You are on page 1of 166

Instructor’s Manual

Containing Solutions to
Over 280 Problems
Selected from

Statistical Mechanics
Third Edition

By
R. K. Pathria and Paul D. Beale

AMSTERDAM · BOSTON · HEIDELBERG · LONDON


NEW YORK · OXFORD · PARIS · SAN DIEGO
SAN FRANCISCO · SINGAPORE · SYDNEY · TOKYO
Academic Press is an imprint of Elsevier
Academic Press is an imprint of Elsevier
225 Wyman Street, Waltham, MA 02451, USA
The Boulevard, Langford Lane, Kidlington, Oxford, OX5 1GB, UK


c 2011 Elsevier Ltd. All rights reserved.

No part of this publication may be reproduced or transmitted in any form or by


any means, electronic or mechanical, including photocopying, recording, or any
information storage and retrieval system, without permission in writing from
the publisher. Details on how to seek permission, further information about the
Publisher’s permissions policies and our arrangements with organizations such
as the Copyright Clearance Center and the Copyright Licensing Agency, can be
found at our website: www.elsevier.com/permissions

This book and the individual contributions contained in it are protected under
copyright by the Publisher (other than as may be noted herein).

Notices

Knowledge and best practice in this field are constantly changing. As new re-
search and experience broaden our understanding, changes in research methods,
professional practices, or medical treatment may become necessary.

Practitioners and researchers must always rely on their own experience and
knowledge in evaluating and using any information, methods, compounds, or
experiments described herein. In using such information or methods they should
be mindful of their own safety and the safety of others, including parties for
whom they have a professional responsibility.

To the fullest extent of the law, neither the Publisher nor the authors, contrib-
utors, or editors, assume any liability for any injury and/or damage to persons
or property as a matter of products liability, negligence or otherwise, or from
any use or operation of any methods, products, instructions, or ideas contained
in the material herein.

ISBN: 978-0-12-416010-1

For information on all Academic Press publications, visit our website: www.elsevierdirect.com
Preface

This instructor’s manual for the third edition of Statistical Mechanics is based
on RKP’s instructor’s manual for the second edition. Most of the solutions
here were retypeset into TeX from that manual. PDB is responsible for the
solutions of the new problems added in the third edition. The result is a manual
containing solutions to some 280 problems selected from the third edition.
The original idea of producing an instructor’s manual first came from RKP’s
friend and colleague Wing-Ki Liu in the 1990’s when RKP had just embarked
on the task of preparing the second edition of Statistical Mechanics.
This should provide several benefits to the statistical mechanics instructor.
First of all, there is the obvious advantage of saving time that one would oth-
erwise spend on solving these problems oneself. Secondly, before one selects
problems either for homework or for an exam, one can consult the manual to
determine the level of difficulty of the various problems and make one’s selection
accordingly. Thirdly, one may even use some of these solved problems, especially
the ones appearing in later chapters, as lecture material, thereby supplementing
the text. We hope that this manual will enhance the usefulness of the text –
both for the instructors and (indirectly) for the students.
We implore that instructors not share copies of any of the material in this
manual with students or post any part of this manual on the web. Students
learn best when they work together and struggle over difficult problems. Readily
available solutions interfere with this crucial aspect of graduate physics training.

R.K.P. San Diego, CA


P.D.B. Boulder, CO

iii
Chapter 1

1.1. (a) We expand the quantity ln Ω(0) (E1 ) as a Taylor series in the variable
(E1 − Ē1 ) and get

ln Ω(0) (E1 ) ≡ lnΩ1 (E1 ) + ln Ω2 (E2 ) (E2 = E (0) − E1 )


= {ln Ω1 (Ē1 ) + ln Ω2 (Ē2 )}+
 
∂ ln Ω1 (E1 ) ∂ ln Ω2 (E2 ) ∂E2
+ (E1 − Ē1 )+
∂E1 ∂E2 ∂E1 E1 =Ē1
( 2 )
1 ∂ 2 ln Ω1 (E1 ) ∂ 2 ln Ω2 (E2 ) ∂E2

+ (E1 − Ē1 )2 + · · · .
2 ∂E12 ∂E22 ∂E1
E1 =Ē1

The first term of this expansion is a constant, the second term van-
ishes as a result of equilibrium (β1 = β2 ), while the third term may
be written as
   
1 ∂β1 ∂B2 2 1 1 1
+ E1 − Ē1 = − + (E1 −Ē1 )2 ,
2 ∂E1 ∂E2 eq. 2 kT12 (Cv )1 kT22 (Cv )2

with T1 = T2 . Ignoring the subsequent terms (which is justified if the


systems involved are large) and taking the exponentials, we readily
see that the function Ω0 (E1 ) is a Gaussian in the variable (E1 − Ē1 ),
with variance kT 2 (Cv )1 (Cv )2 /{(Cv )1 + (Cv )2 }. Note that if (Cv )2 
(Cv )1 — corresponding to system 1 being in thermal contact with a
very large reservoir — then the variance becomes simply kT 2 (Cv )1 ,
regardless of the nature of the reservoir; cf. eqn. (3.6.3).
(b) If the systems involved are ideal classical gases, then (Cv )1 = 23 N1 k
and (Cv )2 = 32 N2 k; the variance then becomes 32 k 2 T 2 · N1 N2 /(N1 +
N2 ). Again, if N2  N1 , we obtain the simplified expression 32 N1 k 2 T 2 ;
cf. Problem 3.18.
1.2. Since S is additive and Ω multiplicative, the function f (Ω) must satisfy
the condition
f (Ω1 Ω2 ) = f (Ω1 ) + f (Ω2 ). (1)

1
2

Differentiating (1) with respect to Ω1 (and with respect to Ω2 ), we get

Ω2 f 0 (Ω1 Ω2 ) = f 0 (Ω1 ) and Ω1 f 0 (Ω1 Ω2 ) = f 0 (Ω2 ),

so that
Ω1 f 0 (Ω1 ) = Ω2 f 0 (Ω2 ). (2)
Since the left-hand side of (2) is independent of Ω2 and the right-hand side
is independent of Ω1 , each side must be equal to a constant, k, independent
of both Ω1 and Ω2 . It follows that f 0 (Ω) = k/Ω and hence

f (Ω) = k ln Ω + const. (3)

Substituting (3) into (1), we find that the constant of integration is zero.
1.4. Instead of eqn. (1.4.1), we now have

Ω ∝ V (V − v0 )(V − 2v0 ) . . . (V − N − 1v0 ),

so that

ln Ω = C + ln V + ln (V − v0 ) + ln (V − 2v0 ) + . . . + ln (V − N − 1v0 ),

where C is independent of V . The expression on the right may be written


as
N −1 N −1 
N 2 v0
  
X jv0 X jv0
C+N ln V + ln 1 − ' C+N ln V + − ' C+N ln V − .
j=1
V j=1
V 2V

Equation (1.4.2) is then replaced by

N 2 v0
 
P N N N v0
= + = 1 + , i.e.
kT V 2V 2 V 2V
 −1
N v0
PV 1 + = NkT .
2V

Since N v0  V, (1 + N v0 /2V )−1 ' 1 − N v0 /2V . Our last result then


takes the form: P (V − b) = NkT , where b = 21 N v0 .
A little reflection shows that v0 = (4π/3)σ 3 , with the result that
 3
1 4π 3 4π 1
b= N· σ = 4N · σ .
2 3 3 2

1.5. This problem is essentially solved in Appendix A; all that remains to be


done is to substitute from eqn. (B.12) into (B.11), to get

(πε∗1/2 /L)3 (πε∗1/2 /L)2


Σ1 (ε∗ ) = V ∓ S.
6π 2 16π
3

Substituting V = L3 and S = 6L2 , we obtain eqns. (1.4.15 and 16).


The expression for T now follows straightforwardly; we get
       
1 ∂ ln Ω k ∂ ln Ω k R+N k Nhν
=k = = ln = ln 1 + ,
T ∂E N hν ∂R N hν R hν E
so that   
hν Nhν
T = ln 1 + .
k E
For E  Nhν, we recover the classical result: T = E/Nk .
1.9. Since the function S(N,V,E) of a given thermodynamic system is an ex-
tensive quantity, we may write
   
V E V E
S(N, V, E) = Nf , = Nf (v, ε) v = ,ε = .
N N N N
It follows that
       
∂S ∂f −V ∂f −E
N =N f +N · 2 +N · 2 ,
∂N V,E ∂v ε N ∂ε v N
       
∂S ∂f 1 ∂S. ∂f 1
V = VN · and E = EN · .
∂V N,E ∂v ε N ∂E N,V ∂ε v N
Adding these expressions, we obtain the desired result.
1.11. Clearly, the initial temperatures and the initial particle densities of the two
gases (and hence of the mixture) are the same. The entropy of mixing may,
therefore, be obtained from eqn. (1.5.4), with N1 = 4NA and N2 = NA .
We get
(∆S)∗ = k[4NA ln(5/4) + NA ln 5]
= R[4 ln(5/4) + ln 5] = 2.502 R,
which is equivalent to about 0.5 R per mole of the mixture.
1.12. (a) The expression in question is given by eqn. (1.5.3a). Without loss of
generality, we may keep N1 , N2 and V1 fixed and vary only V2 . The
first and second derivatives of this expression are then given by
   
N1 + N2 N2 N1 + N2 N2
k − and k − + 2 (1a,b)
V1 + V2 V2 (V1 + V2 )2 V2
respectively. Equating (1a) to zero gives the desired condition, viz.
N1 V2 = N2 V1 , i.e. N1 /V1 = N2 /V2 = n, say. Expression (1b) then
reduces to
 
n n knV1
k − + = > 0.
V1 + V2 V2 V2 (V1 + V2 )
Clearly, (∆S)1≡2 is at its minimum when N1 /V1 = N2 /V2 , and it is
straightforward to check that the value at the minimum is zero.
4

(b) The expression now in question is given by eqn. (1.5.4). With N1 =


αN and N2 = (1 − α)N , where N = N1 + N2 (which is fixed), the
expression for (∆S)∗ /k takes the form

−αN ln α − (1 − α)N ln (1 − α).

The first and second derivatives of this expression with respect to α are
 
N N
[−N ln α + N ln(1 − α)] and − − (2a,b)
α 1−α

respectively. Equating (2a) to zero gives the condition α = 1/2, which


reduces (2b) to −4N . Clearly, (∆S)∗ /k is at its maximum when N1 =
N2 = (1/2)N , and it is straightforward to check that the value at the
maximum is N ln 2.
1.13. Proceeding with eqn. (1.5.1), with T replaced by Ti , it is straightforward
to see that the extra contribution to ∆S, owing to the fact that T1 6= T2 ,
is given by the expression
3 3
N1 k ln (Tf /T1 ) + N2 k ln(Tf /T2 ),
2 2
where Tf = (N1 T1 + N2 T2 )/(N1 + N2 ). It is worth checking that this
expression is always greater than or equal to zero, the equality holding if
and only if T1 = T2 . Furthermore, the result quoted here does not depend
on whether the two gases were different or identical.
1.14. By eqn. (1.5.1a), given on page 19 of the text, we get
3
(∆S)v = Nk ln(Tf /Ti ).
2
Now, since PV = NkT , the same equation may also be written as
    
kT 3 5 2πmkT
S = Nk ln + Nk + ln . (1b)
P 2 3 h2

It follows that
5 5
(∆S)P = Nk ln(Tf / Ti ) = (∆S)V .
2 3
A numerical verification of this result is straightforward.
It should be noted that, quite generally,

(∆S)P T (∂S / ∂T )P CP
= = =γ
(∆S)V T (∂S / ∂T )V CV

which, in the present case, happens to be 5/3.


