You are on page 1of 240

PHYSICS

213 Elements of Thermal Physics


3rd Edition

James P. Wolfe

Department of Physics
University of Illinois at Urbana – Champaign
Copyright © 2010 by James P. Wolfe
Copyright © 2010 by Hayden-McNeil, LLC on illustrations provided
Photos provided by Hayden-McNeil, LLC are owned or used under license

Permission in writing must be obtained from the publisher before any part of this work may be
reproduced or transmitted in any form or by any means, electronic or mechanical, including
photocopying and recording, or by any information storage or retrieval system.

Printed in the United States of America

10 9 8 7 6 5 4 3 2 1

ISBN 978-0-7380-4119-3

Hayden-McNeil Publishing
14903 Pilot Drive
Plymouth, Michigan 48170
www.hmpublishing.com

Wolfe 4119-3 F10


Contents

Preface ......................................................................................................... vii


Definition of Symbols .................................................................................. ix
Table of Constants and Conversion Factors................................................ xi

Introduction and Overview


A. Classical to Quantum Physics ............................................................... xiii
B. Systems with Many Particles ..................................................................xiv
C. Statistics and Entropy..............................................................................xv
D. Road Map for this Course .....................................................................xvii

Chapter 1: Origins of Mechanical Energy


A. Kinetic Energy and Work ......................................................................... 1
B. Extension to Many-Particle Systems........................................................ 3
C. Internal Energy ....................................................................................... 5
D. Potential Energy ...................................................................................... 7
E. Vibrational Energy—Kinetic plus Potential ............................................ 8

Chapter 2: Irreversibility and the Second Law of


Thermodynamics
A. Thermal Energy ..................................................................................... 13
B. Irreversibility of Many-Body Systems .................................................... 14
C. Entropy and the Approach to Equilibrium ............................................ 14
D. Entropy Maximization and the Calculus of Several Variables .............. 17

Chapter 3: Kinetic Theory of the Ideal Gas


A. Common Particles .................................................................................. 21
B. Pressure and Kinetic Energy .................................................................. 22
C. Equipartition Theorem .......................................................................... 23
D. Equipartition Applied to a Solid ........................................................... 25
E. Ideal Gas Law ......................................................................................... 26
F. Distribution of Energies in a Gas ........................................................... 27

Chapter 4: Ideal-Gas Heat Engines


A. The First Law of Thermodynamics ...................................................... 31
B. Quasi-static Processes and State Functions ........................................... 32

iii
Physics 213 Elements of Thermal Physics

C. Isothermal and Adiabatic Processes—Reversibility ............................... 32


D. Entropy of the Ideal Gas—a First Look ................................................ 36
E. Converting Heat into Work ................................................................... 37
F. Refrigerators and Heat Pumps ................................................................ 40

Chapter 5: Statistical Processes I: Two-State Systems


A. Macrostates and Microstates .................................................................. 45
B. Multiple Spins ......................................................................................... 46
C. The Random Walk Problem—Diffusion of Particles ........................... 50
D. Heat Conduction.................................................................................... 54

Chapter 6: Statistical Processes II: Entropy and the Second Law


A. Meaning of Equilibrium ......................................................................... 59
B. Objects in Multiple Bins ........................................................................ 60
C. Application to a Gas of Particles ............................................................ 62
D. Volume Exchange and Entropy ............................................................. 64
E. Indistinguishable Particles ..................................................................... 68
F. Maximum Entropy in Equilibrium ......................................................... 68

Chapter 7: Energy Exchange


A. Model System for Exchanging Energy .................................................. 73
B. Thermal Equilibrium and Absolute Temperature.................................. 78
C. Equipartition Revisited .......................................................................... 79
D. Why Energy Flows from Hot to Cold .................................................. 81
E. Entropy of the Ideal Gas—Temperature Dependence .......................... 82

Chapter 8: Boltzmann Distribution


A. Concept of a Thermal Reservoir ............................................................ 87
B. The Boltzmann Factor............................................................................ 88
C. Paramagnetism ....................................................................................... 91
D. Elasticity in Polymers............................................................................. 94
E. Harmonic Oscillator .............................................................................. 95

Chapter 9: Distributions of Molecules and Photons


A. Applying the Boltzmann Factor ............................................................. 99
B. Particle States in a Classical Gas .......................................................... 100
C. Maxwell-Boltzmann Distribution ........................................................ 102
D. Photons ................................................................................................. 103
E. Thermal Radiation................................................................................ 105
F. Global Warming .................................................................................... 107

Chapter 10: Work and Free Energy


A. Heat Flow and Entropy ........................................................................ 111
B. Ideal Heat Engines ............................................................................... 112
C. Free Energy and Available Work ......................................................... 114
D. Free Energy Minimum in Equilibrium ............................................... 115

iv
Elements of Thermal Physics Physics 213

E. Principle of Minimum Free Energy .................................................. 116


F. Equipartition of Energy ........................................................................ 117
G. Paramagnetism—the Free Energy Approach ...................................... 119

Chapter 11: Equilibrium between Particles I


A. Free Energy and Chemical Potential ................................................... 123
B. Absolute Entropy of an Ideal Gas......................................................... 125
C. Chemical Potential of an Ideal Gas ..................................................... 128
D. Law of Atmospheres ............................................................................. 129
E. Physical Interpretations of Chemical Potential ................................... 130

Chapter 12: Equilibrium between Particles II


A. Ionization of Atoms .............................................................................. 135
B. Chemical Equilibrium in Gases............................................................ 137
C. Carrier Densities in a Semiconductor ................................................. 139
D. Law of Mass Action: Doped Semiconductors .................................... 142

Chapter 13: Adsorption of Atoms and Phase Transitions


A. Adsorption of Atoms on a Solid Surface .............................................. 145
B. Oxygen in Myoglobin ........................................................................... 147
C. Why Gases Condense .......................................................................... 148
D. Vapor Pressure of a Solid ..................................................................... 148
E. Solid/Liquid/Gas Phase Transitions .................................................... 151
F. Model of Liquid–Gas Condensation .................................................... 154

Chapter 14: Processes at Constant Pressure


A. Gibbs Free Energy................................................................................ 157
B. Vapor Pressures of Liquids—General Aspects ..................................... 160
C. Chemical Reactions at Constant Pressure ........................................... 161

Appendices
Appendix 1: Vibrations in Molecules and Solids—Normal Modes ......... 165
Appendix 2: The Stirling Cycle ................................................................ 169
Appendix 3: Statistical Tools ..................................................................... 173
Appendix 4: Table of Integrals .................................................................. 179
Appendix 5: Exclusion Principle and Identical Particles .......................... 181
Appendix 6: Sum over States and Average Energy ................................... 185
Appendix 7: Debye Specific Heat of a Solid ............................................. 189
Appendix 8: Absolute Entropy of an Ideal Gas ........................................ 191
Appendix 9: Entropy and Diatomic Molecules ........................................ 195
Appendix 10: Vapor Pressure of a Vibrating Solid ................................... 199

Solutions to Exercises ............................................................................... 201

Index .......................................................................................................... 215

*May not be covered in Physics 213


v
Physics 213 Elements of Thermal Physics

The central ideas in this course have a wide range of applications.


For example:

← fabrication of materials

chemical reactions →

← biological processes

phase transitions →

← magnetism

electrons and holes in semiconductors →

← converting energy into work

thermal radiation (global warming) →

← thin films and surface chemistry

and much more...

vi
Preface

The unifying concepts of entropy and free energy are essential to the under-
standing of physical, chemical, and biological systems. Recognizing that these
concepts permeate the undergraduate science and engineering curricula, the
Physics Department has created this sophomore-level course dealing with
thermodynamics and statistical mechanics. Starting with a few basic principles,
we introduce practical tools for solving a variety of problems in the areas of
materials science, electrical engineering, chemistry, and biology.

These introductory notes on Thermal Physics are designed to be used in concert


with the Physics 213 Lectures, Discussion Exercises, Homework Problems, and
Laboratory Exercises. The Lectures summarize the principal ideas of the course
with live demonstrations and active-learning exercises. Discussion problems
are solved cooperatively by students. The lab experiments lend reality to the
basic principles.

Exercises at the end of each chapter are designed to complement discussion


and homework problems. Solutions to most Exercises are provided in the back
pages. Appendices (and Chapter 14) are included for students who want to
dig a little deeper into the subjects of this course and gain additional links to
advanced courses.

Acknowledgements
The precursor to this course was first taught in Fall 1997 and Spring 1998 by
Michael Weissman and Dale Van Harlingen. Subsequent versions of the course
were developed by Doug Beck, Michael Weissman, Jon Thaler, Michael Stone,
Paul Debevec, Lance Cooper, Yoshi Oono, Paul Kwiat, and myself. I particularly
wish to thank Mike Weissman, Lance Cooper, Yoshi Oono, Inga Karliner, and
Paul Kwiat for insightful suggestions and corrections to this book.

vii
Physics 213 Elements of Thermal Physics

Reference Texts
In developing material for this course, I have drawn heavily from three excellent books.
For students who wish to extend their knowledge in this area, I highly recommend them:

C. Kittel and H. Kroemer, Thermal Physics, Second Edition (W. H. Freeman,


1980)

D.V. Schroeder, An Introduction to Thermal Physics (Addison-Wesley, 1999)

F. Reif, Fundamentals of Statistical and Thermal Physics (McGraw-Hill, 1965)

The following books also provide very useful perspectives:

Thomas A. Moore, Six Ideas That Shaped Physics, Unit T (McGraw-Hill, 1998)

F. Reif, Statistical Physics, Berkeley Physics Course—Vol. 5 (McGraw-Hill,


1965)

Steven S. Zumdahl, Chemical Principles, 5th Edition, (Houghton Mifflin, 2005)

Professor Gino Segre has written a fascinating historical perspective of the world from
the viewpoint of thermodynamics. It’s a “must read” for science and engineering majors:

Gino Segre, A Matter of Degrees: What Temperature Reveals about the Past
and Future of Our Species, Planet, and Universe (Penguin, USA, 2003).

viii
Elements of Thermal Physics Physics 213

Definition of Symbols
(Alphabetically arranged)

A area, or amplitude of vibration


<a> thermal average of the variable a
 = U/pV = U /NkT constant for ideal gas over some T range,  = 3/2 (mona-
tomic), 5/2 (diatomic)
 = 1/kT shorthand used in Boltzmann factor, exp(–En)
Cv (Cp) heat capacity at constant volume (pressure)
cv = Cv/n molar specific heat (n = # moles)
E energy of a single particle, or a single oscillator
 = hf quantum of energy for an oscillator with frequency f
 = Wby/Qh efficiency of a heat engine operating between Qh and Qc
KE translational kinetic energy (= ½mv2 = p2/2m for a single
particle)
En energy of quantum state labeled n (an integer). (e.g.,
En = n for an oscillator, nB for a spin, h2n2/(2L)2 for a
particle in a box, and –(13.6 eV)/n2 for an H-atom.)
F = U – TS Helmholtz free energy
F force
G = U + pV – TS Gibbs free energy
H = U + pV enthalpy
 = ( + 1)/ adiabatic constant (pV = constant for adiabatic process with
ideal gas)
h,  Planck’s constant = 6.63  10–34 J s,  =h/2
k Boltzmann constant = 1.381  10–23 J/K
ln(x) natural logarithm of x (base e = 2.7183)
log(x) base-10 logarithm of x
x step length in a 1-d random walk process
 mean free path of a particle = v

m wavelength of a standing wave with mode index m
M # bins or cells for one particle, # steps in random walk
0 = (Nup – Ndown) = m total magnetic moment of N spins
 magnetic moment of one spin, or chemical potential
m mass of a particle
m integer = Nup – Ndown or Nleft – Nright in binomial
distribution
N # particles, # oscillators, or # spins
n # moles
n = N/V number density of particles (p = nkT for ideal gas)
nQ quantum density of an ideal gas
NA Avogadro’s constant = 6.02  1023 molecules/mole
p = F/A pressure

ix
Physics 213 Elements of Thermal Physics

p = mv momentum of a particle (distinguish from pressure by usage)


px = mvx x-component of momentum of a particle
Pn probability that a particle is in a quantum state labeled n
P(m) probability of sampling m = Nup – Ndown or Nleft – Nright
P(E), P(x), P(m) probability density (per unit energy, distance, or step)
P(E)dE probability that a particle has energy between E and E + dE
q number of energy quanta in an oscillator
Q heat (positive if inflow, or negative if outflow)
Qh and Qc heat flow to/from hot and cold reservoirs (defined as positive)
R = NAk gas constant = 8.314 J/mol-K
= ln (dimensionless) entropy of a system with microstates
SB Stefan–Boltzmann constant
R(UR) or SR(UR) entropy of a thermal reservoir with energy UR
S = k = k ln( ) conventional entropy (units J/K)
d standard deviation of a distribution
T absolute temperature (Kelvin)
mean collision time for a particle
U energy of a many-particle system
UR energy of a thermal reservoir
V volume
v speed of a particle
# microstates for a many-particle system,
= 1 2 # microstates of 2 combined systems that separately have 1 and
2 microstates
Wby or Won work done by or on a system, Wby = pdV
R(UR) # microstates of a thermal reservoir with energy UR
(E)dE # microstates of a particle with energy between E and E + dE

x
Elements of Thermal Physics Physics 213

Table of Constants
k = 1.381  10–23 J/K Boltzmann’s constant = 8.617  10–5 eV/K
NA= 6.022  1023 Avogadro’s constant
R = 8.314 J/mol K gas constant = NAk
= 0.0821 liter atm/mol K
h = 6.626  10–34 J s Planck’s constant = 4.136  10–15 eV-s
 = h/2 = 1.055  10–34 J s
c = 3.00  108 m/s speed of light
g = 9.80 m/s2 acceleration due to earth’s gravity
me = 9.11  10–31 kg mass of an electron
mp = 1.674  10–27 kg mass of a proton = 1836 me
e = 9.2848  10–24 J/T Electron magnetic moment = 57.95 eV/Tesla
SB = 5.670  10–8 W/m2k4 Stefan–Boltzmann constant

Conversion Factors
1 liter = 103 cm3 = 10–3 m3 1 eV = 1.602  10–19 J
1 Pa = 1 N/m2 1 eV/particle = 96.5 kJ/mol
1 atm = 1.013  105 Pa At 300 K, kT = 0.026 eV
1cal = 4.184 J = energy to raise 1g of H2O by 1 K At 300 K, RT = 2494 J
T(K) = T(°C) + 273 = (5/9)(T(°F) –32) + 273 1 liter-atm = 101.3 J

Temperature Scales

ABSOLUTE CELSIUS FAHRENHEIT

373 K 100°C 212°F


water boils

273 K 0°C 32°F


water freezes

77 K -196°C -320°F
liquid nitrogen
0K -273°C -460°F
absolute zero

"room temperature" ~
~ 293 K ~
~ 20°C ~
~ 68°F

xi
Physics 213 Elements of Thermal Physics

xii
PHYSICS

213
Introduction and Overview

A. Classical to Quantum Physics


Physics seeks to explain and predict the world around us using precise
mathematical tools. The math tools we develop are based on experi-
mental observations. The essential test of a mathematical theory is its
ability to predict the behavior of nature in new situations. The theories
that allowed the engineers and scientists of the 1960s to put a man on
the moon were developed and tested right here on earth. The future
extension of microchips into the sub-micron regime (and the develop-
ment of new data-storage media) will rely upon the creative application
of the present theories of materials.

At the beginning of the 20th century, classical mechanics and electro-


magnetic theory were well developed, so when microscopic particles
such as the electron were discovered, scientists were quick to apply these
proven theories to the new regime. Explaining the atom in planetary
terms soon failed. An electron circulating around a nucleus is continually
accelerating, and an accelerating charge radiates electromagnetic energy
like a radio antenna. Thus, according to classical theories, the electron
should lose its energy and spiral into the nucleus. A resolution of this
dilemma was provided by scientists in the early 1900s who proposed
that an electron behaves more like a wave than a localized particle.

xiii
Physics 213 Elements of Thermal Physics

In essence, the classical orbits are replaced by stationary waves describing the prob-
ability of finding an electron in a certain place. Stationary charge means no radiation,
and therefore the atom is stable. When a particle is confined in space, its wave nature
gives rise to discrete energy levels, such as the electronic energy levels in atoms. Hence
the name “quantum mechanics.”

In this course we shall see that macroscopic (large scale) properties of systems with many
particles depend on the microscopic wave nature of the constituent particles. For example,
the electrical conductivity of the silicon crystals in your watch or computer depends on
the fundamental constant of quantum mechanics, h = Planck’s constant.

B. Systems with Many Particles


This course is an introduction to the physics of many-particle systems, also known as
thermal physics. Traditionally this is the realm of classical thermodynamics, which
approaches the subject from an empirical, or observational, point of view. As the word
suggests, thermodynamics is the study of “heat” and “work.” Although steeped in
mathematical formalism, classical thermodynamics has wide-ranging applications. The
developments of modern machines—including your car, your computer, the plane over-
head, etc.—are based on applications of classical thermodynamics. Modern chemistry
and engineering rely on thermodynamic principles.

Some of the main questions of thermodynamics are: What are the practical limits
in converting heat to work? Does energy always flow from hot to cold? What is the
meaning of hot and cold—quantitatively? What determines the physical properties of a
medium—for example, its heat capacity, its electrical conductivity, or its magnetic prop-
erties? Why does matter undergo phase transitions between gases, liquids, and solids?

Although the empirical laws of classical thermodynamics have wide-ranging utility, a


basic understanding of many-particle systems requires an atomistic approach. Thermal
physics begins at the microscopic level and applies statistical concepts to understand the
macroscopic behavior of matter. The microscopic approach of thermal physics is also
known as statistical mechanics.

Consider the magnitude of the problem: There are 6  1022 atoms in a cubic centimeter
of silicon crystal, and 0.27  1020 molecules in a cubic centimeter of air. Even with the
biggest computer imaginable, you could not predict the motion of an individual particle
in such a complicated system. How, then, can we begin to predict the behavior of the
gas in this room, or the electrical and thermal properties of a solid, or phase transitions
between solids, liquids, and gases?

The answer is that the world around us is governed by the random, statistical behav-
ior of many, many particles. Neither classical mechanics nor quantum mechanics can
predict the properties of many-particle systems without the help of statistical methods.

xiv
Introduction and Overview Physics 213

C. Statistics and Entropy


To appreciate the importance of statistics in describing many-particle systems, consider
the case of 10 gas particles in a two-part container. Initially we put all 10 particles in
the left side:

Now we watch while the particles move around with some thermal energy. As time
proceeds, both sides of the container become populated with particles. If you were to
take snapshots of the system as time progressed you might find the following results:

10

Number in
left side, NL 5

0
Time

As you continue to take snapshots, you would get a pretty good idea what values of NL
you are most likely to observe. If you take many snapshots and tabulate the number of
occurrences for each value of NL, you would find, roughly:

Number of
occurrences

0 1 2 3 4 5 6 7 8 9 10
NL

If you could do the experiment with 100 particles, then the result would be a more
compressed histogram, something like this:

Number of
occurrences

0 10 20 30 40 50 60 70 80 90 100
NL

xv
Physics 213 Elements of Thermal Physics

Intuitively we can understand why the distribution is more compressed for a larger
number of particles. The probability that the left half of this room would contain 90%
of the gas particles is extremely unlikely. A general result is the following: For a total
of N particles, the statistical variation in NL (i.e., the width of the distribution) is about N1/2.
For the 100-particle example above, the width of the distribution is about 10. For N =
1020 the variation is only 1010 particles, or one part in ten billion—an extremely sharp
distribution. That’s why the pressure in this room doesn’t fluctuate significantly.

The above example suggests two things: 1) as systems get bigger, the macroscopic proper-
ties (e.g., the fraction of particles on the left, or the gas pressure on the left side) become
more certain, and 2) in order to quantitatively describe a system, we need to count the
number of ways that particles can distribute themselves. In technical terms, we count
the “number of accessible microstates,” denoted . Counting microstates is a major
topic in statistical mechanics. The “number of occurrences” plotted above are basically
graphs of for various values of NL.

The logarithm of the number of accessible states defines the entropy of a system. More
specifically, entropy is S = k ln, where k is a constant. Entropy is a fundamental prop-
erty of a many-particle system. You may have heard the phrase, “entropy is disorder.”
Disorder is not really a well-defined concept; however, entropy does represent disorder
in the sense that more entropy corresponds to a larger number of possible states.

To get a feeling for the importance of entropy, consider setting up the two-part container
with 10 particles on the left and 90 particles on the right. This situation corresponds to
the arrow on the diagram on page xv. Under the condition NL = 10 there are a limited
number of “microstates” available. There are many more microstates associated with NL
= 50 than NL = 10. The particles will diffuse around, sampling all microstates, until there
is roughly an average of 50 particles on the right and 50 particles on the left. Eventually
there is very little possibility of finding the system with only 10 particles on the left.

There are two fundamental observations that we can make here:

1) Many-particle systems exhibit irreversibility. While it is possible for the system


to revert back to 10 particles on the left, it is extremely unlikely. If we were dealing
with N = 1023 particles, it would take longer than the age of the universe before
NL = N/10 = 1022 is observed. That would be equivalent to the pressure in one of
your lungs suddenly dropping to 0.2 atmospheres.

2) In equilibrium, entropy is maximized. The fundamental postulate of statistical


mechanics is that each available microstate is equally likely. If we initiate a system
in a restricted set of microstates then remove the constraint, the system will redis-
tribute, randomly occupying all accessible microstates. In equilibrium, therefore,
the probability of measuring NL is proportional to the corresponding value of .

These facts are the basis for The Second Law of Thermodynamics. In this course
we will exploit this principle to solve many useful problems.

xvi
Introduction and Overview Physics 213

D. Road Map for this Course


The following section is provided to give you an overview of the contents and goals
of this course. I suggest that you read it briefly now and refer to it frequently as the
course progresses.

Before attacking the statistical aspects of thermodynamics, we will bolster your intuition
about the macroscopic world. In Chapter 1 the concept of internal energy is introduced
for many-particle systems. The ideas of kinetic energy and potential energy are reviewed,
and we consider a system that has both: the harmonic oscillator. The harmonic oscilla-
tor is the basis for vibrations in molecules and solids and comes up often in this course.

The underlying principle of thermal physics is the irreversibility of many-particle sys-


tems. In Chapter 2 we discuss this concept as the Second Law of Thermodynamics and
introduce the concept of entropy. Entropy is a maximum for systems in equilibrium,
leading to an important relationship between entropy and absolute temperature.

In Chapter 3 we investigate the ideal gas—a dilute collection of free particles with neg-
ligible interactions. Kinetic theory provides us with a microscopic model of pressure.
The Equipartition Theorem provides us with a working definition of temperature.
Using these concepts we derive the Ideal Gas Law, pV = NkT.

In Chapter 4, we examine the thermal cycles of ideal gases in the context of heat engines
and discover that the most efficient engine is the Carnot engine. The Carnot cycle is
the standard against which all other thermal cycles are measured. Heat engines that run
in reverse are refrigerators or heat pumps, providing us with many useful applications.

Statistical concepts are introduced in Chapter 5. We begin with a system of spins, pro-
viding the basis for paramagnetism. The spin system clearly illustrates the concepts of
microstates and macrostates. A mathematically similar two-state problem is the random
walk, which is the basis for particle diffusion and heat conduction. The math tools in
this chapter are the binomial and Gaussian distributions.

Chapter 6 extends our statistical tools to systems with multiple bins or cells, allowing us
to treat the particles in an ideal gas. We examine in detail a basic problem of statistical
mechanics: the equilibrium between two systems that exchange volume. The underly-
ing principle is that the most likely configuration of an isolated system corresponds to
a maximum in total entropy, S = k ln .

In Chapter 7 we consider the exchange of energy between two systems, leading to the
general definition of absolute temperature in terms of the derivative of entropy with
respect to energy U,

1 dS
=
T dU ,

at constant volume V and particle number N.

xvii
Physics 213 Elements of Thermal Physics

In Chapter 8 we note that if a large system (a “thermal reservoir”) at temperature T is


in thermal contact with a small system with variable energy U, then the entropy of the
reservoir is just given by a Taylor expansion, So – (dS/dU)U, or in terms of temperature,

Sres = So – U/T,

where So is a constant. We shall see that this relation (and Sres = k ln res) leads directly
to the probability Pn of finding a particle in a quantum state (labeled n) with energy En:

Pn = Ce –E n / kT
,

which is the famous Boltzmann distribution. This basic result is applied to paramagnetic
spins, elasticity in polymers, vibrations in molecules, and electronic states.

In Chapter 9 the Boltzmann approach is used to predict the energy distribution of par-
ticles in an ideal gas—the so-called Maxwell–Boltzmann Distribution. A second applica-
tion is the frequency distribution of thermal radiation, leading to the Stefan–Boltzmann
Law for the power radiated from a hot object, such as a light bulb, your body, or the sun.

The application of these statistical concepts is greatly facilitated by defining what is


known as the “free energy,”

F = U – TS,

and its derivative with respect to particle number at constant V and T,

 = dF/dN,

known as the “chemical potential.” Equilibrium conditions are determined by mini-


mizing the free energy of a system, which leads to simple relations between the chemi-
cal potentials of its subsystems. In Chapters 10–13 we will apply the Principle of Free
Energy Minimum to the following problems:

I. Ideal Gases
II. Paramagnetic Spins
III. Law of Atmospheres
IV. Ionization of Atoms
V. Chemical Equilibrium in Gases
VI. Carrier Densities in Semiconductors
VII. Adsorption of Particles on Surfaces
VIII. Phase Transitions

The problems chosen for this course represent many important processes in the world
around us. You will learn quantitative methods for studying a broad range of physical,
chemical, and biological materials.

xviii
Introduction and Overview Physics 213

Here is some practical information about this book:

a) It is a good idea to read the assigned chapter before lecture and come to class with
questions.

b) Appendices generally are optional reading and some include advanced material that
may be useful in preparation for upper division courses.

c) Chapter 14 (Gibbs free energy) may not be covered in this course, but it may be of
specific interest to chemistry, materials science, and physics majors.

d) Exercise problems are provided at the end of each chapter. Solutions to most prob-
lems are given at the back of the book.

Enjoy your adventure into Thermal Physics!

Jim Wolfe, UIUC

xix
Physics 213 Elements of Thermal Physics

Exercises
1) An appreciation of the concept of microstates can be gained by considering a two-
cell container with 10 distinguishable objects, labeled A through J:

B J I
D
C A
H F E
G

For a total of N objects, the number of ways of arranging the system with NL objects
on the left and NR = N – NL objects on the right is the binomial distribution:

N!
Ω=
N L !N R !

Complete the following table for the above system: (By definition: 0! = 1.)

NL = 0 1 2 3 4 5 6 7 8 9 10
= 210

We say that (NL) is the “number of microstates in the macrostate labeled NL”.
Compare to the graph in Section C. [In reality one must determine for identical
particles (atoms or molecules) in a given volume. Interestingly, the binomial dis-
tribution still applies to identical particles in a two-part container, as discussed in
Chapter 6 (E).]

2) Logarithms and exponents are not just mathematical conveniences; they are an
p integral part of thermal physics. The common integral dx/x equals the natural
logarithm, ln x. Conversely, the derivative of the natural logarithm d(ln x)/dx equals
l/x. Using the ideal gas law, pV = NkT, show that the work W done by an ideal
gas expanding from volume V1 to V2 at constant temperature is NkT ln(V2/V1).
(p = pressure, T = temperature, Nk = constant)
V2

Wby = ∫ pdV
V1
=

xx
CHAPTER

1 Origins of Mechanical Energy

A. Kinetic Energy and Work


Energy may have been the first concept in mechanics that you found
difficult to understand intuitively. In contrast, momentum is not a par-
ticularly difficult concept for anyone who has bumped into something.
The pain or damage from a collision increases with both the mass and
the velocity of the offending object. Momentum is simply mass times
velocity.

Energy, on the other hand, may have a different meaning (or many
meanings) for each of us. Energy is what we are supposed to feel after
eating a certain bowl of cereal in the morning, or what we don’t feel the
day after cramming for a physics exam. Energy is what lights our lights,
warms our dorms, propels our cars and our bodies, and costs money.

Hopefully, your mechanics and E & M courses provided you with a


practical understanding of energy. But, just in case you don’t quite
remember where ½ mv2 came from, here is how the idea crept in your
mechanics course…

1
Physics 213 Elements of Thermal Physics

You were dealing with an object moving in a force field (say the earth’s gravitational
field), and you wanted to predict the velocity of the object at different positions. Well, that’s
just a physicist’s way of saying, “Drop a ball and describe what happens.”

yi

y h
F

yf

To make things simple, we consider motion in one dimension. (No vectors here.)
The gravitational force on the object with mass m is F = mg, where g = 9.8 m/s2 is the
acceleration due to gravity. We invoke Newton’s Second Law (and the definition of ac-
celeration) to describe the motion of the object:

F = ma = m dv/dt = dp/dt, (1-1)

where a and v are the instantaneous acceleration and velocity, and p = mv defines the
momentum. Now, here is where some creative math comes into play. We start with

F dt = m dv (1-2)

and take the rather unpredictable step,

F v dt = m v dv. (1-3)

Why would we want to do that? The reason is that this step allows us to change the
differential on the left from time to distance, namely: v dt = dy. Now, we have,

F dy = m v dv. (1-4)

Now just integrate this equation to get the final result (remembering that vdv = ½ v2):

F dy = ½ mvf2 – ½ mvi2 , (1-5)

where the subscripts i and f refer to the initial and final velocities of the object. Because
F is a constant in this case, the integral on the left becomes mgh, where h is the distance
the ball drops. (The integral is positive because F and dy are in the same direction.)
Now you can set vi = 0 and solve for vf.

Why didn’t we just integrate F dt = m dv to get

mg (tf – ti) = m(vf – vi) ? (1-6)

2
Origins of Mechanical Energy Chapter 1

If the ball is dropped with vi = 0 at ti = 0, then mgtf = mvf. This approach gives us the
final velocity in terms of the final time, which we don’t know. Furthermore, if the force
were a function of position (such as in the proximity of planets or electrical charges),
we would not have enough information to do the integral of Fdt.

The neat trick about multiplying both sides of Eq. (1-2) by v, is that it turns dt into the
differential of a quantity, y, which is the variable specified in the problem (yi = h, yf = 0).

Let’s not lose sight of our goal. We have just seen that

F dy = (½ mv2), (1-7)

where  = “change in,” and the integral form of this equation is required when F is a
function of position. We recognize that the integral on the left is the work done on the
object by the earth. For a general applied force,

Won  F  dr, (1-8)

and for a single particle, we define

KE  ½ mv2 (1-9)

as the kinetic energy of the particle. The relation

Won = (KE) (1-10)

is a very basic result for the motion of a single object, often called the “Work-Energy
Theorem.” In words, the work done on an object by an applied force equals the change
in kinetic energy of the object. In fact, this result is nothing more than the integral form
of Newton’s Second Law, with a couple of new definitions: work and energy.

Equation (1-10) is a highly useful form of Newton’s law. As you have seen in your me-
chanics course, for “conservative forces,” such as those due to gravitation and electric
charges, the work done in moving from point A to point B does not depend on the path
taken. The shape of a roller coaster doesn’t matter in determining the final speed at a
given elevation (assuming no friction). The speed of an object orbiting the earth in an
elliptical orbit is directly related to its distance from the earth.

B. Extension to Many-Particle Systems


If we extend our discussion of energy to a collision between two point-like particles and
assume no external forces, then the left side of Equation (1-10) vanishes and we have,

0 = (KE1) + (KE2) = (KEtot), (1-11)

where the numerical subscripts label the two particles. This is a statement of “conser-
vation of energy” for the simple case where there are no interaction energies in the
system, except briefly during collisions.

3
Physics 213 Elements of Thermal Physics

Recall that there was one other conservation law associated with Newton’s Second Law,
namely the conservation of momentum. Let us assume that there are no external forces
on a system of particles and that Fij is the force on object i due to object j. Newton’s
Zeroth Law dictates that forces appear in equal and opposite pairs: Fij = – Fji. Writing
Newton’s Second Law in vector form,

 Fij = mi dvi/dt = dpi/dt, (Fij = force on i due to j) (1-12)

where the left side is the sum (over j) of all the forces on the object labeled i. If there are
N point objects, then there will be N equations. Adding all of these equations together
and noting that the pairwise forces cancel, we get,

0 =  dpi/dt = d(pi)/dt, (1-13)

implying that the total momentum p = pi must be a constant. Notice that the total
momentum is also the center-of-mass momentum pcm = Mvcm where M = mi and
vcm = mivi/M is the center-of-mass (or average) velocity of the particles.

If a single external force F acts on a system of particles (or any object), then Newton’s
Second Law takes on the form,

F = dpcm/dt = Macm. (1-14)

This is a very general equation, which applies to any system of particles, even those
bonded together in a solid. Consider a force applied to the end of a solid board, as
depicted here:

The dot represents the center of mass of the board, which is situated on a frictionless
table. Applying the force to one end of the board produces a complicated motion involv-
ing both translation and rotation. Notice that the distance the force’s contact point moves is
different from the distance that the center of mass moves. For example, at some specific time
later the force has pulled a distance D and the center of mass of the board has moved
a distance dcm = D – ½ L:

4
Origins of Mechanical Energy Chapter 1

F
F

dcm

(Note: the board is still rotating at this instant of time.) The work done by the constant
force is F times the distance it acts,

Won = FD. (1-15)

What does Newton’s Second Law (Eq. 1-14) tell us about this system? Using our little
trick again,

F vcm dt = M vcm dvcm , (1-16)

but now we have,

F dcm = (KEcm) . (1-17)

The center-of-mass subscripts are very important. dcm is the distance that the center-
of-mass moved, not the distance the force acted, so the left side of this equation is not
equal to the work done on the board! The integral of Newton’s Second Law in this case
is not a work-energy equation. For simplicity, we call Eq. (1-17) the “c.m. equation.”

C. Internal Energy
The Work-Energy Theorem for a many-particle system with internal degrees of free-
dom (e.g., rotation or vibration) is actually a distinct concept from Newton’s Second
Law. It relates work to the change in total energy,

Won = (Total Energy). (1-18)

For the rotating and translating board considered above, the total energy is the sum of
KEcm and rotational energy about the center of mass. Therefore,

FD = (KEcm) + (KErot). (1-19)

The work done on the board is converted totally into translational plus rotational energy.

5
Physics 213 Elements of Thermal Physics

Recall from your mechanics course that rotational energy is ½ Icm2, where Icm is the
moment of inertia of the object and  is the angular speed in radians per second. The
work-energy equation (1-19) plus the cm equation (1-17) allow you to determine the
rotational energy of the board when it reaches the orientation in the second drawing.
(Answer: (KErot) = FL/2.)

In general, the total energy of an object is KEcm + U, where U is the internal energy
of the system, which includes rotational and vibrational motions, as well as potential
energy associated with the binding of molecules and atoms. The generalized work-
energy equation may be written:

Won = (KEcm) + U (1-20)

U is essentially the energy of the system in the cm frame of reference. Consider the mechanics
problem where two blocks collide and stick together (an inelastic collision):

Initially: Finally:

m vi m m m v f = v i /2

Because the total force on the two blocks is zero, their total momentum is constant,
implying mvi = 2mvf. Notice that vf = vcm in the lab frame. Also, acm = 0 implies that vcm
is constant (= vi/2) throughout, so (KEcm) = 0. Therefore, U = 0 in this collision. In
fact, the initial U is not zero for our 2-mass system. We can see this by observing the
collision in the center-of-mass frame:

Initially: Finally:

m v i /2 v i /2 m m m

From the initial state, we see that the internal energy U of the 2-mass system equals ¼
mvi2. In the collision, all of this easily identifiable translational energy is converted into
sound waves and thermal vibrations in the blocks.

The concept of internal energy is particularly important in the study of many-body


systems. In fact, in this course we will rarely deal with the center-of-mass kinetic en-
ergy of a system. We will be concerned almost entirely with the internal energy U by
observing the system in the cm frame. The statement of energy conservation when
(KEcm) = 0 is simply:

Won = U.

6
Origins of Mechanical Energy Chapter 1

D. Potential Energy
In your mechanics course you learned how the work done on an object by a “conser-
vative” force can be treated in terms of potential energy. For example, an object at a
height h from the earth’s surface has a gravitational potential energy PE = mgh. One
must be careful not to double-count work and energy in the Work Energy Theorem,
as I will now demonstrate:

Won = (Total Energy) (1-21)

= (KE) + (PE)

= (½ mv2) – mgh.

The minus sign is because the object lost mgh potential energy in falling a distance h
(chosen positive). However, we have already seen that the work done on the object by
the earth’s gravitational field when the object falls a distance h is

Won = force  distance = mgh. (1-22)

Therefore, we are led to the (incorrect) conclusion that

(½ mv2) = 2mgh (1-23)

in contradiction to our earlier result. Where have we gone wrong?

The reason that we have double-counted mgh is that we have not clearly identified
our system in applying Eq. (1-21). Remember, the procedure in solving a mechanics
problem is:

1) specify the system that you are considering,

2) draw all the external forces on that system, and

3) apply Newton’s Second Law.

If we choose the system as the ball plus the earth, then we realize that there are equal
and opposite forces between the ball and the earth. Thus, we realize that there are no
external forces in this system, but there is the potential energy between the ball and the
earth. So, the left side of the work-energy equation vanishes, and we have,

0 = (½ mv2) – mgh. (1-24)

7
Physics 213 Elements of Thermal Physics

If, on the other hand, we choose the system as the ball alone, then there is no potential
energy (potential energy requires at least two objects interacting, such as the ball and the
earth), and the external force on the system is mg. The work-energy equation becomes,

mgh =  (½ mv2), (1-25)

which, again, is the correct result. A consistent choice of system and external forces
is critical to the solution of a mechanics problem. These ideas carry over to ther-
modynamics problems. For example, in calculating atmospheric pressure at an altitude
h, it is usual to include potential energy mgh per particle in the total internal energy
U, implicitly including the earth in the system.

E. Vibrational Energy—Kinetic plus Potential


Vibrations are important to an understanding of the thermal properties of molecules and
solids. Often we model a molecule or crystal with balls and springs. The balls represent
the atomic cores and the springs represent the binding forces due to the valence electrons.
In your mechanics course you solved the problem of the simple harmonic oscillator:

κ = spring constant

κ m

The equation of motion (Newton’s Second Law) describing this system is

F = m d2u/dt2 = – u , (1-26)

where u(t) = x(t) – L is the displacement of the ball from its rest position, and L is the
length of the spring at rest. A solution to this equation is

u(t) = A sin t, (1-27)

which, when plugged into Equation (1-26), yields the angular frequency  = (/m)1/2,
and frequency f = /2.

The vibrational frequency of a “diatomic molecule” is a little more complicated:

m κ m

8
Origins of Mechanical Energy Chapter 1

To compute the vibrational frequency of this object, we must write equations of motion
for ball 1 and 2, involving their displacements u1 and u2,

m d2u1/dt2 = –(u1 – u2)


(1-28)
m d2u2/dt2 = –(u2 – u1).

These are two coupled differential equations. The solution to this problem is given
in Appendix 1. The angular frequency of vibration of the molecule turns out to be
 = (2/m)1/2.

Now imagine three masses arranged in linear order,

m κ m κ m

This linear triatomic molecule has two compressional “modes of vibration” with two
distinct frequencies:

ω = ( κ/ m) 1/2 ω = (3 κ/ m) 1/2

These vibrations are known as the “normal modes” of the molecule because once the
atoms are started in a normal mode, they will continue vibrating in that mode indefinitely.
Check out Appendix 1 to see how normal mode problems are solved by matrix methods.

In one dimension, N masses connected by springs have N – 1 normal modes of vibration.


In three dimensions, N atoms have 3N – 6 normal modes of vibration (see Appendix
1). For a crystal, the number of atoms N is usually a very, very large number, so we can
say quite accurately that N atoms in a crystal have 3N normal modes. The number
of normal modes of a solid is very important to its thermal properties.

9
Physics 213 Elements of Thermal Physics

In a solid with N = 1022 atoms there are 3  1022 normal modes of vibration. The frequen-
cies range from a few kilohertz to about 1012 Hz. The vibrations of lower frequency we
call “sound waves,” or “ultrasound” in the kHz to MHz range. A vast majority of modes
are well above MHz frequencies, so that thermal energy (which at normal temperatures
distributes randomly among all modes) mostly ends up in these high frequency modes.
Now you can appreciate why atomic vibrations in solids are so important to the study
of thermodynamics. The vibrating solid is characterized by specifying the energy present
in each of its normal modes.

