at</strong>e of the range of validity.<br />

In order to calcul<strong>at</strong>e numerical values we do need to know the constant C2, and<br />

hence we need the details of the model. It is possible th<strong>at</strong> Ginzburg-Landau<br />

theory is valid essentially everywhere for our measurement (apart from giving<br />

the values of the critical exponents), but there are also situ<strong>at</strong>ions where the<br />

excluded range is so large th<strong>at</strong> this theory is never applicable.<br />

Different dimensions.<br />

We can also apply the previous argument for a system in d dimensions. In<br />

th<strong>at</strong> case the pair correl<strong>at</strong>ion length is proportional to r 2−d (see problem in<br />

previous chapter), and the volume V is V = ξ d (T ). This leads to<br />

or<br />

C3 ≪ m 2 0ξ d−2<br />

C4 ≪ θ 2β−(d−2)ν<br />

which is valid even <strong>at</strong> the critical point as long as 2β − (d − 2)ν < 0 or<br />

Marginal dimension.<br />

(5.137)<br />

(5.138)<br />

d > d ∗ = 2(1 + β<br />

) (5.139)<br />

ν<br />

The constant d ∗ above is called the marginal dimension. In Ginzburg-<br />

Landau it has the value d ∗ = 4. The condition d > 4 can be interpreted as<br />

follows. If the number of dimensions is larger than four, mean field theory will<br />

be valid (as long as the conditions described in the necessary versus sufficient<br />

paragraph are met). This is the basis for the famous epsilon expansion, where<br />

all quantities are expanded in a power series in terms of d = d ∗ − ϵ and the limit<br />

d → 3 is taken to get results for the real world.<br />

If we are exactly <strong>at</strong> the marginal dimension, mean field theory does not<br />

necessarily describe the physics of the problem, but the difference can only be<br />

in terms th<strong>at</strong> diverge in a logarithmic fashion, i.e. slower than any arbitrary<br />

power of θ. Details can be quite messy in th<strong>at</strong> case.<br />

5.8 Scaling of transport properties.<br />

Different basis for scaling.


222 CHAPTER 5. CRITICAL EXPONENTS.<br />

We have introduced the scaling hypothesis by using the free energy. This<br />

seems an appropri<strong>at</strong>e choice, since the free energy is the most important quantity<br />

from which all other thermodynamic variables can be derived. Using two basic<br />

scaling parameters, p and q, we were able to derive all critical exponents. The<br />

important observ<strong>at</strong>ion is th<strong>at</strong> there are only two independent variables, and<br />

any set of two exponents can be used as a basic set. Of course, as soon as we<br />

introduce more st<strong>at</strong>e variables, we obtain more possibilities. For example, if<br />

we also allow pressure to vary in our free energy, more complex behavior can<br />

be observed. At first, increasing the pressure in a magnet will simply shift the<br />

critical temper<strong>at</strong>ure, but not change the critical exponents. In th<strong>at</strong> case the<br />

scaling of the pressure is irrelevant, and this can be expressed by setting the<br />

power r in λ r equal to zero. But we could also get things like tri-critical points,<br />

and now the story is different.<br />

The other question one can ask is if the free energy is the only possible<br />

starting point. The answer is no, any function of two variables can be used<br />

as a basis for scaling. One common choice is to start with the pair correl<strong>at</strong>ion<br />

function, and write:<br />

Γ(λ x r, λ y θ) = λΓ(r, θ) (5.140)<br />

and derive all critical exponents from the values of x and y. This is not hard to<br />

do, since we know th<strong>at</strong> the pair correl<strong>at</strong>ion function has the form<br />

with u = 2 − d − η, and therefore<br />

which yields:<br />

and<br />

with solutions<br />

and<br />

Γ(T, r) = f( r<br />

ξ(T ) )ξu (T ) (5.141)<br />

Γ(λ x r, λ y θ) = f(λ x+yν r<br />

ξ(T ) )λ−yuν ξ u (T ) (5.142)<br />

y = − 1<br />

uν =<br />

1<br />

dν − (2 − η)ν<br />

x + yν = 0 (5.143)<br />

−yuν = 1 (5.144)<br />

= 1<br />

2β =<br />

Show th<strong>at</strong> 2β = dν − (2 − η)ν.<br />

p<br />

2(1 − q)<br />

(5.145)


5.8. SCALING OF TRANSPORT PROPERTIES. 223<br />

x = −yν =<br />

2 − α<br />

2dβ =<br />

1<br />

2d(1 − q)<br />

which shows th<strong>at</strong> the two formul<strong>at</strong>ions are equivalent.<br />

Time dependent phenomena.<br />

(5.146)<br />

In Ginzburg-Landau theory we describe the thermodynamically stable st<strong>at</strong>e<br />

when we have applied a field <strong>at</strong> the origin. In a real process we can imagine th<strong>at</strong><br />

<strong>at</strong> time τ = 0 we apply a magnetiz<strong>at</strong>ion <strong>at</strong> the origin in a non-magnetic st<strong>at</strong>e.<br />

The effects of this seed <strong>at</strong> the origin will propag<strong>at</strong>e outwards, and regions away<br />

from the origin will become magnetized. The magnetized region will expand<br />

and finally in the limit τ → ∞ approach the Landau-Ginzburg solution. Since<br />

this process is a non-equilibrium process, it is hard to describe in terms of a free<br />

energy, but one can easily describe the pair correl<strong>at</strong>ion function as a function<br />

of time τ by Γ(r, τ, θ).<br />

Diffusion equ<strong>at</strong>ion.<br />

As we will see l<strong>at</strong>er, the change in magnetiz<strong>at</strong>ion over time is described by<br />

a diffusion equ<strong>at</strong>ion:<br />

( )<br />

∂m<br />

= D<br />

∂τ θ,⃗r<br />

⃗ ∇ 2 m (5.147)<br />

where D is the diffusion constant. In our example where we start with a delta<br />

function <strong>at</strong> the origin, m(r, τ = 0) = M0δ(⃗r) the solution is simple:<br />

m(⃗r, τ) =<br />

M0<br />

(4πDτ) 3<br />

r2 −<br />

e 4Dτ (5.148)<br />

2<br />

where we see th<strong>at</strong> the size of the magnetized area increases like √ τ as expected.<br />

This function m(⃗r, τ) is proportional to Γ(r, τ, θ).<br />

Time scaling.<br />

We now take into account the scaling of the time variable in a straightforward<br />

manner:<br />

Γ(λ x r, λ z τ, λ y θ) = λΓ(r, τ, θ) (5.149)<br />

which introduces a new parameter z and hence one new independent critical<br />

exponent.<br />

Scaling of the diffusion constant.


224 CHAPTER 5. CRITICAL EXPONENTS.<br />

The pair correl<strong>at</strong>ion function also obeys the diffusion equ<strong>at</strong>ion, given by:<br />

( )<br />

∂Γ<br />

= D<br />

∂τ θ,⃗r<br />

⃗ ∇ 2 Γ (5.150)<br />

where the diffusion constant depends on the temper<strong>at</strong>ure, and near the critical<br />

point is described by an exponent κ:<br />

Rel<strong>at</strong>ion between z and κ.<br />

and<br />

D(θ) ∝ θ −κ<br />

The deriv<strong>at</strong>ives in the diffusion equ<strong>at</strong>ion also scale with λ and we have<br />

which leads to<br />

which gives<br />

λ z<br />

(5.151)<br />

( )<br />

∂Γ<br />

(λ<br />

∂τ<br />

x r, λ z τ, λ y ( )<br />

∂Γ<br />

θ) = λ (r, τ, θ) (5.152)<br />

∂τ<br />

λ 2x ⃗ ∇ 2 Γ(λ x r, λ z τ, λ y θ) = λ ⃗ ∇ 2 Γ(r, τ, θ) (5.153)<br />

D(λ y θ) = D(θ)λ 2x−z<br />

κ =<br />

z − 2x<br />

y<br />

= z + 2yν<br />

y<br />

(5.154)<br />

(5.155)<br />

This shows th<strong>at</strong> we can obtain the value of z from a measurement of the diffusion<br />

constant and a knowledge of two other critical exponents.<br />

Consequences.<br />

If the time scaling in the pair correl<strong>at</strong>ion function is not important, and z=0,<br />

we find th<strong>at</strong> κ = 2ν and in Ginzburg-Landau theory this gives κ = 1. Therefore,<br />

we expect the diffusion constant to become infinitely large <strong>at</strong> the critical point.<br />

Any small perturb<strong>at</strong>ion propag<strong>at</strong>es with very large speed near the critical point.<br />

This is, of course, not realistic, and implies th<strong>at</strong> we need to use a rel<strong>at</strong>ivistic<br />

form of the equ<strong>at</strong>ions. Also, the diffusion equ<strong>at</strong>ion has a built-in assumption<br />

th<strong>at</strong> the diffusion is slow. So details near the critical point are again messy,<br />

but the fact th<strong>at</strong> the diffusion constant increases when we approach the critical<br />

point can easily be measured.


5.9. EXTRA QUESTIONS. 225<br />

5.9 Extra questions.<br />

1. Wh<strong>at</strong> is the physics determining the n<strong>at</strong>ure of the critical exponents?<br />

2. Wh<strong>at</strong> is universality?<br />

3. Wh<strong>at</strong> is the meaning of universality?<br />

4. Superconductivity example. Discuss the rel<strong>at</strong>ion between phase and vector<br />

potential.<br />

5. Wh<strong>at</strong> is mean field theory and why is it important?<br />

6. Can we find critical exponents in finite systems?<br />

7. How do we measure critical exponents?<br />

8. Wh<strong>at</strong> are scaled variables?<br />

9. Wh<strong>at</strong> is the meaning of the fluctu<strong>at</strong>ion term in the energy?<br />

10. Wh<strong>at</strong> does scaling of the free energy mean?<br />

11. Why do we only scale the fluctu<strong>at</strong>ion part?<br />

12. Why are there only two independent critical exponents?<br />

13. Wh<strong>at</strong> is the physical meaning of inequalities versus equalities?<br />

14. Wh<strong>at</strong> is the pair correl<strong>at</strong>ion function?<br />

15. Wh<strong>at</strong> does scaling of the pair correl<strong>at</strong>ion function mean?<br />

16. Wh<strong>at</strong> is hyper-scaling?<br />

17. Wh<strong>at</strong> is a marginal dimension?<br />

18. Could we start expansions starting <strong>at</strong> infinite dimensions?<br />

19. Wh<strong>at</strong> are consequences of mean field theory being correct in 4d?<br />

20. Explain time scaling.<br />

21. Can the diffusion constant be infinite?


226 CHAPTER 5. CRITICAL EXPONENTS.<br />

5.10 Problems for chapter 5<br />

Problem 1.<br />

Derive equ<strong>at</strong>ion 5.111.<br />

Problem 2.<br />

Show th<strong>at</strong> equ<strong>at</strong>ion 5.117 for small q is proportional to q η−2 .<br />

Problem 3.<br />

Derive equ<strong>at</strong>ion 5.154<br />

Problem 4.<br />

A sequence of critical exponents ∆l , ∆ ′ l<br />

is defined by<br />

( l ∂ G<br />

∂H l<br />

) ( ) −∆l<br />

T<br />

∝ − 1<br />

T<br />

Tc<br />

( ∂l−1G ∂H l−1<br />

)<br />

T<br />

(5.156)<br />

for l > 0 and T > Tc. The primed exponents are for T < Tc, where the pre-<br />

factor is<br />

(<br />

1 − T<br />

Tc<br />

) −∆ ′<br />

l<br />

. These exponents are called gap exponents. Show th<strong>at</strong><br />

∆ ′ 1 = 2 − α ′ − β and ∆ ′ 2 = β + γ ′ . Show th<strong>at</strong> in mean field theory ∆ ′ l<br />

all values of l.<br />

Problem 5.<br />

= 3<br />

2 for<br />

The scaling laws yield the following identity for the magnetiz<strong>at</strong>ion m as<br />

a function of scaled temper<strong>at</strong>ure θ and magnetic field h: λ q m(λ p θ, λ q h) =<br />

λm(θ, h). In terms of the critical temper<strong>at</strong>ure Tc we have θ =<br />

T −Tc<br />

Tc<br />

. The ex-<br />

ponents p and q are a property of the system under consider<strong>at</strong>ion. We measure<br />

m as a function of h for several temper<strong>at</strong>ures θ. If we plot these measurements<br />

in the m versus h plane we get a series of distinct curves. Suppose we now plot<br />

the d<strong>at</strong>a in a different manner. For a series <strong>at</strong> temper<strong>at</strong>ure θ we plot the d<strong>at</strong>a<br />

points <strong>at</strong> |θ| A h with value |θ| B m. How do we choose A and B in terms of p and<br />

q in such a manner th<strong>at</strong> all d<strong>at</strong>a points now fall on one universal curve?<br />

Problem 6.<br />

The scaling law yields the following identity for the Gibbs free energy per<br />

unit volume g as a function of scaled temper<strong>at</strong>ure θ and magnetic field h:


5.10. PROBLEMS FOR CHAPTER 5 227<br />

g(λpθ, λqh) = λg(θ, h). In terms of the critical temper<strong>at</strong>ure Tc we have θ =<br />

T −Tc . The exponents p and q are a property of the system under consider<strong>at</strong>ion.<br />

Tc<br />

Show th<strong>at</strong> we can write<br />

g(θ, h) = |θ| x G±( h<br />

)<br />

|θ| y<br />

where the plus or minus sign rel<strong>at</strong>es to being above or below the critical<br />

temper<strong>at</strong>ure. Rel<strong>at</strong>e the powers x and y to the powers p and q. Find an<br />

expression for G+ th<strong>at</strong> is consistent with Curies law for T > Tc, th<strong>at</strong> is m =<br />

, and wh<strong>at</strong> does th<strong>at</strong> tell you about the values of p and q.<br />

m0h<br />

T −Tc<br />

Problem 7.<br />

Near a liquid to gas critical point, <strong>at</strong> pressure pc and temper<strong>at</strong>ure Tc we can<br />

write the Gibbs free energy per unit volue, g, in terms of the reduced variables<br />

π = p−pc<br />

T −Tc<br />

and θ = . The dominant part of the free energy obeys the scaling<br />

pc Tc<br />

rel<strong>at</strong>ion:<br />

λg(π, θ) = g(λ z π, λ y θ)<br />

Find the value of the critical exponents γ and γ ′ in terms of z and y.<br />

Problem 8.<br />

Starting with equ<strong>at</strong>ion (5.83) show th<strong>at</strong> one can write<br />

m(θ, h) = |θ| β M±( h<br />

)<br />

|θ| βδ<br />

We use this to plot experimental d<strong>at</strong>a in the form<br />

versus<br />

m(θ, h)<br />

|θ| β<br />

h<br />

|θ| βδ<br />

which has the unknown parameters Tc, β, and δ. If these parameters are<br />

chosen correctly, the d<strong>at</strong>a points fall on two curves, one above Tc and one below.<br />

Using some spreadsheet applic<strong>at</strong>ion and the d<strong>at</strong>a in the table below, make<br />

this plot, and find your best values for the parameters.<br />

The following table (copied from Huang) shows for nickel the magnetiz<strong>at</strong>ion<br />

as a function of temper<strong>at</strong>ure and field:


228 CHAPTER 5. CRITICAL EXPONENTS.<br />

Temper<strong>at</strong>ure (Kelvin) Applied field (Oersted) Magnetiz<strong>at</strong>ion density (emu/g)<br />

678.86 4, 160 0.649<br />

678.86 10, 070 1.577<br />

678.86 17, 775 2.776<br />

653.31 2, 290 0.845<br />

653.31 4, 160 1.535<br />

653.31 6, 015 2.215<br />

653.31 10, 070 3.665<br />

653.31 14, 210 5.034<br />

653.31 17, 775 6.145<br />

653.31 21, 315 7.211<br />

629.53 1, 355 6.099<br />

629.53 3, 230 9.287<br />

629.53 6, 015 11.793<br />

629.53 10, 070 13.945<br />

629.53 14, 210 15.405<br />

629.53 17, 775 16.474<br />

614.41 1, 820 19.43<br />

614.41 3, 230 20.00<br />

614.41 6, 015 20.77<br />

614.41 10, 070 21.67<br />

614.41 14, 210 22.36<br />

614.41 17, 775 22.89<br />

601.14 3, 230 25.34<br />

601.14 6, 015 25.83<br />

601.14 10, 070 26.40<br />

601.14 14, 210 26.83<br />

601.14 17, 775 27.21<br />

Problem 9.<br />

Show th<strong>at</strong> 5.148 is indeed the solution of the diffusion equ<strong>at</strong>ion 5.147


Chapter 6<br />

Transport in<br />

<strong>Thermodynamics</strong>.<br />

Equilibrium thermodynamics demands the absence of macroscopic currents. It<br />

does not require th<strong>at</strong> the m<strong>at</strong>erial is exactly the same everywhere. Energy can<br />

flow back and forth, but we can describe a precise average value of the energy<br />

per unit volume <strong>at</strong> each point in space. In chapter two we have seen th<strong>at</strong> we<br />

can describe the probability of local fluctu<strong>at</strong>ions using the local entropy. We<br />

assumed th<strong>at</strong> measurement times were very long, and th<strong>at</strong> the net effect of<br />

the fluctu<strong>at</strong>ions is zero. We also assumed th<strong>at</strong> fluctu<strong>at</strong>ions <strong>at</strong> different points<br />

in space are not correl<strong>at</strong>ed. In chapter four we discussed how fluctu<strong>at</strong>ions are<br />

localized in space, and how the extent diverges near the critical point. In chapter<br />

five we used this correl<strong>at</strong>ion to describe its effects on the energy and on critical<br />

exponents.<br />

In this chapter we discuss another aspect of correl<strong>at</strong>ions between differences<br />

from average values. When we apply a small external field, we cre<strong>at</strong>e a continuous<br />

change in thermodynamic variables. This gives rise to net macroscopic<br />

currents. Such currents are difficult to describe in general, but here we take one<br />

limit in which we can make predictions. We assume th<strong>at</strong> the currents are so<br />

small, th<strong>at</strong> diffusion times are much larger than fluctu<strong>at</strong>ion times. Th<strong>at</strong> leads<br />

to the concept of local thermodynamic equilibrium.<br />

In this chapter we start with describing transport in general terms, and then<br />

giving examples of thermo-electric effects. In the next section we define the cases<br />

where we can discuss these effects based on thermodynamics. We develop the<br />

transport equ<strong>at</strong>ions, and the equ<strong>at</strong>ion for the gener<strong>at</strong>ion of entropy in such a<br />

non-equilibrium situ<strong>at</strong>ion. Next we follow Onsager, and show how the transport<br />

coefficients are rel<strong>at</strong>ed to the probabilities for local fluctu<strong>at</strong>ions. We then apply<br />

our formulas to cases where electronic transport takes place via diffusion, but<br />

now in the presence of a thermal gradient.<br />

229


230 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

6.1 Introduction.<br />

Wh<strong>at</strong> experimentalists do.<br />

Experimental sciences all have one thing in common. You are trying to<br />

discover something about a system. There are two ways of doing th<strong>at</strong>. You can<br />

sit down and observe the system. This is always the best first step, and one<br />

cannot underestim<strong>at</strong>e the importance of this step. But there comes a point in<br />

time where observing alone does not give much new inform<strong>at</strong>ion on a time scale<br />

of interest. For example, consider a system th<strong>at</strong> shows a very peculiar behavior<br />

one hour every three hundred years. This is longer th<strong>at</strong> our life time, and we<br />

will be imp<strong>at</strong>ient. It is possible for gener<strong>at</strong>ions of scientists to study this system<br />

and write down wh<strong>at</strong> they observe. After a thousand years or so we will have<br />

good inform<strong>at</strong>ion about th<strong>at</strong> system. The ”we” in this context is mankind, not<br />

an individual. Our culture, however, does not have th<strong>at</strong> p<strong>at</strong>ience, and we will<br />

do something about it.<br />

The second step in any experimental discovery is the following. You poke<br />

the system and see how it responds. Here we can make several choices. We<br />

can poke it gently or violently. Wh<strong>at</strong> is the difference? If we poke it gently,<br />

the system will return to its original st<strong>at</strong>e or to a nearby st<strong>at</strong>e which we still<br />

identify as being part of the original system. If we poke it violently, the system<br />

afterwards is completely different from the system before.<br />

Give examples of these two types of investig<strong>at</strong>ions.<br />

One of the problems with distinguishing gentle and violent is th<strong>at</strong> we do not<br />

know a priori where the border between these two is. Of course, we will have<br />

some idea from our careful observ<strong>at</strong>ions in step one. But even th<strong>at</strong> might not be<br />

sufficient. In th<strong>at</strong> case we start with the smallest poke we can do, and observe<br />

the reactions. We can only hope th<strong>at</strong> we did it gentle enough.<br />

How do these general remarks transl<strong>at</strong>e into the realm of physics? Especially<br />

in the context of thermodynamics. Th<strong>at</strong> is not too difficult. For example, we<br />

have a system and increase its temper<strong>at</strong>ure. We wait until the system is again<br />

in equilibrium.<br />

How long do we have to wait?<br />

After the waiting time is over we measure wh<strong>at</strong>ever we want to measure.<br />

Can we measure without perturbing the system?


6.1. INTRODUCTION. 231<br />

From experiments like the one sketched above we find equ<strong>at</strong>ions of st<strong>at</strong>e.<br />

For example, the previous experiment could have given us V (T, p) <strong>at</strong> standard<br />

pressure. We can then increase the pressure, too, and map out the whole curve.<br />

We will probably discover phase transitions in the m<strong>at</strong>erial. At th<strong>at</strong> point we<br />

need to make a judgement. For example, we see the change from ice to vapor.<br />

Is the vapor the same system as the ice or not? The answer depends on the<br />

context of our problem! Although the fact th<strong>at</strong> ice and vapor are different seems<br />

trivial to us, it assumes th<strong>at</strong> we can do a measurement to distinguish between<br />

the two phases! We ”see” the difference, literally. But if we only measure the<br />

volume, we cannot see the difference, apart from an observed jump in volume.<br />

If you are in the situ<strong>at</strong>ion th<strong>at</strong> ice and vapor are the same system, wh<strong>at</strong> if you<br />

further increase the temper<strong>at</strong>ure and the molecules in the vapor break apart?<br />

Gradients.<br />

In the last case, are we still studying the same system?<br />

In the previous subsection we assumed th<strong>at</strong> the change in control variable<br />

(the poke) is uniform over the whole sample. Often th<strong>at</strong> is not true, and we all<br />

know th<strong>at</strong> keeping a temper<strong>at</strong>ure uniform is hard to do. Again, it is a m<strong>at</strong>ter<br />

of tolerance. If the devi<strong>at</strong>ions from uniformity are small enough according to<br />

a specific standard we can ignore them. But we can also turn this around.<br />

Suppose we make the control variable non-uniform on purpose. For example,<br />

we take a block of ice and he<strong>at</strong> one end with a torch. Wh<strong>at</strong> happens? Is it<br />

interesting?<br />

How large is large?<br />

The question th<strong>at</strong> always needs to be answered is how large is the applied<br />

difference of control variable. For example, when we apply a voltage across a<br />

m<strong>at</strong>erial we observe a current. If the voltage is too large, we see a spark. In<br />

between, there can be many interesting types of behavior, too. In non-linear<br />

dynamics you studied the transitions from uniform convection to convection<br />

rolls. There was period doubling en route to chaos. This is all very complic<strong>at</strong>ed,<br />

and not easy to describe. The only part where we have some hope of success is<br />

when the currents are uniform. But even there we need to be careful.<br />

Linear response only.<br />

We are all familiar with Ohm’s law. The current I in response to a voltage<br />

difference V is given by I = V<br />

R . The resistance R for a given piece of m<strong>at</strong>erial<br />

can be measured once and for all, and does not vary for th<strong>at</strong> piece of m<strong>at</strong>erial if


232 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

we change the applied voltage. Hence the resistance R is a m<strong>at</strong>erial’s property<br />

and only depends on the st<strong>at</strong>e of the m<strong>at</strong>erial. For example, it will depend<br />

on the temper<strong>at</strong>ure of the m<strong>at</strong>erial! It also depends on the geometry of the<br />

m<strong>at</strong>erial, however. Make a wire twice as long, and the resistance doubles. This<br />

dependence is easy to scale out, though. In stead of the current I we use the<br />

current density j, the amount of current passing through a unit area. In stead<br />

of the applied voltage we use the electric field E. When the field is constant we<br />

have V = EL where L is the length of the m<strong>at</strong>erial. If the cross sectional area<br />

is A, we arrive <strong>at</strong> j = I EL<br />

RA<br />

A = RA . The quantity L is called the resistivity of the<br />

m<strong>at</strong>erial, and is a property independent of the geometry. It is more useful to<br />

replace it by its inverse, the conductivity σ. Also, we have to realize th<strong>at</strong> both<br />

current and field are vector quantities and hence we write in general:<br />

⃗j = σ · ⃗ E (6.1)<br />

where in the general case σ is a tensor, a three by three m<strong>at</strong>rix in this case. In<br />

cubic m<strong>at</strong>erials, however, the conductivity is simply a number times the unit<br />

m<strong>at</strong>rix. But in layered m<strong>at</strong>erials we know th<strong>at</strong> the conductivity parallel to the<br />

layers is different from the conductivity perpendicular to the layers!<br />

Higher order terms.<br />

In general, the conductivity σ depends on the external field, and this dependence<br />

can be very complic<strong>at</strong>ed. We restrict ourselves to the low field limit. In<br />

th<strong>at</strong> case we can write:<br />

⃗j = σ1 · ⃗ E + σ2 · ⃗ E ⃗ E + · · · (6.2)<br />

and the coefficients in the Taylor expansion are independent of the field. The<br />

second order conductivity is also a m<strong>at</strong>erials property, but is not used very<br />

often. Hence we will assume th<strong>at</strong> this term is not important, and only use<br />

linear response.<br />

Discuss types of experiments th<strong>at</strong> measure higher order response terms.<br />

Typical transport situ<strong>at</strong>ions.<br />

There are a number of cases where transport is important. For example, we<br />

can study the flow of gas through pipes. From a technical point of view, this<br />

is a very important applic<strong>at</strong>ion. The important parameters are pressure and<br />

temper<strong>at</strong>ure. In this scenario one typically follows the fluid while it is flowing<br />

through the system, but th<strong>at</strong> is not necessary. I will not discuss this applic<strong>at</strong>ion<br />

here, since I know the basics only. Th<strong>at</strong> is my shortcoming, however, and does<br />

not reflect <strong>at</strong> all on the importance of the subject!


6.2. SOME THERMO-ELECTRIC PHENOMENA. 233<br />

The applic<strong>at</strong>ion I will use as an example is conductivity in a solid. If we<br />

apply an electric field we will induce a current. In solids the pressure is typically<br />

constant <strong>at</strong> standard pressure, and we will ignore pressure as a variable. This<br />

allows us to discuss the process in terms of densities as shown l<strong>at</strong>er. But temper<strong>at</strong>ure<br />

is an important control parameter, as is magnetic field. We will not<br />

include magnetic field effects extensively, however. This is a detail only, and is<br />

easy to incorpor<strong>at</strong>e, as long as one is aware of extra minus signs. These minus<br />

signs are due to the fact th<strong>at</strong> reversing a current is equivalent to time reversal,<br />

but th<strong>at</strong> only works when one also inverts the magnetic fields.<br />

Steady st<strong>at</strong>e.<br />

Another important point is th<strong>at</strong> we will only discuss transport in the steady<br />

st<strong>at</strong>e. We assume th<strong>at</strong> all transients due to changes in the control parameters<br />

have died out. Hence there are no time dependencies in the applied fields, and<br />

we can write ⃗ E(⃗r) and T (⃗r). Since there are currents, there is time dependence<br />

in corresponding extensive parameters. Something is moving! For example, in<br />

our case we follow the electrons in a solid. The electrons are the vehicle for the<br />

electric current. In a steady st<strong>at</strong>e, the number of electrons in a certain volume is<br />

constant, and equal numbers of electrons are moving in and out. We also assume<br />

th<strong>at</strong> there are enough collisions th<strong>at</strong> these electrons are in thermal equilibrium<br />

with the ions in the solid, and have the same temper<strong>at</strong>ure. Therefore, the length<br />

scale implead by the applied external field should be long compared to the mean<br />

free p<strong>at</strong>h of the electrons. Hence we do not discuss so-called ballistic transport,<br />

where electrons shoot through a medium without interactions. These kind of<br />

processes need a different framework of description.<br />

In our applic<strong>at</strong>ion he<strong>at</strong> is a transport of energy by both the electrons and by<br />

the ions in the solid, via l<strong>at</strong>tice vibr<strong>at</strong>ions. We will not discuss the l<strong>at</strong>ter part,<br />

but real experiments will always have to take th<strong>at</strong> contribution into account!<br />

6.2 Some thermo-electric phenomena.<br />

Electric conductivity.<br />

Electric conductivity is a very familiar example of a transport phenomenon.<br />

It is so familiar, th<strong>at</strong> we easily make mistakes in using conductivity! The naive<br />

approach is to apply a voltage difference to a sample and to measure the current.<br />

Why is this naive? Because we need to connect the sample to the voltage source<br />

and the current meter! These wires have their own resistance, but even more<br />

important, the contact between the wires and the sample have resistance.<br />

An improvement is to reverse the procedure. A known current is driven<br />

through a sample. In a steady st<strong>at</strong>e there is no accumul<strong>at</strong>ion of charge, and we<br />

know th<strong>at</strong> all the current leaving the current source will go through the sample.


234 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

A<br />

B<br />

T1 T2<br />

Figure 6.1: Essential geometry of a thermocouple.<br />

Well, <strong>at</strong> least as long as there are no leaks! We than take two different wires<br />

and connect these to the sample in between the current leads. We measure the<br />

voltage difference between these wires, and from th<strong>at</strong> derive the resistance. We<br />

do not worry about the resistivity of the contacts, since the driving current does<br />

not pass though them, and since the measuring current can be made extremely<br />

small. A very ingenious method indeed!<br />

Can there be a problem? Yes! Several, actually. Because there is resistance,<br />

there will be a production of thermal energy. In such a case we can either<br />

keep the temper<strong>at</strong>ure constant or do the measurement in an adiab<strong>at</strong>ic fashion.<br />

Would th<strong>at</strong> make a difference? It probably will! L<strong>at</strong>er we will show th<strong>at</strong> it<br />

indeed does make a difference, and th<strong>at</strong> the adiab<strong>at</strong>ic conductivity is different<br />

from the constant temper<strong>at</strong>ure conductivity.<br />

Wh<strong>at</strong> else could go wrong? The temper<strong>at</strong>ure in the steady st<strong>at</strong>e of the<br />

experiment might not be uniform! In th<strong>at</strong> case the contacts used for measuring<br />

the voltage difference could be <strong>at</strong> different temper<strong>at</strong>ures, which is known to have<br />

consequences!<br />

Thermocouples.<br />

Consider the setup in figure 6.1. Two different m<strong>at</strong>erials (A and B) are<br />

used. M<strong>at</strong>erial B is cut in the middle, and the ends are connected to input<br />

leads. At these input leads we can either apply a voltage or drive a current.<br />

The connections between m<strong>at</strong>erial A and B are <strong>at</strong> two different temper<strong>at</strong>ures<br />

T1 and T2. During the experiment there will also be he<strong>at</strong> going through both<br />

m<strong>at</strong>erials. Since thermal energy will also be produced everywhere there will be<br />

a he<strong>at</strong> flow out of the m<strong>at</strong>erials into the environment, represented by the arrows<br />

in figure 6.1. We will now use this device in several setups.


6.2. SOME THERMO-ELECTRIC PHENOMENA. 235<br />

Seebeck effect.<br />

We measure the voltage difference V between the input leads when there<br />

is a temper<strong>at</strong>ure difference between the endpoints. In this case there is no<br />

electric current through the thermocouple. The voltage difference depends on<br />

the m<strong>at</strong>erials A and B only, and for small temper<strong>at</strong>ure differences we have<br />

V = −αAB(T2 − T1) (6.3)<br />

where we define the signs in the following manner. If T1 > T2 and if the<br />

thermocouple were connected to a resistor, the coefficient αAB is positive if the<br />

resulting electric current goes from point 1 to point 2 through m<strong>at</strong>erial A and<br />

returns through m<strong>at</strong>erial B and the resistor. If we label the input leads by 1<br />

and 2 according to which end they are connected to, we have V = V2 − V1.<br />

The extra minus sign in the definition of α is due to the distinction between<br />

internal and external currents, which follow an opposite p<strong>at</strong>h. This effect, the<br />

Seebeck effect, was discovered in 1821! Note th<strong>at</strong> this is before the discovery of<br />

Ohm’s law!<br />

Peltier effect.<br />

How can we use the Seebeck effect in a practical manner?<br />

We can also do the opposite. We drive a current I through the thermocouple.<br />

The current is positive if we drive it similar to the previous subsection. The<br />

current is driven into lead 1 and extracted <strong>at</strong> lead 2, if positive. In this case<br />

a positive amount of he<strong>at</strong> Q is gener<strong>at</strong>ed <strong>at</strong> connection 1 between m<strong>at</strong>erials A<br />

and B and the same amount of he<strong>at</strong> is absorbed <strong>at</strong> the other end, connection<br />

2. In addition to this he<strong>at</strong> flow between 2 and 1 there is, of course, also extra<br />

he<strong>at</strong> gener<strong>at</strong>ed all along the p<strong>at</strong>h, because we have both electric and thermal<br />

currents. But the Peltier effect only rel<strong>at</strong>es the he<strong>at</strong> flow between 1 and 2 and<br />

we have<br />

Q = ΠABI (6.4)<br />

This effect was discovered only ten years or so after the Seebeck effect. It has<br />

one very important applic<strong>at</strong>ion. We can use the Peltier effect to cool a sample<br />

connected to lead 2!<br />

Kelvin found an empirical rel<strong>at</strong>ion between the Seebeck effect and the Peltier<br />

effect in 1854:<br />

ΠAB = T αAB<br />

(6.5)


236 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

Thomson he<strong>at</strong>.<br />

Suppose we have both a he<strong>at</strong> flow and an electric current through the thermocouple.<br />

In th<strong>at</strong> case he<strong>at</strong> is gener<strong>at</strong>ed <strong>at</strong> every point along the thermocouple.<br />

This is represented by the arrows in figure 6.1. Suppose the distance along m<strong>at</strong>erial<br />

A (or B) is measured by a parameter x. The he<strong>at</strong> gener<strong>at</strong>ed <strong>at</strong> position<br />

x in an interval ∆x is given by ∆Q, and Thomson found th<strong>at</strong><br />

He also derived (empirically!) th<strong>at</strong><br />

∆Q = τI dT<br />

∆x (6.6)<br />

dx<br />

τA − τB = T dαAB<br />

dT<br />

(6.7)<br />

Note th<strong>at</strong> the coefficients defined so far always depend on the difference between<br />

two m<strong>at</strong>erials. Nowadays we use a reference m<strong>at</strong>erial <strong>at</strong> all temper<strong>at</strong>ures,<br />

and define the absolute thermoelectric coefficients with respect to this standard.<br />

For the absolute values we also have<br />

Magnetic effects.<br />

τ = T dα<br />

dT<br />

Wh<strong>at</strong> is the rel<strong>at</strong>ion between Thomson and Kelvin?<br />

(6.8)<br />

When we apply a magnetic field perpendicular to the thermocouple, the<br />

Lorentz force gives a sideways displacement of the charge carriers. These are<br />

transverse effects. The following four combin<strong>at</strong>ions are useful:<br />

• Hall: Drive an electric current through the m<strong>at</strong>erial and measure the<br />

transverse electric field.<br />

• Nernst: Drive a he<strong>at</strong> flow through the m<strong>at</strong>erial and measure the transverse<br />

electric field.<br />

• Ettingshausen: Drive an electric current through the m<strong>at</strong>erial and measure<br />

the transverse temper<strong>at</strong>ure gradient.<br />

• Righi-Leduc: Drive a he<strong>at</strong> flow through the m<strong>at</strong>erial and measure the<br />

transverse temper<strong>at</strong>ure gradient.<br />

For detailed discussions see books on solid st<strong>at</strong>e physics.


6.3. NON-EQUILIBRIUM THERMODYNAMICS. 237<br />

6.3 Non-equilibrium <strong>Thermodynamics</strong>.<br />

Is it possible?<br />

The words non-equilibrium thermodynamics do not seem to make sense <strong>at</strong><br />

first glance. By definition, thermodynamics describes a m<strong>at</strong>erial in equilibrium,<br />

and it is only then th<strong>at</strong> we are able to have equ<strong>at</strong>ions of st<strong>at</strong>e. If we are not in<br />

equilibrium, some st<strong>at</strong>e functions do not have their correct values (the entropy<br />

is not maximal, or some free energy is not minimal), and there is no correct<br />

m<strong>at</strong>hem<strong>at</strong>ical description.<br />

In order to discuss the situ<strong>at</strong>ion presented above, we need to consider the<br />

microscopic details of the m<strong>at</strong>erial, and follow a system as a function of time by<br />

solving the equ<strong>at</strong>ions of motion. This leads to concepts like chaotic dynamics,<br />

and some very interesting physics. Boltzmann was probably one of the first to<br />

give a beginning of an answer to this problem.<br />

When devi<strong>at</strong>ions are small.<br />

Wh<strong>at</strong> happens when a system is almost in equilibrium? When the devi<strong>at</strong>ions<br />

from equilibrium are small? Perturb<strong>at</strong>ion theory should apply, and a complete<br />

knowledge of the equilibrium st<strong>at</strong>e should be sufficient to understand wh<strong>at</strong> is<br />

happening. This is the situ<strong>at</strong>ion in which we can still use the formulas from<br />

equilibrium thermodynamics, and discuss wh<strong>at</strong> happens when we are not quite<br />

in equilibrium.<br />

Local stability.<br />

Suppose we divide our m<strong>at</strong>erials in small cells, with volume ∆V each. In each<br />

cell, with position ⃗ri we have a an amount of m<strong>at</strong>erial ∆Ni and a temper<strong>at</strong>ure<br />

Ti. We assume th<strong>at</strong> the m<strong>at</strong>erial in each cell is internally in thermodynamic<br />

equilibrium. Therefore, the usual equ<strong>at</strong>ions of st<strong>at</strong>e will give us the values of the<br />

other st<strong>at</strong>e variables pi, µi, and Si, as well as the energy Ui. We assume th<strong>at</strong><br />

we deal with the standard pV T system, if not we need to add one variable for<br />

each additional pair to the set of variables th<strong>at</strong> we need to know. This scenario<br />

is called local stability.<br />

Limits.<br />

In order for local stability to exist, the volume ∆V needs to be large. Wh<strong>at</strong><br />

is large? In general, this volume needs to contain many molecules. On the other<br />

hand, we would like to evalu<strong>at</strong>e quantities like N = ∑<br />

∆Ni via integrals. If we<br />

define the density in cell i by ni = ∆Ni<br />

∆V<br />

we would like to replace<br />

i


238 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

∑<br />

∫<br />

ni∆V ⇒ n(⃗r)d 3 r (6.9)<br />

i<br />

and in order to do so we need to take the limit ∆V → 0, which obviously causes<br />

problems. The answer is again found by assuming th<strong>at</strong> the vari<strong>at</strong>ions in the<br />

density and temper<strong>at</strong>ure are small. This implies th<strong>at</strong>, if we follow the Riemann<br />

sum ∑<br />

ni∆V as a function of decreasing value of ∆V , this sum reaches a stable<br />

i<br />

(converged) value while the volume ∆V is still large on a microscopic scale. At<br />

th<strong>at</strong> point we stop and do not enter the region where ∆V becomes so small th<strong>at</strong><br />

we see the effects of the microscopic structure.<br />

Tests for validity.<br />

The ideas developed in the previous paragraph give us a definition when we<br />

can still use the equ<strong>at</strong>ions of thermodynamics, but also allow for fluctu<strong>at</strong>ions.<br />

It also gives a recipe for a numerical test of this condition. We divide space in<br />

cells and calcul<strong>at</strong>e all quantities with the given cell size. Next, we choose smaller<br />

cells, and recalcul<strong>at</strong>e all our quantities. This gives us a plot of the value of the<br />

quantity versus cell size. By extrapol<strong>at</strong>ing these values to those obtained with<br />

∆V = 0 we find the error in our calcul<strong>at</strong>ions for each cell size. On the other<br />

hand, we also have an error due to the fact th<strong>at</strong> our cells only contain a finite<br />

number of particles. This is an error not included in the calcul<strong>at</strong>ions above, and<br />

has to be added. This error is of the form 1<br />

√ N<br />

V<br />

√ ∆Ni<br />

, which can be approxim<strong>at</strong>ed by<br />

1<br />

∆V . Adding these terms together we get a minimum error <strong>at</strong> some optimal<br />

value of the cell size, and if this minimal error is acceptable, our model is useful.<br />

If not, the sp<strong>at</strong>ial vari<strong>at</strong>ions in the density are too fast. From now on we will<br />

assume th<strong>at</strong> our procedure gives acceptable results according to this test.<br />

Choice of fields.<br />

In a standard pV T system we need to specify three st<strong>at</strong>e variables to completely<br />

identify the st<strong>at</strong>e. At each position ⃗r we fix the volume d 3 r and hence we<br />

need two other st<strong>at</strong>e variables to completely specify wh<strong>at</strong> is happening. Conceptually<br />

the easiest are n(⃗r) and T (⃗r), because we can imagine th<strong>at</strong> these can be<br />

measured very well. But we could also choose quantities like the energy density<br />

u(⃗r), which can have advantages in the theoretical formul<strong>at</strong>ions.<br />

Different coordin<strong>at</strong>e frames.<br />

Another choice which is often made is to use cells th<strong>at</strong> contain a fixed amount<br />

of m<strong>at</strong>erial ∆N. In this case the cells have variable volume d 3 v = n(⃗r)d 3 r, but<br />

all formulas can be derived by simply adding this weight function. All densities<br />

in this case are per mole, and not per unit volume as in the previous case.