5

1.15. For an ideal gas, CP − CV = nR, where n is the number of moles of the
gas. With CP /CV = γ, one gets

CP = γnR / (γ − 1) and CV = nR / (γ − 1).


For a mixture of two ideal gases,
 
n1 R n2 R f1 f2
CV = + = + (n1 + n2 )R.
γ1 − 1 γ2 − 1 γ1 − 1 γ2 − 1

Equating this to the conventional expression (n1 + n2 )R/(γ − 1), we get


the desired result.
1.16. In view of eqn. (1.3.15), E − TS + PV = µN . It follows that

dE − TdS − SdT + PdV + VdP = µdN + Nd µ.


Combining this with eqn. (1.3.4), we get

−SdT + VdP = Nd µ, i.e. dP = (N / V )dµ + (S / V )dT .

Clearly, then,

(∂P / ∂µ)T = N / V and (∂P / ∂T )µ = S / V.

Now, for the ideal gas


( 3/2 )
h2

NkT N
P = and µ = kT ln ;
V V 2πmkT

see eqn. (1.5.7). Eliminating (N/V ), we get


 3/2
2πmkT
P = kT eµ/kT ,
h2

which is the desired expression. It follows quite readily now that for this
system  
∂P 1
= P.
∂µ T kT
which is indeed equal to N/V , whereas
" (  3/2 )#
h2
 
∂P 5 µ 5 N Nk
= P− 2 P = 2 − ln
∂T µ 2T kT V 2πmkT V

which, by eqn. (1.5.1a), is precisely equal to S/V .


Chapter 2

2.3. The rotator in this problem may be regarded as confined to the (z = 0)-
plane and its position at time t may be denoted by the azimuthal angle
ϕ. The conjugate variable pϕ is then mρ2 ϕ̇, where the various symbols
have their usual meanings. The energy of rotation is given by
1
E= m(ρϕ̇)2 = p2ϕ / 2mρ2 .
2

Lines of constant energy in the (ϕ, pϕ )-plane are “straight lines, running
parallel to the ϕ-axis from ϕ = 0 to ϕ = 2π”. The basic cell of area h in
this plane is a “rectangle with sides ∆ϕ = 2π and ∆pϕ = h/2π”. Clearly,
the eigenvalues of pϕ , starting with pϕ = 0, are n~ and those of E are
n2 ~2 /2I, where I = mρ2 and n = 0, ±1, ±2, . . .
The eigenvalues of E obtained here are precisely the ones given by quan-
tum mechanics for the energy “associated with the z-component of the
rotational motion”.
2.4. The rigid rotator is a model for a diatomic molecule whose internuclear
distance r may be regarded as fixed. The orientation of the molecule in

6
7

space may be denoted by the angles θ and ϕ, the conjugate variables being
pθ = mr 2 θ̇ and pϕ = mr 2 sin2 θϕ̇. The energy of rotation is given by

1 1 p2θ p2ϕ M2
E= m(rθ̇)2 + m(r sin θϕ̇)2 = 2
+ 2 2 = ,
2 2 2mr 2mr sin θ 2I

where I = mr 2 and M 2 = p2θ + p2ϕ / sin2 θ .




The “volume”
R0 of the relevant region of the phase space is given by the
integral dp θ dp ϕ dθ dϕ, where the region of integration is constrained
by the value of M . A little reflection shows that in the subspace of pθ
and pϕ we are restricted by an elliptical boundary with semi-axes M and
M sin θ, the enclosed area being πM 2 sin θ. The “volume” of the relevant
region, therefore, is

Zπ Z2π
(πM 2 sin θ)dθ dϕ = 4π 2 M 2 .
θ=0 ϕ=0

The number of microstates available to the rotator is then given by 4π 2 M 2 /h2 ,


which is precisely (M/~)2 . At the same time, the number of microstates
associated with the quantized value Mj2 = j(j + 1)~2 may be estimated as

1 h 2 i  1

3
 
1

1

2
Mj+ 1 − Mj− 1 = j + j+ − j− j+ = 2j + 1.
~2 2 2 2 2 2 2

This is precisely the degeneracy arising from the eigenvalues that the az-
imuthal quantum number m has, viz. j, j − 1, . . . , −j + 1, −j.
2.6. In terms of the variables θ and L(= m`2 θ), the state of the simple pendu-
lum is given by, see eqns. (2.4.9),

θ = (A/`) cos(ωt + ϕ), L = −m`ωA sin(ωt + ϕ),


with E = 21 mω 2 A2 and τ = 2π/ω. The trajectory in the (θ, L)-plane is
given by the equation

θ2 L2
2
+ = 1,
(A/`) (m` ωA)2

which is an ellipse — just like in Fig. 2.2. The enclosed area turns out to
be πmωA2 , which is precisely equal to the product Eτ .
2.7. Following the argument developed on page 68–69 of the text, the number
of microstates for a given energy E turns out to be

 
1
Ω(E) = (R + N − 1)!/ R!(N − 1)!, R= E − N ~ω /~ω. (1)
2
8

For R  N , we obtain the asymptotic result

Ω(E) ≈ RN −1 / (N − 1)!, where R ≈ E / ~ω. (3.8.25a)

The corresponding expression for Γ(E; ∆) would be

(E / ~ω)N −1 ∆ E N −1 ∆
Γ(E; ∆) ≈ · = . (1)
(N − 1)! ~ω (N − 1)!(~ω)N

The “volume” of the relevant region of the phase space may be derived
from the integral
N
Z 0Y N  
X 1 1 2
(dq i dp i ), with kq 2i + p ≤ E.
i=1 i=1
2 2m i

This is equal to, see eqn. (7a) of Appendix C,


  12 N  N N
2 1 πN N 2π E
(2m) 2 N · E = ,
k N! ω N!
p
where ω = k/m. The “volume” of the shell in question is then given by
N N
NE N −1 E N −1 ∆
 
2π 2π
·∆= . (2)
ω N! ω (N − 1)!

Dividing (2) by (1), we see that the conversion factor ω0 is precisely hN .

2.8. We write V3N = AR 3N , so that dV 3N = A · 3NR 3N −1 dR. At the same


time, we have
Z∞ Z∞ N
P N N Z ∞
− ri Y Y
... e i=1 ri2 dr i = e−ri ri2 dr i = 2N . (1)
0 0 i=1 i=1 0

The integral on the left may be written as


Z∞ Z∞
e −R −N
(4π) dV 3N = e−R (4π)−N A·3NR 3N −1 dR = (4π)−N A·3N Γ(3N ).
0 0
(2)
Equating (1) and (2), we get: A = (8π)N /(3N )!, which yields the desired
result for V3N .
The “volume” of the relevant region of the phase space is given by
Z 3N
0 Y Z 0 N
Y
dq i dp i = V N 4πp2i dp i = V N (8π E 3 / c3 )N / (3N )!,

i=1 i=1
9

so that
Σ(n, V, E) = V N (8π E 3 / h3 c3 )N / (3N )!,
which is a function of N and VE 3 . An isentropic process then implies
that VE 3 = const.
The temperature of the system is given by
 
1 ∂(k ln Σ) 3Nk
= = , i.e. E = 3NkT .
T ∂E N,V E

The equation for the isentropic process then becomes VT 3 = const., i.e.
T ∝ V −1/3 ; this implies that γ = 4/3. The rest of the thermodynamics
follows straightforwardly. See also Problems 1.7 and 3.15.
Chapter 3

3.4. For the first part, we use eqn. (3.2.31) with all ωr = 1. We get
( )
k X
−βEr
ln Γ = k ln e + kβU,
N r

which is indeed equal to −(A/T ) + (U/T ) = S.


For the second part, we use eqn. (3.2.5), with the result that
" #
k k X

ln W {nr } = N ln N − ∗
nr ln nr ∗
N N r
X n∗ n∗

n∗

r
= −k ln r = −k ln r .
r
N N N

Substituting for n∗r from eqn. (3.2.10), we get


( )
k ∗
X
−βEr
ln W {nr } = kβhEr i + k ln e ,
N r

which is precisely the result obtained in the first part.

3.5. Since the function A(N, V, T ) of a given thermodynamic system is an


extensive quantity, we may write

A(N, V, T ) = Nf (v, T ) (v = V / N ).

It follows that
         
∂A ∂f −V ∂A ∂f 1
N =N f +N · 2 , and V = VN · .
∂N V,T ∂v T N ∂V N,T ∂v T N

Adding these expressions, we obtain the desired result.

3.6. Let’s go to part (c) P


right away. Our problem here is to maximize
P the
expression S/k = − Pr,s ln Pr,s , subject to the constraints Pr,s =
r,s r,s

10
11

P P
1, Es Pr,s = E and Nr Pr,s = N . Varying P ’s and using the method
r,s r,s
of Lagrange’s undetermined multipliers, we are led to the condition
X
{−(1 + ln Pr,s ) − γ − βEs − αNr } δPr,s = 0.
r,s

In view of the arbitrariness of the δP ’s in this expression, we require that

−(1 + ln Pr,s ) − γ − βEs − αNr = 0

for all r and s. It follows that

Pr,s ∝ exp(−βEs − αNr ).

The parameters α and β are to be determined by the given values of N̄


and Ē.
In the absence of the constraint imposed by N̄ , the parameter α does not
even figure in the calculation, and we obtain

Pr ∝ exp(−βEr ),

as desired in part (b). And if the constraint imposed by Ē is also absent,


we obtain
Pr = const.,
as desired in part (a).
3.7. From thermodynamics,

     2  
∂P ∂V ∂P ∂P
CP − CV = T = −T > 0. (1)
∂T V ∂T P ∂T V ∂V T

From Sec. 3.3,


   
∂A ∂ ln Q
P =− = kT . (2)
∂V N,T ∂V N,T

Substituting (2) into (1), we obtain the desired result.


For the ideal gas, Q ∝ V N T 3N/2 . Therefore, (∂ ln Q/∂V )T = N/V . We
then get
(N/V )2
CP − CV = −k = Nk .
−N/V 2

3.8. For an ideal gas,

(2πmkT )3/2 NkT (2πmkT )3/2


Q1 = V = .
h3 P h3
12

It follows that T (∂ ln Q1 /∂T )P = 5/2; the expression on the right-hand


side of the given equation then is
V (2πmkT )3/2
 
5
ln +
N h3 2
which, by eqn. (3.5.13), is indeed equal to the quantity S/Nk.
P 2 
3.12. We start with eqn. (3.5.5), substitute H(q,p) = pi /2m + U (q) and
i
integrate over the pi 0 s, to get
 3N/2 Z
1 2πmkT
QN (V, T ) = Z N (V, T ), where Z N (V, T ) = e−U (q)/kT d3N q.
N! h2
It follows that, for N  1,
" (  3/2 ) #
h2
A = NkT ln N − 1 − kT ln Z, whence
2πmkT
" (  3/2 ) #  
1 2πmkT 5 ∂ ln Z
S = Nk ln + + k ln Z + kT .
N h2 2 ∂T N,V

Now
e−U/kT (U/kT 2 )d3N q
  R
∂ ln Z kT Ū
kT = R
−U/kT d3N q
= , while
∂T N,V e T
n o Ū
k ln Z = k ln V̄ N e−Ū /kT = Nk ln V̄ − .
T
Substituting these results into the above expression for S, we obtain the
desired result for S. In passing, we note that hHi ≡ A + TS = 23 NkT + Ū .
P
For the second part of the question, we write U (q) = u(rij ), so that
i<j
Y Y
e−βU (q) = e−βu(rij ) = (1 + fij ) ,
i<j i<j

and follow Problems 3.23 and 1.4. The quantity V̄ then appears to be in
the nature of a “free volume” for the molecules of the system.
3.14. a) The Lagrangian is given by
X1 X X
L =K −V = 2
mṙiα − u(rij ) − [uw (riα ) + uw (L − riα )],

2 i<j iα

where i = 1, · · · , N denotes thePparticle number, α = x, y, z denotes the


2
cartesian directions, and rij = α (riα − rjα )2 . The canonical momenta
are
∂L
piα = = mṙiα .
∂ ṙiα
13

The Hamiltonian is given by


X
H = piα ṙiα − L

X p2 X X

= + u(rij ) + [uw (riα ) + uw (L − riα )].