10
Origins of Mechanical Energy Chapter 1

Exercises
1) Check your understanding of Section D by considering a mass m attached to a
spring with spring constant . The spring has an unstretched length L, the force
on the mass is F = –u and the potential energy of the spring is PE = ½ u2, where
u = x – L. Starting with the spring stretched to uo and letting the mass go from rest
to velocity v at displacement u, write down the work-energy equation for the two
cases below, showing that they yield the same relation between u and v:
x
Mass alone:

Mass plus spring:

m
2) As an application of the Work-Energy Theorem and the c.m. equation, consider
the case of a car accelerating from rest, as illustrated below. Assume that the ac-
celerating force is F, which is the horizontal force that the road applies to the tires
(and vice versa). Assume that the tires do not slip.

©Hayden-McNeil, LLC

After the car has moved a distance d,

a) What is the velocity of the car?

b) What was the work done on the car?

c) Where did the energy come from that moved the car?

d) How does that energy enter into your equations?

11
Physics 213 Elements of Thermal Physics

Two problems from a former midterm exam will test your understanding of the concepts
discussed in this chapter:

3) Four balls are rigidly connected together with four rods. At which point should
you apply a constant force F in order to produce the greatest initial acceleration of
the center of mass?

a a) point a

b b) point b

c) either point gives the same cm acceleration.

4) Two 4-kg balls of putty are attached to a string of length 2 meters. A constant force
F = 3 newtons is applied to the center of the string and the balls move without fric-
tion. After the force has pulled a distance of 7 meters, the balls collide and stick.

7m

a) What is the center-of-mass velocity of the balls at the instant that they
collide?

b) What is the thermal energy generated in the collision?

12
CHAPTER

Irreversibility and the Second Law


2 of Thermodynamics

A. Thermal Energy
Consider the problem where a small mass (say, an atom) crashes into
a solid. The solid is represented by a system of masses connected by
springs, and the incident mass sticks:

What we find is that the collision excites not just one normal mode, but
a combination of normal modes. Many frequencies are excited, as you
would discover if you recorded u(t) for any one of the atoms and then
took a Fourier transform to extract the frequency spectrum.

This is a simple model of what happens in an inelastic collision. The


“pure” kinetic energy of a single object is converted into the “compli-
cated” internal energy of the many-particle system. If a mass with initial
velocity vi is incident on an object with a N – 1 similar masses and sticks
to it, then momentum conservation (mvi = (Nm)vf) dictates:

vf = vcm = vi/N (2-1)

The initial energy ½mvi2 equals the final energy ½Nm(vcm)2 + Uvib;
therefore, nearly all the energy of the incident particle is converted into
vibrational energy: Uvib = ½ mvi2 (1 – 1/N). In words:

Incident kinetic energy → Induced vibrational energy (2-2)

13
Physics 213 Elements of Thermal Physics

On the microscopic scale, the masses of atoms are extremely small and interatomic
forces are large, so the frequencies of the normal modes (ranging up to about (/m)1/2)
are very high, typically about 1012 Hz. In an inelastic collision the translational kinetic
energy of a small object colliding with a large object is almost completely converted
into thermal vibrations, which are mostly distributed in the high-frequency modes.

In short, the collision generates “thermal energy.” You may be tempted to call these high
frequency vibrations “heat,” but technically the term heat is reserved for the transfer of
thermal energy from one body to another. Sometimes we slip and call thermal energy
by the name heat because it is a common non-technical usage of the term.

B. Irreversibility of Many-Body Systems


We can now see why an inelastic collision is basically irreversible. The conversion of the
kinetic energy of a single particle into the many, many vibrational modes of a solid does
not take place in reverse. The computer simulation with 20 masses gives us a feeling for
this irreversible process, although with a small number of masses the total energy may
eventually be returned to the incident mass if we wait long enough.

It is an experimental fact that in an inelastic collision between two solids one cannot
recover all of the pure translational energy from the complex thermal energy generated
by the collision, even though that would not be a violation of the First Law of Ther-
modynamics (energy conservation). The reason is that nature also obeys a Second Law of
Thermodynamics regulating energy flow in systems consisting of many particles.

The Second Law of Thermodynamics is quite consistent with our intuition. A hot
object resting on a table does not suddenly cool with the result that the object jumps
into the air. This is not because the total thermal energy of the object is small. You will
see in a discussion exercise that if all the thermal energy in an object initially at room
temperature were converted into center-of-mass energy, the object could indeed jump
to quite a large height (many kilometers!). Of course, this would never happen. The
irreversibility of energy flow is the Second Law of Thermodynamics in action.

We have touched upon a basic question in thermodynamics: how much usable energy is
there in a system? Work can be easily converted to thermal energy by friction, but what
fraction of an object’s thermal energy can be converted into work? The Second Law of
Thermodynamics says that it is impossible to convert thermal energy into work with
100% efficiency. Exactly how much work can be extracted from a vibrating solid, or from a
system of moving gas molecules, is a major problem in the subject of thermodynamics,
which we will treat in the context of heat engines. Indeed, the conversion of work to
thermal energy, and thermal energy to work, is the basic issue of thermodynamics.

C. Entropy and the Approach to Equilibrium


Conversion of work into thermal energy, and vice versa, are fundamental processes of
thermodynamics. Another fundamental process is simply the transfer of thermal energy

14
Irreversibility and the Second Law of Thermodynamics Chapter 2

from one system to another, i.e., the process known as heat and designated Q. Start
with two systems with initial energies U10 and U20 and bring them into thermal contact:

U10 U20 U1f U2f

We know intuitively that thermal energy will flow from one system to the other until an
equilibrium condition is reached. The First Law of Thermodynamics only tells us that
the total energy stays constant. There must be some other property of the system that
tells us how the total energy will be partitioned between the two systems in equilibrium.
That property is the entropy. Entropy is an additive function of the two systems, just
like energy. The basic approach of classical thermodynamics is to postulate that the
total entropy

Stot = S1 + S2 (2-3)

is a maximum in equilibrium. Considering U1 as the free parameter (U2 = Utot – U1),


we have

dStot/dU1 = dS1/dU1 + dS2/dU1 = 0. (2-4)

S tot

U1
f
U1

For this closed system dU1 = –dU2 by conservation of energy. Therefore, we may write
the equilibrium condition as,

dS1/dU1 = dS2/dU2 . (2-5)

The term on the left is a property of system 1 and the term on the right is a property
of system 2. Intuitively, we associate “thermal equilibrium” with an equilibration of
temperatures, so it is natural to define the temperature in terms of dS/dU. To retain
our concept of hot and cold, the most convenient choice is,

1/T  dS/dU, or more precisely, 1/T  (S/U)N,V (2-6)

15
Physics 213 Elements of Thermal Physics

which reminds us that particle number N and volume V are held constant. Therefore,
the equilibrium condition is T1 = T2. By this definition, if T1 > T2 energy will flow from
system 1 to system 2 in order to maximize Stot. Defining the derivative as the inverse of
temperature is consistent with both the maximization of entropy of a closed system and
our intuitive concept that thermal energy flows from a high-T object to a low-T object.

The simplest statement of the Second Law of Thermodynamics is that the entropy
of a closed system either

a) remains constant (if the system is in equilibrium), or

b) increases (if the system is approaching equilibrium).

In mathematical terms, as time proceeds,

Stot  0 (2-7)

which is an alternative statement to the Second Law. Here is a summary of the basic proper-
ties of entropy that will be further developed in this course:

1) Entropy is a property of the system – a “state function” like U, V, N, p, and T, and


unlike heat Q and work W that are energies in transit.

2) Entropy for an isolated system is a maximum in equilibrium.

3) Entropy is increased by heat flow Q into a system at temperature T: S = Q/T.

4) Entropy is proportional to the logarithm of the number of accessible microstates:


S = k ln( ), with k = Boltzmann constant, defining the temperature scale.

Entropy is associated with the hidden motions in many-particle systems, in contrast to the collective
motion of the center-of-mass. Because the entropy of an isolated system always increases
or stays the same, nature exhibits irreversibility. A ball resting on a table does not spon-
taneously convert its thermal energy into center-of-mass energy because that would
mean a decrease in entropy. Nor does heat flow spontaneously from cold to hot objects.

An important task of statistical mechanics is to determine the functional form of the


entropy for an N-particle system with energy U and volume V,

S = S(U, N, V). (2-8)

Knowing this function for the particles in question will enable us to compute the
equilibrium conditions, to describe phase transitions, and to determine the capacity
for doing work. A major aim of this course is to gain a microscopic picture of entropy
for some common systems.

16
Irreversibility and the Second Law of Thermodynamics Chapter 2

D. Entropy Maximization and the Calculus of Several Variables


The Second Law of Thermodynamics says that an isolated many-particle system is in
equilibrium when its entropy is maximized. How do we mathematically define this
maximization condition? The answer involves the calculus of several variables, which
is briefly described below.

First consider the simple function y(x) plotted at the right. A maximum in this function
occurs where a small (nonzero) change in x produces zero change in y. In terms of the y(x)

function’s derivative,

⎛ dy ⎞
Δy = ⎜ ⎟ Δx = 0 . (2-9)
⎝ dx ⎠
x
xm
Because x is nonzero, the maximum occurs when the slope of the curve, dy/dx, equals
zero. That is, setting dy/dx = 0 for the known function y(x) yields the value x = xm.*
Entropy, however, is generally a function of many variables. For example, if we knew the
function S(U,N,V) for a system containing N particles, how would we determine the
equilibrium values of U and V? In the case where N is fixed and U and V are variables,
the condition for maximum entropy is:

⎛ ∂S ⎞ ⎛ ∂S ⎞
ΔS = ⎜ ΔU +  ⎜ ΔV = 0.

⎝ ∂U ⎠ V ⎝ ∂V ⎟⎠ U (2-10)

Quantities like (S/U)V are known as “partial derivatives.” (S/U)V is simply the
derivative of S(U,V) with respect to U, treating V as a constant. Consider the prob-
lem pictured at the right. An ideal gas of atoms is contained in a cylindrical volume, F
VA  h. The container and gas are thermally isolated from the surroundings. Mechani- A
cal equilibrium means that the gas pressure p equals the applied force per unit area:
p = F/A. What are the equilibrium values of V and U for a given force (or p)?
h
In Exercise 3 you will solve this problem given the functional form S(N,U,V) for the
ideal monatomic gas and using the definition of temperature introduced in this chapter.
Amazingly, with Eq. (2-10) you will derive two basic properties of an ideal monatomic
gas: its energy U(N,T) and the ideal gas law pV = NkT.

You won’t see many problems with partial derivatives in this book because multivariable
functions such as S(U,V) or S(U1,U2) can often be reduced to one independent variable
by explicitly stating a constraint such as V = constant or U1 + U2 = constant. In general,
however, partial derivatives are a concise way of describing a property of a system under
specified conditions; e.g., (S/U)NV =1/T.

*For y(x) = 4x – x2 (like the figure), you can easily show that the maximum occurs
at xm = 2.

17
Physics 213 Elements of Thermal Physics

Exercises
1) Considering the definition of temperature in terms of entropy, 1/T  dS/dU, which
of the following diagrams is most reasonable for a many particle system? State your
reason. (Hint: Note that dS/dU is positive in all cases and sketch T(U) for each
case.)

S S S

U U U
(a) (b) (c)

T T T

U U U

2) We shall see that the entropy of an ideal monatomic gas depends on energy as S
= (3/2)Nk ln(U), where N is the number of particles, U is the internal energy, and
k is a constant. By maximizing the total entropy S1 + S2 of two gases in thermal
contact, determine the ratio of their energies in equilibrium. Remember, you need
Stot in terms of a single variable (and Utot = constant). Sketch Stot(U1).

U1 U2
Utot = U1 + U2
N1 = 10 N2 = 40

18
Irreversibility and the Second Law of Thermodynamics Chapter 2

3) Work out the problem posed in Section D: Derive U(N,T) for an ideal monatomic
gas, and the ideal gas law, pV = NkT, assuming that entropy has the form:

S = Nk ln(U3/2V)  constants.

Helpful hints: First write S as a function of U plus a function of V (plus constants):

S=

Take the partial derivatives remembering that d(lnx)/dx = 1/x:

⎛ ∂S ⎞ ⎛ ∂S ⎞
⎜⎝ ⎟ =      ⎜
⎝ ∂V ⎟⎠ U
=
∂U ⎠ V
Notice that one of the partial derivatives is directly related to temperature, giving
U(N,T):

U(N,T) =

Assume that the container has negligible thermal energy and note that dU of the
gas is related to dV by the Work-Energy Theorem. Maximize S to find p(N,V,T):

p(N,V,T) =

This problem illustrates how entropy maximization yields equilibrium conditions.


The equilibrium energy is U(N,T) and the equilibrium volume is V = NkT/p =
NkT(A/F). The entropy of an ideal gas is derived later in this course from micro-
scopic properties.

4) Two objects initially at different temperatures are brought into thermal contact.
Show that heat flow Q from the cold object to the hot object violates the Second
Law:

Q → S =
T1 < T2

19
Physics 213 Elements of Thermal Physics

20
CHAPTER

3 Kinetic Theory of the Ideal Gas

A. Common Particles
The gas that you are breathing is composed of a variety of molecules.
Consulting a reference book on the subject, you would find the fol-
lowing facts:

Molecule Mass/mole Concentration


N2 28 g 78%
O2 32 g 21%
Ar 40 g 0.93%
CO2 44 g 0.033 %
H2 2g trace amounts

(numbers are rounded to two significant figures)

One mole of gas contains NA = 6.022  1023 molecules. NA is known


as Avogadro’s constant and is defined as the number of carbon atoms
in 12g of 12C. So, for example, a nitrogen (14N) molecule has a mass,
m = 28 g / 6.022  1023 = 4.65  10–23g.

21
Physics 213 Elements of Thermal Physics

You may recall from your Chemistry course that one mole of gas, no matter what type
of molecules, occupies 22.4 liters of volume at standard temperature and pressure (STP):
T = 273 K and p = 1.01  105 Pa = 1 atm. This rather remarkable fact follows from the
ideal gas law, to be discussed below.

In theory we define an ideal gas as a “non-interacting gas of molecules.” This means,


for example, that we do not consider the potential energy between molecules. In the
view of classical mechanics, the molecules are tiny hard spheres bouncing elastically off
each other and off the walls of the container. Real gases have significant interactions
between molecules that cause phase transitions to liquids and solids, a topic for later
discussion. Ideal gases don’t condense into liquids or solids.

B. Pressure and Kinetic Energy


Pressure is the force per unit area on a surface. The pressure of an ideal gas depends
on the density of the gas and the average kinetic energy of the particles. The relation
between pressure, density, and translational kinetic energy can be determined by con-
sidering a particle bouncing elastically off the walls of a container of volume V = Ad:

Area A
v2 = vx2 + vy2 + vz2

1
Vx KE = _ mv 2
d 2
(translational)

The round trip time for this particle is to = 2d/vx. Each time the particle hits the piston,
it transfers a momentum 2mvx; therefore, the time-average force on the piston is

F = (mvx)/t = 2mvx /to = mvx2/d.

If the container has many atoms, they will have random velocities, so we designate
brackets, <>, to signify the average, or “mean,” value. By symmetry, all three “mean
square components” of velocity are equal:

<vx2> = <vy2> = <vz2> = <v2>/3.

The average force on the piston due to N atoms is F = Nm<vx2>/d, so the pressure is

p = F/A = Nm<vx2>/V,

where V = Ad is the volume of the container. The average energy of a particle is

<KE> = ½ m<v2> = (3/2)m<vx2>;

22
Kinetic Theory of the Ideal Gas Chapter 3

therefore, pressure, number density (n = N/V), and average kinetic energy are related
by the following formula,

2
p = n < KE > (3-1)
3

Notice that KE represents the translational kinetic energy of a particle.

C. Equipartition Theorem
It is important to realize that in this view of classical particles the average translational
kinetic energy is the sum of three average values,

<½ mv2> = <½ mvx2> + <½ mvy2>+ <½ mvz2>, (3-2)

each of which, by symmetry, must equal the same value. If the particle were a diatomic
molecule, such as N2 and O2, there is also a rotational kinetic energy, with an average
value given by:

<½ I12> + <½ I22> (3-3)

There are two terms because there are two possible axes of rotation normal to the mo-
lecular bond. (The quantum mechanical nature of molecules dictates that the energy
corresponding to a rotational axis along the bond is not significant.)

The surprising consequence of statistical mechanics is that each of the “quadratic terms”
in the energy (e.g., <½ mvx2> and <½ I22>) has the same thermal-average value. This
fact is known as the Equipartition Theorem. Sometimes stated “each quadratic degree
of freedom of the system has exactly the same thermal-average energy,” the Equipartition
Theorem is the classical basis for defining a temperature in terms of the microscopic
motions of the particles. We will derive it later in the course.

We empirically define an absolute temperature T such that each of the quadratic terms
has a thermal-average energy given by

<energy per quadratic term> = ½ kT, (3-4)

where k is the Boltzmann constant, 1.381  10–23 J/K, and T is the absolute temperature
in Kelvin. The Boltzmann constant relates microscopic motion to a practical definition
of temperature, the Kelvin scale. (The consistency of this definition with 1/T = dS/dU
given in Chapter 2 will be shown later.)

Note on definition of temperature scales: By international convention, the Kelvin scale is an


absolute temperature scale (0 K is absolute zero) that takes the triple point of water as exactly
273.16 K. This temperature is 0.01 K above the freezing point of water at atmospheric pressure.
The Celsius scale is defined by: degrees Celsius = T(K) – 273.15. At atmospheric pressure, water
freezes at approximately 273 K (0°C) and boils at approximately 373 K (100°C).

23
Physics 213 Elements of Thermal Physics

Equation (3-4) implies that the monatomic particle has an average thermal energy
of (3/2)kT, the diatomic molecule has a thermal energy equal to (5/2)kT, and so on.
Therefore total thermal energies of monatomic and diatomic gases are

3 3
U= NkT = nRT (monatomic gas) (3-5)
2 2
5 5
U= NkT = nRT (diatomic gas) (3-6)
2 2
where N is the number of molecules in the gas, n = N/NA is the number of moles, and

R = NAk = 8.314 J/(molK)

= 1.987 calorie/(molK) (3-7)

= 0.082 literatm/molK

is the ideal gas constant. 1 calorie = 4.184 J is the heat required to raise the temperature
of 1 gram of water at 1 atmosphere from 14.5°C to 15.5°C. Note that

1 atm = 1.013  105 Pa = 1.013  105 N/m2, and

1 literatm = 101.3 Joules.

In summary, the internal energy of an ideal gas can often be written,

U = NkT = nRT, (3-8)

The coefficients  can be experimentally determined by observing how much energy (in
the form of heat) is required to raise the temperature of the gas by 1 degree at constant
volume, i.e., the heat capacity,

CV = (dU/dT)V = Nk = nR, (3-9)

for a temperature range in which  is constant. The heat capacity per mole, or molar
specific heat, is designated by a lower case letter,

cv = R =   (8.314 J/ Kmol). (3-10)

You might wonder why we did not consider the vibrational energy of the diatomic
molecule. Clearly there is a potential energy associated with the molecular bond that
is quadratic in displacement,

<½ u2>, (3-11)

24
Kinetic Theory of the Ideal Gas Chapter 3

where  is the spring constant of the bond, and u is the stretch of the molecular bond
from its equilibrium value. This perfectly valid contribution to the total thermal en-
ergy of the molecule is actually not observed in the heat capacity of common diatomic
molecules at room temperature. At elevated temperatures, however, the contribution
from molecular vibrations does appear, which for the diatomic molecule increases  in
the heat capacity formula to 7/2, considering internal KE and PE. What is going on?

We will see in Chapter 8 that a minimum thermal energy is required to excite the
vibrational modes of a molecule. For molecules N2 and O2 the thermal energy at 300
K is insufficient to get them vibrating. However, for a molecule such as CO2, which
has low frequency torsional modes, the thermal energy at 300 K is sufficient to excite
these vibrations. Consequently, in a gas of CO2 molecules, vibrations do contribute to
the heat capacity at room temperature.

The lesson is that we must be a bit careful in applying Eqs. (3-8) – (3-10) because  (and
thus the heat capacity) is not necessarily constant over wide temperature ranges. For
example, for the diatomic H2, cv/R changes from 3/2 to 5/2 as rotational modes become
thermally active, and from 5/2 to 7/2 as the vibrational modes become thermally active:

7/2 R
Vibration
cv 5/2 R
Rotation
3/2 R
Ideal Diatomic Gas
Translation

10K 100K 1000K T

D. Equipartition Applied to a Solid


As a natural extension of these ideas, we consider the heat capacity of a solid material.
In Chapter 1 and Appendix 1, we saw what the vibrational modes look like for a col-
lection of masses bonded together by springs, analogous to atomic bonds. Because the
vibrational modes have both kinetic and potential energy components, each contributing
½kT, the Equipartition Theorem applied to solids says:

In the classical limit, each normal mode of vibration in a solid has an average thermal energy of kT.

Because there are 3N vibrational modes in a solid containing N atoms, the internal
energy and heat capacity at constant volume are,

25
Physics 213 Elements of Thermal Physics

U = 3NkT and CV = 3Nk. (3-12)

And the molar specific heat is,

cv = 3R = 25 J/Kmol. (3-13)

That was almost too easy. Does it mean that all solids, no matter what their atomic
constituents or bond strengths, have the same heat capacity? Well, almost all:

• The Equipartition Theorem is valid only at sufficiently high temperatures, a con-


dition which may differ from solid to solid. The specific heat of diamond at room
temperature, for example, is considerably less than 3R. The reason is related to the
one given for molecular vibrations of the H2 gas. We shall examine this effect later
in the course.

• This analysis considers only the contribution of the heavy atomic cores. The kinetic
energies associated with “free electrons” in a metal do contribute to the specific
heat, but their effect is only apparent at very low temperatures.

The specific heat of a solid is often given in units J/Kkg. (Multiply by # of moles/kg.)

E. Ideal Gas Law


Having defined temperature empirically, we continue examining the properties of the
ideal gas. Plugging the Equipartition result for the translational kinetic energy, namely
<KE> = (3/2)kT, into Equation (3-1), we immediately have,

p = nkT (n = N/V). (3-14)

The constant  does not appear in this equation, as it does in the total internal energy,
Equation (3-8). The reason is that the pressure depends only on the translational kinetic
energy, not rotation and vibration. The average translational kinetic energy is equal to
(3/2)kT for any ideal gas (even if CV is changing with T), so the 3/2 in this equation
cancels the 2/3 in Equation (3-1) for all ideal gases. The ideal gas law is commonly
written in the forms,

pV = NkT,
(3-15)
pV = nRT,

where N is the number of molecules and n = N/NA is the number of moles in the gas.
The letter “n” is italicized to distinguish it from n = N/V = number density.

26
Kinetic Theory of the Ideal Gas Chapter 3

F. Distribution of Energies in a Gas


We have just seen the need for statistics in our treatment of the ideal gas. In fact, we did
not deal with the actual distribution of molecular velocities but simply used the average
square velocity to characterize the thermal-average kinetic energy,

<KE> = ½ m<v2> = ½ m(<vx2> + <vy2> + <vz2>). (3-16)

For brevity, set KE = E. The average square of velocity doesn’t tell us the actual distri-
butions of velocity or energy. For example, here are several hypothetical distributions
that could give the same average energy of a molecule, marked by the dashed lines:

P(E) P(E) P(E)

E E E

P(E) is called the probability density. Specifically, P(E)dE is the probability of finding a
given molecule within a small energy range, dE. Therefore, P(E) has units of (energy)–1.
The integral over the entire energy distribution (area under the P(E) curve) equals the
probability of finding the molecule in the container,

 P(E) dE = 1. (3-17)

Later in the course, statistical mechanics will show us that the kinetic energy distribu-
tion of molecules in an ideal gas is described by

P(E) = C E1/2 exp(–E/kT) (ideal gas) (3-18)

which looks something like the distribution plotted at the far right. The prefactor C
is determined by setting  P(E) dE = 1, where the integral runs over all values of E.
(Chapter 9(C)) When dealing with single particles, a common unit of energy is the
electron volt (eV), the amount of energy an electron gains in being accelerated across
an electric potential of 1 volt.

1 eV = 1.6  10–19 coulomb  1 volt = 1.6  10–19 J.

In eV units, the Boltzmann factor is k = 8.617  10–5 eV/K.

Because the probability density P(E) has units of (energy)–1, in order to calculate a
probability you must multiply P(E) by an energy interval. This is not so unreasonable:
the probability of finding a molecule with energy exactly 0.01284759839485738 eV is
essentially zero, but the probability that the particle has an energy between 0.011 and
0.013 eV is a definite number, given approximately by

Probability = P(E) E = P(.012)  (0.002) (3-19)


27
Physics 213 Elements of Thermal Physics

Probability is not a fuzzy quantity. Using the probability density, we can determine
many useful measurable quantities of a many-particle system. For example, the average
number of molecules in the small range E around E is simply

N(E) = N P(E)E, (3-20)

where N is the total number of molecules in the container.

Also, the average value, or mean value, of a quantity X for a given probability distribution is:

<X> =  X P(E) dE. (3-21)

For example, the average energy of a particle in an ideal monatomic gas equals:

<E> = C  E  E1/2 exp(–E/kT) dE. (3-22)

Later we will compute this integral and show that it proves the Equipartition Theorem
for an ideal gas.

Notice that the probability density for the ideal gas contains the factor exp(–E/kT).
This function is known as the Boltzmann factor and occurs in many problems of sta-
tistical mechanics. At this point, I hope that you are asking yourself, “What’s the origin
of the Boltzmann factor?” and “Where does the factor of E1/2 come from?” When you
complete this course you will know the answers to these key questions and many more.

28
Kinetic Theory of the Ideal Gas Chapter 3

Exercises
1) In Chapter 2 we defined temperature in terms of the change in entropy per unit
energy, 1/T = dS/dU. From kinetic theory we have U = (3/2)NkT for the ideal
monatomic gas. Using these relations, determine the change in entropy S2 – S1 for
the ideal gas as its energy increases from U1 to U2 at constant volume. What is the
functional form of entropy for an ideal monatomic gas? (Hint: Eliminate T from
the two equations and integrate, remembering dx/x = ln(x).)

2) a) Assuming that your lung capacity is 2 liters, calculate approximately the total
kinetic energy of the gas in your lungs.

b) What is the weight of 2 liters of N2 gas at 300 K in units of a penny (1 gram)?

c) How high would a penny have to be dropped in order to reach a c.m. kinetic
energy equal to the total energy of gas molecules in your lungs (neglect air
resistance)?

d) Compare the final velocity of the penny to the average velocity of N2 molecules
at 300 K.

e) How much heat is required to raise the gas in your lungs 3 K?

3) If you drop a 1 kg block of aluminum from a height of 1 meter and assume that all
the center-of-mass kinetic energy of the block is converted into heat as it strikes the
floor, what is the rise in temperature of the block? (Al has a molar mass of 12 g.)

29
Physics 213 Elements of Thermal Physics

4) a) Plot a few points of the function x1/2e–x as a function of x and indicate where
the peak and average energies are.

x .1 .25 .5 1.0 1.5 2.0 3.0


x1/2e–x

.4

x 1/2 e -x
.2

0 x = E/kT
0 1 2 3

This is the function describing the distribution of kinetic energies of an ideal gas,
(E/kT)1/2e–E/kT at temperature T, as introduced in this chapter and explained in
detail in Chapter 9.

b) Estimate the probability that the particle will have an energy E between 0.5
kT and 1 kT. [Hint: Approximate P(E)E by a rectangle of width 0.5 kT and
compare its area to the total area under the curve.]

30
CHAPTER

4 Ideal-Gas Heat Engines

A. The First Law of Thermodynamics


When work is done on a system by its surroundings, conservation of
energy takes the following form:

Won = (KEcm) + U. (4-1)

All thermodynamic systems that we will consider have their center of


mass at rest, so (KEcm) = 0, and U is the total energy of the system in
the c.m. frame, i.e., the internal energy. However, if thermal energy is
transferred to the system from its surroundings, there is another input
term on the left. The transfer of thermal energy into the system is known
as heat and is designated by Q. The First Law of Thermodynamics
states: Internal energy is a state function, work and heat are forms of
energy transfer, and total energy is conserved:

Total energy inputs =  (total system energy),


or

Won + Q = U. (FLT) (4-2)

In dealing with engines, it is convenient to define the work done by the


system. In this case, the First Law is written,

U = Q – Wby . (FLT) (4-3)

31
Physics 213 Elements of Thermal Physics

In other books you may find the first law written as U = Q + W, where it’s implied that
W is the work done on the system. In this course we find it is best to simply label W
with a subscript to avoid confusion. We do not bother with a subscript for Q, defining
Q = heat input to the system.

B. Quasi-static Processes and State Functions


In dealing with work and heat engines, we will consider only quasi-static processes of a
homogeneous gas. A “homogeneous” gas is one in which the pressure and temperature
are uniform throughout, and “quasi-static” means that the pressure and temperature
are well-defined at all times. So, an example of a process that is not quasi-static is one
in which the gas piston moves faster than the gas can react. Quasi-static basically means
“slow” compared to the “settling time” of the gas (and its surroundings).

This statement implies that, given sufficient time, the gas will settle into a well-defined
condition, known as equilibrium. In this context, a quasi-static process is one in which
the gas is always very close to equilibrium with itself and its surroundings.

We envision the gas confined to a container with a movable frictionless piston:

State Functions:
N, V, U, p, T

Process energies:
Q, W

Recall that work and heat are energy in transit. They are “process energies” that cause
a change in the state of the system. The internal energy of the gas, the number of
molecules, the volume, the pressure, and the temperature are state functions that are
always well-defined when the system is in equilibrium.

C. Isothermal and Adiabatic Processes—Reversibility


The most important construction for characterizing the thermodynamic processes of
gases is the pV diagram. A quasi-static process can be drawn as a curve on this diagram.
In proceeding from point A to point B on this curve, the work done by the gas is
simply the area under the curve,

Wby =  p dV. (4-4)

The two most important processes that we will study are the isothermal process and
the adiabatic process. In the isothermal process, the gas is always in equilibrium with a ther-
mal reservoir, which can exchange heat Q with the gas. In the adiabatic process, the gas is
thermally isolated from its surroundings.

32
Ideal-Gas Heat Engines Chapter 4

The isothermal process (T = constant), is represented by p = NkT/V,

FLT: W by = Q because
p
U = α Nk Δ T = 0

The adiabatic process (Q = 0), is represented by p = (constant)/V,

FLT: W by = -ΔU because Q = 0


p

(γ = 5/3 for a monatomic gas)

The origin of the relation pV = constant for an adiabatic process follows from the First
Law of Thermodynamics (here, U = – Wby), the Equipartition Theorem, and the ideal
gas law. For infinitesimal changes in T and V,

 Nk dT = – p dV = – (NkT/V) dV

which implies,
 dT/T = – dV/V

  (dT/T) = –  (dV/V)

 ln T = – ln V + constants

33
Physics 213 Elements of Thermal Physics

yielding,

VT = constant, or (adiabatic process)


pV = constant

with  = ( + 1)/. These functional forms apply only in the temperature range where
 and  are constant.

Performed in quasi-static fashion (slowly), the isothermal and adiabatic processes de-
scribed above are reversible. In the isothermal process, involving energy exchange with a
thermal reservoir at the same temperature as the gas, the gas is able to perfectly transfer
heat from the reservoir into work (Q = W), and vice versa, without any change in the
internal thermal energy. In the adiabatic process, pushing the piston down causes the
gas to heat up. As the piston moves back to the original position, the gas does work on
the surroundings at the expense of internal energy, and the gas temperature decreases
to its original value. Perfectly reversible.

It is important to note that in order to determine whether a process is reversible or


not, we must consider any changes in the surroundings. Consider the following process. Is
it reversible?

p2

p1

Obviously, no work was done in this process because there was no volume change.
However, the pressure and temperature have changed by an input of heat. The initial
temperature is T1 = p1V/Nk and the final temperature is T2 = p2V/Nk. In the simplest
case, this heat came from a single thermal reservoir at temperature T2. The gas initially
at T1 was brought into contact with this reservoir, and the gas+reservoir approached
equilibrium by the transfer of heat.

The process described in the last paragraph is clearly irreversible. The gas will never
cool down and give the energy back to the reservoir. There is a fundamental principle
at work here (another statement of the Second Law):

Any process that involves an exchange of heat between two systems at different
temperatures is irreversible.

34
Ideal-Gas Heat Engines Chapter 4

Actually, it is possible to envision a process that looks almost like the one in the graph
above but is reversible. Imagine that the straight line is replaced by a lot of little seg-
ments of alternating isothermal and adiabatic processes:

p2

p1

This series of processes is reversible, but there is a cost to pay. If we use n isothermal
segments, then we need n thermal reservoirs at successively higher temperatures, Tn.
Plus, we would need to remove and connect them before and after each adiabatic
process. [Aside: More practically, one can reversibly transfer energy from a reservoir
at temperature T2 to a gas at T1 by using a second gas and piston. The temperature of
this secondary gas is adjusted adiabatically to match the temperature of the reservoir
or primary gas so that heat is always transferred isothermally. See Carnot cycle below.]

During isothermal and adiabatic processes, some important state functions of the system
remain constant. Let’s see what they are.

Isothermal processes are represented by the family of curves p = NkT/V for different
reservoir temperatures T:

Isotherms (ideal gas):


pV = NkT = constant
p U = constant

T3

T2

T1

35
Physics 213 Elements of Thermal Physics

If the isothermal process is performed slowly, the gas is arbitrarily close to thermal
equilibrium with the reservoir at constant T. The transfer of heat occurs between two
systems at nearly the same temperature. (If the temperature were exactly the same, no
heat would flow.) A slight expansion of the gas (doing work) is accompanied by a slight
cooling (causing heat flow), and this way the gas converts heat from the reservoir to
work on the system holding the piston. The isotherms are constant-energy curves
for the ideal gas.

Is there a quantity like energy that remains constant during an adiabatic expansion or
contraction of the ideal gas? Well, that is the million-dollar question.

Adiabats (ideal gas):


γ
pV = constant
p

S3
S2
S1

The answer is yes. There is a state function of the gas—just as important as internal
energy—which is constant during an adiabatic process. It is none other than the entropy,
designated by S. The adiabats are constant-entropy curves. Let us briefly see how
this happens.

D. Entropy of the Ideal Gas—a First Look


Before statistical mechanics, scientists and engine designers realized that there was a
state function closely associated with the adiabatic process. Consider the First Law of
Thermodynamics for small energy transfers,

dQ = dU + p dV. (4-5)

It is clear from this equation that Q itself is not a state function because pV depends
on the path taken between two points on the pV diagram. (State functions, such as U,
V, p, and T, are well defined for every point on the diagram.) Like work, dQ is not a
“differential” of a function—a fact that is emphasized in some texts by using the symbols
dQ and dW. Using the definition of heat capacity, dU = CV dT, and the ideal gas law,

dQ = CV dT + (NkT/V) dV.

36
Ideal-Gas Heat Engines Chapter 4

Now here is the creative step: If this equation is divided by T, the right side becomes
an exact differential; i.e., it represents a change in a state function S:

dQ/T = CV (dT/T) + Nk (dV/V)

= d(CV ln(T) + Nk ln(V)) (d(lnx)/dx = 1/x)

= dS

with the definition,

S = CV ln(T) + Nk ln(V) + constant. (ideal gas with fixed CV) (4-6)

S is the entropy. Using CV = Nk and  = ( + 1)/, you can show that S = Nk ln(VT)
+ constant = Nk ln(pV) + constant. These equations show that entropy is constant for
an adiabatic process (Q = 0) and that a change in entropy is closely linked to the heat
input to the system:

dS = dQ/T , (4-7)

which is true even if CV is a function of T. The adiabats shown in the figure above are
constant-entropy curves. We shall see later that this logarithmic form of entropy for the
ideal gas has the requisite properties that we postulated for entropy in Chapter 2. We will
discover the microscopic origin of this special property through statistical mechanics.

E. Converting Heat into Work


The principal purpose of a heat engine is to convert heat into work. At the outset, you
might ask what the practical limitations of this process are. After all, we have just seen
that an isothermal process is able to convert heat into work (Q = W), and vice versa, with
100% efficiency! Specifically, for an isothermal process, the work done by an ideal gas is

Vb Vb
dV V
Wby = ∫ pdV = NkT ∫
Va Va
V
= NkT ln b
Va
(4-8)

p W by = area under curve


Isothermal Process: W by = Q
T = constant
(Ideal gas)

V
Va Vb

37
Physics 213 Elements of Thermal Physics

From this property of an isothermal process, one might naively expect that a properly
designed heat engine could have nearly 100% efficiency. Unfortunately, once we have
done work with the expanding piston, we must reset the system so that more work can
be done. Exactly how the system is reset determines the overall efficiency of the heat
engine. Efficiency equals (work done)/(heat used) over a complete cycle of the engine.

All practical heat engines undergo a cyclic (closed loop) process. In Appendix 2 we
consider a specific engine cycle known as the Stirling cycle. It employs the isothermal
and isochoric (constant-volume) processes that we have just considered. As an exercise
you will calculate the efficiency of such an engine working between two thermal reser-
voirs, one at 0°C and the other at 100°C (i.e., 273 K and 373 K). A general diagram for
a heat engine is shown below. In this model, a source of useful energy is represented by
a thermal reservoir at temperature Th and the environment is represented by a thermal
reservoir at temperature Tc. The engine works between these two reservoirs extracting
useful energy from the hot reservoir, doing work with it, and dumping excess heat into
the cold reservoir:

Thermal Reservoir at T h

Qh

Heat Engine: W by = Q h - Q c

Qc

Thermal Reservoir at T c

For reasons that we will see later in the course, the most efficient engine operating
between two temperatures is one that uses only reversible processes (isothermal and
adiabatic). It is called the Carnot Engine. The Carnot engine is the standard by which
we measure all practical engines. No cycle is more efficient than the Carnot cycle.

The Carnot cycle consists of two isothermal and two adiabatic processes:

adiabats

1
Carnot cycle
p

2
Th
4
Tc isotherms
3

38
V
Ideal-Gas Heat Engines Chapter 4

The efficiency of the Carnot cycle can be determined using the relations derived above.
The isothermal work by the gas is

V2
W12 = NkTh ln = Qh
V1 (4-9)
V
W34 = NkTc ln 4 = −Q c .
V3

The input heat Qh from the hot reservoir and the output heat Qc to the cold reservoir
are defined as positive. The adiabatic processes obey

V1 Th = V Tc
α
4
α

V2 Th = V3 Tc
α α
(4-10)

which implies that

V1 V4
= (4-11)
V2 V3
Therefore, from Eqs. (4-9),

Q h Th
= (Carnot cycle) (4-12)
Q c Tc
The total work done by the gas in any closed cycle (not just Carnot) is simply the dif-
ference between the heat absorbed and the waste heat expelled,

Wby = Qh – Qc (in general) (4-13)

because U = 0 for a closed cycle. Therefore, the efficiency of any heat engine is de-
fined by

=
Wby Qh − Qc Q
= = 1− c (in general) (4-14)
heat input Qh Qh

For the Carnot cycle, this result becomes

Tc
=1– . (Carnot efficiency) (4-15)
Th

Notice from Eq. (4-12) that Qh/Th = Qc/Tc in the Carnot cycle. Q/T is something spe-
cial that is conserved in this reversible cycle. Yes, it’s the entropy. The fact that Q/T is
related to an intrinsic property of the system was realized long before the microscopic
origin of entropy was discovered.

39
Physics 213 Elements of Thermal Physics

In working with thermodynamic cycles, you will need to determine the work performed
by the gas for an adiabatic process. You should be able to derive the required equation
by applying the concepts already introduced in this Chapter. The First Law of Ther-
modynamics tells us

U = – Wby , (adiabatic process). (4-16)

For an ideal gas, U = NkT, implying that

Wby = Nk (T1 – T2). Adiabatic work (4-17)

Using the ideal gas law, pV = NkT,

Wby =  (p1V1 – p2V2). Adiabatic work (4-18)

F. Refrigerators and Heat Pumps


The Carnot engine run backward (i.e., reversing the arrows) is a refrigerator or heat
pump. In these devices, we provide work to make heat flow from a cold reservoir to a
hot reservoir. This is not a violation of the Second Law of Thermodynamics, which says
that heat will not spontaneously flow from hot objects to cold objects.