6.3. NON-EQUILIBRIUM THERMODYNAMICS. 239<br />

Sometimes we want to study liquids th<strong>at</strong> support macroscopic currents. At<br />

each point we find a velocity field ⃗v(⃗r) which tells us in which direction the liquid<br />

is flowing <strong>at</strong> th<strong>at</strong> point. It can be very useful to use a coordin<strong>at</strong>e frame th<strong>at</strong> is<br />

moving with the liquid and to follow the amount of m<strong>at</strong>erial in the original cell.<br />

Work between cells.<br />

We need to digress a bit on how we define volume. If we have nothing but<br />

the m<strong>at</strong>erial itself, as in the case of fluid flow, we can only refer to the m<strong>at</strong>erial<br />

itself to define volume. This then leads autom<strong>at</strong>ically to the idea of following<br />

a certain amount of m<strong>at</strong>erial in fluid flow, as described above. When we study<br />

electron transport in solids, though, we have the <strong>at</strong>omic l<strong>at</strong>tice as a reference.<br />

Hence in the cases we will study, volume is defined by the <strong>at</strong>omic l<strong>at</strong>tice. If the<br />

electrons would not interact with the l<strong>at</strong>tice, the gas of electrons could expand or<br />

contract with respect to the l<strong>at</strong>tice, and the amount of electrons in a cell defined<br />

by the <strong>at</strong>omic structure would change because of correl<strong>at</strong>ed motion (pressure<br />

driven) and diffusion (chemical potential driven). In the rest of this chapter we<br />

assume th<strong>at</strong> the electrons interact strongly with the <strong>at</strong>omic l<strong>at</strong>tice. This means<br />

th<strong>at</strong> pressure difference do not lead to correl<strong>at</strong>ed transport of electrons, but to<br />

changes in the <strong>at</strong>omic configur<strong>at</strong>ions. The <strong>at</strong>oms act like pistons. As long as<br />

pressure gradients are small, the <strong>at</strong>oms will not move far because inter<strong>at</strong>omic<br />

forces are strong. So we can assume th<strong>at</strong> the cells defined by the <strong>at</strong>oms have a<br />

fixed volume, and th<strong>at</strong> pressure differences are simply absorbed by the l<strong>at</strong>tice<br />

in reaction.<br />

In our formul<strong>at</strong>ion the size of the cells is fixed, and mechanical work of the<br />

pdV variety is therefore not possible. The only two ways the cells can exchange<br />

energy with each other is to do chemical work µdN, which is an exchange of<br />

m<strong>at</strong>erial between the cells, or use he<strong>at</strong> transfer T dS. Of course, if we would have<br />

chosen cells with a fixed amount of m<strong>at</strong>erial, the situ<strong>at</strong>ion would be different<br />

and we could use mechanical work as well as he<strong>at</strong>.<br />

This situ<strong>at</strong>ion is typical for electronic transport in a solid. The electrons<br />

diffuse through the m<strong>at</strong>erial due to gradients in the chemical potential and/or<br />

temper<strong>at</strong>ure. There are corresponding gradients in the pressure, but those are<br />

compens<strong>at</strong>ed by the pressure due to the l<strong>at</strong>tice. We assume th<strong>at</strong> the l<strong>at</strong>tice<br />

is stiff, and th<strong>at</strong> changes in the l<strong>at</strong>tice pressure have very small effects on the<br />

positions of the <strong>at</strong>oms, so we can assume th<strong>at</strong> we do have small volume elements<br />

th<strong>at</strong> do not change.<br />

As a result, the formulas we will derive in the next section do not apply to<br />

fluid flow, in which there are currents due to pressure gradients, as described<br />

by Bernouilli. But we know th<strong>at</strong> in thermodynamic equilibrium the pressure<br />

gradient is a function of the gradients in the other intensive st<strong>at</strong>e variables, and<br />

we can therefore always express our currents in terms of the other gradients, and<br />

retain a formalism similar to the one to be developed. We just have to be very<br />

careful in making connections between transport coefficients and applied fields.<br />

By restricting ourselves to electron diffusion we avoid those complic<strong>at</strong>ions. See


240 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

textbooks on fluid dynamics for more details.<br />

6.4 Transport equ<strong>at</strong>ions.<br />

Slow processes only.<br />

In the previous section we have developed a model in which we have local<br />

equilibrium, because densities are varying slowly in space. But because there<br />

is a change in density and temper<strong>at</strong>ure between cells, there will be a flow of<br />

m<strong>at</strong>erial and he<strong>at</strong> in order to reach equilibrium. We will assume th<strong>at</strong> this flow<br />

is very slow. In other words, the local densities are changing, but <strong>at</strong> a r<strong>at</strong>e<br />

which is much slower than the r<strong>at</strong>e in which these cells get back to equilibrium<br />

after a small amount of m<strong>at</strong>erial or he<strong>at</strong> has been exchanged. This is a n<strong>at</strong>ural<br />

consequence of the assumption th<strong>at</strong> the system is always in local equilibrium.<br />

If transport is too fast, we cannot maintain local equilibrium.<br />

Continuity equ<strong>at</strong>ion.<br />

If we consider an arbitrary small volume ∆V and follow the change of an<br />

extensive quantity X in th<strong>at</strong> volume, the change d∆X<br />

dt in th<strong>at</strong> quantity is the<br />

sum of the cre<strong>at</strong>ion and/or destruction of X and the flow of X through the<br />

surface ∆S of the volume element. The flow is given by a current density ⃗ JX<br />

by following the amount of X going through an area ∆A with surface normal<br />

ˆn in a time ∆t. The corresponding amount of X flowing through this area is<br />

Therefore, the r<strong>at</strong>e of change in X is<br />

∆X = ⃗ JX · ˆn∆A∆t (6.10)<br />

∫<br />

d<br />

xd<br />

dt ∆V<br />

3 ∫<br />

V = Px(⃗r, t)d<br />

∆V<br />

3 <br />

V − ⃗JX · ˆnd<br />

∆S<br />

2 S (6.11)<br />

where the first term on the right, Px, is the local cre<strong>at</strong>ion (which could be<br />

neg<strong>at</strong>ive) per unit volume and the second term the flow through the surface.<br />

The quantity x is the density of X, or the amount of X per unit volume.<br />

The surface normal in this definition points outwards from the volume element.<br />

Using Gauss’ theorem this gives<br />

∫<br />

d<br />

xd<br />

dt ∆V<br />

3 ∫<br />

V = Pxd<br />

∆V<br />

3 ∫<br />

V −<br />

∆V<br />

and since the volume is arbitrary we have <strong>at</strong> each point in space:<br />

∂x<br />

∂t = Px − ⃗ ∇ · ⃗ JX<br />

⃗∇ · ⃗ JXd 3 V (6.12)<br />

(6.13)


6.4. TRANSPORT EQUATIONS. 241<br />

We have replaced the total deriv<strong>at</strong>ive by a partial deriv<strong>at</strong>ive, since we look <strong>at</strong><br />

the change of x(⃗r, t) <strong>at</strong> each point. The production is sometimes written as a<br />

time deriv<strong>at</strong>ive, too, and in order to distinguish one uses a not<strong>at</strong>ion like<br />

Px = Dx<br />

(6.14)<br />

Dt<br />

This not<strong>at</strong>ion (or vari<strong>at</strong>ions of it), is widely used in physics, and it reflects the<br />

change in quantity X due to both cre<strong>at</strong>ion and transport. In a coordin<strong>at</strong>e frame<br />

moving with the liquid this is the only term th<strong>at</strong> survives, which is a main reason<br />

for choosing such a frame. The continuity equ<strong>at</strong>ion is in this not<strong>at</strong>ion:<br />

Dx<br />

Dt (⃗r, t) = Px(⃗r, t) = ∂x<br />

∂t (⃗r, t) + ⃗ ∇ · ⃗ JX<br />

Is Dx<br />

Dt<br />

a real deriv<strong>at</strong>ive?<br />

Extensive st<strong>at</strong>e properties are conserved.<br />

(6.15)<br />

Consider two adjacent cells. If m<strong>at</strong>erial flows from one cell to the other, the<br />

total amount of m<strong>at</strong>erial does not change. In other words, if a cell loses m<strong>at</strong>erial,<br />

this m<strong>at</strong>erial is not cre<strong>at</strong>ed or destroyed, but transported to neighboring cells.<br />

For example, the energy of one cell changes because energy is flowing in to or<br />

out of the cell. We assume th<strong>at</strong> we do not have any chemical reactions, and<br />

there is no cre<strong>at</strong>ion of energy or m<strong>at</strong>erial. Hence we assume th<strong>at</strong><br />

Since energy and m<strong>at</strong>erial are conserved, we now have<br />

and<br />

Pu(⃗r, t) = 0 (6.16)<br />

Pn(⃗r, t) = 0 (6.17)<br />

⃗∇ · ⃗ JU = − ∂u<br />

∂t<br />

(6.18)<br />

⃗∇ · ⃗ JN = − ∂n<br />

(6.19)<br />

∂t<br />

where u is the energy density and n is the m<strong>at</strong>ter density. Note th<strong>at</strong> the inclusion<br />

of chemical reactions is conceptually not difficult, just tedious, because now we<br />

need to distinguish between several types of m<strong>at</strong>erial and have to include partial<br />

time deriv<strong>at</strong>ives.<br />

Entropy production and fluxes.


242 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

Consider a cell which is divided in to two equal parts by an area ∆A with<br />

surface normal ˆn pointing from the left to the right. The width of each half is<br />

1<br />

2 ∆L. The left side of this cell changes its energy by an amount ∆Ul and its<br />

amount of m<strong>at</strong>erial by ∆Nl. Because we always have local equilibrium, we know<br />

th<strong>at</strong> ∆Ul = Tl∆Sl + µl∆Nl in this part of the cell. If we assume no contact<br />

outside the cell, the m<strong>at</strong>erial and energy need to come from the right part of<br />

the cell, and therefore we have<br />

∆Ul = −∆Ur<br />

∆Nl = −∆Nr<br />

(6.20)<br />

(6.21)<br />

But local equilibrium gives us the conditions th<strong>at</strong> ∆Ur = Tr∆Sr + µr∆Nr for<br />

th<strong>at</strong> half of the cell. Therefore we find th<strong>at</strong><br />

∆Sl + ∆Sr = 1<br />

∆Ul − µl<br />

∆Nl + 1<br />

∆Ur − µr<br />

∆Nr<br />

Tl<br />

or by using the conserv<strong>at</strong>ion laws we find th<strong>at</strong> the entropy production is<br />

∆Sl + ∆Sr = ( 1<br />

Tr<br />

Tl<br />

Tr<br />

− 1<br />

)∆Ur + (<br />

Tl<br />

µl<br />

Tl<br />

Tr<br />

− µr<br />

)∆Nr<br />

Tr<br />

(6.22)<br />

(6.23)<br />

The amount of energy flowing through the area ∆A is ∆Ur and the amount<br />

of m<strong>at</strong>erial ∆Nr, all in a time ∆t. If the local energy current density <strong>at</strong> the<br />

origin is ⃗ JU and the local m<strong>at</strong>erial current density ⃗ JN , then we have<br />

The entropy produced in this process is given by<br />

∆S = ∆Sl + ∆Sr =<br />

∆Ur = ⃗ JU · ˆn∆A∆t (6.24)<br />

∆Nr = ⃗ JN · ˆn∆A∆t (6.25)<br />

(<br />

( 1<br />

−<br />

Tr<br />

1<br />

)<br />

Tl<br />

⃗ JU · ˆn + ( µl<br />

Tl<br />

− µr<br />

)<br />

Tr<br />

⃗ )<br />

JN · ˆn ∆A∆t (6.26)<br />

This can be rel<strong>at</strong>ed to the gradients in the following manner:<br />

1<br />

−<br />

Tr<br />

1<br />

Tl<br />

= ⃗ ( )<br />

1<br />

∇ · ˆn<br />

T<br />

1<br />

∆L<br />

2<br />

(6.27)<br />

and similar for the second term. This gives:<br />

∆S = 1<br />

(<br />

⃗∇<br />

2<br />

( )<br />

1<br />

· ˆn<br />

T<br />

⃗ JU · ˆn − ⃗ ∇<br />

(<br />

µ<br />

)<br />

· ˆn<br />

T<br />

⃗ )<br />

JN · ˆn ∆A∆L∆t (6.28)<br />

and therefore the r<strong>at</strong>e of change of the production of the entropy density s<br />

(entropy per unit volume) for this process is given by:


6.4. TRANSPORT EQUATIONS. 243<br />

Ps = 1<br />

(<br />

⃗∇<br />

2<br />

( )<br />

1<br />

· ˆn<br />

T<br />

⃗ JU · ˆn − ⃗ ∇<br />

(<br />

µ<br />

)<br />

· ˆn<br />

T<br />

⃗ )<br />

JN · ˆn<br />

(6.29)<br />

In this equ<strong>at</strong>ion we have chosen an arbitrary direction of the dividing partition.<br />

Of course, we need to take into account the complete flow. Consider a small<br />

cubic cell with faces orthogonal to the x, y, and z direction. The local entropy<br />

density production for the flow through each of the faces of this cube is given by<br />

the expression above, with either ˆn = ±ˆx, ˆn = ±ˆy, or ˆn = ±ˆz. We can use the<br />

expression for one side of the partition only, since we use the entropy density,<br />

or entropy per unit volume! If we add the effects the flow through all sides of<br />

the cube together we get the total entropy density production in this cubic cell.<br />

It is easy to see th<strong>at</strong><br />

∑<br />

( ⃗ A · ˆn)( ⃗ B · ˆn) = 2AxBx + 2AyBy + 2AzBz = 2 ⃗ A · ⃗ B (6.30)<br />

ˆn<br />

and hence the expression for the total energy density production is<br />

( ( )<br />

Ps = ⃗∇<br />

1<br />

·<br />

T<br />

⃗ JU − ⃗ (<br />

µ<br />

)<br />

∇ ·<br />

T<br />

⃗ )<br />

JN<br />

(6.31)<br />

Note th<strong>at</strong> using a cubic cell is not essential, it just makes the m<strong>at</strong>hem<strong>at</strong>ics much<br />

easier. Also note th<strong>at</strong> we get a factor two, because <strong>at</strong> each point we need to<br />

consider the flow into th<strong>at</strong> point from one side and the flow out of th<strong>at</strong> point<br />

towards the other side!<br />

Entropy flow.<br />

Wh<strong>at</strong> does Ps exactly mean?<br />

Next we define the flow of entropy ⃗ JS by rel<strong>at</strong>ing it to the flow of energy<br />

and the flow of particles.<br />

( )<br />

1<br />

(<br />

µ<br />

)<br />

⃗JS = ⃗JU − ⃗JN (6.32)<br />

T T<br />

As a consequence we find th<strong>at</strong><br />

⃗∇ · ⃗ JS =<br />

( )<br />

1<br />

⃗∇ ·<br />

T<br />

⃗ JU −<br />

(<br />

µ<br />

)<br />

⃗∇ · JN<br />

⃗ +<br />

T<br />

⃗ ∇<br />

( )<br />

1<br />

·<br />

T<br />

⃗ JU − ⃗ ∇<br />

(<br />

µ<br />

)<br />

·<br />

T<br />

⃗ JN<br />

(6.33)<br />

and using the continuity equ<strong>at</strong>ions for the energy and m<strong>at</strong>erial current we obtain:<br />

⃗∇ · ⃗ JS = −<br />

( )<br />

1 ∂u<br />

T ∂t +<br />

(<br />

µ<br />

)<br />

∂n<br />

T ∂t + ⃗ ∇<br />

( )<br />

1<br />

·<br />

T<br />

⃗ JU − ⃗ ∇<br />

(<br />

µ<br />

)<br />

·<br />

T<br />

⃗ JN<br />

(6.34)


244 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

Since we always have local equilibrium, we can use the first law of thermodynamics:<br />

dU = T dS + µdN (6.35)<br />

and show the sum of the first two terms is equal to − ∂s<br />

∂t and hence we find th<strong>at</strong><br />

for the local production of entropy we get<br />

( ( )<br />

Ps = ⃗∇<br />

1<br />

·<br />

T<br />

⃗ JU − ⃗ (<br />

µ<br />

)<br />

∇ ·<br />

T<br />

⃗ )<br />

JN<br />

(6.36)<br />

as we derived before. This justifies th<strong>at</strong> we defined the entropy current correctly.<br />

Transport equ<strong>at</strong>ions.<br />

Why is − ∂s<br />

∂t = − ( )<br />

1 ∂u<br />

T ∂t + ( ) µ ∂n<br />

T ∂t ?<br />

In our example we assume th<strong>at</strong> locally we have a pV T system, and because<br />

we define subregions by constant volume elements, only energy and mass can<br />

flow. In a more general case, we could also include other quantities, like the<br />

flow of magnetiz<strong>at</strong>ion. In this case we have for the first law in each local cell<br />

(with dV = 0) an equ<strong>at</strong>ion of the form:<br />

T dS = ∑<br />

xidXi<br />

which after dividing by T takes the form:<br />

Therefore, we have in the general case:<br />

and<br />

dS = ∑<br />

ϕidXi<br />

i<br />

i<br />

⃗JS = ∑<br />

i<br />

ϕi ⃗ Ji<br />

Ps = ∑<br />

⃗∇ϕi · ⃗ Ji<br />

i<br />

(6.37)<br />

(6.38)<br />

(6.39)<br />

(6.40)<br />

In general, each current will depend on the values of all intensive parameters<br />

and their gradients. We have to keep in mind, however, th<strong>at</strong> we assume th<strong>at</strong><br />

we are close to equilibrium. If all intensive parameters are constant, we are<br />

in equilibrium, and no net macroscopic flows are present. When we are close<br />

to equilibrium, the gradients in the intensive parameters will be small, and as<br />

a consequence we can expand the formulas for the currents in terms of the


6.5. MACROSCOPIC VERSUS MICROSCOPIC. 245<br />

gradients of the intensive parameters. An important case in practice is the<br />

situ<strong>at</strong>ion where the gradients are so small th<strong>at</strong> we need only linear terms. We<br />

will restrict ourselves to th<strong>at</strong> case and only study linear processes. The basic<br />

equ<strong>at</strong>ions are now<br />

⃗Ji = ∑<br />

Lik · ⃗ ∇ϕk<br />

k<br />

(6.41)<br />

where the m<strong>at</strong>rix coefficients Lik are functions of the intensive parameters (or<br />

of the combin<strong>at</strong>ions ϕi) only. The currents we consider here are only currents in<br />

energy, m<strong>at</strong>ter, etc. and not in the entropy. Of course, we can always elimin<strong>at</strong>e<br />

one of these currents in favor of the entropy current. This has the advantage<br />

th<strong>at</strong> we now have he<strong>at</strong> transport included as a separ<strong>at</strong>e entity.<br />

Onsager rel<strong>at</strong>ions.<br />

Are the variables Lik st<strong>at</strong>e functions?<br />

The coefficients Lik in the last formula obey some symmetry rel<strong>at</strong>ions. This<br />

is directly rel<strong>at</strong>ed to the fact th<strong>at</strong> fluctu<strong>at</strong>ions are not completely arbitrary, but<br />

are rel<strong>at</strong>ed to the entropy as described in chapter two. The other ingredient is<br />

time reversal symmetry. The general form of the rel<strong>at</strong>ion in the presence of a<br />

magnetic field is:<br />

Lik(T, µ, ⃗ H) = L T ki(T, µ, − ⃗ H) (6.42)<br />

where the transpose indic<strong>at</strong>es a transpose of the three by three m<strong>at</strong>rix in coordin<strong>at</strong>e<br />

space. Because magnetic fields give velocity dependent forces, we need<br />

to change the sign of the field when we reverse the direction of time. This is<br />

the cause for the minus sign in the last equ<strong>at</strong>ion. There are also additional<br />

minus signs in front of the equ<strong>at</strong>ions for properties th<strong>at</strong> mix magnetic and<br />

non-magnetic transport.<br />

6.5 Macroscopic versus microscopic.<br />

Thermal averages.<br />

In each cell we can define the local values of the extensive variables Xi and<br />

corresponding intensive variables ϕi. As mentioned before, we assume local<br />

equilibrium, and these variables are connected by equ<strong>at</strong>ions of st<strong>at</strong>e. If there<br />

are no external fields, all these values are independent of position and time. If<br />

there is an external field, however, the values in different cells will not be the


246 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

same anymore. For example, we can imagine th<strong>at</strong> there is a temper<strong>at</strong>ure field<br />

T (⃗r, t) present, and we want to study its consequences. Because we assume local<br />

equilibrium, the sp<strong>at</strong>ial and temporal vari<strong>at</strong>ion of this field has to be small. This<br />

implies th<strong>at</strong> we can study only the long wavelength and low frequency response<br />

of a system.<br />

The results will break down when either the wavelength is too small or<br />

the frequency is too high, and microscopic effects start to play a role. In the<br />

next sub-sections we rel<strong>at</strong>e the microscopic and macroscopic views, which also<br />

provides a proof for Onsager’s rel<strong>at</strong>ions. We do this in a general manner, independent<br />

of the microscopic n<strong>at</strong>ure of the m<strong>at</strong>erial. We simply consider the<br />

fluctu<strong>at</strong>ions in each cell. Some of the parameters describing these fluctu<strong>at</strong>ions,<br />

however, can only be calcul<strong>at</strong>ed in a microscopic theory. This is similar to the<br />

situ<strong>at</strong>ion in the Landau-Ginzburg theory as discussed in Chapter 4.<br />

Fluctu<strong>at</strong>ions.<br />

Assume th<strong>at</strong> the equilibrium system is determined by local values X0 i (⃗r, t)<br />

of the extensive parameters in each cell. Then in our non-equilibrium st<strong>at</strong>e we<br />

have in a given cell <strong>at</strong> point ⃗r the values Xi(⃗r, t) = X0 i (⃗r, t) + δXi(⃗r, t). The<br />

fluctu<strong>at</strong>ions are small if our cells contain many <strong>at</strong>oms (which is only possible in<br />

the long wavelength limit of external fields). If we now take a thermal average<br />

in each cell we find:<br />

and<br />

⟨δXi(⃗r, t)⟩ = 0 (6.43)<br />

⟨Xi(⃗r, t)⟩ = X 0 i (⃗r, t) (6.44)<br />

Since a thermal average is equivalent to a time average, we assume th<strong>at</strong> the<br />

fluctu<strong>at</strong>ions are much more rapid than the vari<strong>at</strong>ions in the external fields. In<br />

th<strong>at</strong> case we are able to do a time average of the quantity δXi(⃗r, t) over a time<br />

long enough th<strong>at</strong> it represents the thermodynamic limit, but short on the scale<br />

of the external perturb<strong>at</strong>ions. This is the reason for the low frequency limit.<br />

The fluctu<strong>at</strong>ions describe devi<strong>at</strong>ions from thermal equilibrium, and hence the<br />

quantities Xi(⃗r, t) and corresponding fields ϕi(⃗r, t) do not obey the equ<strong>at</strong>ions<br />

of st<strong>at</strong>e. Only the average quantities X 0 i (⃗r, t) and ϕ0 i<br />

(⃗r, t) do. The transport<br />

equ<strong>at</strong>ions derived in the previous section use these local thermal equilibrium<br />

values, and are defined in th<strong>at</strong> context only.<br />

Correl<strong>at</strong>ion functions.<br />

The fluctu<strong>at</strong>ions mentioned above are correl<strong>at</strong>ed, as we described in chapter<br />

two. Since the decay of fluctu<strong>at</strong>ions follows the laws of physics on the microscopic<br />

level, this correl<strong>at</strong>ion continues in time and we can discuss correl<strong>at</strong>ion<br />

coefficients of the form:


6.5. MACROSCOPIC VERSUS MICROSCOPIC. 247<br />

⟨δXi(⃗r, 0)δXj(⃗r ′ , t)⟩ (6.45)<br />

We can discuss correl<strong>at</strong>ions between fluctu<strong>at</strong>ions <strong>at</strong> difference points, since any<br />

excess in some extensive parameter has to flow away through neighboring cells.<br />

Again, we take averages to avoid all short time thermal fluctu<strong>at</strong>ions, and we<br />

still assume th<strong>at</strong> the cells are large compared to the detailed structure in the<br />

m<strong>at</strong>ter. Since the laws of n<strong>at</strong>ure are time invariant (here we assume th<strong>at</strong> there<br />

are no magnetic fields!) we have:<br />

⟨δXi(⃗r, 0)δXj(⃗r, t)⟩ = ⟨δXi(⃗r, 0)δXj(⃗r, −t)⟩ (6.46)<br />

and since we have transl<strong>at</strong>ional symmetry in time:<br />

⟨δXi(⃗r, 0)δXj(⃗r, t)⟩ = ⟨δXi(⃗r, t)δXj(⃗r, 0)⟩ (6.47)<br />

From this equ<strong>at</strong>ion we subtract ⟨δXi(⃗r, 0)δXj(⃗r, 0)⟩ and in the limit of small t<br />

we find<br />

⟨δXi(⃗r, 0) ∂δXj<br />

(⃗r, 0)⟩ = ⟨<br />

∂t<br />

∂δXi<br />

∂t (⃗r, 0)δXj(⃗r, 0)⟩ (6.48)<br />

Next we consider the current density th<strong>at</strong> is rel<strong>at</strong>ed to the quantity δXi in<br />

our cell, and call this current −→<br />

δJi. By using the symbol δ we make explicitly<br />

clear th<strong>at</strong> this is a microscopic current, rel<strong>at</strong>ed to fluctu<strong>at</strong>ions. Because there<br />

is no local cre<strong>at</strong>ion one has:<br />

This leads to<br />

∂δXj<br />

∂t = −⃗ ∇ · −→<br />

δJj<br />

(6.49)<br />

⟨δXi( ⃗ ∇ · −→<br />

δJj)⟩ = ⟨( ⃗ ∇ · −→<br />

δJi)δXj⟩ (6.50)<br />

where we have dropped the space and time arguments, since all quantities are<br />

<strong>at</strong> the same position ⃗r and same time t. This is important. If we want to<br />

rel<strong>at</strong>e fluctu<strong>at</strong>ions in quantities <strong>at</strong> different positions and times, we need the<br />

probability functions for these correl<strong>at</strong>ions, and we have no idea wh<strong>at</strong> they look<br />

like. On the other hand, probabilities for fluctu<strong>at</strong>ions <strong>at</strong> a given place and time<br />

have been discussed in chapter 2, and hence we do know how to calcul<strong>at</strong>e them.<br />

Therefore, we do have the means to evalu<strong>at</strong>e expressions like the last one.<br />

Fluctu<strong>at</strong>ions in fields.<br />

The local st<strong>at</strong>e of the system is specified by the extensive variables Xi(⃗r, t) =<br />

X 0 i (⃗r, t) + δXi(⃗r, t) in each cell positioned <strong>at</strong> ⃗r. Again, remember th<strong>at</strong> these<br />

cells are large enough to contain enough sufficient m<strong>at</strong>erial to be able to define<br />

local thermal equilibrium, but th<strong>at</strong> they are small on the scale of the vari<strong>at</strong>ions<br />

in the externally applied fields. Because we have


248 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

we have locally th<strong>at</strong><br />

dS = ∑<br />

ϕidXi<br />

i<br />

( )<br />

∂S<br />

ϕi =<br />

∂Xi<br />

(6.51)<br />

(6.52)<br />

where the partial deriv<strong>at</strong>ives only vary one of the extensive variables <strong>at</strong> a time.<br />

We assume local equilibrium, and hence<br />

ϕ 0 ( )<br />

∂S<br />

i = (X<br />

∂Xi<br />

0 1 , X 0 2 , · · ·) (6.53)<br />

<strong>at</strong> each point ⃗r and each time t.<br />

In thermodynamics, intensive quantities like temper<strong>at</strong>ure and pressure are<br />

only defined in the thermodynamic limit, as average quantities. Fluctu<strong>at</strong>ions in<br />

extensive quantities are easy to measure, <strong>at</strong> least in theory. We simply count the<br />

amount of m<strong>at</strong>erial, measure the volume, and measure the energy. Of course,<br />

we often only measure changes, but th<strong>at</strong> does not change the picture. Intensive<br />

quantities like temper<strong>at</strong>ure and pressure are measured using thermal contact<br />

with reservoirs, implying the need of thermal equilibrium. But even though we<br />

cannot technically measure these quantities <strong>at</strong> the level of fluctu<strong>at</strong>ions, we can<br />

easily define them by:<br />

( )<br />

∂S<br />

ϕi(⃗r, t) = (X1(⃗r, t), X2(⃗r, t), · · ·) (6.54)<br />

∂Xi<br />

or<br />

For small fluctu<strong>at</strong>ions we have<br />

ϕi(⃗r, t) = ϕ 0 i (⃗r, t) + ∑<br />

( 2 ∂ S<br />

k<br />

∂Xi∂Xk<br />

)<br />

(X 0 1 , X 0 2 , · · ·)δXk(⃗r, t) (6.55)<br />

δϕi(⃗r, t) = ϕi(⃗r, t) − ϕ 0 i (⃗r, t) = ∑<br />

SikδXk(⃗r, t) (6.56)<br />

where the m<strong>at</strong>rix S is symmetric.<br />

This same m<strong>at</strong>rix also is used in the probability function for small fluctu<strong>at</strong>ions<br />

in the Gaussian approxim<strong>at</strong>ion (see Chapter 2), where we have<br />

W (δX1, δX2, · · ·) = Ae 1<br />

2R<br />

∑<br />

k<br />

ij SijδXiδXj<br />

(6.57)<br />

We divided the change in entropy by the molar gas constant in order to have a<br />

dimensionless quantity in the exponent. Note th<strong>at</strong> in books where the amount<br />

of m<strong>at</strong>erial is measured in particles in stead of moles, one needs to divide by<br />

the Boltzmann constant! This also gives


6.5. MACROSCOPIC VERSUS MICROSCOPIC. 249<br />

( )<br />

∂W<br />

= R<br />

∂δXk<br />

−1 W δϕk<br />

(6.58)<br />

According to the maximum entropy principle, the probability function W<br />

takes its maximal value when all fluctu<strong>at</strong>ions are zero. This implies th<strong>at</strong> the exponent<br />

in the probability function is always neg<strong>at</strong>ive for non-zero fluctu<strong>at</strong>ions.<br />

Since the m<strong>at</strong>rix Sij is real and symmetric, it has a complete set of real eigenvalues.<br />

All these eigenvalues are neg<strong>at</strong>ive, or else there would be fluctu<strong>at</strong>ions<br />

for which the probability did not decrease. When all eigenvalues are neg<strong>at</strong>ive,<br />

the inverse of the m<strong>at</strong>rix Sij exists. We call this m<strong>at</strong>rix Tij and we therefore<br />

find th<strong>at</strong>:<br />

δXj = ∑<br />

Tjiδϕi<br />

and because S is symmetric, T is symmetric too.<br />

Finally, we consider the thermal average of the following quantity:<br />

i<br />

(6.59)<br />

∫<br />

⟨δXiδϕk⟩ = δXiδϕk(δX1, δX2, · · ·)W (δX1, δX2, · · ·)dδX1dδX2 · · · (6.60)<br />

Because of the rel<strong>at</strong>ion derived in 6.58 we have<br />

∫<br />

⟨δXiδϕk⟩ = R<br />

δXi<br />

( ∂W<br />

∂δXk<br />

and integr<strong>at</strong>ion by parts then gives<br />

)<br />

(δX1, δX2, · · ·)dδX1dδX2 · · · (6.61)<br />

⟨δXiδϕk⟩ = −Rδik<br />

This also gives us a represent<strong>at</strong>ion for the m<strong>at</strong>rix T, because<br />

⟨δXiδXj⟩ = ∑<br />

∑<br />

Tjk⟨δXiδϕk⟩ = −kB<br />

k<br />

k<br />

Tjkδik = −kBTji<br />

(6.62)<br />

(6.63)<br />

This also shows the symmetry of the m<strong>at</strong>rix T in an explicit manner. This rel<strong>at</strong>ion<br />

follows, of course, directly from the definition of the probability distribution<br />

in 6.57 given above. Similarly we find th<strong>at</strong><br />

⟨δϕiδϕj⟩ = ∑<br />

Sjk⟨δϕiδXk⟩ = −R ∑<br />

Sjkδik = −RSji<br />

Need for interactions.<br />

k<br />

k<br />

(6.64)<br />

In the previous discussion there is one important element missing. The forces<br />

responsible for currents are the gradients of intensive st<strong>at</strong>e variables, ⃗ ∇ϕi. The


250 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

probabilities for fluctu<strong>at</strong>ions in a cell only depend on the values of the extensive<br />

st<strong>at</strong>e variables Xi. In order to find the value of the gradient in such a field, we<br />

need to consider fluctu<strong>at</strong>ions in adjacent cells. They are uncoupled, however, in<br />

our current approxim<strong>at</strong>ion. Therefore,<br />

⟨(δXi(⃗r) − δXi(⃗r ′ )) 2 ⟩ = ⟨δX 2 i (⃗r)⟩ + ⟨δX 2 i (⃗r)⟩ (6.65)<br />

and the magnitude of this fluctu<strong>at</strong>ion is proportional to the volume of the cell.<br />

If we now consider two adjacent cubic cells of dimension L3 we see th<strong>at</strong> the<br />

typical value of the difference in the density is proportional to L 3<br />

3 and hence<br />

a typical value of the gradient is proportional to √ L. Hence typical values of<br />

gradients of densities are proportional to L−1 and they become very large for<br />

small values of L!<br />

Two things are missing in this approxim<strong>at</strong>ion. First, extensive quantities<br />

are conserved, and if values increase in one cell, the amount has to come from<br />

somewhere, and therefore values in neighboring cells tend to decrease. Second,<br />

there are interactions between the m<strong>at</strong>erial in one cell and in another. There<br />

is always gravity, and there are forces like the van der Waals interaction. In<br />

ionic fluids direct Coulomb interactions are possible. These effects are easily<br />

introduced in a microscopic theory, but then the calcul<strong>at</strong>ions are quite difficult.<br />

In thermodynamics we can do the same as we did in chapter four, however, and<br />

assume th<strong>at</strong> fluctu<strong>at</strong>ions are small and th<strong>at</strong> we can expand in a power series.<br />

The probability of fluctu<strong>at</strong>ions in a cell is now described by:<br />

with<br />

S(δX1, δX2, · · · , ∂aδX1, · · ·) = 1<br />

2<br />

W (δX1, δX2, · · · , ∂aδX1, · · ·) = Ae 1<br />

k B S<br />

∑<br />

ij<br />

SijδXiδXj + 1<br />

2<br />

∑<br />

iajb<br />

Fiajb∂aδXi∂bδXj<br />

(6.66)<br />

(6.67)<br />

The summ<strong>at</strong>ions over a and b are summ<strong>at</strong>ions over Carthesian coordin<strong>at</strong>es.<br />

Because the tensor Fiajb is rel<strong>at</strong>ed to a second order functional deriv<strong>at</strong>ive, it<br />

is symmetric. Therefore, it has a complete spectrum of eigenvalues, and again<br />

all these eigenvalues have to be neg<strong>at</strong>ive in order to guarantee th<strong>at</strong> the st<strong>at</strong>e<br />

without fluctu<strong>at</strong>ions has the highest probability. Because none of the eigenvalues<br />

are zero, the tensor has an inverse.<br />

The definition above is the only one possible in second order, and any microscopic<br />

theory has to reduce to this form in lowest order. The values of Fiajb<br />

can only be calcul<strong>at</strong>ed in a microscopic theory. This is a fundamental difference<br />

with the values of Sij which follow from standard thermodynamics. Note th<strong>at</strong><br />

there are no terms in second order with only one gradient. They disappear<br />

because of symmetry. Also note th<strong>at</strong> this form of theory is in spirit completely<br />

equivalent to Ginzburg-Landau!<br />

Properties for gradients.


6.5. MACROSCOPIC VERSUS MICROSCOPIC. 251<br />

It is easy to derive th<strong>at</strong> we have the following rel<strong>at</strong>ion for the gradients:<br />

⟨∂aδXi<br />

∂∂bδXj<br />

∂t<br />

⟩ = ⟨ ∂∂aδXi<br />

∂bδXj⟩ (6.68)<br />

∂t<br />

using the same reasoning as before. We also define the equivalent of the conjug<strong>at</strong>e<br />

intensive fields in a similar manner as we derived for the fluctu<strong>at</strong>ions in<br />

the fields ϕi by<br />

and we have exactly as before<br />

Onsager’s theorem derived.<br />

Fia = ∑<br />

jb<br />

Fiajb∂bδXj<br />

⟨∂aδXiFjb⟩ = −Rδijδab<br />

(6.69)<br />

(6.70)<br />

We are almost ready to derive the Onsager rel<strong>at</strong>ions. First of all, the r<strong>at</strong>e<br />

of change in the gradient of an extensive quantity is proportional to the current<br />

density pertaining to th<strong>at</strong> quantity:<br />

∂∂aδXi<br />

∂t ∝ δJia (6.71)<br />

which is easy to see, since doubling the current density will double the flow of<br />

m<strong>at</strong>erial and hence double the change in gradient. Next, from our definition of<br />

the fields F we see th<strong>at</strong> they behave like gradients of the intensive st<strong>at</strong>e variables.<br />

Since this is the only way we can define the conjug<strong>at</strong>e fields for fluctu<strong>at</strong>ions,<br />

these are the fields th<strong>at</strong> drive the currents. In the linear approxim<strong>at</strong>ion we have<br />

therefore:<br />

δJia = ∑<br />

jb<br />

LiajbFjb<br />

in the same way as in the previous section. Taken together, we now have:<br />

and hence<br />

⟨(∂aδXi)( d∂bδXj<br />

dt<br />

Similarly we find<br />

∂∂aδXi<br />

∂t<br />

)⟩ = C ∑<br />

kc<br />

∑<br />

= C<br />

jb<br />

LiajbFjb<br />

Ljbkc⟨∂aδXiFkc⟩ = −RLjbia<br />

⟨ ∂∂aδXi<br />

∂bδXj⟩ = −RLiajb<br />

∂t<br />

(6.72)<br />

(6.73)<br />

(6.74)<br />

(6.75)


252 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

and hence<br />

or<br />

Liajb = Ljbia<br />

Lik = L T ki<br />

(6.76)<br />

(6.77)<br />

Therefore, the Onsager rel<strong>at</strong>ions are a general consequence of the microscopic<br />

n<strong>at</strong>ure of m<strong>at</strong>ter, but we do not need to know any details. It is sufficient to<br />

know th<strong>at</strong> the macroscopic laws are the limit of the microscopic description,<br />

and th<strong>at</strong> the microscopic laws have time reversal symmetry. The form 6.66 for<br />

the fluctu<strong>at</strong>ions is simply the most general from we can write down in lowest<br />

order, and this form is sufficient for the purpose of proving Onsager’s rel<strong>at</strong>ions.<br />

The inclusion of magnetic fields is easy, we simply have to change the sign of<br />

all magnetic properties under time reversal. This gives some extra minus signs.<br />

Entropy production.<br />

In the previous section we derived the entropy production when currents are<br />

present in 6.40. When the currents are due to fluctu<strong>at</strong>ions only, this takes the<br />

form<br />

or<br />

Ps = ∑<br />

ia<br />

Ps = ∑<br />

iajb<br />

Fia · δJia<br />

LiajbFiaFjb<br />

(6.78)<br />

(6.79)<br />

and this has to be positive, since entropy is always produced when there are<br />

fluctu<strong>at</strong>ions. Therefore, the eigenvalues of the tensor L have to be positive! Even<br />

though this st<strong>at</strong>ement follows from an analysis of fluctu<strong>at</strong>ions, if can now be<br />

generalized to all currents, since they have the same microscopic underpinning.<br />

For example, consider the case where we only have a gradient in the chemical<br />

potential and only a flow of m<strong>at</strong>erial. In th<strong>at</strong> case we have ⃗ JN = L⃗ ∇ −µ<br />

T =<br />

− L<br />

T ⃗ ∇µ and because of the analysis above, L is positive. Therefore, the mass<br />

current is in the opposite direction of the gradient of the chemical potential, as<br />

is observed.<br />

6.6 Thermo-electric effects.<br />

Transport equ<strong>at</strong>ions for NQ system.