2m i<j iα

The canonical pressure can be written


∂H 1 ∂H 1 X 0 1
P =− =− 2 =− 2 u (L − riα ) = (Fx + Fy + Fz ) .
∂V 3L ∂L 3L iα w 3L2

This is clearly the instantaneous force per unit area on the right, back,
and top walls.
b) The cartesian coordinates for the scaled position inside the box are
siα = riα /L so the Lagrangian becomes
X1 X X
L = mL2 ṡ2iα − u(Lsij ) − [uw (Lsiα ) + uw (L − Lsiα )].

2 i<j iα

In this case the canonical momenta are


∂L
p̃iα = = mL2 ṡiα .
∂ ṡiα
This leads to a Hamiltonian of the form
X p̃2 X X
H = iα
+ u(Ls ij ) + [uw (Lsiα ) + uw (L − Lsiα )],

2mL2 i<j iα

with canonical pressure is


1 X p̃2iα
P =+
3L2 iα mL3
1 X 0
− u (Lsij )sij
3L2 i<j
1 X 0
− [u (Lsiα )siα + u0w (L − Lsiα )(1 − siα )].
3L2 iα w

Converting back to normal cartesian coordinates and momenta gives


2 X p2iα
P =+
3V iα 2m
1 X 0
− u (rij )rij
3V i<j
1 X 0
− [u (riα )riα + u0w (L − riα )(L − riα )].
3V iα w
14

The first term is (2/3)(N/V ) times the kinetic energy per particle so is
O(N ). The second term is (1/3)(N/V ) times the virial per particle so is
also O(N ). On the other hand, third term is proportional to the force on
the walls divided by the volume so is O(N 2/3 ) which is negligible in the
thermodynamic limit.
Comparing to equation (3.7.15) for the average pressure we see that
* +
P 1 X
=1− u0 (rij )rij .
nkT 3N kT i<j w

3.15. Here, QN (V, T ) = (1/N !)QN


1 (V, T ), while

Z∞
V · 4πp2 dp 8πV 1
Q1 (V, T ) = e−βpc = 3 3 3,
h3 h β c
0

which yields the desired result for QN . The thermodynamics of the system
now follows straightforwardly.
As regards the density of states, the expression
Z∞
8πV 1
Q1 (V, T ) = e−βε g(ε)dε =
h3 β 3 c3
0

leads to
4πV 2
g(ε) = ε
h3 c3
for a single particle, while the expression for QN (V, T ) leads to
N
E 3N −1

1 8πV
g(E) =
N! h3 c3 Γ(3N )

for the N -particle system; cf. the expression for Σ(E) derived in Prob-
lem 2.8.
3.17. Differentiate the stated result with respect to β, to get
Z  
∂U
− H(U − H) e−βH dω = 0.
∂β

This means that  


∂U
− HU + H 2 = 0,
∂β
which amounts to the desired result: hH 2 i − hHi2 = −(∂U/∂β).
15

3.18. We start with eqn. (3.6.2), viz.


P 2 −βEr
Er e
∂U r
= − P −βEr + U 2 , (1)
∂β e
r

and differentiate it with respect to β, keeping the Er fixed. We get

∂2U ∂U
= hE 3 i − hE 2 ihEi + 2U .
∂β 2 ∂β

Substituting for (∂U/∂β) from eqn. (1), we get

∂2U
= hE 3 i − 3hE 2 iU + 2U 3 ,
∂β 2

which is precisely equal to h(E −U )3 i. As for ∂ 2 U/∂β 2 , we note that, since


   
∂U ∂U
= −kT 2 = −kT 2 CV ,
∂β Er ∂T V
 2       
∂ U 2 ∂ 2
 2 2 2 ∂CV
= −kT −kT CV = k T 2TC V + T .
∂β 2 Er ∂T V ∂T V

Hence the desired result.


For the ideal classical gas, U = 23 NkT and CV = 32 Nk , which readily yield
the stated results.
P P
3.19. Since G = qi pi , Ġ = (q̇i pi + qi ṗi ). Averaging over a time interval τ ,
i i
we get

t+τ t+τ
G(t + τ ) − G(t)
Z Z
1 X 1
(q̇i pi + qi ṗi )dt = Ġdt = . (1)
τ i
τ τ
t t

For a finite V and finite E, the quantity G is bounded ; therefore, in the


limit τ → ∞, the right-hand side of (1) vanishes. The left-hand side
then gives * +
X
(q̇i pi + qi ṗi ) = 0.
i

which leads to the desired result.


3.20. The virial of the noninteracting system, by eqn. (3.7.12), is −3PV . The
contribution from interparticle interactions, by eqn. (3.7.15), is given by
the “expectation value of the sum of the quantity −r(∂u/∂r) over all pairs
of particles in the system”. If u(r) is a homogeneous function (of degree
n) of the particle coordinates, this contribution will be −nU , where U is
16

the mean potential energy (not the internal energy) of the system. The
total virial is then given by

V = −3PV − nU .

The relation K = − 12 V still holds, and the rest of the results follow
straightforwardly.
3.21. All systems considered here are localized. The pressure term, therefore,
drops out, and we are left with the result

n n
K= U= E.
2 n+2
Example (a) pertains to n = 2, while examples (b) and (c) pertain to
n = −1. In the former case, K = U = 12 E; in the latter, K = − 21 U = −E.
The next problem pertains to n = 4.
3.22. Note that a force proportional to q 3 implies a potential energy proportional
to q 4 . Thus

1 2
H= p + cq 4 (c > 0).
2m
It follows that
R∞ 2
  e−βp /2m
(p2 /2m)dp
1 2 −∞ 1
p = R∞ = ;
2m 2β
e−βp2 /2m dp
−∞

for the values of these integrals, see eqns. (13a) of Appendix B. Next,
R∞ 4
e−βcq (cq 4 )dq
−∞ ∂
hcq 4 i = R∞ =− ln I(B),
∂β
e−βcq 4 dq
−∞

where I(β) denotes the integral in the denominator. It is straightforward


to see that I(β) is proportional to β −1/4 , whence hcq 4 i = 1/4β, which
proves the desired result.

3.23. The key to this derivation is writing the partition function in terms of
position integrals over scaled coordinates. Assume a cubic box of size L
and volume V = L3 . The scaled position for particle i is si = r i /L. The
17

partition function is
 
Z
1 X
QN (V, T ) = exp −β u(rij ) dN r
λ3N N ! i<j
 
N Z
V X
= exp −β u(V 1/d sij ) dN s.
λ3N N ! i<j

Now the pressure is


 
∂A
P =−
∂V N,T
     
N Z X
kT  N QN βV X
= −  V 1/d sij u0 (V 1/d sij ) exp −β u(V 1/d sij ) dN s .
QN V dV λ3N N ! i<j i<j

This can be simplified by going back to integrals over the normal position
variables to give equation (3.7.15).
3.24. By eqn. (3.7.5), we have, for a single particle,
* 3 +
X
pi q̇i = 3kT . (1)
i=1

The left-hand side of (1) is the expectation value of the quantity p · u, i.e.
pu which, for a relativistic particle, is equal to m0 u2 (1 − u2 /c2 )−1/2 . The
desired result follows readily.
In the non-relativistic limit (u  c), one obtains: 12 m0 u2 ≈ 32 kT ; in the

extreme relativistic limit (u → c), one obtains: hmc 2 i ≈ 3kT . Note that,
in the latter case, m0 c2 is negligible in comparison with mc 2 , so there is
no significant difference between the kinetic energy and the total energy
of the particle.
3.25. For the first part of this problem, see Sec. 6.4 — especially the derivation
of the formula (6.4.9). For the second part, equate the result obtained in
the first part with the one stated in eqn. (3.7.5).
3.26. The multiplicity w(j){= (j + s − 1)!/j!(s − 1)!} arises from the variety
of ways in which j indistinguishable quanta can be divided among the
s dimensions of the oscillator: j = j1 + . . . + js ; this is similar to the
calculation done on page 68–69 of the text.
h iN
(s) (s)
As for the partition function, QN (β) = Q1 (β) , where

(j + s − 1)!
e−β (j+ 2 s)~ω
(s)
X 1
Q1 (β) =
j=0
j!(s − 1)!
1
= e− 2 sβ~ω (1 − e−β~ω )−s .
18

Calculation of the various thermodynamic quantities is now straightfor-


ward. The results are found to be essentially the same as for a system of
sN one-dimensional oscillators. However, since
(s) (1)
QN (β) = QNs (β),

the chemical potential µs will turn out to be s times µ1 .


3.28. (a) When one of the oscillators is in the quantum state n, the energy left
for the remaining (N − 1) oscillators is E − n + 21 ~ω; the corre-
sponding number of quanta to be distributed among these oscillators
is R−n; see eqn. (3.8.24). The relevant number of microstates is then
given by the expression (R − n + N − 2)!/(R − n)!(N − 2)!. Combined
with expression (3.8.25), this gives

(R − n + N − 2)! (R + N − 1)!
pn = ÷ . (1)
(R − n)!(N − 2)! R!(N − 1)!
It follows that
pn+1 R−n R n̄
= ' = .
pn R−n+N −2 R+N n̄ + 1
By iteration, pn = p0 {n̄/(n̄ + 1)}n .
Going back to eqn. (1), we note that
N −1 N 1
p0 = ' =− ,
R+N −1 R+N n̄ + 1
which completes the desired calculation.
(b) The probability in question is proportional to gN −1 (E − ε), i.e. to
3
(E − ε) 2 (N −1)−1 . For 1  N , this is essentially proportional to
3
(1 − ε/E) 2 N and, for ε  −E, to e−3N ε/2E .
3.29. The partition function of the anharmonic oscillator is given by
Z∞ Z∞
p2
 
1 −βH 2 3 4
Q1 (β) = e dq dq H= + cq − gq − fq .
h 2m
−∞ −∞
p
The integration over p gives a factor of 2πm/β. For integration over q,
we write
 
−βcq 2 β(gq 3 +fq 4 ) −βcq 2 3 4 1 2 3 4 2
e e =e 1 + β(gq + fq ) + β (gq + fq ) + . . . ;
2
the integration then gives
r r r
π 3 π 1 2 2 15 π
+ βf · + β g · + ....
βc 4 β 5 c5 2 8 β 7 c7
19

It follows that
r
15g 2
 
π 2m 3f
Q1 (β) = 1+ + + ... ,
βh c 4βc2 16βc3

so that
3f 15g 2
ln Q1 (β) = const. − ln β + + + ...,
4βc2 16βc3
whence
1 3f 15g 2
U (β) = + 2 2+ + ...
β 4β c 16β 2 c3
and
3f k 2 T 15g 2 k 2 T
C(β) = k + + + ....
2c2 8c3
Next, the mean value of the displacement q is given by
Z ∞ Z ∞ Z ∞ Z ∞
hqi = exp(−βH)q dp dq / exp(−βH)dpdq.
−∞ −∞ −∞ −∞