To appreciate these processes, consider the following questions (circle your answers
and we will see how good your intuition is):

1) A heat pump can extract heat from the cold outdoors to warm the inside of your
house. a) True, b) False.

2) The amount of work needed to overcome a heat leak Q from the inside of the house
to the colder outdoors is

a) less than Q,

b) equal to Q,

c) greater than Q.

The following diagrams show the distinction between a heat engine and a refrigerator
(used to cool your food) or a heat pump (used to warm your house in winter or cool
it in summer). For a refrigerator, room temperature is Th, and for a heat pump, room
temperature is Th in winter or Tc in summer. And, yes, as improbable as it sounds, a heat
pump can extract heat from a cold winter day. All it takes is work, usually provided in
the form of electrical energy.

40
Ideal-Gas Heat Engines Chapter 4

Q h and Q c defined
as positive.
Arrows (and signs
Qh in equations) show
Qh
p p direction of flow.

Th Th

Tc Tc
Qc V Qc V

Heat Engine Refrigerator or Heat Pump

Here is an example of a refrigerator problem: A refrigerator keeps the food cold at 5°C
despite a heat leakage of 100 J per second (= 100 W), which is compensated by Qc = 100 J
per second. Assuming that it has a Carnot efficiency and that the ambient temperature
is 20°C, what electrical power is required to run this device?

We begin the fridge problem by making a simple sketch like the one shown at the right. Qh
The First Law tells us that the work done on the gas over the entire cycle is:

Won = Qh – Qc . (4-19)
W on
In these cyclic problems, we use this energy-conservation equation to calculate power
(energy flow) in joules/sec = watts. For a Carnot cycle,

Qc Tc Q c = Q leak
= (4-20)
Q h Th
In order to keep the fridge at a constant Tc = 5°C, we must remove Qc = 100 J per second
(offsetting the 100 W leakage). Therefore we write:

⎛T ⎞
Won = Qc ⎜ h − 1⎟ (4-21)
⎝ Tc ⎠
Remembering to use Kelvin units, you will find that only 5.4 J of input work per second
(5.4 watts of power) is required to overcome the 100 W leakage. Amazing. Sounds like
a violation of some law, but remember we are using the work simply to transfer energy
from one reservoir to another. Energy conservation is not violated. The result is cor-
rect. A more refined way of showing the energy flow for refrigerator or heat pump is:

41
Physics 213 Elements of Thermal Physics

Fridge Problem: Heat Pump Problem:

Kitchen (20˚C) Thermal Reservoir at T h House (20˚C)

Qh heat leak = Q h

W on
heat leak = Q c
Qc

Fridge (5˚C) Thermal Reservoir at T c Outside (0˚C)

Notice that the input “cross sections” add up to the output cross section (Won  Qc =
Qn). Reverse the arrows and change Won to Wby and you have a heat engine.

Now, do you want to reconsider your answer to question 2 above? Let’s say that we need
to offset 100 watts heat leakage through the windows with Qh = 100 J per second from
the heat pump in order to keep the interior of the house at 20°C when the outside air
is at 0°C. What electrical work must be done to overcome this leakage?

Complete the sketch at the left following the rules: a) the hot reservoir is always at the
top and the cold reservoir at the bottom, b) the arrows decide the directions of energy
flow, so all quantities are positive, c) depending on the problem, the heat leak could be
from the hot reservoir (Qleak = Qh) or to the cold reservoir (Qleak = Qc), d) the form of
the conservation equation (Won = Qi – Qj) follows from the flow diagram.

As before, write the conservation equation in terms of the one known heat flow (Qc or
Qh) and use the Carnot ratios:

W = Q( )

The answer I get is 6.8 watts. Interesting, eh? Only 6.8 watts of input power are needed
to overcome 100 watts of leakage. Did you answer the question correctly?

42
Ideal-Gas Heat Engines Chapter 4

Exercises
1) Work through each process of the Stirling cycle and determine the heat transferred
to the gas Qi, the work done by the gas Wi, and the internal energy change
Ui. Assume that the gas is 0.1 mole of argon and Th = 373 K, Tc = 273 K and
Vb/Va = 2. Fill out the table in units of joules, and check that the First Law of
Thermodynamics, Qi – Wi = Ui, is obeyed in each process. Determine the efficiency,
 = work/heat-input = (W2 + W4)/(Q1 + Q2).

1 2

Th
3
4 Tc isotherms

Va Vb V

Process (i) Qi Wi Ui


1
2
3
4

43
Physics 213 Elements of Thermal Physics

2) Nitrogen gas initially at 300 K and atmospheric pressure in a volume of 1 liter is


adiabatically expanded to 2 liters, then isothermally returned to 1 liter. How much
heat is required to restore the gas to its initial conditions? Suggestion: Draw the
p-V diagram and complete the following tables:

point p V T
a
b
c

process Q Wby
1
2
3

3) a) Using the FLT (Eq. (4-5)) and the ideal gas law show that heat capacity at constant
volume is CV = (dU/dT)V and heat capacity at constant pressure is Cp = CV + nR.

b) For an ideal gas with U = nRT show that Cp = CV , with  = ( + 1)/.

c) The temperature of a thermally isolated ideal gas increases by a factor of 1.219


when its volume is decreased by a factor of 2. What are CV and Cp for one mole
of this gas (i.e., the specific heats cv and cp)? Describe the gas.

44
CHAPTER

Statistical Processes I:
5 Two-State Systems

A. Macrostates and Microstates


The word “state” is probably the most widely used term in the study of
thermodynamics. In fact, there are two quite different ways to describe
the state of a system in equilibrium. The first is its macrostate, which is
a specification of its large-scale properties, such as U, V, p, T. The second
is its microstate, which is a detailed specification of the condition of
all of its particles. We will illustrate these two important concepts with
two examples: a system of spins, and an ideal gas of particles.

The problem boils down to statistics, like a game of dice. When you
roll two dice, each with equal probability of turning up 1 through 6,
is it more probable that their sum will be “2” or “7”? Answer: There is
only one way of rolling “2” (1+1), but there are 6 ways of rolling “7”
(1+6, 2+5, 3+4, 4+3, 5+2, and 6+1). Therefore, rolling “7” is six times
more likely than rolling “2”. In statistics, each possible roll (say 2 + 5) is
a microstate, and the collection of all those combinations which yield a
sum of “7” is a macrostate.

A simple application of statistics is flipping a coin. In N tosses, what is


the probability of getting Nh heads and Nt = N – Nh tails? To answer
this problem, we have to know the number of ways one can throw Nh heads

45
Physics 213 Elements of Thermal Physics

in N tosses, which we denote as (N, Nh). This problem is commonly known as the “n
choose m” problem (here “N choose Nh”) and has the solution:

N! N!
Ω( N, N h ) = =
N h ! N t ! N h !( N − N h )!

where, N! is defined by the example: 5! = 5  4  3  2  1 = 120. This equation


represents the Binomial Distribution for two-state systems.

There are many important applications of the binomial distribution that do not involve
gambling. In the Introduction, we considered the behavior of particles in a two-cell
box. In this chapter, we will examine a system of spins (pointing up and down) and the
random walk (left and right steps). The latter problem is directly related to diffusion
of atoms in gases and solids. We begin with the spin problem.

B. Multiple Spins
Magnetism is an important property of materials with countless applications. Just con-
sider the uses in your immediate vicinity: solenoids, power transformers, watch motors,
audio headphones and speakers, tape drives, disk drives, etc., etc.

Magnetism has its origin in the spin of an electron, with an associated magnetic mo-
ment. If the electron, with magnetic moment , is placed in a magnetic field, quantum
mechanics tells us that it will be pointed either parallel or anti-parallel to the magnetic
field. The lowest energy state is when the moment is pointing along the field (this actu-
ally corresponds to the electron spin pointing opposite the field, but we will adhere to
the usual convention of using the term “spin up” meaning “moment up”):

or
B
Energy = - μ B Energy = μ B
"spin up" "spin down"

Just how electron spins “add up” or “cancel out” in an atom, a molecule or a solid is the
subject for advanced courses. Here we simply consider a collection of non-interacting
spins, each with moment . As an example, imagine the following arrangement of 9
spins, labeled by their site:

B
1 2 3 4 5 6 7 8 9

46
Statistical Processes I: Two-State Systems Chapter 5

From now on we will omit the numbers and use the position of the spin to designate
its site.

You may wonder why all the moments in the state shown above are not pointing along
the magnetic field, which is the lowest energy state of the system. Well, this is how we
happened to set up the particular system, which is isolated from other sources of energy.
In this system, two electrons of opposite spin can undergo a mutual spin flip,

but a single spin flip is not allowed because it would require adding or subtracting a
magnetic energy of 2B from the system.

A specific arrangement of moments such as the one shown above is known as a mi-
crostate of the 9-spin system: the orientation of each one of the 9 spins is specified. In
contrast, the macrostate of the system is a specification of the total magnetic moment
of the system:

0 =  (±)
(5-1)
= 5 – 4 = 1

for the situation depicted above. The total magnetic moment is typically what we would
measure in an experiment.

Notice that there are many possible microstates for a given macrostate of the system.
In particular, for 0 = 1, several other microstates are:

The total number of microstates for “5 spins up” and “4 spins down” is

9!
= 126
5! 4 !

47
Physics 213 Elements of Thermal Physics

In statistical language this combination is called “9 choose 5.” Each of these 126 micro-
states provides exactly the same total magnetic moment, or macrostate.

Not all values of the macroscopic parameter 0 have the same number of microstates.
For example, there is only one microstate for 0 = 9:

In summary, we designate a macrostate by:

(N, Nup)

where Nup designates the number of “up spins.” The number of down spins is obviously
N – Nup. All of the microstates in a given macrostate have the same total magnetic mo-
ment and the same total energy. The number of accessible microstates for a given
macrostate (N, Nup) is:

N! N!
Ω( N, N up ) = = (5-2)
N down N up ! N down ! N up !( N − N up )!
N up
The macrostate may be specified either by Nup or by the total magnetic moment 0,
which is proportional to the integer m = Nup – Ndown.
m
0 m  = (Nup – Ndown)  = (2Nup – N) , (5-3)

where m is the “spin excess.” The product m is the total moment along the field direc-
tion. The total energy of the spin system is U = – 0  B = – 0B = –mB. You can easily
show that may also be written in terms of m:

N!
Ω( m) = (5- 4)
⎛ N + m ⎞ ⎛ N − m⎞
⎜⎝ ⎟ !⎜ ⎟!
2 ⎠ ⎝ 2 ⎠

This statistical distribution is known as the Binomial Distribution because of its 2-state
nature (spin up/spin down). Continuing with our analysis of the 9-spin problem, we can
make the following plot of the number of microstates for the ten macrostates:

48
Statistical Processes I: Two-State Systems Chapter 5

Ω(N,N up )
or
Ω(m)

Number of up spins: N up = 0 1 2 3 4 5 6 7 8 9
Spin Excess: m = M/μ = -9 -7 -5 -3 -1 1 3 5 7 9
# microstates: Ω(m) = 1 9 36 84 126 126 84 36 9 1

(Notice that Nup ranges from 0 to N in steps of 1 and m ranges from -N to +N in steps of 2)

We can now convert this distribution of microstates into a probability distribution by


assuming the fundamental postulate of statistical mechanics, namely that every acces-
sible microstate is equally likely. With this assumption, what is the probability of finding
a total magnetic moment of 0= 3? Write your answer here:

P(m) = P(3) =

Give it a try before reading on. (This calculation corresponds to the limit B → 0. We
will consider non-zero fields in a later chapter.)

You probably computed this by dividing the number of microstates in the “m = 3” mac-
rostate by the total number of microstates. As you can verify by adding up the numbers
in the 9-spin problem, the total number of microstates for a binomial distribution turns
out to be:

Σ Ω( N , N up ) = Σ Ω( m ) = 2N (5-5)
N up m

Can you give a simple argument why the total number of microstates (including all the
macrostates) of an N-spin problem is 2N?

You should find that the probability of finding 0= 3 is 0.164. In other words, there
is 16.4% chance of finding exactly 3 spins up. Now we can write a formula for P(m) in
terms of (m) and N:

P(m) = (m)/2N (5-6)

In the discussion section you will be asked to repeat this exercise for a different num-
ber of spins. Convince yourself of the validity of the (N, Nup) or (m) functions by
counting the possible microstates for each macrostate.

49
Physics 213 Elements of Thermal Physics

For large numbers of spins, it is very useful to approximate the Binomial distribution
by a Gaussian distribution. As detailed in Appendix 3, the probability becomes:

(m) = (2/πN) 1/2 exp(-m 2 /2N)

(2/πN) 1/2

.607(2/πN) 1/2

m
0 N 1/2
Give it a try for the nine-spin problem. For example, I find (3) = 2N P(m) = 82.6,
which is pretty close to the exact value given in the plot of (m) above. The agreement
is even better for larger N. Moreover, for N > 100 or so, your calculator will probably
choke on the factorials, requiring you to use the Gaussian form.

C. The Random Walk Problem—Diffusion of Particles


This classic problem is often stated in terms of a drunk staggering away from a lamppost.
He takes M steps of length x, but each step is in a random direction along the sidewalk.

This one-dimensional random walk is mathematically equivalent to the spin problem.


The step size x corresponds to the spin moment , and the number of right and left
steps correspond to Nup and Ndown. The total displacement x = m x of the drunk from
the lamppost after M steps corresponds to the total magnetic moment 0 = m for N
spins. We change N to M here because in problems dealing with particles it is usual to
represent the number of particles by N. M = total number of steps.

The drunk is used as an example to get your attention in the potentially soporific field
of statistics. In fact, there are many useful problems involving the mathematics of the
random-walk problem, especially when we expand it to three dimensions. In this course,
we will consider the diffusion of N molecules in the atmosphere and the diffusion of N
electrons and impurity atoms in a semiconductor.

You can probably guess that, after M steps, the mean (i.e., average) displacement away
from the lamppost at x = 0 is:

<x> = <si> = 0, (5-7)

where the step size si = ± x and i ranges from 1 to M in the sum. This result doesn’t
mean that the drunk will always end up back at the lamppost after M steps; but it does

50
Statistical Processes I: Two-State Systems Chapter 5

mean that, if he repeats this random process night after night; on average he will end
up equally to the right or to the left of the lamppost. At least the dude isn’t driving.

For a random-walk process with a constant step size, the mean square displacement after
M steps is,

<x2> = Mx2 , (5-8)

as proven in Appendix 3. As indicated in the last section, for a large number of trials
(many drunks, or one drunk many times), each taking many steps, the binomial distri-
bution is well approximated by a Gaussian function:

N(x)

No

σd = <x 2 > 1/2 = M 1/2 x


= standard deviation

x
0

Here is how we interpret this graph: If we observed this behavior for N drunks, record-
ing each of their displacements after M steps, then N(x) equals the number of drunks
that ended up at position x.

This is quite similar to the problem of a molecule released at a time t = 0 and at a posi-
tion (0,0,0) in a room filled with some other type of gas. The molecule travels with an
average thermal speed v, scatters randomly from the other molecules, and at time t ends
up at position (x,y,z). The thermal speed is given by Equipartition: ½ mv2 = (3/2) kT.

(x,y,z)
(0,0,0)

51
Physics 213 Elements of Thermal Physics

We take  to be the average distance traveled between collisions, or “mean free path.”
Because 2 = x2 + y2 + z2 and all three directions are equivalent, we see that 2 = 3x2.
The average x-projection of this ballistic path, x =  /31/2, is roughly equivalent to the
drunk’s step size on the sidewalk, and similarly we assume that each collision random-
izes the molecule’s direction.

The mean time between collisions is = /v. After an elapsed time t, the average number
of collisions is M = t/ . As shown in Appendix 3(D), the mean square displacement for
random step sizes is <x2> = 2 Mx2, in contrast to Mx2 for a constant step size (Eq. 5-8).
Therefore,

<x2> = 2 (t/ ) x2 . (5-9)

If we release N molecules at time t = 0 and position (0,0,0) and then plot their posi-
tions at a time t, we will find a Gaussian distribution for the number of molecules per
unit distance,

N(x) = N P(x) = N (2 d2)–1/2 exp(–x2 /2 d2). (5-10)

This distribution has a mean square displacement that depends linearly on time,

d2 = <x2> = (2x2/ ) t = (22/3 ) t. (5-11)

It is customary to define a diffusion constant as

D = (2/3 ) = v  /3 , (5-12)

using  = v for the latter form. The root-mean-square (rms) displacement of the molecules
at time t is

xrms(t) = d(t) = (2 D t)1/2. (5-13)

This is the basic result of a diffusion process. The size of the cloud of diffusing particles
increases as the square root of the time. The formula for the expanding distribution is,
2
x
N –
N(x,t) = e 4Dt , (5-14)
4Dt

which is plotted for several times after the molecules are released:

52
Statistical Processes I: Two-State Systems Chapter 5

t=0
t=0.3
N(x,t) t=1
t=3
t=10

-10 -8 -6 -4 -2 0 2 4 6 8 10

It is interesting and useful to notice that the expanding Gaussian distribution is a solu-
tion to the following differential equation:

dN d 2N
=D 2 (5-15)
dt dx

where N = N(x,t). You will be asked to prove this as an Exercise. This famous equation
is known as the diffusion equation.

The rms displacement along x of a particle in time t is (2Dt)1/2. Diffusion in 3 dimensions


yields a square displacement, <r2> = <x2> + <y2> + <z2> = 6Dt, by symmetry. The rms
diffusion radius after time t is

rrms(t) = (6Dt)1/2 . (5-16)

Notice that the mean volume occupied by the expanding cloud of diffusing molecules is

Vol  (4/3) r3rms  4(6Dt)3/2. (5-17)

If N particles are deposited initially at r  0, then the average density of particles de-
creases as

nav(t)  N/Vol  N/4(6Dt)3/2 (5-18)

53
Physics 213 Elements of Thermal Physics

The average number of particles remaining in a small volume dV at the origin after
time t is approximately navdV, assuming dV << Vol.

The equations (5-16 thru 5-18) apply only when the mean diffusion distance is much
larger than the size of the initial region into which the particles were deposited. If the
particles were deposited in a Gaussian distribution with standard deviation o, then
the expanding distribution in one dimension has the mean square deviation d2 = o2 +
2Dt, which is also valid at early times. Also, these equations are valid only when d and
rrms are much larger than .

Sometimes you will want to know the typical time it takes for particles to diffuse a given
distance. First find the diffusion constant in terms of the mean free path and thermal
velocity, then use equation (5-13) or (5-16) to solve for a typical time t for a particle to
arrive at distance xrms(t) or rrms(t). There are a lot of practical problems involving dif-
fusion, such as the migration of impurities in a semiconductor chip, or the diffusion of
carriers in a semiconductor. The few exercises given below will give you some experi-
ence with these important technological problems.

D. Heat Conduction
The second process that allows a system to approach equilibrium is thermal conduc-
tion. This process has similarities to particle propagation (or “particle conduction”) in
the presence of an external force field, such as for electrical conduction of electrons in
a semiconductor or metal. In solids at normal temperatures the principal entities con-
ducting heat are packets of vibrational energy, also known as phonons. Both particle
conduction and heat conduction are based on the random-walk process and therefore
have similar mathematical descriptions.

Heat flows from a hot region to a cold region if there is a temperature gradient. A tem-
perature gradient in the x direction given by dT/dx will cause a heat current density
(energy transfer per unit area per unit time) designated by Jx. In fact, the rate of energy
flow is proportional to the temperature gradient, as given by the heat conduction law:

dT
J x = −κ (5-19)
dx

where  is known as the thermal conductivity and depends on the material being
considered and even the temperature. Can you see why a positive gradient in T gives
a negative Jx? (Answer: Heat flows from high T to low T.)

54
Statistical Processes I: Two-State Systems Chapter 5

For problems where the change in temperature across the material is small compared to
the absolute temperature, the magnitude of the heat-current density (Eq. (5-19)) becomes:

J =  T/x, (5-20)

where we drop the sign, knowing which way heat will flow. Let’s calculate the heat leak
for a “thermopane” window of area A = 1.5 m2 and an air-gap thickness of 1 cm. Assume
that there is a temperature difference of 20 K between the outside and inside and that
the thermal conductivity of air is air = 0.03 W/m K. With these parameters, we find,

J  A = (0.03 W/m K)(20 K)(1.5 m2)/(.01 m) = 90 watts.

This is approximately the heat loss we assumed when we calculated the power required
to drive an ideal heat pump (Chapter 4).

55
Physics 213 Elements of Thermal Physics

Exercises
1) Flipping coins has the same binomial description as the random-walk and spin-1/2
systems. a) What is the probability of getting exactly 5 heads in 8 tosses? b) What is
the probability of getting exactly 400 heads in 800 tosses? (Hint: Use the Gaussian
approximation.)

2) Show that

N down N!
Ω( m ) =
N up ⎛ N + m ⎞ ⎛ N − m⎞
⎜⎝ ⎟ !⎜ ⎟!
2 ⎠ ⎝ 2 ⎠
m
is consistent with (N, Nup) = N! / Nup! Ndown! .

3) Compare the exact relation P(m) = (m) /2N, using the (m) above, with the
Gaussian approximation,

P (m) = (2/N)1/2 (exp(–m2/2N)),

for a few values of m, with N = 40:


P (m)
m 0 4 8
exact P (m)
approx. P (m)
m

4) Verify that Eq. (5-14) is a solution to the diffusion equation, Eq. (5-15).

56
Statistical Processes I: Two-State Systems Chapter 5

5) A few atoms of Argon (atomic weight = 40) are released from a storage tank into
a room full of air at 300 K. Assume that the Ar atoms have an average energy of
(3/2)kT and a mean free path of 0.1 micrometer. Determine the diffusion constant
of the Ar gas and the rms displacement of the atoms from their point of origin at 1
second and at 1 hour after their release. Plot a few points of rrms(t) to get a feeling
for the t1/2 function.

r rms (cm)
2.5

2.0

1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 t (s)

6) Researchers measuring statistical events often determine the full-width at half-


maximum (FWHM) of a Gaussian distribution. Calculate the relationship between
the FWHM and d.

e -x /2 σd
2 2

FWHM .5

x
0 x1

57
Physics 213 Elements of Thermal Physics

7) When light strikes a pure semiconductor such as silicon, electrons are promoted
from a nearly filled valence band to a nearly empty conduction band. Conduction
electrons produced by the light near the crystal surface diffuse as an ideal gas at
300 K and randomly drop back into the valence band with an average lifetime of
0  1 s. a) Complete the table below to compare the electron gas with N2
gas in air at 300 K, assuming a collision time (with phonons) of  0.01 ns.
b) Calculate the average depth an electron diffuses from the crystal surface during
its lifetime of 1 s. (Hint: Use the Equipartition Theorem for average velocity.)

light crystal

v   D
N2
500 m/s 0.2 ns 0.1 m 1.67  10–5 m2/s
Molecules
electron
0.01 ns
gas

A snapshot of a diffusing gas in a semiconductor at T2 K is reproduced below.


(D.P. Trauernicht and J.P. Wolfe, Physics Department, University of Illinois)

58
CHAPTER

Statistical Processes II:


6 Entropy and the Second Law

A. Meaning of Equilibrium
The fundamental postulate of statistical mechanics is that in equilibrium
all accessible microstates of an isolated system are equally likely. A basic
problem of thermodynamics is to determine how volume, energy, and
particles are distributed in various physical systems.

Consider the example given in Section C of the Introduction. N par-


ticles are contained in a box that is divided into two sections, like the
picture above with NL particles on the left and NR = N – NL particles
on the right. The box is isolated from its surroundings, and the particles
move freely between the two sides. What is the most probable value
of NL? Statistical mechanics says that to solve this problem, we need
to compute the number of microstates for each macrostate defined
by a specific NL.

The probability of observing a given number NL for a system of particles


in equilibrium is proportional to (NL). One of the most fascinating
and important results of statistical mechanics is that for large numbers
of particles, the probability distribution becomes extremely sharp, as
depicted in the following figure:

59
Physics 213 Elements of Thermal Physics

Number of
microstates,
Ω(N L )

NL
0 N
Equilibrium value

Although equilibrium corresponds to many possible values of NL, the distribution is so


sharp that only those NL very near the peak in the distribution are likely to be observed.
Therefore, we can usually say with great precision that there is only one “equilibrium
value” of NL. (Very precise experiments on some systems are able measure the fluctua-
tions around the mean value, i.e., the width of the distribution.) For large N, equilibrium
values of macroscopic parameters, such as gas pressure and total magnetic moment, are
extremely well-defined.

In the last chapter, we considered the binomial distribution of electron spins. Let us now
consider an ideal gas of N particles, confined to a box of volume V, which in general is
not described by a binomial distribution. We eventually need to know how the number
of microstates depends on volume V, particle number N, and energy U. In this chapter we
will examine the dependence of on volume and particle number. In Chapter 7 we
will see how depends on energy.

B. Objects in Multiple Bins


To introduce the counting statistics, we consider a simple system of 4 bins that are oc-
cupied by objects (or particles) labeled A and B:

B A

We allow the two objects to occupy the same bin or different bins; that is, we allow
“unlimited occupancy.”

You should find a systematic way of determining how many different ways there are of
arranging 2 distinguishable particles in 4 bins. Here is some working space:

60
Statistical Processes II: Entropy and the Second Law Chapter 6

You will find that there are 16 possibilities. In fact, you should be able to reason that for
the case of M bins and N particles, the total number of arrangements (i.e., the number
of microstates) is simply

= MN. N distinguishable particles, M unlimited occupancy bins (6-1)

In real life, microscopic particles of a given type are indistinguishable. Every electron
is identical* to every other electron. For a given isotope, every sodium atom is exactly
like every other sodium atom. Every nitrogen molecule … and so on. Let’s see how our
statistical counting would be different if the particles were identical:

A A

How many distinct microstates can you find with these two identical particles? Again
you may assume unlimited occupancy of the bins. Here is some working space:

* For atoms and molecules we use the terms “identical” and “indistinguishable” interchangeably.

61
Physics 213 Elements of Thermal Physics

You will find that the number of microstates is reduced to = 10. The general formula
for this case is:

( N + M − 1)! N identical particles, M unlimited-


Ω=
( M − 1)! N ! occupancy bins (6-2)

If the number of bins is large and much greater than the number of particles (M >> N),
this result simplifies to (see argument in Appendix 5):

MN N identical particles, M unlimited-


Ω≈
N! occupancy bins, N<<M. (low density) (6-3)

In Appendix 5 we consider also the situation where a bin can hold only one object at a
time (single occupancy). A summary of results for follows:

Unlimited Occupancy Single Occupancy


M!
Distinguishable MN
( M − N )!

( N + M − 1)! M!
Identical
( M − 1)! N ! ( M − N )! N !

In this course we will be concerned primarily with the low-density limit (N << M). In
this limit, the formulas for are:

= MN for distinguishable particles


(6-4)
= MN/N! for indistinguishable particles

C. Application to a Gas of Particles


Imagine a gas of N particles in a box of volume V. What is the number of microstates?
We ignore the kinetic energy of the particles for the moment in order to determine
just the volume dependence of .

Each particle in the gas can be anywhere in the box,

Volume V
N par ticles

62
Statistical Processes II: Entropy and the Second Law Chapter 6

so we imagine the box divided up into equal cells of volume V.

δV

The total number of cells in the box is

M = V/V. (6-5)

Assume for this example that there are N distinguishable particles. The number of
microstates is:

= MN = (V/V)N . (6-6)

If the box has a volume of 1 liter (1000 cm3) and the size of a cell is chosen to be 1 cm3,
then V/V = 1000. If there are 10 particles in the box, the total number of microstates is:

= (1000)10 = 1030.

Astronomical! The basic fact is that the number of microstates grows extremely rapidly
(exponentially) with the number of particles, and also very rapidly with V. Just double
V and changes by a factor of 2N = 210 = 1024. (If the particles were indistinguishable,
then we would need to divide by a factor of 10! = 3,628,800. becomes 2.76 1023,
which is still immense.)

Those numbers are for 10 particles in 1000 cells. Imagine the number of microstates
for 1020 particles in 1023 cells (still the low-density limit):

M N = (10 23 )10 = 10 2310


20 20
(6-7)

which is an inconceivably large number. Even if we take the logarithm of this number,
we end up with a huge number:

log10 = 23  1020 (6-8)

Obviously dealing with the logarithm of the number of microstates is a much more
manageable task than dealing with itself. Generally we will use the natural logarithm
(base e = 2.72). It is the “natural” choice because d(lnx) /dx  l/x.

63
Physics 213 Elements of Thermal Physics

An interesting consequence of taking the logarithm is that changing the number of


hypothetical cells in the box doesn’t have a large effect on our counting. For example,
if we take 100 times as many cells, then log10 = 25  1020, which is only an 11%
increase. The properties that we derive from counting microstates do not usually de-
pend significantly on the number of cells chosen. In the next section we will see that
the equilibrium condition depends only on d /dV. The V term (cell size) drops out.

D. Volume Exchange and Entropy


Consider the following basic problem. We have two ideal gas systems that can exchange
volumes by moving a partition. What is the most likely final position of the partition
(i.e., its equilibrium value)?

V1 N1 V2 N2

The numbers of particles N1 and N2 are fixed, but V1 and V2 are allowed to vary such
that the total volume Vtot = V1 + V2 is constant. Because V2 = Vtot – V1 we specify the
macrostate of this system by V1.

In reality, the microstates of an ideal gas depend on V, N, and U. In this chapter, we


consider only how the microstates depend on V and N, leaving the energy dependence
of to Chapter 7. The purpose of this simplification is to build some intuition about
entropy and equilibrium (i.e., entropy maximization) with simpler math.

Denote 1 and 2 as the number of microstates in the left and right chamber, respec-
tively. For each microstate in the left volume, there are 2 possible microstates in the
right volume, so the total number of microstates for the combined system is the product,

= 1 2 . (6-9)

Taking the natural logarithm of , we have,

ln( ) = ln( 1) + ln( 2) . (6-10)

How does the total number of microstates depend on the variable, V1? To keep the
math simple, consider distinguishable particles first:

⎛V ⎞
Ω1 = M1N1
with M1 = ⎜ 1 ⎟ = the number of cells in V1
⎝ δV ⎠ (6-11)

⎛V ⎞
Ω2 = M N2 2 with M 2 = ⎜ 2 ⎟ = the number of cells in V2
⎝ δV ⎠

64
Statistical Processes II: Entropy and the Second Law Chapter 6

Therefore,

V1N V2N
1 2

Ω= = (constant ) V1N (Vtot – V1 )N


1 2
(6-12)
(δV)N
or,
ln( ) = N1 ln(V1) + N2 ln(Vtot – V1) + constant. (6-13)

The maximum in the number of total microstates is given by,

dΩ d ln(Ω )
= 0, or equivalently =0 (6-14)
dV1 dV1
because log( ) increases monotonically with . Because d(lnV)/dV = 1/V, using the
logarithm form gives us the result immediately:

N1 N 2 N1 N 2
− =0 or − (6-15)
V1 V2 V1 V2

This result is just what we should have expected. The most probable state is where the
density of particles on the left is equal to the density of particles on the right. Or, to
answer our original question, the most probable ratio of volumes is

V1 N1
= .
V2 N 2 (6-16)

The same result holds for identical particles, where Ωi = M iN /N i! . The added term in
i

(6-12) doesn’t contain V1; therefore, d(ln )/dV1 is unaffected.

Notice that finding the most likely macrostate is equivalent to maximizing the sum of
logarithms:

ln( ) = ln( 1) + ln( 2), (6-17)

which is a sum of a property of system 1 and a property of system 2. Recalling our earlier
discussion in the Introduction and Chapter 2, these were just the properties we postu-
lated for entropy, S: it is an additive function of the two systems, and the maximum of S
with respect to the free variable (here V1) determines the most likely macrostate. Thus,
our definition of entropy from the statistical mechanics point of view is:

S = k ln( ) , (6-18)

where k is a constant (the Boltzmann constant, 1.381  10-23 J/K) that defines the ab-
solute temperature scale. Temperature comes into play when we consider exchange of
energy between the two systems, as discussed in Chapters 2(C) and 7(B). For our present
volume-exchange problem, we again see the additive property of this state function:

S = S1 + S2 . (6-19)

65
Physics 213 Elements of Thermal Physics

In statistical problems it is often convenient to drop the Boltzmann constant and de-
fine the “dimensionless entropy,” = ln( ). It’s relation to the usual “thermodynamic
entropy” is S = k , and, of course, for the two-system problem, = 1 + 2.

Because the number of accessible states depends on volume (and energy) roughly to
the Nth power, the magnitude of = ln( ) is generally close to the number of particles, as
illustrated in Section F. This is a useful fact to remember in your dealings with entropy.

A numerical example of volume-exchange, shown below, involves six cells and a movable
partition blocking the transfer of three distinguishable particles. Allowing for multiple
occupancy of the bins and counting states, you can easily show that = 1 + 2 = 3.22,
3.47, 3.30, 2.77 and 1.61 for the 5 possible positions of the partition. A table for your
results and plot of the total entropy are given below:

V1 V2



V
A C B

1 2 3 4 5

σ
Most likely value of V1
1 2 = 1 2 = ln if partition were allowed
1 52 25 3.22 4 to move freely
2
3 3
4
2
5
1

0 V1 / δV
1 2 3 4 5
On the next page we make a similar plot with the conventional entropy S  k , showing
how S1 and S2 add up to the total entropy S for a fixed V1. For a freely moving parti-
tion, the probability of observing a particular value of V1 is proportional to the product
1 2. The most probable V1 is where 1 2 and S(V1) are a maximum. For two systems
with many cells and many particles (e.g., two real gases), the probability function is
sharply peaked, and it is reasonable to call V1 at the peak the “equilibrium value” of V1.

66
Statistical Processes II: Entropy and the Second Law Chapter 6

Most probable configuration

S = S1 + S2
Entropy

S2 S 1 = kN 1 ln (V 1 /δV)
S 2 = kN 2 ln ((V tot -V 1 )/δV)

S1

V1
δV V tot -δ V

For larger systems (bigger N’s and V’s), and ln( ) are much more sharply peaked:

"Equilibrium value" of V 1
Ω 1 r Ω 2
( ∝ Probability
of finding V 1 )

V1

In summary, we have defined the conventional entropy S = k from a microscopic


viewpoint: entropy is a constant times the logarithm of the number of accessible states. The
dependence of entropy on V for an ideal gas of N particles has the form:

S = k ln( ) = Nk ln(V/V)
= Nk ln(V) – Nk ln(V) (6-20)

The value of entropy depends on the choice of cell size V; however, the derivative of
entropy with respect to volume (used above to determine the most probable configura-
tion) does not. That is, dS/dV = Nk/V. The derivative is well defined when the partition
is allowed to move in infinitesimal increments; i.e., real life.

67
Physics 213 Elements of Thermal Physics

E. Indistinguishable Particles
In the Introduction, we considered a system of N distinguishable particles labeled A, B,
C, D, etc., that are able to move between two cells:

B A
E D
C G
F

At any one time, the number of particles on the left is NL and the number of particles
on the right is NR. Because the total number of particles is fixed (NL + NR = N), the
macrostate of this system is labeled by:

(N, NL) (6-21)

We have already solved this problem mathematically. Because each particle can exist in
only one of two cells, the resulting distribution is a Binomial Distribution. The number
of microstates for the macrostate (N, NL) is written as:

N!
(N, NL) = (6-22)
N L !( N − N L )!
To generalize this problem to a real box with N indistinguishable particles, the box (with
volume 2V) must be divided up into many cells. As shown in the last section,

VN VN
L R
VN
Ω(N, N L ) ∝ i ∝
N L ! N R! N L ! N R ! (6-23)

with NL + NR = N. It is quite interesting that the functional form 1/NL!NR! for the
many-cell problem with indistinguishable particles is the same as that of the binomial
distribution for the two-cell problem with distinguishable particles, illustrated in Exercise
1 of the Introduction.

F. Maximum Entropy in Equilibrium


Using this example of particle exchange, we can now quantitatively understand the
concepts of equilibrium and irreversibility in terms of entropy maximization, i.e., the
Second Law of Thermodynamics. In the classical picture, particles occupy microscopic
cells of volume V in the box. To simplify the formulas, let’s use microliter units for V
and set V = 1 microliter. Also, for this statistical problem, let’s use the dimensionless
form of entropy = S/k.

68
Statistical Processes II: Entropy and the Second Law Chapter 6

Initially prepare the system by blocking the passage of particles between the two sides
(each with volume V) and then loading all N particles into the left half of the box. From
Eqs. (6-4) and (6-5), the number of accessible states is = VN/N!, giving the entropy
of this initial configuration as:

= ln( ) = N ln V – ln N!

Now open a passageway between the two compartments. Initially, there is a non-
equilibrium distribution. As the particles scatter and diffuse to the right side they
will visit more and more microstates of the unconstrained system. The second law of
thermodynamics tells us that eventually the system will reach a situation where each
microstate of the total system is equally likely to be occupied.
In this case, = (2V)N/N! and the entropy is,

= N ln(2) + N ln V – ln N!

The increase in entropy by opening the passageway is

 = N ln(2) = 0.69  1020 for N = 1020 particles.

The key idea is that when a macroscopic constraint is removed (such as a partition in the
container), the entropy of the system generally increases dramatically. Replacing the partition
here does not return the entropy to its former value. This is the essence of irreversibility
in a thermodynamic system. (See also Exercise 4.)

In calculating the entropy and equilibrium conditions of systems with large numbers,
a very useful approximation comes into play:

ln(N!)  N ln(N) – N,

which is known as Stirling’s approximation. This formula is discussed at the end of


Appendix 5.

Finally, recall that entropy has three natural variables: U, N, and V. In this chapter we
considered how entropy depends on V and N. In Chapter 2 we saw that thermal equilib-
rium (maximum S) involves temperatures, defined by 1/T = (S/U)V,N. Just as thermal
equilibrium between two systems implies equal temperatures, mechanical equilibrium
between two systems means equal pressures. p1 = p2 must correspond to:

(S1/V1) = (S2/V2) ,

69
Physics 213 Elements of Thermal Physics

where the U’s and N’s are held constant. Because S has units J/K and pressure has units
J/m3 the general definition of pressure is:

p = T (S/V)U,N,

which is derived at the beginning of Chapter 10. You can test out this definition for the
ideal gas by using the volume dependence of entropy that we derived in this chapter,
S = k ln(VN) + constants. Do you recover the ideal gas law?

70
Statistical Processes II: Entropy and the Second Law Chapter 6

Exercises
1) Using Stirling’s approximation, compute log10 for 1020 identical particles in 1023
cells, and compare to Eq. (6-8).

2) A box with 4 bins contains four particles separated by a partition shown as the dark
line. The particles on either side of the partition can only occupy bins on that side
of the partition. That is, the particles cannot cross the partition. The bins allow
multiple occupancies. As shown, the particles are distinguishable.

One possible microstate:

C
A D
B
1 2 3

a) With the partition fixed in the central position, what is the total dimensionless
entropy of the system?

b) At which fixed position of the partition is the total entropy a maximum?

c) If the constraint of fixed partition is removed and the partition is allowed to


move among all three positions, what is the total entropy of the system?

3) Consider N = 1000 identical particles distributed in a box of volume V and assume


that the number of cells in the box is much larger than N. By what factor does the
total number of microstates increase if the volume of the box is increased by just
1% (a tiny amount)? Assume that the cell size V does not change.

71
Physics 213 Elements of Thermal Physics

4) Entropy of Mixing: Many practical problems deal with the mixing of two materials.
It is not surprising that if you mix the contents of two bottles of different gases, or
two beakers of different solutions, the total entropy will increase. Let’s see how much.
Start with two separated helium and argon gases, each with N atoms and volume V:

SHe =

SAr =

Calculate the change in entropy if the partition is removed. Be sure to as-


sume that atoms of a given type are identical, and use Stirling’s Approximation,
ln N!  N ln N - N.

Sinitial = Sfinal =

S = entropy of mixing =
(note: many terms cancel)

Now calculate the change in entropy if the two gases were initially the same type
(say He):

S =

What statement involving the Second Law could you say about these two experi-
ments? If all He atoms were not strictly identical (no N! term), would you get the
same result?