6.6. THERMO-ELECTRIC EFFECTS. 253<br />

We now apply the transport theory developed in the second section to a<br />

case where we have gradients in both the temper<strong>at</strong>ure and chemical potential.<br />

We have transport of m<strong>at</strong>erial and energy, or in different words, transport of<br />

m<strong>at</strong>erial and he<strong>at</strong>. Therefore we call this a NQ system. We repe<strong>at</strong> the equ<strong>at</strong>ions<br />

for the currents and entropy production:<br />

Ps =<br />

( )<br />

1<br />

⃗JS = ⃗JU −<br />

T<br />

( ( )<br />

⃗∇<br />

1<br />

·<br />

T<br />

⃗ JU − ⃗ ∇<br />

The transport equ<strong>at</strong>ions are in this case:<br />

and<br />

(<br />

µ<br />

)<br />

⃗JN (6.80)<br />

T<br />

(<br />

µ<br />

)<br />

·<br />

T<br />

⃗ JN<br />

⃗JU = LUU · ⃗ ∇ 1<br />

T + LUN · ⃗ ∇ −µ<br />

T<br />

)<br />

(6.81)<br />

(6.82)<br />

⃗JN = LNU · ⃗ ∇ 1<br />

T + LNN · ⃗ ∇ −µ<br />

(6.83)<br />

T<br />

In many cases it is useful to replace the energy flow by the he<strong>at</strong> flow. because<br />

we have ¯ dQ = T dS we define<br />

⃗JQ = T ⃗ JS = ⃗ JU − µ ⃗ JN<br />

(6.84)<br />

which more or less st<strong>at</strong>es th<strong>at</strong> the he<strong>at</strong> flow is the total flow of energy minus<br />

the flow of potential energy, represented by the last term in the equ<strong>at</strong>ion above.<br />

Using the he<strong>at</strong> flow, we find:<br />

( ( )<br />

Ps = ⃗∇<br />

1<br />

·<br />

T<br />

⃗ JQ + 1<br />

T ⃗ ∇(−µ) · ⃗ )<br />

JN<br />

(6.85)<br />

and<br />

or<br />

⃗JQ = (LUU − µLNU ) · ⃗ ∇ 1<br />

T + (LUN − µLNN ) · ⃗ ∇ −µ<br />

T<br />

⃗JQ = (LUU − µLNU − µLUN + µ 2 LNN ) · ⃗ ∇ 1<br />

T<br />

Similarly we have<br />

⃗JN = (LNU − µLNN) · ⃗ ∇ 1<br />

T<br />

which can be written in the form<br />

1<br />

⃗JN = L11<br />

T · ⃗ ∇(−µ) + L12 · ⃗ ∇ 1<br />

T<br />

(6.86)<br />

− 1<br />

T (LUN − µLNN ) · ⃗ ∇µ (6.87)<br />

− 1<br />

T LNN · ⃗ ∇µ (6.88)<br />

(6.89)


254 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

with<br />

1<br />

⃗JQ = L21<br />

T · ⃗ ∇(−µ) + L22 · ⃗ ∇ 1<br />

T<br />

L11 = LNN = L T 11<br />

L12 = LNU − µLNN<br />

L21 = LUN − µLNN = L T 12<br />

L22 = LUU − µLNU − µLUN + µ 2 LNN = L T 22<br />

(6.90)<br />

(6.91)<br />

(6.92)<br />

(6.93)<br />

(6.94)<br />

These last equ<strong>at</strong>ions can be written in the form of a similarity transform<strong>at</strong>ion:<br />

( L11 L12<br />

L21 L22<br />

)<br />

=<br />

( 1 0<br />

−µ 1<br />

) ( LNN LNU<br />

LUN LUU<br />

) ( 1 −µ<br />

0 1<br />

which implies th<strong>at</strong> the transformed tensor Lij is also positive definite.<br />

Units.<br />

)<br />

(6.95)<br />

It is useful to take time to consider the dimensions of all quantities involved,<br />

and which units we are employing. In thermodynamics all m<strong>at</strong>erial is counted<br />

in moles, not per particle. The transform<strong>at</strong>ion is easy, though, and involves<br />

Avogadro’s number NA. The current density in equ<strong>at</strong>ion 6.82 is measured in<br />

Jm −2 s −1 , while the current density in 6.83 is measured in Mm −2 s −1 . The<br />

temper<strong>at</strong>ure is in K and the chemical potential is in JM −1 , where we use the<br />

symbol M for moles. From 6.82 we then find the following for the transport<br />

coefficients:<br />

and similarly:<br />

Coefficient Dimensions<br />

LUU<br />

LUN<br />

LNU<br />

LNN<br />

JKm−1 s−1 MKm −1s−1 MKm −1s−1 M 2J −1Km−1s−1 Coefficient Dimensions<br />

L22<br />

L21<br />

L12<br />

L11<br />

JKm−1 s−1 MKm −1s−1 MKm −1s−1 M 2J −1Km−1s−1


6.6. THERMO-ELECTRIC EFFECTS. 255<br />

Electrical and thermal conductivity.<br />

For the rest of this chapter we confine ourselves to a specific case. Suppose<br />

we are describing the behavior of electrons in a metal. In th<strong>at</strong> case the flow of<br />

m<strong>at</strong>erial is equivalent to a flow of electrical current, and the he<strong>at</strong> flow is just the<br />

he<strong>at</strong> flow due to the electrons. Note th<strong>at</strong> in any experiment we also have a he<strong>at</strong><br />

flow due to the ions. This part of the he<strong>at</strong> flow has to be analyzed separ<strong>at</strong>ely,<br />

and should not be included in the Onsager reciprocity rel<strong>at</strong>ions! Basic quantities<br />

one defines are the electrical conductivity σ, which rel<strong>at</strong>es the electrical current<br />

density ⃗j to the electric field ⃗ E, and the thermal conductivity κ which rel<strong>at</strong>es<br />

the he<strong>at</strong> current density ⃗ JQ to the temper<strong>at</strong>ure gradient. The only contribution<br />

of interest to the chemical potential is the Coulomb potential, and we have:<br />

⃗j = −NAe ⃗ JN<br />

with −e the charge of an electron. This gives<br />

(6.96)<br />

⃗E = 1<br />

NAe ⃗ ∇µ (6.97)<br />

⃗j = N 2 A e2<br />

T L11 · ⃗ E + NAe<br />

T 2 L12 · ⃗ ∇T (6.98)<br />

⃗JQ = − NAe<br />

T L21 · ⃗ E − 1<br />

T 2 L22 · ⃗ ∇T (6.99)<br />

where the parameters Lij only account for the effects of the electrons.<br />

The entropy production expressed in these quantities is:<br />

Ps = − 1<br />

T 2 ⃗ ∇T · ⃗ JQ + 1<br />

T ⃗ E · ⃗j (6.100)<br />

Since the he<strong>at</strong> production is T Ps we see immedi<strong>at</strong>ely th<strong>at</strong> the second term is<br />

the standard he<strong>at</strong> production as described in electricity theory. The first term<br />

is the he<strong>at</strong> produced when a he<strong>at</strong> flow goes to a lower temper<strong>at</strong>ure.<br />

The electrical conductivity tensor σ is defined <strong>at</strong> constant temper<strong>at</strong>ure by<br />

⃗j = σ · ⃗ E, ⃗ ∇T = 0, and this gives<br />

σ = N 2 A e2<br />

T L11 (6.101)<br />

and has dimensions C 2 N −1 m −2 s −1 as needed. Remember th<strong>at</strong> the electric<br />

field has dimensions NC −1 . Also, because of the Onsager rel<strong>at</strong>ion we have th<strong>at</strong><br />

σ = σ T .<br />

The thermal conductivity tensor κ for the electrons is defined by the he<strong>at</strong><br />

current due to the electrons as a function of the thermal gradient, when there is<br />

no electrical current, or ⃗ JQ = −κ· ⃗ ∇T , ⃗j = 0. Because there is no electric current<br />

there is an electric field counteracting the force due to the thermal gradient:


256 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

or<br />

and hence<br />

which gives<br />

0 = N 2 A e2<br />

T L11 · ⃗ E + NAe<br />

T 2 L12 · ⃗ ∇T (6.102)<br />

⃗E = − 1<br />

NAeT L−1<br />

11 L12 · ⃗ ∇T (6.103)<br />

⃗JQ = 1<br />

T 2 (L21 · L −1<br />

11 L12 · ⃗ ∇T − L22 · ⃗ ∇T ) (6.104)<br />

and again because of Onsager we have κ = κ T .<br />

Cubic m<strong>at</strong>erials are easy.<br />

κ = 1<br />

T 2 (L22 − L21L −1<br />

11 L12) (6.105)<br />

In cubic m<strong>at</strong>erials all tensors in Carthesian space are simple numbers times<br />

the unit m<strong>at</strong>rix. Therefore, we have only three numbers th<strong>at</strong> characterize the<br />

transport, L11, L12 = L21, and L22. In th<strong>at</strong> case the electrical and thermal<br />

conductivity are given by:<br />

and we have<br />

κ = 1<br />

σ = N 2 A e2<br />

T L11 (6.106)<br />

T 2 (L22 − L 2 21L −1<br />

11<br />

) (6.107)<br />

σκ = N 2 A e2<br />

T 3 (L11L22 − L 2 21) (6.108)<br />

and since the m<strong>at</strong>rix of transport coefficients is positive definite, we see th<strong>at</strong> in<br />

this case:<br />

σκ > 0 (6.109)<br />

which, of course, can be easily generalized, and holds in all cases, when we<br />

interpret the gre<strong>at</strong>er than to mean being positive definite. This rel<strong>at</strong>ion is no<br />

surprise, since electrical currents always follow the electric field, and thermal<br />

currents always follow the inverse temper<strong>at</strong>ure gradient. But here we have<br />

rel<strong>at</strong>ed these common observ<strong>at</strong>ions to the production of entropy, which gives the<br />

fact th<strong>at</strong> the m<strong>at</strong>rix of transport coefficients is positive definite. The electrical<br />

conductivity is positive because of Newton’s law. The thermal conductivity is<br />

positive because of the second law of thermodynamics.


6.6. THERMO-ELECTRIC EFFECTS. 257<br />

Thermopower.<br />

Equ<strong>at</strong>ion 6.102 has interesting consequences. If there is a temper<strong>at</strong>ure gradient<br />

across a conductor, and if the end points are not connected and no current<br />

can flow, there will be a particle flow inside the m<strong>at</strong>erial causing a buildup of<br />

charge <strong>at</strong> the endpoints. The particle flow will stop when the electric force balances<br />

the thermal force, and the resultant electric field follows from 6.102, as<br />

given before:<br />

⃗E = − 1<br />

NAeT L−1<br />

11 L12 · ⃗ ∇T (6.110)<br />

The constant of proportionality is the absolute thermoelectric power tensor ϵ,<br />

defined by ⃗ E = ϵ ⃗ ∇T , and it follows th<strong>at</strong><br />

ϵ = − 1<br />

NAeT L−1<br />

11 L12<br />

(6.111)<br />

Note th<strong>at</strong> the n<strong>at</strong>ure of the charge carriers is reflected in the sign of ϵ. The<br />

carrier concentr<strong>at</strong>ion will be less <strong>at</strong> higher temper<strong>at</strong>ures. Therefore, if the charge<br />

carriers are positive, the low temper<strong>at</strong>ure part will have a larger positive charge,<br />

and the resulting field will be towards the high temper<strong>at</strong>ure. For neg<strong>at</strong>ive<br />

charged carriers this is the opposite.<br />

Suppose we have a m<strong>at</strong>erial with end points <strong>at</strong> temper<strong>at</strong>ures T1 and T2. No<br />

current flows, and therefore the difference in potential between the end points<br />

is<br />

∫ 2<br />

V2 − V1 = − ⃗E · d⃗r = −<br />

1<br />

∫ 2<br />

1<br />

d⃗r · ϵ · ⃗ ∇T (6.112)<br />

If we now take two different m<strong>at</strong>erials and connect the end points <strong>at</strong> one end,<br />

there will be a potential difference between the unconnected end points <strong>at</strong> the<br />

other end. The larger the difference in thermoelectric power between the two<br />

m<strong>at</strong>erials, the larger the voltage difference.<br />

A very simple expression can be obtained if we assume th<strong>at</strong> the temper<strong>at</strong>ures<br />

are very close together, and the m<strong>at</strong>erial is cubic. Then we have<br />

V2 − V1 ≈ (T1 − T2)ϵ (6.113)<br />

where the tensor is now reduced to a number times the unit m<strong>at</strong>rix. If we<br />

measure the voltage difference <strong>at</strong> the disconnected end points <strong>at</strong> T1 we find:<br />

V x<br />

1 − V y<br />

1 ≈ (T1 − T2)(ϵ y − ϵ x ) (6.114)<br />

which shows how a thermocouple can be used to measure a temper<strong>at</strong>ure difference.<br />

If we compare with formula 6.3 we see th<strong>at</strong>:<br />

αAB = ϵ A − ϵ B<br />

(6.115)


258 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

Reformul<strong>at</strong>ion of the transport equ<strong>at</strong>ions.<br />

The equ<strong>at</strong>ions for the electric and he<strong>at</strong> current densities can now be modified<br />

by elimin<strong>at</strong>ing the transport coefficients in terms of the measurable quantities<br />

we used in the previous subsections. We obtain:<br />

L11 = T<br />

N 2 σ (6.116)<br />

Ae2 T 2<br />

L12 = −NAeT L11 · ϵ = − σϵ (6.117)<br />

NAe<br />

L21 = L T 12 = −<br />

which yields the following equ<strong>at</strong>ions:<br />

Other thermoelectric effects.<br />

2 T<br />

NAe ϵT σ T<br />

(6.118)<br />

L22 = T 2 κ + T 3 ϵ T σ T ϵ (6.119)<br />

⃗j = σ · ⃗ E − σϵ ⃗ ∇T (6.120)<br />

⃗JQ = T ϵ T σ ⃗ E − (κ + T ϵ T σϵ) ⃗ ∇T (6.121)<br />

When a current flows through a m<strong>at</strong>erial <strong>at</strong> constant temper<strong>at</strong>ure, it causes<br />

an electric field across the m<strong>at</strong>erial, which in turn causes a he<strong>at</strong> flow. In this<br />

scenario we have<br />

⃗JQ = T ϵ T⃗j (6.122)<br />

for ⃗ ∇T = 0. If we force a current through two different m<strong>at</strong>erials connected<br />

in series, the he<strong>at</strong> flow will be different in the two m<strong>at</strong>erials. Therefore, <strong>at</strong> the<br />

junction he<strong>at</strong> is produced or absorbed. The amount is<br />

∆Q = T ∆ϵ T⃗j (6.123)<br />

This is the Peltier effect. By comparing with equ<strong>at</strong>ion 6.4 we see th<strong>at</strong><br />

ΠAB = T ∆ϵ (6.124)<br />

for cubic m<strong>at</strong>erials, which is the rel<strong>at</strong>ion found by Kelvin in an empirical manner.<br />

He<strong>at</strong> gener<strong>at</strong>ion in resistors.<br />

We all know the formula for the gener<strong>at</strong>ion of he<strong>at</strong> in a resistor, where power<br />

production is proportional to ⃗j · ⃗ E. This formula, however, assumes constant


6.6. THERMO-ELECTRIC EFFECTS. 259<br />

temper<strong>at</strong>ure. Wh<strong>at</strong> if there is also a temper<strong>at</strong>ure gradient in the resistor?<br />

We can find the answer quite easily. Assume th<strong>at</strong> we drive a current density<br />

⃗j through a resistor, and th<strong>at</strong> there is a temper<strong>at</strong>ure gradient ⃗ ∇T . These two<br />

quantities determine completely the st<strong>at</strong>e of the system, because of the transport<br />

equ<strong>at</strong>ions. The electric field in the resistor follows from 6.120:<br />

⃗E = ρ · ⃗j + ϵ ⃗ ∇T (6.125)<br />

where ρ is the resistivity tensor, the inverse of the conductivity tensor. The<br />

electric field is caused by both a charge transport due to the applied current<br />

and a charge transport due to the thermal gradient. The he<strong>at</strong> current follows<br />

from 6.121 and is<br />

⃗JQ = T ϵ T (⃗j + σϵ ⃗ ∇T ) − (κ + T ϵ T σϵ) ⃗ ∇T = T ϵ T⃗j − κ ⃗ ∇T (6.126)<br />

The entropy production follows from 6.100:<br />

Ps = − 1<br />

T 2 ⃗ ∇T · (T ϵ T⃗j − κ ⃗ ∇T ) + 1<br />

T (ρ · ⃗j + ϵ ⃗ ∇T ) · ⃗j (6.127)<br />

which leads to production of thermal energy <strong>at</strong> a r<strong>at</strong>e<br />

P = T Ps = ⃗j · ρ · ⃗j + 1<br />

T ⃗ ∇T · κ · ⃗ ∇T (6.128)<br />

as expected. The first term is the Joule he<strong>at</strong>, the second term the he<strong>at</strong> gener<strong>at</strong>ed<br />

while keeping a temper<strong>at</strong>ure gradient.<br />

Now we will perform a different experiment. We set up a resistor and fix<br />

the temper<strong>at</strong>ures <strong>at</strong> the end points, under conditions where there is no electric<br />

current. This sets up a he<strong>at</strong> flow in the resistor, and sets up a temper<strong>at</strong>ure profile<br />

in the resistor. Since he<strong>at</strong> is flowing, and he<strong>at</strong> is gener<strong>at</strong>ed, the temper<strong>at</strong>ure<br />

profile would change unless we keep each point in contact with a he<strong>at</strong> reservoir.<br />

This is exactly wh<strong>at</strong> we do in the following experiment. We take a conductor,<br />

and <strong>at</strong> each point of the conductor we make thermal contact with a reservoir,<br />

in such a manner th<strong>at</strong> the temper<strong>at</strong>ure of the conductor <strong>at</strong> each point remains<br />

<strong>at</strong> a constant value. Because of the gener<strong>at</strong>ion of he<strong>at</strong> and the he<strong>at</strong> currents,<br />

this means th<strong>at</strong> <strong>at</strong> each point of the conductor there is a he<strong>at</strong> flow into or out<br />

of the reservoir, needed to keep the temper<strong>at</strong>ure <strong>at</strong> th<strong>at</strong> point constant.<br />

We now ask the following question. In the previous set-up, if we now add an<br />

electric current in the conductor, wh<strong>at</strong> is the additional he<strong>at</strong> flow into or out<br />

of the reservoir needed to keep the temper<strong>at</strong>ure the same? In order to calcul<strong>at</strong>e<br />

this, we need to calcul<strong>at</strong>e the change in internal energy <strong>at</strong> each point.<br />

The energy transport is given by ⃗ JU = ⃗ JQ + µ ⃗ JN and the local change in<br />

energy follows from<br />

∂u<br />

∂t = −⃗ ∇ · ⃗ JU<br />

(6.129)<br />

In a steady st<strong>at</strong>e there is a constant profile in the particle density (no accumul<strong>at</strong>ion<br />

of charge!), and we have


260 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

and hence<br />

0 = ∂n<br />

∂t = −⃗ ∇ · ⃗ JN<br />

(6.130)<br />

∂u<br />

∂t = −⃗ ∇ · ⃗ JU = − ⃗ ∇ · ⃗ JQ − ( ⃗ ∇µ) ⃗ JN = − ⃗ ∇ · ⃗ JQ + ⃗ E · ⃗j (6.131)<br />

Using equ<strong>at</strong>ions 6.125 and 6.126 we get<br />

and since ⃗ ∇ · ⃗j = 0 this leads to<br />

∂u<br />

∂t = −⃗ ∇ · (T ϵ T⃗j − κ ⃗ ∇T ) + (ρ · ⃗j + ϵ ⃗ ∇T ) · ⃗j (6.132)<br />

∂u<br />

∂t = ⃗j · ρ · ⃗j + ⃗ ∇ · (κ ⃗ ∇T ) − T ( ⃗ ∇ϵ T ) · ⃗j (6.133)<br />

and since the thermopower only changes because the temper<strong>at</strong>ure changes we<br />

write this in the form<br />

∂u<br />

∂t = ⃗j · ρ · ⃗j + ⃗ ∇ · (κ ⃗ ∇T ) − T ( ⃗ ∇T ) · dϵT<br />

dT · ⃗j (6.134)<br />

Suppose there is no current. Than the change in internal energy is given by<br />

∂u0<br />

∂t = ⃗ ∇ · (κ ⃗ ∇T ) (6.135)<br />

and this determines the he<strong>at</strong> flow with the reservoir <strong>at</strong> each point. When there<br />

is a current, the additional he<strong>at</strong> flow is therefore:<br />

∂uj<br />

∂t = ⃗j · ρ · ⃗j − T ( ⃗ ∇T ) · dϵT<br />

dT · ⃗j (6.136)<br />

The first term is easy, and is the Joule he<strong>at</strong> gener<strong>at</strong>ed in each point in the<br />

conductor. In order for the conductor to remain <strong>at</strong> a constant temper<strong>at</strong>ure<br />

profile this Joule he<strong>at</strong> needs to go into the reservoir. The second term is a new<br />

he<strong>at</strong> source, and this is called the Thomson he<strong>at</strong>. It is linear in both the thermal<br />

gradient and the electric current and for small currents domin<strong>at</strong>es the Joule he<strong>at</strong>!<br />

The strength of this he<strong>at</strong> source is directly rel<strong>at</strong>ed to the temper<strong>at</strong>ure deriv<strong>at</strong>ive<br />

of the thermo-power.<br />

Comparing with equ<strong>at</strong>ion 6.6 we see th<strong>at</strong><br />

τ = T dϵ<br />

dT<br />

(6.137)<br />

for cubic m<strong>at</strong>erials, again in agreement with wh<strong>at</strong> Kelvin found empirically.<br />

Wh<strong>at</strong> happened to the minus sign when we went to the equ<strong>at</strong>ion above?


6.6. THERMO-ELECTRIC EFFECTS. 261<br />

Adiab<strong>at</strong>ic conductivity.<br />

A recurring theme in thermodynamics is th<strong>at</strong> the values of measured properties<br />

depend on the conditions under which experiments are performed. If we<br />

measure the conductivity of a m<strong>at</strong>erial <strong>at</strong> constant temper<strong>at</strong>ure we find the<br />

property σ. It is also possible to perform the experiment under adiab<strong>at</strong>ic conditions,<br />

in this case demanding th<strong>at</strong> ⃗ JQ = 0. Equ<strong>at</strong>ion 6.121 then gives<br />

and equ<strong>at</strong>ion 6.120 then leads to<br />

T ϵ T σ ⃗ E = (κ + T ϵ T σϵ) ⃗ ∇T (6.138)<br />

⃗j = σ · ⃗ E − σϵ(κ + T ϵ T σϵ) −1 T ϵ T σ ⃗ E (6.139)<br />

and hence the adiab<strong>at</strong>ic conductivity is given by<br />

σad = σ(1 − ϵ(κ + T ϵ T σϵ) −1 T ϵ T σ) (6.140)<br />

As always, we need to define which are the st<strong>at</strong>e variables we have under control.<br />

Note th<strong>at</strong> for very large temper<strong>at</strong>ures the adiab<strong>at</strong>ic conductivity goes to zero!<br />

Wh<strong>at</strong> does it mean th<strong>at</strong> the adiab<strong>at</strong>ic conductivity is zero?<br />

External versus total chemical potential.<br />

We rel<strong>at</strong>ed the gradient in the chemical potential to the electric field by<br />

⃗E = 1<br />

NAe ⃗ ∇µ (6.141)<br />

but this is obviously only the gradient of the external chemical potential. We<br />

also know from thermodynamics th<strong>at</strong><br />

SdT − V dp + Ndµ = 0 (6.142)<br />

Now assume th<strong>at</strong> the pressure is always constant, in which case we set dp = 0.<br />

Therefore we have for the gradients:<br />

or<br />

S ⃗ ∇T + N ⃗ ∇µtot = 0 (6.143)<br />

S ⃗ ∇T + NNAe ⃗ E + N ⃗ ∇µint = 0 (6.144)<br />

Therefore, if we apply an electric field and a temper<strong>at</strong>ure gradient, we autom<strong>at</strong>ically<br />

gener<strong>at</strong>e a gradient in the internal chemical potential, and hence a<br />

gradient in the amount of m<strong>at</strong>erial! This is a warning sign when we want to


262 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.<br />

compare thermodynamics and microscopic calcul<strong>at</strong>ions. A similar situ<strong>at</strong>ion occurs<br />

in dielectrics, where on a microscopic level we calcul<strong>at</strong>e the response of the<br />

system to a change in the total field, including the induced dipole field, driven<br />

by an applied field. In our case, a microscopic calcul<strong>at</strong>ion would give results of<br />

the form<br />

⃗jN = Rint · ⃗ ∇µint + Rext · ⃗ ∇µext + RT ⃗ ∇T (6.145)<br />

We then use the rel<strong>at</strong>ion 6.144 to distribute the effects of the internal chemical<br />

potential over the gradient contributions for temper<strong>at</strong>ure and applied field in<br />

order to get the response functions needed in thermodynamics.<br />

6.7 Extra questions.<br />

1. How can we take l<strong>at</strong>tice changes as a result of pressure gradients into<br />

account?<br />

2. Is it possible to construct a theory in which pressure also drives electron<br />

transport?<br />

3. How can we expand transport theory to include spin transport?<br />

4. How do we include he<strong>at</strong> transport due to the l<strong>at</strong>tice?<br />

6.8 Problems for chapter 6<br />

Problem 1.<br />

Consider the he<strong>at</strong> current density jQ and electric current density j through<br />

a straight metal wire of length d. The end points of the wire are connected to<br />

electrically insul<strong>at</strong>ing contacts, which can be held <strong>at</strong> fixed temper<strong>at</strong>ures. We<br />

assume th<strong>at</strong> the potential and temper<strong>at</strong>ure gradients are constant in the wire.<br />

The electric current density j and he<strong>at</strong> current density jQ are given by j =<br />

σE − σϵ dT<br />

dx and jQ = T ϵσE − (κ + T σϵ2 ) dT<br />

dx , where all coefficients are numbers,<br />

not tensors, and functions of temper<strong>at</strong>ure. (a) Wh<strong>at</strong> is the difference in electric<br />

potential between the end points as a function of the temper<strong>at</strong>ure difference?<br />

(b) The he<strong>at</strong> current density can be written in the form jQ = K × (T2 − T1).<br />

Find K in terms of κ, σ, and ϵ.<br />

Problem 2.<br />

For a copper to iron thermocouple we have the following d<strong>at</strong>a. When one<br />

junction is held <strong>at</strong> the temper<strong>at</strong>ure of melting ice, the emf <strong>at</strong> the other junction<br />

as a function of temper<strong>at</strong>ure is


6.8. PROBLEMS FOR CHAPTER 6 263<br />

with<br />

E = α1t + 1<br />

2 α2t 2 + 1<br />

3<br />

α3t 3<br />

α1 = −13.403 × 10 −6 , α2 = +0.0275 × 10 −6 , α3 = +0.00026 × 10 −6<br />

and t in degrees centigrade.<br />

Find the Peltier he<strong>at</strong> transferred <strong>at</strong> a junction <strong>at</strong> the temper<strong>at</strong>ure of (a)<br />

20 ◦ C, (b) 200 ◦ C, by a current of 3 mA in 5 minutes.<br />

If the current is 1 µA, <strong>at</strong> wh<strong>at</strong> temper<strong>at</strong>ure is the Peltier he<strong>at</strong> equal to zero?<br />

Problem 3.<br />

Consider the he<strong>at</strong> current density jq and electric current density je through<br />

a straight metal wire of length L. We assume th<strong>at</strong> the potential and temper<strong>at</strong>ure<br />

gradients are constant in the wire. The electric current density je and he<strong>at</strong><br />

current density jq are given by je = σE − σϵ dT<br />

dx and jq = T ϵσE − (κ + T σϵ2 ) dT<br />

dx ,<br />

where all coefficients are scalars, not tensors, and are functions of tempera-<br />

ture. The he<strong>at</strong> production is given by Pq = σ−1j 2 + κ dT<br />

T ( dx )2 . Find the electric<br />

conductivity as measured when there is no he<strong>at</strong> flow in the m<strong>at</strong>erial (adiab<strong>at</strong>ic<br />

conductivity), in terms of the quantities given in the formulas, and find the limit<br />

of the value of the adiab<strong>at</strong>ic conductivity for large temper<strong>at</strong>ures. Calcul<strong>at</strong>e the<br />

r<strong>at</strong>io of the he<strong>at</strong> production for a given electric current density in the adiab<strong>at</strong>ic<br />

case and in the normal constant temper<strong>at</strong>ure case.<br />

Problem 4.<br />

Using equ<strong>at</strong>ion (6.36) estim<strong>at</strong>e the he<strong>at</strong> production in the Niagara falls,<br />

assuming the temper<strong>at</strong>ure is constant when the w<strong>at</strong>er is falling.


264 CHAPTER 6. TRANSPORT IN THERMODYNAMICS.


Chapter 7<br />

Correl<strong>at</strong>ion Functions.<br />

Correl<strong>at</strong>ion functions are normally introduced in st<strong>at</strong>istical mechanics. It is<br />

easy to define the instantaneous positions of single <strong>at</strong>oms. It is also easy to<br />

define a function th<strong>at</strong> measures the position of two <strong>at</strong>oms simultaneously. After<br />

performing the appropri<strong>at</strong>e averages one can then construct probability distributions<br />

for finding distances between <strong>at</strong>oms.<br />

In <strong>at</strong>omic calcul<strong>at</strong>ions appropri<strong>at</strong>e averages typically means taking both sp<strong>at</strong>ial<br />

and temporal averages. We assume th<strong>at</strong> such averages yield the thermodynamical<br />

averages we need, and th<strong>at</strong> temper<strong>at</strong>ure follows. This is often the<br />

case, which is why introducing correl<strong>at</strong>ion functions on the <strong>at</strong>omic level is very<br />

useful.<br />

The calcul<strong>at</strong>ions mentioned above are, by construction, short range in n<strong>at</strong>ure.<br />

Long range correl<strong>at</strong>ions are small, and we know th<strong>at</strong> far away events<br />

should be uncorrel<strong>at</strong>ed. The zeroth order approxim<strong>at</strong>ion for large distances is<br />

therefore easy. The question arises wh<strong>at</strong> we can say about the first order approxim<strong>at</strong>ion<br />

how the correl<strong>at</strong>ion functions approach this limit. If we assume th<strong>at</strong><br />

long range fluctu<strong>at</strong>ions exist, but th<strong>at</strong> we always have local equilibrium, we can<br />

make st<strong>at</strong>ements about the approach to the uncorrel<strong>at</strong>ed st<strong>at</strong>e. This sets limits<br />

on correl<strong>at</strong>ions functions, and can be used to connect correl<strong>at</strong>ion functions with<br />

thermodynamical properties. This is the focus of the current chapter.<br />

Long range behavior in real space corresponds to low frequency behavior<br />

in Fourier space. Therefore, we are asking questions about low frequency (and<br />

hence low energy) excit<strong>at</strong>ions of the system. If our system is in thermal equilibrium,<br />

there are limit<strong>at</strong>ions on these low energy excit<strong>at</strong>ions.<br />

7.1 Description of correl<strong>at</strong>ed fluctu<strong>at</strong>ions.<br />

Phenomenological theories.<br />

In the previous chapters we have already mentioned the importance of in-<br />

265


266 CHAPTER 7. CORRELATION FUNCTIONS.<br />

teractions between parts of a m<strong>at</strong>erial a number of times. Without interactions,<br />

all m<strong>at</strong>erials would obey the ideal gas equ<strong>at</strong>ion of st<strong>at</strong>e, and we know th<strong>at</strong> to<br />

be not true. When we discussed the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e, we showed<br />

th<strong>at</strong> the additional parameters rel<strong>at</strong>e to interactions, an <strong>at</strong>tractive long range<br />

part and a strongly repulsive hard core interaction. In fact, we can say th<strong>at</strong> all<br />

devi<strong>at</strong>ions from the ideal gas law give us inform<strong>at</strong>ion about the interactions in<br />

a m<strong>at</strong>erial. The question remains, how to extract this inform<strong>at</strong>ion and how to<br />

interpret it. Van der Waals is just one simple model. The ultim<strong>at</strong>e answer can<br />

only be found in st<strong>at</strong>istical mechanics, where we calcul<strong>at</strong>e the partition function<br />

on a microscopic level, and derive equ<strong>at</strong>ions of st<strong>at</strong>e. A discussion on this<br />

level pertains to single phases. Even though the van der Waals equ<strong>at</strong>ion gives a<br />

phase transition, for each set of st<strong>at</strong>e variables only one phase is present, with<br />

the exception of the transition itself, where there is a mixture of both phases.<br />

Details of this mixture are not specified.<br />

Landau-Ginzburg theory is another way of incorpor<strong>at</strong>ing the effects of interactions.<br />

Here we recognize th<strong>at</strong> fluctu<strong>at</strong>ions exist, and th<strong>at</strong> the origin of<br />

these fluctu<strong>at</strong>ions is beyond the standard equilibrium theory. We model the<br />

effects of the fluctu<strong>at</strong>ions by a gradient term, and investig<strong>at</strong>e the consequences.<br />

This leads to the idea of a correl<strong>at</strong>ion length, and the divergence of physical<br />

properties <strong>at</strong> the phase transition. The theory is an improvement, but still too<br />

simple because it is always equivalent to mean field theory. This was elabor<strong>at</strong>ed<br />

upon in the chapter on critical exponents, where the effects of the ”real” free<br />

energy associ<strong>at</strong>ed with fluctu<strong>at</strong>ions was analyzed and rel<strong>at</strong>ions between critical<br />

exponents were developed.<br />

Finally, in the chapter on transport theory we used the existence of fluctu<strong>at</strong>ions<br />

to show why Onsager’s rel<strong>at</strong>ions are valid. Again, we used the gradient of<br />

st<strong>at</strong>e variables as a characteristic of local equilibrium. Any microscopic theory<br />

will have such terms in the macroscopic limit, and therefore Onsager’s rel<strong>at</strong>ions<br />

have to be valid for all m<strong>at</strong>erials.<br />

Correl<strong>at</strong>ions between fluctu<strong>at</strong>ions.<br />

In this chapter we will discuss the idea of interactions from a different perspective.<br />

We again divide space into cells, which are small on the macroscopic<br />

scale, but are large enough for the thermodynamic limit to be obeyed in each<br />

cell. We know th<strong>at</strong> the density in each cell fluctu<strong>at</strong>es around its average value,<br />

and we have used the local entropy to find the probability of such fluctu<strong>at</strong>ions.<br />

In order to discuss this we need to distinguish between instantaneous values<br />

of the density in a cell and its thermal average. Assume th<strong>at</strong> in thermal<br />

equilibrium the density in a cell <strong>at</strong> ⃗r is given by n(⃗r), in mols per cubic meter.<br />

The equilibrium density can be position dependent if we have an external<br />

potential which depends on position. For a homogeneous system we have, of<br />

course, n(⃗r) = N<br />

V , where N is the total amount of m<strong>at</strong>erial in the system and<br />

V the total volume.<br />

If we do an instantaneous measurement we find a value ρ(⃗r) for the density


7.1. DESCRIPTION OF CORRELATED FLUCTUATIONS. 267<br />

the cell. Because the density can fluctu<strong>at</strong>e, this is not equal to the equilibrium<br />

value, but we have<br />

⟨ρ(⃗r)⟩ = n(⃗r) (7.1)<br />

where the brackets denote taking the thermal average, or local equilibrium value.<br />

Now consider a measurement of the density <strong>at</strong> two different places, done<br />

simultaneously. We find a value ρ(⃗r)ρ(⃗r ′ ) In a simple minded theory without<br />

long range interactions fluctu<strong>at</strong>ions in neighboring cells are not correl<strong>at</strong>ed, and<br />

we have for the equilibrium value<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩ = ⟨ρ(⃗r)⟩⟨ρ(⃗r ′ )⟩ , ⃗r ̸= ⃗r ′<br />

(7.2)<br />

for fluctu<strong>at</strong>ions in different cells. For a homogeneous system the value is ( )<br />

N 2.<br />

V<br />

For the rest of the discussion we will assume th<strong>at</strong> we have a homogeneous system,<br />

unless st<strong>at</strong>ed otherwise.<br />

When we consider fluctu<strong>at</strong>ions in the same cell, however, we have to use the<br />

results from a previous chapter. Equ<strong>at</strong>ion 2.148 shows th<strong>at</strong><br />

and with<br />

⟨(∆N) 2 ⟩ = −kB<br />

( )<br />

∂S<br />

∂N U,V<br />

( ( 2 ∂ S<br />

∂N 2<br />

) ) −1<br />

V,U<br />

= − µ<br />

T<br />

(7.3)<br />

(7.4)<br />

the equ<strong>at</strong>ion for the fluctu<strong>at</strong>ions in density is quite complex, because temper<strong>at</strong>ure<br />

fluctu<strong>at</strong>es too.<br />

Density fluctu<strong>at</strong>ions only.<br />

We can simplify our model by assuming th<strong>at</strong> the system is <strong>at</strong> constant temper<strong>at</strong>ure<br />

T. Hence the only fluctu<strong>at</strong>ions we consider are fluctu<strong>at</strong>ions in the<br />

density. This is quite reasonable. Energy fluctu<strong>at</strong>es between the cells, but<br />

the transport of energy is a transport of momentum. M<strong>at</strong>erial oscill<strong>at</strong>es back<br />

and forth and the collisions transport energy. The transport of m<strong>at</strong>erial itself,<br />

though, needs a transport of mass, and this is typically slower. Hence if we<br />

look <strong>at</strong> the low frequency limit the effects of density fluctu<strong>at</strong>ions are the<br />

last too disappear. In other words, on the time scales where our description is<br />

useful, energy fluctu<strong>at</strong>ions have already died out, and we can assume a constant<br />

temper<strong>at</strong>ure everywhere.<br />

This point is not obvious. We have seen in chapter two th<strong>at</strong> we can integr<strong>at</strong>e<br />

out the energy fluctu<strong>at</strong>ions, and obtain a probability distribution function for<br />

the density fluctu<strong>at</strong>ions, with a variance depending on both fluctu<strong>at</strong>ions. Here<br />

we follow a different recipe, and assume th<strong>at</strong> the energy fluctu<strong>at</strong>ions are coupled<br />

to the outside world though the temper<strong>at</strong>ure reservoir. This means th<strong>at</strong> for a


268 CHAPTER 7. CORRELATION FUNCTIONS.<br />

given density change the average change in energy is now biased in order to<br />

keep the temper<strong>at</strong>ure the same. This changes the distribution of the density<br />

fluctu<strong>at</strong>ions, and makes the variance the simple function discussed in the next<br />

paragraph. Note th<strong>at</strong> with volume and temper<strong>at</strong>ure specified, the value of the<br />

density and the condition of local equilibrium together determine the value of<br />

the pressure and chemical potential. This means th<strong>at</strong> the pressure and chem-<br />

ical potential ) also fluctu<strong>at</strong>e, but are rel<strong>at</strong>ed by simple partial deriv<strong>at</strong>ives, like<br />

. The entropy fluctu<strong>at</strong>es, too, in a dependent manner!<br />

( ∂p<br />

∂N<br />

T,V<br />

At constant temper<strong>at</strong>ure the fluctu<strong>at</strong>ions are now given by<br />

⟨(∆N) 2 ( )<br />

∂N<br />

⟩ = kBT<br />

∂µ<br />

Because changes in intensive st<strong>at</strong>e variables are rel<strong>at</strong>ed by<br />

we see th<strong>at</strong><br />

T,V<br />

(7.5)<br />

0 = SdT − V dp + Ndµ (7.6)<br />

( )<br />

∂N<br />

∂µ T,V<br />

= N<br />

V<br />

( )<br />

∂N<br />

∂p T,V<br />

and the triple partial deriv<strong>at</strong>ive product rel<strong>at</strong>ion then tells us<br />

( )<br />

∂N<br />

∂p<br />

( )<br />

∂N<br />

= −<br />

∂V<br />

( )<br />

∂V<br />

∂p<br />

( )<br />

∂N<br />

=<br />

∂V<br />

T,V<br />

T,p<br />

T,N<br />

V κT<br />

T,p<br />

(7.7)<br />

(7.8)<br />

The partial deriv<strong>at</strong>ive <strong>at</strong> constant T and p is easy to find. If we change the<br />

volume <strong>at</strong> constant T and p, the amount of m<strong>at</strong>erial has to change in the same<br />

way, since we know th<strong>at</strong> N(V, T, p) = V n(T, p). In other words, if all intensive<br />

st<strong>at</strong>e variables are constant, the density is constant, and hence<br />

Combining all this leads to<br />

( ∂N<br />

∂V<br />

⟨(∆N) 2 ⟩ =<br />

)<br />

T,p<br />

= N<br />

V<br />

( ) 2<br />

N<br />

VcellkBT κT<br />

V<br />

(7.9)<br />

(7.10)<br />

where the volume involved is the volume of the local cell, hence the subscript.<br />

and the fluctu<strong>at</strong>ions in amount of<br />

Note th<strong>at</strong> for an ideal gas κT = 1<br />

p<br />

m<strong>at</strong>erial in mols is equal to Ncell<br />

NA<br />

= V<br />

NRT<br />

, which is independent of temper<strong>at</strong>ure!<br />

We can also include time, and discuss the correl<strong>at</strong>ion between fluctu<strong>at</strong>ions<br />

<strong>at</strong> different times. The time scale is set by the time needed for a local cell to<br />

reach equilibrium, time is also divided into cells. In th<strong>at</strong> case we have<br />

⟨ρ(⃗r, t)ρ(⃗r ′ , t ′ )⟩ =<br />

( ) 2<br />

N<br />

, ⃗r ̸= ⃗r<br />

V<br />

′<br />

or t ̸= t ′<br />

(7.11)


7.1. DESCRIPTION OF CORRELATED FLUCTUATIONS. 269<br />

⟨ρ(⃗r, t)ρ(⃗r, t)⟩ =<br />

( ) 2 (<br />

N kBT κT<br />

V Vcell<br />

where the l<strong>at</strong>ter follows using ⟨n 2 ⟩ = ⟨(∆n) 2 ⟩ + (⟨n⟩) 2 .<br />

Typical distances and times.<br />

)<br />

+ 1<br />

(7.12)<br />

How small can we choose our cell? We need enough m<strong>at</strong>erial in a cell to allow<br />

the cell to behave like a macroscopic system. Wh<strong>at</strong> is enough? Th<strong>at</strong> depends on<br />

our required precision. For example, if we have a cubic cell with sides of 10 −6 m,<br />

such a cell in a solid contains about 10 10 <strong>at</strong>oms, which is fairly large and gives<br />

fluctu<strong>at</strong>ions in particle numbers of order 10 −5 . In a gas, there are fewer <strong>at</strong>oms,<br />

say 10 6 , but this still has fluctu<strong>at</strong>ions much smaller than one percent. So we<br />

can say th<strong>at</strong> we probably should have distances larger than 1 µm. In terms of<br />

wavelengths of perturb<strong>at</strong>ions, we can say th<strong>at</strong> we need wavelength larger than<br />

1 µm.<br />

Wh<strong>at</strong> about time scales? In order to go to equilibrium we need energy to<br />

move back and forth in a cell, and a typical speed would be the speed of sound.<br />

Say th<strong>at</strong> is about 1000 m/s. The time it takes to travel 1 µm is 10 −9 s. So if<br />

we restrict ourselves to times larger than 1 µs we are in a good regime. In a<br />

frequency domain, this means frequencies below the MHz scale.<br />

Note th<strong>at</strong> these are just estim<strong>at</strong>es, and experiment only can tell us if macroscopic<br />

theories are sufficient or if we need a microscopic description. But wh<strong>at</strong><br />

is sufficient? Errors less than one percent? In addition, <strong>at</strong> low temper<strong>at</strong>ures we<br />

have to consider the effect of quantum fluctu<strong>at</strong>ions as well as thermodynamic<br />

fluctu<strong>at</strong>ions, and the former can domin<strong>at</strong>e the system.<br />

Using Dirac delta functions<br />

The two equ<strong>at</strong>ions we used for correl<strong>at</strong>ions between fluctu<strong>at</strong>ions can be combines<br />

using Kronecker delta symbols:<br />

⟨ρ(⃗r, t)ρ(⃗r ′ , t ′ )⟩ =<br />

( ) 2 (<br />

N kBT κT<br />

δc(⃗r, ⃗r<br />

V Vcell<br />

′ )δc(t, t ′ )<br />

) + 1<br />

(7.13)<br />

In this equ<strong>at</strong>ion we assume th<strong>at</strong> we can divide space into cells and these cells<br />

are small on a macroscopic scale, but large enough th<strong>at</strong> we can assume local<br />

equilibrium. We will use this equ<strong>at</strong>ion on a macroscopic scale only, and in th<strong>at</strong><br />

case we can replace summ<strong>at</strong>ions over cells by integr<strong>at</strong>ions. Because ∫ δ(⃗r)d 3 r =<br />

1 we can replace the Kronecker delta by a delta function times the cell volume,<br />

since on a macroscopic scale d 3 r = Vcell.<br />

For the Kronecker delta in time we can do the same, but we need to decide<br />

on a cell in time. In our current formalism we assume th<strong>at</strong> density fluctu<strong>at</strong>ions<br />

are slow, because they involve transport of m<strong>at</strong>erial. If a cell changes N, going<br />

to equilibrium involves the transport of thermal energy. We assume th<strong>at</strong> energy


270 CHAPTER 7. CORRELATION FUNCTIONS.<br />

transport is fast. We associ<strong>at</strong>e th<strong>at</strong> with a relax<strong>at</strong>ion time, τQ, and replace<br />

the Kronecker delta in time by a delta function times this relax<strong>at</strong>ion time.<br />

Altern<strong>at</strong>ively, we can use a relax<strong>at</strong>ion frequency defined by ωQτQ = 1. Adding<br />

these elements together we arrive <strong>at</strong> the macroscopic equ<strong>at</strong>ion:<br />

⟨ρ(⃗r, t)ρ(⃗r ′ , t ′ )⟩ =<br />

( ) 2 (<br />

N kBT κT<br />

δ(⃗r − ⃗r<br />

V ωQ<br />

′ )δ(t − t ′ )<br />

) + 1<br />

(7.14)<br />

In case the density is independent of time, which happens when we have<br />

time-independent external potentials, we need to revert back to the Kronecker<br />

delta in time and get<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩ =<br />