In the desired approximation, we get


Z∞ , Z∞
−βcq 2 4 2
hqi ' βg e q dq e−βcq dq
−∞ −∞
r r
3 π π 3g
= βg · / = .
4 β 5 c5 βc 4βc2

3.30. The single-oscillator partition function is now given by


2
e−β (n+ 2 )~ω+βx(n+ 2 )
X 1 1

Q1 (β) = .
n=0

For x  1, we may write



"  2 #
X
−β (n+ 12 )~ω 1
Q1 (β) = e 1 + βx n + ~ω + . . . .
n=0
2

With u = β~ω, the sums involved are


∞   −1
X
−u(n+ 21 ) 1
S1 (u) = e = 2 sinh u , and
n=0
2
∞ 2  −1 
d2
    
1 1 1 1
e−u(n+ 2 ) n +
X 1
2
S2 (u) = = S1 (u) = 4 sinh u coth u − .
n=0
2 du 2 2 2 2
20

It follows that
ln Q1 = ln[S1 + xuS2 + . . .] ' ln S1 + xu(S2 /S1 )
      
1 1 1 1
= − ln 2 sinh u + xu coth2 u − .
2 2 2 2
The first part of this expression leads to the standard results (3.8.20 and
21). The second part may, for simplification, be expressed as a power
series in u, viz.
u3
 
2 u
x + + + ... .
u 12 120
The resulting contribution to the internal energy per oscillator turns out
to be
u2
 
2 1
x~ω − − − . . .
u2 12 40
and the corresponding contribution to the specific heat is given by
4 u3
 
xk + + ... .
u 20

3.31. This problem is essentially the same as Problem 3.32, with g1 = g2 =


1, ε1 = 0 and ε2 = ε.
3.32. We use formula (3.3.13), with Pr = p1 /g1 for each of the states in group
1 and p2 /g2 for each of the states in group 2. We get
    
p1 p1 p2 p2
S = −k g1 ln + g2 ln . (1)
g1 g1 g2 g2
(a) In thermal equilibrium,
gi e−βεi
pi = P −βεi (i = 1, 2).
gi e
i

With x = β(ε2 − ε1 ), we have: p1 = g1 /(g1 + g2 e−x ) and p2 =


g2 /(g1 ex + g2 ). Substituting these results into (1), we obtain
 
g1 −x
 g2 x
S=k ln g1 + g2 e + ln (g1 e + g2 ) .
g1 + g2 e−x g1 ex + g2
Writing the first log as ln g1 + ln{1 + (g2 /g1 )e−x } and the second log
as ln g1 + x + ln{1 + (g2 /g1 )e−x }, we obtain the stated expression
for S.
(b) With Q = g1 e−βε1 + g2 e−βε2 , it is straightforward to see that
A = −kT ln{g1 e−βε1 + g2 e−βε2 } and
U = {g1 ε1 e−βε1 + g2 ε2 e−βε2 }/{g1 e−βε1 + g2 e−βε2 }.
The formula S = (U − A)/T then leads to the desired result.
21

(c) As T → 0, x → ∞ and S indeed tends to the value k ln g1 . This


corresponds to the fact that the probabilities p1 and p2 in this limit
tend to the values 1 and 0, respectively.

3.35. The partition function of the system is given by


1 N V
QN = Q , where Q1 = 3 · Z,
N! 1 λ
Z being the factor that arises from the rotational/orientational degrees of
freedom of the molecule:
" ( )#
p2θ p2ϕ dp θ dp ϕ dθdϕ
Z
Z = exp −β + 2 − µE cos θ
2I 2I sin θ h2
Zπ  1/2  1/2
2πI 2πI sin2 θ cos θ 2πdθ
= eβµE
β β h2
0
I 2 sinh(βµE)
= · .
β~2 βµE
The study of the various thermodynamical quantities of the system is now
straightforward.
Concentrating on the electrical quantities alone, we obtain for the net
dipole moment of the system
 
N ∂ ln Z 1
Mz = N hµ cos θi = = N µ coth (βµE) − ;
β ∂E βµE
cf. eqns. (3.9.4 and 6). For βµE  1,
1
Mz ≈ N µ · βµE.
3
The polarization P , per unit volume, of the system is then given by

P ≈ nµ2 E / 3kT (n = N/V ),

and the dielectric constant ε by


E + 4πP 4πnµ2
ε= ≈1+ .
E 3kT
The numerical part of the problem is straightforward.
3.36. The mean force hFi between the two dipoles is given by
 R −βU
(−∂U / ∂R) sin θd θdϕ · sin θ0 dθ0 dϕ0

∂U e
hFi = − = R (1)
∂R e−βU sin θdθdϕ · sin θ0 dθ0 dϕ0
1 ∂
= ln Z . (2)
β ∂R
22

where Z denotes the integral in the denominator of (1). At high temper-


atures, we may write
Z  
1 2 2
Z= 1 − βU + β U − . . . sin θ dθdϕ · sin θ0 dθ0 dϕ0 .
2

The linear term vanishes on integration and we are left with

1 2 (µµ0 )2
Z  
0 0 0 2
Z= 1+ β {2 cos θ cos θ − sin θ sin θ cos(ϕ − ϕ )} − . . .
2 R6
sin θd θd ϕ · sin θ0 dθ0 dϕ0
1 2 (µµ0 )2
   
2 1 1 2 2 1
= 16π 1 + β 4· · −0+ · · − ... .
2 R6 3 3 3 3 2

It follows that
1 (µµ0 )2
ln Z = const. + β 2 − ...
3 R6
and hence, at high temperatures,

(µµ0 )2
hFi ≈ −2β R̂.
R7

3.37. By eqns. (3.9.17 and 18), we have, for a single dipole,

J
P
(gµB m) exp(βgµB mH )
m=−J
µ̄z = J
.
P
exp(βgµB mH )
m=−J

At high temperatures, the exponential may be approximated by (1 +


βgµB mH ) which yields, to the leading order in H,

µ̄z = βg 2 µ2B Hm2 .

One readily obtains for the Curie constant (per unit volume) of the system

CJ = N0 g 2 µ2B / k m2 .


Writing m = J cos θ, one obtains the desired result.


For the second part, we simply note that, for a given J,
J
m2
P
m=−J J(J + 1)(2J + 1)/3 1
m2 = = = J(J + 1).
2J + 1 2J + 1 3
23

3.38. Treating m as a continuous variable, the partition function of a magnetic


dipole assumes the form
Z J
2
Q1 (β) = eβgµB Hm dm = sin h (βgµB JH );
−J βgµB H

cf. eqn. (3.9.5). It is clear that this approximation will lead essentially to
the same results as the ones following from the Langevin theory — except
for the fact that the role of µ will be played by gµB J, which should be
contrasted with the expression (3.9.16) of the quantum theory.
3.40. (a) By definition, CH = T (∂S/∂T )H and CM = T (∂S/∂T )M . Now
       
∂S ∂S ∂S ∂M
= + ; (1)
∂T H ∂T M ∂M T ∂T H

at the same time, dA ≡ dU − TdS − SdT = HdM − SdT , with the


result that (∂H/∂T )M = −(∂S/∂M )T . Equation (1) then becomes
       
∂S ∂S ∂H ∂M
= − . (2)
∂T H ∂T M ∂T M ∂T H

Multiplying (2) by T , we obtain the desired result for CH − CM .


(b) The Curie law implies that M = CH /T . This means that (∂H/∂T )M =
H/T , while (∂M/∂T )H = −CH /T 2 . It follows that CH − CM =
CH 2 /T 2 .

3.42. Let N1 (N2 ) be the number of dipoles aligned parallel (opposite) to the
field. Then
N1 + N2 = N, while − N1 ε + N2 ε = E.
It follows that
1 1
N1 = (N − E/ε), N2 = (N + E/ε).
2 2
The number of microstates associated with this macrostate is given by
N!
Ω(N, E) =  1 1
2 (N − E/ε) ! 2 (N + E/ε) !

The entropy of the system is then given by the expression


     
1 E 1 E
S = k ln Ω ≈ k N ln N − N− ln N−
2 ε 2 ε
    
1 E 1 E
− N+ ln N+ ,
2 ε 2 ε

which is essentially the same as eqn. (3.10.9).


24

For the temperature of the system, we get


   
1 ∂S k N − E/ε
= = ln ,
T ∂E N 2ε N + E/ε

which agrees with eqn. (3.10.8).

3.43. The partition function of this system is given by the usual expression
(3.5.5), except for the fact that the Hamiltonian of the system is now a
function of the quantities pj + (ej / c)A(rj ), and not of the pj as such.
However, on integration over any component of pj , from −∞ to +∞, we

obtain the same standard factor 2πmkT — regardless of the value of
the corresponding component of A. The partition function is, therefore,
independent of the applied field and hence the net magnetization of the
system is zero.
P
3.44. The Shannon information for a single message is given by I1 = − r Pr ln Pr
where Pr is the a priori probability of message r from among all Ω pos-
sible messages. The maximum information is obtained from varying the
probabilities,
P using a Lagrange multiplier µ to maintain the normalization
P
r r = 1, and demanding the solution is stationary.
!
X X
0 = δI1 − µδ Pr = − δPr [ln pr − 1 − µ].
r r

This implies the Pr = const, i.e. all messages are equally likely. Therefore
Pr = 1/Ω , which gives I1 = ln Ω. Any other set of probabilities gives
smaller information per message.
Keeping to the general cases in which probabilities of individual mes-
sages messages do not need to be equal, consider a sequence of two mes-
sages. The a priori probability of message r followed by message r0 is
Prr0 = Pr Pr0 Grr0 . The quantity Grr0 is the correlation between the two
messages. A value of Grr0 greater than unit implies that the first message r
increases the probability of finding the second message r0 above
P Pr0 . The
two Pmessage probabilitiesP have the followingPproperties: r Prr = Pr
0 0

and r0 Prr = Pr , i.e.


0
r Pr Grr = 1 and
0
r 0 Pr Grr = 1. The infor-
0 0

mation contained in two messages is given by


X X
I2 = Prr0 ln Prr0 = Pr Pr0 Grr0 ln (Pr Pr0 Grr0 ).
rr 0 rr 0

Expanding the logarithm and using the above summation properties gives
 
X X 1
I2 = 2I1 − Pr Pr0 Grr0 ln Grr0 = 2I1 + Pr Pr0 Grr0 ln .
0 0
Grr0
rr rr
25

Now, using ln x ≤ x − 1 for all x > 0, we get


X
I2 ≤ 2I1 + Pr Pr0 [1 − Grr0 ] = 2I1 .
rr 0

The information contained in two correlated messages is reduced compared


to sum of the information contained in two uncorrelated messages. Anal-
ysis of the first 65536 digits of π results in an information per character
of I1 ≈ 2.3 = ln 10. That makes sense because the characters 0, · · · , 9 are
evenly distributed in the digital representation of π. Furthermore, since
the digits of π are uncorrelated, the information per pair of characters is
I2 ≈ 4.6 = 2I1 . Analysis of the first 15, 000 characters of A Chrismas
Carol by Charles Dickens gives I1 ≈ 3.08 ≈ ln 21.75. This value is rea-
sonable since most of the characters are lower case letters of the alphabet
and blanks. The nonuniformity of the distribution of letters reduces the
information below ln 27 . When analyzed two characters at a time, the
information is I2 ≈ 5.45 ≈ 2 ln 15.25. The strong correlations between
characters in English text reduces the information well below 2I1 .
Chapter 4

4.1. By eqns. (4.1.9), (4.3.10) and (4.1.8), we get


X
Pr,s ln Pr,s ≡ hln Pr,s i = −αN̄ − β Ē − βPV
r,s

= (µN̄ − U − PV )/kT . (1)

Since µN̄ = G = U + PV − TS , the right-hand side of (1) equals −S/k.