72
CHAPTER

7 Energy Exchange

A. Model System for Exchanging Energy


The ideas of the previous two chapters will carry over nicely into the
present problem: How do two systems come into thermal equilibrium
with each other? In exchanging volume and particles, it was useful to
consider discrete quantities of volume V and discrete changes in par-
ticle number N = 1. The approach is no different with the exchange
of energy.

Again to simplify the mathematics, we would like to “quantize” the


problem by choosing indivisible packets of energy that are swapped
between systems. In fact, matter on the microscopic scale is quantized
into discrete energy levels and therefore undergoes discrete changes in
energy. However, for most systems the energy levels are not uniformly
spaced.

Nevertheless, let us imagine that an object in our thermodynamic sys-


tem has a discrete set of equally spaced energy levels. (Those of you who
have played with numerical calculations on a computer should have no
trouble imagining this.) Adding or subtracting energy from this system
amounts to giving or taking an integer number of discrete packets E of
energy, which for brevity we write as . Here is the energy level diagram:

73
Physics 213 Elements of Thermal Physics

En = nε

3ε Convention in this book:


E = single-particle energy
2ε U = energy of a many-particle system

Assume that the energy levels continue this way to infinity. Because this is the energy
diagram for a single particle, we use the letter E for energy. When we discuss a system
of more than one particle, we will revert to the letter U for energy.

The particles in an ideal gas don’t have equally spaced quantum levels like the one
above. In fact, a particle in a (1-dimensional) box has energies given by E = Cn2, with
n = 1,2,3… The outer electrons in an atom are described by energy levels like that of
hydrogen, E = –(13.6 eV)/n2. There is one basic system, however, that does have equally
spaced energy levels on the microscopic scale. It is the simple harmonic oscillator:

E = n ε, with n = 0,1,2,3....
m
ε = hf (h = Planck's constant)
κ Frequency f = (κ /m)1/2/2π

In reality, there is a “zero-point energy” of ½ hf added to this scale that we will ignore.
The harmonic oscillator problem has wide applications in the vibrations of molecules,
solids, and electromagnetic waves, as we shall see.

From here on we will refer to a system with equally spaced levels as an “oscillator.”
Consider two oscillators* in thermal contact with each other, meaning that they can
exchange energy in units of . So, for example, an exchange of one packet of energy
may look like:

* The oscillators are “identical,” but they are “distinguishable” by their labeled position (A, B,…), as in
the earlier problem with spins.
74
Energy Exchange Chapter 7

Before exchange After exchange of ε

4ε 4ε

3ε 3ε

2ε 2ε

1ε 1ε

0 0
Osc A Osc B Osc A Osc B

Note that there are a total of 7 energy quanta in this system, U = 7. In general we
will write U = q, where q = # quanta = # energy packets. Now you determine all the
possible microstates of the two-oscillator system assuming that the total energy is U =
EA + EB = 5 (five energy quanta).





0
Osc A Osc B Osc A Osc B Osc A Osc B Osc A Osc B Osc A Osc B Osc A Osc B

Yes, there are 6 microstates, so = 6. Now work out the microstates for three coupled
oscillators with a total energy U = EA + EB + EC = 3 (three quanta). Here is the diagram
for 3 oscillators:

EA EB EC


EA = n  = En

EB = i 
2ε EC = j 

n, i, j = integers
0
Oscillator A B C

75
Physics 213 Elements of Thermal Physics

I suggest that you group the possibilities according to the energy En = n in oscillator A:

En

3ε = for 3 oscillators with three


energy quanta (U = 3ε )

Continue this type of counting and fill out the following graph for 3 oscillators with
U = 0, , 2, 3, and 4: (Note: (0) = 1)

20

15


10

0 U/ε = q
0 1 2 3 4 = # quanta

76
Energy Exchange Chapter 7

To repeat, we use the symbol q = U/ for the total number of energy packets, or quanta,
in the system. This problem is like the bin problem we saw earlier, although here N is
the # oscillators (or “bins”). In this case we want to know how many ways there are to
arrange q identical packets of energy among N oscillators. The answer is:

(q + N – 1)!
Ω= (7-1)
(N – 1)!q!
If there are many more packets of energy than oscillators (q>>N), known as the “classical
limit,” the number of microstates is approximately given by (see Appendix 5):

qN – 1
≈  q N – 1  U N – 1. (7-2)
(N – 1)!

In particular, as the energy U of the three-oscillator system increases, approaches the quadratic
form,  U2. This result applies to a single three-dimensional oscillator (illustrated at
the right) with energy E; i.e.,  E2. Likewise, for a collection of N three-dimensional
oscillators with total energy U the number of microstates has an energy dependence
N  U3N-1. Clearly the number of microstates increases very rapidly with both U
and N.

Finally, using your work above, plot the probability Pn of observing oscillator A in a
state with energy En for various n:

Pn

3 oscillators and U = 3ε
1

0.5

En
0 1 2 3 xε

The lesson in this last exercise is that if you consider an individual system as a part of
a larger system, the probability of finding the small system in a state with energy En
decreases with increasing En. This is a very basic principle. Can you put in words why
it is true? Look at the number of microstates of the total system as a function of En.

77
Physics 213 Elements of Thermal Physics

B. Thermal Equilibrium and Absolute Temperature


We now consider the energy exchange problem in more general terms. We denote the
energy, the number of microstates, and the entropy of two systems in thermal contact
as follows:

1 (U 1 ) 2 (U 2 )
U1 U2
S 1 = k ln 1 S 2 = k ln 2

U 1 + U 2 = U tot (total energy of the isolated system)

How is energy partitioned between U1 and U2? The equilibrium (most probable) value
of the energy U1 is found by maximizing the total number of microstates of the system.
Equivalently we can maximize the total entropy,

S = S1 + S2 . (7-3)

The math is similar to that for volume exchange,

dS1 dS2
+ = 0
dU1 dU1

dU1  – dU2 (7-4)

dS1 dS2
− = 0
dU1 dU 2

The resulting condition for equilibrium is,

dS1 dS2
= (7-5)
dU1 dU 2

The left side of this equation is a property of system 1 and the right side is a property
of system 2. We know, however, from macroscopic thermodynamics that the basic
condition of thermal equilibrium is that the temperatures of the two systems are equal.
Consequently the slope of the entropy curve, dS/dU, must be directly related to our
concept of temperature. The absolute temperature is defined by the relation,

1 dS
=
T dU (7-6)

78
Energy Exchange Chapter 7

In general, S = S(U, N, V), so T is defined by a partial derivative,

1 ⎛ dS ⎞
=⎜ ⎟ (7-7)
T ⎝ dU ⎠ N,V

meaning that the derivative is taken while holding N and V constant. Now we show
that k in S = k ln( ) is the Boltzmann constant in the Equipartition Theorem.

C. Equipartition Revisited
Let us see if this definition of temperature is consistent with what we know about a
large collection of harmonic oscillators. Recall that the number of microstates for N
oscillators with total energy U = q is a rapidly increasing function of N and U. We
consider the limit where the average oscillator has much more energy than the quantum
of energy , i.e., q >> N. In this case,

= (constant) U N–1 (7-8)

as discussed with Equation (7-2). Therefore, the entropy for this system is,

S = k ln( ) = (N – 1)k ln(U) + const. (7-9)

Taking the derivative of the entropy with respect to energy, we see that,

dS ( N − 1) k
=
dU U (7-10)

For a large number of oscillators, N – 1 may be replaced by N. The entropy and its
derivative are plotted below:

S dS/dU

Nk In U
Nk / U

U U

A graph of kT reveals an interesting result for this limit of U >> N (i.e., q >> N),

79
Physics 213 Elements of Thermal Physics

-1 U/N
kT = k ( )
dS
dU

In this limit, kT is simply the average energy per oscillator, U/N. Therefore, the total
energy of N harmonic oscillators is related to T by,

U = NkT. (7-11)

This is the average energy of N oscillators postulated by the Equipartition Theorem:


recall that the energy of the harmonic oscillator has two quadratic terms, each with
½ kT thermal energy. So for N one-dimensional oscillators,

U = N < ½ mv2 + ½ u2> = N kT. (7-12)

The constant k in the statistical definition of temperature, Equation (7-6), is indeed the
Boltzmann constant introduced with the Equipartition Theorem.

[A technical detail: In most cases the difference between the most probable energy U
and the average energy <U> is negligible. However, for small N our analysis yields
U = (N–1)kT, which is not the Equipartition result. For N = 1, the most probable single-
oscillator state actually occurs at E = 0. (See your graph at the end of Section A.) The
average energy, however, is greater than zero because states with E > 0 have non-zero
probability of being occupied. In an exercise of Chapter 8, you can show that <E> =
kT for the single oscillator. Therefore, <U> = NkT for all values of N, assuming that
kT >> level spacing .]

80
Energy Exchange Chapter 7

D. Why Energy Flows from Hot to Cold


The statistical definition of the temperature of a system can be represented graphically
as follows:

S1
Slope = dS1/dU1 = 1/T1

dS 1

dU 1

U1

The logarithmic dependence of S = k on energy is quite universal for many-particle


systems because we have demanded that the entropy 1) is an additive property of a
system, and 2) it increases monotonically with the number of accessible states, which
rises exponentially with particle number. In precise terms, entropy (like energy) is an
extensive quantity: its value doubles when two identical systems are put together. (Inten-
sive quantities like T and p do not change when two identical systems are combined.)
Notice that the number of accessible states, , is not an extensive property because it
increases (exponentially) with the number of particles.

The slope of the logarithmic curve is 1/T, which decreases monotonically as the energy
of the system increases. This behavior for all types of systems is consistent with our
intuition that temperature increases with increasing internal energy.

As in the case of volume exchange, the condition for thermal equilibrium between two
systems can be represented graphically by the logarithmic plot:

S1 + S2

Entropy
S2

S1

U1
U1* U tot
Most probable U 1

81
Physics 213 Elements of Thermal Physics

The probability of observing system 1 with a particular energy U1 is proportional to


(U1) = e (U1), which for large systems is a very sharply peaked function. Hence, the
most probable value of U1 can be confidently called the equilibrium value.

The equilibrium value of U1 occurs where the slopes of the S1 and S2 curves (with re-
spect to U1 and U2) are equal. If the energy of system 1 were initially higher, say at the
energy denoted by U1*, the temperature of system 1 would be higher (smaller slope)
than for equilibrium, and vice versa for system 2. To reach equilibrium, energy would
flow from system 1 (at higher T) to system 2 (at lower T) until the temperatures are
equal. Because we want the convention that energy flows from “high T” to “low T,” we
have defined T as the inverse of the slope.

The fact that energy flows in the direction that maximizes entropy, however, is not a
convention. It is the Second Law of Thermodynamics, which we now understand in
terms of the most probable condition of a many-particle system in equilibrium.

The shape of the logarithmic function makes all this possible. Because the slope of
increases towards infinity as U approaches the ground-state energy, it is always possible
to match the slopes of two systems so that thermal equilibrium is attained.

E. Entropy of the Ideal Gas—Temperature Dependence


The natural variables for entropy are U, V, and N. However, many problems that we will
deal with involve a system of interest ( ) in good thermal contact with a much larger
system ( R) at temperature T (i.e., a “thermal reservoir”). For these cases it is useful to
describe the entropy of system in terms of T rather than U. From the definition of
absolute temperature, we can determine how the entropy of a gas or solid depends on
temperature. Assuming that the volume is fixed, we rearrange (7-6) as

dU
dS = (7-13)
T
Since no work is done,

dU = CV dT (7-14)

leading to,

C V dT
dS = (7-15)
T

Assuming constant CV and integrating this differential, we find the change in entropy
as the temperature is changed from T1 to T2,

S2 – S1 = CV (lnT2 – lnT1)

S = CV ln (T2/T1). (constant CV) (7-16)

82
Energy Exchange Chapter 7

This relation also works well for solids because their volumes are relatively constant
as a function of temperature. For constant-pressure processes in a gas or solid one can
substitute CP for CV. For an ideal gas with constant CV we can now include the volume
dependence of the entropy, from Chapter 6,

S = CV ln(T) + Nk ln(V) + constants. (7-17)

This equation can only be used to determine differences in entropy during a thermo-
dynamic process: if both V and T are changing quasi-statically for an ideal monatomic
gas (CV = (3/2)Nk),

Sf – Si = (3/2)Nk ln(Tf /Ti) + Nk ln(Vf /Vi). (7-18)

There are no unknown constants here, and the result is independent of units used for
V. The absolute entropy of the ideal gas is calculated in Chapter 11. Notice that this
relation implies S(U,V,N) = Nk ln(U3/2V) + constants*, where U is the average thermal
energy U = (3/2)NkT of the ideal monatomic gas. The instantaneous energy of a gas with
many particles is extremely close to its average energy (in contrast to widely changing
energy of a single particle), as discussed at the beginning of Chapter 6.

* This formula was used in Exercise 3 of Chapter 2. It can be shown that for large N, fluctuations in the
energy of the gas have little effect on its entropy.

83
Physics 213 Elements of Thermal Physics

Exercises
1) Consider a system of 4 oscillators with a total energy of 6; that is, 6 quanta shared
among 4 oscillators. a) What is the average energy per oscillator? ____________
b) Using the exact formula for , determine the probability distribution for a single
oscillator in this system by completing the following table and plotting the result.
(Check your numbers with the sums given.) Here En is the energy of the selected
oscillator, and UR is the energy of the remaining three oscillators.

En UR R Pn PnEn
0

2
3
4
5
6
R = 84 Pn = 1

0.5 Compute the average energy


of an oscillator with the
0.4
formula, <E> = ∑ PnEn and
Pn
0.3 mark the average energy
with a vertical line.
0.2

0.1

0
0 1 2 3 4 5 6 xε

84
Energy Exchange Chapter 7

2) Consider a collection of 100 oscillators, each with an average of 10 quanta. a) By


what factor would change if the total energy were increased by a factor of 2?
b) By what factor would change if one more oscillator were added to the original
system without changing the total energy?

3) An ideal monatomic gas at an initial pressure of p1 = 1 atm and temperature


T1 = 300 K expands adiabatically from 1 liter to 10 liters.

1
P

2
V

Consider the relation: S2 – S1 = (3/2)Nk ln(T2/T1) + Nk ln(V2/V1).


a) What is the change in entropy due to the volume change alone, ignoring any
effects due to changing internal energy?

b) What is the final temperature?

c) What is the change in entropy due solely to the change in temperature, ignor-
ing the entropy change due to volume change?

d) Considering a) and c), what is the total entropy change in this adiabatic
expansion process? Is the result surprising?

85
Physics 213 Elements of Thermal Physics

4) a) What is the entropy change for 2 moles of ideal diatomic gas cooled from
300 K to 200 K at constant V? b) at constant p?

5) The Einstein model of a solid considers each atom in a harmonic potential


well with 3 degrees of freedom; therefore, N = 1022 atoms correspond to 3 
1022 oscillators. Let f = 1  1012 Hz. The average energy of each oscillator is
kT with T = 300 K.
a) Find the average number of quanta q for each oscillator.

b) Determine = ln for the solid. (Stirling’s eqn.: ln (N!)  N lnN – N.)

The following graphing exercises are provided for those who wish to gain more insights into
entropy maximization and spin temperature:

6) a) Consider a single-particle system with number of accessible states 1(U) = C1


U in thermal contact with a 4-particle reservoir with number of accessible states
R(U) = CR (10 – U)4. For simplicity set C1 and CR equal to unity. The total
energy of the two systems is 10. Plot the individual entropies 1 and R and
the total entropy (U) = ln 1(U) R(10 – U) for values of U between 1 and 9.

b) Determine analytically what value of U gives the maximum S(U). Check your
answer against the graph.

7) a) Plot (m) = 100 exp (–m2/25) between m = –10 and 10. This Gaussian func-
tion mimics the number of accessible states of a multiple-spin system (with
m = U/B), as we have seen in Chapter 5 and will study later in Chapter 10.

b) Plot the entropy, (m) = ln (m), of this model system between m = –10 and
10. What is this shape called?

c) Calculate and sketch the temperature of this spin system (Note: d /dU = (1/B)
 d /dm). Can you explain the unusual dependence? Do you think that nega-
tive temperatures are physically attainable in the laboratory?

86
CHAPTER

8 Boltzmann Distribution

A. Concept of a Thermal Reservoir


Consider the following situation: A small system consisting of one oscil-
lator is brought into thermal contact with a larger system (a “thermal
reservoir”) consisting of three oscillators:

En Ei + Ej + Ek = UR = energy of reservoir


Total energy Utot = En + UR

R = # microstates of reservoir

Small Reservoir
system

The basic question asked here is, “What is the probability that the small
system is found in a state with a particular energy En?” Because each
microstate is equally likely, the probability is proportional to the number
of ways En can occur, which is simply the number of microstates of the
reservoir when it has energy (Utot – En). To compute the probability, we

87
Physics 213 Elements of Thermal Physics

must divide R(Utot – En) by the total number of microstates of the system, considering
all possible values of En,

ΩR Ω R ( U tot - E n )
Pn = = ∝ Ω R ( U tot - E n )
Ωtot ΣΩ R ( U tot - E m ) (8-1)
m

Results for Utot = 3 are easily determined from your previous exercises in Chapter
7. Your graph of (U) for 3 oscillators becomes R(UR) here. Notice the value of the
denominator, tot =  R = 20. The results are:

En UR = Utot – En R(UR) Pn = R / R
0 3 10 10/20 = 0.5
 2 6 6/20 = 0.3
2  3 3/20 = 0.15
3 0 1 1/20 = 0.05
20

Pn

0.5

En
0 1 2 3 xε

Why the decrease? In order for the energy of the small system to be increased by en-
ergy , that energy must be taken away from the reservoir. The number of microstates
of the reservoir decreases rapidly as its energy decreases (as in your plot of (U) for 3
oscillators) so the probability of observing the single oscillator in a state with energy
En decreases rapidly with increasing En. If you follow this reasoning, you understand a
basic result of statistical mechanics. It will allow you to predict the average energies of
oscillators, atoms, molecules, spins, etc., in thermal equilibrium.

B. The Boltzmann Factor


Now let us generalize this problem to any single-particle system with energy levels En
in thermal contact with a many-particle system with energy UR. Again, n identifies a
single state of the small system, and R stands for thermal reservoir. A thermal reservoir

88
Boltzmann Distribution Chapter 8

is a system that has sufficiently large heat capacity that its temperature remains virtu-
ally unchanged when contacted by a small system of interest. (d R/dUR  constant.)

UR
En

The small system need not have equally spaced levels like the oscillator, but we assume
that it has discrete energy levels labeled by the index n. This is a reasonable assumption
because all confined particles have quantized levels due to the wave nature of matter.

The entropy of the reservoir is written as,

SR = k ln R (8-2)

The number of microstates of the reservoir R(UR) depends on the energy of the small
system because

UR = Utot – En (8-3)

with Utot = total energy (a constant). As illustrated in the former problem, the probability
that the small system is in a state labeled n (with energy En) is,

ΩR ( U tot –E n )
Pn = ∝ ΩR (U tot –E n )
ΣΩR ( U tot –E m ) (8-4)
m

where the sum over states in the denominator is the normalizing factor (= 20 in the
above example).

R(UR) varies rapidly with energy UR so we can more accurately approximate the en-
tropy, SR = k ln R, which varies more slowly:

S R (U tot - E n )

En
Utot

89
Physics 213 Elements of Thermal Physics

Because the small system is likely to extract a very small fraction of the reservoir’s en-
ergy in equilibrium, En << Utot, we therefore need only consider the region enclosed
by the circle:

S o - E n /T

So

SR

As indicated in the drawing, the entropy of the reservoir near En = 0 can be approxi-
mated by a straight line,

SR = So – (dSR/dUR) En + … (Taylor Series)

SR  So – En/T (8-5)

where T is the temperature of the reservoir, which is a constant for small En.

Probability is directly proportional to , not S. Entropy and the number of microstates


are related by SR = k ln R, or equivalently

ΩR = eS R /k
. (8-6)

Combining the above equations, we have,

ΩR = eS o /k – E n /kT

(8-7)
=e So /k
e –E n /kT
.

The first factor is a property of the reservoir (its number of microstates when it has all
the energy) and is a constant. Thus, the probability that a state labeled n (with energy
En) is occupied is simply,

Pn = C e − βE . n
(  1/kT) (8-8)

90
Boltzmann Distribution Chapter 8

This is the famous Boltzmann distribution. The “constant” C is determined from the
normalization condition Pn = 1. A plot of this general result

∑P n = C∑ e −E n = 1
n n
C
1 1
C= =
∑e − E n
Z
n
Pn
Z = ∑ e −E n (“sum over states”)
n

En

looks a lot like our prior results for small oscillator systems. It basically says that the num-
ber of accessible states of a reservoir decreases exponentially with the energy removed,
as long as the energy removed is small compared to the total energy of the reservoir.

This is an extremely important result that you will use in a variety of problems. Why
is it so important? In addition to telling us the likelihood of finding a small system in a
given state (for example, an excited state of an atom or molecule), the Boltzmann factor,
exp(–En), is used to determine the thermally-averaged values of various properties of
the system. We now consider three practical examples.

C. Paramagnetism
In Chapter 5 we counted the microstates associated with a system of magnetic moments
in a magnetic field:

Spin energy:
E down = +μB
B
ΔE = 2μB
E up = -μB

Note that the spins pointing in the direction of the magnetic field (i.e., “up”) have the
lower energy. This system corresponds to (spin-½) electrons whose moments have a
magnitude given by

 = B = 9.2848  10–24 J/T. (electron) (8-9)

91
Physics 213 Elements of Thermal Physics

The unit of magnetic field strength (or flux density) is Tesla (T). An atom or molecule
with unpaired electron spins has a magnetic moment in units of B. A collection of such
atoms with little or no interaction between spins is known as a paramagnetic system.

In this section we will determine the total magnetic moment of a paramagnetic system
in contact with a thermal reservoir at temperature T. For a system of N spins with Nup
moments pointing in the direction of the magnetic field, and Ndown moments pointing
opposite the field, the total moment in the direction of the field is,

Ndown 0 = (Nup – Ndown)  = m  (8-10)


Nup
and the energy of the spin system is,
m
U = – 0 B = – mB . (8-11)

If the spins are in contact with a thermal reservoir at temperature T, we can compute the
probability that a particular spin will be pointing up or down by using the Boltzmann
distribution:

= CeμB/kT
– E up /kT
Pup = Ce
(8-12)
Pdown = Ce − E down /kT
= Ce −μB/kT .

The constant C is determined from the condition Pup + Pdown = 1, giving,

1
C= .
( eμB/kT + e − μB/kT ) (8-13)

Because Nup = NPup and Ndown = NPdown, the total moment of the spin system at tem-
perature T is

0 = N(Pup – Pdown)
(8-14)
μB/kT − μB/kT
(e −e )
= Nμ .
( eμB/kT + e − μB/kT )

This combination of exponentials is defined as the hyperbolic tangent,

0 = N tanh(B / kT) . (8-15)

92
Boltzmann Distribution Chapter 8

Here is a graph of this function:

0 μB/kT
0 1

We see that the total moment “saturates” at N, corresponding to all spins pointing
along the field.

How strong must the magnetic field be to significantly “polarize” an electron spin
system at 300 K? To find out, we plug in the numbers that make B/kT = 1 (marked
by the cross in the drawing):

B = kT/ = (1.38  10–23J/K)(300 K)/ (9.27  10–24 J/T) (8-16)

= 447 Tesla.

This field strength is about ten times larger than that produced by the largest labora-
tory magnet made today, which indicates that the practical region for paramagnetic
systems at room temperature is the small linear region near the zero in the graph. We
can approximate the hyperbolic tangent in this region by:

tanh x  x for small x, (8-17)

which leads directly to the equation,

N 2B B
0 ∝ . (8-18)
kT T

This inverse-temperature dependence of the total moment (known as Curie’s Law) is


the signature of a paramagnetic system at normal temperatures. By lowering the tem-
perature to a few degrees Kelvin, however, it is possible to almost completely polarize a
system of electron spins in a laboratory field (0  N). In the low temperature regime
we must revert to the exact expression for 0, Equation (8-14).

93
Physics 213 Elements of Thermal Physics

D. Elasticity in Polymers
Ever wonder how a rubber band works? It’s a little different from most solids because
it is comprised of long-chain polymers with many segments. Let us imagine that a seg-
ment can point either parallel or antiparallel to the chain. This simple model of the
rubber band considers the segments of the polymer to be randomly oriented in an “up”
or “down” position, sort of like the drawing below,

Weight: w =mg

The polymer-plus-weight is mathematically equivalent to the spin problem, with mag-


netic field replaced by gravitational field.

Segment
orientation

gravitational
field, g

E = 2aw

In the polymer case, a force, w = mg, stretches the polymer to length L. The change in
gravitational potential energy when one of the segments flips is:

E = 2aw , (8-19)

where a equals the segment length. (This energy is provided by the thermal reservoir.)
We note the correspondence B → aw with the spin system and write for the thermal-
average length of the polymer:

Na 2 w
< L > = Na tanh(aw / kT)  (8-20)
kT

94
Boltzmann Distribution Chapter 8

(Compare to 0 in the last section.) The length is proportional to the weight and inversely
proportional to T in the high-T limit. Obviously there is not a perfect correspondence
with the spins because the polymer has a finite length at zero force and the segment
picture is oversimplified, but this model gives us useful insights into how the rubber
band works. Its elasticity is not due to the stretching of bonds, as in the case of ordinary solids,
but it comes from the thermodynamics of the long-chain polymer.

One amazing consequence is that the total length of the weighted polymer is predicted
to shrink as the temperature is raised. You can do this experiment yourself with a rubber
band, a weight, and an “entropy gun” (such as the one you use to dry your hair). This
contraction effect is opposite the thermal expansion of ordinary solids. The orientations
of the segments become more random as the temperature is raised, causing the rubber
band to shrink. More generally, many types of polymers tend to curl up as temperature
increases.

E. Harmonic Oscillator
An important example of Boltzmann statistics is a simple harmonic oscillator in contact
with a thermal reservoir at temperature T. The distribution of states for the simple
harmonic oscillator is almost the simplest imaginable:

En = nε
0 ε 2ε 3ε 4ε 5ε 6ε . . .

In Appendix 6, we show that the sum over states in this case is

1
Z = Σe − βnε = ( = 1/kT) (8-21)
n 1 − e − βε
and that the average energy of a harmonic oscillator in contact with a thermal reservoir is:

ε
<E>= βε
(8-22)
e −1
Here is a plot of this function:

<E>
<E> ~
~ kT

kT
ε/2

95
Physics 213 Elements of Thermal Physics

At high temperatures ( << 1), the exponent can be represented by the first and second
term of a Taylor series expansion,

e  1 + ,

and the average energy becomes <E>  –1 = kT. The total vibrational energy of a
system of N oscillators is U = N<E> = NkT, as predicted by the Equipartition Theo-
rem. However, at low temperatures the energy decreases more slowly with T. What is
happening at low temperatures?

Essentially when kT <  the thermal excitations are “frozen out.” If kT <<  the prob-
ability that the n = 1 state is occupied is (from Eq. (8-8)),

P1 = e– /(1 + e– + …)  e– ( = 1/kT) (8-23)

which decreases rapidly as T decreases. The evolution from quantum to classical statistics
is shown schematically for several temperatures:

Increasing temperature

E E E

kT < ε 4ε kT ≈ ε 4ε kT > ε 4ε

3ε 3ε 3ε

2ε 2ε 2ε

1ε 1ε 1ε

0 0 0
Pn Pn Pn

The total thermal energy of N diatomic molecules is given by U = N<E>, so the heat
capacity is

Cv = dU/dT = N d<E>/dT. (8-24)

The slope of the <E> graph diminishes at low T; therefore, the contribution to the heat
capacity due to vibrations drops to zero at zero temperatures:

96
Boltzmann Distribution Chapter 8

Cv (vibrations)

Freeze-out of
vibrations
Nk

kT
ε/2

Strongly bonded molecules like N2, O2, and H2 have large vibrational frequencies;
therefore,  = hf >> kT at 300 K and the vibrational contribution to their specific heat
is small. On the other hand, CO2 and H2O molecules have low frequency torsional
vibrations that are thermally active at room temperature (  kT). For this reason the
heat capacity of these polyatomic gases has a significant vibrational contribution at
room temperature. Now we know the physical origin of the temperature dependence
of the specific heat (cv = R) for an ideal gas, as discussed in Chapter 3(C). We shall see
in the next chapter that the low-frequency normal modes of CO2 play a key role in the
global warming of our planet.

Probing the Boltzmann Factor: e-E/kT


The derivation of the Boltzmann factor in Eq. (8-8) may seem a little magical. If so,
you may wish to gain some intuition about how this fundamental relation comes
about by considering the following example: A 6-particle reservoir of harmonic oscil-
lators initially has energy equal to 10, assuming  = 1. The number of available states
of the reservoir depends on how much energy U a single oscillator draws from it; i.e.,
R = (10 – U)5. Show that d R/dU = 1⁄2 for the reservoir at U = 0 and generate two
simple graphs using your calculator:
a) Plot R = (10 – U)5 from U = 0 to 5.
b) Plot R = ln R over the range U = 0 to 9, and show that near U = 0, a good ap-
proximation is R = ln R(0) - U/2, implying that R(U) = R(0)e-U/2.
c) Sketch this function on your original graph … a good approximation?

R ln R

12
1 X 105

0 U 0 U
0 1 2 3 4 5 0 1 2 3 4 5 6 7 8 9

97
Physics 213 Elements of Thermal Physics

Exercises
1) a) Compute the ratio Pup/Pdown for electron spins at T = 2 K and B = 1 Tesla.

b) At what magnetic field would this ratio equal 10?

c) Compute the total moment for 1023 spins at T = 2 K and B = 1 Tesla.

2) With Equation (8-22), show that in the “classical limit” of high temperatures the
energy of N harmonic oscillators is U = NkT.

3) Calculate the average energy for a harmonic oscillator with level spacing
 = 0.5  10–20 J at a temperature of 300 K.

4) A hypothetical molecule has 4 low-lying states spaced by an energy


En  = 0.5  10–20 J.
a) What is the average population of each level for a collection of 100 molecules
_______ 3 at 300 K?

_______ 2

_______ 

_______ 0

b) What is the average energy of one of these molecules? Compare your answer
to that of problem 3.

98
CHAPTER

Distributions of Molecules
9 and Photons

A. Applying the Boltzmann Factor


When we consider a small system in contact with a thermal reservoir at
temperature T, the Boltzmann factor tells us the probability of finding
the small system in a particular state (labeled n),

Pn = Ce – E n /kT
(9-1)

where En is the energy of that state and C is the normalization constant


determined by setting the sum over all probabilities Pn equal to 1. Of-
ten we wish to know the probability, P(E)E, that the small system has
energy between E and E + E, where P(E) is the probability density
(probability per unit energy). To answer this question, we must find the
number of states in the energy range E to E + E, then multiply by the
probability that each state is occupied.

For a 1-dimensional harmonic oscillator (e.g., a diatomic molecule) the


distribution of energy levels is very simple, so the probability density
is also very simple:

99
Physics 213 Elements of Thermal Physics

E

P(E) = C e -E/kT
(1-d oscillator)
E

1d osc.

In this case, the number of states with energy between E and E + E is simply E/,
which is independent of E. The normalization constant C is found by setting  P(E)dE
= 1, which yields C = 1/kT using an integral given in Appendix 4.

In the Einstein model of a solid, interactions between atoms are ignored and each atom
is considered to act as a three-dimensional oscillator, as illustrated below. In Chapter 7
we found that the number of states associated with independent oscillations along x, y,
and z increases as the square of the total energy: (E)  E2. Therefore, the probability
density for a vibrating atom in a solid has the form:

P(E) = C E 2 e -E/kT
(3-d oscillator)

The normalization constant in this case is not the same as for the single harmonic oscil-
lator. Doing the integral gives C = 1/2(kT)3, again referring to Appendix 4.

B. Particle States in a Classical Gas


In order to determine the distribution of energies of particles in an ideal gas, we must
again determine how many states there are between E and E + E. In Chapter 6 we
considered the microstates associated with a gas of particles in a container of volume
V. We divided the volume up into a grid of cells with volume V, and we assumed that

100
Distributions of Molecules and Photons Chapter 9

the classical particle can be localized in a cell at position r. The picture looks something
like this (imagine a 3-dimensional grid):

y
Volume V

δV = dxdydz

r
z x

To completely determine the particle’s state, we must also specify its momentum. Mo-
mentum space is divided up the same way that we sectioned real space. Now here is
the important point: Because E = p2/2m, a sphere in this momentum space represents
a constant-energy surface described by the equation, px2 + py2 + pz2 = p2 = 2mE, where
p = (2mE)1/2 is the radius of the sphere:

py

Constant-energy
surface

px

101
Physics 213 Elements of Thermal Physics

So, how many states there are for particles with energy between E and E + E? The
relevant volume in momentum space is a spherical shell:

py

px

Δp

The number of momentum states with energies between E and E is proportional to


the volume between two constant-energy spheres:

1  vol. of shell  4p2p  4(2mE)p

Using p = (2mE)1/2, we relate p and E:

p = dp E  E–1/2 E
dE
Combining these two equations, we have the desired result,

1  E1/2 E

The “1” stands for “one particle.”

C. Maxwell–Boltzmann Distribution
All the states within the thin shell have nearly the same energy E and nearly the same
Boltzmann factor exp(–E), with  = 1/kT. Therefore, the probability of finding a
particle with energy in the range E + E is

P(E)E  1 exp(–E).

With 1  E1/2 E, the probability density for a particle in an ideal gas is:

P(E) = C E1/2 e–E


(9-2)

102
Distributions of Molecules and Photons Chapter 9

A plot of this function looks like this:

P(E)
E1/2

(β = 1/kT)

e −βE

1/ kT
E = 1/2 mv 2
2

This is the famous Maxwell–Boltzmann Distribution of particle energies in an ideal


gas. By setting dP(E)/dE = 0 you will find that the peak in the distribution occurs at the
energy Epeak = (1/2) kT. The constant of proportionality is determined by integrating
from zero to infinity:

C E1/2 e–E dE = 1

The substitution E = x2 and an integral given in Appendix 4 yields C = 23/2/1/2. Another


integral in Appendix 4 provides us with the average energy of a particle at temperature T:

<E> = C E P(E) dE = C E3/2 e–E dE = (3/2) kT, (9-3)

which is precisely the prediction of the Equipartition Theorem for the ideal monatomic
gas. Finally, the average energy of a monatomic gas of N atoms is:

U = N <E> = (3/2) NkT (9-4)

This derivation of <E> from first principles justifies the Equipartition Theorem (½kT
for each quadratic degree of freedom) for the ideal monatomic gas. It is a result of
Boltzmann statistics, which comes from maximizing the entropy of the gas-plus-reservoir.

D. Photons
Another important application of Boltzmann statistics is the radiation of electromagnetic
energy. How much energy is radiated from a hot object such as a light bulb, your body,
or the sun? We begin this problem by considering electromagnetic waves confined to
a box (or “cavity”) at temperature T. The waves are naturally generated by the jiggling
of electronic charges in the walls of the cavity. An equilibrium is set up between the
moving electrons and the electromagnetic waves.

103
Physics 213 Elements of Thermal Physics

We know from our E&M course that electric fields vanish inside a metal; therefore, the
waves have a null very close to the wall boundaries. This means that there are discrete
wavelengths determined by the size of the cavity:

E
x

y
B

The allowed wavelengths of these standing waves are


m = 2L/m (9-5)

with m = an integer (the “mode index”). The corresponding frequencies are,

f = c/
m (9-6)

where c is the velocity of light. “m” is the number of half wavelengths fitting in the cavity.

Now here’s the new physics: Quantum field theory dictates that only certain energies
are allowed for each wave, just like our harmonic oscillator with an energy level spac-
ing given by  = hf,

En = n  = n hf (9-7)

where h = 6.626  10–34 J s is Planck’s constant and n = 1,2,3… Strange as it may seem,
we can picture the standing wave as having only certain allowed amplitudes correspond-
ing to the quantized energies above.

En = n ε = n hf
n=
4ε 3
Electric field, E
3ε m=3
2

1ε 1

104
Distributions of Molecules and Photons Chapter 9

The quantum of electromagnetic energy is  = hf. We refer to this packet of electro-


magnetic energy as a “photon” and say that there are 1, 2, or 3 photons occupying the EM
mode shown above. Each mode (with wavelength
m) has a different frequency, so n(f) is
designated as the number of photons in a particular mode with frequency f. The energy
in that mode is simply n(f)  hf.

E. Thermal Radiation
From our discussion of the harmonic oscillator in Chapter 6, we know that the average
energy in a harmonic oscillator mode with frequency f is:

hf
<E>= βhf
(9-8)
e −1
which is known as the Planck Distribution. Now we are faced with adding the contribu-
tions from all possible modes. Since there are an infinite number of possible frequencies,
f = c/
m, the total electromagnetic energy in the cavity is formally obtained by sum-
ming over all these modes,

hf
U = Σ<E> =Σ (9-9)
f f e  hf 1
In practice, the counting of modes is similar to the counting of accessible states for the
ideal gas. For a 3-dimensional cavity there are three mode indices, mx, my, and mz. We
make the following sketch (imagine 3-d):

my

shell of radius
1
m = (mx2 + mx2 + mx2) /2

mx

Each mode has an


average energy
hf
<E>=
eβhf - 1

The frequency f of a mode with indices mx, my, mz equals mc/2L, where m = (mx2 + mx2
+ mx2)1/2 is the radius of a constant-frequency sphere. The volume of a shell of radius m
and width m is 4m2 m, and because m = (2L/c)f, we have m = (2L/c) f. Therefore,
the number of modes between f and f + f is proportional to f 2 f.

105
Physics 213 Elements of Thermal Physics

Taking into account both the number of modes between f and f + df and the average
energy hf in a given mode, the electromagnetic energy for all modes with frequencies
between f and f + df is,

f3
u(f )df ∝ df
eβhf − 1

where u(f ) is the electromagnetic energy per unit frequency, sketched below,

u(f)

Planck Radiation Law

hf
2.8 kT

The energy 2.8 kT at the peak is determined by setting du(f)/df = 0. Taking into account
the constant factors and integrating u(f )df , the total electromagnetic energy contained
in the cavity of volume V = L3 at temperature T is,

8π 5 (kT)4
U=V
15 (hc)3 (9-10)

This is the famous Stefan-Boltzmann Law of Radiation. This important prediction


tells how objects at temperature T radiate electromagnetic energy. The total power
per unit area radiated by a perfect radiator (a so-called “blackbody”) at temperature T
is proportional to U and is given by,

JU = SB T4
with (9-11)
SB = 5.670  10–8 W/m2 K4

A numerical example of thermal radiation is provided as an Exercise for this chapter.

Real solids are not perfect electromagnetic radiators. This fact is usually taken into
account with a single parameter, the emissivity . The radiated power per area becomes
JU =  SB T4, where  depends on the material and quality of the surface. In equilibrium
the emission from a solid must equal its absorption from the environment. Perfect
emitters ( = 1) are also perfect absorbers; hence, the name “Black Body Radiation” is
associated with Eq. (9-11).

106
Distributions of Molecules and Photons Chapter 9

Remember that at normal temperatures the peak in the radiation curve is in the infrared
(IR). A strong absorber in the IR may appear brightly colored in the visible region due to
electronic transitions. This discussion leads us directly to a critical environmental issue.

F. Global Warming
One of the most pressing issues of our time is dealing with the increasing temperature
of the earth due to so-called “greenhouse gases.” Politics aside, we now have the tools
to see for ourselves what the basic scientific issues are. We begin by calculating the
temperature of the earth using the Planck Radiation Law and a few simple assumptions.
We only need a few well-known numbers: the radii of the earth RE and sun RS, the
distance R between earth and sun, and the surface temperature of the sun, TS.