( ) 2<br />

N<br />

(kBT κT δ(⃗r − ⃗r<br />

V<br />

′ ) + 1) (7.15)<br />

7.2 M<strong>at</strong>hem<strong>at</strong>ical functions for correl<strong>at</strong>ions.<br />

Pair distribution function.<br />

The discussion in the previous section assumes th<strong>at</strong> each cell goes to equilibrium<br />

by exchanging he<strong>at</strong> and m<strong>at</strong>ter with a reservoir. We assume th<strong>at</strong> we<br />

know th<strong>at</strong> the values of the temper<strong>at</strong>ure and the chemical potential in each cell<br />

are given, and the standard way to specify is to use external reservoirs. But<br />

th<strong>at</strong> is not the correct picture in this case. Each cell interacts only with its<br />

neighbors, and thermal energy and m<strong>at</strong>ter have to come from these neighboring<br />

cells. If the amount of m<strong>at</strong>erial in our cell increases, the amount of m<strong>at</strong>erial in<br />

the neighboring cell decreases! The same is true for thermal energy.<br />

We are looking <strong>at</strong> time scales for which thermal equilibrium is established.<br />

Hence the exchange of thermal energy is fast, and we have de facto local equilibrium<br />

in all cells, and we can assume th<strong>at</strong> all cells are <strong>at</strong> the same temper<strong>at</strong>ure<br />

T . Therefore, is we measure an instantaneous value of the density ρ(⃗r) the measurement<br />

is slow enough th<strong>at</strong> this represents a local equilibrium value, and th<strong>at</strong><br />

the values of the density, temper<strong>at</strong>ure, and volume of the cell then determine<br />

the value of the chemical potential in the cell. If we now compare neighboring<br />

cells, we find a difference in chemical potential, which will lead to currents and<br />

in the end to the establishment of global thermal equilibrium.<br />

This implies, however, th<strong>at</strong> there is correl<strong>at</strong>ion between the values of the<br />

density in neighboring cells, and hence we cannot assume th<strong>at</strong> the thermal<br />

averages over neighboring cells are independent. In addition, the presence of<br />

interactions changes the notion th<strong>at</strong> fluctu<strong>at</strong>ions are independent. Th<strong>at</strong> also<br />

leads to correl<strong>at</strong>ions between different cells.<br />

We define the pair distribution function g by:<br />

⟨ρ(⃗r, t)ρ(⃗r ′ , t ′ )⟩ =<br />

( ) 2<br />

N<br />

g(⃗r, t, ⃗r<br />

V<br />

′ , t ′ ) (7.16)


7.2. MATHEMATICAL FUNCTIONS FOR CORRELATIONS. 271<br />

where we require th<strong>at</strong> the position variables are NOT in the same cell. If the<br />

system were not homogeneous, we would change this to<br />

⟨ρ(⃗r, t)ρ(⃗r ′ , t ′ )⟩ = n(⃗r, t)n(⃗r ′ , t ′ )g(⃗r, t, ⃗r ′ , t ′ ) (7.17)<br />

In homogeneous systems this function is only a function of the distances. In<br />

th<strong>at</strong> case we have no time dependence of the variables. We will now restrict<br />

ourselves to th<strong>at</strong> case, and write:<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩ =<br />

( ) 2<br />

N<br />

g(⃗r − ⃗r<br />

V<br />

′ ) (7.18)<br />

and use the same symbol g, but now with two arguments. Note th<strong>at</strong> the pair<br />

distribution function is a dimensionless quantity.<br />

It is important to realize th<strong>at</strong> this defines the pair distribution function as a<br />

thermodynamic function. Its definition relies on local equilibrium, and its values<br />

will depend on the thermodynamic st<strong>at</strong>e of the whole system. In other words,<br />

its values will depend on how we measure it, in the same way as the value of<br />

the he<strong>at</strong> capacity depends on working <strong>at</strong> constant temper<strong>at</strong>ure or entropy. This<br />

point should not be overlooked when one makes connections with microscopic<br />

definitions in st<strong>at</strong>istical mechanics!<br />

If we now require th<strong>at</strong> the pair distribution function is continuous in the<br />

limit of zero distance, we need to modify the definition of this function when<br />

the sp<strong>at</strong>ial arguments are in the same cell. Correl<strong>at</strong>ions in one cell are taken into<br />

account by fluctu<strong>at</strong>ions, but the values are now different because neighboring<br />

cells are correl<strong>at</strong>ed. Hence we cannot use the same expression as we derived for<br />

the uncorrel<strong>at</strong>ed system and therefore we write:<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩ =<br />

( ) 2<br />

N<br />

(g(⃗r − ⃗r<br />

V<br />

′ ) + F δ(⃗r − ⃗r ′ )) (7.19)<br />

with F an unknown function of temper<strong>at</strong>ure.<br />

Because of our conditions of local equilibrium, this definition only makes<br />

sense for distances and time differences which are large enough so we have<br />

local equilibrium. In a Fourier picture, this means th<strong>at</strong> we can only describe<br />

the long wavelength and low frequency limit. A microscopic theory is needed<br />

to give the pair distribution functions for all distances and times, but it is<br />

important to realize th<strong>at</strong> in the limit mentioned above the microscopic results<br />

should agree with the concepts we develop here. Our goal for this chapter<br />

is therefore to derive some rel<strong>at</strong>ions th<strong>at</strong> the pair distribution functions have<br />

to obey, based on simple thermodynamic properties. This will again give us<br />

limits for a microscopic theory. Also, many experiments are performed under<br />

conditions of local equilibrium, and for such experiments the results derived in<br />

this chapter should be valid. The key to deciding the validity is the question<br />

wh<strong>at</strong> is a large wavelength and wh<strong>at</strong> is a low frequency.<br />

In a general thermodynamical approach, where we ignore details on the<br />

<strong>at</strong>omic structure, the pair distribution function 7.27 depends only on the difference<br />

vector ⃗r − ⃗r ′ . In a gas or liquid one can simplify this even further, because


272 CHAPTER 7. CORRELATION FUNCTIONS.<br />

such fluids are isotropic, and the pair distribution function depends only on the<br />

scalar distance |⃗r −⃗r ′ |. In a solid, however, this symmetry is always broken. For<br />

example, one can study layered m<strong>at</strong>erials with properties th<strong>at</strong> are very different<br />

in plane as compared to out of plane. Therefore, we will study the more general<br />

form. Additional simplific<strong>at</strong>ions due to the isotropic n<strong>at</strong>ure of fluids can be<br />

derives easily.<br />

It is also important to realize th<strong>at</strong> the averages we are taking are thermodynamical<br />

or macroscopic averages. Th<strong>at</strong> is quite different from the quantum<br />

mechanical averages one includes on a microscopic scale. For example, in a two<br />

particle system we can rel<strong>at</strong>e the joint probability of finding a particle <strong>at</strong> place<br />

one and the other <strong>at</strong> place two directly to the wave function. In thermodynamics,<br />

we talk about many particles and averages due to the fact th<strong>at</strong> all these<br />

particles interact. The road from the microscopic behavior to the macroscopic<br />

setting is not clearly defined, and is a very interesting topic of study. Again, for<br />

another course.<br />

Since we study the pair distribution function as a thermodynamic quantity,<br />

we need to be precise which other st<strong>at</strong>e variables are kept constant, and which<br />

are not. In our simple case we assume a constant temper<strong>at</strong>ure T. This means<br />

th<strong>at</strong> the pressure will fluctu<strong>at</strong>e, since the pressure follows from the equ<strong>at</strong>ion of<br />

st<strong>at</strong>e, given the density and temper<strong>at</strong>ure. Remember, we assume local equilibrium!<br />

If we also allow the energy to fluctu<strong>at</strong>e, then the temper<strong>at</strong>ure will<br />

fluctu<strong>at</strong>e, too. Th<strong>at</strong> is a more general case. One could restrict this by assuming<br />

a constant pressure, which might be appropri<strong>at</strong>e in certain situ<strong>at</strong>ions. But<br />

no m<strong>at</strong>ter wh<strong>at</strong> the restrictions are, like always in thermodynamics one has to<br />

know wh<strong>at</strong> they are, and be aware of the fact th<strong>at</strong> different external constraints<br />

yield different results for thermodynamic quantities like the pair distribution<br />

function.<br />

The value of F can be chosen as follows. Suppose we integr<strong>at</strong>e ⃗r ′ in the<br />

correl<strong>at</strong>ion function over a very small volume ν around ⃗r. Th<strong>at</strong> leads to<br />

∫<br />

ν<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩d 3 r ′ =<br />

( ) 2 (∫<br />

N<br />

g(⃗r − ⃗r<br />

V ν<br />

′ )d 3 r ′ )<br />

+ F<br />

(7.20)<br />

Now consider an ideal gas, where there are no interactions between particles, in<br />

which case we have g = 1 and hence<br />

∫<br />

ν<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩d 3 r ′ =<br />

Next we integr<strong>at</strong>e ⃗r over the same volume to get<br />

∫<br />

ν<br />

∫<br />

ν<br />

( ) 2<br />

N<br />

(ν + F ) (7.21)<br />

V<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩d 3 rd 3 r ′ ( ) 2 (<br />

Nν<br />

= 1 +<br />

V<br />

F<br />

)<br />

ν<br />

(7.22)<br />

This procedure is allowed, even for a very small volume ν, since we are<br />

working in a time independent theory. By definition, we can wait infinitely long<br />

for our measurement, and hence we can establish thermal equilibrium. But wh<strong>at</strong>


7.2. MATHEMATICAL FUNCTIONS FOR CORRELATIONS. 273<br />

happens if we make the volume ν so small th<strong>at</strong> it is comparable with particle<br />

size? In th<strong>at</strong> case we can have only zero or one particle in our little volume,<br />

and we have, using Avogadro’s number NA:<br />

∫<br />

ν<br />

∫<br />

⟨ρ(⃗r)NAρ(⃗r ′ )NA⟩d 3 rd 3 r ′ ∫<br />

= ⟨ρ(⃗r)NA⟩d 3 N<br />

r = NA ν (7.23)<br />

V<br />

ν<br />

since 0 2 = 0 and 1 2 = 1. Hence in this limit we have<br />

or<br />

N<br />

V NA<br />

ν<br />

( ) 2 (<br />

Nν<br />

ν = 1 +<br />

V<br />

F<br />

)<br />

ν<br />

V<br />

NNAν =<br />

(<br />

1 + F<br />

)<br />

ν<br />

(7.24)<br />

(7.25)<br />

F = V<br />

− ν (7.26)<br />

NNA<br />

In this last equ<strong>at</strong>ion the first term on the right hand side is the average<br />

volume for one particle in the gas, and the second term is the volume of a<br />

particle itself, which is much smaller, and can be ignored for point particles.<br />

Note th<strong>at</strong> this leaves some questions for real particles in solids! Hence we have<br />

found a value for F , based on the knowledge th<strong>at</strong> m<strong>at</strong>ter is indeed made of<br />

<strong>at</strong>oms, and th<strong>at</strong> two <strong>at</strong>oms cannot be <strong>at</strong> the same place.<br />

This finally leads to our definition of the pair distribution function:<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩ =<br />

Pair correl<strong>at</strong>ion function.<br />

( ) 2 (<br />

N<br />

g(⃗r − ⃗r<br />

V<br />

′ ) + V<br />

δ(⃗r − ⃗r<br />

NNA<br />

′ )<br />

)<br />

(7.27)<br />

In many cases we are interested in the correl<strong>at</strong>ions between fluctu<strong>at</strong>ions.<br />

This follows quite easily from equ<strong>at</strong>ion 7.27:<br />

⟨(ρ(⃗r) − N<br />

V )(ρ(⃗r′ ) − N<br />

)⟩ =<br />

V<br />

and the function<br />

( ) 2 (<br />

N<br />

g(⃗r − ⃗r<br />

V<br />

′ , t − t ′ ) − 1 + V<br />

δ(⃗r − ⃗r<br />

NNA<br />

′ )<br />

)<br />

(7.28)<br />

h(⃗r − ⃗r ′ , t − t ′ ) = g(⃗r − ⃗r ′ , t − t ′ ) − 1 (7.29)<br />

is commonly called the pair correl<strong>at</strong>ion function. Without correl<strong>at</strong>ions between<br />

fluctu<strong>at</strong>ions, this function is zero unless we consider the fluctu<strong>at</strong>ions in one cell<br />

<strong>at</strong> one time.


274 CHAPTER 7. CORRELATION FUNCTIONS.<br />

It is instructive to evalu<strong>at</strong>e the following integral<br />

∫<br />

H = d 3 rh(⃗r) (7.30)<br />

where we take the correl<strong>at</strong>ion function <strong>at</strong> equal time, and omit the time argument.<br />

If we go back to equ<strong>at</strong>ion 7.27 we can integr<strong>at</strong>e both ⃗r and ⃗r ′ over the<br />

total volume and get<br />

or<br />

∫<br />

V<br />

∫<br />

V<br />

⟨ρ(⃗r)ρ(⃗r ′ )⟩d 3 rd 3 r ′ =<br />

⟨N 2 ⟩ =<br />

( ) 2 ( ∫<br />

N<br />

V g(⃗r)d<br />

V<br />

V<br />

3 r +<br />

( ) 2 ( ∫<br />

N<br />

V g(⃗r)d<br />

V<br />

V<br />

3 r +<br />

V 2<br />

NNA<br />

)<br />

) 2 V<br />

N<br />

(7.31)<br />

(7.32)<br />

Assume we have a system <strong>at</strong> temper<strong>at</strong>ure T and chemical potential µ. In<br />

th<strong>at</strong> case we have<br />

and hence<br />

or<br />

which leads to<br />

⟨(∆N) 2 ⟩ =<br />

( ) 2<br />

N<br />

V kBT κT + N<br />

V<br />

2 =<br />

V kBT κT + V 2 ∫<br />

= V<br />

( ) 2<br />

N<br />

V kBT κT<br />

V<br />

( ) 2 ( ∫<br />

N<br />

V g(⃗r)d<br />

V<br />

V<br />

3 r +<br />

V<br />

V 2<br />

g(⃗r)d 3 r +<br />

NNA<br />

V 2<br />

NNA<br />

)<br />

(7.33)<br />

(7.34)<br />

(7.35)<br />

H = kBT κT − V<br />

NNA<br />

(7.36)<br />

This is consistent with the equ<strong>at</strong>ions of st<strong>at</strong>e for an ideal gas, which has<br />

κT = 1<br />

p and hence H = 0. For an ideal gas we have g = 1 and h = 0 indeed.<br />

The equ<strong>at</strong>ion above is called the compressibility equ<strong>at</strong>ion of st<strong>at</strong>e. We can take<br />

a deriv<strong>at</strong>ive to get a rel<strong>at</strong>ed form:<br />

∫<br />

NA d 3 ( )<br />

∂h<br />

r<br />

∂T<br />

( )<br />

∂RT κT<br />

(⃗r, T ) =<br />

∂T<br />

(7.37)<br />

N,V<br />

which is another useful form of this equ<strong>at</strong>ion of st<strong>at</strong>e! Note th<strong>at</strong> in the equ<strong>at</strong>ion<br />

7.37 we use h(T, µ, V ; ⃗r), but th<strong>at</strong> the deriv<strong>at</strong>ive is <strong>at</strong> constant N!<br />

There are two useful limits of the equ<strong>at</strong>ion of st<strong>at</strong>e in equ<strong>at</strong>ion 7.36. In<br />

liquids and solids the compressibility is much smaller than th<strong>at</strong> of an ideal gas,<br />

and we have:<br />

N,V


7.3. ENERGY CONSIDERATIONS. 275<br />

H ≈ − V<br />

NNA<br />

Liquids and solids (7.38)<br />

which means th<strong>at</strong> fluctu<strong>at</strong>ions are reduced from ideal gas values because <strong>at</strong>oms<br />

cannot be <strong>at</strong> the same place. The other case occurs near phase transitions,<br />

where κT → ∞, in which case we have<br />

H → ∞ Near phase transitions (7.39)<br />

consistent with the fact th<strong>at</strong> fluctu<strong>at</strong>ions now have a very long range.<br />

7.3 Energy consider<strong>at</strong>ions.<br />

Energy of fluctu<strong>at</strong>ions.<br />

The m<strong>at</strong>erial in different cells interact with each other. There is an energy<br />

associ<strong>at</strong>ed with this interaction, which we call Ueff (⃗r, ⃗r ′ ). It is important to<br />

realize th<strong>at</strong> this is an effective interaction, since we look <strong>at</strong> the interactions<br />

between cells which are in local equilibrium. This interaction is much smaller<br />

than the real <strong>at</strong>omic interaction between the <strong>at</strong>oms in the cell, since screening<br />

and similar many body effects are included in this interaction. If the fluctu<strong>at</strong>ions<br />

in different cells where uncorrel<strong>at</strong>ed, this interaction would be zero. But since<br />

there is correl<strong>at</strong>ion, some interaction survives. Since we study the interaction<br />

between cells, this is clearly a long range interaction. In addition, we can assume<br />

th<strong>at</strong> the interaction depends on the difference between the positions only.<br />

The energy associ<strong>at</strong>ed with the fluctu<strong>at</strong>ions can be easily evalu<strong>at</strong>ed. If the<br />

energy for the homogeneous system is Uh then the energy due to the fluctu<strong>at</strong>ions<br />

is given by (since this is equal time, we leave out the time variable):<br />

Uf (T, V, N)) = 1<br />

∫<br />

2<br />

d 3 ∫<br />

r<br />

d 3 r ′ Ueff (⃗r − ⃗r ′ )⟨(ρ(⃗r) − N<br />

V<br />

N<br />

)(ρ(⃗r) − )⟩ (7.40)<br />

V<br />

The factor one half is needed to avoid double counting. Also, we need to be<br />

careful how to tre<strong>at</strong> the integr<strong>at</strong>ion when both position variables describe the<br />

same cell. Part of th<strong>at</strong> energy is already included in the thermodynamic description<br />

of the system <strong>at</strong> constant density. This can be done by replacing the<br />

correl<strong>at</strong>ion function by the pair correl<strong>at</strong>ion function without the term depending<br />

on F . Another way of looking <strong>at</strong> this is to do the integr<strong>at</strong>ion, and observe th<strong>at</strong><br />

the term with F looks like F Ueff (0), which can be absorbed in Uh. Hence we<br />

have<br />

Uf (T, V, N)) =<br />

2 N<br />

2V 2<br />

∫<br />

d 3 ∫<br />

r<br />

d 3 r ′ Ueff (⃗r − ⃗r ′ )h(⃗r − ⃗r ′ ) =<br />

N 2<br />

2V<br />

∫<br />

d 3 rUeff (⃗r)h(⃗r)<br />

(7.41)


276 CHAPTER 7. CORRELATION FUNCTIONS.<br />

The energy is a function of the total volume, total number of particles, and<br />

temper<strong>at</strong>ure. This needs some explan<strong>at</strong>ion. We originally defined fluctu<strong>at</strong>ions<br />

in a large cell of given volume, amount of m<strong>at</strong>erial, and energy. But then we<br />

assumed th<strong>at</strong> energy fluctu<strong>at</strong>ions are fast and th<strong>at</strong> the energy is always the<br />

thermal equilibrium value for the fluctu<strong>at</strong>ing density <strong>at</strong> the temper<strong>at</strong>ure for the<br />

whole system. In order for th<strong>at</strong> to be meaningful, it means th<strong>at</strong> we replace our<br />

boundary condition for the large cell by a condition of constant temper<strong>at</strong>ure.<br />

But we still assume th<strong>at</strong> the total amount of m<strong>at</strong>erial in the large volume is<br />

constant. We could also allow th<strong>at</strong> to change, by working <strong>at</strong> a given chemical<br />

potential.<br />

The way we have defined the pair correl<strong>at</strong>ion function is as a thermodynamic<br />

quantity, and it will depend on the st<strong>at</strong>e variables of the whole system. So we<br />

either have hN(T, V, N, ⃗s) or hµ(T, V, µ, ⃗s) and as a function of position these<br />

are different functions! Again, for the compressibility equ<strong>at</strong>ion of st<strong>at</strong>e 7.36 we<br />

need to use the pair distribution function as a function of µ!<br />

It can be useful to rewrite this by subtracting and adding a typical energy:<br />

Uf (T, V, N)) = V<br />

2<br />

( ) 2 ∫<br />

N<br />

V<br />

1<br />

2 V<br />

d 3 s (Ueff (⃗s) − Uave) h(T, V, N; ⃗s)+<br />

( ) 2 ∫<br />

N<br />

Uave<br />

V<br />

and using the fluctu<strong>at</strong>ion equ<strong>at</strong>ion of st<strong>at</strong>e we get<br />

Uf (T, V, N)) = 1<br />

2 V<br />

1<br />

2 V<br />

( ) 2 ∫<br />

N<br />

V<br />

If we now define the average energy by<br />

∫<br />

we have<br />

d 3 sh(⃗s) (7.42)<br />

d 3 s (Ueff (⃗s) − Uave) h(⃗s)+<br />

( ) 2<br />

N<br />

(RT κT −<br />

V<br />

1<br />

n )Uave<br />

NA<br />

d 3 ∫<br />

sUeff (⃗s)hN (T, V, N, ⃗s) = Uave(T, V, N)<br />

Uf (T, V, N) = 1<br />

2 V<br />

( ) 2<br />

N<br />

(RT κT −<br />

V<br />

1<br />

n )Uave<br />

NA<br />

(7.43)<br />

d 3 shN (T, V, N, ⃗s) (7.44)<br />

(7.45)<br />

There are two interesting applic<strong>at</strong>ions of this equ<strong>at</strong>ion. For an ideal gas we<br />

have κT = 1 1<br />

p = nRT and hence Uf (T, V, N) ≡ 0. This is not correct, however.<br />

Because the integral of the pair correl<strong>at</strong>ion function is zero, we cannot define<br />

the average energy (it would be infinite), and we are left with the basic formula<br />

in terms of the integral over the effective interaction.


7.3. ENERGY CONSIDERATIONS. 277<br />

Uave<br />

2N NA .<br />

Finally, we consider the effects of making a change in volume by changing<br />

coordin<strong>at</strong>es according to ⃗r ′ = λ⃗r. This gives V ′ = λ3V and<br />

For a solid we have κT ≈ 0 and hence Uf (T, V, N) ≈ − 1<br />

Uf (T, V ′ , N)) =<br />

2 N<br />

2V ′<br />

∫<br />

V ′<br />

d 3 r ′ Ueff (⃗r ′ )h(⃗r ′ ) (7.46)<br />

The additional pressure due to this interaction follows from pf = − ( ∂U<br />

∂V<br />

which gives in the limit λ → 1<br />

pf = − Uf (T, V ′ , N)) − Uf (T, V, N))<br />

V ′ − V<br />

)<br />

S,N<br />

(7.47)<br />

The difference in internal energy is<br />

N 2<br />

2V ′<br />

∫<br />

V ′<br />

d 3 r ′ Ueff (⃗r ′ )h(⃗r ′ 2 ∫<br />

N<br />

) − d<br />

2V V<br />

3 rUeff (⃗r)h(⃗r) (7.48)<br />

Now we rewrite the integr<strong>at</strong>ion in the first term to an integr<strong>at</strong>ion over V<br />

N 2<br />

2V ′<br />

∫<br />

d<br />

V<br />

3 2 ∫<br />

N<br />

rUeff (λ⃗r)h(λ⃗r) − d<br />

2V V<br />

3 rUeff (⃗r)h(⃗r) (7.49)<br />

This is rearranged in the form<br />

N 2<br />

2V ′<br />

∫<br />

d<br />

V<br />

3 ( ) 2 2 ∫<br />

N N<br />

r {Ueff (λ⃗r)h(λ⃗r) − Ueff (⃗r)h(⃗r)} + − d<br />

2V ′ 2V V<br />

3 rUeff (⃗r)h(⃗r)<br />

(7.50)<br />

For the first part we need to recognize th<strong>at</strong> we require the difference <strong>at</strong> constant<br />

entropy. Th<strong>at</strong> requires th<strong>at</strong> all rel<strong>at</strong>ive distributions remain the same, and th<strong>at</strong><br />

in the first integral we can replace h(λ⃗r) by h(⃗r). But th<strong>at</strong> is not sufficient,<br />

since we need to keep the total amount of fluctu<strong>at</strong>ions constant as well. This<br />

means th<strong>at</strong> we need to replace h(λ⃗r) by λ3h(⃗r), because this keeps the density<br />

constant. Therefore we have<br />

N 2<br />

2V ′<br />

∫<br />

d<br />

V<br />

3 r { λ 3 Ueff (λ⃗r) − Ueff (⃗r) } ( ) 2 2 ∫<br />

N N<br />

h(⃗r) + − d<br />

2V ′ 2V V<br />

3 rUeff (⃗r)h(⃗r)<br />

(7.51)<br />

N 2<br />

2V ′<br />

∫<br />

d<br />

V<br />

3 rλ 3 [⃗r · ⃗ 2 N<br />

∇Ueff (⃗r)]∆λh(⃗r) +<br />

2V ′<br />

∫<br />

d<br />

V<br />

3 r(λ 3 − 1)Ueff (⃗r)h(⃗r)<br />

V ′<br />

V<br />

( ) 2 2 ∫<br />

N N<br />

+ − d<br />

2V ′ 2V V<br />

3 rUeff (⃗r)h(⃗r) (7.52)<br />

Because λ3 = we see th<strong>at</strong> the second and third term cancel. Next we use<br />

and get<br />

∆V<br />

V<br />

= 3 ∆λ<br />

λ


278 CHAPTER 7. CORRELATION FUNCTIONS.<br />

N 2 ∫<br />

∆V d<br />

6V 2<br />

V<br />

3 rλ[⃗r · ⃗ ∇Ueff (⃗r)]h(⃗r) (7.53)<br />

In the limit ∆V going to zero λ goes to one, and we see th<strong>at</strong> the additional<br />

pressure due to the fluctu<strong>at</strong>ions is given by<br />

2 N<br />

pf = −<br />

6V 2<br />

∫<br />

d 3 rλ[⃗r · ⃗ ∇Ueff (⃗r)]h(⃗r) (7.54)<br />

V<br />

This is a very useful formula and can be generalized to much smaller length<br />

scales and <strong>at</strong>omic interactions, as long as we keep in mind th<strong>at</strong> the formula is<br />

derived using thermal equilibrium conditions. On smaller length scales we have<br />

to be careful how to define such conditions!


PART III<br />

Additional M<strong>at</strong>erial<br />

279


280


Appendix A<br />

Questions submitted by<br />

students.<br />

Students were asked to submit 10 question they would like to see included in<br />

the notes, similar to the questions already included, shown in italics. Errors are<br />

due to my editing. Some questions were left out because of duplic<strong>at</strong>ion.<br />

A.1 Questions for chapter 1.<br />

• Wh<strong>at</strong> is the high temper<strong>at</strong>ure reservoir for an internal combustion engine?<br />

• Are the rel<strong>at</strong>ions T<br />

T0 = 1 − ηC assured, derived, or observed?<br />

• Is there a standard or limit to which real gases can follow the ideal gas<br />

approxim<strong>at</strong>ions?<br />

• The energy of an ideal gas is D<br />

2 NRT , where D is the number of dimensions?<br />

Why is there a factor one-half for each dimension?<br />

• Is there an extensive variable th<strong>at</strong> we have not yet figured out how to<br />

measure, and how will this ignorance affect existing laws?<br />

• Work can be done in many different ways. Do different mechanisms for<br />

he<strong>at</strong> transfer lead to different T dS values?<br />

• Our bodies have more or less the same temper<strong>at</strong>ure, but do give of different<br />

amounts of he<strong>at</strong>. Can this be used to calcul<strong>at</strong>e the disorder in our bodies<br />

and predict human defects?<br />

• In a system th<strong>at</strong> is not in equilibrium, wh<strong>at</strong> can we say about deterministic<br />

behavior?<br />

• Do we have the same thermodynamic laws on the moon?<br />

281


282 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Are the laws of thermodynamics valid in acceler<strong>at</strong>ed systems?<br />

• Does the second law of thermodynamics hold for our mind? Could we<br />

cre<strong>at</strong>e more than we have learned?<br />

• Is it always possible to find other reversible processes th<strong>at</strong> have the same<br />

initial and final st<strong>at</strong>e?<br />

• The third law st<strong>at</strong>es th<strong>at</strong> lim<br />

T →0 S(T ) = 0, but does this not lead to contradictions<br />

when we also have ∆S = ∆Q<br />

T ?<br />

• How can we explain the second law of thermodynamics when an electric<br />

field acts on a dielectric and rearranges the dipoles?<br />

• Is the third law derived from experimental observ<strong>at</strong>ions?<br />

• Wh<strong>at</strong> would our world look like <strong>at</strong> T=0?<br />

• Are st<strong>at</strong>e variables always dependent on other st<strong>at</strong>e variables?<br />

• Give an example of a system th<strong>at</strong> can be made to be in thermodynamic<br />

equilibrium<br />

• Are there systems th<strong>at</strong> are not in equilibrium but th<strong>at</strong> can be modelled<br />

using thermodynamic equilibrium assumptions?<br />

• Wh<strong>at</strong> would be a consequence of the zeroth law being wrong?<br />

• Are there easy rel<strong>at</strong>ions for exact differentials when we have more than<br />

two variables? E.g. when is ¯ dg = h(x, y, z)dx + k(x, y, z)dy + j(x, y, z)dz<br />

exact?<br />

• Wh<strong>at</strong> is the importance of reversibility?<br />

• Can the efficiency be exactly zero?<br />

• If entropy change is irreversible, wh<strong>at</strong> does th<strong>at</strong> imply for the f<strong>at</strong>e of the<br />

universe?<br />

• If in the equ<strong>at</strong>ion lim<br />

N→∞ lim S = 0 we could interchange the limits, wh<strong>at</strong><br />

T →0<br />

would the consequences be?<br />

• How would things change if entropy was reversible?<br />

• Give some examples of reversible processes<br />

• Give some examples of irreversible processes<br />

• Give some examples of other extensive thermodynamic variables<br />

• Give some examples of other intensive thermodynamic variables


A.1. QUESTIONS FOR CHAPTER 1. 283<br />

• Find the exact differential df for f(x, y) = 3x y − 6yx 2 and y = y(u, v)<br />

• Wh<strong>at</strong> are the consequences of having a Carnot engine with η > 1?<br />

• Wh<strong>at</strong> are the consequences of having a Carnot engine with η < 0?<br />

• Wh<strong>at</strong> happens to the efficiency of a Carnot engine when T1 is finite and<br />

T2 → 0?<br />

• If the entropy of the universe is always increasing, is the universe a reversible<br />

process?<br />

• Why can’t we measure entropy directly?<br />

• How do you know when you have enough st<strong>at</strong>e variables?<br />

• Why does a process have to be reversible in order to study it in thermodynamics?<br />

• Why are the laws of thermodynamics numbered 0 to 3 and not 1 to 4?<br />

• In defining he<strong>at</strong> and work flow, which sign convention do you prefer and<br />

why?<br />

• How would you st<strong>at</strong>e the second law?<br />

• Why does the cycle used in the Carnot cycle produce the most efficient<br />

engine?<br />

• Why would you ever use a less efficient engine?<br />

• Why will a completely isol<strong>at</strong>ed system end up in equilibrium?<br />

• Why can’t we reach zero temper<strong>at</strong>ure?<br />

• If you would be designing a temper<strong>at</strong>ure scale, wh<strong>at</strong> would you base it on<br />

and why?<br />

• The universe is cooling down; can it reach zero temper<strong>at</strong>ure?<br />

• Is the big bang a reversible process?<br />

• Why would someone want to study macroscopic systems?<br />

• Are there any st<strong>at</strong>e variables describing a piece of iron th<strong>at</strong> we have not<br />

yet mentioned?<br />

• Give examples of st<strong>at</strong>e functions<br />

• Give an example of a spontaneous process<br />

• Wh<strong>at</strong> is the difference between Kelvin’s and Clausius’ st<strong>at</strong>ements of the<br />

second law?


284 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Why is temper<strong>at</strong>ure considered a st<strong>at</strong>e variable, but he<strong>at</strong> not?<br />

• Why can’t we measure temper<strong>at</strong>ure directly?<br />

• Why does the error in the integr<strong>at</strong>ion when adding many Carnot cycles to<br />

describe an arbitrary process approach zero when ∆T → 0 for each cycle?<br />

• Why is the T dS term necessary in the first law?<br />

• Why does the entropy increase when a system reaches equilibrium in a<br />

spontaneous process?<br />

• Wh<strong>at</strong> is a system?<br />

• How can you tell the st<strong>at</strong>e a system is in?<br />

• Is the n<strong>at</strong>ural evapor<strong>at</strong>ion/rain cycle an engine?<br />

• Does he<strong>at</strong> flow?<br />

• Why is the combin<strong>at</strong>ion of an electric motor and gener<strong>at</strong>or not capable of<br />

being more efficient than a Carnot engine?<br />

• Wh<strong>at</strong> makes an engine less efficient than a Carnot engine?<br />

• Considering the example of combining two Carnot engines to get a Carnot<br />

engine oper<strong>at</strong>ing between arbitrary temper<strong>at</strong>ures, does this show th<strong>at</strong> temper<strong>at</strong>ure<br />

is a st<strong>at</strong>e variable?<br />

• Are there consequences to the universe from the inequality of irreversible<br />

processes?<br />

• Does the maximiz<strong>at</strong>ion of entropy have any universal consequences?<br />

( )<br />

∂U<br />

• How does the st<strong>at</strong>ement lim<br />

T →0<br />

∂T = 0 about the limit of the partial<br />

V,N<br />

deriv<strong>at</strong>ive agree/disagree with the quantum mechanical st<strong>at</strong>ement th<strong>at</strong><br />

the energy of an oscill<strong>at</strong>or is never <strong>at</strong> the bottom of the well?<br />

A.2 Questions for chapter 2.<br />

• Why is the minimum energy principle important?<br />

• Wh<strong>at</strong> is the usefulness of writing all of these free energies in differential<br />

form?<br />

• How many Maxwell rel<strong>at</strong>ions can there be?<br />

• Why are response functions used in experiment?<br />

• If your finger had a small he<strong>at</strong> capacity, does this mean your finger would<br />

burn less if you touched a hot stove?


A.2. QUESTIONS FOR CHAPTER 2. 285<br />

• Wh<strong>at</strong> is the usefulness of writing S as a function of T, V, and X?<br />

• If the compressibility <strong>at</strong> constant entropy was less than zero, wh<strong>at</strong> consequences<br />

would this have on engines?<br />

• Based on physical grounds, why is the product in 2.102 less than or equal<br />

to 0?<br />

• Wh<strong>at</strong> are the possible uses of the magnetic susceptibility <strong>at</strong> constant temper<strong>at</strong>ure<br />

and the magnetic susceptibility <strong>at</strong> constant entropy?<br />

• Wh<strong>at</strong> would be some of the effects on closed systems if the expect<strong>at</strong>ion<br />

value of the product of delta U and delta N was zero.<br />

• In the third case in the beginning of chapter 2,how can we be sure th<strong>at</strong><br />

the equilibrium th<strong>at</strong> has been reached by suspending the mass is the same<br />

equilibrium without it?<br />

• Why we can choose all the independent variables to be extensive?<br />

• Wh<strong>at</strong> is more fundamental in Maxwell rel<strong>at</strong>ions?Is it the m<strong>at</strong>hem<strong>at</strong>ical<br />

rules or the physical rel<strong>at</strong>ions?<br />

• Are equ<strong>at</strong>ions 2.31 and 2.32 valid for irreversible process?<br />

• Explain the three lines after 2.38 about compressibility?<br />

• Prove equ<strong>at</strong>ion 2.52?<br />

• How can we explain in physical words equ<strong>at</strong>ion 2.84 which rel<strong>at</strong>es different<br />

response functions?<br />

• How our analysis of a magnetic system will change if we have a nonlinear<br />

magnetic response (ferromagnetic m<strong>at</strong>erial)?<br />

• How can we include the higher order susceptibilities in our analysis?<br />

• Is it possible to determine ω in equ<strong>at</strong>ion 2.144 by macroscopic analysis<br />

only?<br />

• Is the correl<strong>at</strong>ion between fluctu<strong>at</strong>ion in U and N trivial (reasonable)?<br />

• Are all free energies a minimum of their internal degrees of freedom when<br />

in equilibrium?<br />

• Why are the extensive entropy second deriv<strong>at</strong>ives rel<strong>at</strong>ed to a no-root<br />

condition?<br />

• Does the value − V<br />

4p<br />

in 2.136 have any special significance?<br />

• Why are the quantities in 2.144 a Gaussian?<br />

• Why does the exponent in 2.148 contain a factor of 1<br />

2 ?


286 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Should the fluctu<strong>at</strong>ions going to zero in the thermodynamic limit surprise<br />

you?<br />

• Is the correl<strong>at</strong>ion of the fluctu<strong>at</strong>ions a consequence of the uncertainty<br />

principle or does it represent a more general uncertainty rel<strong>at</strong>ionship?<br />

• When will the partial deriv<strong>at</strong>ives cancel in a special way and why?<br />

• Why do the deriv<strong>at</strong>ive permut<strong>at</strong>ions change sign?<br />

• In a non-equilibrium condition is it possible to have a neg<strong>at</strong>ive he<strong>at</strong> capacity<br />

and wh<strong>at</strong> would it mean?<br />

• Since there are several different ways to write the energy in thermodynamics,<br />

is there something fundamentally missing from thermodynamics<br />

th<strong>at</strong> is not missing from classical mechanics?<br />

• We have these things called Maxwell Rel<strong>at</strong>ions for different thermodynamics<br />

variables. Can we formul<strong>at</strong>e such rel<strong>at</strong>ions for other disciplines?<br />

• When doing flow in magnetics, or other solid st<strong>at</strong>e phenomenological processes,<br />

we run into vortices. These seem to be a special form of fluctu<strong>at</strong>ions.<br />

Are they?<br />

• Wh<strong>at</strong> are the implic<strong>at</strong>ions to the theory if we were ever to find a neg<strong>at</strong>ive<br />

compressibility or he<strong>at</strong> capacity.<br />

• Can we really make a closed system, and why would we want to, since we<br />

can’t interact with it.<br />

• Most of the inconsistencies in thermodynamics can be ironed out with<br />

quantum mechanics, but the first uses of thermodynamics were well before<br />

quantum mechanics, how did those scientists work out the problems?<br />

• With the three systems th<strong>at</strong> we discussed early in the chapter, only one<br />

was experimental, but all three were different, why?<br />

• Can you derive a general form of the Maxwell rel<strong>at</strong>ions th<strong>at</strong> we can then<br />

just plug the different variables into?<br />

• Why do we rely so much on the ideal gas law when it only works in<br />

equilibrium, and most systems we study are not in equilibrium?<br />

• How difficult is it to find Maxwell rel<strong>at</strong>ions for more than two variables?<br />

• How can a system exchange he<strong>at</strong> with the outside world, yet remain <strong>at</strong><br />

constant entropy?<br />

• Why are the SdT − V dp + Ndµ + Xdx terms not st<strong>at</strong>ed when expressing<br />

dU?<br />

• In equ<strong>at</strong>ion 2.66 is ( )<br />

∂U<br />

∂S V,N,X =<br />

( )<br />

∂U<br />

∂Sl V,N,X = Tl ?


A.3. QUESTIONS FOR CHAPTER 3. 287<br />

• Wh<strong>at</strong> is the general rule for figuring out where the minus signs come in<br />

when taking products of partial deriv<strong>at</strong>ives of many variables.<br />

• Does the rel<strong>at</strong>ion Cp<br />

CV<br />

= κT<br />

κS<br />

hold <strong>at</strong> very low temper<strong>at</strong>ures?<br />

• Why are all the other intensive-extensive variable pairs called work yet<br />

the temper<strong>at</strong>ure - entropy pair referred to as he<strong>at</strong> exchange.<br />

• For the inequality derived before 2.75, of the form ax 2 + bx + c ≥ 0, would<br />

imaginary roots have any physical meaning?<br />

• Can we say anything about the value of µ in the limit N<br />

V<br />

• Wh<strong>at</strong> is energy?<br />

→ 0?<br />

• How can you tell which thermodynamic potential is appropri<strong>at</strong>e?<br />

• In the first case (no friction) on page 52, how long will the oscill<strong>at</strong>ions<br />

last? Why?<br />

• Why is he<strong>at</strong> always gener<strong>at</strong>ed in an irreversible process?<br />

• Why are Legendre transform<strong>at</strong>ions useful?<br />

• Prove th<strong>at</strong> the Gibbs potential is a st<strong>at</strong>e function and th<strong>at</strong> its differential<br />

is exact.<br />

• How did the grand potential get its name?<br />

A.3 Questions for chapter 3.<br />

• Is energy the only factor to determine the phase st<strong>at</strong>e, wh<strong>at</strong> about other<br />

quantum factors?<br />

• How many phase transitions do we have?<br />

• Can the special rel<strong>at</strong>ivity and time play any role here?<br />

• Why is figure 3.13 impossible?<br />

• Can we give the phase transition an analysis based on chaos?<br />

• For the van der Waals equ<strong>at</strong>ion, if b=0 and T=an/R then p=0, why?<br />

• Why does the integral 3.34 yield A1-A2 ?<br />

• How will the van der Waals equ<strong>at</strong>ion be changed if we include magnetiz<strong>at</strong>ion?<br />

• Do we have any universal constants associ<strong>at</strong>ed with the van der Waals<br />

equ<strong>at</strong>ion?


288 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Why is the volume of the ice bigger than th<strong>at</strong> of w<strong>at</strong>er?<br />

• Are there any m<strong>at</strong>erials where the phase boundary for solid to liquid has<br />

a neg<strong>at</strong>ive slope?<br />

• Are there other systems where the energy of interest is not the Gibbs<br />

energy?<br />

• Why does the phase transition in Gibbs free energy always jump down?<br />

• Using stability and the model phase diagram in p − V space, can you<br />

explain why the triple point is where it is?<br />

• If we made a three dimensional graph of G(p,T) for 2 phases wh<strong>at</strong> would<br />

we see?<br />

• Can the Van Der Waals equ<strong>at</strong>ion be derived from a virial expansion?<br />

• Why look <strong>at</strong> the asymptotic forms?<br />

• Wh<strong>at</strong> do neg<strong>at</strong>ive pressures in the Van Der Waals equ<strong>at</strong>ion mean?<br />

• How important is scaling?<br />

• Why does crossing a phase boundary cause a jump in the value of the<br />

volume?<br />

• Do you think th<strong>at</strong> the phase boundary between solid and liquid will vanish<br />

if we can reach a high enough pressure?<br />

• Why is it often better to go around the critical point instead of crossing<br />

through the phase boundary?<br />

• Why do inaccessible regions only show up in phase diagrams when extensive<br />

variables are involved?<br />

• Is there anything besides the indices of refraction th<strong>at</strong> we can use to<br />

distinguish a liquid from a gas?<br />

• How are the forbidden regions of a V-T diagram rel<strong>at</strong>ed to the corresponding<br />

p-T graph?<br />

• Why do systems in forbidden regions phase separ<strong>at</strong>e?<br />

• Why is the eutectic point important?<br />

• Why are phase transitions important to thermodynamics?<br />

• Why is it important to include entropy as one of the st<strong>at</strong>e variables in our<br />

model system?<br />

• Wh<strong>at</strong> importance does the triple point have in industry?