Hence the result.
4.2. According to the grand canonical ensemble theory,
( )
X
PV = kT ln z Nr QNr (V, T ) . (1)
Nr

Now, the largest term in the sum pertains to the value N ∗ , of Nr , which
is determined by the condition

{Nr ln z − ln QNr }Nr =N ∗ = 0.
∂Nr
By Sec. 3.3, this is equivalent to the statement: z = exp(µ∗ /kT ), where
µ∗ is the chemical potential of the given system in a canonical ensemble
(with N = N ∗ ). If we replace the sum in (1) by its largest term, we
would get
PV ≈ N ∗ µ∗ − A∗ = P ∗ V,
where P ∗ is the pressure of the system in the canonical ensemble (with
N = N ∗ ). How different would P be from P ∗ depends essentially on
how different the particle density n̄ is from n∗ — a question thoroughly
discussed in Sec. 4.5.
4.3. The probability distribution in question is the binomial distribution

N (0) !
 
N N (0) −N V V
P (N, V ) = p q p = (0) , q = 1 − (0) .
N !(N (0) − N )! V V

26
27

We note that
(0)
N
X (0)
P (N, V ) = (q + p)N = 1.
N =0

For part (i), we have


(0)
N
X (
N̄ = N P (N, V ) = N (0) p(q + p)N 0)−1
= N (0) p, while
N =0
(0)
N
X
N (N − 1) = N (N − 1)P (N, V ) = N (0) (N (0) − 1)p2 (q + p)N (0)−2 = N (0) (N (0) − 1)p2 .
N =0

It follows that

N 2 =N (N − 1) + N̄ = (N (0) p)2 − N (0) p2 + N (0) p, whence


(∆N )2 ≡ N 2 − N̄ 2 = N (0) p(1 − p), etc.

For part (ii), we shift the origin to N = N (0) p, write

N = N (0) p + x, N (0) − N = N (0) q − x

and examine the function

ln P (x) = ln N (0) !−ln(N (0) p+x)!−ln(N (0) q−x)!+(N (0) p+x) ln p+(N (0) q−x) ln q.

Since N (0) p and N (0) q are both  1, we apply Stirling’s formula, ln v ! ≈


v ln v − v, and get (after some reduction)
   
(0) x (0) x
ln P (x) ≈ −(N p + x) ln 1 + (0) − (N q − x) ln 1 − (0) .
N p N q

For x  N (0) p and N (0) q, we expand this expression in powers of x, with


the result that ln P (x) ≈ −x2 /2N (0) pq. It follows that the distribution
P (x), under the stated conditions, is a Gaussian, with (∆N )2 = N (0) pq.
For part (iii), we write

N (0) (N (0) − 1) . . . (N (0) − N + 1) N (0)


P (N ) = p (1 − p)N −N .
N!
Now, if p  1 and N  N (0) , we obtain the Poisson distribution

[N (0) ]N N −N (0) p (N̄ )N −N̄


P (N ) ≈ p e = e ,
N! N!

with (∆N )2 = N̄ .
28

4.4. For obvious reasons,

e−αNr e−βEs
P
X s z Nr QNr (V, T )
P (Nr ) = Pr,s = = . (1)
s
Q(α, β, V ) Q(z, V, T )

For an ideal classical gas, see Sec. 4.4,


N
N̄ λ3

1 V
z= , QN = , Q ≡ ePV /kT = eN̄ . (2)
V N! λ3

Substituting (2) into (1), we get

(N̄ )N −N̄
P (N ) = e ,
N!

which is a Poisson distribution, with (∆N )2 = N̄ .


We note that the variance of N , calculated from the general formula
(4.5.3), also turns out to be the same:
      
2
∂ N̄ ∂ N̄ ∂ zV V
(∆N ) = − =z =z = z 3 = N̄ .
∂α T,V ∂z T,V ∂z λ3 T,V λ

4.5. The first term on the right-hand side of (4.3.20) may be written as
  "      #
∂q ∂q ∂q ∂µ
kT = kT +
∂T z,V ∂T µ,V ∂µ T,V ∂T z,V
 
∂q N̄
= kT + kT · · k ln z (for µ = kT ln z ).
∂T µ,V kT

Equation (4.3.20) then reduces to


   
∂q ∂
S = kT + kq = k (Tq) .
∂T µ,V ∂T µ,V

Note that this result is directly related to the formula, see Problem 1.16,

d(PV ) = PdV + Nd µ + SdT ,

whence  

S= (PV ) .
∂T µ,V

4.6. The Gibbs free energy is

G(N, P, T ) = −kT ln (YN (P, T )) .


29

For example
  Z ∞
∂G 1
= V Q(N, V, T )e−βP V = hV i.
∂P N,T YN (P, T ) 0

The ideal gas gives


1
YN (P, T ) = N +1
,
(βP λ3 )
G(N, P, T ) ≈ N kT ln βP λ3 ,

 
∂G N kT
V = = .
∂P N,T P

4.10. The partition function of the adsorbed molecules, assumed noninteracting,


is given by

N0 !
QN (N0 , T ) = g(N )aN = aN [a = a(T )]. (1)
N !(N0 − N )!
Using Stirling’s formula (B.29), we get

ln QN ≈ N0 ln N0 − N ln N − (N0 − N ) ln(N0 − N ) + N ln a,

with the result that


∂ ln QN N
µ = −kT = kT ln . (2)
∂N (N0 − N )a

Alternatively, the grand partition function of the system consisting of all


N0 sites (of which some are empty while others are occupied by a single
molecule) is given by

Q(z, N0 , T ) = [Q(z, 1, T )]N0 = [1 + za(T )]N0 ; (3)

see eqn. (4.4.15), with Nr = 0 or 1. Note that expression (3) could also be
N0
obtained by using the standard definition Q(z, N0 , T ) = z N QN (N0 , T )
P
N =0
and employing expression (1) for QN . The mean value of N now turns
out to be
∂ za 1 N̄
N̄ = z ln Q = N0 , whence z = , (4)
∂z 1 + za a N0 − N̄
which agrees with (2).
4.11. By eqn. (4) of the preceding problem, the fraction θ of the adsorption sites
that are occupied is given by
N̄ za 1 θ
θ= = , whence z = . (1)
N0 1 + za a1−θ
30

Now, if the molecules in the adsorbed phase are in equilibrium with those
in the gaseous phase, then their fugacity z would be equal to the fugacity
zg of the gaseous phase. The latter is given by eqns. (4.4.5 and 29),
whereby
Pg h3
zg = . (2)
kT (2πmkT )3/2
Equating (1) and (2), we obtain the desired result

θ 1 (2πmkT )3/2
Pg = × kT .
1 − θ a(T ) h3

4.12. From eqn. (4.5.1), we get


 
∂ N̄
= −NE + N̄ Ē.
∂β α,V

The left-hand side here is equal to, see eqns. (4.5.3 and 12),
       
∂ N̄ ∂U ∂U ∂ N̄
−kT 2 = −kT = −kT
∂T z,V ∂µ T,V ∂ N̄ T,V ∂µ T,V
 
∂U
=− (∆N )2 .
∂ N̄ T,V

Hence the result.


4.13. With µ fixed (as it is in the grand canonical ensemble),

(∆J)2 = (∆E)2 − 2µ(∆E)(∆N ) + µ2 (∆N )2 .

Substituting from eqn. (4.5.14) and from the previous problem, we get
( 2   )
2 2 ∂U ∂U 2
(∆J) = h(∆E) ican + − 2µ + µ (∆N )2 ,
∂ N̄ T,V ∂ N̄ T,V

which is the desired result.

4.14. The Clausius–Clapeyron equation (4.7.7) can be integrated to give


  
L 1 1
Pσ (T ) = Pσ (T0 ) exp − ,
k T0 T

where T0 = 373 K, Pσ (T0 ) = 1 atm and

L (2260 kJ/kg)(18 kg/kmol)


= = 4890 K.
k (6.02 × 1026 kmol−1 )(1.38 × 10−23 J/K)
31

This gives Pσ (273 K) ' 0.0082 atm and Pσ (473 K) ' 16 atm. The experi-
mental values are 0.006 atm and 15.3 atm respectively.

4.15. The correct value for the latent heat of sublimation near the triple point
is 2833 kJ/kg. Following the solution to problem 4.14,
  
L 1 1
Pσ (T ) = Pσ (T0 ) exp − ,
k T0 T

where T0 = 273 K, Pσ (T0 ) = 612 Pa and

L (2833 kJ/kg)(18 kg/kmol)


= = 6138 K.
k (6.02 × 1026 kmol−1 )(1.38 × 10−23 J/K)

This gives Pσ (193 K) ' 0.055 Pa which corresponds nearly exactly with
the experimental value.
4.16. The slope of the melting line is

dPm Lm (80 cal/g)(4.18 J/cal)(106 cm3 /m3 )


= ' ' −1.3 × 107 Pa/K.
dT T (∆v) (273 K)(−0.09 cm3 /g)

This gives Tm (100 atm) = −0.77◦ C.


4.17. The slopes at the triple point are of the form
 
dPσ si − sj ∆y
= = ,
dT ij vi − vj ∆x

so the vectors

[(v1 − v2 )x̂ + (s1 − s2 )ŷ] + [(v2 − v3 )x̂ + (s2 − s3 )ŷ] + [(v3 − v1 )x̂ + (s3 − s1 )ŷ] = 0

sum to zero. This makes the third vector the negative of the sum of the
first two vectors in each case, guaranteeing the stated geometry.
32

4.18. The liquid–vapor lines will appear much like in figure 6.2 but the liquid
branch will extend to P = 0. The upper end of the solid–liquid lines will
appear as in figure 6.2 but the lines will end at Ps .

4.19. Since p1 (µσ (T ), T ) = p2 (µσ (T ), T ) on the coexistence line


       
∂p1 dµσ ∂p1 ∂p2 dµσ ∂p2
+ = + ,
∂µ T dT ∂T µ ∂µ T dT ∂T µ
which gives
dµσ s1 − s2 −L
=− = .
dT n1 − n2 T ∆n
4.20. The liquid–vapor lines will appear much like in figure 6.2 but with the
liquid branch ending abruptly at Pt . The liquid side of solid–liquid lines
will start at Pt and extend upward as in figure 6.2 but the solid side of
the solid–liquid transition will be to the right of the liquid line (since the
solid has lower density) and will extend to P = 0.
Chapter 5

5.1. On transformation, a given operator  would become

 √ √   √ √ 
1/ √2 1/√2 a11 a12 1/√2 −1/√ 2
Â0 = Û ÂÛ −1 = .
−1/ 2 1/ 2 a21 a22 1/ 2 1/ 2

Equations (2), (3) and (4) of Sec. 5.3 would then be replaced by
     
1 0 0 −i 0 −1
σ̂x0 = , σˆy0 = , σ̂z0 = ,
0 −1 i 0 −1 0
 
1 cosh(βµB B) − sinh(βµB B)
ρ̂0 = βµB B
e + e−βµB B − sinh(βµB B) cosh(βµB B)

and
1
hσz0 i = Tr (ρ̂0 σ̂z0 ) = · 2 sinh(βµB B) = tanh(βµB B),
eβµB B + e−βµB B
with no change in the final result.
5.2. For a formal solution to this problem, see Kubo (1965), problem 2.32, pp.
178–80.
5.4. If we use the unsymmetrized wave function (5.4.3), rather than the sym-
metrized wave function (5.5.7), the density matrix of the system turns out
to be, cf. eqn. (5.5.11),
X 2 2
h1, . . . , N |e−β Ĥ |10 , . . . , N 0 i = e−β~ K /2m {uk1 (1) . . . ukN (N )} u∗k1 (10 ) . . . u∗kN (N 0 )