Because radiative power per unit area (i.e., flux) from a point source falls off as 1/R2, the
sun’s radiative flux at the earth’s surface is (RS/R)2 times the immense flux at the sun’s
surface. A calculation of the earth’s surface temperature is quite straightforward:

Surface Temperature of the Earth from the Radiation Law

R Rs

Rs = 7 x 108 meters
Earth
R = 1.5 x 1011 meters
TE Ts Sun Ts = 5800K

JR  RE2 = JE 4RE2

JR = 4 JE
JS = Sun’s flux at its surface = SBTS4 2
⎛R ⎞
2
SBTS4  ⎜ s ⎟ = 4 SBTE4
JR = “Sun’s flux at Earth” = SBTS4  ⎛⎜ s ⎞⎟
R ⎝ R⎠
⎝ R⎠
1/ 2
JE = Earth’s flux at its surface = SBTE4 ⎛R ⎞
TE = ⎜ s ⎟ TS = 280 K
⎝ 2R ⎠

From these simple assumptions, we predict the steady-state surface temperature of the
earth to be 280 K, or about 45°F. In fact, the average surface temperature of the earth

107
Physics 213 Elements of Thermal Physics

is about 290 K—a comfortable 60°F for life as we know it. Our calculation has left out
two important factors:
a) A significant fraction (about 30%) of the radiation from the sun is reflected (or
scattered) from the earth’s atmosphere. Because T  (flux)1/4, this reduction in
sun’s flux at the earth’s surface causes about 8% reduction in TE from what we just
calculated. The corrected TE is about 250 K, or 0°F. That would not sustain life as
we know it.
b) Fortunately, counteracting this effect, the earth’s atmosphere also acts as a sort of
blanket, reflecting back some of the radiation that the earth emits. This is known
as the “greenhouse effect,” after the clear-glass buildings that keep plants warm
even in the winter. A greenhouse works because the frequencies f (and thus photon
energies hf) of the radiation from the sun and the earth are quite different:

Planck Radiation Law (not to scale)

u(f) sun

earth
hf = photon
2.8 kT E 2.8 kT s energy

Photons from the sun have frequencies (or wavelengths,


= c/f) mainly in the visible
and ultraviolet (UV) range of the electromagnetic spectrum, whereas photons from the
earth have frequencies mainly in the infrared (IR). Much of the sun’s light is transmit-
ted through the glass roof of a greenhouse, warming the plants and earth inside, but
the glass is not so transparent to the infrared radiation emitted from within. Like the
greenhouse roof, our atmosphere reflects a portion of the infrared radiation from the
earth’s surface.

The “greenhouse gases” that are most effective in reflecting infrared radiation back to
the earth’s surface are H2O and CO2. As we learned in Chapter 3(C) and 8(E), CO2 has
low-frequency vibrations that contribute to its heat capacity even at room temperature.
These torsion, or bending, vibrations (see below) have normal mode frequencies in
the infrared, almost exactly at the peak in the earth’s radiation spectrum. The higher the
concentration of CO2 in the atmosphere, the larger the backflow of thermal energy to the earth.

108
Distributions of Molecules and Photons Chapter 9

En = n

CO2
Photon
0

The importance of the greenhouse effect on a planet’s climate can be realized by com-
paring to our nearest neighbors: Mars has a thin atmosphere with few greenhouse gases,
and its temperatures range from 70°F in the day to -130°F at night! Venus has lots of
CO2, and its atmosphere averages 800°F! These facts and the following are drawn from
a chapter on Global Warming in Professor Segre’s book, cited in the Preface.

As we all know, our planet’s animals take in O2 and generate CO2, and plants turn H2O
+ CO2 into O2. About 80% of the generated CO2 stays in the atmosphere and 20% is
absorbed by the ocean. Seashells are CaCO2. An alarming rise of the CO2 concentra-
tion in our atmosphere, however, is caused by human activities such as fossil fuel usage.
Studies of ice cores show that CO2 is now about 360 parts per million (ppm) of the
atmosphere, compared to 227 ppm in 1750, before the industrial revolution. Scientific
studies suggest that a doubling of the CO2 concentration in the atmosphere—possibly
occurring before the end of this century—would lead to about 10°F increase in the
earth’s surface temperature. The changes in the earth’s ecosystem would be immense.

The North Pole is ice over water, and melted ice does not displace water. Ice melted
in the Antarctic (a land base), however, increases oceanic water. With melting of the
West Atlantic Ice Sheet (WAIS) in the Antarctic, oceans would rise 15 feet, leading to
the disappearance of most ports. For example, LA and much of Florida would be gone.

At the time of this writing, the most comprehensive scientific study on global warming
has been conducted by the Intergovernmental Panel on Climate Change. IPCC, 2007:
Summary for Policymakers deals with the physical science issues and can be accessed on
the Web. The 2006 movie An Inconvenient Truth by Nobel Laureate Al Gore provides
important, graphic details about global warming and should not be missed. In April
2010, the U.S. Environmental Protection Agency (EPA) published an excellent report
on climate change at http://epa.gov/climatechange/index.html.

109
Physics 213 Elements of Thermal Physics

Exercises
1) As discussed in this chapter, the energy distribution for an ideal gas in a box is given
by P(E) = C E1/2 e–E/kT, as illustrated below. It is also possible to confine atoms by
creative use of electromagnetic fields. The harmonic oscillator potential, U(r) =
½ r2, sketched below acts like a spring to confine the atoms. Such “atom traps”
are used by physicists to cool atoms to nanokelvin temperatures and observe their
unusual quantum behavior.

E E

r
Box potential P(E) H.O. potential P(E)

a) The energy distribution for a collection of harmonic oscillators (or atoms in


a H.O. well) at temperature T is given by P(E) = C E2 e-E/kT. Using this func-
tion and an integral given in Appendix 4, compute the average energy <E> =
 E P(E) dE of a particle trapped in the harmonic oscillator potential.

b) Show that this result is consistent with the Equipartition Theorem for a particle
trapped in the harmonic oscillator potential U(r) = ½ r2 = ½ (x2 + y2 + z2). Don’t
forget to count the quadratic terms in the kinetic energy of the particle.

2) a) Calculate the power radiated by a 1-kg sphere of aluminum at 20°C due to


electromagnetic radiation. The density of aluminum is 2.7g/cm3. [Hint: Com-
pute the surface area of the sphere and use the radiation formula.]

b) How rapidly do you think this sphere would cool if it were in outer space?
Calculate how much time would it take for the Al sphere to cool from room
temperature to freezing (T  20 K).

110
CHAPTER

10 Work and Free Energy

A. Heat Flow and Entropy


Given the statistical interpretation of entropy, we will now re-examine
the thermodynamic processes of an ideal gas. Does entropy provide any
insights into heat engines?

First recall the relationship between entropy and heat flow in a reversible
process, S = Q/T, which was derived at the end of Chapter 4(D) for
the ideal gas. This relation between entropy and heat actually extends
beyond the ideal gas, as we will now show. For a fixed particle number,
the differential of S = S(U,V) is quite generally written

⎛ ∂S ⎞ ⎛ ∂S ⎞
dS = ⎜ dU + ⎜ dV .

⎝ ∂U ⎠ V ⎝ ∂V ⎟⎠ U

This is Eq. (2-10) for infinitesimal changes in S, U, and V. The derivative p


(S/U)V equals 1/T by the definition of temperature. The derivative
(S/V)U can be determined by considering the figure at the right. For
constant p, the First Law gives dU = –pdV when dQ = 0. The equilib-
rium condition (maximum entropy) is
U, V
⎛ 1⎞ ⎛ ∂S ⎞ ⎛ p ⎛ ∂S ⎞ ⎞
dS = – ⎜ ⎟ pdV + ⎜
⎝ ∂V ⎟⎠ U
dV = ⎜ – + ⎜ ⎟ dV = 0,
⎝ T⎠ ⎝ T ⎝ ∂V ⎠ U ⎟⎠

111
Physics 213 Elements of Thermal Physics

yielding (S/V)U = p/T. Quite generally, the change in entropy for a change in volume
at constant energy is p/T. We now have a general relationship between the differentials
of the three state functions U, S, and V:

TdS = dU + pdV ,

which is named the Thermodynamic Identity. Combining this result with the First Law
of Thermodynamics, dU = dQ – pdV, we have,

dS = dQ/T for any reversible process.

We first saw this result in Chapter 4D. Remember that S = S (U,V) is a state function, so
dS is a small change in that function. Q is a process energy, so dQ is the small amount
of thermal energy added to the system, as discussed after Eq. (4-5). For larger changes
at constant T, the equation becomes S = Q/T.

B. Ideal Heat Engines


Recall the Carnot Cycle discussed in Chapter 4:

Qh
1 2 adiabat

Th
4 3
Tc
Qc

Vb Va Vc Vd V

The area enclosed by this closed cycle is simply the work done by the engine over one
cycle.

Wby = p dV . (10-1)

There is another useful way of representing this cycle: Plot the gas entropy vs. tem-
perature, remembering that S = Q/T on the isotherms and S = 0 on the adiabats:

112
Work and Free Energy Chapter 10

3
S2

Qc 4 W by 2

S1
1

T
Tc Th

Note that Qc = TcS, Qh = ThS, and Qh = Qc + Wby from energy conservation. This
graph shows clearly that the entropy extracted from the hot reservoir (process 2) equals the
entropy expelled to the cold reservoir (process 4):

Qh Qc
= .
Th Tc (10-2)

When heat is transferred in the Carnot engine, the gas and thermal reservoirs are
ideally always at the same temperature, so the change in the total-system entropy
Stot = –Qh/Th + Qc/Tc + Sgas is zero. (Qc and Qh are defined as positive, and Sgas =
0 for a closed cycle.)

By definition, the efficiency of any heat engine is

Wby Q h − Qc Q
ε= = = 1− c , (10-3)
Qh Qh Qh

which, using Eq. (10-2), leads to the efficiency of the Carnot engine,

=1–
Tc Carnot efficiency (10-4)
.
Th
The Second Law says that the entropy of any isolated system cannot decrease, i.e.,
Stot  0. A reversible process is one in which the entropy of the total system (engine
plus reservoirs) does not change: Stot = 0. On an adiabatic curve (Q = 0), the total en-
tropy of the system does not change because energy and volume contributions to Sgas
cancel out: dS = dQ/T = (dU – pdV)/T = 0. For an isothermal branch with heat flow Q,
the change in entropy of the gas is equal and opposite to the change in entropy of the
reservoir because Tgas = Tres:

Sgas + Sres = Q/Tgas – Q/Tres = 0 (10-5)

113
Physics 213 Elements of Thermal Physics

corresponding to a vertical line on the S-T diagram of the gas. The heat flow must be
small enough that a negligible temperature gradient is generated at the interface. This
is the prime drawback of a Carnot engine: it cannot be cycled quickly and still keep
its efficiency. You’ve probably noticed that “speed” and “quality” are inversely related,
even in life!

All real engines have  < Carnot because, Stot > 0 for any real engine. (The proof is simple
and is left as an Exercise for you.) Three common sources of unwanted entropy genera-
tion are 1) friction, which represents the excitation of vibrational degrees of freedom in
the parts of the engine, 2) the flow of heat between two parts of the system that are not
at the same temperature, and 3) a process not carried out in a quasi-static fashion. As
an example of the latter process, consider a piston that is allowed to move very rapidly,
so that the pressure is not uniform in the gas. An extreme case is “free expansion”:

Imagine that the piston is massless so it takes no work to move it. If the volume changes
instantly by a factor of 2, the gas would eventually fill the volume without doing work (U
= 0). For free expansion the gas temperature is unchanged and the entropy changes by

S = Nk ln(V2/V1) = Nk ln(2) = 0.63 Nk. (10-6)

Due to these types of irreversible processes (see also Chapters 2(B), 6(F)), the efficien-
cies of all real engines are less than that of the Carnot engine.

C. Free Energy and Available Work


Let’s use the energy in a hot brick to drive a Carnot engine. The brick is initially at a
temperature Th, and the environment is at temperature Tc. The Carnot efficiency of the
engine will decrease as the brick cools, but initially the efficiency is given by:

 = 1 – Tc/Th (10-7)

brick Th
Qh
W by
Qc
room Tc

This means that for a small heat input dQh (defined positive) from the brick, the Carnot
engine performs a small amount of work:

dWby = dQh(1 – Tc/Th) (10-8)


114
Work and Free Energy Chapter 10

During this process the energy of the brick changes by dU = – dQh and its entropy
changes by dS = –dQh/Th. Therefore the work done is:

dWby = – dU + TcdS . (10-9)

Notice that the temperature of the environment enters into the work. This equation
applies to each step of the process. When the brick finally reaches Tc, the total work
accomplished is;

Wby = –U + Tc S = –F with F = U – TcS , (10-10)

where F is the “Helmholtz Free Energy” of the brick referenced to the temperature
of the environment at a fixed Tc. In this course, we simply use the term “free energy”
to represent F. U, S and F are properties of the brick. In essence, for this reversible
process, free energy is converted into work:

Wby = Available work = –(change in F) = Fi – Ff . (10-11)

To calculate the total work performed by the engine, one must calculate U and S.
For solid matter at normal temperatures, the heat capacity C is relatively constant,
implying U = CT, but the entropy change depends on temperature, requiring us to
perform an integral:

S = dQ/T = C dT/T = C ln(Tf / Ti) (10-12)

A numerical example is provided in the Exercises at the end of this chapter.

The essential physics of Equation (10-10) is that even for the most efficient engines,
Second
some heat is lost. An energy –U is supplied by the energy source, but an amount Qc
Law in
= TcS is lost to the environment. The free energy F of the energy source (brick or
action.
gasoline) accounts for this loss: F = U – TcS. Got it?

D. Free Energy Minimum in Equilibrium


What if the brick were initially colder than the environment? Can we extract positive
work from this system? Equation (10-8) tells us that the differential work is dWby =
(dQh/Th)(Th – Tc), where Th is now the environmental temperature and Tc is the (vari-
able) temperature of the brick. As in the hot brick case, dQh and (Th – Tc) are positive,
providing positive work.

300K Th
Qh
W by
Qc

brick Tc

115
Physics 213 Elements of Thermal Physics

In general, all it takes is two systems with different temperatures to produce positive
work. In the cold brick case, the energy to do work is extracted from the environment.
Indeed, there are legitimate proposals to produce work (e.g., electrical energy) by run-
ning engines between our environmental temperatures and colder temperatures deep
in the sea; however, the idea is not economically feasible at the present time.

How is the work done by the engine related to the free energy change of the cold brick?
Here we notice that U = Qc and S = Qc/Tc = Qh/Th for the brick; therefore,

Wby = Qh – Qc (closed cycle)

= ThS – U (Carnot engine)

= – F = Fi – Ff (Free energy decrease)

Notice that in both hot-brick and cold-brick cases, free energy of the brick is converted
into work; that is, Wby = Fi – Ff . This means that, even though its temperature is ris-
ing, the free energy of the cold brick decreases as it approaches equilibrium. We can
graphically display this fact with a diagram:

Fi

Ff Ti
200K 300K 400K (initial temperature
of the brick)

Tenv

This analysis shows that the free energy of the brick is a minimum when it is in
equilibrium with its environment at Tenv. In general,

F = U – TenvS.

E. Principle of Minimum Free Energy


F also plays an important role in determining the equilibrium conditions of systems in
which work is not involved, and in which the system is always in thermal equilibrium
with a reservoir at temperature T (= Tenv in the above equation). Consider a small system
with initial energy U = 0 brought into contact with a large thermal reservoir. The small
system gains entropy S(U) by extracting energy U from the reservoir. The temperature
of the reservoir does not change significantly because U << UR; however, the reservoir
does lose entropy when it supplies energy to the small system. The equilibrium value

116
Work and Free Energy Chapter 10

of U occurs when the total entropy of the system Stot is a maximum. When the small
system has extracted energy U, the entropy of the reservoir is SR = SO – (dSR/dUR)U =
SO – U/T, where SO is the entropy of the reservoir when U = 0. You’ve seen this result
before: the heat flow Q from the reservoir equals U, so its entropy change is –Q/T.

The total entropy is the sum of the entropies of the reservoir and the small system:

U
Stot = SR + S ≈ SO − +S
T

U  TS
Stot = SO –
T

F
Stot = SO 
T

The last equation shows that maximizing the total entropy with respect to an internal
variable is equivalent to minimizing the Free Energy of the small system:

Stot –F/T .

The following pictures illustrate this idea:

Total Entropy Free Energy


Stot = S - U/T + const. F = U - TS

internal
variable

internal
variable

equilibrium value equilibrium value

The reason that we shift our attention from the total entropy to the free energy is that
the free energy is determined by the properties of the system of interest, with the single
parameter T reflecting the influence of the thermal reservoir.

F. Equipartition of Energy
Remember the Equipartition Theorem? We postulated that each of the quadratic terms
in the system’s energy has a thermal-average value given by

<energy per quadratic term> = ½ kT,

where k is the Boltzmann constant, 1.381  10–23 J/K, and T is the absolute temperature
in Kelvin. This theorem was stated in Chapter 2 without proof.

117
Physics 213 Elements of Thermal Physics

We can now justify this statement using our knowledge of statistical mechanics. In
equilibrium, a system in contact with a thermal reservoir has a minimum free energy.
Consider a small system in contact with a reservoir at temperature T:

U T

From energy conservation alone, U could be any value between zero and the total
system energy. Basically, we want to calculate the average amount of energy U that the
reservoir is willing to give up to the small system. To answer this, we simply minimize
the free energy

F(U) = U – T S(U) (10-13)

with respect to U. Statistical mechanics has provided us with S(U) = k ln (U) for a va-
riety of systems. In general the number of accessible states (U) increases exponentially
with the number of particles.

In Chapter 7(B) we found that (U) = CUN for N one-dimensional oscillators and
(U) = CU3N for N three-dimensional oscillators, assuming systems with many particles
(N >> 1). In Chapter 7(E) we showed that (U) = CU3N/2 for an ideal monatomic gas,
with the same approximation. For these and other many-particle systems, the state
functions look something like this:

F
U

-TS
minimum (dF/dU = 0)

With (U) = CUN, the free energy of the one-dimensional (1D) oscillator is

F(U) = U – kT (N lnU + lnC), (10-14)

118
Work and Free Energy Chapter 10

so the derivative with respect to the total energy U of the oscillators is

dF/dU = 1 – NkT/U. (10-15)

Setting dF/dU = 0 and solving for U leads to the equilibrium energy,

U = NkT , (10-16)

which is the same result postulated by the Equipartition Theorem, assuming ½ kT


kinetic energy and ½ kT potential energy per particle.

NkT is the average energy that the reservoir supplies to the N-oscillator system in
equilibrium. Fascinating, isn’t it, that our result from counting microstates, (U) =
CUN, came from the assumption of equally spaced levels for the oscillator, a fact that
is rooted in quantum mechanics.

You can easily apply this method to determine the average energy of N three-dimensional
oscillators (the Einstein model of a solid) and the ideal monatomic gas. Here’s a table
for your results (again assuming N >>1):

System (U) S(U) dF/dU Average U


1D
UN Nk lnU 1 – NkT/U NkT
Oscillator
3D
U3N
Oscillator
Monatomic
U3N/2
gas

Compare your results to the Equipartition results, Eqns. (3-5) and (3-12).

G. Paramagnetism—the Free Energy Approach


The spin-½ system is a particularly good example of the interplay between energy and
entropy in the principle of free energy minimum. Consider a collection of N spins with
magnetic moment  that are in contact with a thermal reservoir at temperature T.
In Chapter 5 we found that the number of accessible microstates for large N has the
Gaussian form with a standard deviation d = N1/2 ,

(m) = C exp(– m2/2N), (10-17)

where m = Nup – Ndown is the “spin excess.” The corresponding entropy of the spin
system with spin excess m is therefore,

S(m) = k ln( ) = k (ln(C) – m2/2N) , (10-18)

which is simply an inverted parabola, drawn as follows:


119
Physics 213 Elements of Thermal Physics

S(m)

m
-N 0 +N

Notice that the point of maximum entropy corresponds to maximum disorder in spin
orientation. The energy of the spin system in magnetic field B is plotted as follows:

U(m) = -mμB

+N μ B
0 N
m

-N μ B

In some sense, there is a competition between S and U: the number of accessible spin
states is a maximum at m = 0, but the number of reservoir states is increased when it takes
energy from the spins. As usual, the total entropy of spins-plus-reservoir is maximized
in equilibrium. Equivalently the free energy is minimized with respect to U (or m),

dF/dm = 0, (10-19)

where,

F(m) = U(m) – TS(m). (10-20)

Graphically this effect is represented by combining the last two plots, which we do
here for two different temperatures. Notice that the entropy parabola is plotted upside
down and is multiplied by T:

120
Work and Free Energy Chapter 10

+N μ B m equilibrium
U

0 N
m

-N μ B
F(m) =
U - T1S
U - T2S

As the temperature of the reservoir is raised from T1 to T2, the minimum in free energy
moves closer to m = 0, which corresponds to completely random spin orientation, or
maximum disorder. Increasing T enhances the effect of the spin entropy S(m). The
equilibrium value of m, and thus the total magnetic moment 0 = m, is deter-
mined by minimizing F(m).

By plugging U(m)/and S(m) into Eq. (10-20) and setting dF/dm = 0, you can easily solve
for the equilibrium value of m and obtain Curie’s law for the total magnetic moment,

Nμ 2B B
0= ∝ , (10-21)
kT T
as derived in Chapter 8 by considering the average magnetic moment of a single spin
in the high temperature limit. This example shows the underlying importance of free
energy and the minimization principle.

121
Physics 213 Elements of Thermal Physics

Exercises
1) Let’s say that you have “invented” a heat engine with efficiency  = (1–Qc/Qh) larger
than that of a Carnot engine, c = (1–Tc/Th). Before you rush off to the Patent Of-
fice, consider the total entropy change in a heat engine, Stot = –Qh/Th +Qc/Tc, and
show that your engine would decrease the entropy of the universe (Stot < 0) and
thus violate the Second Law of Thermodynamics. (The Patent Office doesn’t have
Einstein any more, but they do know about the Second Law.)

2) What is the amount of work that can be extracted from a 75 kg container of boiling
water, initially at T = 393 K and allowed to cool to room temperature, 293 K, while
driving a Carnot engine? Assume that the specific heat of water is 4184 J/kg K.

3) a) Using free energy, calculate the work done by a Carnot engine as a hot brick
cools from 450 K to 300 K. The heat capacity of the brick is C = 1 kJ/K.
b) Calculate the work done by a Carnot engine as a cold brick warms from 150
K to 300 K.

4) In this chapter we introduced the link between free energy and available work us-
ing the cycle of a heat engine. Free energy can also be understood by considering
just one step in the process: isothermal expansion. Consider a cylinder containing
a monatomic ideal gas at constant T doing quasi-static work on an outside body:

Thermal
force = p x area
Reservoir

With U = (3/2)NkT and S = Nk ln(nQV/N), show that the work done by the gas in
expanding from Vi to Vf equals its reduction in free energy, - F = Fi – Ff .

F = U – TS =

Vf

Wby = ∫ p dV =
Vi

The free energy F(V,T) can be thought of as the work required to compress the
gas (at constant T) from infinite volume to volume V.
122
CHAPTER

11 Equilibrium between Particles I

A. Free Energy and Chemical Potential


In the present chapter, we consider the equilibrium between systems that
can exchange particles. A wide variety of important problems involve
particle exchange between systems at temperature T, for example,
ionization of atoms, dissociation of molecules, chemical reactions, car-
riers in semiconductors and metals, and, more broadly, the science and
engineering of materials.

We will now introduce the general methods for solving these great
problems. As usual, an isolated system is in equilibrium when its total
entropy Stot is a maximum. If the isolated system consists of a small
system in thermal equilibrium with a reservoir at temperature T, then
equilibrium is determined by a minimum in the free energy F of the
small system. This point is summarized in the following equations:

F ΔF
Stot = SO  or ΔStot = − (11-1)
T T
where SO (a constant) is the entropy that the reservoir would have if it
had all of the energy (highly unlikely).

If two systems in equilibrium with a reservoir at temperature T are


allowed to exchange particles, the equilibrium condition is found by
setting the derivative of F = F1 + F2 with respect to N1 equal to zero:

dF dF1 dF2 (dN2–dN1) (11-2)


= − =0
dN 1 dN 1 dN 2
123
Physics 213 Elements of Thermal Physics

Pictorially, we have:

F = F1 + F2

N1

dF/dN 1 =0
V1 V2
equilibrium value

The derivative of free energy with respect to particle number is so important in de-
termining an equilibrium condition that we define a special name and symbol for it:

dFi
i   chemical potential of subsystem “i” (11-3)
dN i

For two subsystems exchanging particles (one for one), Equation (11-2) shows that the
condition for “chemical equilibrium” is:

 1 = 2 (11-4)

The chemical potential of a system equals the change in free energy when one particle is
added to the system at constant volume. If the two systems are in equilibrium, exchanging
a particle doesn’t change the total free energy; i.e., 1 = 2. In summary, equilibrium
for this simple system is determined by:

Maximum total entropy Stot


→ Minimum free energy F (11-5)
→ Equal chemical potentials 

Note the similarities to what we have already learned:

Energy exchange ↔ T1 = T2 ↔ thermal equilibrium


Volume exchange ↔ p1 = p2 ↔ mechanical equilibrium (11-6)
Particle exchange ↔ 1 = 2 ↔ chemical equilibrium

In all of the systems that we wish to study, at least one of the subsystems is an ideal
gas. Therefore, we must now “bite the bullet” and determine the absolute entropy of
an ideal gas, so that we can determine its free energy and chemical potential. The final
result is μ = kTln(n/nQ) with n = N/V and nQ given in Eq. (11-11). Applications of this result
begin with Section D.

124
Equilibrium between Particles I Chapter 11

B. Absolute Entropy of an Ideal Gas


In Chapter 7 we derived the form of S(N,V,T) for an ideal gas in contact with a thermal
reservoir under the assumption that the particles behave “classically.” In other words,
we assumed that a particle can have any position and momentum, to an arbitrary ac-
curacy. Nature on the atomic scale is not like that. Position and momentum are linked.

In the early part of the 20th century, scientists discovered that microscopic particles such
as electrons and nuclei behave more like waves than billiard balls. One of the statements
of this fact is contained in the Heisenberg Uncertainty Principle. It says that a particle’s
position and momentum cannot be perfectly defined at the same time. It was found that
the uncertainties in momentum and position are limited by the relation,

px x   , (11-7)

where  = h/2 with h = 6.63  10–34 J-s (Planck’s constant). The reason for this un-
certainty in momentum and position is that the particles on the atomic scale behave
like delocalized waves. The uncertainty principle was an indication of the surprising
fact that a particle has a wavelength that is inversely proportional to its momentum:


= h/p , (11-8)

known as the “de Broglie wavelength” of the particle. (Notice the similarity between
this equation and the Uncertainty Principle, if p is replaced by p and
2 by x.)

In a course on quantum mechanics you will see that the square of the wave amplitude
is a measure of the probablity (per unit volume) of finding the particle at a particular
position in space. A particle is viewed as a “wave packet” with poorly defined momentum
and position, as the Uncertainty Principle predicts:

δpx /δx
δx

Armed with this knowledge, we can understand that there is a minimum volume into
which a particle can be confined, depending on its momentum. The minimum cell size
is roughly
3. This is roughly the cell size V used for counting microstates in statisti-
cal mechanics.

Recall that if a box of volume V is divided up into cells of volume V, the total number
of cells is M = V/V. The number of spatial microstates (neglecting momentum) for
one particle is M, and the number of spatial microstates for N identical particles is:

MN
Ω=
N!

125
Physics 213 Elements of Thermal Physics

It is very useful at this stage to employ two “number densities”:

Density of cells: nc = 1/V Density of particles: n = N/V

not to be mistaken with n, the number of moles of gas. With these definitions and the
Stirling approximation, ln N! = N lnN – N, valid for large numbers, you can easily find
the entropy associated with the positions of the particles,

⎛ n ⎞
S = k lnΩ = Nk ⎜ ln c + 1⎟ .
⎝ n ⎠

This analysis has two deficiencies: 1) it doesn’t account for momentum states, and 2) it
uses an arbitrary cell density nc = 1/V. These two deficiencies can be roughly rectified
by assuming that the thermal average energy of a particle is p2/2m = (3/2)kT and taking
a cell density equal to 1/
3 with
= h/p. These mixed classical/quantum considerations
yield a “cell density”:

nc  1/
3 = (p/h)3 = (3mkT/h2)3/2 .

This analysis gives us a good idea of the essential physics, and it turns out to be accurate
to within a factor of about 3. As derived in more advanced texts, the precise entropy of
an ideal monatomic gas turns out to be:

⎛ n 5⎞
S = Nk ⎜ ln Q + ⎟ ,
⎝ n 2⎠ (11-9a)

where

3/ 2
⎛ mkT ⎞
nQ = ⎜
⎝ 2π 2 ⎟⎠
(11-9b)

is known as the “quantum density,” which we identify as the density of quantum cells—the
number of thermally accessible cells per unit volume. This equation for the entropy of
an ideal gas is known as the Sackur-Tetrode (ST) equation. Considering the V and
T dependences (remember, n = N/V), we recover our classical result:

3
S = Nk ln(V) + Nk ln(T) + function(N). (11-10)
2

The term Nk ln(nQ/n) in the ST equation, plus S = k ln , suggest that the number of
accessible states is roughly (nQ/n)N, where nQ/n is the total number of quantum cells
per particle. When n << nQ, the average distance between particles is much greater than

and the particles behave more like billiard balls. In summary,

n / nQ << 1 classical regime (low density)

n / nQ  1 quantum regime (high density).


126
Equilibrium between Particles I Chapter 11

In the classical regime the probability of any particular cell (or “state”) being occupied is
much less than one. In the quantum regime, each state is likely to be occupied by more
than one particle, causing interference between wave-like particles. In Appendix 8 the
absolute entropy for some common gases is calculated and compared to experimental
measurements. The agreement is found to be excellent for monatomic gases. Let’s now
consider a numerical example.

For a monatomic gas of Ar at room temperature and pressure (T = 300 K and p = 1 atm):

m = 40g/(6.022  1023) = 6.64  10–26 kg


 = h/2 = 1.055  10–34 J-s
k = 1.381  10–23 J/K
T = 300 K

yielding,
nQ = 2.47  1032 /meter3.

The ideal gas law gives a density of

n = N/V = p/kT = 105/(1.381  10–23)(300) = 2.45  1025/meter3 .

Therefore, we find the ratio

nQ/n = 1.02  107

for Ar gas at room temperature and pressure. The particle density is about one ten-
millionth that of the quantum density for this temperature. Because n << nQ, we do not
have to worry about the wavelike properties of the Ar atoms. The entropy for one mole
of Ar gas at 300 K and 1 atm is:

S = NAk (ln(1.02  107) + 2.5) = R(16.1 + 2.5) = 18.6 R = 155 J/mol-K.


(R = 8.314 J/mol-K)

To facilitate calculations with other gas atoms, we can combine constants in Eq. (11-
-9b) to get:
3/ 2
⎛ m T ⎞
n Q = 9.88 × 10 meter ⎜
29 –3
⎟ , (11-11a)
⎝ m p 300 K ⎠

where mp is the mass of a proton, which is essentially the mass of a hydrogen atom, mH.
To about 1% accuracy, you may use:

3/ 2
⎛ m T ⎞
nQ1030 meter–3 ⎜⎝ m 300 K ⎟⎠ , (11-11b)
H

which can be remembered because it corresponds to a de Broglie wavelength of about


1Å for H-atoms at T = 300 K. (1Å = 10-10 m) The ratio m/mH is the atomic (or molecular)
weight of the particle: m/mH = 40 for Ar and 28 for N2.
127
Physics 213 Elements of Thermal Physics

Also useful is a formula for the particle density (from the ideal gas law):

p ⎛ p 300 K ⎞
n= = 2.45 × 1025 meter −3 ⎜ ,
kT ⎝ 1atm T ⎟⎠ (11-12)

which you can easily verify using 1atm = 1.013  105 Pa and k = 1.381  10-23J/k

C. Chemical Potential of an Ideal Gas


In Chapters 11 through 13, we will apply the Principle of Minimum Free Energy to a
variety of practical problems involving the equilibrium between two or more subsystems.
In each case, at least one of the subsystems is an ideal monatomic gas, whose entropy is:

⎛ n 5⎞ ⎛ n V 5⎞
S = Nk ⎜ ln( Q ) + ⎟ = Nk ⎜ ln( Q ) + ⎟ ,
⎝ n 2⎠ ⎝ N 2⎠ (11-13)

with U = (3/2)NkT, the free energy, F = U – TS, of the monatomic gas is,

3 ⎛ n V 5⎞
F= NkT − NkT ⎜ ln( Q ) + ⎟ ,
2 ⎝ N 2⎠ (11-14)

which simplifies to the compact form:

⎛ N ⎞
F = NkT ⎜ ln( ) − 1⎟ .
⎝ nQ V ⎠ (11-15)

This equation gives the free energy in terms of its natural variables (N, V, T).

The chemical potential of an ideal monatomic gas is obtained by taking the derivative
of F = kT(N lnN – N ln(nQV) – N) with respect to N,  = (F/N)V,T:

⎛ n ⎞
μ = kT ln ⎜ ⎟
⎝ nQ ⎠ (11-16)

where we have substituted n = N/V as the density of particles. Notice that for the pic-
torial example in Section A, 1 = 2 implies n1 = n2, which is the obvious equilibrium
condition: equal densities (and pressures pi = nikT).

For one mole of Ar gas at p = 1 atm and T = 300 K, we have kT = 0.026 eV, nQ = 1030
 (40)3/2 m–3 and n = 2.45  1025 m–3, yielding a chemical potential,

  (0.026 eV) ln(9.8  10–8)  – 0.42 eV

which is plotted as the point in the following graph of (T):

128
Equilibrium between Particles I Chapter 11

300 K
0 T

p = 1 atm

- 0.4 eV

The curve in the graph above shows how  varies as the temperature is changed while
holding the pressure constant at 1 atm. Specifically, (T)  kT ln(p/nQkT) with p 
1 atm.

Remember that the chemical potential is the change in free energy when one particle is added
to the system at constant volume. If two subsystems with the same  exchange a particle, F
remains unchanged (a minimum), implying that the two subsystems are in equilibrium.

Notice that  is negative for the classical regime, n/nQ << 1. The reason for negative
 is that the TS term dominates F = U – TS and S increases when a particle is added
to the system. In the above example, when one particle is added to the gas, the average
energy increases by (3/2)kT = +0.039 eV, but the free energy changes by   –0.4 eV.
Obviously for dilute gases at ordinary temperatures, the entropy term (–TS) dominates
F. At very low temperatures this classical model of a gas fails, because nQ approaches n.

We will now apply these ideas to some practical problems. Unless stated otherwise, we
assume ideal monatomic gases. The extension to diatomic gases is discussed in Appendix
9. For the diatomic gas, nQ in the chemical potential must be replaced by nQZint, where
Zint represents the number of states associated with internal motions of a molecule. We
will ignore this complication in this course.

D. Law of Atmospheres
If you have ever visited mile-high Denver, you know that the air is a bit thinner than in
central Illinois. It is said that baseballs fly farther in Denver Stadium. We all know that
atmospheric pressure decreases as we go up in altitude, and now you are in a position
to calculate this effect. As a simple representation of this problem, we connect a tube N2
between two equal volumes of monatomic ideal gas, one on the ground and the other at
an altitude h. Gas can flow freely through the tube, but the total gas in the tube is far
less than that contained in the two volumes. The problem is to determine the relative
pressures of the gases in the two boxes. We assume that the temperature is the same at h
both elevations. Chemical equilibrium implies 1 = 2.

N1

129
Physics 213 Elements of Thermal Physics

In this case, the molecules in the upper box each have a potential energy of mgh;
therefore,

U1 = (3/2)N1kT and U2 = (3/2)N2kT + N2(mgh).

The mgh term carries through to the free energy, and to the chemical potential, 2 =
dF2/dN2 (Eq. (11-16)):

⎛n ⎞ ⎛n ⎞
μ1 = kT ln ⎜ 1 ⎟ μ 2 = kT ln ⎜ 2 ⎟ + mgh
⎝ nQ ⎠ ⎝ nQ ⎠ (11-17)

Setting 1 2 immediately yields,

kT ln(n2/n1) = – mgh

(notice that the nQ’s cancel) implying,

n2/n1 = e–mgh/kT

The ideal gas law, p = (N/V)kT = nkT, tells us that p2/p1 = n2/n1, yielding the final result:

p2
= e − h / ho (11-18)
p1

where ho = kT/mg is the characteristic height at which the atmospheric pressure de-
creases by 1/e from that at sea level. This formula is known as the Law of Atmospheres
because it applies to other planets, too. For nitrogen gas (m = 28  1.674  10–27 kg)
at 293 K, you will find,

ho = 8.8 km.

The pressure in Denver (at h = 1.6 km) is less than in central Illinois by a factor of
exp(–1.6/8.8) = 0.83. Yep, a 17% reduction in gas density would mean a few more hom-
ers, unless they moved the fence back.

E. Physical Interpretations of Chemical Potential


There are several ways of viewing chemical potential. In one sense it is a real potential
to do work. For an ideal gas (or atoms in solution) at a given temperature, the chemi-
cal potential is just kT times the logarithm of density:  = kT ln(n), minus a constant.
Consider the following situation:

High  Low 

130
Equilibrium between Particles I Chapter 11

A piston moves isothermally against an external force from volume V1 to V2. The work
done per particle may be simply viewed as the change in chemical potential,

Work per particle = – = –kT ln(n2/n1) = kT ln(V2/V1).

Not surprisingly, for N particles this is the isothermal work we derived in Chapter 4.
It is just another way of looking at the problem.

In mechanics, we know that work is associated with a force applied over a distance. If a
potential energy is involved, force is the negative gradient in potential energy. In one
dimension, F = –d(PE)/dx. For an ideal gas there is no potential energy associated with
a density gradient, yet there appears to be a net force pushing particles towards lower
densities. We can view this situation in terms of an effective force (sometimes called
“diffusive force”) equal to the negative gradient of chemical potential. Imagine particles
released from a wall at high density and diffusing to lower density. These particles could
be moving in a solution or another type of gas. At one instant of time the distribution
of particles might look like this:

high μ x low μ

After a scattering event, each particle is assumed to scatter randomly from other particles,
or elastically from the wall. At a given position x, the chemical potential of this gas has
the form μ = kT ln(n), and the local drift motion appears to be driven by an effective
force per particle given by

Feff = -dμ/dx = -(kT/n) dn/dx ,

where we have used the chain rule for derivatives and d(ln(n))/dn = 1/n. In the example
shown, the effective force per particle is in the positive x direction because the chemi-
cal potential is decreasing from left to right, implying dn/dx is negative. This relation
between diffusive force and gradient in density can be verified by kinetic theory and is
left as an Example problem for those interested.

131
Physics 213 Elements of Thermal Physics

Most importantly, gradients in chemical potential can do chemical work, which can
generate electrical power. In the next chapter we will show the use of chemical potential
in chemical reactions, such as:

2H2 + O2 ↔ 2H2O + energy

high  ↔ low  + energy

which is the basis for the hydrogen fuel cell, which generates a voltage V with the re-
combining of H2 and O2 gases:

H2O (steam)

porous
H2 electrodes O2
solute

By previously separating out the hydrogen and oxygen gases (at a cost in energy) large
differentials in chemical potential are created, which are used to produce electrical
work. The hydrogen fuel cell has been used as a power source in the space shuttle and
is being developed as an automotive power source. Of course, the input work needed
to separate the gases always exceeds the output work from the cell.

132
Equilibrium between Particles I Chapter 11

Exercises
1) a) Show that the chemical potential of helium gas at T = 300 K and p = 1 atm is
 = –0.33 eV.

b) Compute U, S, and F for 1 mole of He gas at 1 atm and 300 K.

U S F 
He

2) a) Using the ST equation and heat capacity C  Q/T = TdS/dT, show that
the molar specific heats, c = C/n , for an ideal monatomic gas are cv = (3/2)R at
constant volume and cp = (5/2)R at constant pressure.

b) Using Equation (A8-4) in Appendix 8, show that cv = (5/2)R and cp = (7/2)R


for the ideal diatomic gas near 300 K, as discussed in Chapter 5.

133
Physics 213 Elements of Thermal Physics

3) What is the pressure on the top of Mt. Everest, at an altitude of 27,000 feet? Assume
that the average temperature in the atmosphere is 270 K.

4) Using kinetic theory, prove that the effective force per particle in a density gradient
is given by:

kT dn
Feff = −
n dx

You will find it useful to define a volume V = A·x and compute the net force on
this layer due to the pressure p on the left and right. Use the ideal gas law, p = nkT.
Here is a picture of the situation:

p(x) p(x + Δx)

Δx

5) Derive the law of atmospheres assuming that for a stationary gas distribution the
effective diffusive force Feff = -(kT/n) dn/dh at height h from the planet’s surface
cancels out the force of gravity, mg.