A.3. QUESTIONS FOR CHAPTER 3. 289<br />

• How would we find Vc and Tc for w<strong>at</strong>er?<br />

• Why do we need weights in the equ<strong>at</strong>ion ∑ wiNi = N on page 122?<br />

• Why are there regions in which the cBT diagram which are not allowed?<br />

• Why is the ideal gas law an approxim<strong>at</strong>ion?<br />

• Why are some of the intermedi<strong>at</strong>e st<strong>at</strong>es unstable?<br />

• Wh<strong>at</strong> is the difference between mechanically stable and thermodynamically<br />

stable st<strong>at</strong>es?<br />

• Can w<strong>at</strong>er be he<strong>at</strong>ed over 100 degrees Celsius?<br />

• For p-V-T systems not in equilibrium, why can equ<strong>at</strong>ions of st<strong>at</strong>e like<br />

p=f(V,T) not exist?<br />

• In the p-T diagram for solid Ce in 3.2 with axis interchanged, does the fact<br />

th<strong>at</strong> this curve is not concave downwards contradict the ’consequences for<br />

second order deriv<strong>at</strong>ives’ discussed in chapter 2?<br />

• Where would a plasma ’fit in’ on a phase diagram like fig 3.1?<br />

• If a system were prepared <strong>at</strong> a given Magnetiz<strong>at</strong>ion M0 <strong>at</strong> high temper<strong>at</strong>ure<br />

and then cooled to a temper<strong>at</strong>ure which will force the st<strong>at</strong>e of the<br />

system into an inaccessible region, will the system phase separ<strong>at</strong>e?<br />

• If so, will the magnetiz<strong>at</strong>ion be in different directions, have different magnitudes,<br />

or both.<br />

• In a system with 4 independent st<strong>at</strong>e variables, wh<strong>at</strong> is the maximum<br />

number of phases th<strong>at</strong> can coexist?<br />

• We see th<strong>at</strong> neg<strong>at</strong>ive pressures are allowed by the Van der Waals equ<strong>at</strong>ion<br />

of st<strong>at</strong>e. Are neg<strong>at</strong>ive temper<strong>at</strong>ures also allowed?<br />

• During the binodal decomposition, why are time scales important?<br />

• In trying to achieve stability, does a system necessarily become mechanically<br />

stable first, then thermodynamically stable, or can these two processes<br />

occur simultaneously, incrementally.<br />

• Various groups of m<strong>at</strong>erials have common molecular sizes, interaction energies<br />

etc. within their groups. Are there various, commonly used, virial<br />

expansions for each such group?<br />

• When mixing two m<strong>at</strong>erials, we get a binary phase diagram. Wh<strong>at</strong> would<br />

the diagram look like if we were to mix three or more m<strong>at</strong>erials?<br />

• Why use the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e when it is limited in the<br />

systems th<strong>at</strong> it can describe?


290 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Are there systems with only meta-stable st<strong>at</strong>es?<br />

• Theoretically, we can have neg<strong>at</strong>ive pressure, why can’t we measure it?<br />

• Can we make a m<strong>at</strong>erial th<strong>at</strong> exists in an unstable region, but will not<br />

easily break down?<br />

• Why doesn’t neg<strong>at</strong>ive pressure correspond to a physical st<strong>at</strong>e?<br />

• When are non-unique energy solutions better <strong>at</strong> describing a system?<br />

• Can pressure be multi-valued for a single temper<strong>at</strong>ure?<br />

• Can temper<strong>at</strong>ure be multi-valued for single values of volume?<br />

• Can volume be multi-valued for single values of temper<strong>at</strong>ure?<br />

A.4 Questions for chapter 4.<br />

• Is there really no long-range order in a liquid?<br />

• Can we have spontaneous broken symmetries in a non-degener<strong>at</strong>e system?<br />

• Landau’s theory does not converge <strong>at</strong> Tc, but can’t we modify the theory<br />

so th<strong>at</strong> it does converge?<br />

• Can there be third order phase transitions, or is th<strong>at</strong> forbidden by the free<br />

parameters?<br />

• Can both he<strong>at</strong> capacities be modelled by the same response?<br />

• Wh<strong>at</strong> happens when there is a positive term in the fourth order sign of<br />

the expansion?<br />

• Is there a place where all three roots will have the same value of f for the<br />

same m?<br />

• Can we make our expansions work to get rid of anisotropies in the system?<br />

• How important are inhomogeneities in a system?<br />

• If the correl<strong>at</strong>ion length diverges <strong>at</strong> Tc, why is it so important in the<br />

theory?<br />

• Why do we need a non-equilibrium theory to describe the effects of fluctu<strong>at</strong>ions<br />

in a system?<br />

• How is snow different from ice?<br />

• Why must an order parameter be extensive?<br />

• Why do we measure the values of an order parameter via the associ<strong>at</strong>ed<br />

generalized force instead of measuring it directly?


A.4. QUESTIONS FOR CHAPTER 4. 291<br />

• Why do we compare free energies to determine if the st<strong>at</strong>e is an equilibrium<br />

st<strong>at</strong>e?<br />

• Prove th<strong>at</strong> there are no singularities in st<strong>at</strong>e functions for finite systems.<br />

• Where do you see hysteresis in E-and-M?<br />

• Why are there large fluctu<strong>at</strong>ions in the order parameter near the phase<br />

transition?<br />

• Why is it so hard to describe the system <strong>at</strong> the critical temp?<br />

• Why do small changes affect the whole system <strong>at</strong> the critical temp?<br />

• Can we describe phase transition by complex functions?<br />

• Can we describe phase transition by group theory?<br />

• Wh<strong>at</strong> is the effect of pressure on Landau form of free energy?<br />

• Is Landau theory equivalent to a unified theory in particle physics?<br />

• Is there any theory th<strong>at</strong> describes exactly wh<strong>at</strong>’s happening <strong>at</strong> a phase<br />

transition?<br />

• Is it possible to define b(T) in such away to have the correct exponent in<br />

(4.12)?<br />

• Wh<strong>at</strong>’s the meaning of nucle<strong>at</strong>ion and growth?<br />

• Do you know any real system with first order transition?<br />

• Is the description for fluctu<strong>at</strong>ions for macroscopic systems rel<strong>at</strong>ed to the<br />

uncertainty principle for microscopic systems?<br />

• Is the correl<strong>at</strong>ion length rel<strong>at</strong>ed to the susceptibility?<br />

• By wh<strong>at</strong> mechanism does a tap on a beaker of supercooled w<strong>at</strong>er cause it<br />

to suddenly crystalize?<br />

• Is it possible to apply our external field over such a large time scale th<strong>at</strong><br />

we can ignore fluctu<strong>at</strong>ions even near a phase transition?<br />

• Can we use Cauchy’s 1st theorem to evalu<strong>at</strong>e our Landau function <strong>at</strong> the<br />

critical point?<br />

• Is our ”Divide and conquer” technique for considering fluctu<strong>at</strong>ions analogous<br />

to our ”equivalent surface and volume charges” for considering dielectric<br />

samples?<br />

• Wh<strong>at</strong> physical consequences would there be if our ”f” in eq’n 4.37 were<br />

neg<strong>at</strong>ive, ie. fluctu<strong>at</strong>ions lowered our energy st<strong>at</strong>e?


292 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Wh<strong>at</strong> is our sigma*dm equivalent to supercooling and superconductivity?<br />

• Would analytic continu<strong>at</strong>ion help the evalu<strong>at</strong>ion of our Landau expansion<br />

<strong>at</strong> the critical point?<br />

• For the change in local magnetic field causing a perturb<strong>at</strong>ion in the magnetiz<strong>at</strong>ion<br />

everywhere; does this perturb<strong>at</strong>ion happen instantaneously or<br />

does it travel <strong>at</strong> the speed of light?<br />

• Does causality play a role in this perturb<strong>at</strong>ion?<br />

• At r=0 it is reasonable to speak of E=-gradient(Potential), why then is<br />

an expansion of free energy in terms of gradients of the order parameter<br />

not valid near Tc.<br />

• Why do experiments give different results than the Van Der Waals model<br />

<strong>at</strong> the phase transition?<br />

• Wh<strong>at</strong> would be a suitable order parameter for describing the phase transition<br />

between snow and ice?<br />

• Why is it important to compare the idea of a critical point in thermodynamics<br />

to the idea of a critical point in m<strong>at</strong>hem<strong>at</strong>ics?<br />

• Wh<strong>at</strong> kind of expansion would be needed to describe f exactly?<br />

• Would this analysis be appropri<strong>at</strong>e for approxim<strong>at</strong>ing the n<strong>at</strong>ure of a bar<br />

magnet?<br />

• Is the transition from liquid w<strong>at</strong>er to ice a second or first order phase<br />

transition?<br />

• At which point on the hysteresis will a butterfly flapping its wings 1000<br />

miles away cause the system to go from a meta-stable st<strong>at</strong>e to a stable<br />

st<strong>at</strong>e?<br />

• Wh<strong>at</strong> m<strong>at</strong>hem<strong>at</strong>ical oper<strong>at</strong>ions were used to derive (4.38)?<br />

• Why does rot<strong>at</strong>ional symmetry cause the solutions to be of the form<br />

ul(r)Ylm(ω)/r?<br />

• Wh<strong>at</strong> parameters would tell us whether a system is ice or snow?<br />

A.5 Questions for chapter 5.<br />

• Why in deriving the pair correl<strong>at</strong>ion function we assume a spherical symmetry?<br />

• Can we replace h and n by one critical exponent?


A.5. QUESTIONS FOR CHAPTER 5. 293<br />

• Is the law described by the correl<strong>at</strong>ion function obeying the third Newton<br />

law?<br />

• Wh<strong>at</strong> is the physical meaning of the fact th<strong>at</strong> the susceptibility is proportional<br />

to the q=0 component of the Fourier transform of the pair correl<strong>at</strong>ion<br />

function?<br />

• Are in reality the dependence of the response functions with respect to T<br />

is the same above and below the critical temper<strong>at</strong>ure<br />

• Wh<strong>at</strong> is the critical field?<br />

• Is there a mean field theory associ<strong>at</strong>ed with the first phase transition?<br />

• How can we be sure th<strong>at</strong> it is enough to tre<strong>at</strong> the system near the critical<br />

temper<strong>at</strong>ure in a perturb<strong>at</strong>ion theory?<br />

• How can Ginzburg Landau theory give valid measurements while wrong<br />

critical exponents?<br />

• Wh<strong>at</strong> is the maximum number of critical exponents?<br />

• The critical exponents shown have simple rel<strong>at</strong>ions to T and Tc. How<br />

can this be possible, when we know a st<strong>at</strong>e is dependent on <strong>at</strong> least two<br />

variables.<br />

• I have seen the reduced p,T,V written as t=T/Tc, etc. We do not do th<strong>at</strong><br />

in this chapter. Which is the correct version.<br />

• Since st<strong>at</strong>istical mechanics can get us to the heart of phase transitions,<br />

why bother doing this with thermodynamics.<br />

• How many other critical exponents are there.<br />

• Is there a simpler way to find our descriptions other than scaling.<br />

• How do we rel<strong>at</strong>e the functionals th<strong>at</strong> come out of scaling to experimental<br />

d<strong>at</strong>a.<br />

• Critical exponents can be hard to get out of st<strong>at</strong>e equ<strong>at</strong>ions, and they do<br />

not seem to give any new inform<strong>at</strong>ion. Why use them.<br />

• How can critical exponents get you back to the phase diagram. I just<br />

don’t see the connection.<br />

• If the exponents are the same for T¿Tc or T¡Tc, why do we bother to<br />

define them separ<strong>at</strong>ely.<br />

• Which is the most powerful and useful critical exponent.<br />

• Why would gamma not equal gamma prime?


294 APPENDIX A. QUESTIONS SUBMITTED BY STUDENTS.<br />

• Wh<strong>at</strong> would happen if Tc and pc could be neg<strong>at</strong>ive?<br />

• Why must delta M ¡¡ M?<br />

• How can you have a 2 dimensional system?<br />

• Why must a realistic description of a phase transition near a critical point<br />

include the effects of fluctu<strong>at</strong>ions?<br />

• Why does the correl<strong>at</strong>ion length become very large near a critical point?<br />

• Where else do we see scaling and universality?<br />

• Why are critical exponents so difficult to measure?<br />

• How can you measure a pair correl<strong>at</strong>ion function?<br />

• Why can we use any function of two variables as a basis for scaling?<br />

• Why do fluctu<strong>at</strong>ions occur near the critical point?<br />

• Would an engineer want to know the critical exponents of a steam engine?<br />

• Why is it important to know the mean field approach?<br />

• Wh<strong>at</strong> is the usefulness of using rel<strong>at</strong>ive variables?<br />

• Is the scaling function a m<strong>at</strong>hem<strong>at</strong>ical convenience or does it have physical<br />

significance?<br />

• If critical exponents are so hard to measure, wh<strong>at</strong> is their usefulness?<br />

• Could we have described the change in magnetiz<strong>at</strong>ion over time by something<br />

other than the diffusion equ<strong>at</strong>ion?<br />

• Would it be plausible to use a rel<strong>at</strong>ivistic form of the equ<strong>at</strong>ions?<br />

• Why does the scaling theory give nu=nu’?<br />

• Wh<strong>at</strong> would happen if we used the compressibility as the response function?


Appendix B<br />

Summaries submitted by<br />

students.<br />

Students were asked to submit a paragraph describing each chapter, written<br />

in a way they would like to see it included in an introductory description of<br />

the notes. Therefore these paragraphs reflect both the understanding of the<br />

m<strong>at</strong>erial and the exposition of the m<strong>at</strong>erial.<br />

B.1 Summaries for chapter 1.<br />

In thermodynamics one describes a macroscopic system, with many degrees of<br />

freedom, by st<strong>at</strong>e variables (which are time-averaged). For st<strong>at</strong>e variables to be<br />

useful in predicting intermedi<strong>at</strong>e values, thermodynamics concerns itself only<br />

with reversible processes in thermodynamic equilibrium in the thermodynamic<br />

limit. St<strong>at</strong>e variables are governed by the laws of thermodynamics. Laws zero<br />

and three give us fixed references. Law one is the conserv<strong>at</strong>ion of energy and<br />

law two tells us th<strong>at</strong> if all the processes are reversible we have the best scenario,<br />

otherwise entropy increases.<br />

<strong>Thermodynamics</strong> is the study of macroscopic systems in equilibrium. We use<br />

thermodynamics to help define experimental variables such as temper<strong>at</strong>ure and<br />

pressure. Using thermodynamics we can set up simple experiments to test our<br />

theories, and we can use our experimental results to find simple rel<strong>at</strong>ionships<br />

for our variables.<br />

Chapter 1 introduces extensive and intensive thermodynamical variables, talks<br />

about wh<strong>at</strong> they mean, talks about reversibility and irreversibility, and how<br />

quasi-st<strong>at</strong>ic systems work in thermodynamics. Chapter 1 also goes over the four<br />

laws of thermodynamics, and introduces some m<strong>at</strong>hem<strong>at</strong>ical formalism. The<br />

end of chapter 1 talks about entropy in the context of the third law. There are<br />

295


296 APPENDIX B. SUMMARIES SUBMITTED BY STUDENTS.<br />

also definitions of wh<strong>at</strong> we mean by macroscopic systems, st<strong>at</strong>e, thermodynamic<br />

equilibrium, and other thermodynamic vocabulary terms.<br />

Rules, rules, rules. Everything in life seems to be governed by rules. The best<br />

way to understand any system, from a society to a steam engine, is by looking<br />

<strong>at</strong> the laws th<strong>at</strong> govern its behavior. In this chapter we will learn the language<br />

of thermodynamics and study the laws th<strong>at</strong> apply to thermodynamic systems.<br />

The first law included he<strong>at</strong> as a form of energy and st<strong>at</strong>es th<strong>at</strong> the total energy<br />

of a system is conserved. The second law puts restrictions on the transfer of<br />

he<strong>at</strong> or the exchange of he<strong>at</strong> into work. It defines a preferred direction of time<br />

and tells us th<strong>at</strong> no engine can be more efficient than a Carnot engine. The<br />

third law says th<strong>at</strong>, in the thermodynamic limit, when the temper<strong>at</strong>ure goes to<br />

zero so does the entropy. It is thus impossible to reach T=0K using a reversible<br />

process in a finite number of steps. Finally, the zeroth law, which is so trivial<br />

it is often overlooked, st<strong>at</strong>es th<strong>at</strong> if two systems are in equilibrium with the<br />

same system, they are also in equilibrium with each other. This simple fact is<br />

wh<strong>at</strong> allows us to define universal standards for temper<strong>at</strong>ure, pressure, and all<br />

the other descriptors of our system. With these laws in hand we can begin to<br />

understand the behavior of thermodynamic systems throughout the land.<br />

This chapter will quickly review the 4 basic laws of thermodynamics with an<br />

emphasis on how the laws affect the real wold measurements. It will use the basic<br />

concepts of the Carnot engine to illustr<strong>at</strong>e how one might go about developing<br />

such measurements. The significance, limits, and possibilities will be touched<br />

upon as the framework is laid down.<br />

Reversible processes are important. There are four laws of thermodynamics,<br />

numbered 0 through 3. These laws hold for all systems. The Carnot engine is<br />

the most efficient engine there is. Entropy is an important quantity which is<br />

rel<strong>at</strong>ed to he<strong>at</strong> flow. You can only define entropy well in the thermodynamic<br />

limit. A Carnot engine is a good way to do work and to measure temper<strong>at</strong>ure.<br />

In thermodynamics we consider a macroscopic system and our variables are<br />

global averages. The thermodynamic variables th<strong>at</strong> uniquely determine the<br />

thermodynamic st<strong>at</strong>e are called st<strong>at</strong>e variables. To rel<strong>at</strong>e the different variables<br />

to each other we use two kinds of rel<strong>at</strong>ions. The first are the thermodynamic<br />

laws, which hold for all systems. The zeroth law discusses the conditions for<br />

thermal equilibrium. The first law is a st<strong>at</strong>ement about the conserv<strong>at</strong>ion of<br />

energy. The second law is a st<strong>at</strong>ement about maximal entropy in equilibrium.<br />

The third law is putting the entropy to be zero <strong>at</strong> zero absolute temper<strong>at</strong>ure.<br />

The second type of rel<strong>at</strong>ions are the equ<strong>at</strong>ions of st<strong>at</strong>e, which depend on our<br />

specific system. The equ<strong>at</strong>ions of st<strong>at</strong>e can be used only when the system is<br />

in equilibrium. For this point it is important to define two types of processes,


B.2. SUMMARIES FOR CHAPTER 2. 297<br />

reversible and irreversible.<br />

B.2 Summaries for chapter 2.<br />

Energy is stored in a gas. There are many different free energies because there<br />

are many different variables involved in a process. Response functions such as<br />

he<strong>at</strong> capacities are the most important way to obtain inform<strong>at</strong>ion. Thermodynamic<br />

processes have many variables involved. Hence, the calculus of many<br />

variables is used. When two systems come into contact, they are able to exchange<br />

energy by exchanging their extensive variables. Fluctu<strong>at</strong>ions play a role<br />

in m<strong>at</strong>erials.<br />

Chapter two introduces the minimum energy principle and discusses the consequences<br />

of things changing. The internal energy of a closed system is a minimum<br />

as a function of all parameters th<strong>at</strong> are not fixed by external constraints. Internal<br />

energy also gives an upper bound for the amount of work th<strong>at</strong> can be<br />

done in an irreversible process. One way to see the consequences of changing<br />

variables is via free energy. There are many different constructions for free energies,<br />

such as the Helmholtz free energy, the Gibbs potential, enthalpy, and<br />

the grand potential. Each of these formul<strong>at</strong>ions is useful in different situ<strong>at</strong>ions,<br />

depending on the system we are describing and the set of n<strong>at</strong>ural variables used<br />

with th<strong>at</strong> system. Once you decide which set of variables to use, you can use<br />

response functions, such as he<strong>at</strong> capacities, compressibilities, and the coefficient<br />

of thermal expansion to describe the effect of allowing your variables to<br />

change. Response functions help us describe the rel<strong>at</strong>ionships between different<br />

thermodynamic variables and tell us a lot about how a system behaves.<br />

The amount of energy which can be released by a gas depends on the process by<br />

which the gas is doing work. When trying to measure this energy one needs to<br />

decide which variables can, and should be controlled and which ones should be<br />

kept constant (eg. Volume may be difficult to measure and Entropy impossible<br />

to measure directly, so we may want to keep these variables constant in our<br />

measurements). Various ”Free Energies” then help us to express the amount of<br />

work a system can do with our desired variables held constant. By definition,<br />

these Free Energy st<strong>at</strong>e functions have exact differentials, and as a result, lead<br />

to equivalence rel<strong>at</strong>ions between various pairs of partial deriv<strong>at</strong>ives of many<br />

st<strong>at</strong>e variables with respect to other st<strong>at</strong>e variables (Maxwell Rel<strong>at</strong>ions). At<br />

equilibrium there is minimum energy and maximum entropy with respect to the<br />

concerned variables. This fact also leads to conditions on st<strong>at</strong>e variables. All<br />

intensive variables are the same for an isol<strong>at</strong>ed system in thermal equilibrium.<br />

With Newtonian Mechanics we use the Lagrangian and Hamiltonian to get the<br />

equ<strong>at</strong>ions of motion. With the Equ<strong>at</strong>ions of motion we have a fundamental


298 APPENDIX B. SUMMARIES SUBMITTED BY STUDENTS.<br />

understanding of the system th<strong>at</strong> we are working with. In <strong>Thermodynamics</strong> it<br />

is the Energy th<strong>at</strong> we use to get a fundamental understanding of the system.<br />

The internal energy is the basic form of energy th<strong>at</strong> we use. If we begin with<br />

the internal energy U, we can do successive transforms, or use the Maxwell<br />

rel<strong>at</strong>ions to find properties of the parameters th<strong>at</strong> define the system. Using this<br />

method we can arrive <strong>at</strong> the equ<strong>at</strong>ion of st<strong>at</strong>e for the system, which will define<br />

the system. We can also use this to find other measurable parameters, such as<br />

the compressibilities and capacitances.<br />

We will start the chapter by introducing the minimum internal energy and maximum<br />

entropy principles. Next we will generalize the minimum energy principle<br />

to other kinds of thermodynamic potentials (free energies). We will see th<strong>at</strong> the<br />

thermodynamic potentials th<strong>at</strong> describe our process depend on the variables<br />

under control. For each type of process we define the appropri<strong>at</strong>e thermodynamic<br />

potential, among those we will define the Internal Energy (U),Helmholtz<br />

(F),Gibbs (G),Entropy (H) and Grand Potential (Ω). Applying a m<strong>at</strong>hem<strong>at</strong>ical<br />

rules to derive the Maxwell rel<strong>at</strong>ions which connect changes in different st<strong>at</strong>e<br />

variables in different processes. Response functions like he<strong>at</strong> capacity, compressibility,<br />

and coefficient of thermal expansion, help us to obtain inform<strong>at</strong>ion<br />

about the thermodynamic system. Again we will use m<strong>at</strong>hem<strong>at</strong>ical concepts<br />

(partial deriv<strong>at</strong>ives ) to find rel<strong>at</strong>ions between the response functions and with<br />

the help of the minimum energy condition of equilibrium we will derive the<br />

inequalities for the he<strong>at</strong> capacities and compressibilities.Then we will look <strong>at</strong><br />

the stability requirements on thermodynamic functions other than the internal<br />

energy. When it comes to magnetic system we must be careful in our analysis<br />

and include the work term comes from the field cre<strong>at</strong>ed in the m<strong>at</strong>erial. Finally<br />

we will briefly discuss a small fluctu<strong>at</strong>ion in a thermodynamic limit and derive<br />

a very important probability rule.<br />

This chapter begins by introducing the minimum energy principle. This is used<br />

to lead into the different free energies and their definitions. Now with the five<br />

free energies defined along with the corresponding differential equ<strong>at</strong>ions, their<br />

linear n<strong>at</strong>ures (Gibbs-Duhem rel<strong>at</strong>ions) are explored. These linear rel<strong>at</strong>ions lead<br />

to several important properties such as dependence of some intensive variables<br />

on other variables. Maxwell rel<strong>at</strong>ions are introduced. Next, the basic m<strong>at</strong>erial<br />

properties such as he<strong>at</strong> capacity or compressibility are defined. A m<strong>at</strong>h<br />

review of multi-variable calculus is added to the list of tools next. Finally, with<br />

these tools, various rel<strong>at</strong>ionships between m<strong>at</strong>erial properties and solutions to<br />

some problems are examined. During the applic<strong>at</strong>ions to solutions a perturb<strong>at</strong>ional<br />

approach involving the role of fluctu<strong>at</strong>ions in m<strong>at</strong>erials is introduced.<br />

A large number of tools are assumed understood, such as the Gaussian n<strong>at</strong>ure<br />

of Maxwellian functions. After this chapter, there are a large number of problems<br />

solvable in thermodynamics with the tools presented. The main body of<br />

thermodynamics is presented here.


B.3. SUMMARIES FOR CHAPTER 3. 299<br />

B.3 Summaries for chapter 3.<br />

Discontinuities in equ<strong>at</strong>ions of st<strong>at</strong>e manifest themselves in the real world as<br />

phase transitions. Phase transitions enable us to tell when a m<strong>at</strong>erial becomes<br />

superconductive, or when it’s magnetic properties suddenly change, or where<br />

general ”jumps” occur in values of other extensive st<strong>at</strong>e variables. In a simple<br />

p-V-T system, the ideal gas law (PV=nRT) doesn’t help us to explain phase<br />

transitions, so we resort to virial expansions which take interaction between<br />

molecules; molecules actually taking up volume etc., into consider<strong>at</strong>ion. Van<br />

der Waals postul<strong>at</strong>ed one such ”not-so-ideal gas” equ<strong>at</strong>ion, valid for both liquid<br />

and gas phase. Stability requirements (partial (p) w.r.t. (v) ¡=0) and the<br />

minimum energy principle, help us to determine which values of p, V and T<br />

are allowable, from our <strong>at</strong> first seemingly multi-valued Van der Waals equ<strong>at</strong>ion<br />

solution.<br />

Chapter 3 addresses the issue of phase transitions, investig<strong>at</strong>ing discontinuities<br />

in the equ<strong>at</strong>ions of st<strong>at</strong>e. This is best seen by examining phase diagrams. The<br />

lines marked on the p-V diagrams represent the coexistence curves where a<br />

system can be in more than one st<strong>at</strong>e <strong>at</strong> a given pressure and temper<strong>at</strong>ure.<br />

When two such lines meet, we have the triple point, the p and T values <strong>at</strong><br />

which all 3 phases coexist. For multi-phase diagrams, the maximum number of<br />

phases which can coexist is one more than the number of independent intensive<br />

st<strong>at</strong>e variables needed to describe the system. Crossing a phase boundary leads<br />

to a change in st<strong>at</strong>e. This change causes a jump in the volume and the other<br />

extensive variables, while the energy (but not its deriv<strong>at</strong>ive) remains continuous.<br />

If a system is in a forbidden region of a phase diagram involving an extensive<br />

variable, it will phase separ<strong>at</strong>e. The Clausius-Clapeyron rel<strong>at</strong>ion tells us the<br />

slope of the coexistence curve in a p-V diagram (which is always positive) and<br />

can be expressed in terms of l<strong>at</strong>ent he<strong>at</strong>. The jump in volume of a phase<br />

transition is also seen in the van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e in which V is not<br />

a unique function of p and T. However, the van der Waals equ<strong>at</strong>ion allows for<br />

st<strong>at</strong>es with neg<strong>at</strong>ive pressures, which are never stable, so this equ<strong>at</strong>ion is only<br />

a good model for a liquid to gas phase transition, ending in a critical point.<br />

In this chapter phase transitions are explored. The boundary is defined. The rel<strong>at</strong>ionship<br />

<strong>at</strong> the boundary is m<strong>at</strong>hem<strong>at</strong>ically modelled in the Clausius-Clapeyron<br />

rel<strong>at</strong>ionship. L<strong>at</strong>ent he<strong>at</strong>s are defined in the context of this rel<strong>at</strong>ion. Further,<br />

the limit<strong>at</strong>ions on the possible coexistent phases are informally defined. Some<br />

applic<strong>at</strong>ions are described during the rest of the chapter while introducing additional<br />

concepts. Binary systems with chemical reactions are explored. The<br />

Van Der Waals equ<strong>at</strong>ion is next. The Virial expansion is introduced. The rest<br />

of the chapter deals with the behavior and validity conditions of the Van Der<br />

Waals equ<strong>at</strong>ion. The asymptotic form is explored to achieve insight.


300 APPENDIX B. SUMMARIES SUBMITTED BY STUDENTS.<br />

Phase transitions are discontinuities in the equ<strong>at</strong>ions of st<strong>at</strong>e. In the p versus<br />

T diagram, all three phases(solid, liquid, and gas) can coexist <strong>at</strong> the triple<br />

point. Also, the phase boundary between liquid and gas vanishes <strong>at</strong> the critical<br />

point. Two distinct phases can coexist <strong>at</strong> the phase boundary. We find a phase<br />

boundary by looking <strong>at</strong> the Gibbs free energy for two phases. The number of<br />

phases th<strong>at</strong> can coexist depends on the number of independent st<strong>at</strong>e variables.<br />

The Van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is a better approxim<strong>at</strong>ion than the ideal<br />

gas law. A line which separ<strong>at</strong>es mechanically stable st<strong>at</strong>es from mechanically<br />

unstable st<strong>at</strong>es is a spinodal. A line th<strong>at</strong> separ<strong>at</strong>es a thermodynamically metastable<br />

st<strong>at</strong>e with a stable st<strong>at</strong>e is a binodal.<br />

In this chapter we will analyze the phase transition phenomena from a macroscopic<br />

point of view. Looking <strong>at</strong> a P-V-T system, we will first analyze the P-T<br />

phase diagram and describe the triple point and the critical point. Then we<br />

will analyze the V-T diagram which (like P-V diagram) characterized by the<br />

forbidden regions. Gibbs free energy analysis yields the Clausius-Clapeyron rel<strong>at</strong>ion,<br />

which gives a quantit<strong>at</strong>ive description of the phase boundaries. Next we<br />

will generalize our conclusions to a multi phase boundaries and give the binary<br />

phase diagrams as an applic<strong>at</strong>ion. Here the point where all three phases can<br />

coexist is called eutectic point. To give a more quantit<strong>at</strong>ive description of the<br />

phase transition we need to know the equ<strong>at</strong>ion of st<strong>at</strong>e. As a first try we will<br />

express it in a virial expansion form. The Van Der Waals equ<strong>at</strong>ion of st<strong>at</strong>e is a<br />

model which is valid for the change from the gas to the liquid phase. Here we<br />

will give a quantit<strong>at</strong>ive analysis for such important model.<br />

In chapter 3 we looked <strong>at</strong> the different ways to represent the phases of m<strong>at</strong>erials.<br />

We saw th<strong>at</strong> simple m<strong>at</strong>erials can have complex even convoluted phase diagrams.<br />

Simple binary m<strong>at</strong>erials can have highly unstable regions, and there are places<br />

in all phases diagrams where m<strong>at</strong>erials can simply not exist. We saw th<strong>at</strong> there<br />

are co-existence curves for all m<strong>at</strong>erials where we can have equal amounts of<br />

both phases. There are triple points in all phase diagrams where we can have<br />

solid-liquid-gas co-existing <strong>at</strong> the same time. There are however, no points<br />

on any phase diagrams where we can get four phases coming together <strong>at</strong> the<br />

same time. Phase diagrams can be simple or complex, but they give a good<br />

description of the system.<br />

B.4 Summaries for chapter 4.<br />

We use order parameters, which are st<strong>at</strong>e variables, to distinguish among phases<br />

of a system. A useful way to study the free energy per unit volume of a magnetic<br />

system is to expand it in terms of the magnetiz<strong>at</strong>ion of the system. Landau studied<br />

one such expansion, which only included even powers of the magnetiz<strong>at</strong>ion.<br />

This expansion, however, is not valid for when a system is <strong>at</strong> a phase transition.<br />

Ginzberg and Landau modified this expansion to include fluctu<strong>at</strong>ions. This


B.4. SUMMARIES FOR CHAPTER 4. 301<br />

modified form, however, is still incorrect precisely <strong>at</strong> the phase transition.<br />

Wh<strong>at</strong> happens <strong>at</strong> a phase transition? All the techniques and theories discussed<br />

so far enable us to model simple equilibrium systems. But wh<strong>at</strong> about the<br />

fluctu<strong>at</strong>ions and discontinuities <strong>at</strong> and around phase transitions? Chapter 4<br />

presents a good framework to analyze first and second order transitions with the<br />

Landau-Ginzburg theory. In order to examine these transitions, we must first<br />

determine an order parameter, an extensive st<strong>at</strong>e variable th<strong>at</strong> uniquely defines<br />

which phase the system is in. To measure the value of the order parameter<br />

we need to look <strong>at</strong> the generalized force associ<strong>at</strong>ed with it. Landau’s method<br />

describes phase transitions by parameterizing the free energy as a function of<br />

the order parameter. Because of the essential singularity <strong>at</strong> Tc, Landau’s power<br />

series fails to converge so it can’t explain wh<strong>at</strong> happens <strong>at</strong> the phase transition,<br />

but it is good for classifying kinds of phase transitions by comparing the st<strong>at</strong>es <strong>at</strong><br />

both sides of the transition. The order of a phase transition can be defined by its<br />

continuity. If it is possible to find a continuous p<strong>at</strong>h around a critical point from<br />

one phase to another (much like the p<strong>at</strong>h around an essential singularity with<br />

a branch cut) it is a second order transition. If there is no continuous p<strong>at</strong>h,<br />

it’s a first order transition. In first order transitions, the value of the order<br />

parameter is discontinuous <strong>at</strong> the critical temper<strong>at</strong>ure. If we go continuously<br />

through a co-existence curve, we end up in a meta-stable st<strong>at</strong>e. Near a phase<br />

transition, fluctu<strong>at</strong>ions become an issue. Ginzburg adapted Landau’s theory to<br />

better account for the effects of fluctu<strong>at</strong>ions by including sp<strong>at</strong>ial vari<strong>at</strong>ions in<br />

the order parameter. Ginzburg and Landau further improved the theory with<br />

an expansion of the free energy in terms of gradients of the order parameter.<br />

This provides a good description of the st<strong>at</strong>e of the system much closer to the<br />

critical temper<strong>at</strong>ure.<br />

We will start this chapter by defining the order parameter and how to measure<br />

it. Using a continuity concepts we will define the second order phase transition.<br />

Then we will show the similarity between the coexistence curve and the branch<br />

cut in a complex function. Giving an overview of a spontaneously broken symmetry<br />

before introducing the Landau theory of phase transition. Unlike the<br />

van der Waals equ<strong>at</strong>ion, the Landau theory of phase transition is more general<br />

one. Here we will analysis the Landau form of the free energy for the second<br />

and first order phase transition, and describe the behavior of the entropy, he<strong>at</strong><br />

capacity and susceptibility near the critical temper<strong>at</strong>ure. Here we will give another<br />

definition of the phase transition order, based on the continuity of the<br />

order parameter <strong>at</strong> the critical temper<strong>at</strong>ure. Trying to give more accur<strong>at</strong>e description<br />

near the phase transition we will study the Ginzburg-Landau theory<br />

for fluctu<strong>at</strong>ions.<br />

In chapter 4 we began the analysis th<strong>at</strong> will lead us to critical phenomena. We<br />

noticed th<strong>at</strong> phase transitions could be modelled by order parameters. Order


302 APPENDIX B. SUMMARIES SUBMITTED BY STUDENTS.<br />

parameters are st<strong>at</strong>e variables th<strong>at</strong> give the critical phenomena of the system.<br />

In the case of chapter 4 we looked <strong>at</strong> magnetiz<strong>at</strong>ion using Landau-Ginzburg<br />

theory, but we know th<strong>at</strong> this model can be used for other spontaneous broken<br />

symmetries.<br />

To describe the phase of a system we need a parameter which: 1) differenti<strong>at</strong>es<br />

between st<strong>at</strong>es of a system, while 2) being measured does not change the system<br />

(requires zero field for measurement). Such a st<strong>at</strong>e variable is termed an order<br />

parameter. When an order parameter and its external field couple to zero,<br />

the free energy is usually symmetric as a function of order parameter; so when<br />

we have a nonzero order parameter, there are <strong>at</strong> least 2 st<strong>at</strong>es for the system<br />

with the same energy. The act of measurement breaks this symmetry. Landau<br />

devised a model to parameterize the free energy as a function of order parameter.<br />

With an expansion of the free energy w.r.t order parameter we can look <strong>at</strong><br />

minima and see which values correspond to stable st<strong>at</strong>es. At 1st order phase<br />

transitions, the Gibbs function is continuous, but deriv<strong>at</strong>ives (e.g. (g/T)p =-s<br />

and (g/p)T=v) are discontinuous. At 2nd order phase transitions, G and first<br />

deriv<strong>at</strong>ives are continuous, but second deriv<strong>at</strong>ives are discontinuous. We look<br />

for phase transitions <strong>at</strong> discontinuities. . Landau’s expansion doesn’t converge<br />

<strong>at</strong> the critical points (essential singularities). Fluctu<strong>at</strong>ions become important<br />

near phase transitions. Fluctu<strong>at</strong>ions in a system causes parts of a system to<br />

be stable and other parts meta-stable, causing phase separ<strong>at</strong>ion of the system.<br />

Ginzburg, tried to account for fluctu<strong>at</strong>ions by making the order parameter a<br />

function of position. His model is still not valid <strong>at</strong> the critical point, but allows<br />

us to analyze the system much closer to it.<br />

B.5 Summaries for chapter 5.<br />

Critical exponents show how fast response functions diverge near the critical<br />

point. The mean field approach ignores the effects arising from fluctu<strong>at</strong>ions in<br />

the magnetiz<strong>at</strong>ion. When a system is far away from the critical point, mean<br />

field theory is a good approxim<strong>at</strong>ion. It is not a good approxim<strong>at</strong>ion, however,<br />

close to the critical point. The theory of Ginzberg and Landau gives a better<br />

description of a system near the critical point. We use scaling, which has<br />

consequences.<br />

Wh<strong>at</strong> is happening near the critical point and how can we describe the thermodynamic<br />

st<strong>at</strong>e there? We start by defining the four common critical exponents<br />

and discuss the universality behind th<strong>at</strong>. Landau model for superconductivity<br />

will be a good example for second order phase transition. Then we discuss the<br />

mean field theory, derive the critical exponents for Landau theory, and show<br />

th<strong>at</strong> van der waals model is a mean field model. Near the critical point the<br />

systems’ behavior determined by the fluctu<strong>at</strong>ion term of the free energy, which<br />

should be homogenous. As a consequence of the scaling we reduce the number


B.5. SUMMARIES FOR CHAPTER 5. 303<br />

of independent critical exponents. Next we apply the scaling on the pair correl<strong>at</strong>ion<br />

function and define the hyper scaling. Then we explore the validity of<br />

Ginzburg Landau theory. We end the chapter by introducing the time scaling<br />

in the correl<strong>at</strong>ion function.<br />

Critical exponents help us describe different r<strong>at</strong>es of divergence of response<br />

functions, near the critical point. These critical exponents are a unique characteriz<strong>at</strong>ion<br />

of phase transitions - i.e. Similar critical exponents indic<strong>at</strong>e similar<br />

processes <strong>at</strong> phase transition. Mean field models produce critical exponents of<br />

- b=1/2; a=1; g=0. Van Der Waals used such a model in his equ<strong>at</strong>ion of st<strong>at</strong>e,<br />

and these critical exponents were verified. Mean field theory is most useful far<br />

away from the critical point. In scaling theory we describe a non-equilibrium<br />

st<strong>at</strong>e with fluctu<strong>at</strong>ions. Including gradient terms in the expansion of free energy,<br />

improves our model near the critical point. The pair correl<strong>at</strong>ion function<br />

describes the average of these fluctu<strong>at</strong>ions. Near the critical point only x is important.<br />

No m<strong>at</strong>ter wh<strong>at</strong> you do though, everything still diverges <strong>at</strong> the critical<br />

point and we generally need to throw this whole mean field idea out the window.<br />

Chapter 5 discusses the n<strong>at</strong>ure of phase transitions th<strong>at</strong> end in a critical point<br />

by focusing on the properties <strong>at</strong> the critical point (which also determine wh<strong>at</strong><br />

happens away from the critical point). In order to characterize the divergent<br />

behavior due to fluctu<strong>at</strong>ions near the critical point, critical exponents are introduced.<br />

They show the speed of the divergence. Critical exponents are a<br />

unique characteriz<strong>at</strong>ion of the physics of phase transitions and are used to classify<br />

groups of similar transitions with identical exponents. Mean field theory<br />

ignores the details of fluctu<strong>at</strong>ions by averaging them out. Any mean field theory<br />

will always yield the same critical exponents. MFT is a good first approxim<strong>at</strong>ion<br />

for many systems whose measured critical exponents are close to the universal<br />

exponents predicted by the theory. Once the critical temper<strong>at</strong>ure and pressure<br />

are known, the van der Waals equ<strong>at</strong>ion is unique. We can then use rel<strong>at</strong>ive<br />

variables to write a parameter-free version of this equ<strong>at</strong>ion to derive values of<br />

the critical exponents. If we change the values of the temper<strong>at</strong>ure of the field<br />

slightly, we see a scaled version of the original picture because the free energy due<br />

to fluctu<strong>at</strong>ions near a critical point is a homogeneous function of its variables.<br />

Near the critical point, the correl<strong>at</strong>ion length is the only important length scale,<br />

so the energy density exhibits hyper-scaling there. Mean field theory works far<br />

away from the critical point. Closer to this point, the model is improved by<br />

including gradient terms in the expansions of the free energy. Very close to the<br />

critical point, however, we need a different theory.