K
−β~2 (k12 +...+kN
2
)/2m
X
uk1 (1)u∗k1 (10 ) . . . ukN (N )u∗kN (N 0 )
  
= e
k1 ,...,kN
 
N n o
Y X 2 2
=  e−β~ kj /2m ukj (j)u∗kj (j 0 )  .
j=1 kj

33
34

Replacing the summation over kj by an integration, one gets [see the


corresponding passage from eqn. (5.5.12) to (5.5.14)]
 3N/2  
−β Ĥ 0 0 m m 2 2

h1, . . . , N |e |1 , . . . , N i = exp − ξ + . . . + ξN ,
2πβ~2 2β~2 1
(1)

where ξj = rj − r0j . The diagonal elements of the density matrix then are

h1, . . . , N |e−β Ĥ |1, . . . , N i = (m/2πβ~2 )3N/2 = 1/λ3N , (2)


where λ is the mean thermal wavelength of the particles. The structure of
expressions (1) and (2) shows that there is no spatial correlation among
the particles of this system.
The partition function now turns out to be
vN
Z
1 3N
QN (V, T ) ≡ Tr (e−β Ĥ ) = d r = ,
λ3N λ3N
with no Gibbs’ correction factor.
5.5. By eqn. (5.5.17), we have

  1
QN (V, T ) ≡ Tr e−β Ĥ = ZN (V, T ),
N !λ3N
where
Z X
ZN (V, T ) = {. . .}d3N r. (1)
P
P
In the zeroth approximation, = 1; see eqn. (5.5.19). So, ZN (V, T ) =
P
V N . In the first approximation,
X X X 2 2
=1± fij fji = 1 ± e−2πrij /λ . (2)
P i<j i<j

If λ is much smaller than the mean interparticle distance, we may write


 
X Y 2 2
 Y  X 
≈ 1 ± e−2πrij /λ = e−βvs (rij ) = exp −β vs (rij ) ,
 
P i<j i<j i<j

which leads to the desired result.


For the second part, we substitute (2) into (1) and integrate over the posi-
tion coordinates of the particles. We obtain, on assembling contributions
from all pairs of particles,
N (N − 1) V · λ3
ZN (V, T ) = V N ± · V N −2 3/2 .
2 2
35

The case N = 2 corresponds to eqn. (5.5.25) for Q2 (V, T ). For N  1


and N λ3  V , we may write
N
λ3 N λ3
  
ZN (V, T ) = V N 1 ± N 2 5/2 ≈ V N 1 ± 5/2 .
2 V 2 V

It follows that
N λ3
   
V
ln QN (V, T ) ≈ −N ln N + N + N ln + N ± 5/2 ,
λ3 2 V

whence
N 2 λ3 1 λ3
 
P ∂ ln QN N 1
≡ ≈ ∓ 5/2 2 = ∓ 5/2 2 ,
kT ∂V N,T V 2 V v 2 v

where v = V /N ; cf. eqns. (7.1.13) and (8.1.17).


5.7 and 8. For solutions to these problems, consult the references cited in Notes 10
and 11.
Chapter 6

6.1. We start with eqn. (6.1.19) and write it in the form


X n∗i
    
gi gi 
S=k n∗i ln + n ∗
i − ln 1 − a . (1)
i
n∗i a gi

Now, setting all gi = 1 and identifying (n∗i /gi ) with hnε i, see eqns. (6.1.18a)
and (6.2.22), we get
X 
1
 
S=k −hnε i lnhnε i + hnε i − ln(1 − ahnε i) . (2)
ε
a

Choosing a = −1 or +1, we obtain the desired results.


Next we have to verify that
( )
X X X
S = −k pε (n) ln pε (n) = −k hln pε (n)i. (3)
ε n ε

Substituting for pε (n) from eqn. (6.3.10) into (3) leads to the desired
result (2), with a = −1; substituting from eqn. (6.3.11) instead leads to
the desired result (2), with a = +1.
6.2. In the B.E. case, see eqn. (6.3.10),

pε (n) = (1 − r)rn [r = hnε i/(hnε i + 1); n = 0, 1, 2, . . .].

It follows that

X
hnε i = (1 − r) nr n = r/(1 − r),
n=0
X∞
n2ε = (1 − r) n2 rn = r(1 + r)/(1 − r)2 , so that


n=0
n2ε − hnε i2 = r/(1 − r)2 = hnε i + hnε i2 .


(1)

36
37

In the F.D. case, see eqn. (6.3.11),


1

2 X
nε = n2 pε (n) = pε (1) = hnε i, so that
n=0

2
nε − hnε i2 = hnε i − hnε i2 (2)
In the M.B. case, see eqn. (6.3.12), one can readily see that
X hnε in −hnε i X hnε in−2
hnε (nε − 1)i = n(n − 1) e = hnε i2 e−hnε i = hnε i2 , so that
n
n! n
(n − 2)!

2
nε − hnε i2 = hnε i. (3)

For the second part, we note, from eqn. 6.2.22, that


hnε i−1 = e(ε−µ)/kT + a.

Differentiating this result with respect to µ, we get


 
∂hnε i 1 (ε−µ)/kT 1
−hnε i−2 =− e =− [hnε i−1 − a].
∂µ T kT kT
It follows that  
∂hnε i
kT = hnε i − ahnε i2 . (4)
∂µ T
Comparing (4) with our previous results (1)–(3), and with formula (6.3.9),
we infer that, quite generally,

2
nε − hnε i2 = kT [∂hnε i/∂µ]T .

6.3. Starting with eqn. (6.2.15), we now have


" ` #
Y X Y  1 − (ze −βε )`+1 
Q(z, V, T ) = (ze −βε nε
) = ,
ε n =0 ε
1 − ze −βε
ε

so that
X
q(z, V, T ) = [ln{1 − (ze −βε )`+1 } − ln{1 − ze −βε }];
ε

cf. eqn. (6.2.17). It follows that


 
1 ∂q
hnε i = −
β ∂ε z,T,all other ε
−βε ` −βε
(` + 1)(ze ) (ze ) ze −βε
=− −βε
+
1 − (ze )`+1 1 − ze −βε
1 `+1
= −1 βε −
z e − 1 (z −1 eβε )`+1 − 1
For ` = 1, we obtain the Fermi-Dirac result; for ` → ∞ and z −1 eβε > 1
[see eqn. (6.2.16a)], we obtain the Bose-Einstein result.
38

6.4. To determine the state of equilibrium of the given system, we minimize


its free energy, U − TS , under the constraint that the total number of
particles, N , is fixed. For this, we vary the particle distribution from n(r)
to n(r) + δn(r) and require that the resulting variation

e2 n(r)δn(r0 ) + n(r0 )δn(r)


Z Z Z
0
δ(U − TS ) = drdr + e δn(r)ϕext (r)dr
2 |r − r0 |
Z
+ kT [1 + ln n(r)]δn(r)dr = 0,
R
while δN = δn(r)dr is, of necessity, zero. Introducing the Lagrange
multiplier λ, our requirement takes the form

n(r0 )
Z  Z 
0
e2 dr + eϕext (r) + kT [1 + ln n(r)] − λ δn(r)dr = 0.
|r − r0 |

Since the variation δn(r) in this expression is arbitrary, the condition for
equilibrium turns out to be

n(r0 )
Z
e2 dr0 + eϕext (r) + kT ln n(r) − µ = 0, (1)
|r − r|0

where µ = λ − kT .
Introducing the total potential ϕ(r), viz.

n(r)0
Z
ϕ(r) = ϕext (r) + e dr0 , (2)
|r − r0 |

condition (1) takes the Boltzmannian form

n(r) = exp[{µ − eϕ(r)}/kT ]. (3)

Choosing n(r) to be n0 at the point where ϕ(r) = 0, eqn. (3) may be


written as
n(r) = n0 exp[−eϕ(r)/kT ]. (4)
With ϕext (r) given, the coupled equations (2) and (4) together determine
the desired functions n(r) and ϕ(r).
6.5. The (un-normalized) distribution function for the variable ε in this prob-
lem is given by
f (ε)dε ∼ e−βε ε1/2 dε,

where use has been made of expression (2.4.7) for the density of states of
a free particle. It is now straightforward to show that

β −5/2 Γ(5/2) 3 2 =
β −7/2 Γ(7/2) 15
ε̄ = −3/2
= and ε −3/2
= 2
.
β Γ(3/2) 2β β Γ(3/2) 4β
39

It follows that
q
(ε2 − ε̄2 )
p
(∆ε)r.m.s. (3/2β 2 ) p
≡ = = (2/3).
ε̄ ε̄ (3/2β)

6.6. We have to show that, for any law of distribution of molecular speeds [say,
F (u)du],

R∞ R∞
u F (u)du u−1 F (u)du
0 0
R∞ · R∞ ≥ 1, i.e.
F (u)du F (u)du
0 0
2
Z∞ Z∞
∞
Z
u F (u)du · u−1 F (u)du ≥  F (u)du  .
0 0 0

For this, we employ Schwarz’s inequality (see Abramowitz and Stegun,


1964),
 b 2
Z Zb Zb
 f (x)g(x)dx  ≤ [f (x)]2 dx · [g(x)]2 dx ,
a a a

which holds for arbitrary functions f (x) and g(x) — so long as the integrals
exist; the equality holds
p if and only if f (x)p = c g(x), where c is a constant.
Now, with f (u) = uF (u) and g(u) = u−1 F (u), we obtain the desired
result.
For the Maxwellian distribution,
1 2
F (u)du ∼ e− 2 βmu u2 du.

It is then straightforward to see, with the help of the formulae (B.13), that
 1/2  1/2
I3 8 −1 I1 2βm
hui = = and hu i= = ,
I2 πβm I2 π

whence huihu−1 i = 4/π, in conformity with the inequality stated.

6.7. For light emitted in the x-direction, only the x-component of the molecular
velocity u will contribute to the Doppler effect. Moreover, for ux 
c, (ν − ν0 )/ν0 ' ux /c, which means that (λ − λ0 )/λ0 ' −ux /c. Now,
the distribution of ux among the molecules of the gas is governed by
the Boltzmann factor exp − 12 mu 2x /kT ; the distribution of λ in the light


emerging from the window will, therefore, be determined by the factor


exp − 12 mc 2 (λ − λ0 )2 /λ20 kT .

40

6.8. The partition function QN (β) = (1/N !)QN1 (β), where Q1 (β) is given by
Z   2 
1 p
Q1 (β) = 3 exp −β + mgz dp x dp y dp z dxdydz
h 2m
3/2
1 − e−βmgL

2πm
= · A , (1)
βh2 βmg

A being the area of cross-section of the cylinder. In the limit L → ∞,


 3/2
2πm A
Q1 (β) = ∝ T 5/2 . (2)
βh2 βmg

The thermodynamic properties of the system now follow straightforwardly.