134
CHAPTER

12 Equilibrium between Particles II

A. Ionization of Atoms
The next challenge for our F-minimum principle is the ionization of
atoms. For the first time we are dealing with particles of different masses,
so each will have its own quantum density. The simplest process we can
imagine is atomic hydrogen ionizing into a proton and an electron:

H ↔ p+ + e– (12-1)

The ionization energy is simply the binding energy of the hy-


drogen atom,  = 13.6 eV. Considering that the average trans-
lational kinetic energy of a gas molecule in this room is only
(3/2)kT = (3/2)(0.025 eV) = 0.0375 eV, this ionization process is not go-
ing to occur often by random atomic collisions. Also because hydrogen
easily binds into H2 molecules at room temperature, we must imagine a
high temperature gas (say at the sun’s surface) in order to get a significant
number of free hydrogen atoms. (We will consider dissociation of H2
in an Exercise at the end of this chapter.)

In this problem there are three components (H, p, e), so the simple
equilibrium condition for two components (1 = 2) must be general-
ized. Consider a volume V with NH, Np, Ne numbers of particles at
temperature T. Equilibrium corresponds to a minimum in total free
energy (like Eq. (2–10) for entropy),

⎛ ∂F ⎞ ⎛ ∂F ⎞ ⎛ ∂F ⎞
F = ⎜ N H + ⎜ ⎟ N p + ⎜ N e = 0 , (12-2)

⎝ ∂N H ⎠ ⎝ ∂N p ⎠ ⎝ ∂N e ⎟⎠
135
Physics 213 Elements of Thermal Physics

where we have suppressed the subscripts (V, T, Nj, Nk) on the partial derivatives. From
the general definition of the chemical potential, μi = (F/Ni), and the relation between
number changes for the reaction: NH: Np: Ne:: 1:-1:-1, we have:

H – p – e = 0.
(12-3a)

This is a simple result to remember:

H = p + e for the reaction H ↔ p + e. (12-3b)

The chemical potential of the hydrogen atom incorporates the large electrostatic interac-
tion between electron and proton producing the binding energy  = 13.6 eV. Ignoring
the electrostatic interaction between well separated particles in this low-density neutral
gas, the three chemical potentials are:

⎛ n ⎞ ⎛ np ⎞ ⎛ n ⎞
μ H = kT ln ⎜ H ⎟ − Δ μ p = kT ln ⎜ ⎟ μ e = kT ln ⎜ e ⎟ (12-4)
⎝ n QH ⎠ ⎝ n Qp ⎠ ⎝ n Qe ⎠

Using the equilibrium condition (12-3), we immediately have the final result for the
gas densities (ni = Ni/V):

nH n QH Δ / kT
= e = K(T) , (12-5)
n p n e n Qp n Qe

where K(T), defined by this equation, is known as the “equilibrium constant” because
it does not depend on the particle densities. Of course, np = ne , so the equation can be
simplified to the following form:

nH
= K(T) .
n 2p (12-6)

Multiplying by nH and inverting, the fraction of ionized atoms is found to be:

np 1
= . (12-7)
nH n H K(T)

This interesting result shows that the fraction of ionized hydrogen depends not only
on the temperature but also on the density of hydrogen atoms. At a given temperature,
ionization becomes more probable at lower hydrogen densities. This is why there are many
ions in outer space, despite the low T.

A simple way of understanding this effect is that at low particle densities it is less likely
that a free electron and a proton will collide and form a hydrogen atom. The effect is
sometimes called “entropy ionization” because two particles (e + p) have more entropy

136
Equilibrium between Particles II Chapter 12

than one (H). You might have thought it was just the kinetic energy per particle that
determined the fraction of ionized atoms. We see that entropy is equally important.
Equilibrium involves a balance between energy and entropy (F = U – TS), as seen in
Chap. 10(G).

B. Chemical Equilibrium in Gases


The equilibrium between chemical species is another important application of the Prin-
ciple of Minimum Free Energy. As before, we are interested in the relative densities (or
concentrations) of the different species in a container of volume V. We consider only
equilibrium between gases, leaving the subject of phase transformations in liquids and
solids to later discussion. A common example of a gas reaction—introduced in your first
chemistry course—is the synthesis of ammonia from nitrogen and hydrogen:

N2 + 3H2 ↔ 2NH3 . (12-8)

The energy of two NH3 molecules is lower by 2  0.9 eV than the energy of one N2
molecule plus three H2 molecules. As Professor Zumdahl of UIUC points out in his
book, Chemical Principles, ammonia synthesis is a key process in plant life and in the
production of fertilizers. Plants require nitrogen to produce amino acids. It is much
easier to remove nitrogen atoms from ammonia than from triply-bonded N2 in the
air. Nature has sophisticated ways of breaking down N2 and making nitrogen available
to plants (e.g., lightning and ammonia-producing bacteria). We won’t need to invoke
lightning or bacteria here because we are concerned with finding equilibrium concentra-
tions, not reaction rates.

Let us simplify the notation by writing the chemical equation as:

aA + bB ↔ cC , (12-9)

where for the ammonia reaction a,b,c equal the coefficients 1,3,2 and A,B,C denote the
molecular types. We label the particle numbers NA, NB, and NC and particle densities
nA, nB, and nC. Conservation of atoms means that a change in one of the numbers de-
termines the changes in both the other numbers: NA: NB: NC:: a : b : –c. Following
the procedure in the last section, we find:

aA + bB = cC for the reaction aA + bB ↔ cC . (12-10)

Assuming that each component acts as a nearly ideal gas,

nA n n
akT ln( ) + bkT ln( B ) = ckT ln( QC ) − cΔ . (12-11)
n QA n QB nC
With a little algebra, we find the basic relation between particle densities:

c
n cC n QC
= ecΔ /kT = K(T) (12-12)
n aA n bB n aQA n QB
b

137
Physics 213 Elements of Thermal Physics

where K(T) is the equilibrium constant, neglecting internal motions. In dealing with
chemical reactions, it is common to use n = N/V in units of

moles per volume (usually mol/Liter) “concentration”

rather than

particles per volume (usually particles/meter3) “particle density.”

Multiplying 1 particle/m3 by (1 m3/1000 L) and (1 mol/6.0221023 particles), we get the


useful conversion factor: 1 particle/m3 = 1.661
10–27 mol/L.

In the mole-per-liter units, Eqs. (11-11) and (11-12) become:

3/ 2
⎛ m T ⎞
n Q = 1641 mol/L ⎜ ⎟
⎝ m p 300 K ⎠ (12-13)

p ⎛ p 300 K ⎞
n= = 4.06 × 10 –2 mol / L ⎜
kT ⎝ 1 atm T ⎟⎠ (12-14)

where mp is the mass of a proton. In chemistry, concentration in mol/L is denoted by


brackets, nN2 = [N2]. We use the notation nN2, regardless of units.

Using our practical formulas for quantum densities of monatomic gases, you can show
that the quantum densities for these gases at T = 300 K are roughly:

gas m/mp nQ (per m3) nQ (mol/L)


H2 2 2.79
1030 4.64
103
N2 28 1.46
1032 2.43
105
NH3 17 6.93
1031 1.15
105

Using these values, we find

2
n QNH3
= 5.45 × 10−7 L2 /mol 2 . (12-15)
n QN2 n3QH2

With 2/kT = 0.9 eV/.026 eV = 38.5, we evaluate the exponential factor:

e2/kT = 1.1  1015. (12-16)

Multiplying these two factors, we find the following equilibrium constant for the
ammonia reaction is,

K(300 K) = 6  108 L2/mol2 (neglecting internal motions)

138
Equilibrium between Particles II Chapter 12

Notice that small uncertainties in  and T can cause large differences in the predicted
K(T), due to the exponential factor. Reducing the temperature from 300 K to 290 K
(a 3% change) increases K(T) by a factor of about four! The internal motions of the
molecules modify both the energy and the entropy of the gas. The effect of internal
motions (rotation and vibration) on K is discussed in Appendix 10.

Let us use the equilibrium constant that we derived to estimate the equilibrium con-
centrations in a reaction. Start with a gas of NH3 at T = 300 K and p = 1 atm. From
Eq. (12-14), one atmosphere of NH3 has the concentration nNH3 = 4.06  10–2 mol/L.
Hydrogen and nitrogen are formed in the ratio nH2 = 3nN2 from the chemical equation,
so the concentration of N2 is estimated as follows,

n 2NH3
= 6 × 108 L2 /mol 2 (12-17)
( n N2 )(3n N2 )3

( 4.06 × 10−2 mol/L)2


( n N2 )4 =
27 × 6 × 108 L2 /mol 2
yielding,

nN2 = 5.65  10–4 mol/L. (12-18)

Therefore, N2 and H2 occur at roughly 1.4% and 4.2% of the concentration of NH3
under equilibrium conditions. The initial pressure of NH3 is only slightly reduced at
300 K. Obviously, in the laboratory one should use the empirically measured equilib-
rium constants. More is said about this problem in Chapter 14 (optional reading). In
that chapter, we consider cases where pressure, not volume, is held constant. That cor-
responds to systems in expandable containers (or under ambient pressure conditions)
rather than fixed volumes.

C. Carrier Densities in a Semiconductor


We now consider a very different physical system with our general mathematical tools:
mobile charge carriers a semiconductor. A crystal such as silicon has bands of closely
spaced energy levels, separated by an energy gap. At T = 0 the crystal is an insulator
because a filled band cannot conduct current. Schematically we draw the energy bands:

Conduction band (at T = 0


unpopulated with electrons)

Energy gap, Δ

electron Valence band (at T = 0


hole totally filled with electrons)

139
Physics 213 Elements of Thermal Physics

In a defect-free semiconductor, for every electron promoted to the conduction band,


there remains a “hole” in the valence band. A hole acts like a particle, too, because a
nearby electron in the valence band can jump into the hole, causing the hole to move.
It is really electrons that carry current, but holes are a way of keeping track of the net
motion of electrons in the nearly filled valence band.

Considering our previous use of the Boltzmann factor to populate excited energy levels,
you might expect that the number of electrons thermally excited from the valence band
to the conduction band is simply Nconduction = Nvalenceexp(–/kT). As found in the cases
of atomic ionization and chemical reactions, the result is not so simple. Free energy is
involved.

Incredible as it may seem, the distributions of states near the conduction-band minimum
and near the valence-band maximum have essentially the same form as that of an ideal
gas. Because the atomic cores are periodically arranged on a crystal lattice, an electron
can actually travel across many lattice sites without scattering (hence, the term “nearly
free electron”). The effect of the lattice is to modify the mass of the electron, i.e., its
acceleration under an applied force. Here we use the symbols me and mh for the “ef-
fective masses” of electrons and holes.

Now we minimize the free energy of this 2-component gas (just like Eq. (12-2) with
i = (F/Ni)):

F = eNe + hNh = 0. (12-19)

Because an electron and hole in the pure semiconductor are created at the same time,
Ne equals Nh and Ne = Nh, leading to the equilibrium condition:

e + h = 0. (12-20)

Ignoring interactions at low densities, we treat the electrons and holes as two ideal
gases. In terms of the densities ne = Ne/V and nh = Nh/V, the chemical potentials are:

⎛ n ⎞ ⎛ n ⎞
μ e = kT ln ⎜ e ⎟ + Δ μ h = kT ln ⎜ h ⎟
⎝ n Qe ⎠ ⎝ n Qh ⎠ (12-21)

with nQe = 2(mekT/2 2)3/2 and nQh = 2(mhkT/2 2)3/2 as the respective quantum
densities. The prefactor 2 signifies that electrons and holes can have their spin either
“up or down.” Generally the effective masses of an electron and a hole are not equal.
Setting e = –h, the equilibrium condition becomes:

nenh = nQenQhe– / kT , (12-22)

140
Equilibrium between Particles II Chapter 12

which simplifies with the definition nQ  (nQenQh)1/2 to:

nenh = nQ2 e– / kT . (12-23)

For a defect-free semiconductor, the number of electrons in the conduction band equals
the number of holes in the valence band; thus,

ne = nh = ni , (12-24)

where ni is known as the “intrinsic carrier density”:

ni = nQe– / 2kT . (12-25)

The carrier masses and band gaps are determined by various experimental methods.
Given the m’s and , one can calculate the density of free carriers at any temperature.
The quantum densities have roughly the same order of magnitude as the quantum
density of an electron in free space. Using Eq. (11-11) with 300 K we have:

3/ 2
⎛ me ⎞
nQe = 2  1030 meter–3 ⎜ ⎟ = 2.51  1025 meter–3 (12-26)
⎝ mp ⎠
where mp/me = 1836 is used, and the factor of 2 is from the two spin states.

The following table gives the measured energy gap and quantum density (nQ = (nQenQh)1/2),
and the corresponding electron (hole) density Eq. (12-25) for three common semicon-
ductors at T = 300 K:

material (eV) nQ (meter–3) ni (meter–3)


Si 1.14 1.72
1025 5.2
1015
Ge 0.67 7.21
1024
GaAs 1.43 2.63
1024 3.0
1012

I have left a blank for you to fill in. It is helpful to recognize that kT = 0.026 eV at
300 K. The intrinsic carrier densities vary rapidly with temperature. At room temperature
the calculated densities of thermally excited carriers are much lower than the density of
atoms in the crystal. For example, silicon has a density of 5
1028 atoms/m3. In effect,
only 1 in 1013 silicon atoms contribute an electron to the conduction band at 300 K.

141
Physics 213 Elements of Thermal Physics

D. Law of Mass Action: Doped Semiconductors


Most applications of semiconductors require them to conduct current. The conduc-
tion and valence bands contain excellent free-particle-like states in which the carriers
behave like an ideal gas, but, as we have just seen, the intrinsic density of free carriers is
extremely low. In order to raise the carrier density up to useful levels we must add just
the right kind of impurity atoms. This process is called doping.

Silicon is a group-IV element; i.e., it has a valence of 4. If we replace one of the silicon
atoms with an atom of valence 5, such as phosphorus (P), then it turns out that the extra
valence electron is only weakly bound to the impurity atom. At room temperature, this
“extra electron” is easily excited into the conduction band, leaving the phosphorus atom
“ionized” with a net charge of +e .

Our previous analysis still holds:

nenh = nQ2 e– / kT , (12-27)

which, with the definition of intrinsic carrier density, ni, simplifies to:

nenh = n2i . (12-28)

This is called the Law of Mass Action for a semiconductor crystal. It says that if the
electron density is increased by doping, the density of holes must decrease, because the
intrinsic density ni is fixed by the temperature and the physical parameters of the semi-
conductor. The effect is striking: A common doping level for Si is about one atom in
104, or about 5  1024 phosphorous atoms/meter3. At 300 K those extra phosphorous
electrons go directly into the conduction band, causing the hole density to decrease
to the value

( )
2
n 2i 5.2 × 1015
nh = = = 5.4 × 106 /meter 3 (12-29)
ne 5 × 1024

In this case the hole density in the doped crystal is roughly 18 orders of magnitude
smaller than the electron density! The situation is reversed if impurity atoms with
valence 3 (e.g., boron atoms) are substituted for Si atoms. Then the holes become the
“majority carriers.” Selective impurity doping allows the engineer to fabricate “n” and
“p” layers essential for diodes and transistors.

142
Equilibrium between Particles II Chapter 12

Exercises
1) At what temperature would a gas of nitrogen at 1 atmosphere pressure reach the
quantum density? (Assume for this problem that nitrogen is a monatomic gas.)
Knowing the properties of nitrogen, do you think that this condition (n = nQ) is
attainable for an ideal gas of N2?

2) Determine the density of ionized hydrogen atoms on the surface of the sun, assum-
ing that non-ionized hydrogen atoms exist at a density of nH = 1  1023/m3, and
the surface temperature of the sun is 6000 K.

3) Write down the relations between chemical potentials (e.g., A = B) for each of
the following gas reactions:

2H2O ↔ 2H2 + O2
N2 + 3H2 ↔ 2NH3
H2 ↔ 2H

4) Let us dope the silicon crystal with Nd phosphorus atoms and assume that the
temperature is high enough that all are ionized. The crystal must remain neutral,
so there is a constraint on the number of electrons and holes:

Ne = Nh + Nd (12-30)

showing the negative electrons on the left and the positive holes and ionized donors
on the right. In general, to determine the number of electrons and holes, we must
solve two equations in two unknowns:

ne nh = ni2 and ne – nh = nd

where nd = Nd/V is the density of donors.

a) Use these two equations to solve for ne in terms of Nd and ni.

(Over)

143
Physics 213 Elements of Thermal Physics

b) Compute ne for Si when nd = 1014 m–3 and 1017 m–3. Do your answers make
sense for these cases where nd % ni and nd & ni, respectively?

5) a) Taking into account molecular rotations, find the formula for equilibrium
constant K(T) for dissociation of hydrogen molecules: H2 ↔ 2H. Because H2
is a diatomic molecule, nQ is replaced by nQZint in the formula for its chemical
potential, as described in Appendix 10. Your answer should be in terms of the
quantum densities for H and H2, the binding energy  for H2, and Zint.

b) Determine how many H atoms are present in one cubic meter of H2 at T =


600 K and p = 1 atm. Use  = 4 eV and Zint  kT/r, where r = 0.03 eV is the
quantum of rotational energy.

6) We have seen that entropy tends to a maximum simply because that corresponds
to the maximum number of microstates. In the sense that “disorder” means a large
number of possibilities, large entropy (many possible microstates) means large
disorder. Gases are more disordered than solids and liquids; therefore gases have
higher entropy.

The entropy of an isolated system either stays the same (in equilibrium) or increases
(approaching equilibrium). So the entropy of the universe (the ultimate isolated
system) is always increasing, a situation sometimes referred to as “the heat death
of the universe.”

Now here is the puzzle: If the entropy of the universe is always increasing, how can
there be increasing order in the world? How did life originate from an inorganic
world? How can a (low entropy) flower grow from a disordered array of dirt, chemi-
cals, heat, and light? How can silicon chips be created out of sand? How would you
answer this line of questioning? (We will consider these issues in the next chapter.)

144
CHAPTER

Adsorption of Atoms and


13 Phase Transitions

A. Adsorption of Atoms on a Solid Surface


From the adsorption of pollutant gases by the platinum metal in your
car’s catalytic converter to the binding of oxygen by hemoglobin in
your blood—from materials physics to biophysics—the interactions of
gas molecules with solids and biomolecules impact our lives. We shall
see how free energy F (or, equivalently chemical potential,  = dF/dN)
controls the balance of these important processes.

Consider first the adsorption of atoms on a solid. Imagine a surface with


M binding sites, each of which can hold one particle with a binding en-
ergy . The surface is inside a container of volume V at temperature T.

Volume V, Temperature T

Ng = # atoms in gas
Ns = # atoms on surface
N = total # atoms = Ns + Ng
M = # binding sites

145
Physics 213 Elements of Thermal Physics

It is useful to work in terms of the gas pressure p = nkT instead of the gas density
n = N/V. We define a “quantum pressure” pQ = nQkT. Since n/nQ = p/pQ, the chemical
potential of atoms in the gas takes the form,

⎛ p ⎞
μ g = kT ln ⎜ ⎟
⎝ pQ ⎠ (13-1)

Recalling Eq. (11-11) for nQ, a useful form of pQ = (m/22)3/2(kT)5/2 is:

3/ 2
⎛ m⎞ ⎛ T ⎞
5/ 2

pQ = 4.04 × 10 atm ⎜4
⎟ ⎜⎝ ⎟
⎝ mp ⎠ 300 K ⎠ (13-2)

where m = atomic mass, and mp = mass of proton. pQ represents the hypothetical pres-
sure of an ideal gas at a density nQ. It’s about one million atmospheres for Ar at 300K.

The Helmholtz free energy of the molecules adsorbed on the surface is,

FS = US – TSS = –NS – T k ln( )

M!
with Ω= (13-3)
(M − N S )!N S!

representing the number of ways to arrange NS identical particles on M single-occupancy


sites (Chapter 6). Taking the logarithm of ,

ln = ln M! – ln NS! – ln (M – NS)!. (13-4)

The derivative of ln requires the simple result d(ln N!)/dN = ln N, which you can
intuit or prove as an Exercise in this chapter. We find,

d(ln Ω ) ⎛ M − NS ⎞
= − ln N S + ln( M − N S ) = ln ⎜
⎝ N S ⎟⎠
(13-5)
dN S
leading to the chemical potential of an atom on the surface:

dFS ⎛ M − NS ⎞
μS = = − Δ − kT ln ⎜ .
dN S ⎝ N S ⎟⎠ (13-6)

In equilibrium g = s , so the result is:

M – N S pQ − Δ /kT
= e . (13-7)
NS p

146
Adsorption of Atoms and Phase Transitions Chapter 13

We wish to know the fraction of occupied sites f = NS/M at a given temperature. It is


helpful to define a characteristic pressure p0 = pQe–/kT. Equation (13-7) now becomes
(1–f )/f = p0/p, which we can solve for f:

p
f= (13-8)
p + p0
By adding or subtracting gas atoms to the container, we can control the gas pressure
p = NgkT/V. The fraction of occupied surface sites ranges from zero to one as p is in-
creased. This result is graphically represented by plotting f vs. p for several values of T.

1
f=
fraction of T1
T2
occupied sites T3

0.5

0 p
po(T2)

The characteristic pressure for temperature T2 is indicated by the cross. Notice that
when p = p0 half of the surface sites are occupied. As the pressure is raised at a fixed
temperature a larger fraction of sites will be occupied. A numerical example is provided
as an Exercise at the end of the chapter.

B. Oxygen in Myoglobin
Important biological processes involve the interaction of biomolecules with atmospheric
gases. The protein hemoglobin in your blood is the carrier of oxygen from your lungs to
the cells in your body. A similar molecule, myoglobin, handles oxygen in your muscles.
Myoglobin has one binding site for oxygen, with a binding energy   0.5 eV.

If we define M = number of myoglobin molecules, Ns = number of oxygen molecules


bound to myoglobin, and f = NS/M as the fraction of O2 molecules that are bound, then
the math is identical to the problem we just solved. Oxygen is a diatomic molecule, so
its chemical potential is modified slightly from the atomic case (Eq. (11-16)) due to
molecular rotations. pQ is replaced by pQZint, where Zint  3Zrot  3(kT/r) with r =
.00036 eV, as discussed in Appendix 9. The chemical potential is,

⎛ p ⎞ (13-9)
μ g = kT ln ⎜ ⎟,
⎝ pQ Z int ⎠

147
Physics 213 Elements of Thermal Physics

leading to p0 = pQ Zint e–/kT in Eq. (13-8). As an exercise you will be asked to estimate
the binding energy of oxygen to a myoglobin molecule.

C. Why Gases Condense


The transformations between gases, liquids, and solids govern the natural world around
us. From your knowledge of entropy, you might well ask why phase transitions even oc-
cur. It is generally true that gases have more accessible states (they occupy more volume)
than liquids or solids, so you might expect that the approach to equilibrium always results
in a gas. In other words, if entropy is always increasing, why do (lower entropy) liquids
and solids even exist? Indeed, how did life originate in an entropy-increasing world?

The simple answer to that question is that entropy need not increase everywhere. There
may be regions of lower entropy (a snowflake) and higher entropy (water vapor) as long
as the total entropy of an isolated system is increasing. A low-temperature weather front
increases its entropy by removing entropy (i.e., heat) from water vapor to condense a
water droplet or snowflake.

This simple answer is not totally satisfying because it only says that condensation can
happen, not that it will happen (as it does on a regular basis). The better explanation is
in terms of free energy, which explicitly contains not only entropy of a small system
but also the attractive forces (binding energies) between atoms and molecules. Minimi-
zation of free energy for a system of particles in thermal contact with a reservoir (say,
the environment) requires that atoms become molecules, and molecules condense into
liquids and solids, at sufficiently low temperatures. The inhomogeneity of nature (yes,
even life) is driven by the action of free energy, which incorporates energy and entropy.

We have already seen the effect of free-energy minimization in the production of


molecules from atoms, the adsorption of atoms onto a solid surface, and the binding
of oxygen in biomolecules. It is a natural step further to understand the basics of phase
transformations. We begin with an elementary example—equilibrium between a gas
and solid—which provides us with a basic concept of thermodynamics: vapor pressure.

D. Vapor Pressure of a Solid


Consider an ideal gas in contact with a simple solid—one that has binding energy but no
internal motions (i.e., no entropy). As in the surface-adsorption case, we put N atoms in
a container of constant volume and at temperature T. Assume that an atom condensed
in the solid loses an energy , which must be absorbed by the gas and eventually the
reservoir. We are simplifying this problem by ignoring the fact that an atom on the
surface is not completely surrounded by other atoms and is somewhat less bound than
an atom in the bulk of the solid. (For liquids, this effect is characterized by a “surface
tension” that leads to the instability of small droplets.)

148
Adsorption of Atoms and Phase Transitions Chapter 13

Volume V, Temperature T

Ng = # atoms in gas
Ns = # atoms in solid
N = total # atoms = Ns + Ng

Under these simplifying assumptions, F = Fg + Fs. Setting dF/dNg = 0, and using


dNg = –dNs we have the solid-gas equilibrium condition,

s = g. (13-10)

Because the solid has no internal motions, its free energy is simply Fs = – Ns. (Perfect
order, zero entropy.) The chemical potentials of the solid and gas are:

s = dFs/dNs = – ,
(13-11)
⎛ p ⎞
μ g = kT ln ⎜ ⎟
⎝ pQ ⎠

where pQ is the quantum pressure given by Eq. (13-2). The equilibrium condition im-
mediately yields the pressure in the container,

p = pQe– / kT = pvapor (13-12)

As emphasized by the subscript, this is the vapor pressure of the solid at temperature T—
the pressure at which the rate of evaporation exactly equals the rate of adsorption. For
equilibrium to exist, we do not have the option of making the pressure of the gas any
value we choose! If the pressure is initially higher than the vapor pressure, pvapor(T), the
solid will grow. If the pressure is initially lower than the vapor pressure, the solid will
shrink, possibly to zero. We can illustrate this process by two diagrams. First we draw
(T) for solid and gas at three different pressures, showing the equilibrium condition
(s = g) as a dot. Next we plot the resulting “coexistence temperatures” on a p-T phase
diagram. The vapor-pressure curve p(T) is also known as a “phase equilibrium curve”
or a “coexistence curve.”

149
Physics 213 Elements of Thermal Physics

0 T

μ g = -kT ln
( ) (
pQ
p
p Q ∝ T 5/2
)

-Δ μ s = -Δ

p1 p2 p3

p = p Q e -Δ/kT
p3 Solid region "Coexistence curve"

Gas region
p2
p1
T

If initially we place a cube of ice in an empty freezer, the water molecules evaporate
(“sublime”) from its surface, eventually producing a pressure of pvapor(T). A tiny chip
of ice, however, may not have enough molecules to support this pressure alone, so it
evaporates away completely, corresponding to the region to the right of the coexistence
curve. Realize also that the calculated vapor pressure is only a “partial pressure,” con-
tributing to the total atmospheric pressure in the container.

What about the solid materials around us? For example, a silicon crystal in your com-
puter. What vapor pressure does it try to maintain? Will it evaporate away? Fortunately,
the binding energy of an atom in a solid is a few electron volts, and at 300 K the thermal
energy per atom is only about kT = 0.026 eV, so it would take a huge thermal fluctuation
to kick an atom out of the solid. If this unlikely event happened, the ejected atom would
probably just diffuse off into the atmosphere with little chance of returning to the solid.

150
Adsorption of Atoms and Phase Transitions Chapter 13

Do we have to worry about the silicon in our computer chips evaporating away? No,
because the rate of evaporation of ordinary solids is extremely slow at room temperature.
Also, the vapor pressure for an ordinary solid at 300 K with   3 eV is extremely small:

pQe–/kT = (4.04  104 atm)(28)3/2 e–3/.026 = 4.4  10–44 atm. (13-13)

This vapor pressure is much less than the gas pressure in outer space. In this simple
model, even if you put the solid in a vacuum chamber at T = 1000 K, only a few evapo-
rated molecules would be required to sustain equilibrium. The solids around us are not
likely to evaporate away.

The moral of this story is that we are surrounded by solid materials that are significantly
out of contact with their vapor phases. They were produced either by nature or by man
under much more extreme conditions of temperature and density than they are now
experiencing. The silicon molecules in your computer or watch were grown into a near
perfect crystal from a hot silicon liquid, which in turn came from chemical processing
of SiO2—sand! As the hot Si liquid is cooled, free-energy minimization drives Si atoms
from high-entropy liquid to low-energy solid. On a grander scale, this is how nuclei of
atoms formed from elementary particles in the early universe.

E. Solid/Liquid/Gas Phase Transitions


We can extend these ideas to more general cases. A real solid has internal modes of
oscillation (vibrations) which give it a non-zero entropy considered in Appendix 10. An
atom in a liquid has less binding energy than one in a solid, but the liquid generally has
more entropy (it is highly disordered, plus vibrational modes). We can schematically
represent these ideas with the following drawings:

μ μ

0 T 0 T

μg

Liquid-solid
Solid-gas equilibrium
equilibrium -ΔL
-Δs -Δs
μs μs

μL

Now let’s put all three branches together. Here are 4 graphs at different pressures,
which will help us determine the phase diagram. Notice how the chemical potential of
the gas passes through the other two branches. At one (p,T) all three phases coexist. In
the graphs, Tc represents a coexistence temperature at the given pressure.

151
Physics 213 Elements of Thermal Physics

T T T

Tc Tc Tc Tc

μ p1 p2 p3

T p
Phase Diagram:
Tc p4
p4 Liquid
Gas μg p3 Solid
μ
Liquid μL p2
Gas
Solid μs
p1
T

In the (T) diagrams, we have plotted the three branches of chemical potential at suc-
cessively higher pressures: p1 through p4. The pressure p2 and corresponding coexistence
temperature Tc mark the “triple point” of the system. The actual phase diagram for a
given system depends on the shapes of (T) for the liquid and gas phases, which in turn
depend on the particular “structure” of the liquid or solid.

Recall that  is the change in free energy of a particular phase when a particle is added to
it. Notice that there are crossings of the branches that do not correspond to an equilib-
rium state, because the free energy of the system is lowered by choosing a lower branch
at that temperature. Let’s look more closely at the case with a constant pressure p3.

Phase Diagram:
T p

Tc1 Tc2 Solid

Liquid

p3
μ(p3,T) Gas

T
Tc1 Tc2

152
Adsorption of Atoms and Phase Transitions Chapter 13

The heavy line on the left indicates the equilibrium (p3,T) as one increases the tem-
perature at a constant pressure. The system proceeds from pure solid, to solid/liquid
coexistence at Tc1, to pure liquid, to liquid/gas coexistence at Tc2, and finally to pure gas.
The trajectory for this process on the p-T phase diagram is shown at the right. This
process is achieved by adding heat to the system while keeping the pressure constant.
Let’s examine in detail the liquid/gas part of the transition:

Heater

The system travels from points 0 through 3 on the following diagrams:

Solid T
Liquid gas
liquid 3
0 3
liquid + gas

Tc2
Gas
1 2

0 heat
T input
Tc2 L Q

Notice that a lot happens at the coexistence point, TC2. With the addition of heat, the
system is being converted from pure liquid (at point 1) to pure gas (at point 2). The
amount of heat required to make this complete conversion is known as the “heat of
transformation” or “latent heat” of the liquid, denoted L.

Because the process is conducted at constant pressure, the latent heat is not simply the
energy needed to free the atoms from their bonds to the liquid, i.e., the difference in
internal energies of the liquid and gas phases. While the liquid/gas conversion proceeds,
work is being done on the environment (shown above as the gravitational force of a
weight, but usually atmospheric pressure). The First Law of Thermodynamics tells us

Q = heat input = U + pV , (13-14)

153
Physics 213 Elements of Thermal Physics

showing that the heat source must provide energy for doing work, as well as for breaking
bonds. Because so many processes such as phase transitions and chemical reactions are
conducted at constant pressure, it is common to tabulate the state function;

H = U + pV , (13-15)

known as the “enthalpy” of a substance, usually given per mole of substance. Enthalpy
is understandably known as the “heat content” of a material. If Hl and Hg are the en-
thalpies of the particular liquid and gas at a standard pressure, then Hlg = Hg – Hl =
latent heat of the transition.

Hlg is the energy required to evaporate all the liquid at its boiling point. Here is a table
of the latent heats and entropy changes (S = Hlg /Tc) per mole for some common
substances at atmospheric pressure:

Latent Heats of Evaporation


Gas Boiling Pt. (K) H (J/mol) S (J/molK)
Helium 4.2 92 22
Nitrogen 77.3 5610 72
Oxygen 90.1 6820 76
H 2O 373 44000 119

F. Model of Liquid–Gas Condensation


So far we have considered relatively low densities of non-interacting particles, which
obey the ideal gas law:

pV = NkT . (13-16)

Liquids and solids form because there are attractive interactions between atoms or
molecules. When the temperature of the gas is reduced, the attractive interactions can
cause the gas to condense into a liquid, where the molecules pack together densely in
order to lower their potential energy. The effects of interactions between particles is
incorporated neatly in the van der Waals equation of state (see the Exercise at the end
of this chapter):

⎛ N 2a ⎞
⎜⎝ p + V 2 ⎟⎠ ( V − Nb ) = NkT . (13-17)

Here, V is the volume of the container, and Nb is total space displaced by the N finite-
sized particles, each having volume b. The term N2a/V2 causes reduction in pressure p
(for a given N and V) due to the attractive potential between particles.

154
Adsorption of Atoms and Phase Transitions Chapter 13

It is particularly informative to observe the system at constant temperature. At low


temperatures and high densities the corrections to the ideal gas law greatly distort the
isothermal curves of the ideal gas:

pV NkT

T3

p1 T2
2 1 T1
Variable Force

Consider the lowest curve, where the system is in contact with a thermal reservoir at
temperature T1, and V is controlled by adjusting the force. The system pressure p =
force/A (which includes patm) is recorded as V is varied. In the region between points
1 and 2, the gas condenses into a liquid of constant density, and there is a coexistence
between liquid and gas. As the volume is decreased from point 1 to point 2, the fraction
of liquid increases until, at point 2, only liquid remains in the container. The region
below the dashed curve in the figure corresponds to the mixed phase, or coexistence,
region of the phase diagram.

This type of experiment allows us to determine the complete “equation of state,” p(V,T),
of a particular substance. However, we have no way of directly measuring either the
internal energy change U or the entropy change S in isothermal experiments, be-
cause we cannot measure the flow of heat between the system and the thermal reservoir.
However, our earlier experiment—adding heat at constant pressure—provided us with
the latent heat of the transition as well as the phase boundaries in the p-T plane. Little
by little, bit by bit, nature reveals itself.

155
Physics 213 Elements of Thermal Physics

Exercises
1) Prove that d(lnN!)/dN = ln N, which was used in the problem where atoms were
bound to sites on a solid surface.

2) a) Determine the fraction of surface sites occupied with Ar atoms if the site energy
is  = 0.3, 0.4, and 0.5 eV, assuming a gas pressure of 1 atm and T = 300 K.

b) Show that the crossover from near-zero to near-full occupancy occurs when
 = –. That is, f = ½ when the chemical potential equals the energy of a bound
atom. For a particular surface,  is fixed and  is controlled by the pressure.

3) For a body temperature of 310 K and a partial pressure of oxygen in the muscle of
0.5 atm, let us imagine that the fraction of myoglobin sites occupied by oxygen is
50%. Determine the binding energy  for myoglobin in this hypothetical example.
Be sure to replace pQ with pQZint in the formula for the chemical potential of O2,
where Zint  230 for O2, as given in Appendix 9.

4) A 1 kg block of ice at 273 K melts completely in 6 hours. a) Using the latent heat
of fusion for water as 3.335  105 J/kg, determine the average power provided by
the surroundings at 293 K. b) Calculate the total change in entropy for the melt-
ing process, neglecting any heat rise in the water.

5) Recall the free energy of an ideal gas given by Eq. (11-15). By minimizing free
m energy, derive the gas law of an interacting gas of particles under the assumptions:
1) each particle takes up a volume b, so the volume available to the gas becomes
V – Nb, and 2) there is an attractive interaction between particles. Assume that
h the average potential energy felt by a single particle is proportional to the aver-
age density of other particles, PEint = –aN/V; therefore, the total potential energy
for N particles is –aN2/V. The total potential energy that must be included in F is
PE grav = mgh = pV PEgrav + PEint. The result is the van der Waals equation which is responsible for
liquid-gas condensation!
156
CHAPTER

14 Processes at Constant Pressure*

A. Gibbs Free Energy


In previous chapters we have minimized Helmholtz free energy at
constant volume to determine equilibrium conditions. Frequently, phase
transitions and chemical reactions are observed under conditions of
constant pressure (e.g., atmospheric pressure). In the case of constant
pressure, there is a similar useful function known as the Gibbs free energy.

In the case of fixed volume, a small system is in thermal contact with


a reservoir whose entropy depends on the energy it provides to the
small system: SR = S0 – U/T, where 1/T = (dSR/dUR)V and dU = –dUR.
Consider now that the small system is in contact with a thermal and
pressure reservoir:

Reservoir,
entropy = S R
volume = V R
p, T
p

small system,
entropy = S,
volume = V

157
*Optional reading in Physics 213.
Physics 213 Elements of Thermal Physics

The entropy of the reservoir depends on both the energy U and volume V of the small
system. For simplicity, let’s assume that the reservoir is an ideal gas with pVR = NRkT.
Because VR >> V, we assume that p is a constant. Entropy depends on volume as SR
 NRk lnVR; therefore, (dSR/dVR)T = NRk/VR = p/T. The entropy of the reservoir is:

SR = S0 – (1/T)U – (p/T)V , (14-1)

where S0 is the reservoir’s entropy when U = 0 and V = 0. The total entropy is:

Stot = SR + S = S0 – (U + pV – TS)/T ,

or,
Stot = S0 – G/T , (14-2)

where we have defined,

G = U + pV – TS (14-3)

as the Gibbs free energy. In terms of enthalpy, G = H – TS. In essence, the Gibbs
energy takes into account the amount of work required when the small system changes
its volume at constant pressure: G = U + pV – TS.

Equation (14-2) shows that for systems at constant pressure and temperature, a mini-
mum in the Gibbs energy defines a maximum in Stot, and therefore determines the
equilibrium conditions. In short,

Stot = –G/T . (14-4)

Let us examine the ideal gases that we considered in Chapter 11, but now under the
condition of constant pressure. What is G for an ideal gas? From the above definition
of Gibbs free energy, we see that G = F + pV, and F has the form (Eq.(11-15)),

⎛ ⎛ n ⎞ ⎞
F = NkT ⎜ ln ⎜ ⎟ − 1⎟ . (14-5)
⎝ ⎝ nQ ⎠ ⎠
Using pV = NkT and changing densities to pressures, n/nQ = p/pQ, we have

⎛ ⎛ p ⎞ ⎞
G = F + pV = NkT ⎜ ln ⎜ ⎟ − 1⎟ + NkT .
⎝ ⎝ pQ ⎠ ⎠ (14-6)

Amazingly, the NkT terms cancel; therefore, the Gibbs free energy of an ideal mona-
tomic gas takes on a wonderfully simple form:

⎛ p ⎞
G = NkT ln ⎜ ⎟ = Nμ pQ = nQkT (14-7)
⎝ pQ ⎠

158
Processes at Constant Pressure Chapter 14

where  is the chemical potential defined in Chapter 11, and pQ is the quantum pres-
sure of the gas. Notice also that ln(p/pQ) does not depend on the number of molecules
N, so that G(N,p,T) = N (p,T). (These simplifications occur also for a diatomic gas,
for which U = (5/2)NkT, because the entropy in that case has a 7/2 replacing 5/2. See
Appendix 9.)

Chemical potential is the Gibbs free energy per particle:  = G/N. This fact is consistent
with the definition of chemical potential as the increase in Helmholtz free energy
when one molecule is added at constant volume:  = dF/dN. In that case, the addi-
tion of a molecule increases the pressure, which is a consequence of dealing with the
Helmholtz free energy. Chemists tabulate G because most reactions are performed at
athmospheric (or constant) pressure. The relation between G and  is a little simpler
than between F and .