304 APPENDIX B. SUMMARIES SUBMITTED BY STUDENTS.


Appendix C<br />

Solutions to selected<br />

problems.<br />

C.1 Solutions for chapter 1.<br />

Problem 1.<br />

The general form is ¯ df = A(x, y)dx + B(x, y)dy and is an exact differential if<br />

and only if ( )<br />

∂A<br />

=<br />

∂y x<br />

Therefore we have<br />

Part 1: A(x, y) = x2 , B(x, y) = y2 and<br />

( )<br />

∂A<br />

∂y<br />

( )<br />

∂B<br />

∂x<br />

Hence an exact differential. Using<br />

we find<br />

f(x, y) = f(0, 0) +<br />

x<br />

y<br />

( )<br />

∂B<br />

∂x y<br />

= 0<br />

= 0<br />

∫ x<br />

A(x<br />

0<br />

′ , 0)dx ′ ∫ y<br />

+ B(x, y<br />

0<br />

′ )dy ′<br />

f(x, y) = f(0, 0) + 1<br />

3 x3 + 1<br />

3 y3<br />

Part 2: A(x, y) = y2 , B(x, y) = x2 and<br />

( )<br />

∂A<br />

∂y<br />

x<br />

305<br />

= 2y


306 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Hence NOT an exact differential.<br />

( )<br />

∂B<br />

= 2x<br />

∂x y<br />

Part 3: A(x, y) = 3x2y2 , B(x, y) = 2x3y + y4 and<br />

( )<br />

∂A<br />

= 6x<br />

∂y x<br />

2 y<br />

( )<br />

∂B<br />

= 6x<br />

∂x<br />

2 y<br />

Hence an exact differential. Using<br />

we find<br />

Problem 2.<br />

h(x, y) = h(0, 0) +<br />

y<br />

∫ x<br />

A(x<br />

0<br />

′ , 0)dx ′ ∫ y<br />

+ B(x, y<br />

0<br />

′ )dy ′<br />

h(x, y) = h(0, 0) + x 3 y 2 + 1<br />

5 y5<br />

¯dq = xydx + 1<br />

x dy<br />

With the test given in problem 1 we have A(x, y) = xy , B(x, y) = 1<br />

x and<br />

( )<br />

∂A<br />

= x<br />

∂y<br />

( ∂B<br />

∂x<br />

Hence NOT an exact differential. Now use:<br />

)<br />

y<br />

x<br />

= − 1<br />

x 2<br />

¯ds = f(x)xydx + f(x) 1<br />

x dy<br />

With the test given in problem 1 we have A(x, y) = f(x)xy , B(x, y) = f(x) 1<br />

x<br />

and ( )<br />

∂A<br />

= f(x)x<br />

∂y<br />

x<br />

( )<br />

∂B<br />

= −f(x)<br />

∂x y<br />

1<br />

x2 + f ′ (x) 1<br />

x<br />

Hence in order th<strong>at</strong> this is an exact differential we need<br />

f(x)x = −f(x) 1<br />

x 2 + f ′ (x) 1<br />

x


C.1. SOLUTIONS FOR CHAPTER 1. 307<br />

f ′ (x)<br />

f(x) = x2 + 1<br />

x<br />

ln(f(x)) = C + 1<br />

3 x3 + ln(x)<br />

Where C is a constant. This gives, with D a constant:<br />

To check this answer we note<br />

f(x) = Dxe 1<br />

3 x3<br />

ds = Dx 2 ye 1<br />

3 x3<br />

dx + De 1<br />

3 x3<br />

With the test given in problem 1 we have A(x, y) = Dx 2 ye 1<br />

3 x3<br />

and ( )<br />

∂A<br />

∂y<br />

( )<br />

∂B<br />

∂x<br />

which is an exact differential indeed.<br />

Problem 3.<br />

Part a.<br />

x<br />

y<br />

= Dx 2 e 1<br />

3 x3<br />

= Dx 2 e 1<br />

3 x3<br />

dU = sin(x + y)dx + sin(x + y)dy<br />

, B(x, y) = De 1<br />

3 x3<br />

With the test given in problem 1 we have A(x, y) = sin(x + y) , B(x, y) =<br />

sin(x + y) and<br />

( )<br />

∂A<br />

= cos(x + y)<br />

∂y x<br />

( )<br />

∂B<br />

= cos(x + y)<br />

∂x<br />

which shows th<strong>at</strong> dU is an exact differential. We have<br />

Part b.<br />

y<br />

U(x, y) = U0 − cos(x + y)<br />

¯dW = sin(x) cos(y)dx


308 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

With the test given in problem 1 we have A(x, y) = sin(x) cos(y) , B(x, y) = 0<br />

and ( )<br />

∂A<br />

= − sin(x) sin(y)<br />

∂y x<br />

( )<br />

∂B<br />

= 0<br />

∂x<br />

which shows th<strong>at</strong> ¯ dW is NOT an exact differential.<br />

Part c. From the first law dU = ¯ dQ − ¯ dW we get<br />

Part d.<br />

Step 1 → 2<br />

¯dQ = [sin(x + y) + sin(x) cos(y)] dx + sin(x + y)dy<br />

W1,2 =<br />

y<br />

U(0, 0) = U0 − 1<br />

U(απ, 0) = U0 − cos(απ)<br />

U(απ, απ) = U0 − cos(2απ)<br />

U(0, απ) = U0 − cos(απ)<br />

∫ απ<br />

Hence Q1,2 = 2 − 2 cos(απ)<br />

Step 2 → 3<br />

0<br />

U2 − U1 = 1 − cos(απ)<br />

sin(x) cos(0)dx = − cos(απ) + 1<br />

U3 − U2 = cos(απ) − cos(2απ)<br />

W2,3 = 0 if x constant no work done!<br />

Hence Q2,3 = cos(απ) − cos(2απ)<br />

Step 3 → 4<br />

U4 − U3 = cos(2απ) − cos(απ)<br />

W3,4 =<br />

∫ 0<br />

απ<br />

sin(x) cos(απ)dx = (−1 + cos(απ)) cos(απ)<br />

Hence Q3,4 = cos(2απ) − cos(απ) − (1 − cos(απ)) cos(απ)<br />

Step 4 → 1<br />

U1 − U4 = cos(απ) − 1<br />

Hence Q4,2 = cos(απ) − 1<br />

Adding all components:<br />

W4,1 = 0 if x constant no work done!<br />

Wtotal = cos 2 (απ) − 2 cos(απ) + 1 = (1 − cos(απ)) 2


C.1. SOLUTIONS FOR CHAPTER 1. 309<br />

Qtotal = cos 2 (απ) − 2 cos(απ) + 1 = (1 − cos(απ)) 2<br />

Indeed Wtotal − Qtotal = 0 as needed for a cyclic process. This means th<strong>at</strong> Q is<br />

always positive, work is always produced, it is a he<strong>at</strong> engine. The only exception<br />

is α = 2 where nothing happened, Q=0.<br />

Problem 4.<br />

Step 1: Initial st<strong>at</strong>e (p0, T0, V0), final st<strong>at</strong>e (p0, T1, 2V0)<br />

Step 2: Initial st<strong>at</strong>e (p0, T1, 2V0) , final st<strong>at</strong>e (p1, 1<br />

2T1, 2V0)<br />

Step 3: Initial st<strong>at</strong>e (p1, 1<br />

2T1, 2V0) , final st<strong>at</strong>e (p1, T2, V0)<br />

Step 4: Initial st<strong>at</strong>e (p1, T2, V0) , final st<strong>at</strong>e (p0, T0, V0)<br />

The equ<strong>at</strong>ions for the ideal gas are pV = NRT and U = 3pV<br />

. This gives<br />

2<br />

T1 = 2T0 for step 1, p1 = 1<br />

2p0 for step 2, T2 = 1<br />

4T1 = 1<br />

2T0 for step 3, and<br />

p1 p0<br />

= T2 T0<br />

for step 4. The last equ<strong>at</strong>ion is consistent with the previous three.<br />

Hence we have<br />

Step 1 ∆U = 3<br />

2 p0V0<br />

Step 2 ∆U = 3<br />

2 (p1 − p0)2V0 = − 3<br />

2 p0V0<br />

Step 3 ∆U = − 3<br />

2 p1V0 = − 3<br />

4 p0V0<br />

Step 4 ∆U = 3<br />

2 (p0 − p1)V0 = 3<br />

4 p0V0<br />

Indeed the total change in internal energy is zero. It is easy to calcul<strong>at</strong>e the<br />

work done in all steps.<br />

Step 1 W = p0V0, hence Q = 5<br />

2 p0V0.<br />

Step 2 W = 0 , hence Q = − 3<br />

2 p0V0<br />

Step 3 W = −p1V0 = − 1<br />

2 p0V0, hence Q = − 5<br />

4 p0V0.<br />

Step 4 W = 0 , hence Q = 3<br />

4 p0V0<br />

This gives Wtotal = 1<br />

2p0V0 , Qin = 13<br />

Wtotal = Qin + Qout.<br />

The efficiency is η = W 2 = Qin 13<br />

The Carnot efficiency would be ηC = 1 − T2<br />

T1<br />

Problem 5.<br />

Part a.<br />

U(S, V ) = S2<br />

V gives<br />

T =<br />

4 p0V0, Qout = − 11<br />

( )<br />

∂U<br />

∂S V<br />

4 p0V0 which indeed obeys<br />

3 = 4 , which is much larger.<br />

= 2 S<br />

V


310 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

p = −<br />

( )<br />

∂U<br />

=<br />

∂V S<br />

S2<br />

V 2<br />

which gives the equ<strong>at</strong>ion of st<strong>at</strong>e 4p = T 2 . Therefore, <strong>at</strong> constant T we have<br />

constant p. Also, <strong>at</strong> constant S we see th<strong>at</strong> T V = 2S, constant.<br />

Part b.<br />

Carnot cycle<br />

Step 1: (p1, V1) → (p2, V2) isothermal <strong>at</strong> Th<br />

Step 2: (p2, V2) → (p3, V3) adiab<strong>at</strong>ic Th → Tl<br />

Step 3: (p3, V3) → (p4, V4) isothermal <strong>at</strong> Tl<br />

Step 4: (p4, V4) → (p1, V1) adiab<strong>at</strong>ic Tl → Th<br />

1 → 2<br />

2 → 3<br />

3 → 4<br />

4 → 1<br />

This gives<br />

W1,2 =<br />

∫ V2<br />

V1<br />

pdV = 1<br />

4 T 2 h(V2 − V1)<br />

Q1,2 = Th(S2 − S1) = Th( 1<br />

2 ThV2 − 1<br />

2 ThV1) = 1<br />

2 T 2 h(V2 − V1)<br />

W2,3 =<br />

∫ V3<br />

V2<br />

pdV = S 2 2<br />

W3,4 =<br />

∫ V4<br />

V3<br />

∫ V3<br />

V2<br />

(<br />

1 1<br />

dV =<br />

V 2 2 ThV2<br />

) 2<br />

( 1<br />

−<br />

V2<br />

1<br />

)<br />

V3<br />

Q2,3 = 0<br />

pdV = 1 2<br />

Tl (V4 − V3)<br />

4<br />

Q3,4 = Tl(S4 − S3) = Tl( 1<br />

2 TlV4 − 1<br />

W4,1 =<br />

∫ V1<br />

V4<br />

Wtotal = 1<br />

4 T 2 h(V2 − V1) + 1<br />

pdV = S 2 4<br />

∫ V1<br />

V4<br />

4 T 2 hV2(1 − V2<br />

2 TlV3) = 1<br />

2<br />

T 2<br />

l (V4 − V3)<br />

(<br />

1 1<br />

dV =<br />

V 2 2 TlV4<br />

) 2<br />

( 1<br />

−<br />

V4<br />

1<br />

)<br />

V1<br />

Q4,1 = 0<br />

V3<br />

) + 1 2<br />

Tl (V4 − V3) +<br />

4 1<br />

4<br />

2<br />

Tl V4(1 − V4<br />

V1<br />

For the adiab<strong>at</strong>ic steps we have V2Th = V3Tl and V4Tl = V1Th which gives<br />

T 2 V2<br />

hV2 V3<br />

= T 2<br />

l V3. Therefore<br />

Wtotal = 1<br />

4 T 2 h(V2 − V1) + 1<br />

4 T 2 hV2 − 1<br />

4<br />

2<br />

Tl V3 + 1 2<br />

T<br />

1<br />

2 T 2 h(V2 − V1) + 1 2<br />

Tl (V4 − V3)<br />

2<br />

4<br />

l (V4 − V3) + 1 2<br />

Tl V4 −<br />

4 1<br />

4 T 2 hV1 =<br />

)


C.1. SOLUTIONS FOR CHAPTER 1. 311<br />

Qin = 1<br />

2 T 2 h(V2 − V1)<br />

Qout = 1 2<br />

Tl (V4 − V3)<br />

2<br />

and Wtotal = Qin + Qout indeed. The efficiency is<br />

η = W<br />

2 Tl = 1 +<br />

Qin T 2 V4 − V3<br />

h<br />

V2 − V1<br />

which can be simplified because Tl(V4 − V3) = Th(V1 − V2) and gives<br />

as expected.<br />

Problem 6.<br />

η = 1 − Tl<br />

This problem is along the same lines as the previous one, with dU = T dS+HdM<br />

and work done by changing the magnetic field. Given is th<strong>at</strong> U = f(T ) , or<br />

U(S, M) = f(T (S, M)) and also<br />

M = CH<br />

T . In principle, this gives us U by using<br />

( )<br />

∂U<br />

T = = f<br />

∂S<br />

′ (T )<br />

M<br />

Th<br />

( ∂T<br />

∂S<br />

and a similar equ<strong>at</strong>ion for H. This approach is not needed here, though.<br />

Carnot cycle<br />

Step 1: (H1, M1) → (H2, M2) isothermal <strong>at</strong> Th<br />

Step 2: (H2, M2) → (H3, M3) adiab<strong>at</strong>ic Th → Tl<br />

Step 3: (H3, M3) → (H4, M4) isothermal <strong>at</strong> Tl<br />

Step 4: (H4, M4) → (H1, M1) adiab<strong>at</strong>ic Tl → Th<br />

Note: assume M2 < M1, so in step 1 we do work on the outside and he<strong>at</strong> flows<br />

in.<br />

1 → 2<br />

2 → 3<br />

∫ M2<br />

W1,2 = −<br />

M1<br />

HdM = − Th<br />

C<br />

U1,2 = 0<br />

∫ M2<br />

M1<br />

)<br />

M<br />

MdM = − Th<br />

2C (M 2 2 − M 2 1 )<br />

Q1,2 = U1,2 + W1,2 = − Th<br />

2C (M 2 2 − M 2 1 )<br />

U2,3 = f(Tl) − f(Th)<br />

Q2,3 = 0<br />

W2,3 = −U2,3 + Q2,3 = f(Th) − f(Tl)


312 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

3 → 4<br />

4 → 1<br />

∫ M4<br />

W3,4 = −<br />

M3<br />

U3,4 = 0<br />

HdM = − Tl<br />

∫ M4<br />

MdM = −<br />

C M3<br />

Tl<br />

2C (M 2 4 − M 2 3 )<br />

Q3,4 = U3,4 + W3,4 = − Tl<br />

2C (M 2 4 − M 2 3 )<br />

U4,1 = f(Th) − f(Tl)<br />

Q4,1 = 0<br />

W4,1 = −U4,1 + Q4,1 = f(Tl) − f(Th)<br />

This shows th<strong>at</strong> the total change in U is indeed zero and th<strong>at</strong><br />

hence<br />

Wtotal = − Th<br />

2C (M 2 2 − M 2 1 ) − Tl<br />

2C (M 2 4 − M 2 3 )<br />

Qin = − Th<br />

2C (M 2 2 − M 2 1 )<br />

η = W<br />

= 1 +<br />

Qin<br />

Tl M<br />

Th<br />

2 4 − M 2 3<br />

M 2 2 − M 2 1<br />

Finally, we need to reduce the last factor. We need to connect the change in M<br />

along the adiab<strong>at</strong>ic line:<br />

C<br />

2<br />

M 2 3 − M 2 2 = 1<br />

∫<br />

M3MdM =<br />

2 M2<br />

C<br />

∫ 3<br />

2 2<br />

∫ 3<br />

2<br />

( 1<br />

∫ 3<br />

C<br />

dU − dS) =<br />

T 2 2<br />

∫ Tl<br />

1 C<br />

dU =<br />

T 2 Th<br />

H<br />

dM =<br />

T<br />

1<br />

T f ′ (T )dT<br />

a similar expression can be found for the other adiab<strong>at</strong>ic process and we get<br />

or<br />

This gives<br />

as expected.<br />

M 2 3 − M 2 2 = M 2 4 − M 2 1<br />

M 2 3 − M 2 4 = M 2 2 − M 2 1<br />

η = 1 − Tl<br />

Th


C.1. SOLUTIONS FOR CHAPTER 1. 313<br />

Problem 7.<br />

In step 1, adiab<strong>at</strong>ic expansion. V0 → 2V0 and T0 → Tint. VT constant, hence<br />

Tint = 1<br />

2 T0<br />

In step 2, isothermal compression. 2V0 → V0 <strong>at</strong> constant T.<br />

Hence after 1 step T1 = 1<br />

2 T0. This means th<strong>at</strong> after n steps<br />

and we never reach zero.<br />

Problem 8.<br />

Part a.<br />

Initially U = U1 + U2 = C1T1 + C2T2.<br />

Finally U = (C1 + C2)Tf .<br />

No energy exchange, hence<br />

Tn = 1<br />

T0<br />

2n Tf = C1T1 + C2T2<br />

C1 + C2<br />

Part b.<br />

∆S = ∫ 1<br />

T dU holds for each system. Therefore<br />

∫ Tf 1<br />

∆Stotal =<br />

T1 T C1dT<br />

∫ Tf 1<br />

+<br />

T2 T C2dT =<br />

This gives<br />

C1(ln(Tf ) − ln(T1)) + C2(ln(Tf ) − ln(T2))<br />

∆Stotal = (C1 + C2) ln(Tf ) − C1 ln(T1) − C2 ln(T2)<br />

We know th<strong>at</strong> (C1 + C2)Tf = C1T1 + C2T2 and we need to show th<strong>at</strong> (C1 +<br />

C2) ln(Tf ) > C1 ln(T1) + C2 ln(T2)<br />

This follows from the n<strong>at</strong>ure of the logarithm. It is a concave function. In other<br />

words, if z = cx + (1 − c)y with c between 0 and 1, meaning th<strong>at</strong> z is between<br />

x and y, we have ln(z) > c ln(x) + (1 − c) ln(y).<br />

Problem 16.<br />

Consider the following differential:<br />

¯df = g(x, y, z)dx + h(x, y, z)dy + k(x, y, z)dz<br />

Wh<strong>at</strong> are the necessary and sufficient conditions for ¯ df to be exact?<br />

If the function f is a st<strong>at</strong>e function f(x, y, z) we have


314 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

from which it follows th<strong>at</strong><br />

( )<br />

∂f<br />

g(x, y, z, ) =<br />

∂x<br />

( )<br />

∂f<br />

h(x, y, z, ) =<br />

∂y<br />

( )<br />

∂f<br />

k(x, y, z, ) =<br />

∂z<br />

( )<br />

∂g<br />

∂y<br />

( )<br />

∂g<br />

∂z<br />

( )<br />

∂h<br />

∂z<br />

x,z<br />

x,y<br />

x,y<br />

( )<br />

∂h<br />

=<br />

∂x<br />

( )<br />

∂k<br />

=<br />

∂x<br />

( )<br />

∂k<br />

=<br />

∂y<br />

since in the second order deriv<strong>at</strong>ives of f we can interchange the order of the<br />

deriv<strong>at</strong>ives. Once these rel<strong>at</strong>ions are obeyed, rel<strong>at</strong>ions for higher order deriv<strong>at</strong>ives<br />

follow. Hence the rel<strong>at</strong>ions given above are the primary rel<strong>at</strong>ions. They<br />

are necessary.<br />

Suppose th<strong>at</strong> these rel<strong>at</strong>ions are true, can we derive th<strong>at</strong> the differential is exact?<br />

Yes, follow the line from the notes and define<br />

which gives<br />

H(x, y, z) =<br />

( )<br />

∂H<br />

∂y<br />

∫ x<br />

0<br />

∫ x<br />

dx<br />

0<br />

′ g(x ′ ∫ y<br />

, y, z) + dy<br />

0<br />

′ h(0, y ′ ∫ z<br />

, z) + dz<br />

0<br />

′ k(0, 0, z ′ )<br />

x,z<br />

dx ′<br />

( ) ∫ x<br />

∂H<br />

= dx<br />

∂z x,y 0<br />

′<br />

∫ x<br />

dx ′<br />

( )<br />

∂k<br />

∂x<br />

0<br />

∫ x<br />

0<br />

y,z<br />

x,z<br />

x,y<br />

y,z<br />

y,z<br />

x,z<br />

( )<br />

∂H<br />

= g(x, y, z)<br />

∂x y,z<br />

∫ x<br />

= dx<br />

0<br />

′<br />

( )<br />

∂g<br />

∂y<br />

( )<br />

∂h<br />

∂x<br />

(x<br />

x,z<br />

′ , y, z) + h(0, y, z) =<br />

(x<br />

y,z<br />

′ , y, z) + h(0, y, z) = h(x, y, z)<br />

( )<br />

∂g<br />

(x<br />

∂z x,y<br />

′ , y, z) +<br />

(x<br />

y,z<br />

′ , y, z) +<br />

dx ′<br />

( ∂k<br />

∂x<br />

)<br />

∫ y<br />

0<br />

dy ′<br />

∫ y<br />

0<br />

dy ′<br />

( ∂k<br />

∂y<br />

)<br />

( )<br />

∂h<br />

(0, y<br />

∂z x,y<br />

′ , z) + k(0, 0, z) =<br />

(0, y<br />

x,z<br />

′ , z) + k(0, 0, z) =<br />

(x<br />

y,z<br />

′ , y, z) + k(0, y, z) = k(x, y, z)


C.1. SOLUTIONS FOR CHAPTER 1. 315<br />

Consider an arbitrary p<strong>at</strong>h (ϕ(t), ψ(t), λ(t)) from (x0, y0, z0) to (x1, y1, z1). The<br />

functions ϕ,ψ, and λ are arbitrary except th<strong>at</strong> ϕ(0) = x0, ϕ(1) = x1, etc.<br />

Following the differential along this p<strong>at</strong>h gives<br />

∆f =<br />

∫ 1<br />

0<br />

(<br />

g(ϕ(t), ψ(t), λ(t)) dϕ<br />

(t) + h(ϕ(t), ψ(t), λ(t))dψ<br />

dt dt (t)+<br />

k(ϕ(t), ψ(t), λ(t)) dλ<br />

dt (t)<br />

)<br />

dt<br />

We define H(t) = H(ϕ(t), ψ(t), λ(t)) and find similar to the notes th<strong>at</strong><br />

∆f =<br />

∫ 1<br />

0<br />

dH<br />

dt = H(1) − H(0)<br />

dt<br />

which is p<strong>at</strong>h independent. Note th<strong>at</strong> we used all three conditions. Conditions<br />

on higher order deriv<strong>at</strong>ives can also be used, since our conditions can be derived<br />

by integr<strong>at</strong>ion. We need to make the extra assumption, though, th<strong>at</strong> the higher<br />

order deriv<strong>at</strong>ives do exist.<br />

Problem 17.<br />

Consider the following examples:<br />

¯df = 3x 2 y 2 zdx + 2x 3 yzdy + 6x 3 y 2 dz<br />

¯dg = 2xy 2 z 2 dx + 2x 2 yz 2 dy + 2x 2 y 2 zdz<br />

Are these differentials exact? If so, find the corresponding st<strong>at</strong>e function.<br />

First case:<br />

( ) 2 2 ∂3x y z<br />

= 6x<br />

∂y x,z<br />

2 yz<br />

( ) 2 2 ∂3x y z<br />

= 3x<br />

∂z x,y<br />

2 y 2<br />

( ) 3 ∂2x yz<br />

= 6x<br />

∂x y,z<br />

2 yz<br />

( ) 3 ∂2x yz<br />

= 2x<br />

∂z x,y<br />

3 y<br />

( ) 3 2 ∂6x y<br />

= 18x<br />

∂x y,z<br />

2 y 2<br />

( ) 3 2 ∂6x y<br />

= 12x<br />

∂y<br />

3 y<br />

x,z


316 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

which shows th<strong>at</strong> only the first of the three conditions is obeyed, hence this is<br />

not an exact differential.<br />

Second case:<br />

( ) 2 2 ∂2xy z<br />

= 4xyz<br />

∂y x,z<br />

2<br />

( ) 2 2 ∂2xy z<br />

= 4xy<br />

∂z x,y<br />

2 z<br />

( ) 2 2 ∂2x yz<br />

= 4xyz<br />

∂x y,z<br />

2<br />

( ) 2 2 ∂2x yz<br />

= 4x<br />

∂z x,y<br />

2 yz<br />

( ) 2 2 ∂2x y z<br />

= 4xy<br />

∂x y,z<br />

2 z<br />

( ) 2 2 ∂2x y z<br />

= 4x<br />

∂y<br />

2 yz<br />

which shows th<strong>at</strong> all three conditions are obeyed, hence this is an exact differential.<br />

It is not hard to guess th<strong>at</strong> the st<strong>at</strong>e function is given by<br />

Problem 18.<br />

x,z<br />

g(x, y, z, ) = g0 + x 2 y 2 z 2<br />

A system is completely described by the four st<strong>at</strong>e variables T,S,p, and V.<br />

1. Show th<strong>at</strong> the general form of the energy equ<strong>at</strong>ion is U = V f( S<br />

V ).<br />

2. If U = S α V 1−α find the equ<strong>at</strong>ions of st<strong>at</strong>e p(S,V) and T (S,V).<br />

3. The entropy S is not always easy to deal with, and we often elimin<strong>at</strong>e this<br />

variable. In the previous case, solve for p(T,V). Wh<strong>at</strong> is special in this<br />

equ<strong>at</strong>ion, and why is th<strong>at</strong> the case?<br />

4. Calcul<strong>at</strong>e the efficiency of a Carnot engine using a m<strong>at</strong>erial with the energy<br />

equ<strong>at</strong>ion U = S α V 1−α .<br />

We can always write U in the form U = V f(S, V ). Since U, S, and V are both<br />

extensive we have when we multiply all extensive variables by an arbitrary scale<br />

variable x:<br />

xU = xV f(xS, xV )


C.1. SOLUTIONS FOR CHAPTER 1. 317<br />

Combined with the original expression this gives<br />

f(S, V ) = f(xS, xV )<br />

for all non-zero values of x. This means th<strong>at</strong> f can only depend on the r<strong>at</strong>ion<br />

S<br />

V .<br />

From the first law:<br />

we have<br />

which leads to<br />

Note: this gives T S − pV = U !!<br />

The first equ<strong>at</strong>ion gives<br />

dU = T dS − pdV<br />

( )<br />

∂U<br />

T =<br />

∂S<br />

( )<br />

∂U<br />

p = −<br />

∂V<br />

U =<br />

S<br />

V<br />

= α U<br />

S<br />

= (α − 1) U<br />

V<br />

T S<br />

α<br />

U = pV<br />

α − 1<br />

( ) α−1<br />

S<br />

T = α<br />

V<br />

( ) α<br />

S<br />

p = (α − 1)<br />

V<br />

Since the left side is intensive, the right side can only depend on the r<strong>at</strong>io of S<br />

and V.<br />

We find<br />

( ) α<br />

α−1 T<br />

p = (α − 1)<br />

α<br />

hence p(T,V) does not depend on V. This is not surprising, since by scaling V we<br />

need th<strong>at</strong> p(T, xV ) = p(T, V ), and using x = 0 (which is allowed here, different<br />

from the case above!) shows th<strong>at</strong> p(T, V ) = p(T, 0) which is independent of V.<br />

In a Carnot cycle:<br />

Step 1: Isothermal expansion (p1, V1) to (p2, V2) <strong>at</strong> Th.<br />

Since the temper<strong>at</strong>ure is constant, the pressure is constant, and we have<br />

W1,2 = ph(V2 − V1)


318 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

and hence<br />

U2 − U1 = ph(V2 − V1)<br />

α − 1<br />

Q1,2 = αph(V2 − V1)<br />

α − 1<br />

Step 2: Adiab<strong>at</strong>ic expansion (ph, V2) to (pl, V3).<br />

The entropy is constant, Sh. Obviously we have<br />

Step 3 (similar to step 1):<br />

Step 4 (similar to step 2):<br />

Hence the total he<strong>at</strong> in is<br />

and the total work done is<br />

Q2,3 = 0<br />

W2,3 = U2 − U3 = (Th − Tl)Sh<br />

α<br />

W3,4 = pl(V4 − V3)<br />

Q3,4 = αpl(V4 − V3)<br />

α − 1<br />

Q4,1 = 0<br />

W4,1 = U4 − U1 = (Tl − Th)Sl<br />

α<br />

Qin = αph(V2 − V1)<br />

α − 1<br />

W = ph(V2 − V1) + pl(V4 − V3) + (Th − Tl)(Sh − Sl)<br />

α<br />

Using (from the equ<strong>at</strong>ions for the energy)<br />

we find<br />

or<br />

W =<br />

pV =<br />

α − 1<br />

T S<br />

α<br />

α − 1<br />

α Th(Sh<br />

α − 1<br />

− Sl) +<br />

α Tl(Sl − Sh) + (Th − Tl)(Sh − Sl)<br />

α


C.1. SOLUTIONS FOR CHAPTER 1. 319<br />

(No surprise, total he<strong>at</strong> exchange!!!)<br />

and<br />

which gives<br />

Problem 19.<br />

W = (Th − Tl)(Sh − Sl)<br />

Qin = Th(Sh − Sl)<br />

η = W<br />

=<br />

Qin<br />

Th − Tl<br />

Th<br />

= ηCarnot<br />

Consider the following cyclic process (Brayton), with an ideal mono-<strong>at</strong>omic gas<br />

as the working m<strong>at</strong>erial :<br />

1. Adiab<strong>at</strong>ic expansion from pA, V1, T1 to pB, V2, T2.<br />

2. Isobaric expansion from pB, V2, T2 to pB, V3, T3.<br />

3. Adiab<strong>at</strong>ic contraction from pB, V3, T3 to pA, V4, T4.<br />

4. Isobaric contraction from pA, V4, T4 to pA, V1, T1.<br />

Calcul<strong>at</strong>e the efficiency in terms of pA and pB. Compare this with the Carnot<br />

efficiency (indic<strong>at</strong>e the temper<strong>at</strong>ures used, and why you choose these references).<br />

Step 1: S constant (see notes)<br />

Step 2: p constant<br />

Step 3:<br />

Step 4:<br />

W1,2 = U1 − U2 = 3<br />

2 (V1pA − V2pB)<br />

W2,3 = pB(V3 − V2)<br />

W3,4 = 3<br />

2 (V3pB − V4pA)<br />

W4,1 = pA(V1 − V4)<br />

W = 5<br />

2 (pA(V1 − V4) + pB(V3 − V2))<br />

In step 2 the change in energy follows from U = 3<br />

2pV and is


320 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

U3 − U2 = 3<br />

2 pB(V3 − V2)<br />

Hence the energy flow in this step is<br />

Therefore<br />

In the adiab<strong>at</strong>ic steps we have<br />

Qin = Q2,3 = 5<br />

2 pB(V3 − V2)<br />

η = W<br />

= 1 +<br />

Qin<br />

pA(V1 − V4)<br />

pB(V3 − V2)<br />

pAV γ<br />

1<br />

pBV γ<br />

3<br />

= pBV γ<br />

2<br />

= pAV γ<br />

4<br />

with γ = 5<br />

3 , and hence V1V3 = V2V4 as in the notes.<br />

Also<br />

and hence<br />

or<br />

V3 = V4<br />

V2 = V1<br />

η = 1 −<br />

η = 1 −<br />

( pA<br />

pB<br />

( pA<br />

pB<br />

( pA<br />

pB<br />

( pA<br />

During the constant entropy parts we have<br />

Hence we have<br />

pB<br />

3 −<br />

V ∝ T 2<br />

p ∝ T 5<br />

2<br />

) 1<br />

γ<br />

) 1<br />

γ<br />

) γ−1<br />

γ<br />

η = 1 − T1<br />

T2<br />

which is a Carnot-like efficiency. However, we have an even lower temper<strong>at</strong>ure,<br />

T3 available, and it would be more efficient to run a Carnot engine between T1<br />

and T3!<br />

) 2<br />

5


C.2. SOLUTIONS FOR CHAPTER 2. 321<br />

C.2 Solutions for chapter 2.<br />

Problem 1.<br />

Part a.<br />

The Helmholtz Free Energy<br />

Part b.<br />

Ftotal = F (T, V1, N1) + F (T, V − V1, N − N1)<br />

This should be a minimum as a function of the internal parameter V1<br />

( ∂Ftotal<br />

∂V1<br />

This is zero if<br />

) ( )<br />

( )<br />

( )<br />

∂F<br />

∂F<br />

∂(V − V1)<br />

= (T, V1, N1) + (T, V − V1, N − N1)<br />

∂V<br />

∂V<br />

∂V1<br />

p(T, V1, N1) = p(T, V − V1, N − N1)<br />

where we used ( )<br />

∂F<br />

∂V = p.<br />

T,N<br />

Hence the pressure left and right needs to be the same.<br />

For the second order deriv<strong>at</strong>ives we find<br />

( 2 ∂ Ftotal<br />

∂V 2<br />

) ( 2 ∂ F<br />

=<br />

1 ∂V 2<br />

)<br />

( 2 ∂ F<br />

(T, V1, N1) +<br />

∂V 2<br />

)<br />

(T, V − V1, N − N1)<br />

where the second term now has a plus sign again. This has to be positive! In<br />

general, we can take V1 to be half V and we find th<strong>at</strong> we need<br />

or<br />

1<br />

The left hand side is equal to<br />

in the notes.<br />

Problem 2.<br />

−V κT =<br />

( 2 ∂ F<br />

∂V 2<br />

)<br />

> 0<br />

T,N<br />

( )<br />

∂p<br />

−<br />

∂V T,V<br />

V κT<br />

( ) ( ) ( )<br />

∂V ∂V ∂V<br />

= +<br />

∂p T,N ∂p S,N ∂S p,N<br />

> 0<br />

which is positive indeed, as has been shown<br />

( )<br />

( ) ( )<br />

∂S<br />

∂V ∂V<br />

= −V κS−<br />

∂p T,N<br />

∂S p,N ∂T p,N


322 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

where we used a Maxwell rel<strong>at</strong>ion based on the Gibbs Free Energy G(T,p,N).<br />

Therefore<br />

( )<br />

∂V<br />

κT − κS = α<br />

∂S p,N<br />

Using a Maxwell rel<strong>at</strong>ion based on the Enthalpy H(S,p,N) we find<br />

( )<br />

∂V<br />

∂S<br />

( )<br />

∂T<br />

=<br />

∂p<br />

( )<br />

∂T<br />

= −<br />

∂S<br />

( )<br />

∂S<br />

∂p<br />

p,N<br />

S,N<br />

where we also used the identity connecting three partial deriv<strong>at</strong>ives. The first<br />

partial deriv<strong>at</strong>ive is<br />

( )<br />

∂T<br />

=<br />

∂S p,N<br />

T<br />

Cp<br />

and the second is<br />

p,N<br />

( ) ( )<br />

∂S<br />

∂V<br />

= − = V α<br />

∂p T,N ∂T p,N<br />

as we used before. Combining all of this:<br />

Problem 3.<br />

T V α2<br />

κT − κS =<br />

Part a.<br />

Basic variables T, V, µ hence use the Grand Potential Ω = U − T S − µN.<br />

Part b.<br />

for RT


C.2. SOLUTIONS FOR CHAPTER 2. 323<br />

Part d.<br />

If S is infinity for µ = 0 th<strong>at</strong> is certainly also true <strong>at</strong> T=0 and µ = 0. Wh<strong>at</strong> is<br />

wrong?<br />

The formula only is valid for RT 0<br />

= 2S2<br />

> 0<br />

V 3<br />

= − 2S<br />

< 0<br />

V 2<br />

( 2 ∂ U<br />

∂S2 )<br />

V<br />

( )<br />

∂U<br />

∂S V<br />

= 2S<br />

V<br />

( )<br />

∂U<br />

=<br />

∂V S<br />

S2<br />

V 2<br />

( 2 ∂ U<br />

∂V 2<br />

)<br />

S<br />

Which gives 4p = T 2 , not surprising since we have only two intensive variables<br />

and they cannot be independent.<br />

Therefore U(T, V ) = 1<br />

4T 2V and CV = 1<br />

2T V = S ( of course! ).<br />

Also Cp does not exist, since we cannot change T while keeping p constant.<br />

The volume depends on pressure according to V = S √<br />

p and κS = − 1<br />

( )<br />

∂V<br />

V ∂p<br />

S =<br />

1<br />

2p<br />

Also here κT does not exist since we cannot change p <strong>at</strong> constant T.<br />

Finally, α cannot be defined, since we cannot change T <strong>at</strong> constant p.


324 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Problem 6.<br />

Hence<br />

Problem 7.<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

=<br />

V N<br />

( 2 ∂ U<br />

∂S2 )<br />

( 2 ∂ S<br />

∂U 2<br />

)<br />

V N<br />

V N<br />

( ) −1 ∂T<br />

∂U<br />

( )<br />

∂T<br />

=<br />

∂S<br />

V N<br />

V N<br />

= − 1<br />

T 2 CV<br />

= T<br />

CV<br />

( 2 ∂ U<br />

∂S2 )<br />

= −<br />

V N<br />

1<br />

T C2 V<br />

< 0<br />

Part a.<br />

( )<br />

∂S<br />

CM = T<br />

∂T MN<br />

( )<br />

∂S<br />

CH = T<br />

∂T HN<br />

( ) ( ) ( )<br />

∂S<br />

∂S ∂M<br />

= T<br />

+ T<br />

∂T MN ∂M T N ∂T<br />

( )<br />

∂S<br />

CH − CM = T<br />

∂M<br />

( )<br />

∂M<br />

∂T<br />

( )<br />

∂H<br />

= −T<br />

∂T<br />

( )<br />

∂M<br />

∂T<br />

or<br />

Part b.<br />

CH − CM = T<br />

T N<br />

( )<br />

∂H<br />

∂M T N<br />

HN<br />

( )<br />

∂M<br />

∂T HN<br />

MN<br />

( )<br />

∂M<br />

=<br />

∂T HN<br />

T<br />

χT<br />

HN<br />

HN<br />

( ) 2<br />

∂M<br />

∂T HN<br />

( )<br />

∂M<br />

χS =<br />

∂H SN<br />

( )<br />

∂M<br />

χT =<br />

∂H T N<br />

( ) ( ) ( )<br />

∂M ∂M ∂S<br />

=<br />

+<br />

∂H SN ∂S HN ∂H T N<br />

( ) ( )<br />

∂M ∂S<br />

χT − χS =<br />

∂S HN ∂H T N<br />

( )<br />

∂M<br />

∂S HN<br />

( ) ( )<br />

∂M ∂T<br />

=<br />

∂T HN ∂S HN<br />

( )<br />

∂M<br />

=<br />

∂T HN<br />

T<br />

CH<br />

( )<br />

∂S<br />

∂H<br />

( )<br />

∂M<br />

=<br />

∂T<br />

T N<br />

HN


C.2. SOLUTIONS FOR CHAPTER 2. 325<br />

and hence<br />

χT − χS =<br />

( )<br />

∂M<br />

∂T HN<br />

T<br />

CH<br />

( )<br />

∂M<br />

∂T HN<br />

Part c.<br />

Using the internal energy U(S,M) we have for stability<br />

( 2 ∂ U<br />

0 <<br />

∂S2 ) ( )<br />

∂T<br />

=<br />

=<br />

∂S<br />

T<br />

CM<br />

or CM > 0. Also<br />

or χS > 0.<br />

From parts a and b:<br />

0 <<br />

MN<br />

MN<br />

( 2 ∂ U<br />

∂M 2<br />

) ( )<br />

∂H<br />

=<br />

=<br />

SN ∂M SN<br />

1<br />

χS<br />

χT (CH − CM ) = CH(χT − χS)<br />

which gives χT CM = CHχS or CH and χT have the same sign.<br />

Finally, we need the mixed inequality.<br />

The right hand side is T<br />

CM χS<br />

and<br />

( )2 ( 2 2 ∂ U ∂ U<br />

<<br />

∂M∂S N ∂M 2<br />

)<br />

SN<br />

( 2 ∂ U<br />

∂S2 )<br />

MN<br />

while the left hand side is<br />

( ) ( ) ( ) ( )<br />

2 ∂ U ∂T<br />

∂T ∂S<br />

=<br />

=> −<br />

= −<br />

∂M∂S N ∂M SN ∂S MN ∂M T N<br />

T<br />

CM<br />

gives<br />

or<br />

( ) ( )<br />

∂H<br />

∂H<br />

= −<br />

∂T MN ∂M T N<br />

T<br />

CMχS<br />

><br />

[ ( )<br />

∂M T ∂T HN<br />

CM χT<br />

CH − CM<br />

χT<br />

] 2<br />

< CM<br />

χS<br />

( )<br />

∂M<br />

∂T HN<br />

= T (CH − CM )<br />

C 2 M χT<br />

= CH<br />

χT<br />

( )<br />

∂H<br />

∂T MN<br />

from which it follows th<strong>at</strong> CM<br />

χT > 0 or χT > 0 and since CH has the same sign,<br />

we have CH > 0. Parts a and b then give χT > χS and CH > CM.<br />

Part d.


326 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

From M = CNH<br />

T we get χT = CN<br />

T<br />

From U = 3<br />

2NkT we get CM = 3<br />

2Nk With ( )<br />

∂M<br />

CNH<br />

∂T = − HN T 2 we get using parts a and b:<br />

( 2 T CNH<br />

CH − CM =<br />

CN T 2<br />

) 2<br />

= CNH2<br />

T 2<br />

or CH = 3 MH<br />

2Nk + T<br />

and also<br />

or<br />

Problem 10.<br />

Part A.<br />

Part B.<br />

Part C.<br />

Part D.<br />

Part E.<br />

This means<br />

χT − χS = T<br />

χS = CN<br />

T −<br />

( )<br />

∂S<br />

=<br />

∂V T N<br />

∆Q = T ∆S = T<br />

CH<br />

( CNH<br />

T 2<br />

C 2 NH 2<br />

) 2<br />

3<br />

2 kT 3 + CH 2 T<br />

( )<br />

∂p<br />

=<br />

∂T V N<br />

NR<br />

V<br />

∆W = p∆V = NRT<br />

V ∆V<br />

( )<br />

∂S<br />

∆V =<br />

∂V T N<br />

NRT<br />

V ∆V<br />

∆U = ∆Q − ∆W = 0<br />

( )<br />

∂U<br />

= 0<br />

∂V T N<br />

or U = g(T, N).<br />

Part F.<br />

is intensive and cannot depend on N, therefore U = Nf(T ).<br />

U<br />

N<br />

Problem 11.