In particular, U turns out to be 25 NkT and hence Cv = 52 Nk . The extra
contribution comes from the potential energy of the system, which also
rises with T . Note, from eqns. (1) and (2), that the effective height of the
gas molecules is (1 − e−βmgL )/βmg which for small heights is essentially
L itself but for large heights is essentially kT/mg — making the total
potential energy of the gas equal to NkT.
6.9. Correction to the first printing of third edition: the correct Hamiltonian
is
p2r (p2 − mr2 ω)2 p2 mr2 ω 2
H (pr , pθ , pz , r, θ, z) = + θ 2
+ z − .
2m 2mr 2m 2
This gives for the partition function

2πH R βmr2 ω 2 βmR2 ω 2


Z      
2πHkT
Q1 (V, T ) = 3 exp rdr = 3 exp −1
λ 0 2 λ mω 2 2

In the limit of small rotation rate, this becomes Q1 = πHR2 /λ3 = V /λ3
as expected.
The density is determined from hδ(z − z1 )δ(θ − θ1 )δ(r − r1 )/ri. This gives

βmω 2 r2
 
n(r) = n(0) exp .
2

Since the 238 UF6 molecules are heaver, their concentration is enhanced at
r = R, while the concentration of the 235 UF6 is enhanced near r = 0. The
ratio at r = 0 is given by
" #
m235 N235 exp 21 βm238 ω 2 R2 − 1

n235 (0)
=
m238 N238 exp 12 βm235 ω 2 R2 − 1

n238 (0)
 
m235 N235 1
≈ exp β (m238 − m235 ) ω 2 R2 .
m238 N238 2
41

A value of ωR = 500 m/s gives a 16% enhancement compared to the input


fraction. Drawing the uranium hexafluoride gas from near the center of
the cylinder results in a sample that is isotopically enhanced with 235 U
compared to the input concentration. This process may be repeated as
often as needed to achieve the isotopic fraction needed.
6.10. Consider a layer of the gas confined between heights z and z + dz . For
hydrostatic equilibrium, we must have
P (z + dz ) + ρgdz = P (z),
where ρ is the mass density of the gas. In differential form, one gets
dP /dz = −ρg = (−mg/kT )P. (1)
(a) If T is uniform, eqn. (1) can be readily integrated, with the result
ln P = −(mg/kT )z + const., (2)
which yields the desired formula: P (z) = P (0) exp(−mgz /kT ).
(b) If, on the other hand, the equilibrium is attained adiabatically, then
T is related to P ; in fact, T ∝ P (γ−1)/γ . We now get
dT γ − 1 dP γ − 1 mg
= =− dz . (3)
T γ P γ kT
This means that T now decreases essentially linearly with height.
The pressure P and the density ρ go hand in hand with T — varying
as T γ/γ−1 and T 1/γ−1 , respectively.
6.11. (a) For the given system,
2 1/2
+m20 c2 )
f (p)dp = const.e−βε(p) (4πp2 dp) = C e−βc(p p2 dp.
The normalization constant C is determined by the condition
Z Z∞
2 1/2
+m20 c2 )
f (p)dp = C e−βc(p p2 dp = 1.
0

Substituting p = m0 c sinh θ, we get for the left-hand side of this


equation
Z∞
2
C e−βm0 c cosh θ
m30 c3 sinh2 θ cosh θ dθ
0
Z∞
 
2 2
e−βm0 c cosh θ e−βm0 c cosh θ
= C m30 c3  sinh θ cosh θ|∞
0 + cosh(2θ)dθ
−βm0 c2 βm0 c3
0
2 −1
= Cm 30 c3 · (βm0 c ) 2
K2 (βm0 c ).
Equating this result with 1, we obtain the desired expression for C.
42

(b) Using the limiting forms


(
(π/2x)1/2 e−x (x  1)
K2 (x) ≈ ,
2/x2 (x  1)

we obtain, rather straightforwardly, the nonrelativistic and the ex-


treme relativistic limits of the distribution.
(c) Since

dε m0 c2 d(cosh θ)
u= = = c tanh θ,
dp m0 cd (sinh θ)
Z∞
2
hpui = C {m0 c2 sinh θ tanh θ} e−βm0 c cosh θ m30 c3 sinh2 θ cosh θ dθ
0
Z∞
2
=C m40 c5 e−βm0 c cosh θ
sinh4 θ dθ.
0

Once again, integrating by parts (this time twice), we obtain

hpui = C m40 c5 · 3(βm0 c2 )−2 K2 (βm0 c2 ).

Substituting for C, we obtain: hpui = 3/β — regardless of the sever-


ity of the relativistic effects and in conformity with the results of
Secs. 3.7 and 6.4.
6.12. Ordinarily, when a molecule is reflected from a stationary wall that is
perpendicular to the z-direction, the z-component of its velocity u simply
changes sign, i.e. u0z = −uz . If the wall is receding at velocity v in the
direction of its normal, the above result changes to (u0z − v) = −(uz − v),
so that u0z = −(uz − 2v). This results in a change in the translational
energy of the molecule which, for small v, is given by
1 1
∆ε = mu 02 2
z − mu z ' −2mu z v.
2 2
If A is the area of the wall, the net change in the energy of the gas, in
43

time δt, is then given by, cf. eqn. (6.4.10),


Z∞ Z∞ Z∞
δE = Aδt · n (−2mu z v)uz f (u)du x du y du z
ux =−∞ uy =−∞ uz =−∞

Z2π Zπ/2 Z∞
= −Avδt · n {2mu 2 cos2 θf (u)}(u2 sin θ du dθ dφ)
φ=0 θ=0 u=0
Z∞
1
= −δV · n {mu 2 f (u)}(4π u2 du)
3
0
2 2 Ek
= −δV · nε̄k = −δV · , (1)
3 3 V
where Ek is the total kinetic energy of the gas. Note that, since the gas
continues to be in a state of (quasi-static) equilibrium, the change δE (even
though it originates in the translational motion of the molecules) becomes
eventually a change in the internal energy U of the gas (which may well
have contributions from degrees of freedom other than translational). If
U = aE k , we may write
1 2 δV
δU = − Ek
δEk = . (2)
a 3a V
Next, since PV = (2/3)Ek , we get
δP δV δEk 2 δV
+ = =− . (3)
P V Ek 3a V
Re-arranging (3) and integrating it, we obtain the desired result.
In the extreme relativistic case, the factor 2/3 is replaced throughout by
1/3, leading to the alternate value of γ.
6.13. We refer to expression (6.4.11) of the text. For part (a) of the question,
we integrate only over u and ϕ, to get

1
dR θ = n(ū/4π) · 2π sin θ cos θdθ = nū sin θ cos θdθ.
2
For part (b), we integrate only over θ and ϕ, to get
 m 3/2 2
dR u = nπ · f (u)u3 du, where f (u)|M.B. = e−mu /2kT .
2πkT
For part (c), we refer to expression (6.4.10) instead and get
 m 3/2 Z∞ Z∞ Z∞
2 2 2
RE = n e−m(ux +uy +uz )/2kT uz du x du y du z
2πkT √
ux =−∞ uy =−∞ uz = 2E/m
 1/2
kT
=n e−E/kT .
2πm
44

It follows that
 1/2   
RE (T2 ) T2 −E 1 1
= exp − .
RE (T1 ) T1 k T2 T1

With T1 = 300 K, T2 = 310 K and E = 10−19 J, this ratio turns out to


be about 2.2.
6.14. (a) We start by calculating the kinetic energy associated with the z-
component of the motion of the effused molecules. Proceeding as
Section 6.4 of the text, we get [see eqn. (6.4.11)]
π/2
R R∞
  (u3 cos3 θ)f (u)u2 sin θdudθ
1 1 1 0 0
mu 2z 2 2
= mhu cos θi = m
2 2 2 π/2
R R∞
(u cos θ)f (u)u2 sin θdudθ
0 0
1 hu3 i
= m ;
4 hui

note that the averages on the right-hand side are taken over the gas
inside the vessel. It is not difficult to show, see the corresponding
calculation in Problem 6.6 and the formulae (B.13b), that

hu3 i I5 4
= = ,
hui I3 βm

so that 21 mu 2z , for the effused molecules, = 1/β = kT . The kinetic



energy associated with the x- and y-components of the molecular


motion will be the same as inside the vessel, viz. 21 kT each. It
follows that the mean energy ε of an effused molecule is 2 kT.
(b) Assuming quasi-static equilibrium, the relations E = (3/2)NkT and
P = NkT /V will continue to hold for the gas inside the vessel. How-
ever, in view of the result obtained in part (a), we shall also have
 
dE d 3 dN
≡ NkT = 2kT .
dt dt 2 dt

It follows that
dT 1 dN
= and hence T ∝ N 1/3 ;
T 3 N
it further follows that P ∝ N 4/3 .
As for explicit variations with t, we make use of eqn. (6.4.13) and write
 1/2
dN 1 1 aN 8kT
= − anhui = − .
dt 4 4 V πm
45

Combining the last two results, we get


 1/2
dT 1a kT
=− dt,
T 3V 2πm

so that T = T0 (1+ct)−2 , where c = (a/6V )(kT 0 /2πm)1/2 . The variations


of N and P with t follow straightforwardly.

6.15. If nH is the number of holes per unit area of the surface of the balloon (of
radius r), a the area of each hole and t the duration of the leak, then the
total number of molecules leaking is given by
1 h i
∆N = nū · nH (4πr2 )at ū = (8kT /πm)1/2 .
4
The fraction of the molecules leaking is thus given by
 1/2
∆N 1 8kT
= · nH (4πr2 )at.
N 4V πm

Since V = (4π/3)r3 , we get


 1/2
∆N r 2πm
nH = · .
N 3at kT
2
Substituting the data given, we obtain: nH ' 187 holes/m .
6.16. The rate of effusion of molecules from side A to side B, through a hole of
cross-section S, is given by the expression
r
1 8kT A PA
RA→B = nA S=√ S;
4 πmA 2πmA kT A
the same from side B to side A is given by
r
1 8kT B PB
RB→A = nB S=√ S.
4 πmB 2πmB kT B
In the stationary state, these two expressions will be equal — which leads
to the condition of dynamic equilibrium

PA /PB = (mA TA /mB TB )1/2 .

If the two gases are samples of the same gas, the condition simplifies to

PA /PB = (TA /TB )1/2 .


46

6.18. The (un-normalized) velocity distribution for a pair of molecules is given by


2 2
F (u1 , u2 )d3 u1 d3 u2 ∼ e− 2 βm(u1 +u2 ) d3 u1 d3 u2 .
1

We define the relative velocity, v, and the velocity of the centre-of-mass,


V, in the usual manner, viz.
1
v = u2 − u1 , V = (u1 + u2 ).
2
This results in a new distribution for the variables v and V:
1 2 2
F (v,V)d3 vd 3 V ∼ e− 4 βmv d3 v · e−βmV d3 V.

It is now straightforward to show that hvi = (16/πβm)1/2 = 2hui, √ while
hv2 i = (6/βm) = 2hu2 i. The latter result implies that vr.m.s. = 2 ur.m.s ..
We note that, since

v2 = u21 + u22 + u22 − 2u1 · u2 ,

hv2 i = 2hu2 i, regardless of the law of distribution of velocities — so long


as it is isotropic, making hu1 · u2 i = 0.
6.19. The (un-normalized) joint distribution for the molecular energies ε1 and
ε2 is

1/2 1/2
f (ε1 , ε2 )dε1 dε2 ∼ e−β(ε1 +ε2 ) ε1 ε2 dε1 dε2 .

To obtain the desired distribution, we set ε2 = E − ε1 and integrate over


all relevant values of ε1 , with the result that

ZE
−βE
P (E)dE ∼ e {ε1 (E − ε1 )}1/2 dε1 dE
0
−βE
∼e E 2 dE ;

cf. eqns. (3.4.3) and (3.5.16), with N = 2. It is now straightforward to


check that
β −4 Γ(4) 3
hEi = −3 = .
β Γ(3) β

6.20. The relative fraction of the excited atoms in the given sample of the helium
gas would be 3e−βε1 , where
hc
βε1 = ' 38.22.
kT λ1

The desired fraction turns out to be extremely small — about 7 × 10−17 .