Quite generally a 3-component reaction is given by,

aA + bB ↔ cC . (14-8)

The total Gibbs free energy for this reaction becomes,

G = NAA + NBB + NCC . (14-9)

Considering the relations between the dNi, as we did for Eqs. (12-2) and (12-10), the
equilibrium condition, dG = 0, implies,

aA  bB – cC = 0. (14-10)

For reactions between ideal gases, the chemical potentials are

⎛ p ⎞
μ i = kT ln ⎜ i ⎟ − Δ i , (14-11)
⎝ pQi ⎠
where i is the energy released (binding energy) when a molecule is formed from the
reactants. Commonly i is set to zero for elemental molecules such as N2 and H2. For
polyatomic gases, pQ must be replaced by pQZint, where Zint accounts for the internal
motions of the molecules. (See Appendix 9.)

The equilibrium relations between the chemical potentials are the same as derived in
the constant-volume case (Chapter 12), but now the ’s are written in terms of “partial
pressures” pi = nikT, rather than densities ni. In an Exercise at the end of this chapter,
you are asked to determine the form of the equilibrium constant when using partial
pressures, rather than densities or concentrations (see Eqs. (12-13) – (12-15)).

159
Physics 213 Elements of Thermal Physics

B. Vapor Pressures of Liquids—General Aspects


Real solids and liquids are usually not so “easy” to model as the previous examples.
However, a little math manipulation using the equilibrium condition (l = g) between
phases yields a very general property for the vapor pressure pvapor(T) of any liquid or
solid, which we abbreviate as p(T). If we expand the differential for one of the phases
by using the product rule for differentials,

G = U + (pV) –  (TS)
= U + Vp + pV – ST – TS (14-12)

and substitute the thermodynamic identity (FLT plus Q = TS),

U = TS – pV (14-13)

we are left with,

G = Vp – ST . (14-14)

At this point it is useful to define G, V, and S per mole of material, denoted g, v, and
s. Eq. (14-14) becomes g = v p – s T. Now consider the coexistence curve p(T).
Because l = g along the curve, we have l = g and gl = gg, yielding,

vl p – sl T = vg p – sg T. (14-15)

Rearranging terms,

Δp sg − s l
= . (14-16)
ΔT v g − v l
With vg – vl = v the volume change per mole, and s = hlg/T the entropy change per
mole, the slope of a liquid-gas phase boundary takes on the general form,

dp h lg
= , (14-17)
dT T v
where hlg ≡ Hlg/NA is the latent heat per mole. Because the liquid is much denser
than the gas, the volume change per mole of material is given by the ideal gas law: V/n
= RT/p. The above equation, known as the Clausius-Clapeyron equation, is very useful.
By measuring how the boiling point depends on pressure (i.e., following along the phase
boundary), one can determine the latent heat of a liquid. An example is given in the
exercises for this chapter. Alternatively, if one knows the latent heat, one can determine
the pressure dependence of the boiling point of the liquid.

In most cases a substance expands when solid turns into liquid (v > 0) and the pro-
cess requires a positive heat of transformation (Hlg > 0). Therefore, the slope of the
coexistence curve dp/dT is positive. Water is unusual in that it contracts upon melting,
implying that dp/dT is negative. To illustrate this difference, the phase diagrams for
water and carbon dioxide are plotted below, along with a table of critical pressures and
temperatures:
160
Processes at Constant Pressure Chapter 14

H2O CO 2
p p

Liquid Liquid
Pcp
Solid Solid
Ptp

Gas Gas

T T
Ttp Tcp

pcp Tcp ptp Ttp


H2O 218 atm 374°C .006 atm .0098°C
CO2 72.8 atm 31°C 5.1 atm –56.6°C
(Values from Zumdahl’s “Chemical Principles.”)

C. Chemical Reactions at Constant Pressure


Consider the following example involving the gas reaction, N2 + 3H2 ↔ 2NH3. What is
the energy required to produce NH3 from N2 and H2, under constant pressure? From
our discussion of enthalpy, the total thermal energy required to cause this reaction is
H = HNH3 – (HH2 + HN2).

Chemical tables do not give absolute values of H for a substance. Instead they list the
standard enthalpy of formation Hfo, which is defined as the energy required to form 1
mole of a compound from its elements, with all substances in their standard states (T = 298 K,
p = 1 atm). Thus, Hfo = zero for elemental gases (e.g., N2, H2, O2, Ar), liquids (Hg, Br)
and solids (Cu, Al). Here is a subset of entries from the Appendix in Zumdahl’s book
(so = molar entropy; and Gfo = Hfo – Tso is the Gibbs energy required to form one
mole of the substance):

Substance and Hfo so Gfo


state (kJ/mol) (J/mol-K) (kJ/mol)
N2 (gas) 0 192 0
H2 (gas) 0 131 0
NH3 (gas) –46 192.5 –16.5
O2 (gas) 0 205 0
H 2O (gas) –242 189 –229
H2O (liquid) –286 70 –237

161
Physics 213 Elements of Thermal Physics

In Chapter 12, we assumed that an energy 2 = 0.9 eV (= 1.44  10–19J) is released


when two molecules of NH3 are created from one molecule of N2 and 3 molecules of
H2. This value is consistent with the empirical data given in the above table. First
realize that the internal energy change is related to the standard enthalpy change by
Uo = Ho – (pV).

Since each ammonia molecule has binding energy , the energy needed to create 1 mole
(NA = 6.023  1023 molecules) of NH3 in a “vacuum” is:

Uo = – NA = – NA  (1.44  10–19J)/2 = – 43.4 kJ.

However, to create 2 moles of NH3 from 1 mole of N2 and 3 moles of H2, all at 1 atm,
there must be a factor of 2 decrease in volume because 4 moles of H2 and N2 gas are
replaced by 2 moles of NH3. (Remember that one mole of an ideal gas has the same
volume as a mixture of ideal gases adding up to one mole.) Therefore, to produce 2
moles of NH3, we have (pV) = –2RT, and for 1 mole of NH3,
p = 1 atm

(pV) = – 1RT = – 8.314 J  298 K = – 2.5 kJ,

for a total enthalpy of,

Hfo = U + (pV) = – 43.4 kJ – 2.5 kJ = – 45.9 kJ,

which is consistent with the value given in the table. In fact, I obtained the value of
2 = 0.9 eV from the empirically measured Hfo = – 46 kJ/mol given in this table,
working the problem in reverse.

162
Processes at Constant Pressure Chapter 14

Exercises
1) In chemical reactions, the law of mass action is frequently given in terms of partial
pressures of the components, pi = NikT/V = niRT/V, where ni = Ni/NA. For the
reaction aA  bB → cC, the equilibrium relation is:

pcC
= K p (T),
paA pbB

where Kp denotes the equilibrium constant applicable to partial pressures. In Chapter


11 we estimated the (concentration-based) equilibrium constant for the ammonia
reaction to be K(300 K) = 6  108 L2/mol2. Show what the general relation is be-
tween K and Kp and evaluate Kp(300 K) for the ammonia reaction.

2) A hypothetical liquid boils at 40°F in atmospheric pressure, and it boils at 45°F


when the pressure is raised to 1.1 atm. Use the Clausius-Clapeyron equation (14-17)
to determine the latent heat of this liquid.

3) Using the table given in Section C, determine the amount of energy 2 (in eV)
required to form two molecules of H2O from two molecules of H2 and one molecule
of O2. This is twice the binding energy of the water molecule.

4) a) Using the Clausius-Clapeyron equation (14-17), derive an equation for the


vapor pressure p(T) of a liquid in terms of the gas constant R and the latent
heat per mole L.

b) With a vapor pressure of the form (13-12) derived for a solid-gas phase transi-
tion, show the relationship between the latent heat per mole L and the bind-
ing energy  of a molecule in the liquid, thus connecting the macroscopic and
microscopic worlds.

163
Physics 213 Elements of Thermal Physics

164
APPENDIX

Vibrations in Molecules and Solids—


1 Normal Modes*

In Chapter 1 we introduced the problem of vibrations in a diatomic


molecule, simply represented by the following diagram (with a spring
replacing the actual binding forces):

m κ m  = spring constant

To compute the vibrational frequency of this object, we wrote two


coupled differential equations for the displacements u1 and u2 of ball
1 and 2,

m d2u1/dt2 = –(u1 – u2)

m d2u2/dt2 = –(u2 – u1).

These equations are turned into two linear equations in two unknowns
by assuming solutions of the form, u1 = A1sin t and u2 = A2sin t, where
A1 and A2 are the vibrational amplitudes of atoms 1 and 2. Rearranging
terms a little, we have,

( – m 2) A1 –  A2 = 0

– A1 + ( – m 2) A2 = 0

*Optional reading for Physics 213 165


Physics 213 Elements of Thermal Physics

These two equations can be written in matrix form:

⎛ κ − mω 2 −κ ⎞ ⎛ A1 ⎞ ⎛ 0 ⎞
⎜ ⎜
2 ⎟ ⎜
⎟ =⎜
⎝ −κ κ − mω ⎠ ⎝ A 2 ⎟⎠ ⎝ 0 ⎟⎠

The two linear equations have a non-zero solution only if the determinant of the matrix
equals zero. Dividing by  and defining X = m2/, we see that the matrix has the form:

(1 − X ) −1
=0
−1 (1 − X )

With the definition of the determinant, given at the end of the appendix, you will find
the solutions X = 0 and 2, giving  = 0 and (2/m)1/2. The solution  = 0 corresponds
to the center-of-mass motion of the molecule.

The adventurous reader may follow this procedure in order to determine the resonant
frequencies associated with the triatomic linear molecule:

m κ m κ m

This exercise is a little more complicated (involving a 33 matrix), but the general
procedure is extremely important. The definition of a 33 determinant is given at the
end of this chapter. The three coupled equations of motion are:

m d2u1/dt2 = –(u1 – u2)

m d2u2/dt2 = –(u2 – u1) – (u2 – u3) (A1-1)

m d2u3/dt2 = –(u3 – u2)

What we are doing here is finding the “normal modes” of a system of oscillators, i.e.,
the natural resonances of the system. You will find that there are two compressional
resonant frequencies associated with the triatomic linear molecule. In general, a system
of N masses connected by springs and allowed to vibrate in one dimension will
have N – 1 normal modes, each with a distinct frequency. N masses in 3 dimensions
have 3N degrees of freedom. For a solid, these 3N degrees of freedom equal (3 c.m.
translational) + (3 rotational) + (3N–6 vibrational).

Finding the normal-mode frequencies is the first step in completely solving the prob-
lem, which would also tell us the ratios of amplitudes A1:A2: .. :AN of the atoms in a
particular normal mode. In this course, we will not solve for the amplitudes, but the
computer program demonstrated in lecture shows the results for various situations. You

166
Vibrations in Molecules and Solids—Normal Modes Appendix 1

may experiment with this normal-mode simulation, called NORMALMODE, written


by Professor Jim Smith at UIUC and loaded onto the computers in the lab room.

Professor Smith’s computer simulation shows the actual motions associated with the
normal modes for (up to) 20 masses. The results are quite amazing. The normal modes
turn out to be standing waves. See the figure for N = 5 at the end of this section. For
20 atoms, there are 19 different wavelengths for these standing waves, corresponding
to the N – 1 = 19 different frequencies. The wavelengths of the vibrational waves are:


n = 2L/n, (A1-2)

where L = Na is roughly the length of the line of atoms and n is an integer ranging
from 1 to 19. The displacement of atoms in the nth standing wave is roughly given by:

u(x,t) = A (sin knx) (sin nt), (A1-3)

where kn = 2/
n is the wavenumber, and n’s are determined by finding the roots of
the 20th order equation (“diagonalizing” the 2020 matrix).

You may notice that the equation for a standing wave is a little different from that of a
traveling wave,

u(x,t) = A sin (knx – nt). (A1-4)

In fact, a standing wave is the superposition of two traveling waves, one moving in the
+x direction, sin (knx – nt), and one moving in the –x direction, sin (knx + nt), as you
can verify by using the trigonometric identity, sin A + sin B = 2sin((A+B)/2) sin((A–B)/2).
The speed of each traveling wave is v = n/kn = fn
n.

Definition of 22 and 33 determinants:

A B C
A B
= AD − BC, D E F = AEI + BFG + CDH – AFH – BDI – CEG
C D G H I

Exercises
1) Imagine a spring that stretches 6 cm when a 1 kg mass is hung on it. Its spring
constant is  = . Is the “spring constant” between the two hydrogen
atoms in a hydrogen molecule larger or smaller than this? Take a
guess, then compute H2 using the measured vibrational frequency of the diatomic
molecule H2: f = /2 = 6.5  1013 Hz.

2) Calculate the normal mode frequencies of a hypothetical diatomic molecule bonded


to a solid:

167
κ m κ m
Physics 213 Elements of Thermal Physics

No r ma l Mo d e s fo r N = 5

L = Na

At rest

d i s pl ac ement u

n = 1 λ1 = 2L

n = 2 λ2 = L

n = 3 λ 3 = 2 L /3

n = 4 λ 4 = L /2

168
APPENDIX

2 The Stirling Cycle

The purpose of a heat engine is to convert heat into work. A straight-


forward heat engine that finds itself in some practical applications is the
Stirling engine. The cycle of the Stirling engine has four steps, which
we will take one at a time. Throughout, W = work by the gas.

We start with the gas at room temperature, Tc. In Step 1, we fix the
volume of the gas cylinder and immerse it in boiling water. The idea is
to give it some energy to do work in the next process.

Step 1

p
p = NkT/V
Va

Q1
Th

Tc

Boiling water, Th = 373 K Va V

169
Physics 213 Elements of Thermal Physics

This is clearly an irreversible process, because heat does not flow from cold to hot.
Because no work is done (pV = 0), the first law tells us:

Q1 = U = Cv (Th – Tc) =  Nk (Th – Tc). (Nk = nR)

In Step 2, the hot gas performs isothermal work by pushing out the piston.

Step 2

p
Vb
W2
Va
Q2
Th

Tc

Boiling water, Th = 373 K Va Vb V

The work done in step 2 by the gas is:

Vb Vb
dV V
W2 = ∫ pdV = NkTh ∫ = NkTh ln b
Va Va V Va

Because the gas’s temperature is held constant, its internal energy does not change.
From the First Law:

U2 = Q2 – W2 = 0.

The heat transferred by the hot reservoir to the gas is

Vb
Q2 = W2 = NkTh ln
Va
The total heat transferred from the hot reservoir in steps 1 and 2 is

Vb
Qh = Q1 + Q2 = Nk(Th – Tc ) + NkTh ln
Va
We have successfully converted thermal energy from the hot reservoir into work. Now
we must reset the system to its original condition. In Step 3, we begin re-setting the
gas temperature by again fixing the volume and taking the container out of the hot
bath into the room:
170
The Stirling Cycle Appendix 2

Step 3

p
Vb

Q3

Th

Tc

Room temperature, Tc = 293 K Va Vb V

This is an irreversible process. No work is done. The heat transferred to the environ-
ment is Nk(Th – Tc ).

Finally, we arrive at Step 4, which recompresses the gas at low temperature:

Step 4

p
Vb
W4
Va

Q4
Th

Tc

Room temperature, Tc = 293 K Va Vb V

The work required in Step 4 to reset the gas isothermally at Tc is given by

Va Va
dV V
W4 = ∫ pdV = NkTc ∫ = NkTc ln a
Vb Vb V Vb

171
Physics 213 Elements of Thermal Physics

Note that the work done by the gas is negative because VbVa; i.e., we do work on the
gas. The heat transfer in step 4 is calculated as in step 2.

The total work done by the gas in the entire cycle is:

Vb
Wby = W2 + W4 = Nk (Th – Tc) ln
Va
The efficiency of any engine is

work done by the system Wby


ε =
heat extracted from thee hot reservoir Qh

For the Stirling cycle, we have the final result

Vb Vb
Nk(Th − Tc )ln (Th − Tc )ln
Va Va
ε= =
V Vb
αNk(Th − Tc ) + NkTh ln b α (Th − Tc ) + Th ln
Va Va

For a volume ratio Vb/Va = 2, you can show that the efficiency is 14.6%. Notice that as
the volume ratio becomes large, the efficiency of the Stirling cycle approaches that of
the Carnot cycle,  = 1 – Tc/Th (= 21% in this case).

172
APPENDIX

3 Statistical Tools

A. Statistical Definitions and Concepts


The statistics of many particles involves probabilities. We introduce our
discussion of statistics by considering the results of a hypothetical exam:

Number of 6
papers with 5
score s,
4
N(s)
3

0
0 1 2 3 4 5

Score, s

173
Physics 213 Elements of Thermal Physics

This graph tells the final results of the exam. There are no uncertainties or probabilities
involved. There is, however, an experiment that you can do with this distribution that
involves probabilities. Take all the papers, shuffle them, and ask, for example, “What is
the probability that I will pick out a paper with an exam score of 2?” The exam distri-
bution given above now becomes a probability distribution. The probability of picking
out an exam with score s is:

P(s) = N(s) /  N(s) = N(s) / N, (A3-1)

where N is the total number of exam papers. (N = 20 in the example.) A plot of the
probability of choosing a paper with score s is therefore,

.45

.40
P (s)
.35

Probability of choosing .30


a paper with score s .25

.20

.15

.10

.05

0
0 1 2 3 4 5

Score, s

Here we are dealing with the actual probability, rather than a probability density P(s),
like P(E) considered in Chapter 3. The probability of choosing a score s is P(s), where
P(s) is a pure number ranging between 0 and 1.

The probabilities summed over all possible values of s add to 1:

Σ P( s ) = 1. (A3-2)
s

In the above example, the sum is from s = 0 to 5. The average value of s is:

< s > = Σ sP( s ) (A3-3)


s

which, you can verify, equals 3.60 for the above example. Notice that sometimes a bar
over the variable is used instead of < >. I mention this because both designations are
used in the literature.

174
Statistical Tools Appendix 3

Finally, we need a measure of the width of the distribution. The “deviation” defined by
s = s – <s> has the average value:

< Δs > = Σ( Δs )P ( s ) = 0 , (A3-4)


s

because s can be positive or negative. To gauge the width of the distribution, we must
define the standard deviation, or root-mean-square deviation,

σ d = < ( Δs )2 > = [ Σ (s)2 P(s) ]1/2 . (A3-5)


s

In the example above, verify that d = 1.07 by summing from s = 0 to s = 5.

Sometimes it is convenient to describe a discrete system with a continuous function.


We simply evaluate the function at discrete points.

P (s) P(s)

Notice the difference between probability P(s) and probability density P(s), which is
by definition a continuous function. They are related by P(s) = P(s) s, where s is the
interval between the discrete values of s.

You may be worried about dealing with the statistics of continuous variables because
it involves integrals rather than sums. In fact, sums are often more difficult to compute
than integrals, unless you have a computer program that happily computes either sums
or integrals. Generally, when we give you a problem with integrals, we will tell you
how to evaluate the particular integral. Appendix 4 is a table of all the integrals we will
need in this course.

Finally, just a few general statements about probabilities. There is no uncertainty about
a probability distribution. It is a very definite function (e.g., P(E) = CE1/2exp(–E/kT)
for the ideal gas) or result (e.g., the distribution of grades given above). The fuzziness
comes about when one uses these distribution functions to predict the occurrence of
a given event. For example, if one asks “in a given experiment, how many molecules
will I find within a certain range of energies?” or “after 100 tries, how many times will
I pick an exam with a score of 3 from the stack of exams?”, the answers to these ques-
tions have an uncertainty.

If you could count the molecules within the range E, the result would be pretty close
to N P(E) E. Likewise, if you “shuffle the pile and choose an exam” 100 times, a
histogram of the results will closely resemble the P(s) graph above, multiplied by the
number of experiments, 100.

Such is the way of statistics.

175
Physics 213 Elements of Thermal Physics

B. The Gaussian Distribution


A particularly useful function in statistics is the Gaussian function, which may be writ-
ten as a function of the continuous variable x:

2
e -x 1

.607 Area under curve = π 1/2

x
0 1

A Gaussian probability density is written:

P(x) C
P(x) = Ce -x /2 σ d
2 2

.607 C Area under curve = C (2π σ d 2 ) 1/2 = 1

x
0 σd
(A3-6)

where d = (<(x)2>)1/2 is the standard deviation of the distribution, as proven below. A


displacement of x = d from the peak of the distribution corresponds to P( d) = C e–1/2 =
0.607 C. A probability density must be normalized by integrating over all possible x
values:

+

∫ P(x)dx = 1.
−

Consulting the table in Appendix 4, we find,

+ 1/ 2
⎛ π⎞
∫−  e dx = ⎜⎝ a ⎟⎠ ,
2
− ax

(A3-7)

which yields the normalization constant

C = (2 d2)–1/2 . (A3-8)

176
Statistical Tools Appendix 3

The probability density is therefore,

1 2
/ 2σ 2d
P( x ) = e− x .
2πσ 2 (A3-9)
d

The mean square displacement is given by,

< x2 > =  x2 P(x) dx. (A3-10)

Consulting our integral table again, we use

+ 1/ 2
⎛ π ⎞
∫−  x e dx = ⎜⎝ 4a3 ⎟⎠
2
2 − ax
(A3-11)

to show

< x2 > = d2 . (A3-12)

The d in the Gaussian formula is indeed the root mean square (rms) deviation, or
standard deviation, as stated earlier.

C. Gaussian Approximation to the Binomial Distribution


For the case of N spins at B = 0, we can approximate the binomial distribution with
a Gaussian function by setting x → m, where m = Nup – Ndown. The probability density
(A3-9) becomes,

1 2
/ 2σ d2
P( m ) = e− m
2σ 2
d

However, for a given problem, m is either an odd or even integer depending on whether
N is odd or even, i.e., m = 2. Thus, in order to evaluate the probability P(m) of observ-
ing the spin excess m, we multiply the probability density by the interval m = 2 (see
Section A):

P(m) = P(m) 2 and d2 = < m2 > = N

leading to the final result,

1 2
P( m ) = e − m / 2N
2

177
Physics 213 Elements of Thermal Physics

D. Standard Deviations Applied to the Random Walk


The standard deviation d is determined from,

d2  < x2 > = ∫ x2 P(x) dx.

In applying the Gaussian formula to a binomial distribution, we must determine d in


terms of the binomial “step.” For the random walk problem, the net displacement for
M equal-length steps is x = si where the step size si = ± x. Therefore,

<x2> = <(si) (sj)> = <(si2)> + <(sisj)> .

The second term vanishes due to the equal number of + and – terms, leaving the first
term with exactly M terms, all equaling <si2> = x2. So,

d2 = <x2> = Mx2

This analysis explains why statistical fluctuations in a “random walk” are proportional
to the square root of the number of “steps,” d  M½.

A binomial distribution with large M is well approximated by a Gaussian probability


density, Eq.(A13-9) evaluated at discrete points, e.g., x = mx (m = integer):

M steps of constant size x and net displacement x:


M left P(x) with x = (Mright – Mleft)x = mx and d2 = Mx2
M right
or,
M steps of mean size x and net displacement x:
m P(x) with x = si, (si variable) and d2 = 2 Mx2

The factor of 2 in the latter case comes about because the probability of a step length
s (always positive) has the form p(s) = (1/ x)exp(–s/ x). From the integral table in Ap-
pendix 4, we find

∫ p(s) ds = 1

<s> = ∫ s p(s) d s = x, and

<s2> = ∫ s2 p(s) d s = 2 x2 ,

where x is the mean step length. We therefore conclude

< si2> = <si2> = 2x2 = 2 Mx2 ,

as stated in Section C of Chapter 5.

178
APPENDIX

4 Table of Integrals

Here is a table of commonly used integrals involving the Boltzmann,


Gaussian, and Planck distributions. a and b are constants.

1
∫e
− bx
dx =
0
b

2!
∫x e
2 − bx
dx =
0
b3

3!
∫x e
3 − bx
dx =
0
b4

n!
∫x e
n − bx
dx = Positive integer n, b > 0
0
b n+1


1 1 π
∫e
− ax 2
dx =
∫e
− x2
dx = π
2 0
2 a
0



1 π To find ∫E½ e–bE dE,
∫x e
2 − ax 2
dx =
∫ xe dx =
2
− ax

0
2a 0
4 a 3/ 2 etc., use E = x2

 
1 3 π
∫ x e dx = ∫x e
2
3 − ax 4 − ax 2
dx =
0
2a 2 0
8a 5/ 2

x3 π4
∫0 e x − 1 15
dx =
179
Physics 213 Elements of Thermal Physics

180
APPENDIX

Exclusion Principle and Identical


5 Particles

Each quantum-mechanical state of an electron can hold only a single


electron. Multiple occupancy is forbidden. For example, the 1s orbital
of an atom has two possible states for an electron: 1s↑ and 1s↓, where
the ↑ and ↓ represent the electron spin. The Pauli Exclusion Principle
prohibits more than one electron from occupying each state. This is the
basis for the periodic table of atoms, which have electronic configura-
tions 1s22s23p6… according to the exclusion principle. Single occu-
pancy is the rule for electrons (and for all particles with half-integral
spin, known as fermions).

Let us now apply this single-occupancy rule to our prior example with
bins (Chapter 6) and see what happens. Begin with two distinguishable
particles, A and B and figure out the possibilities:

181
Physics 213 Elements of Thermal Physics

= ______ (distinguishable particles, single occupancy)

Your answer should be 12. The result for the general case of M bins and N particles is:

M!
Ω= N distinguishable particles, M single occupancy bins
( M − N )!

As an example, imagine 4 distinguishable particles arranged in 10 single-occupancy bins:

A B C D

As we have seen, the total number of possible arrangements is

M! 10!
Ω= = = 5040.
( M − N )! 6!

If we now change to 4 identical particles, we have overcounted the number of arrange-


ments. Consider the four specific bins occupied by A, B, C, and D above. What is the
total number of ways those bins can be occupied?

A can be in 4 positions Axxx xAxx xxAx xxxA


Each of these has 3 possibilities for B Bxx xBx xxB
Each of these has 2 possibilities for C Cx xC
Leaving only one spot for D D,

yielding 4! = 24 permutations of ABCD. Therefore the number of total arrangements


for identical particles is:

= 5040/24 = 210.

182
Exclusion Principle and Identical Particles Appendix 5

In general, changing from distinct to identical particles reduces the number of micro-
states by N!.

M!
Ω= N identical particles, M single occupancy bins
( M − N )! N !

This result is exact. For large N and M and in the dilute gas limit, N<<M, this latter
result becomes approximately equal to the unlimited-occupancy case for identical particles
(Equation (6-3)),

M! MN
≈ .
( M − N )! N ! N !

To see that this latter result is reasonable, consider the case M = 100, N = 4:

M! 100 × 99 × 98 × 97 M 4
= ≈
( M − N )! N ! 4 × 3× 2× 1 4!

In using this approximation, we are restricted to the “low density limit,” also known as
the “classical limit.” This result should come as no surprise because in the low-density
limit (N<<M), there is little chance for particles to attempt to occupy the same state.
In considering the ideal gas at low densities we use the latter form for because it is
algebraically simpler.

In dealing with factorialized numbers, there is a very useful formula known as Stirling’s
formula, valid for N>>1,

N ! = 2πN N N e − N

ln(N!) = N ln N – N + (1/2) ln(2N) ,

which, for large numbers, is well approximated by

ln(N!)  N ln N – N .

183
Physics 213 Elements of Thermal Physics

184
APPENDIX

6 Sum over States and Average Energy

As we have seen in Chapters 8 and 9, the “constant” C multiplying the


Boltzmann factor in Eq. (8-8) is not really a constant. The probabilities
must all add up to 1, so C depends on both the distribution of states
and the temperature. Consider a hypothetical distribution of states at
the following two temperatures:

C1
Pn = C1exp(-βEn) low T

E
E1 E2 E3 E4 . . . . . . .

Pn = C2exp(-βEn) higher T
C2

E
E1 E2 E3 E4 . . . . . . .

185
Physics 213 Elements of Thermal Physics

In each case, C must be chosen such that the sum of probabilities over all states equals
one: Cexp(–En) = 1. This means that C = 1/exp(–En). It is convenient to define
the “the sum over states” or “partition function”,

Z = Σ e − βE n

n (A6-1)

such that,

e −βE n

Pn = . (A6-2)
Z

Z = Z(T) represents the number of states that are likely to be occupied at temperature
T, just as (U) represents the number of accessible states for an isolated system with
energy U. If the function Z(T) is known, you can determine the average energy of the
system by noticing what happens when you take the derivative of Z with respect to :

dZ
= − Σ E n e − βE n

dβ n

The average energy is defined by

1
< E > = Σ E n Pn = Σ E n e – βE ;n

n Zn
Therefore,

1 dZ
<E>=− . ( = 1/kT) (A6-3)
Z dβ

The harmonic oscillator discussed in Chapter 8 has a simple ladder of energy levels,

En = nε
0 ε 2ε 3ε 4ε 5ε 6ε . . .

In this case the sum over states becomes:

Z = Σ e − β E n = Σ e − β nε = Σ x n
n n n

(A6-4)
where
x  e–

186
Sum over States and Average Energy Appendix 6

The sum over states is evaluated using a neat little math trick:

Z = Σn x n = 1 + x + x 2 + x 3 +  = 1 + x(1 + x + x 2 +  ) = 1 + xZ (A6-5)

which yields the result,

1 (A6-6)
Z=
1− x

Using Equations (A6-3) and (A6-6), the average energy of a harmonic oscillator in
contact with a thermal reservoir becomes:

ε
<E> = βε
(A6-7)
e −1

187
Physics 213 Elements of Thermal Physics

188
APPENDIX

7 Debye Specific Heat of a Solid*

The statistical counting of phonons in a solid is almost identical to that


for photons, given in Chapter 9. The main differences in these two cases
are 1) we must replace the velocity of light with the velocity of sound,
and 2) there is an upper limit to the vibrational frequencies of a solid,
fmax  vsound/a, where a is the atomic spacing in the lattice. In contrast,
photon frequencies have no upper limit.

At low temperatures the total vibrational energy of a solid increases


as U  T4 but when kT becomes comparable to  = hfmax, the thermal
energy assumes the dependence U = 3nRT given by the Equipartition
Theorem. The crossover between these regimes occurs at a temperature
where kT is roughly equal to hfmax.

Consequently, the solid at high temperature has a constant specific heat


given by 3R, whereas at low temperature the solid has a specific heat
(per mole) of the form,

cv = (1/n) dU/dT  234 R (T/TD)3 (A7-1)

*Reference material not generally covered in Physics 213

189
Physics 213 Elements of Thermal Physics

where TD is known as the Debye temperature, characteristic of a particular solid. These


ideas are sketched below for a crystal of diamond.

Cv

3R

Diamond

Cv ∝ T3 ( Carbon
Atomic Weight = 12g/mol )
atoms

T
0 500K 1000K 1500K

The equation above shows that the specific heat of a solid achieves about one-half of
its high temperature value (i.e., 234R(T/TD)3 = 3R/2) at approximately (1.5/234)1/3 =
19% of the Debye temperature. The Debye temperature depends on the masses of the
atoms in the solid and the strength of the atomic bonds. For typical solids, the Debye
temperature is in the range 100–500 K, so the specific heat of most solids at room
temperature is close to the Equipartition value 3R.

For crystalline diamond, having light carbon atoms and stiff covalent bonds, the maxi-
mum phonon frequency is high and the corresponding Debye temperature is measured
to be 2230 K. This implies that for diamond, the crossover temperature is about 400
K, as indicated in the graph above.

Exercise
1) How much heat must be removed from a 1-gram crystal of diamond in order to
lower its temperature by 10 K when its initial temperature is, a) 1000 K, b) 100 K,
c) 10 K? Give a physical explanation for your results.

190
APPENDIX

8 Absolute Entropy of an Ideal Gas

In Chapter 11, we showed the entropy of a classical gas of indistinguish-


able particles has an entropy associated with its spatial coordinates,

⎛ n ⎞
S = k ln Ω = Nk ⎜ ln c + 1⎟
⎝ n ⎠

where nc = 1/V is the density of cells chosen, and n = N/V is the density
of particles. When the wave nature of the particles is taken into account,
the entropy of the ideal gas is given by the Sakur-Tetrode equation,

⎛ n 5⎞
S(N,V,T) = Nk ⎜ ln Q + ⎟ , (A8-1)
⎝ n 2⎠

with, the quantum density.


3/ 2
⎛ mkT ⎞
nQ = ⎜
⎝ 2π 2 ⎟⎠
(A8-2)

191
Physics 213 Elements of Thermal Physics

Practical forms for n and nQ (to about 1% accuracy) are:

p ⎛ p 300 K ⎞
n= = 2.45 × 1025 meter −3 ⎜
kT ⎝ 1atm T ⎟⎠

3/ 2
⎛ m T ⎞
−3
n Q = 10 meter ⎜
30

⎝ m p 300 K ⎠

where m/mp is the molecular weight of a particle, and mp  mH.

Sometimes it is convenient to deal with the entropy per mole of material, i.e., the en-
tropy for NA = 6.022  1023 particles. With Nk = nR from Chapter 5, the molar entropy,
s = S/n, of an ideal monatomic gas is,

⎛ n 5⎞
s = R ⎜ ln( Q ) + ⎟  strans
⎝ n 2⎠ (A8-3)

where R = 8.314 J/mol-K is the gas constant. This entropy applies only to translational
degrees of freedom, and so is denoted strans. Diatomic molecules have rotational and
vibrational degrees of freedom that also contribute to the total entropy, as we will con-
sider in Appendix 9.

Below we tabulate the calculated quantum density and entropy for two monatomic gases
at their boiling temperature for p = 1 atm. Also listed are the measured entropies under
these conditions, as discussed below. We also list the translational entropy, strans (Eq.
(A8-3)), for some diatomic and polyatomic gases at 300 K and 1 atmosphere. We wish to
compare the calculated entropy strans to the experimentally measured total entropy smeas:

strans smeas
gas m/mp T (K) nQ (per m3)
(J/mol-K) (J/mol-K)
Ar 40 87.3 3.93  1031 129.5 129.8
monatomic
Ne 20 27.2 2.42  1030 96.5 96.4
H2 2 300 2.79  1030 118 131
diatomic N2 28 300 1.46  1032 150 192
O2 32 300 1.79  1032 152 205
H2O 18 300 7.55  1031 145 189
polyatomic
NH3 17 300 6.93  1031 144 193

192
Absolute Entropy of an Ideal Gas Appendix 8

The excellent agreement between experiment and theory for the two monatomic gases
confirms the validity of statistical mechanics and quantum mechanics. The discrepancy
between Strans and Smeas for diatomic and polyatomic molecules implies that they gain
entropy from degrees of freedom that we have not accounted for, i.e., their “internal”
motions.

The agreement between experiment and theory for the monatomic gases is pretty
amazing. Think about it. The measurement of absolute entropy requires the experi-
menter to cool the particles down to near-zero Kelvin, where they form a solid. Small
increments of heat Q are added to the system and the temperature T is recorded at
each step. The measured increments in entropy, S = Q/T, are added up as the solid
heats up, turns into a liquid, warms further, and vaporizes into gas. The values of smeas
in the tables above represent the sum, S = (Q/T) per mole, over this torturous path
involving two phase transitions. Yet, the end result is accurately predicted by the statisti-
cal properties of an ideal gas of quantum particles!

For diatomic molecules at room temperature, the internal motions are largely associ-
ated with rotations. In Appendix 9, we discuss the entropy associated with translation
plus rotations of diatomic molecules such as N2, and H2, which has the form:

⎛ n 7⎞
S = Strans + Srot = Nk ⎜ ln( Q Z r ) + ⎟
⎝ n 2⎠ (A8-4)

where Zr  kT/r and r is a quantum of rotational energy. Zr roughly equals the aver-
age number of rotational energy levels that are occupied per particle.

Exercise
1) a) Compute the absolute molar entropy of a gas of N2 at 300 K, assuming no internal
motions of the molecules, and compare your answer to the number given in the table
on page 192. b) Now compute the absolute entropy at 300 K by including internal
rotations of the molecules. Use Equation (A8-4) and assume that the quantum of
rotation is r = 0.00026 eV. Is it closer to the measured value in the table? 10% is
a reasonable discrepancy considering the approximations made.

193
Physics 213 Elements of Thermal Physics

194
APPENDIX

9 Entropy and Diatomic Molecules*

In this section we will consider the entropy of ideal diatomic gases. In


Chapter 11 we stated that the entropy of a monatomic gas with particle
density n = N/V is

⎛ n 5⎞
S = k ln(Ω ) = Nk ⎜ ln Q + ⎟ ,
⎝ n 2⎠

where nQ = (mkT/2 2)3/2. The monatomic gas has only translational


degrees of freedom, so this formula is also denoted Strans.

In addition to translating, diatomic molecules can rotate and vibrate.


Quantum mechanics dictates that these motions have discrete energy
levels. The entropy associated with these motions involves a sum over
these quantum states. Entropy is an additive property, so we simply add
these contributions to the translational entropy: S = Strans + Srot + Svib.

First consider vibrations of a diatomic molecule with mass 2m and


spring constant . The quantum-mechanical harmonic oscillator has rotation
equally spaced energy levels, En = nv = nhf, with f = (2/m)1/2/2 the
frequency of vibration. In Chapter 5, we defined q = U/v as the number
of energy quanta in the system. The number of accessible states for a
system of N oscillators is,
vibration
( q + N − 1)! q N
Ω= ≈ . (A9-1)
( N − 1)!q ! N!

*Reference material generally not covered in Physics 213. 195


Physics 213 Elements of Thermal Physics

In the last step we have assumed that N and q are large numbers and that q >> N, the
“classical limit.” Taking the logarithm and using the Stirling approximation, ln(N!) 
NlnN – N,

⎛ q ⎞ ⎛ U ⎞
S vib ≈ Nk ⎜ ln( ) + 1⎟ = Nk ⎜ ln( ) + 1⎟
⎝ N ⎠ ⎝ Nε v ⎠
(A9-2)
⎛ kT ⎞
S vib ≈ Nk ⎜ ln( ) + 1⎟ when KT > ε v
⎝ εv ⎠

In the last step, we used the result U  NkT from Chapter 8, as also predicted by the
Equipartition theorem (Chapter 3).

[Detail: The ratio (kT/v) in the above equation is actually an approximation for the
exact sum over states for the harmonic oscillator, Zvib = 1/(1–e–/kT), which does not
assume kT > v. This “sum over states” was derived in Appendix 6. For most diatomic
molecules at 300 K, kT > v ; therefore, Zvib  1 and <E>/kT < 1, implying Svib is small
compared to other contributions to the entropy.]

Rotational energy levels are determined by the quantization of angular momentum. It


turns out that the number of states with energy less than E equals N(E) = E/r. (For
more details, see the discussion by Schroeder, referenced in the Preface.) This result
is similar to the result for vibrations, N(E) = E/v. The rotational entropy for diatomic
molecules has a form similar to the vibrational entropy,

Srot = Nk(ln(Zr) + 1) (A9-3)

where Zr  kT/r when kT >> r.

Rotational energy levels are generally more closely spaced than those of vibration. For
example, for hydrogen (H2), v  .27 eV and r  .015 eV, compared to kT = .026 eV
at 300 K. For molecules like H2 and N2 only rotations (not vibrations) are thermally
active at 300 K, and the total entropy has the form

⎛ n 7⎞
S = Strans + Srot = Nk ⎜ ln( Q Z r ) + ⎟ (A9-4)
⎝ n 2⎠

and Zr  kT/r when kT >> r. See Kittel and Kroemer (ref. Preface), p. 83, for a more
exact treatment.

196
Entropy and Diatomic Molecules Appendix 9

Finally, there may be spin degrees of freedom. As discussed in Chapter 5, electrons have
spin = ½, which gives rise to two energy levels, E = ± B (“spin-up” and “spin-down”)
in a magnetic field B. For N spins the total number of states, , equals 2N; therefore,

Sspin1/2 = k ln = Nk ln(2).

In general,

Sspin = Nk lnZspin. (A9-5)

For B = 0, Zspin equals the number of spin levels, 2S + 1, where S is the total spin quantum
number. To determine the spin S of an atom or molecule, one must consider the number
of unpaired electrons. H2 and N2 have no unpaired electrons, so Sspin = Nk ln(1) = 0.
For O2, the ground molecular level has two unpaired electrons with S = 1, providing
three magnetic sublevels, ms = -1, 0, +1, that lead to Sspin1 = Nk ln(3).