C.2. SOLUTIONS FOR CHAPTER 2. 327<br />

The initial st<strong>at</strong>e with N moles of m<strong>at</strong>erial has pressure p, temper<strong>at</strong>ure T, and<br />

volume V. Hence we can calcul<strong>at</strong>e the energy U, the entropy S, and the chemical<br />

potential µ. The final st<strong>at</strong>e has volume V ′ , amount of m<strong>at</strong>erial N ′ = N<br />

moles. Because there is no energy exchange, the final st<strong>at</strong>e has energy U ′ = U.<br />

Because we know three st<strong>at</strong>e variables in the final st<strong>at</strong>e, all other variables can<br />

be calcul<strong>at</strong>ed. Note th<strong>at</strong> the entropy S will change, even though there is no he<strong>at</strong><br />

exchange.<br />

Part A.<br />

For the ideal gas we have U = 3<br />

2 NRT , and hence T ′ = T . Using the ideal gas<br />

law we get p ′ V ′ = NRT ′ = NRT = pV , or p ′ = p V<br />

V ′ . The temper<strong>at</strong>ure stayed<br />

the same, but the pressure went down.<br />

Part B.<br />

Assume th<strong>at</strong> we take a reversible p<strong>at</strong>h from the initial to the final st<strong>at</strong>e. Because<br />

the entropy changes, this will be a p<strong>at</strong>h along which he<strong>at</strong> is exchanged and work<br />

is done, but in such a manner th<strong>at</strong> these two cancel. We need to calcul<strong>at</strong>e:<br />

We have<br />

T ′ − T =<br />

∫ V ′<br />

( ) ( )<br />

∂T<br />

∂T<br />

= −<br />

∂V U,N ∂U V,N<br />

V<br />

( )<br />

∂T<br />

dV<br />

∂V U,N<br />

( )<br />

∂U<br />

= −<br />

∂V T,N<br />

1<br />

CV<br />

( )<br />

∂U<br />

∂V T,N<br />

which we use since CV was given. Since we know partial deriv<strong>at</strong>ives of U <strong>at</strong><br />

constant values of extensive st<strong>at</strong>e variables, we write<br />

( )<br />

∂U<br />

=<br />

∂V T,N<br />

( )<br />

∂U<br />

+<br />

∂V S,N<br />

( )<br />

∂U<br />

∂S V,N<br />

( )<br />

∂S<br />

= −p + T<br />

∂V T,N<br />

( )<br />

∂S<br />

∂V T,N<br />

Since we do not like the partial deriv<strong>at</strong>ives of S we use Maxwell to get:<br />

and hence<br />

Finally:<br />

T ′ − T =<br />

( ) ( ) 2 ∂S<br />

∂ F<br />

= −<br />

=<br />

∂V T,N ∂T ∂V N<br />

∫ V ′<br />

V<br />

( )<br />

∂p<br />

=<br />

∂T V,N<br />

NR<br />

V − Nb<br />

( )<br />

∂U<br />

= −p +<br />

∂V T,N<br />

NR<br />

= aN2<br />

V − Nb V 2<br />

( )<br />

∂T<br />

∂V U,N<br />

2 ∫ ′<br />

V<br />

2aN<br />

dV = −<br />

3NR V<br />

which is neg<strong>at</strong>ive indeed, assuming th<strong>at</strong> a > 0.<br />

( )<br />

1 2aN 1 1<br />

dV = −<br />

V 2 3R V ′ V


328 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Problem 16.<br />

Part A.<br />

Part B.<br />

since S(0,V)=0 and hence<br />

Part C.<br />

( )<br />

∂S<br />

∂T V<br />

= 1<br />

T<br />

S(T, V ) =<br />

S(T, V ) =<br />

∫ T<br />

0<br />

∫ T<br />

0<br />

( )<br />

∂U<br />

∂T V<br />

= 2cV<br />

( )<br />

∂S<br />

dT<br />

∂T V<br />

2cV dT = 2cV T<br />

U = T S − pV ⇒ pV = T S − U = cV T 2<br />

or p = cT 2 . Note this is the example we used before starting with U(S, V ) = S2<br />

V<br />

and c = 1<br />

4 .<br />

Problem 17.<br />

Part A.<br />

At constant volume and number of moles dU = T dS − pdV + µdN = T dS and<br />

hence CV = T ∂S ∂U 3<br />

∂T = ∂T = 2NR. Part B.<br />

S(T, V, N)−S(T0, V, N) =<br />

Part C.<br />

∫ T<br />

T0<br />

( ) ∫ T<br />

∂S<br />

CV<br />

dT =<br />

∂T V,N T0 T<br />

lim S(T, V, N) = −∞<br />

T →0<br />

3<br />

dT = NR(ln(T )−ln(T0))<br />

2<br />

which is against the third law. Hence an ideal gas cannot be the correct description<br />

of a gas <strong>at</strong> very low temper<strong>at</strong>ure.<br />

Part D.<br />

H = U + pV = 5<br />

2 NRT<br />

Part E.<br />

At constant pressure and number of moles dH = T dS + V dp + µdN = T dS and<br />

hence CV = T ∂S ∂H 5<br />

∂T = ∂T = 2NR.


C.2. SOLUTIONS FOR CHAPTER 2. 329<br />

Problem 18.<br />

The mass of a mole of gas is m, and hence <strong>at</strong> height h the gravit<strong>at</strong>ional potential<br />

energy is mgh per mole. The total chemical potential is<br />

µtotal = µintrinsic + mgh<br />

which is constant in equilibrium. Therefore,<br />

and<br />

which leads to<br />

Problem 19.<br />

RT ln( n<br />

n<br />

nG<br />

nG<br />

) + mgh = C<br />

= e C−mgh<br />

RT<br />

mgh<br />

−<br />

n(h) = n(0)e RT<br />

Part A.<br />

For an adiab<strong>at</strong>ic process we have pV γ = constant, and the work done per mole<br />

is therefore<br />

W =<br />

∫ 2<br />

1<br />

pdV =<br />

∫ 2<br />

1<br />

c c<br />

(<br />

dV = V<br />

V γ 1 − γ<br />

1−γ<br />

2<br />

which is (with the ideal gas law pV = NRT ):<br />

W = 1<br />

1 − γ (pV2 − pV1) = R<br />

1 − γ (T2 − T1)<br />

− V 1−γ<br />

)<br />

1<br />

Remember this is per mole! Using the value of γ for w<strong>at</strong>er we get<br />

and hence<br />

W = 3R (T1 − T2)<br />

f = W<br />

H<br />

= 300R<br />

H<br />

Part B.<br />

The gain in height is 100m. The gain in potential energy is<br />

The amount of he<strong>at</strong> needed is<br />

∆Ug = (M + Mfuel + Mw<strong>at</strong>er) 100g


330 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Q = 1<br />

f ∆Ug = Hg<br />

3R (M + Mfuel + Mw<strong>at</strong>er)<br />

and hence the mount of fuel mf used is:<br />

mf = Hg<br />

3JR (M + Mfuel + Mw<strong>at</strong>er)<br />

Part C.<br />

Similar, the amount of w<strong>at</strong>er mw used is:<br />

Problem 20.<br />

mw = g<br />

3R (M + Mfuel + Mw<strong>at</strong>er)<br />

Part A.<br />

At room temper<strong>at</strong>ure, air (di<strong>at</strong>omic molecules) has its rot<strong>at</strong>ional degrees of<br />

freedom activ<strong>at</strong>ed, but not its vibr<strong>at</strong>ional, and hence we have cV = 5<br />

2R and<br />

cp = 7<br />

2R. Part B.<br />

The he<strong>at</strong> needed to cool N moles of air is Q = Ncp∆T .<br />

The amount of air exchanged per day is Va = 4 × 8 × 0.75m3 , which is about<br />

0.28 liters per second, or about Na = 0.013 moles per second. This requires an<br />

energy Qa = Nacp∆T per second. Hence:<br />

Qa = 0.013 × 7<br />

× 8.31 × 18J<br />

2<br />

or Qa is about 7 Joules per second.<br />

The efficiency of the refriger<strong>at</strong>or is η = W<br />

18<br />

W and this is about +Qa 300 . Hence the<br />

refriger<strong>at</strong>or uses about 120 Joules per second to oper<strong>at</strong>e.<br />

Part C.<br />

Now we also need to cool and condense w<strong>at</strong>er. The gas going into the refriger<strong>at</strong>or<br />

is now 90% dry air and 10% w<strong>at</strong>er <strong>at</strong> room temper<strong>at</strong>ure. Therefore, we need to<br />

cool and condense about 0.0013 moles of w<strong>at</strong>er per second. For w<strong>at</strong>er cp = 4R<br />

and hence the energy needed to cool the w<strong>at</strong>er is 0.1 × 4 7 7 or about 0.8 J per<br />

2<br />

second. This is an additional 0.1 J/s compared to the dry case.<br />

In addition, we need to condense 0.0013 moles of w<strong>at</strong>er per second. Since one<br />

mole of w<strong>at</strong>er is 18g, this is about 0.023 g per second, which requires 2.26 kJ<br />

per gram, or 52 J per second. This is by far the largest extra cost. Using the<br />

efficiency we find th<strong>at</strong> in this case the refriger<strong>at</strong>or needs an extra 870 Joules per<br />

second!<br />

Problem 21.<br />

Part A.


C.2. SOLUTIONS FOR CHAPTER 2. 331<br />

An amount of he<strong>at</strong> Q flows from the solid into the Carnot engine, an amount<br />

Q’ flows from the Carnot engine into the room temper<strong>at</strong>ure reservoir, and an<br />

amount of work W is done on the Carnot refriger<strong>at</strong>or. The efficiency is η =<br />

−W<br />

−Q ′ = TR−T<br />

. This gives<br />

TR<br />

Q = Q ′ − W =<br />

( )<br />

1<br />

− 1 W =<br />

η<br />

T<br />

TR − T W<br />

The amount of he<strong>at</strong> one needs to extract to change the temper<strong>at</strong>ure from T to<br />

T − ∆T is C∆T = AT 3 ∆T and hence the amount of work th<strong>at</strong> needs to be<br />

done to do this is<br />

∆W = (TR − T ) AT 2 ∆T<br />

Because ∆T is defined in the neg<strong>at</strong>ive T direction we find therefore<br />

dW<br />

dT = − (TR − T ) AT 2<br />

and hence the amount of work needed to cool the solid to zero temper<strong>at</strong>ure is<br />

Part B.<br />

and hence<br />

Part C.<br />

Part D.<br />

W =<br />

∫ 0<br />

TR<br />

( )<br />

∂S<br />

∂T V<br />

∆U =<br />

∆S =<br />

∫ 0<br />

TR<br />

− (TR − T ) AT 2 dT = A<br />

12 T 4 R<br />

= 1<br />

T<br />

∫ 0<br />

TR<br />

CdT =<br />

( )<br />

∂U<br />

∂T V<br />

= C<br />

T<br />

= AT 2<br />

AT 2 dT = − 1<br />

3 AT 3 R<br />

∫ 0<br />

TR<br />

AT 3 dT = − 1<br />

4 AT 4 R<br />

∆UR = −∆U + W = 1<br />

3 AT 4 R<br />

∆SR = −∆S = 1<br />

3 AT 3 R<br />

Note th<strong>at</strong> indeed ∆UR = TR∆SR.<br />

Problem 22.<br />

Part A.


332 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Part B.<br />

( ) ( ) 2 ∂S<br />

∂ F<br />

= −<br />

=<br />

∂V T,N ∂T ∂V N<br />

and hence<br />

∆Q = T<br />

( )<br />

∂S<br />

∆V<br />

∂V T,N<br />

( )<br />

∂p<br />

=<br />

∂T V,N<br />

N<br />

V R<br />

(<br />

1 + B(T ) N<br />

)<br />

+<br />

V<br />

N<br />

V RT B′ (T ) N<br />

V<br />

∆Q = p∆V +<br />

N 2<br />

V 2 RT 2 B ′ (T )<br />

Part C.<br />

The first term is the normal work done on the environment, the second term<br />

must therefore represent a change in internal energy!<br />

Problem 23.<br />

Suppose th<strong>at</strong> during the process the pressure and temper<strong>at</strong>ure in the tank are<br />

p and T . A small amount of air, ∆N moles, goes from the line into the tank.<br />

This air expands to pressure p and temper<strong>at</strong>ure T” and then exchanges he<strong>at</strong> to<br />

get to a common temper<strong>at</strong>ure.<br />

The initial internal energy of the air in the tank is N1u1, the final internal energy<br />

of the air in the tank is N2u2, where we have p1V = N1RT1 and p2V = N2RT2.<br />

There is an amount of he<strong>at</strong> Q added to the gas, which is supplied by the wall.<br />

The amount of m<strong>at</strong>erial added to the tank is N2 − N1 and by adding this<br />

m<strong>at</strong>erial we added energy to the system. Because the air is taken out of the<br />

line <strong>at</strong> constant pressure the amount of energy available to add to the tank<br />

is (N2 − N2)h ′ where h ′ is the molar enthalpy h ′ = u ′ + p ′ v ′ and v ′ is the molar<br />

volume in the compressed air line, p ′ v ′ = RT ′ . Hence<br />

N2u2 = N1u1 + Q + (N2 − N1)h ′<br />

which gives the rel<strong>at</strong>ion in the problem.<br />

This gives<br />

and<br />

Using cp = cV + R we find<br />

N2(u2 − u ′ − RT ′ ) = N1(u1 − u ′ − RT ′ ) + Q<br />

u2 − u ′ = cV (T2 − T ′ )<br />

N2(cV T2 − cpT ′ ) = N1(cV T1 − cpT ′ ) + Q


C.2. SOLUTIONS FOR CHAPTER 2. 333<br />

The amount of he<strong>at</strong> is supplied by the wall, but since the temper<strong>at</strong>ure of the<br />

wall increases this is a neg<strong>at</strong>ive number, with<br />

Q = −mtcs(T2 − T1)<br />

Using the equ<strong>at</strong>ion of st<strong>at</strong>e we arrive <strong>at</strong> an equ<strong>at</strong>ion for T2:<br />

p2<br />

T2<br />

V (cV T2 − cpT ′ ) = p1<br />

V (cV T1 − cpT ′ ) − mtcsR(T2 − T1)<br />

T2<br />

T1<br />

′<br />

′<br />

T T<br />

V (cV p2 − cpp2) = V (cV p1 − cpp1) − mtcsRT ′ ( T2<br />

T<br />

define x1 = T1<br />

T ′ and x2 = T2<br />

T ′ and we get<br />

T1<br />

)<br />

T ′<br />

′ − T1<br />

V (cV p2 − 1<br />

cpp2) = V (cV p1 − 1<br />

cpp1) − mtcsRT ′ (x2 − x1)<br />

x2<br />

cV (p2 − p1) + cp( p1<br />

(cV − cp)(p2 − p1) + cp(p1( 1<br />

x1<br />

x1<br />

x1<br />

− p2 RT<br />

) + mtcs<br />

x2<br />

′<br />

V (x2 − x1) = 0<br />

− 1) − p2( 1<br />

x2<br />

RT<br />

− 1)) + mtcs<br />

′<br />

V (x2 − x1) = 0<br />

We will get an approxim<strong>at</strong>e solution (exact is possible, too, since this is a second<br />

order equ<strong>at</strong>ion). Assume th<strong>at</strong> x1 = 1 + δ1, x2 = 1 + δ2, are close to 1, and keep<br />

only linear terms.<br />

RT<br />

(cV − cp)(p2 − p1) + cp(−p1δ1 + p2δ2) + mtcs<br />

′<br />

V (δ2 − δ1) = 0<br />

T<br />

δ2(mtcs<br />

′ 7<br />

+<br />

V 2 p2)<br />

T<br />

≈ δ1(mtcs<br />

′ 7<br />

+<br />

V 2 p1) + (p2 − p1)<br />

Divide by p ′ and define y1 = p1<br />

p ′ , y2 = p2<br />

p ′ :<br />

T<br />

δ2(mtcs<br />

′<br />

p ′ 7<br />

+<br />

V 2 y2)<br />

T<br />

≈ δ1(mtcs<br />

′<br />

p ′ V<br />

+ 7<br />

2 y1) + (y2 − y1)<br />

This expresses everything in dimensionless quantities. The quantity mtcs p ′ V is<br />

about 12. Hence we have<br />

14δ2 ≈ δ113 + 0.4<br />

which gives T2 about 304K or 32C and N2 about three times N1.<br />

Problem 24.<br />

Part A.<br />

T ′


334 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

where m is the mass density, M=mV.<br />

Part B.<br />

and hence<br />

F (T, V, N) = Fideal(T, V, N) − V 2 3m2 G<br />

5R<br />

V = 4π<br />

3 R3<br />

F (T, V, N) = Fideal(T, V, N) − 3m2 G<br />

5<br />

( ) 1<br />

3 4π<br />

V<br />

3<br />

5<br />

3<br />

( )<br />

∂F<br />

p = −<br />

=<br />

∂V T,N<br />

NRT<br />

V + m2 ( ) 1<br />

3 4π<br />

G V<br />

3<br />

2<br />

3<br />

κT = − 1<br />

V<br />

( )<br />

∂V<br />

∂p T,N<br />

( )<br />

∂p<br />

= −<br />

∂V T,N<br />

NRT<br />

V 2 + 2m2G 3<br />

κT = −<br />

V<br />

( ∂p<br />

∂V<br />

1<br />

)<br />

T,N<br />

Part C.<br />

stability needs positive κ and hence<br />

=<br />

NRT ≥ 2m2 G<br />

3<br />

NRT<br />

V<br />

( ) 1<br />

3 4π<br />

1 −<br />

V 3<br />

3<br />

1<br />

− 2m2 G<br />

3<br />

( ) 1<br />

3 4π<br />

V<br />

3<br />

5<br />

3<br />

This is not obeyed <strong>at</strong> low temper<strong>at</strong>ure of large volume!<br />

Problem 25.<br />

We have per mole th<strong>at</strong><br />

which gives<br />

or<br />

s(T ) =<br />

∫ 300<br />

0<br />

γT<br />

T<br />

( ) 1<br />

4π 2<br />

3<br />

3 V 3<br />

hs<br />

dT +<br />

300 +<br />

∫ 600<br />

cp<br />

300 T dT<br />

s(600K) = 300γ + hs<br />

300 + cp ln(2)


C.2. SOLUTIONS FOR CHAPTER 2. 335<br />

5.5cal/degree/mole = 1.5cal/degree/mole + 0.5cal/degree/mole + cp/ln(2)<br />

cp ≈ 5cal/degree/mole<br />

The molar specific he<strong>at</strong> cp for a mono-<strong>at</strong>omic gas is 5<br />

2<br />

R and for a di<strong>at</strong>omic gas<br />

<strong>at</strong> least 7<br />

2R, since rot<strong>at</strong>ional degrees of freedom are active <strong>at</strong> 600K. Hence the<br />

gas is mono-<strong>at</strong>omic.<br />

Problem 26.<br />

A substance has the following properties:<br />

(I) The amount of m<strong>at</strong>erial is fixed <strong>at</strong> N moles.<br />

(II) At constant temper<strong>at</strong>ure T0 the work W done by it on expansion from a<br />

volume V0 to a volume V is<br />

(III) The entropy is given by<br />

W = NRT0 ln( V<br />

S = NR V0<br />

V<br />

( T<br />

T0<br />

V0<br />

)<br />

) x<br />

In these expressions T0, V0, and x are fixed constants.<br />

(A) Calcul<strong>at</strong>e the Helmholtz free energy of the system.<br />

dF = −SdT − pdV<br />

F (T0, V ) − F (T0, V0) = −W = −NRT0 ln( V<br />

∫ T<br />

F (T, V ) − F (T0, V ) = −<br />

F (T0, V ) = F0 − NRT0 ln( V<br />

T0<br />

SdT ′ ∫ T<br />

= −<br />

T0<br />

V0<br />

)<br />

NR V0<br />

V<br />

V0<br />

)<br />

( T ′<br />

T0<br />

) x<br />

dT ′<br />

( ( ) )<br />

x+1<br />

V0 1 T<br />

F (T, V ) = F (T0, V ) − NRT0<br />

− 1<br />

V x + 1 T0<br />

F (T, V ) = F0 − NRT0 ln( V<br />

V0<br />

( ( ) )<br />

x+1<br />

V0 1 T<br />

) − NRT0<br />

− 1<br />

V x + 1 T0


336 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

(B) Find the equ<strong>at</strong>ion of st<strong>at</strong>e for the pressure p in terms of V , T , and N.<br />

( )<br />

∂F<br />

p = − =<br />

∂V T<br />

NRT0<br />

V<br />

NRT0 V0<br />

−<br />

V V<br />

( ( ) )<br />

x+1<br />

1 T<br />

− 1<br />

x + 1 T0<br />

(C) Find the work done on expansion <strong>at</strong> constant T, where T is arbitrary.<br />

W =<br />

∫ V<br />

V0<br />

Problem 27.<br />

pdV = NRT0 ln( V<br />

V0<br />

)+NRT0<br />

( V0<br />

V<br />

The he<strong>at</strong> capacity <strong>at</strong> constant pressure is given by<br />

( )<br />

∂S<br />

Cp = T<br />

∂T<br />

and the coefficient of thermal expansion is given by<br />

α = 1<br />

( )<br />

∂V<br />

V ∂T<br />

The amount of m<strong>at</strong>erial N is always kept constant.<br />

If we consider S(T, p) we have<br />

( ) ( )<br />

∂S ∂S<br />

dS = dT + dp<br />

∂T p ∂p T<br />

(A) Show th<strong>at</strong> this leads to<br />

T dS = T<br />

T dS = CpdT − αT V dp<br />

p<br />

p<br />

( )<br />

∂S<br />

dT + T<br />

∂T p<br />

T dS = CpdT + T<br />

) ( ( ) )<br />

x+1<br />

1 T<br />

− 1<br />

− 1<br />

x + 1 T0<br />

( )<br />

∂S<br />

dp<br />

∂p T<br />

( )<br />

∂S<br />

dp<br />

∂p T<br />

( ) ( ) ( )<br />

2 ∂S ∂ G ∂V<br />

= −<br />

− = −αV<br />

∂p T ∂T ∂p = ∂T p<br />

T dS = CpdT − T αV dp


C.2. SOLUTIONS FOR CHAPTER 2. 337<br />

One can show th<strong>at</strong> (but you do NOT have to!)<br />

where X is a finite constant.<br />

αV<br />

lim<br />

T →0 Cp<br />

= X<br />

(B) Show th<strong>at</strong> this has the consequence th<strong>at</strong> in an adiab<strong>at</strong>ic process one can<br />

never reach T = 0 by a finite change in pressure.<br />

At very low temper<strong>at</strong>ures<br />

CpdT = αV T dp<br />

dT = αV<br />

T dp<br />

dT<br />

T<br />

Cp<br />

= Xdp<br />

ln(T ) − ln(T ′ ) = X(p − p ′ )<br />

Therefore, if T → 0 the pressure difference goes to infinity.<br />

Problem 28.<br />

A container of volume V0 is divided into two equal parts by a removable wall.<br />

There are N0 moles of gas in equilibrium <strong>at</strong> temper<strong>at</strong>ure Ti on the left, the right<br />

side is empty. The container is isol<strong>at</strong>ed, no he<strong>at</strong> goes through. The internal<br />

wall is suddenly removed and the gas can expand freely until it finally fills up<br />

the whole container.<br />

(A) Is the gas in equilibrium just after the wall has been removed?<br />

No, it is not, it wants to expand and fill up all space.<br />

(B) Is the free expansion a reversible process?<br />

No, the process is not slow, the beginning is not in equilibrium, and only<br />

the final st<strong>at</strong>e is in equilibrium.<br />

(C) Does the entropy of the gas increase, decrease, or stay the same?<br />

The entropy increases (maximum entropy principle)<br />

(D) Does the energy of the gas increase, decrease, or stay the same?<br />

Stay the same, because there is no he<strong>at</strong> exchange, and no work done on<br />

the outside world.


338 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Assume th<strong>at</strong> we have a different, reversible process th<strong>at</strong> brings a gas <strong>at</strong> constant<br />

internal energy from a temper<strong>at</strong>ure Ti and volume 1<br />

2V0 to a volume V0 and<br />

temper<strong>at</strong>ure Tf . At each stage of this reversible process the gas is described by<br />

an internal energy function of the form U = u(T, V, N).<br />

(E) Is this reversible process adiab<strong>at</strong>ic?<br />

No, it is not. The internal energy is constant, but now work is done and<br />

he<strong>at</strong> has to flow in.<br />

(F) Express the response function ( )<br />

∂T<br />

∂V in terms of partial deriv<strong>at</strong>ives of<br />

U,N<br />

the energy function u(T, V, N).<br />

There are several approaches, for example, use the product rule:<br />

( ) ( )<br />

∂T<br />

∂T<br />

= −<br />

∂V U,N ∂U V,N<br />

( )<br />

∂U<br />

= −<br />

∂V T,N<br />

[ (∂U )<br />

∂T<br />

V,N<br />

] −1 ( )<br />

∂U<br />

∂V T,N<br />

(G) A mono-<strong>at</strong>omic ideal gas has u(T, V, N) = 3<br />

2NRT . How much does the<br />

temper<strong>at</strong>ure change in a free expansion from volume 1<br />

2V0 to V0?<br />

For the equivalent reversible process we have ( )<br />

∂U<br />

∂V = 0, and hence<br />

( )<br />

T,N<br />

∂T<br />

∂V = 0, which means th<strong>at</strong> the temper<strong>at</strong>ure T does not change. Since<br />

U,N<br />

the equivalent process brings us to the same final st<strong>at</strong>e, the free expansion<br />

also has no temper<strong>at</strong>ure change.<br />

Problem 29.<br />

Under adiab<strong>at</strong>ic compression a gas changes its temper<strong>at</strong>ure when the pressure is<br />

changed, ( in) a reversible manner. The response function describing this situ<strong>at</strong>ion<br />

∂T is Ω = ∂p<br />

S,N .<br />

( )<br />

T V α<br />

1 ∂V<br />

Show th<strong>at</strong> Ω = with α = + Cp V ∂T p,N and Cp the he<strong>at</strong> capacity <strong>at</strong> constant<br />

pressure.<br />

Note the sign error in the problem, α does not have a minus sign. Using the<br />

product rule we find:<br />

( ) ( ) ( )<br />

∂T<br />

∂T ∂S<br />

Ω =<br />

= −<br />

∂p S,N ∂S p,N ∂p T,N<br />

( )<br />

∂S<br />

Cp = T<br />

∂T p,N<br />

( ) ( ) ( )<br />

2 ∂S<br />

∂ G<br />

∂V<br />

= −<br />

= − = −V α<br />

∂p T,N ∂T ∂p N ∂T p,N


C.2. SOLUTIONS FOR CHAPTER 2. 339<br />

Combining this gives the required result.<br />

For an ideal mono-<strong>at</strong>omic gas αT = 1 and Cp = 5<br />

2NR. Find the function f<br />

defined by T = f(p) in this adiab<strong>at</strong>ic process.<br />

( )<br />

∂T<br />

=<br />

∂p<br />

T V α<br />

= 2V<br />

5NR<br />

Using the ideal gas law:<br />

This can be integr<strong>at</strong>ed to give:<br />

S,N<br />

Cp<br />

pV = NRT<br />

( )<br />

∂T<br />

=<br />

∂p S,N<br />

2pV 2NRT<br />

=<br />

5pNR 5pNR<br />

5 log(T ) = 2 log(p) + C(S, N)<br />

(<br />

p<br />

f(p) =<br />

p0<br />

) 2<br />

5<br />

where we changed the constant of integr<strong>at</strong>ion to p0.<br />

Problem 30.<br />

The Jacobian J is defined by<br />

J(x, y; u, v) =<br />

Show th<strong>at</strong> J(T, S; p, V ) = 1.<br />

Three approaches:<br />

use<br />

which gives<br />

J(T, S; p, V ) =<br />

use<br />

J(T, S; p, V ) =<br />

∂(x, y)<br />

∂(u, v) =<br />

( )<br />

∂x<br />

∂u v<br />

∂(T, S)<br />

∂(p, V ) =<br />

( )<br />

∂T<br />

∂p V<br />

( )<br />

∂S<br />

=<br />

∂V T<br />

( )<br />

∂T<br />

∂p V<br />

( )<br />

∂S<br />

+<br />

∂V p<br />

( ) ( )<br />

∂S ∂T<br />

−<br />

∂V T ∂p V<br />

( )<br />

∂y<br />

−<br />

∂v u<br />

= 2T<br />

5p<br />

( )<br />

∂x<br />

∂v u<br />

( ) ( )<br />

∂S ∂T<br />

−<br />

∂V p ∂V p<br />

( )<br />

∂S<br />

( )<br />

∂p<br />

∂p V ∂V T<br />

( )<br />

∂y<br />

∂u v<br />

( )<br />

∂S<br />

∂p V<br />

( ) ( ) ( )<br />

∂S ∂p ∂T<br />

−<br />

∂p V ∂V T ∂V p<br />

( )<br />

∂S<br />

∂p V


340 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

and get<br />

Mr. Maxwell gives:<br />

and hence<br />

A second route:<br />

which gives<br />

( ) ( ) ( )<br />

∂T ∂p ∂T<br />

= −<br />

∂p V ∂V T ∂V p<br />

J(T, S; p, V ) =<br />

( )<br />

∂T<br />

∂p V<br />

( ) ( ) 2 ∂S<br />

∂ F<br />

= − =<br />

∂V T ∂T ∂V<br />

J(T, S; p, V ) =<br />

J(T, S; p, V ) =<br />

J(T, S; p, V ) =<br />

( )<br />

∂T<br />

∂p V<br />

( )<br />

∂S<br />

∂V T<br />

( )<br />

∂p<br />

∂T V<br />

( )<br />

∂p<br />

∂T V<br />

= 1<br />

∂(T, S) ∂(T, S) ∂(T, V )<br />

=<br />

∂(p, V ) ∂(T, V ) ∂(p, V )<br />

( )<br />

∂S<br />

( )<br />

∂T<br />

∂V T ∂p V<br />

where we note th<strong>at</strong> the following partial deriv<strong>at</strong>ives are zero:<br />

( )<br />

∂T<br />

= 0<br />

∂V T<br />

Mr. Maxwell applies again, and the solution follows the end of the previous one.<br />

Third route:<br />

Take a single area in the p-V plane enclosed by a line L.<br />

∫ ∫<br />

dpdV = W<br />

A<br />

which is the work done in a cyclic process along L.<br />

Each point in A corresponds to a point (T, S) and hence maps A into A’<br />

∫ ∫<br />

A ′<br />

dT dS = Q<br />

is the he<strong>at</strong> into the system in the same process. Since the energy did not change<br />

we have W = Q, or<br />

∫ ∫ ∫ ∫<br />

dpdV =<br />

A<br />

A ′<br />

dT dS


C.2. SOLUTIONS FOR CHAPTER 2. 341<br />

for arbitrary areas A and corresponding areas A’. But the definition of the<br />

Jacobian tells us<br />

∫ ∫<br />

∫ ∫<br />

J(T, S; p, V )dpdV =<br />

A<br />

A ′<br />

dT dS<br />

and hence we have<br />

∫ ∫<br />

∫ ∫<br />

J(T, S; p, V )dpdV = dpdV<br />

A<br />

A<br />

for arbitrary areas A. This can only be true if the Jacobian is one.<br />

Problem 31.<br />

A fluid undergoes an adiab<strong>at</strong>ic throttling process. This means th<strong>at</strong> the fluid is<br />

pressed through a porous plug, with initial pressure p1 and final pressure p2.<br />

(A) Show th<strong>at</strong> the enthalpy of the fluid is constant in this process.<br />

As a result of this process, the temper<strong>at</strong>ure changes (remember when you pump<br />

air in your bicycle tire, or when you let the air escape?). The amount of change<br />

is measured by the Joule-Thomson coefficient ηJT , defined by<br />

( )<br />

∂T<br />

ηJT =<br />

∂p<br />

(B) Show th<strong>at</strong> ηJT = V (T α − 1)<br />

Cp<br />

(C) Wh<strong>at</strong> is the value for an ideal gas?<br />

Consider an amount of m<strong>at</strong>erial N taken from region 1 to region 2.<br />

The initial volume is V1, the final volume V2. The equipment th<strong>at</strong> keeps the<br />

pressure constant has to do work when these volumes change. The work done<br />

by this equipment on region 1 is p1V1, while the work done in region 2 is −p2V2<br />

(neg<strong>at</strong>ive!). There is no he<strong>at</strong> exchange, and the energy of the amount of m<strong>at</strong>erial<br />

changes from U1 to U2. Energy conserv<strong>at</strong>ion yields:<br />

H<br />

U2 − U1 = p1V1 − p2V2<br />

from which we get H1 = U1 + p1V1 = H2.<br />

( )<br />

∂T<br />

ηJT =<br />

∂p<br />

We always keep N constant, and leave out th<strong>at</strong> variable for the remainder of the<br />

problem. There are many ways to proceed. For example, we use the fact th<strong>at</strong><br />

deriv<strong>at</strong>ives <strong>at</strong> constant H are a problem, and write<br />

H,N


342 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

( ) ( )<br />

∂T ∂T<br />

= −<br />

∂p H ∂H p<br />

We have, with dH = T dS + V dp,<br />

which gives<br />

Cp = T<br />

( )<br />

∂H<br />

= −<br />

∂p T<br />

( )<br />

∂S<br />

=<br />

∂T p<br />

ηJT = − 1<br />

Again, from dH = T dS + V dp we find<br />

( )<br />

∂H<br />

= T<br />

∂p<br />

T<br />

Cp<br />

[ (∂H ) ] −1 (∂H )<br />

∂T<br />

( )<br />

∂H<br />

∂T p<br />

( )<br />

∂H<br />

∂p T<br />

( )<br />

∂S<br />

+ V<br />

∂p T<br />

Mr. Maxwell comes in handy again:<br />

( ) ( ) ( )<br />

2 ∂S ∂ G ∂V<br />

= − = − = −V α<br />

∂p T ∂T ∂p ∂T p<br />

and hence<br />

ηJT = − 1<br />

Cp<br />

(V − T V α)<br />

For an ideal gas we have αT = 1 and hence ηJT = 0. An ideal gas escaping<br />

from a tyre would not cause any cooling, or pumping an ideal gas into a tyre<br />

would not cause any he<strong>at</strong>ing!<br />

If we use<br />

(virial expansion), we find th<strong>at</strong><br />

pV α<br />

NRT<br />

pV = NRT (1 + N<br />

V B2(T ))<br />

pV<br />

NRT<br />

If we write α = 1<br />

T (1 + ∆α) we find<br />

= 1 + N<br />

V B2(T )<br />

pV<br />

− = −αN<br />

NRT 2 V B2(T ) + N<br />

V B′ 2(T )<br />

pV<br />

∆α = −(1 + ∆α)N<br />

NRT V B2(T ) + T N<br />

V B′ 2(T )<br />

(1 + N<br />

V B2(T ))∆α = −(1 + ∆α) N<br />

V B2(T ) + T N<br />

V B′ 2(T )<br />

p<br />

∂p<br />

T


C.3. SOLUTIONS FOR CHAPTER 3. 343<br />

In the low density limit we only need terms in first order:<br />

∆α ≈ − N<br />

V B2(T ) + T N<br />

V B′ 2(T ) = N<br />

V (T B′ 2(T ) − B2(T ))<br />

and by measuring the Joule-Thomson coefficient we can obtain the second virial<br />

coefficient<br />

C.3 Solutions for chapter 3.<br />

Problem 1.<br />

Note:<br />

p<br />

RT<br />

( )<br />

∂p<br />

∂T V N<br />

= N<br />

V<br />

= RN<br />

V<br />

(1 + R<br />

this can be neg<strong>at</strong>ive, if 2T < T0 − p0V<br />

NR .<br />

We have<br />

( )<br />

∂p<br />

∂V T N<br />

= RT (− N<br />

V<br />

p0<br />

(1 + R<br />

(T − T0) N<br />

V )<br />

p0<br />

p0<br />

(2T − T0) N<br />

V )<br />

2<br />

R 2N<br />

− (T − T0) )<br />

2 V 3<br />

for mechanical stability this has to be neg<strong>at</strong>ive, which gives<br />

or<br />

or<br />

(− N R<br />

−<br />

V 2<br />

(1 + R<br />

p0<br />

p0<br />

(T − T0)<br />

2 2N<br />

) < 0<br />

V 3<br />

(T − T0) 2N<br />

) > 0<br />

V<br />

(T0 − T )2NR < V p0<br />

which is always true if T > T0 . In the stable region we also have<br />

T0 − p0 V<br />

R N < T0 − 2(T0 − T ) = 2T − T0 < 2T<br />

and hence the isotherms are not crossing in the region of mechanical stability.<br />

The line separ<strong>at</strong>ing the stable and unstable region follows from<br />

NR<br />

p0V =<br />

1<br />

, T < T0<br />

2(T0 − T )


344 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

and hence<br />

T<br />

p = p0 , T < T0<br />

4(T0 − T )<br />

This shows th<strong>at</strong> the instability region starts <strong>at</strong> V → 0, p → ∞, T ↑ T0 and runs<br />

to V = 2NRT0<br />

, p = 0, forT → 0.<br />

p0<br />

There are no meta-stable st<strong>at</strong>es.<br />

Problem 2.<br />

Part a.<br />

or<br />

p = NRT aN<br />

e− V RT<br />

V − Nb<br />

pV<br />

NRT =<br />

V<br />

aN<br />

e− V RT<br />

V − NB<br />

in the limit V → ∞ the right-hand side goes to one.<br />

Part b.<br />

or<br />

which is zero if<br />

( )<br />

∂p<br />

= −<br />

∂V T N<br />

NRT<br />

aN<br />

e− V RT +<br />

(V − Nb) 2 NRT aN<br />

e− V RT<br />

V − Nb<br />

( )<br />

∂p<br />

∂V T N<br />

1<br />

= p(−<br />

(V − Nb)<br />

aN<br />

V 2RT =<br />

1<br />

(V − Nb)<br />

For the second order deriv<strong>at</strong>ive we have<br />

+ aN<br />

V 2 RT )<br />

aN<br />

V 2 RT<br />

([A])<br />

( 2 ∂ p<br />

∂V 2<br />

) ( ) (<br />

∂p<br />

1 aN<br />

=<br />

− +<br />

T N ∂V T N (V − Nb) V 2 ) (<br />

1 aN<br />

+ p<br />

− 2<br />

RT (V − Nb) 2 V 3 )<br />

RT<br />

which gives <strong>at</strong> the point where the first deriv<strong>at</strong>ive is zero th<strong>at</strong> we also need<br />

Dividing A by B gives<br />

V − Nb = V<br />

2<br />

or Vc = 2Nb. Inserting this in A gives<br />

1 aN<br />

(<br />

= 2<br />

(V − Nb) 2 V 3 ) ([B])<br />

kT


C.3. SOLUTIONS FOR CHAPTER 3. 345<br />

aN<br />

4N 2 b 2 RT<br />

= 1<br />

Nb<br />

or RTc = a<br />

4b .<br />

Finally, from the equ<strong>at</strong>ion of st<strong>at</strong>e we get pc = a<br />

4b2 e−2 .<br />

Part c.<br />

Similar to van der Waals.<br />

Problem 3.<br />

part a.<br />

Spherical drops V = 4π<br />

3 R3 , A = 4πR2 , and hence dV = 4πR2dR, dA = 8πRdR<br />

and dA = 2<br />

RdV .<br />

This gives<br />

dU = T dS − (p − σ 2<br />

)dV + µdN<br />

R<br />

and peff = p − 2σ<br />

R .<br />

part b.<br />

Equilibrium gives pout = peff = pin − 2σ<br />

R .<br />

part c.<br />

µg(p, T ) = µl(p + 2σ<br />

, T )<br />

R<br />

where p is the pressure in the gas.<br />

part d.<br />

Left hand side:<br />

Right hand side<br />

Reducing this to the volume<br />

or<br />

N∆µg = ∆G = Vgdp − SgdT<br />

N∆µl = ∆G = Vl(1 − 2σ<br />

R 2<br />

N∆µl = ∆G = Vl(1 − 6σ<br />

VlR<br />

( )<br />

∂R<br />

)dp − SldT<br />

∂p T N<br />

( )<br />

∂Vl<br />

)dp − SldT<br />

∂p T N<br />

N∆µl = ∆G = Vl(1 + 6σκT<br />

)dp − SldT<br />

R<br />

where we introduced the isothermal compressibility of the liquid. Along the<br />

coexistence line the changes are the same and hence<br />

Vgdp − SgdT = Vl(1 + 6σκT<br />

R<br />

)dp − SldT


346 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

or<br />

which gives<br />

(Sg − Sl)dT = (Vg − Vl(1 + 6σκT<br />

R ))dp<br />

dpco<br />

dT =<br />

Sg − Sl<br />

Vg − Vl(1 + 6σκT<br />

R )<br />

part e.<br />

With changing p and R INDEPENDENTLY we have for the left hand side<br />

and for the right hand side<br />

which gives<br />

or pco increases when R decreases.<br />

Problem 4.<br />

N∆µg = ∆G = Vgdp<br />

N∆µl = ∆G = Vl(dp − 2σ<br />

dR)<br />

R2 2σ<br />

(Vg − Vl)dp = −Vl dR<br />

R2 part a.<br />

The variables th<strong>at</strong> can change are the temper<strong>at</strong>ure and the Helium partial pressure,<br />

two in total, and hence three phases can coexist.<br />

part b.<br />

The total pressure is p. Therefore we have<br />

µice(p, T ) = µw<strong>at</strong>er(p, T ) = µgas(p − pHe, T )<br />

If the first two are independent of p, this fixes T and hence the temper<strong>at</strong>ure of<br />

the triple point does not change. From the last we find th<strong>at</strong> the pressure <strong>at</strong> the<br />

triple point is simple increased by the Helium partial pressure.<br />

Problem 10.<br />

Part A.<br />

In case Tg ̸= Tl he<strong>at</strong> will flow from one to the other until thermal equilibrium is<br />

reached and the temper<strong>at</strong>ure is the same everywhere. If pg ̸= pl one phase will<br />

do work on the other by changing the volumes until equilibrium is reached for<br />

which the pressure is the same everywhere.<br />

Part B.