Adding the terms S = Strans + Srot + Svib + Sspin involves multiplying the arguments of the
logarithms, i.e., replacing nQ with nQZint, or pQ with pQZint, where Zint = ZvibZrotZspin.
Zint represents the contribution of internal motions to the number of thermally acces-
sible states for a molecule. This procedure also applies to F, G and chemical potential
. Specifically,
⎛ n ⎞ ⎛ p ⎞
μ = k ln ⎜ ⎟ = k ln ⎜ ⎟ (A9-6)
⎝ n Q Z int ⎠ ⎝ pQ Z int ⎠

with n = N/V. The following table gives rough values of ’s and Z’s for common diatomic
molecules at 300 K, where kT = 0.026 eV:

Zint =
molecule m/mp r (eV) Zrot  kT/r Zvib
ZvibZrotZspin
H2 2 .015 2 1 2
N2 28 .00026 100 1  100
O2 32 .00036 72  1.07  230

We can now see if internal motions account for the difference in Smeas and Strans that
we tabulated in Appendix 8. Adding the contributions from rotation and spin using the
values of Zint in the table yields the results: (Recall molar entropy, s = S/n = R ln(Zint))

strans +
gas Zint strans srot smeas
srot
H2 2 118 14 132 131
N2 100 150 47 197 192
O2 230 152 54 206 205
[all entropies in units of (J/mol-K)]

197
Physics 213 Elements of Thermal Physics

The measured values of entropy are taken from the Appendix in Zumdahl’s text Chemi-
cal Principles.

The estimates of Strans + Srot given here should not be taken too seriously. The as-
sumption Zr  kT/rot tends to overestimate this contribution to the entropy. The
approximation kT/rot >> 1 is not well met for H2. Considering the approximations,
the agreements within a few percent are quite acceptable. Mainly, however, this table
shows that the internal motions of the molecules contribute significantly to the total
entropy of a molecular gas.

Internal motions also affect the equilibrium constants in chemical reactions. Basically,
one must replace nQ with nQZint for each of the molecules, where Zint represents the
number of states due to internal motions. For the reaction aA + bB ↔ cC, this produces
another factor,

Z cint C
Z aint A Z bint B

multiplying the right side of Eq. (12-12), and included in k(T).

Internal motions affect the heat capacities of ideal gases. Specifically, Cv = Q/T =
TS/T. When rotations are present, the temperature dependence of S is Nk ln(T)5/2
because nQ has a T3/2 dependence and Zrot  T . Therefore,

Cv = T dS/dT = (5/2)Nk ,

as predicted by the Equipartition Theorem for diatomic molecules.

198
APPENDIX

10 Vapor Pressure of a Vibrating Solid*

In Chapter 13 we derived the vapor pressure of a simple solid—one with


no internal motions or entropy. Real solids have internal motions and
therefore significant entropy. It is interesting to see how the entropy of
a vibrating solid affects the vapor pressure. Do you think that allowing
vibrations in the solid will Increase or Decrease the vapor pressure of the
solid? Take a guess: I D . Let’s do the problem. It requires us
to know the entropy associated with oscillations.

Assume that the solid has a “typical” vibration energy v and that
kT > v. We considered the harmonic oscillator in Chapters 7 and 8.
The energy of Ns harmonic oscillators able to vibrate in 3 dimensions
is Uvib = 3NskT. In Appendix 9 we show that the entropy is,

⎛ kT ⎞
S vib ≈ 3N S k ⎜ ln( ) + 1⎟ (A10-1)
⎝ εv ⎠
As considered in Chapter 13, an atom in the solid is bound by ; there-
fore, U = –Ns  3NskT. The Helmholtz free energy, F = U – TS, for
the solid is:

⎛ ⎛ kT ⎞ ⎞
FS = − N S Δ + 3N S kT − 3N S kT ⎜ ln ⎜ ⎟ + 1⎟
⎝ ⎝ εv ⎠ ⎠

⎛ kT ⎞
= − N S Δ − 3N S kT ln ⎜ . (A10-2)
⎝ ε v ⎟⎠

199
*Reference material generally not covered in Physics 213.
Physics 213 Elements of Thermal Physics

We differentiate Fs to get the chemical potential:

dFS ⎛ kT ⎞ ⎛ kT ⎞
μS = = − Δ − 3kT ln ⎜ ⎟ = − Δ − 3kT ln ⎜ . (A10-3)
dN S ⎝ εv ⎠ ⎝ ε v ⎟⎠

Setting s equal to the chemical potential of the gas, g = kT ln(p/pQ), we find the vapor
pressure of the harmonic solid with binding energy ,

⎛ ε ⎞
p vapor = pQ ⎜ v ⎟ e − Δ /kT .
⎝ 3kT ⎠ (A10-4)

Vibrations have produced an extra factor, not present in Eq. (13-12). For typical solids
at 300 K, the ratio v/3kT ranges from 0.1 to 0.3. Therefore, vibrations in the solid
cause a decrease in the vapor pressure. How can that be? Doesn’t the added motion in
the solid increase the possibility of knocking an atom into the gas, thereby increasing
the vapor pressure?

The answer is in the influence of entropy. By allowing vibrations, we have created more
states in the solid. More states in the solid mean more probability of finding atoms in
the solid, and thus lower vapor pressure. This effect is illustrated graphically by plot-
ting the chemical potential for several pressures, and plotting the corresponding phase
diagram (as in Chapter 13).

0 T

μ g = -kT ln
( ) (
pQ
p
p Q ∝ T 5/2
)

- μ s = - no vibrations

kT
p1 p2 p3 μ s = - - 3kTln ( )
εv
with vibrations

p = p Q e -/kT no vibrations
p3 Solid region
with vibrations

Gas region
p2
p1
T

200
Solutions to Exercises

Introduction
1) 1, 10, 45, 120, 210, 252, 210, 120, 45, 10, 1

2) W = ∫pdV = NkT(1/V)dV = NkTln(V2/V1).

Chapter 1
1) Using Won = (KEcm) + U, where U is the PE when the spring is
included in the system:

Mass alone: ∫ Fdu = mvf2/2 + 0


Mass plus spring: 0 = mvf2/2 + (u2 – uo2)/2
Because ∫ Fdu = –∫ u du = –(u2 – uo2)/2, the equations are equiva-
lent.

2) a) vcm = (2Fd/m)1/2 from c.m. Eq., not Work-Energy Eq.

b) Considering the “point” of contact between tire and road, can


you see that the force applied by the earth acts through zero dis-
tance? The earth isn’t moving, and a force acting through zero
distance does no work. If the earth were providing the energy
for the car to move (by doing work), we would not need fossil
fuel. c) Internal chemical energy U converted to c.m. energy.
d) Work-Energy Eq.: Won = (KEcm) + U = 0 implies U =
mv2cm/2.

201
Physics 213 Elements of Thermal Physics

3) Believe Equation (1-14). In one of the two cases, a constant force acts through a
greater distance and imparts rotational motion, but the center-of-mass accelera-
tion…

4) a) c.m. Eq.: Fdcm = mv2cm/2 = 18 J, vcm = 2.12 cm/s.


b) Work-Energy Eq: FD = mv2cm/2 + U = 21 J, U = 21 J – 18 J = 3 J.

Chapter 2
1) Consider how T behaves as energy is added to the system. Only one dependence
is reasonable.

T T T

U U U
(a) (b) (c)

2) Stot = (3/2)k (10 ln(U1) + 40 ln(U2)) , U2 = Utot – U1

dStot/dU1 = (3/2)k (10/U1 – 40/U2) = 0 in equilibrium

U2 = 4 U1 in equilibrium (maximum entropy)

Stot

S2

S1

U1

Equilibrium U tot

3) S = (3/2)Nk ln(U) + Nk ln(V) + constants


(S/U)V = (3/2)Nk/U; (S/V)U = Nk/V
(S/U)V = 1/T; 1/T = (3/2)Nk/U gives U = (3/2)NkT for ideal monatomic gas.

S = (S/U)V dV + (S/V)U dU and dU = -pdV


S = (1/T) dU + (Nk/V) dV = (-(1/T)p + (Nk/V)) dV.

202
Solutions to Exercises Physics 213

As in Eq. (2-9), for non-zero dV, the coefficient must vanish in equilibrium
(S = 0):
(1/T)p = (Nk/V) reduces to pV = NkT, the ideal gas law.

4) S = S1 + S2 . For the process shown, S = dU1/T1 + dU2/T2 = -Q/T1 + Q/T2


S = Q(1/T2 – 1/T1) is negative for T1 < T2 implying decreasing entropy, which
is in violation of the Second Law of Thermodynamics.

Chapter 3
1) dS/dU = 1/T = (3/2) Nk/U
dS = (3/2) Nk dU/U Integrating:
S2 – S1 = (3/2) Nk (ln(U2) – ln(U1))
S = (3/2) Nk ln(U) + constant for an ideal monatomic gas

2) a) Ideal gas law: pV = NkT = (105 Pa)(2  10–3 m3) = 200 J = 2 liter-atm/K
Diatomic gas: U = (5/2) NkT = 500 J Assume pure N2 below:

b) n = pV/RT = (1atm)(2 liter)/(.082 liter-atm/mol-K)(300 K) = .0813 mol


For N2 gas, mgas = (.0813 mol)(28 g/mol) = 2.3 g = 2.3 “pennys”

c) mgh = 500 J, h = (500 J)/(10–3 kg)(10 m/s2) = 50,000 meters

d) penny: (1/2)mvf2 = 500 J, vf = (1000 J/.001 kg)1/2 = 1000 m/s


gas: (1/2)m<v2> = (3/2)NkT (translational energy only)
average speed: v = (3kT/m)1/2 = 517 m/s using
m = (.028 kg/mol)/(6.022  1023atom/mol) = 4.65  10–26 kg and
kT = (1.38  10–23J/K)(300 K) = 4.14  10–21 J. The purpose of this exercise is
to give a sense for how fast the molecules are moving.

e) U = cv T = (5/2) Nk (3 K) = 5 J

3) mgh = CV T. CV = 3nR = 3(8.314 J/mol-K)(1 mol/.012 kg)(1 kg) = 2079 J/K.
T = 10 J / (2079 J/K) = .005 K. Not much.

4)
E/kT .1 .25 .5 1.0 1.5 2.0 3.0
(E/kT)1/2e–E/kT .286 .389 .428 .368 .273 .191 .086

203
Physics 213 Elements of Thermal Physics

Maxwell-Boltzmann Distribution
.4 for an ideal gas of particles

(E/kT)1/2e-E/kT
.2

0 E/kT
0 1 2 3

average at 1.5 kT

peak at .5 kT

The area in the rectangle is about ¼ the total area under the curve. Therefore the
probability is in the range 20–30%. In Chapter 9 we will see how to compute the
probability more accurately with integrals.

Chapter 4
1) This process is described in Appendix 2. Here T = 100 K.
nR = (3/2)(.1 mol)(8.314 J/mol-K) = 1.247 J/K, ln(Vb/Va) = .693

Stage Q W U
1 124.7 J 0 124.7 J
2 215.0 J 215.0 J 0
3 – 124.7 J 0 – 124.7 J
4 – 156.7J – 156.7J 0

 = 17% between boiling and freezing water

2) For a diatomic gas,  = 5/2 and  = ( + 1)/ = 7/5.


U = nRT = pV nR = pV/T = .00333 l-atm/K

p
a
FLT:
3 1 ΔU = Q - W
c
b
2

204
Solutions to Exercises Physics 213

Process 1: VbTb = VaTa → Tb = Ta(Va/Vb)1/ = 227 K


(Q = 0) paVa = paVb → pb = pa(Va/Vb) = .379 atm
Wby = –U = (paVa – pbVb) = .605 l-atm

Process 2: pbVb/Tb = pcVc/Tc → pc = pb(Vb/Vc) = .758 atm


(U = 0) Wby = nRTb ln(Vc/Vb) = –.524 l-atm = Q

Process 3: Q = U = nR(Ta – Tc) = .608 l-atm (Tc = Tb)


(Wby = 0)

3) a) dQ = dU + pdV. For non-interacting (ideal) gases, U = U(N,T).


At constant V, dV = 0. Therefore, CV = (dQ/dT)V = dU/dT.
For an ideal gas, V = nRT/p, so Cp = dU/dT + p(dV/dT) = CV + nR even for
CV(T).

b) CV = nR and Cp = CV + nR = nR + nR = ( + 1)nR = ( + 1)CV/.

c) Diatomic gas in the high-T limit where rotations and vibrations are active.

Chapter 5
1) a) (8,5) = 8!/5!3! = 8  7 = 56
P(8,5) = (8,5)/28 = 56/256 = .219, or about 22%.
b) P(400) = (2/N)1/2 exp(–0/2N) = (2/N)1/2 = .028, or about 3%.

This is a distribution with a width (<m2>)1/2 = N1/2 = 28.3.

2) m = 2Nup – N, so Nup = (N+m)/2 and Ndown = N – Nup = (N–m)/2.

5) For the Ar with mass m = .040 kg/6.022  1023 = 6.64  10–26 kg,
v = (3kT/m)1/2 = (3  1.38  10–23  300/6.64  10–26) = 432 m/s
 = 0.1 m = 10–7 m
D = v  / 3 = (432 m/s )(10–7)/3 = 1.44  10–5 m2/s
Diffusion distance in 1 sec = (6Dt)1/2 = (6  1.44  10–5  1)–1/2 = .94 cm
Diffusion distance in 3600 sec = .94 cm  (3600)1/2 = 56 cm

2.5

2.0

1.5
rrms(cm)
1.0

0.5

0.0 t (s)
0 1 2 3 4 5 6

205
Physics 213 Elements of Thermal Physics

6) exp(–x12/2 d2) = ½
x12/2 d2 = ln 2
x1 = (2ln 2)1/2 d = 1.177 d
FWHM = 2x1 = 2.355 d
2 2 σ 2
e -x d

0.5

x
0 x1

7) a) For the electrons with mass m = 9.11  10–31 kg,


v = (3kT/m)1/2 = (3  1.38  10–23  300/9.11  10–31) = 1.16  105 m/s
 = v = (1.16  105)(10–11) = 1.16 m
D = v  / 3 = (1.16  105 m/s )(1.16  10–6)/3 = .045 m2/s
b) Diffusion distance in lifetime = (2D 0)1/2 = (2  .045  10–6)–1/2 = .3 mm

v   D
N2
500 m/s 0.2 ns 0.1 m 1.67  10–5 m2/s
Molecules
electron 1.16  105
0.01 ns 1.16 m .045 m2/s
gas m/s

For comparison, in 1s an N2 molecule would diffuse approximately (2D 0)1/2 =


(2  1.67  10–5  10–6)–1/2 = 5.8 m, or 50 times smaller distance than the elec-
tron, even though the N2 molecule collides 20 times more often. The reason for
the quicker diffusion of the electron is that it is much lighter than the molecule.

Chapter 6
1) log (MN/N!) = NlogM  Nlog N  N = N(log(M/N)  1)
= 1020(log (103)  1) = 4  1020
As expected, less than 23  1020 for distinguishable particles

2) a) Counting states or using the formula: = 1 2 = 23 × 21 = 16, = ln(16) =


2.77.
b) We expect position 3 because the number of particles per cell are the same on
both sides. (See calculation in part c for a justification.)
c) = 1331 + 2321 + 3311 = 3 + 16 + 27 = 46; = ln(46) = 3.83
If no partition, = 4 ln 4 = 5.55. Removing constraints increases S.
206
Solutions to Exercises Physics 213

3) Number of cells = M = V/V; = MN/N!


f/ i = (Vf/Vi)N = (1.01Vf /Vi)N = (1.01)1000 = 21,000
The number of microstates increases very rapidly with V, even for just 1000 particles.

4) = VN/N!; therefore, SHe = k(N lnV – ln N!) = SAr .


Sinitial = 2k(N lnV – ln N!) and Sfinal = 2k(N ln(2V) – ln N!).
ln(2V) = lnV + ln2; therefore, S = 2Nk ln2 = entropy of mixing.

For identical gases: Sinitial = 2k(N lnV – ln N!) and Sfinal = k(2N ln(2V) – ln(2N)!)
With ln(2N)! = 2N ln2N – 2N, we have Sfinal = k(2N ln(2V) – 2N ln(2N) + 2N)
Sfinal = 2k(N(lnV + ln2) – N(lnN + ln2) + N) = 2k(N lnV– N lnN + N) = Sinitial
Adding and removing a partition does not change the gas (or S) in this case. Because
entropy does not change, the process is reversible, consistent with the Second Law.

Without the N! term in the latter case, Sinitial = 2k(N lnV) and Sfinal = k(2N ln(2V)),
leading to the unphysical result that S = 2Nk ln2. All atoms of the same type are
identical to one another, a fact that was not appreciated in J.W. Gibbs time, leading
some scientists to reject his (now famous) theory of statistical mechanics. (Gibbs
Paradox)

Chapter 7
1)
0.5

0.4
.33
0.3 .250
Pn .179
0.2 .121
.071
0.1 .037
.012
0.0
0 1 2 3 4 5 6 ×ε

2) a) By a factor of about 299 = 6.3  1029. b) a factor of 11.

3) Units: p1 = 1 atm = 105 Pa, V1 = 1 liter = 10–3 m3 = V2/10


Ideal gas law: Nk = pV/T = 100/300 = 1/3.
a) Sv = Nk ln(V2/V1) = (1/3)(ln 10) = .767 J/K.
b) VT = constant with  = 3/2, so T2 = T1(V1/V2)2/3 = 64.6 K.
c) ST = CV ln (T2/T1) = – .767 J/K with CV = (3/2)Nk.
d) S = Nk ln(V2/V1) + CV ln (T2/T1) = 0
Entropy change is zero in an adiabatic process (Q = 0)

207
Physics 213 Elements of Thermal Physics

4) a) S = ncvln(Tf /Ti) = (5/2) nR ln(2/3) = 2.5 (2) (8.314) (–.405) = – 16.96 J/K
b) S = ncp ln(Tf /Ti) = (7/2) nR ln(2/3) = 3.5 (2) (8.314) (–.405) = – 23.57 J/K

5) a) kT is the average energy per oscillator and  = hf is the energy of a quantum, so


kT/ = 6.25 quanta/oscillator, for a total of q = 3N  6.25 = 1.88  1023 quanta in
the solid. b) From = q3N–1/(3N–1)!, ignore the 1’s, and use Stirling’s relation,
= ln = 3N(ln(q/3N) +1) = 8.5  1022. Notice that the amount of entropy
for a vibrating solid at room temperature is roughly equal to the number of
oscillators (3N) times the logarithm of the # of quanta per oscillator (ln 6.25).

Chapter 8
1) a) B/kT = (9.27  10–24 J/T)(1T)/(1.38  10–23 J/K)(2K) = 0.336
Pup/Pdown =e2B/kT = e0.672 = 1.96. There are nearly twice as many spins pointing up
as down under these conditions. Nup/Ndown = Pup/Pdown = 1.96. b) B/kT =
ln(10) = 2.3, B = (2.3/2) kT/ = 3.42 Tesla. c) M = N2B/kT = 0.336 N =
0.31 J/T. Also, N tanh(B/kT) = 0.300 J/T.

2) e  1 +  → <E> = /(e – 1)  –1 = kT, U  N<E> = NkT

3) <En> =  / (e/ kT–1)


kT = 1.38  10–23  300 = .414  10–20 J /kT = .5/.414 = 1.21
<En> =  /(3.345 –1) = (.426)  = .213  10 J
–20

4) a) kT = 1.38  10–23  300 = .414  10–20 J /kT = .5/.414 = 1.21


define a shortand, X = e–/ kT = .299
Pn = e–En/ kT /  e–En/ kT  e–En/ kT = 1 + X + X2 + X3 = 1.415

Probability: Po = 1/1.415 = .707 Population: N0 = 100 Po = 70.7


Probability: P1 = .299/1.415 = .211 Population: N1 = 100 P1 = 21.1
Probability: P2 = .0893/1.415 = .063 Population: N2 = 100 P2 = 6.31
Probability: P3 = .0267/1.415 = .019 Population: N3 = 100 P3 = 1.9
Check that sum of N’s equals 100: 99.9

b) <En> =  Pn En = .707  0 + .211   + .063  2 + .019  3 = .394


This result is a little lower than the harmonic oscillator case (.426) in which
there is some population of levels at higher energy levels than 3.

1
<En>

Pn

En
0 ε 2ε 3ε
208
Solutions to Exercises Physics 213

Chapter 9
1) a) First determine C from the normalization condition:
1 =  P(E) dE = C  E2 e–E/kT dE = C (kT)3 2! giving C = 1/2(kT)3

Average energy:
<E> =  E P(E) dE = C  E3 e–E/kT dE = C (kT)4 3! = 3 kT

b) The energy of a particle is (1/2)m(vx2 + vy2 + vz2) + (1/2)(x2 + y2 + z2).


Equipartition says that each quadratic term gets (1/2) kT, totaling 3 kT.

2) a) The volume of the sphere is V = M/ = 1kg/(2.7g/cm3)


= 3.7  10–4 m3 = 4R3/3. The surface area is A = 4R2
= 2.49  10–2 m2, so the total energy radiated per second is
JU  A = (5.670  10–8 W/m2 K4 T4) (293 K)4 (2.49  10–2 m2) = 10.4 W

b) At 300 K, the specific heat of aluminum is a constant at 3R = 3 


(8.314 J/mol-K) = 24.9 J/mol-K and the number of moles of Al in the 1 kg
sphere is n = (1 mole/0.027 kg )(1 kg) = 37 moles, so the heat capacity near
300K is Cv = 3nR = 922 J/K. Assuming an average temperature of 283 K in the
above equation, the energy loss of the sphere per second is,

U/t = Cv T/t = 9.1 watts

Therefore we find the time to cool 20 K is,

t = Cv (T) / (9.1 W) = (922 J/K) (20 K) / (9.1 W)


 2000 seconds = 33 minutes.

Chapter 10
1)  > c implies Qc/Qh > Tc/Th and thus Qc/Tc > Qh/Th, which, in turn, implies that
Stot = –Qh/Th +Qc/Tc  0. Entropy decreases, against the law.

An example of a real engine is one in which there is a heat leak directly between
the hot and the cold reservoir. The total entropy change due to this heat leak is S
= –Qleak/Th +Qleak/Tc = Qleak(1/Tc – 1/Th) = cQleak/Tc. It is straightforward to show
that the work done by the real engine equals the work done by a Carnot engine
minus the heat loss due to friction and leakage, i.e., Wby = cQhtot – TcS.

2) Without much thought, you might have blindly plugged into the formulas and
obtained the following results:
Wby =  Qh for a Carnot cycle (most efficient engine).
 = 1 – Tc/Th = 1 – 393/293 = .254
Qh = cv M T = 4184  75  100 = 31.4 megajoules
Wby = .254  3.14  107 = 7.97  106 J = 8.0 megajoules (Wrong!)

The  and W computed above are incorrect because the difference in temperature
between the hot and cold reservoirs is varying. This isn’t such a trivial problem, is it?
Try again, realizing that Th and the efficiency are changing with time. Set Th = T:
209
Physics 213 Elements of Thermal Physics

Wby =   dQ = –  (1 – 293/T) cvM dT integrating from 393 to 293.


= – cvM (  dT -  (293/T) dT )
= – cvM (– 100 K – 293 ln(293/393) )
= – cvM (– 100 K + 86 K ) = 4.38  106 = 4.38 megajoules

Do you see why the output is only about half the previous (erroneous) result? The
average T (and therefore Qh) is roughly half that used in the incorrect calculation.

The next problem shows a simpler approach with free energy.

3) Wby = –F = –U + TenvS = –CT + TenvC ln(Tf/Ti), Tenv = 300 K, C=1 kJ/k
a) Wby = 150 J –121.64 J = 28.36 J
b) Wby = –150 J +207.94 J = 57.94 J

4) ΔF = ΔU − TΔS = (3 / 2) NkΔT − NkTΔ(ln( n Q V / N ))


ΔT = 0 Δ(ln( n Q V / N )) = Δ(ln( V )) = NkT(lnVf - lnVi )
ΔF = − NkT ln(Vf /Vi )
Vf Vf

Wby = ∫ pdV = ∫ ( NkT / V)dV = NkT ln(V /V )f i


Vi Vi

Chapter 11
1) a) n = 2.45  1025 m–3 nQ = 9.88  1029 m–3 (4)3/2 = 7.9  1030 m–3
 = kT ln(n/nQ) = (0.026 eV)(–12.7) = –0.33 eV.
b) U = (3/2) nRT = (3/2)(8.314 J/K)(300 K) = 3741 J
S = nR(ln(nQ/n) + 5/2) = (8.314)(ln(3.22  105) + 2.5) = 126 J/K
F = U – TS = 3741 J – (300 K)(126 J/K) = – 34.1 kJ

2) a) At constant volume nQV = (mkT/22)3/2V  T3/2.


s = S/n = Rln(nQ/n) + const.
s(T) = Rln(T3/2) + const. = (3/2)Rln(T) + const.
cV = T(ds/dT) = T (3/2)R/T = (3/2)R
At constant pressure, V = NkT/p and nQV/N  T5/2
s(T) = (5/2) Rln(T) + const.
cV = T(ds/dT) = T (5/2)R/T = (5/2)R
b) At constant V, (nQV/N)(kT/r)  T5/2
s(T)  (5/2)Rln(T), cV = (5/2)R
At constant pressure, (nQV/N)(kT/r)  T7/2
s(T)  (7/2)Rln(T), cV = (7/2)R

3) ho = kT/mg = (1.381  10–23)(270)/(28  1.674  10–27)(9.8) = 8117 meters


p = (1 atm) exp(–104/8117) = 0.29 atm. You need a breathing apparatus.

210
Solutions to Exercises Physics 213

dp dn
4) Flayer = A(p(x) – p(x +x)) = − A Δx = − kT A Δx
dx dx
Flayer Flayer kT dn dμ
Feffective = = =− =− Q.E.D.
N nAΔx n dx dx
5) Feff – mg = 0, –(kT/n) dn/dh = mg, dn/dh = –(mg/kT)n = –n/ho , ho= kT/mg
Pressure p = nkT, so dp/dh = –(p/ho) with the solution p = poexp(-h/ho).

Chapter 12
1) nQ = 1030(28)3/2(T/300)3/2 m–3 n = 2.45  1025 (300/T) m–3
nQ/ n = 6.0  10 (T/300) = 1 when T = 0.6 K. N2 is not an ideal gas at T = 0.6
6 5/2

K. It liquifies at 77 K. The ideal gas equations do not apply.

2) Use nQ = 1030 m–3 (m/mp)3/2(T/300 K)3/2 me/mp = 1/1836


Notice that nQH = nQp to good approximation, neglecting spin.

3) 2H2O = 2H2 + O2 N2 + 3H2 = 2NH3 H2 = 2H

4) a) ne2 – nend – ni2 = 0 ne = nd/2 + (nd2 + 4ni2)1/2


b) ni = 5.2  1015 m–3 at 300 K, and nd = 1014 m–3, so ne = 5.25  1015 m–3
When nd << ni then ne  ni. Impurities have little effect.
If nd = 1017 m–3 then ne = 1.0027  1017 m–3.
When nd >> ni then ne  nd. Impurities have a big effect.

5) a) n1 / n 2 = ⎡⎣ n Q1 / n Q 2Z int ⎤⎦ e
2 2
= K (T)
– Δ / kT

b) n1 = sqrt {(1.225  10 )(10 )(.03/.052)e–76.9} = 5.3  1010 /m3


25 30

Chapter 13
1) Because N! = 1 2 3 … N, then lnN! = ln1 + ln2 + …lnN. Taking dN = 1, we see
that d(logN!)/dN = lnN! – ln(N–1)! = ln(N!/(N–1)!) = lnN.

2) a) pQ = nQkT = (9.88  1029)(40)3/2 (1.38  10–23)(300) = 1.03  1012 Pa


= 1.02  107 atm e–/kT = 9.75  10–6, 2.08  10–7, 4.45  10–2
f = 1/(1 + po) = 0.01, 0.321, 0.957
b) po = 1 implies f = Ns/M = ½, so the left side of Eq. (13-7) equals 1, and p/pQ =
exp(–/kT). The chemical potential of the gas is  = kT ln(p/pQ) = –. In this
case,  = (.026 eV)(–16.1) = –0.42 eV.

3) pQ = (4  104 atm) (32)3/2 (310/300)5/2 = 7.86  106 atm


pQZint = (7.86  106 atm)  211 = 1.8  109 atm
f = ½ means p = po = 0.2 atm kT = (8.617  105)(310) = .0267 eV
po = pQZinte–/kT = 0.2 atm  = kT ln(pQZint/po) = 0.61 eV

4) a) Q = 3.335  105 J/kg  1 kg = 3.34  105 J, t = 6  3600 = 21600 s


Power = Q/t = 3.34  105 J / 21600 s = 15 watts.
b) Stot = SH2O + Sroom
= 3.34  105 J (1/273 K – 1/293 K) = 83.5 J/K
211
Physics 213 Elements of Thermal Physics

5) F = U – TS = NkT(ln(N/nQ(V–Nb)) – 1) + pV – aN2/V (mgh = pV)


dF/dV = –NkT/(V–Nb) + p + aN2/V2 = 0 →
(p + aN2/V2)(V – Nb) = NkT
(van der Waals equation)

Chapter 14
pcQ n cQ
1) = a b (kT)c −a− b
C C
because pQ = nQkT
paQA pQb n QA n Q
B B

Kp = (kT)–2 K for ammonia (p in Pascal units)


Kp = (RT)–2 K for ammonia (p in atmosphere units)
RT = .0821 L-atm/mol-K  300 K = 24.6 L-atm/mol, (RT)2 = 607 (L-atm/mol)2
Kp = 6  108 L2/mol2 /607 (L-atm/mol)2 = 1  106 atm–2

2) First we must take care of all the units. A 5° Fahrenheit change corresponds to
5  (5/9) = 2.8° Centigrade or Kelvin change, and the average temperature of 42°F
equals 279 K. One atmosphere is 105 Pascal, so p = 104 Pa. The latent heat is
L1 = T (p/T) (RT/p)
= (279 K)2 (104 Pa/2.8 K) (8.314 J/mol-K)/(105 Pa)
= 23 kJ/mol.

3) 2H2 + O2 ↔ 2H2O, 2 moles of H2O from 3 moles of reactants implies (pV) = 1RT,
or (pV) = 0.5 RT = 1.25 kJ for 1 mole of H2O.
Uo = Ho – (pV) = –242 kJ/mol – 1.25 kJ/mol = –243 kJ/mol = 2.56 eV

4) a) dp/dT = L/TVm = Lp/RT2


dp/p = (L/R)dT/T2
Integrating: ln(p) = (L/R)(1/T)  constant
Solution: p = (const.) exp (L/RT)

b) L/RT = /kT and nR = Nk


Result: nL = N = total binding of N molecules

Appendix 1
1)  = mg/x = 163 Nt/m. H2 = m2/2 = (1.67  10–27 kg)(2  6.5  1013)2/2 
140 Nt/m. I chose this spring to give you a feeling for H2.

2)  = ((3 ±
5)/2/m)1/2 = (.38/m)1/2 and (2.62/m)1/2.

212
Solutions to Exercises Physics 213

Appendix 7
1) CV = n cv and n = 1g / (12g/mol) = .083 mole.
a) Considering the plot of cv given in the text, 1000 K is in the high temperature
limit where cv = 3R (see Chapter 5 for derivation).
CV = .083  3  8.314 J/mol-K = 2.08 J/K
Q = CV T = 2.08 J/K  10 K = 20.8 J

b) From the graph this is roughly in the low temperature (quantum) limit,
CV = n 234 R (T/TD)3 = (.083) 234  8.31  (T/2230)3 = 1.45 (10–8 T3)
Q =  CV dT = 1.45  10–8  T3 dT = 1.45  10–8 [T24 – T14]/4
Q = 3.63  10–9 [1004 – 904] = .125 J

c) Also in the low temperature limit,


Q = 3.63  10–9 [104 – 04] = 3.63  10–5 J
which is 570,000 times less heat than needed to make the 10 K change at
1000 K. Most of the vibrational modes are in their ground level, and a tem-
perature rise cannot promote them to the next quantum level.

Appendix 8
1) a) n = 2.45  1025 m–3 nQ/n = 5.98  106 at 300 K.
strans = R(ln(nQ/n) + 5/2) = 8.314(15.6 + 2.5) = 18.1R = 150.5 J/mol-K.
b) kT/r = .026/.00026 = 100. srot = R(ln(kT/r)+1) = 5.6R = 46.6 J/mol-K.
strans + srot = 150.5 + 46.6 = 197 J/mol-K.
Very close to smeas = 192 J/mol-K.

213
Physics 213 Elements of Thermal Physics

214
Index

Accessible states
Absolute ................................................................................. 125, 127, 191
Harmonic oscillator ............................................................... 105, 186, 195
Ideal gas ............................................................. 67, 85, 105, 118, 126, 148
Adiabatic process ..............................................................ix, 32–37, 40, 85, 207
Adsorption ........................................................................... xviii, 145, 148–149
Ammonia reaction ........................................................................................137
Average value ...................................................................... 22–23, 28, 174–175
Avogadro constant .................................................................................ix, xi, 21

Binomial distribution ....................................ix, xx, 46, 48–51, 60, 68, 177–178


Boltzmann constant....................................... ix, xi, 23, 79–80, 91–92, 117, 185
Boltzmann distribution ............................................................. xviii, 91–92, 98
Boltzmann factor ........................................ ix, 27–28, 88, 91, 99, 102, 140, 185
Bricks doing work.................................................................................114–116

c.m. equation .............................................................................................. 5, 10


Carnot cycle
Efficiency ............................................................. 38–39, 41, 172, 209–210
Carrier densities ...........................................................................................141
Center of mass .................................................................................. 4–5, 10, 31
Chemical potential .......................................................................................124
Diatomic gas .......................................................................... 129, 133, 159
Interpretation.........................................................................................130
Liquid .............................................................................................151–152
Monatomic gas.......................................................... ix, 128–129, 133, 143
Solid ....................................................................... 145, 151–152, 160, 200
Classical thermodynamics ..................................................................xiv, 15, 68

215
Physics 213 Index

Clausius–Clapeyron equation ..............................................................160, 163


Coexistence curve .................................................................................150, 160
Curie’s law ...................................................................................... 93, 121–122

Debye solid ...........................................................................................189–190


Diatomic molecules ........................................................................................24
Abundances ..............................................................................................21
Entropy .................................................................. 192–193, 195–196, 198
Thermal energy ................................................................................. 25, 96
Diffusion
Differential equation ...............................................................................53
Electrons in silicon ..................................................................................57
Diffusive force ......................................................................................131–132

Einstein solid ..........................................................................................86, 119


Energy ..............................................................................................................1
Bands ......................................................................................................139
Center of mass ...........................................................................................6
Internal................................................................................................. 6, 24
Kinetic ........................................................................................................2
Potential .....................................................................................................7
Rotational...................................................................................................6
Vibrational .................................................................................................8
Energy levels ................................................................................................. xiv
1D electron in a box ................................................................................74
H–atom ....................................................................................................74
Harmonic oscillator ........................................... 74, 99, 186, 195–196, 208
Enthalpy ...........................................................................ix, 154, 158, 161, 163
Entropy
And heat .......................................................... x, xvii, 36, 39, 111, 144, 155
Conventional ....................................................................................... x, 79
Diatomic gas ....................................................................................85, 129
Molar................................................................................ 29, 161, 192–193
Monatomic gas........ 18, 29, 83, 85, 103, 119, 126–129, 158, 192, 195, 202
Vibrational ............................................................... 14, 114, 151, 192, 196
Volume exchange ..........................................................xvii, 64, 78, 81, 124
Entropy ionization .......................................................................................136
Entropy maximization ........................................................................ 16, 64, 68
Equilibrium ....................................................................................................32
Chemical ................................................................................................137
F–minimum ...................................................................................116, 124
Statistical ..................................................................................................59

216
Index Physics 213

Equilibrium constant....................................................................................136
H–ionization ..........................................................................................136
H2–dissociation ......................................................................................143
With internal motions ...................................................................138, 199
Equilibrium value ...........................................................................................60
Equipartition Theorem ..................................................... 23, 80, 103, 117

First Law of Thermodynamics (FLT)............................................................31


Free energy (F) .....................................................................................115–117
Ideal gas .................................................................................................128
Paramagnet ............................................................................................119
Surface states..........................................................................................146
Free expansion ..............................................................................................114
Fuel cell ........................................................................................................132
Fundamental postulate of statistical mechanics................................ xvi, 49, 59

Gaussian distribution ................................................... 50, 52–54, 57, 176–177


Gibbs free energy ......................................................................ix, xix, 157–159
Gibbs paradox.................................................................................................71

Harmonic oscillator ........................................................................... 74–77, 95


Average energy...............................................................................186–187
Sum over states .............................................................. 185–187, 195–196
Heat ................................................................................................................31
And engines........................................................................................31–44
And entropy ..............................................................xvii, 37, 111–115, 154
Heat capacity ..................................................................................................24
Constant pressure ....................................................................................44
Harmonic oscillator .................................................................................86
Ideal gas ...................................................................................................25
Solid .........................................................................................................25
Heat conduction .............................................................................................54
Heat current (J) ........................................................................................54–55
Heat pump .......................................................................................... 40, 42, 55

Ideal gas law............................................................................................26, 122


Integrals (table).............................................................................................179
Irreversibility .................................................................. xvi–xvii, 13–16, 68–69
Isothermal process....................................................................................33–37
217
Physics 213 Index

Latent heat ........................................................... 153–156, 160, 163, 211–212


Law of atmospheres .............................................................................129–130
Law of mass action ...............................................................................142, 163
Liquid–gas condensation......................................................................154–156
Logarithms ..................................................................................................... xx

Magnetic moment .............................................................. 46–50, 92, 119–121


Magnetism ......................................................................................................46
Maxwell–Boltzmann distribution.........................................................102–103
Mean free path .......................................................................................52, 178
Microstates ........................................xvi, xx, 45–47, 59–65, 68–69, 75–79, 181
Moles .................................................................................................. 21, 24, 26
Momentum conservation ........................................................................... 4, 13
Myoglobin .................................................................................... 147–148, 156

Newton’s Second Law .......................................................................... 2–5, 7–8


Normal modes ...................................................................... 9, 13–14, 165–167

Occupancy rules .....................................................................................62, 181

Paramagnetism .......................................................................................91, 119


Phase diagrams .....................................................................................150–155
Solid–gas ................................................................................................150
Solid–liquid ............................................................................................151
Solid–liquid–gas .............................................................................152, 161
Photons .................................................................................................103, 105
Planck Radiation Law ..................................................................................106
Planck’s constant ............................................................................ 74, 104, 125
Polymer ....................................................................................................94–95
Pressure ..........................................................................................................22
Probability ........................................................................xvi, 27–29, 45, 49–50
Probability density ..................................................... 27–28, 99–102, 174–178
Process energies .............................................................................................32

218
Index Physics 213

q–formula for harmonic oscillator .................................................................77


Quadratic terms......................................................................................23, 117
Quantum density .................................................. 126–127, 141, 143, 191–192
Quantum mechanics........................................... xiv, 46, 64, 119, 125, 193, 195
Quantum pressure ........................................................................ 146, 149, 159
Quasi–static processes ....................................................................................32

Random walk ..........................................................................................50, 178


Refrigerator ..............................................................................................40–42
Reversible process .......................................................... 34, 111–115, 206–207

Second Law of Thermodynamics ...................... xvi, 14, 16, 40, 68–69, 82, 122
Specific Heat ..........................................................ix, 24, 26, 97, 122, 189–190
State functions .......................................................................... 32, 36, 112, 118
Statistical mechanics........................xiv, 23, 27–28, 49, 59, 71, 79, 88, 118, 125
Statistical tools..............................................................................................173
Stefan–Boltzmann Law ................................................................................106
Stirling cycle ................................................................................... 43, 169–172
Stirling’s approximation ................................................................... 69–71, 183
Sum over states ................................................................................... 74–77, 95

Temperature
Definition ..................................................................xvii, 15, 18, 23, 78–82
Scales ........................................................................................................23
Thermal conductivity ...............................................................................54–55
Thermal equilibrium .......................................15, 73, 78, 81–82, 116, 123–124
Thermal Radiation ...............................................................................105–106
Thermal reservoir ................. 33–34, 36, 38, 87–88, 95, 99, 116–119, 123, 157
Thermodynamic identity .....................................................................112, 160

Van der Waals equation ........................................................................154, 156


Vapor pressure ...................................................... 148–151, 160, 163, 199–200

219
Physics 213 Index

Work .............................................................................................................3–5
Adiabatic ................................................................................ 32–34, 36–40
Engines ................................................................ 31–32, 38, 112, 114–116
Isothermal .................................................................... 32, 36–39, 131, 170

220

You might also like