C.3. SOLUTIONS FOR CHAPTER 3. 347<br />

Vl = ∂Gl<br />

∂p<br />

Nl + Ng = N0<br />

Vl + Vg = V0<br />

pVg = NgRT<br />

= Nl<br />

∂µl<br />

(p, T )<br />

∂p<br />

Part C.<br />

Multiply the fourth equ<strong>at</strong>ion by p and substitute the third and fourth equ<strong>at</strong>ion<br />

in the second, leads to<br />

or after elimin<strong>at</strong>ing Ng<br />

Nl + Ng = N0<br />

Nlp ∂µl<br />

∂p + NgRT = pV0<br />

Nl(p ∂µl<br />

∂p − RT ) = pV0 − N0RT<br />

Part D.<br />

In order for the solution to be valid we need 0 ≤ Nl ≤ N0 and 0 ≤ Vl ≤ V0.<br />

Problem 11.<br />

In the vicinity of the triple point of a m<strong>at</strong>erial the vapor pressure of the liquid<br />

is given by log( p<br />

20T0<br />

) = 10 − p0 T . The vapor pressure of the solid is given by<br />

log( p<br />

25T0<br />

) = 15 − p0 T .<br />

(A) Wh<strong>at</strong> are the temper<strong>at</strong>ure and the pressure of the triple point in terms of<br />

T0 and p0?<br />

The two equ<strong>at</strong>ions above give the coexistence curves gas-liquid and gassolid.<br />

At the triple point they meet. Hence<br />

10 − 20T0<br />

Tt<br />

= 15 − 25T0<br />

Tt<br />

Tt = T0<br />

pt = p0e −10


348 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

(B) Using the Clausius-Clapeyron equ<strong>at</strong>ion, show th<strong>at</strong><br />

dpco<br />

dT<br />

d log(pco)<br />

dT<br />

= l<br />

T ∆v<br />

= dpco<br />

dT<br />

1<br />

pco<br />

d log(pco)<br />

dT = l<br />

pcoT ∆v .<br />

(C) Is the volume per mole of the gas larger or smaller than th<strong>at</strong> of the liquid,<br />

assuming th<strong>at</strong> T0 and p0 are both positive?<br />

For the gas liquid curve we have:<br />

dpco<br />

dT<br />

= 20T0<br />

T 2<br />

which is positive. Since we need he<strong>at</strong> to go from liquid to gas, this shows<br />

th<strong>at</strong> the molar volume of the gas is larger.<br />

(D) Assume th<strong>at</strong> the values of the l<strong>at</strong>ent he<strong>at</strong> are independent of temper<strong>at</strong>ure<br />

near the triple point, and th<strong>at</strong> the molar volume of the gas is much larger<br />

than the molar volume of the liquid and of the solid. Calcul<strong>at</strong>e the r<strong>at</strong>ios<br />

lSL : lLG : lSG, where S,L, and G stand for solid, liquid, and gas.<br />

At the triple point we have:<br />

But we also have:<br />

which gives:<br />

and hence the r<strong>at</strong>io is:<br />

Problem 12.<br />

lLG = Tt∆v dpco<br />

dT = 20vG t<br />

lSG = Tt∆v dpco<br />

dT = 25vG t<br />

lSG = lSL + lLG<br />

lSL = 5v G t<br />

T0<br />

Tt<br />

lSL : lLG : lSG = 1 : 4 : 5<br />

T0<br />

Tt<br />

T0<br />

Tt


C.3. SOLUTIONS FOR CHAPTER 3. 349<br />

Grüneisen parameter.<br />

In many solids we can approxim<strong>at</strong>e the entropy by S = Nf(T/TD) where the<br />

Debye temper<strong>at</strong>ure TD is a function of the molar volume only, TD = g( V<br />

N ). The<br />

Grüneisen parameter γg is defined by<br />

Show th<strong>at</strong><br />

γg = − V<br />

NTD<br />

γg =<br />

V α<br />

CV κT<br />

dTD<br />

d V<br />

N<br />

For many solids the Grüneisen parameter is almost constant. If we assume th<strong>at</strong><br />

it is a constant, show th<strong>at</strong> we have the equ<strong>at</strong>ion of st<strong>at</strong>e<br />

P V = γgU + Nh( V<br />

N )<br />

This is the Mie-Grüneisen equ<strong>at</strong>ion of st<strong>at</strong>e.<br />

α = 1<br />

V<br />

Combining everything<br />

or<br />

( )<br />

∂S<br />

CV = T<br />

∂T V,N<br />

= NT<br />

f<br />

TD<br />

′ ( T<br />

)<br />

TD<br />

κT = − 1<br />

( )<br />

∂V<br />

V ∂p T,N<br />

( )<br />

∂V<br />

∂T<br />

= 1<br />

( ) 2 ∂ G<br />

V ∂p∂T<br />

= − 1<br />

(<br />

∂S<br />

V ∂p<br />

p,N<br />

( ∂S<br />

∂V<br />

)<br />

( ∂V<br />

∂p<br />

α = − 1<br />

V<br />

T,N<br />

( )<br />

∂S<br />

∂V T,N<br />

= − NT<br />

T 2 f<br />

D<br />

′ ( T<br />

TD<br />

CV = NT<br />

α = κT<br />

TD<br />

NT<br />

V TD<br />

)<br />

T,N<br />

) dTD<br />

d V<br />

N<br />

N<br />

1<br />

N<br />

γgf ′ ( T<br />

= κT<br />

TD<br />

( ∂S<br />

∂V<br />

)<br />

)<br />

T,N<br />

T,N<br />

NT<br />

= γgf<br />

V TD<br />

′ ( T<br />

)<br />

TD<br />

f ′ ( T<br />

) =<br />

TD<br />

α NT V TD<br />

κT γg TD NT<br />

γg =<br />

V α<br />

CV κT<br />

)


350 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

From these equ<strong>at</strong>ions we find<br />

Integr<strong>at</strong>ion gives<br />

V α<br />

= γgCV<br />

κT<br />

( ) ( )<br />

∂S<br />

∂S<br />

V<br />

= γgT<br />

∂V T,N ∂T V,N<br />

( ) ( )<br />

∂S ∂U<br />

T<br />

=<br />

∂T V,N ∂T V,N<br />

( ) ( ) ( )<br />

2 ∂S<br />

∂ F ∂p<br />

= −<br />

=<br />

∂V T,N ∂T ∂V N ∂T<br />

( ) ( )<br />

∂p<br />

∂U<br />

V<br />

= γg<br />

∂T<br />

∂T<br />

V,N<br />

V p = γgU + I(V, N)<br />

V,N<br />

V,N<br />

But since I is extensive we can write it in the form Nh( V<br />

N ).<br />

Problem 13.<br />

Murnaghan equ<strong>at</strong>ion.<br />

The isothermal bulk modulus BT is given by BT = −V<br />

( )<br />

∂p<br />

∂V<br />

T<br />

= 1 . The bulk<br />

κT<br />

modulus can often be approxim<strong>at</strong>ed by BT (p, T ) = BT (0, T ) + βT p where βT is<br />

independent of pressure.<br />

Show th<strong>at</strong> this leads to the equ<strong>at</strong>ion of st<strong>at</strong>e<br />

p(V, T ) = BT<br />

( [V0<br />

(0, T )<br />

V<br />

βT<br />

] βT<br />

where V0 is the volume <strong>at</strong> p = 0 and temper<strong>at</strong>ure T.<br />

− 1<br />

( )<br />

∂p<br />

−V<br />

∂V T<br />

= BT (0, T ) + βT p<br />

The solution of the homogeneous equ<strong>at</strong>ion<br />

( )<br />

∂p<br />

−V<br />

∂V<br />

= βT p<br />

T<br />

)


C.3. SOLUTIONS FOR CHAPTER 3. 351<br />

is<br />

The general solution is therefore<br />

which leads after substitution to<br />

or<br />

After defining C = BT (0,T )<br />

V βT<br />

βT<br />

0<br />

Problem 14.<br />

p = CV −βT<br />

p = p0 + CV −βT<br />

βT (p − p0) = BT (0, T ) + βT p<br />

p0 = − BT (0, T )<br />

βT<br />

we get the required answer.<br />

Derive the virial coefficients Bj(T ) for the Van der Waals equ<strong>at</strong>ion of st<strong>at</strong>e.<br />

Comment on the sign of the second virial coefficient.<br />

Compare with<br />

and get<br />

p<br />

RT<br />

p<br />

RT<br />

(p + a<br />

p<br />

RT<br />

N 2<br />

= − a<br />

RT<br />

)(V − Nb) = NRT<br />

V 2<br />

= − a<br />

RT<br />

= − a<br />

RT<br />

p<br />

RT =<br />

N 2<br />

+<br />

V 2 N<br />

N 2 N<br />

+<br />

V 2 V<br />

N 2 N<br />

+<br />

V 2 V<br />

j=1<br />

V − Nb<br />

(<br />

1 − N<br />

V b<br />

∞∑<br />

j=0<br />

∞∑<br />

( ) j<br />

N<br />

Bj(T )<br />

V<br />

B1(T ) = 1<br />

B2(T ) = b − a<br />

RT<br />

) −1<br />

( ) j<br />

N<br />

b<br />

V<br />

j


352 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Bj(T ) = b j−1<br />

for j > 2<br />

The sign of the second coefficient is positive for RT > a 27<br />

b = 8 RTc. This indic<strong>at</strong>es<br />

th<strong>at</strong> <strong>at</strong> high temper<strong>at</strong>ure the pressure is higher than in an ideal gas, which is<br />

caused by the excluded volume term. If the temper<strong>at</strong>ure is low, the pressure<br />

is lower than in an ideal gas, due to the <strong>at</strong>tractive interactions. At very high<br />

temper<strong>at</strong>ures the thermal motion is so strong th<strong>at</strong> the <strong>at</strong>tractive interactions<br />

do not play a role anymore.<br />

Remember the Joule-Thomson coefficient?<br />

ηJT = N<br />

(T B ′ 2 − B2) = N<br />

Cp<br />

Cp<br />

( 2a<br />

RT<br />

− b)<br />

This is positive for T < 27<br />

4 Tc. Hence for low temper<strong>at</strong>ures the gas he<strong>at</strong>s up when<br />

compressed, but for high temper<strong>at</strong>ures actually cools down when compressed!<br />

Problem 15.<br />

The Gibbs free energy of pure m<strong>at</strong>erials A and B is given by NAµA(p, T ) and<br />

NBµB(p, T ). The entropy of mixing for m<strong>at</strong>erials A and B is −NAR log( NA<br />

NA+NB )−<br />

NBR log( NB<br />

NA+NB<br />

), while the interaction energy between A and B is given by<br />

A(T, p) NANB<br />

NA+NB . Show th<strong>at</strong> a system where we mix NA moles of A and NB<br />

moles of B is stable for 2RT > A and th<strong>at</strong> there are forbidden regions when<br />

2RT < A.<br />

G(p, T, NA, NB) = NAµA(p, T ) + NBµB(p, T ) + Gmix<br />

Gmix = A(T, p) NANB<br />

+<br />

NA + NB<br />

NART log( NA<br />

) + NBRT log(<br />

NA + NB<br />

NB<br />

)<br />

NA + NB<br />

The Gibbs energy per mole (divide by NA + NB) in terms of the concentr<strong>at</strong>ion<br />

c of B <strong>at</strong>oms ( c = NB ) is<br />

NA+NB<br />

g(p, T, c) = (1 − c)µA(p, T ) + cµB(p, T ) + A(T, p)c(1 − c)+<br />

cRT log(c) + (1 − c)RT log(1 − c)<br />

This has to be convex as a function of c.


C.4. SOLUTIONS FOR CHAPTER 4. 353<br />

This is positive if<br />

( 2 ∂ g<br />

∂c2 )<br />

= −2A(p, T ) + RT<br />

T,p<br />

1 1<br />

+ RT<br />

c 1 − c<br />

−2A(p, T ) + RT 1 1<br />

+ RT > 0<br />

c 1 − c<br />

2A(p, T )c(1 − c) < RT (1 − c) + RT C = RT<br />

Hence phase separ<strong>at</strong>ion occurs when<br />

RT < 2Ac(1 − c)<br />

The value of the right hand side is maximal for c = 1<br />

2 and unstable regions first<br />

form in the middle, when RT < 1<br />

2A. C.4 Solutions for chapter 4.<br />

Problem 1.<br />

∂f<br />

∂m = m(b(T ) + cm + dm2 ) with b(T ) = b0(T − T0). Extrema <strong>at</strong> m = 0 and<br />

m = m± = 1<br />

2d (−c ± √ c2 − 4db(T )).<br />

If T > T0 then b(T ) > 0 and m+ > m− > 0, as long as 4db(T ) < c2 , and then<br />

we have minima <strong>at</strong> m+ and 0. If T = T0 the extrema <strong>at</strong> m− and 0 coincide, and<br />

for T < T0 we have m+ > 0 > m− and now m=0 is a maximum. The energy<br />

differences are:<br />

f(m+, T ) − f(0, T ) = m 2 +( 1 1<br />

b(T ) +<br />

2 3 cm+ + 1<br />

4 dm2 +)<br />

Using b(T ) + cm+ + dm 2 + = 0 to elimin<strong>at</strong>e the linear term we get<br />

f(m+, T ) − f(0, T ) = m 2 +( 1 1<br />

b(T ) −<br />

6 12 dm2 +)<br />

which is neg<strong>at</strong>ive if 2b(T ) < dm2 + = −b(T ) − cm+, or 3b(T ) < −c 1<br />

2d (−c +<br />

√<br />

c2 − 4db(T ). This gives<br />

These two values are equal if<br />

6db(T )<br />

−c + c < √ c 2 − 4db(T )<br />

( 6db(T )<br />

−c + c)2 = c 2 − 4db(T )


354 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

and hence<br />

36d 2 b 2 (T )<br />

c 2<br />

− 12db(T ) = −4db(T )<br />

36d 2 b 2 (T ) = 8c 2 db(T )<br />

b(T ) = 2c2<br />

9d<br />

Tc = T0 + 2c2<br />

9db0<br />

This is larger than T0 and hence the phase transition is first order, the minimum<br />

<strong>at</strong> m+ becomes lower th<strong>at</strong> the minimum <strong>at</strong> 0.<br />

We also have<br />

which gives<br />

or<br />

f(m−, T ) − f(0, T ) = m 2 −( 1 1<br />

b(T ) −<br />

6 12 dm2−) f(m−, T ) − f(m+, T ) = 1<br />

6 b(T )(m2 − − m 2 +) − 1<br />

12 d(m4 − − m 4 +)<br />

f(m−, T ) − f(m+, T ) = 1<br />

12 (m2 − − m 2 +)(2b(T ) − d(m 2 − + m 2 +))<br />

Since 2b(T ) < dm 2 + we have<br />

2b(T ) − d(m 2 − + m 2 +) < −dm 2 − < 0<br />

and since m− − m+ = − 1<br />

√<br />

d c2 2 − 4db(T )0 we also have m− − m2 + = (m− −<br />

m+)(m− + m+) < 0. This gives f(m−, T ) − f(m+, T ) > 0 and hence if there is<br />

a minimum <strong>at</strong> m− it is a local minimum, and there is no other phase transition.<br />

Finally, for b(T ) = 2c2<br />

9d we have m+ = 1<br />

√<br />

2d (−c + c2 − 4 2c2<br />

9 ) or<br />

since<br />

√ c 2<br />

9<br />

Problem 2.<br />

= − 1<br />

3 c.<br />

m+(Tc) = −2c<br />

3d<br />

∂f<br />

∂m = b(p, T )m + c(p, T )m3 + dm 5 = 0<br />

gives the loc<strong>at</strong>ion of the extrema, b(p, T ) + c(p, T )m 2 0 + dm 4 0 = 0.


C.4. SOLUTIONS FOR CHAPTER 4. 355<br />

In case one, for c(p, T ) = c > 0, the d-term does not change wh<strong>at</strong> we had before,<br />

and the critical temper<strong>at</strong>ure is defined by b(p, Tc) = 0.<br />

In case two, for c(p, T ) = c < 0, we have the same situ<strong>at</strong>ion as in the first order<br />

transition discussed in the notes, with equ<strong>at</strong>ions<br />

This leads to<br />

and finally:<br />

f(m0, p, T ) − f(0, p, T ) = 0<br />

b(p, T ) + c(p, T )m 2 0 + dm 4 0 = 0<br />

1 1<br />

b(p, T ) +<br />

2 4 cm20 + 1<br />

6 dm40 = 0<br />

1 1<br />

b(p, T ) +<br />

2 2 cm20 + 1<br />

2 dm40 = 0<br />

m 2 0 = − 3c<br />

4d<br />

For part c we get th<strong>at</strong> for p > p0 the second order phase transition is always <strong>at</strong><br />

Tc = T0. For p < p0 we have m2 0 = 1<br />

2d (−c + √ c2 − 4db), or − c<br />

√<br />

1<br />

4d = 2d c2 − 4db,<br />

which gives<br />

−c = 2 √ c 2 − 4bd<br />

16bd = 3c 2<br />

16b0p(T − T0) = 3c0T 2 (p − p0) 2<br />

This clearly shows th<strong>at</strong> for p = p0 we have T = T0 and hence the curve Tc(p)<br />

for p < p0 connects smoothly to the curve Tc(p) = T0 for p > p0.<br />

In addition, when ∆p = p − p0 is small we have in lowest order<br />

or<br />

16b0p0∆T = 3c0T 2 0 (∆p) 2<br />

∆T<br />

∆p = 3c0T 2 0<br />

∆p<br />

16b0p0<br />

which shows th<strong>at</strong> the slope of the Tc(p) curve for p < p0 approaches 0 <strong>at</strong> the<br />

connection point, and hence also is continuous.<br />

Problem 3.<br />

We need to find the extrema and solve


356 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

and<br />

0 = ∂f<br />

= b(T )mx +<br />

∂mx<br />

1<br />

2 cmxm 2 y + dm 3 x<br />

0 = ∂f<br />

∂my<br />

= b(T )my + 1<br />

2 cmym 2 x + dm 3 y<br />

Clearly for b(T ) > 0 the only solution is mx = my = 0. For b(T ) < 0 the<br />

minimum follows from<br />

and<br />

b(T ) + 1<br />

2 cm2 y + dm 2 x = 0<br />

b(T ) + 1<br />

2 cm2 x + dm 2 y = 0<br />

Subtracting these two equ<strong>at</strong>ions yields:<br />

( 1<br />

2 c − d)(m2 x − m 2 y) = 0<br />

which shows th<strong>at</strong> for c ̸= 2d we get mx = my. These quantities then follow<br />

from<br />

b(T ) + ( 1<br />

2 c + d)m2 x = 0<br />

which is just wh<strong>at</strong> we had for a second order phase transition.<br />

Using the fact th<strong>at</strong> m 2 = m 2 x + m 2 y we find<br />

f(mx, my, T ) = a(T ) + 1<br />

2 b(T )m2 + 1<br />

4 dm4 + 1<br />

4 (c − 2d)m2 xm 2 y<br />

If c = 2d the last term disappears, and we can find m as before, but do not<br />

know the r<strong>at</strong>io of mx and my. In other words, in this case the two equ<strong>at</strong>ions<br />

we had before became dependent.<br />

Problem 4.<br />

( )<br />

∂f<br />

= m(b0(T − T0) + cm<br />

∂m T<br />

2 + dm 4 ) = 0<br />

has a solution m=0, which is a minimum for T > T0.<br />

b0(T − T0) + cm 2 + dm 4 = 0<br />

has a solution if c2 ≥ b0(T − T0)d or T ≤ T0 + c2<br />

b0d . The energy <strong>at</strong> this solution<br />

is the lowest if T < Tc with


C.4. SOLUTIONS FOR CHAPTER 4. 357<br />

b0(Tc − T0) = 3c2<br />

16d<br />

(see text, equ<strong>at</strong>ion 4.32 ). Therefore we have<br />

(a) T > T0 + c2<br />

b0d , m=0 only solution<br />

(b) T0 + c2<br />

b0d > T > T0 + 3c2<br />

16b0d<br />

two solutions, m=0 is stable.<br />

(c) T0 + 3c2<br />

16b0d > T > T0, two solutions, non-zero m stable.<br />

(d) T0 > T , one solution, non-zero.<br />

Hence this looks like the first order transition in the van der Waals model, where<br />

the local minima only exist for a short range of values.<br />

Problem 5.<br />

( )<br />

∂f<br />

= m<br />

∂m T<br />

3 (c0(T − T0) + dm 2 ) = 0<br />

has a solution m=0, which is a minimum for T > T0.<br />

c0(T − T0) + dm 2 = 0<br />

has a solution for T < T0 given by<br />

√<br />

c0(T0 − T )<br />

m =<br />

d<br />

This is again a second order phase transition. The magnetiz<strong>at</strong>ion is again proportional<br />

to the square root of the temper<strong>at</strong>ure difference.<br />

The specific he<strong>at</strong> per unit volume follows from c = −T d2f dT 2 and is<br />

For T < T0 we have<br />

or<br />

or<br />

f(T ) = a(T ) + m 4<br />

f(T ) = a(T ) + c2 0(T0 − T ) 2<br />

T > T0, c = −T d2 a<br />

dT 2<br />

d 2<br />

(<br />

1<br />

4 c0(T − T0) + 1<br />

6 dm2<br />

)<br />

(<br />

1<br />

4 c0(T − T0) + 1<br />

6 dc0(T0 )<br />

− T )<br />

d<br />

f(T ) = a(T ) + c3 0(T − T0) 3<br />

12d 2


358 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

which gives<br />

c = −T d2a dT 2 − T c30(T − T0)<br />

4d2 and in this case the specific he<strong>at</strong> is continuous <strong>at</strong> the phase transition (but its<br />

deriv<strong>at</strong>ive is not). Also, again in this case the specific he<strong>at</strong> has an extra positive<br />

term below the phase transition.<br />

The susceptibility follows from<br />

χ −1 =<br />

( 2 ∂ f<br />

∂m2 )<br />

= 3c0(T − T0)m<br />

T<br />

2 + 5dm 4<br />

which is zero for T > T0. This means th<strong>at</strong> for these temper<strong>at</strong>ures a very small<br />

field will cause a large magnetiz<strong>at</strong>ion!<br />

For T < T0 we have<br />

χ −1 = 2c2 0(T − T0) 2<br />

d<br />

and hence the susceptibility diverges strongly <strong>at</strong> the transition temper<strong>at</strong>ure.<br />

Problem 6.<br />

The equ<strong>at</strong>ion to solve is again<br />

We solve this by doing a Fourier expansion<br />

∫<br />

ϕ(r) =<br />

∆ϕ + αϕ = − h0<br />

δ(⃗r) = 0 (C.1)<br />

f<br />

d d kψ(k)e ı⃗ k·⃗r<br />

where we used the fact th<strong>at</strong> we have a spherical solution. The delta function is<br />

given by<br />

δ(⃗r) = 1<br />

(2π) d<br />

∫<br />

d d ke ı⃗k·⃗r and hence we get<br />

and<br />

ψ(k)(−| ⃗ k| 2 + α) = − h0<br />

f<br />

ϕ(r) = h0<br />

f(2π) d<br />

∫<br />

d d k<br />

1<br />

1<br />

(2π) d<br />

k 2 − α eı⃗ k·⃗r<br />

where we have to keep in mind th<strong>at</strong> α < 0. We set again α = −ξ −2 and get


C.4. SOLUTIONS FOR CHAPTER 4. 359<br />

ϕ(r) =<br />

2 h0ξ<br />

f(2π) d<br />

∫<br />

d d 1<br />

k<br />

ξ2k2 + 1 eı⃗ k·⃗r<br />

Introducing a new variable ⃗q = r ⃗ k we get (the integr<strong>at</strong>ions are always over the<br />

whole infinite space, hence the domain of the integr<strong>at</strong>ion does not change):<br />

or<br />

ϕ(r) =<br />

2 h0ξ<br />

f(2π) drd ∫<br />

2<br />

2−d h0ξ<br />

ϕ(r) = r<br />

f(2π) d<br />

∫<br />

d d 1<br />

q<br />

ξ2q2r −2 + 1 eı⃗q·ˆr<br />

d d 1<br />

q<br />

ξ2q2 eı⃗q·ˆr<br />

+ r2 This shows th<strong>at</strong> in the limit r → 0 ϕ is proportional to r2−d , because the r<br />

dependence in the integral disappears.<br />

In general, we get the solution by doing a contour integr<strong>at</strong>ion in the complex k<br />

plane, and we pick up contributions from the poles <strong>at</strong> q = ±ırξ−1 . This shows<br />

r −<br />

th<strong>at</strong> the exponential behavior is of the form e ξ indeed. The correl<strong>at</strong>ion length<br />

has the same form as in 3 dimensions.<br />

Problem 7.<br />

From the equ<strong>at</strong>ions<br />

we see th<strong>at</strong><br />

T > Tc : S(T ) = − da<br />

dT<br />

T < Tc : S(T ) = − da<br />

dT<br />

(C.2)<br />

1 db<br />

−<br />

2 dT m2 (T ) (C.3)<br />

∆S = 1 db<br />

2 dT (Tc)m 2 (Tc)<br />

and hence the l<strong>at</strong>ent he<strong>at</strong> follows immedi<strong>at</strong>ely. Note th<strong>at</strong> a measurement of the<br />

l<strong>at</strong>ent he<strong>at</strong> together with a measurement of the magnetic moment gives us the<br />

slope of the second order coefficient in this model!<br />

Problem 8.<br />

The angle between ⃗ S and h<strong>at</strong>n is θ and hence<br />

f(T ; S, θ) = a(T ) + 1<br />

2 b0(T − T0)S 2 + 1<br />

2 c0(T − 1<br />

2 T0)S 2 cos 2 (θ) + 1<br />

4 dS4<br />

Therefore, if T > 1<br />

2T0 the angle dependent term is minimal for θ = 1<br />

2π and ⃗ S<br />

will be orthogonal to ˆn. In th<strong>at</strong> case we have


360 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

f(T ; S, θ = 1<br />

1<br />

π) = a(T ) +<br />

2 2 b0(T − T0)S 2 + 1<br />

4 dS4<br />

which is as studied before. For T > T0 we have S=0, and for T < T0 we have a<br />

non-zero solution, and a second order phase transition.<br />

On the other hand, if T < 1<br />

2T0 the angle dependent term will be minimal if ⃗ S<br />

is parallel to ⃗n ( or anti-parallel), and the energy is<br />

(<br />

1<br />

f(T ; S, θ = 0, π) = a(T ) +<br />

2 b0(T − T0) + 1<br />

2 c0(T − 1<br />

2 T0)<br />

)<br />

S 2 + 1<br />

4 dS4<br />

which gives an even larger non-zero solution for S. Hence this system shows two<br />

second order phase transitions. At T0 from zero S to non-zero S orthogonal to<br />

ˆn, and <strong>at</strong> 1<br />

2 T0 the non-zero S switches to parallel to ˆn.<br />

Problem 9.<br />

( )<br />

∂f<br />

=<br />

∂m T<br />

2<br />

R(T +<br />

m0<br />

1<br />

8 T0) sinh( 2m<br />

Using sinh(2x) = 2 sinh(x) cosh(x) we get<br />

which is zero if<br />

or<br />

m0<br />

) − 1<br />

RT0 sinh(<br />

m0<br />

m<br />

)<br />

m0<br />

( )<br />

∂f<br />

=<br />

∂m T<br />

R<br />

(<br />

4(T +<br />

m0<br />

1<br />

8 T0) cosh( m<br />

)<br />

) − T0 sinh(<br />

m0<br />

m<br />

)<br />

m0<br />

The first condition gives m=0.<br />

The second condition gives<br />

sinh( m<br />

) = 0<br />

m0<br />

4(T + 1<br />

8 T0) cosh( m<br />

cosh( m<br />

which has a solution when T0 ≥ 8T .<br />

The second order deriv<strong>at</strong>ive is<br />

and for m=0 this is<br />

( 2 ∂ f<br />

∂m2 )<br />

=<br />

T<br />

4<br />

m2 R(T +<br />

0<br />

1<br />

m0<br />

m0<br />

) − T0 = 0<br />

) = 2T0<br />

T0 + 8T<br />

8 T0) cosh( 2m<br />

m0<br />

) − 1<br />

m2 RT0 cosh(<br />

0<br />

m<br />

)<br />

m0


C.4. SOLUTIONS FOR CHAPTER 4. 361<br />

( 2 ∂ f<br />

∂m2 )<br />

(m = 0) =<br />

T<br />

4<br />

m2 R(T +<br />

0<br />

1<br />

8 T0) − 1<br />

m2 RT0 =<br />

0<br />

4R<br />

m2 (T −<br />

0<br />

1<br />

8 T0)<br />

which is a maximum when 8T ≤ T0.<br />

Therefore, we have a critical temper<strong>at</strong>ure 1<br />

8 T0 and a second order phase transition.<br />

Problem 10.<br />

0 =<br />

( )<br />

∂f<br />

=<br />

∂m T<br />

3<br />

2 b0(T − T0) √ m + cm 3<br />

which has solutions m=0 or 3<br />

2b0(T − T0) + cm2√m = 0.<br />

The l<strong>at</strong>ter only occurs for T < T0.<br />

The second order deriv<strong>at</strong>ive is<br />

( 2 ∂ f<br />

∂m2 )<br />

=<br />

T<br />

3<br />

4 b0(T − T0) 1<br />

√ + 3cm<br />

m 2<br />

and diverges <strong>at</strong> m=0. The form near m=0 is simple, however. For T > T0 there<br />

is a minimum ( positive infinite slope ) and for T < T0 it is a maximum. At the<br />

transition both solutions are degener<strong>at</strong>e. Hence we have again a second order<br />

phase transition <strong>at</strong> T0.<br />

The magnetic moment is<br />

.<br />

Problem 11.<br />

m =<br />

(<br />

3b0<br />

2c (T0<br />

) 2<br />

5<br />

− T )<br />

Consider a model for a second order phase transition in a magnetic field. The<br />

Helmholtz free energy per unit volume is given by<br />

f(T ; m) = a(T ) + 1<br />

2 b(T )m2 + 1<br />

c(T )m4<br />

4<br />

Assume c(T ) = c > 0, a positive constant, and b(T ) = b0(T − Tc) with b0 > 0.<br />

The magnetiz<strong>at</strong>ion density is still considered to be an internal parameter, which<br />

should be chosen to minimize the Helmholtz free energy. The condition is now<br />

( )<br />

∂f<br />

= H<br />

∂m T<br />

Calcul<strong>at</strong>e m(T,H) (hint: use Maple or equivalent) and plot m(H) for various<br />

temper<strong>at</strong>ures and m(T) for various fields.


362 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

has one or three real solutions.<br />

General case:<br />

H = b(T )m + cm 3<br />

z 3 + a2z 2 + a1z + a0 = 0<br />

q = 1<br />

3 a1 − 1<br />

9 a2 2<br />

r = 1<br />

6 (a1a2 − 3a0) − 1<br />

27 a3 2<br />

One real root if q3 + r2 > 0, three if q3 + r2 < 0<br />

b(T )<br />

In our case a2 = 0, a1 = c , a0 = − H<br />

c .<br />

Hence one root if<br />

It is possible to find the roots:<br />

q =<br />

b(T )<br />

3c<br />

r = H<br />

2c<br />

( ) 3<br />

b(T )<br />

+<br />

3c<br />

( ) 2<br />

H<br />

> 0<br />

2c<br />

−(b(T )) 3 < 27<br />

4 H2 c<br />

(Tc − T ) 3 < 27<br />

4b3 H<br />

0<br />

2 c<br />

(<br />

s± = r ± (q 3 + r 2 ) 1<br />

2<br />

) 1<br />

3<br />

z1 = (s+ + s−) − 1<br />

3 a2<br />

z2 = − 1<br />

2 (s+ + s−) − 1<br />

3 a2<br />

√<br />

3<br />

+ ı<br />

2 (s+ − s−)<br />

z3 = − 1<br />

2 (s+ + s−) − 1<br />

3 a2<br />

√<br />

3<br />

− ı<br />

2 (s+ − s−)<br />

Hence when we have only one root, we get (with a2 = 0):


C.4. SOLUTIONS FOR CHAPTER 4. 363<br />

m =<br />

⎛<br />

⎝ H<br />

2c +<br />

√ (b(T )<br />

3c<br />

) 3<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

0.8<br />

0.6<br />

0.4<br />

1<br />

0.2<br />

0<br />

0<br />

1<br />

2<br />

T 3<br />

4<br />

5<br />

1<br />

2<br />

H<br />

Figure C.1: m versus T-H<br />

+<br />

( H<br />

2c<br />

When we have three roots we have<br />

and<br />

s± =<br />

⎛<br />

⎝ H<br />

2c<br />

This can be plotted via Maple.<br />

Problem 12.<br />

) 2<br />

⎞<br />

⎠<br />

1<br />

3<br />

+<br />

<br />

<br />

(<br />

± ı<br />

b(T )<br />

<br />

3c<br />

⎛<br />

3<br />

⎝ H<br />

2c −<br />

) 3<br />

z1 = 2ℜs+<br />

z2 = −ℜs+ + √ 3ℑs+<br />

z3 = −ℜs+ − √ 3ℑs+<br />

Consider the following model of a tri-critical point:<br />

4<br />

√ (b(T )<br />

3c<br />

( ) ⎞ 2<br />

H<br />

<br />

+ ⎠<br />

2c<br />

f(m, T, p) = 1<br />

2 a(T, p)m2 + 1<br />

4 b(T, p)m4 + 1<br />

6 cm6<br />

where c is a positive constant. The functions a and b are given by<br />

1<br />

3<br />

) 3<br />

+<br />

( H<br />

2c<br />

) 2<br />

⎞<br />

⎠<br />

1<br />

3


364 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

T − Tc p − pc<br />

a(T, p) = A + B<br />

Tc pc<br />

T − Tc p − pc<br />

b(T, p) = C + D<br />

Tc pc<br />

Depending on the p<strong>at</strong>h in p-T space, along p = f(T ) with pc = f(Tc), the<br />

behavior of the response functions and the order parameter takes on a different<br />

form. For example, the behavior of the magnetiz<strong>at</strong>ion m(p, T ) always varies<br />

like (Tc − T ) β , but the value of β depends on the p<strong>at</strong>h taken. The two distinct<br />

regions with different values of β are called critical and tri-critical.<br />

(A) Calcul<strong>at</strong>e the β this system.<br />

(B) Draw the phase diagram in p-T space.<br />

(C) Draw the critical and tri-critical regions in this phase diagram.<br />

(D) Give an explan<strong>at</strong>ion of these results, write it in such a way th<strong>at</strong> a new,<br />

incoming gradu<strong>at</strong>e student could understand it.<br />

( )<br />

∂f<br />

(m, T, p) = m<br />

∂m T,p<br />

( a(T, p) + b(T, p)m 2 + cm 4) = 0<br />

leads to either m = 0 or m 2 = 1<br />

2c<br />

(<br />

−b(T, p) ± √ b2 )<br />

(T, p) − 4a(T, p)c . This<br />

requires th<strong>at</strong> b2 (T, p) − 4a(T, p)c > 0.<br />

( 2 ∂ f<br />

∂m2 )<br />

(m, T, p) = a(T, p) + 3b(T, p)m<br />

T,p<br />

2 + 5cm 4<br />

indic<strong>at</strong>es th<strong>at</strong> m = 0 is a minimum when a(T, p) > 0. If this is the case both<br />

non-zero solutions for m 2 (if they exist) are positive only if b(T, p) < 0, and<br />

starting <strong>at</strong> the origin m = 0 we have a sequence minimum, maximum, minimum<br />

in f(m). If a(T, p) > 0 and b(T, p) > 0 the only solution is m = 0. If, on the<br />

other hand , a(T, p) < 0 then one of the solutions for m 2 is neg<strong>at</strong>ive, hence only<br />

one remains, and the sequence is maximum, minimum in f(m).<br />

The energy <strong>at</strong> the value m 2 = 1<br />

2c<br />

where we used<br />

(<br />

−b(T, p) + √ b 2 (T, p) − 4a(T, p)c<br />

f(m, T, p) = 1<br />

12 m2 ( 4a(T, p) + b(T, p)m 2)<br />

1<br />

6 a(T, p)m2 + 1<br />

6 b(T, p)m4 + 1<br />

6 cm6 = 0<br />

from the zero of the first deriv<strong>at</strong>ive. This is neg<strong>at</strong>ive when<br />

4a(T, p) + b(T, p)m 2 < 0<br />

)<br />

is


C.4. SOLUTIONS FOR CHAPTER 4. 365<br />

(when m 2 > 0 or b(T, p) < 0 ), which gives<br />

or<br />

8a(T, p)c − b 2 (T, p) + b(T, p) √ b 2 (T, p) − 4a(T, p)c < 0<br />

b(T, p) √ b 2 (T, p) − 4a(T, p)c < b 2 (T, p) − 8a(T, p)c<br />

Since b(T, p) < 0 this is always true if b 2 (T, p) − 8a(T, p)c > 0, which is always<br />

true if a(T, p) < 0. On the other hand, if b 2 (T, p) − 8a(T, p)c < 0 we need (note<br />

the change from < to > !!)<br />

or<br />

or<br />

b 2 (T, p) ( b 2 (T, p) − 4a(T, p)c ) > ( b 2 (T, p) − 8a(T, p)c ) 2<br />

12b 2 (T, p)a(T, p)c > 64a 2 (T, p)c 2<br />

3b 2 (T, p)a(T, p) > 16a 2 (T, p)c<br />

which is never true when a(T, p) < 0, as expected. If a(T, p) > 0 this leads to<br />

3b 2 (T, p) > 16a(T, p)c<br />

If this is true, certainly 4b 2 (T, p) > 16a(T, p)c and the minimum does exist.<br />

Also, this extends smoothly into the region b 2 (T, p) > 8a(T, p)c.<br />

In summary, if a(T, p) < 0 we always have a non-zero magnetiz<strong>at</strong>ion, and no<br />

local minimum <strong>at</strong> m = 0. If a(T, p) > 0 and b(T, p) > 0 the only solution is<br />

m = 0. If b(T, p) > 0 the line a(T, p) = 0 gives the line of second order phase<br />

transitions. If b(T, p) < 0 the line 3b 2 (T, p) = 16a(T, p)c gives a line of first<br />

order phase transitions.<br />

The analysis is now easiest in reduced coordin<strong>at</strong>es, π = p−pc<br />

pc<br />

which gives<br />

a(T, p) = Aθ + Bπ<br />

b(T, p) = Cθ + Dπ<br />

But we can make it even easier, if we use as coordin<strong>at</strong>es<br />

x = Aθ + Bπ<br />

y = Cθ + Dπ<br />

and θ =<br />

T −Tc<br />

, Tc<br />

In these coordin<strong>at</strong>es we have for x < 0 a non-zero magnetiz<strong>at</strong>ion with m=0 a<br />

maximum. For x > 0 and y > 0 we have m=0, and hence the line x = 0, y > 0<br />

gives a line of second order phase transitions. For x > 0 and y < 0 we have a


366 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Figure C.2: Figure 1.<br />

first order phase transition <strong>at</strong> x = 3<br />

16cy2 . In the region y < 0 and 0 < x < 3<br />

m=0 is a local minimum. See figure 1.<br />

The non-zero magnetiz<strong>at</strong>ion is given by the solutions of<br />

m 2 = 1<br />

(<br />

−b(T, p) +<br />

2c<br />

√ b2 )<br />

(T, p) − 4a(T, p)c<br />

16c y2<br />

If we approach the tri-critical point along b = 0 and a ↑ 0 we have m2 (<br />

=<br />

1 + √ )<br />

4|a(T, p)|c . Along b = 0 we have Dπ = −Cθ and hence |a| = |A −<br />

2c<br />

BC<br />

D ||θ| which gives m2 ∝ √ |θ|, or β = 1<br />

4 . On the other hand, if we approach<br />

the tri-critical point along a = 0 , b ↑ 0 we have m2 = |b|<br />

1<br />

c . This gives β = 2 .<br />

It is easy to show th<strong>at</strong> only along this straight line the critical exponent is 1<br />

2 ,<br />

any other line will have the a-term in the square root domin<strong>at</strong>ing near the tricritical<br />

point. If we approach the critical point along a parabolic line of the<br />

form b2 = 4acQ, with Q < 0, then the moment will look like<br />

m 2 = 1<br />

(<br />

√ )<br />

2c<br />

−b(T, p) + |b|<br />

Q − 1<br />

Q<br />

and now the exponent is always 1<br />

2 . This is equivalent to approaching the tricritical<br />

point along a = 0 and hence the critical region is really the line a = 0<br />

for b < 0.<br />

Problem 13.<br />

Consider the following model Helmholtz free energy per unit volume:<br />

f(T ; m) = a(T ) + e T 0 −T<br />

T 0 e −m2<br />

+ ce m2


C.4. SOLUTIONS FOR CHAPTER 4. 367<br />

where a(T ) is an arbitrary function of temper<strong>at</strong>ure, and T0 and c are positive<br />

constants. Show th<strong>at</strong> this system undergoes a second order phase transition,<br />

calcul<strong>at</strong>e the critical temper<strong>at</strong>ure Tc, and find the critical exponent β.<br />

Which is zero if m = 0 or<br />

( )<br />

∂f<br />

(<br />

= 2m −e<br />

∂m T<br />

T0−T T0 e −m2<br />

+ ce m2)<br />

T0 − T<br />

T0<br />

e T 0 −T<br />

T 0 e −m2<br />

= ce m2<br />

− m 2 = ln(c) + m 2<br />

2m 2 = T0 − T0 ln(c) − T<br />

T0<br />

Hence if T > T0(1 − ln(c)) there is only one solution and the system has zero<br />

magnetiz<strong>at</strong>ion. If T < T0(1 − ln(c)) there are three solutions, and because the<br />

energy goes to infinity for infinite m the solutions now are minimum, maximum,<br />

minimum. Since m = 0 is the middle solutions, it is now a maximum. The<br />

magnetiz<strong>at</strong>ion is therefore in th<strong>at</strong> case:<br />

m = ±<br />

√<br />

T0 − T0 ln(c) − T<br />

2T0<br />

which goes continuously to zero. The critical temper<strong>at</strong>ure is<br />

and the critical exponent is β = 1<br />

2 .<br />

Problem 14.<br />

Tc = T0(1 − ln(c))<br />

A magnetic system has a spontaneous magnetiz<strong>at</strong>ion below a critical temper<strong>at</strong>ure<br />

Tc, given by m2 = m2 0 Tc−T<br />

. The system is in a st<strong>at</strong>e <strong>at</strong> temper<strong>at</strong>ure<br />

Tc<br />

Ti = 2Tc and magnetiz<strong>at</strong>ion mi = 1<br />

2m0. This system is cooled down to very low<br />

temper<strong>at</strong>ure, but <strong>at</strong> constant magnetiz<strong>at</strong>ion. Describe the st<strong>at</strong>e of the system<br />

<strong>at</strong> this very low temper<strong>at</strong>ure. Calcul<strong>at</strong>e relevant quantities as a function of<br />

temper<strong>at</strong>ure.<br />

Define a temper<strong>at</strong>ure Ts by<br />

m(Ts) = mi<br />

m 2 Tc − Ts<br />

0 =<br />

Tc<br />

1<br />

4 m20


368 APPENDIX C. SOLUTIONS TO SELECTED PROBLEMS.<br />

Tc − Ts = 1<br />

4 Tc<br />

Ts = 3<br />

4 Tc<br />

For T > Ts nothing special is happening, the system is uniform, and a magnetic<br />

field is required to produce the magnetiz<strong>at</strong>ion. At T = Ts the required magnetic<br />

field is zero. When T < Ts the system will phase separ<strong>at</strong>e, part of it will<br />

have a neg<strong>at</strong>ive magnetiz<strong>at</strong>ion. No field is required. If the fraction of neg<strong>at</strong>ive<br />

magnetiz<strong>at</strong>ion is c, we have<br />

√ √<br />

Tc − T<br />

Tc − T<br />

mi = c(−m0 ) + (1 − c)(m0 )<br />

Tc<br />

√<br />

1<br />

Tc − T<br />

= (1 − 2c)<br />

2 Tc<br />

√ √<br />

Tc − Ts<br />

Tc − T<br />

= (1 − 2c)<br />

Tc<br />

1 − 2c =<br />

√ Tc − Ts<br />

Tc − T<br />

Tc<br />

Tc

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!