47

6.21. We extend the treatment of Problem 3.14 to the reaction AB + CD ↔


AD + CB and obtain, in equilibrium,
nAD nCB fAD fCB
= = K(T ).
nAB nCD fAB nCD
For the given reaction,
2
fHD
K(T ) = ,
fHH fDD
where each f is a product of three factors — the translational, the rota-
tional and the vibrational.
Now, for a heteronuclear molecule like HD we have, at high temperatures,
 3/2
mHD kT 2IHD kT kT
fHD ≈ V · · ,
2π~2 ~ 2 ~ωHD

while for a homonuclear molecule like HH we have instead


 3/2
mHH kT IHH kT kT
fHH ≈ V · · ;
2π~2 ~2 ~ωHH

see Note 11 of the text. It follows that, at high temperatures,

m3HD 2
IHD 2
IHD ωHH ωDD
K(T ) ≈ 4 3/2 3/2
· · · 2 . (1)
mHH mDD IHH IDD IHH IDD ωHD

Assuming the internuclear distances to be the same, the I’s here will be
proportional to the reduced masses of the molecules; the ω’s, on the other
hand, are inversely proportional to the square roots of the reduced masses.
Accordingly,
2
IHD ωHH ωDD µ3HD {mH mD /(mH + mD )}3
· 2 = 3/2 3/2
= 3/2 1 3/2 . (2)
IHH IDD ωHD µHH µDD 1
2 mH 2 mD

At the same time,

m3HD (mH + mD )3
3/2 3/2
= . (3)
mHH mDD (2mH )3/2 (2mD )3/2

Substituting (2) and (3) into (1), we see that K(T ) ≈ 4.

6.23. The potential V (r) is minimum at r = r0 , which determines the equilib-


rium value of r. Accordingly, the quantum of the rotational motion of
the molecule is ~2 /2I, where I = µr02 . This gives for Θr the expression
~2 /2µr02 k = ~2 /mr 20 k because the reduced mass µ in this case is equal to
m/2. Substituting the given data, Θr turns out to be about 75 K. This
gives a fairly clear idea of the “temperature range” where the rotational
48

motion of the hydrogen molecules begins to contribute towards the specific


heat of the gas.
Next we expand V (r) in the neighborhood of r = r0 and write

V (r) = −V0 + (V0 /a2 )(r − r0 )2 + . . . .

This gives an ω equal to (2 V0 /µa2 )1/2 = (4V0 /ma 2 )1/2 and hence a Θv
equal to ~(4V0 /ma 2 )1/2 /k. Substituting the given data, Θv turns out to
be about 6260 K. Again, this gives a fairly clear idea of the “temperature
range” where the vibrational motion of the hydrogen molecules begins to
contribute towards the specific heat of the gas.

6.24. The effective potential of a diatomic molecule (including both rotation


and vibration) is given by

1 ~2
V (r) = −V0 + µω 2 (r − r0 )2 + J(J + 1).
2 2µr2

The equilibrium value of r is obtained by minimizing V (r), with the result

~2 ~2
(req − r0 ) = J(J + 1) ' J(J + 1).
µ2 ω 2 req
3 µ2 ω 2 r03

It follows that
2
~2

∆r0 Θr
' 2 2 4 J(J + 1) = 4 J(J + 1).
r0 µ ω r0 Θv

Using data from the preceding problem, we find that for a hydrogen
molecule the fractional change in r0 is O(10−3 ).

6.25. The occupation number NJ is proportional to (2J + 1) e−εJ /kT . It fol-


lows that
N0 1 N1 3
= e−(ε0 −ε2 )/kT , = e−(ε1 −ε2 )/kT .
N2 5 N2 5
Substituting the given data, we get
N0 1 N1 3
= e−1.086 ' 0.0675, = e−0.760 ' 0.2806;
N2 5 N2 5
in other words,

N0 : N1 : N2 :: 0.050 : 0.208 : 0.742.

6.29. The various contributions to the molar specific heat of the gas at 300 K are:

(i) translational — the amount being (3/2)R.


49

(ii) rotational — since the characteristic values of the parameter Θr in


this case are of the order of 10 K, these degrees of freedom may
be treated classically, which yields a contribution of (3/2)R; see
eqn. (6.5.42).
(iii) vibrational — here, the parameters Θv are such that the various con-
tributions have to be calculated quantum-mechanically, using formula
(6.5.44). We find that

Θ1,2 Θ3,4 Θ5 Θ6
' 16.00, ' 4.56, ' 16.37, ' 7.80,
T T T T
with the result that only modes 3 and 4 make appreciable contributions to
the specific heat of the gas; it turns out that each of these contributions is
about 0.22R. The contribution from mode 6 is about 0.02R, while those
from modes 1,2 and 5 are entirely negligible.
The net result is: 3.46R.
P
6.30. Equation (6.6.3) can be written α να µα = 0 where the stioichiometric
coefficients are understood to be positive if they appear on the right hand
side of equation (6.6.1) and negative if on the left. Using equation (6.6.5)
gives
X 
να α + kT ln nα λ3α − kT ln jα = 0,
 
α

which gives
  
X nα
να α + kT ln n0 λ3α − kT ln jα + kT ln

= 0.
α
n0

Rearranging gives
 
X nα X 
να α + kT ln n0 λ3α − kT ln jα
 
να ln = −β
α
n 0 α
X
= −β να µ(0)
α ,
α

(0)
where µα is the chemical potential of species α at temperature T and
standard density n0 . Equation (6.6.6) follows from exponentiating both
sides.
6.31. Equation (6.6.11) gives
s
[CO] 1
=
[CO2 ] K(T ) [O2 ]
50

where
 
(0) (0) (0)
K(T ) = exp −2βµCO2 + 2βµCO + βµO2 .

For the parameters given in the problem K(1500 K) = 4 × 1010 which


yields [CO] / [CO2 ] ' 5 × 10−5 = 50 ppm, while K(600 K) = 1.7 × 1040
which yields [CO] / [CO2 ] ' 7 × 10−20 which yields a negligible [CO] con-
centration.
6.32. The equilibrium constant for N2 + O2 → 2NO is
2
[NO] j2 λ3 λ3
= K(T ) = e−β∆ε NO N26 O2 .
[N2 ] [O2 ] jN2 jO2 λNO
The internal partition functions are of the form
  
 T for kT  ~ω
j=  Θr 
T kT


Θr ~ω for kT  ~ω

which leads to
   
 e−β∆ε ΘN22ΘO2 303
3/2 323/2 for kT  ~ω
Θ 28
K(T ) =  NO    
 e−β∆ε ΘN22ΘO2 303
3/2 3/2
ωN2 ωO2
2 for kT  ~ω
Θ NO 28 32 ω NO

6.33. The equilibrium relation is


[CO2 ]H2 O]2
=K
[CH4 ][O2 ]2
Let [excess] be the initial excess amount of O2 above stoichiometry, and
[unburned] be the unburned amount of CH4 . Then
3
[CO2 ]H2 O]2 4 ([CH4 ]0 − [unburned])
= 2 =K
[CH4 ][O2 ]2 [unburned] ([excess] + 2[unburned])
Since K  1, at the stoichiometric point [excess] = 0 so
[CH4 ]0
[unburned] ≈ .
K 1/3
On the lean side of the stoichiometric point [excess] > 0 so

4[CH4 ]20
[unburned] ≈ .
[excess]K
Finally, on the rich side of the stoichiometric point [excess] < 0 so
[excess]
[unburned] ≈ − .
2
51

6.34. Equation (6.6.3) gives µNa = µNa+ + µe where

µNa = −εb − kT ln 2 + kT ln nNa λ3Na ,




µNa+ = kT ln nNa+ λ3Na ,




µe = −kT ln 2 + kT ln ne λ3e ,


where b is the ionization energy of Na. These lead to


nNa
= eβb λ3e .
nNa+ ne

If the total density is n0 = nNa + nNa+ , the ionized fraction f = nNa+ /n0 ,
and the system is charge neutral, then
1−f
= eβb n0 λ3e = s,
f2
which has solution

1 + 4s − 1
f= .
2s
Chapter 7

7.2. With N0  N , eqn. (7.1.8) reads

nλ3 = g3/2 (z) = z + 2−3/2 z 2 + 3−3/2 z 3 + 4−3/2 z 4 + . . . , (1)

where n is the particle density. To invert this series, we write

z = c1 (nλ3 ) + c2 (nλ3 )2 + c3 (nλ3 )3 + c4 (nλ3 )4 + . . . (2)

and substitute into (1). Equating coefficients of like powers of (nλ3 ) on


the two sides of the resulting equation, we get

1 = c1 , 0 = c2 + 2−3/2 c21 , 0 = c3 + 2−3/2 · 2c1 c2 + 3−3/2 c31 ,


0 = c4 + 2−3/2 c22 + 2c1 c3 + 3−3/2 · 3c21 c2 + 4−3/2 c41 , . . . .


It follows that

c1 = 1, c2 = −2−3/2 , c3 = (1/4) − 3−3/2 ,


c4 = 5.6−3/2 − 5.2−9/2 − (1/8), . . .

We now write eqn. (7.1.7) in the form

PV 1
= (z + 2−5/2 z 2 + 3−5/2 z 3 + 4−5/2 z 4 + . . .)
NkT nλ3
and substitute expression (2) into it. This leads to the desired result
(7.1.13), with

a1 = c1 = 1, a2 = c2 + 2−5/2 c21 = −2−5/2 ,


a3 = c3 + 2−5/2 · 2c1 c2 + 3−5/2 c31 = (1/8) − 2.3−5/2 ,
a4 = c4 + 2−5/2 c22 + 2c1 c3 + 3−5/2 · 3c21 c2 + 4−5/2 c41


= 3.6−3/2 − 5.2−11/2 − (3/32),

in agreement with the values quoted in expressions (7.1.14).

52
53

7.3. By eqns. (7.1.24) and (7.1.26), nλ3 = g3/2 (z) while nλ3c = ζ(3/2). It
follows that

 −2  −2/3
T λ g3/2 (z)
≡ = .
Tc λc ζ(3/2)
The right-hand side of this equation may be approximated with the help
of formula (D.9), with the result that
−2/3
ζ(3/2) − 2π 1/2 α1/2 + . . . 4π 1/2 α1/2

T
= ≈1+ ,
Tc ζ(3/2) 3ζ(3/2)
valid for α  1 and hence for T & Tc . The desired result now follows
readily.
7.4. By eqn. (7.1.7), P = cT 5/2 g5/2 (z), where c is a constant. Differentiating
this result with respect to T at constant P, we get
 
5 3/2 5/2 ∂g5/2 (z) ∂z
0 = cT g5/2 (z) + cT ,
2 ∂z ∂T P
so that  
∂z 5 g5/2 (z)
=− .
∂T P 2T {∂g5/2 (z)/∂z}
Using the recurrence relation (D.10), we get the desired result
 
1 ∂z 5 g5/2 (z)
=− . (1)
z ∂T P 2T g3/2 (z)

Now, CP = T (∂S/∂T )P,N and CV = T (∂S/∂T )V,N . In view of the


fact that S, at constant N, is a function of z only, see eqn. (7.1.44a),
we may write
       
∂S ∂z ∂S ∂z
CP = T and CV = T .
∂z N ∂T P ∂z N ∂T v
It follows that
CP (∂z/∂T )P
γ= = .
CV (∂z/∂T )v
Substituting from eqn. (1) above and from eqn. (7.1.36), we obtain the
desired result

CP /CV = (5/3) g5/2 (z)g1/2 (z)/{g3/2 (z)}2 .


 

For T  Tc , which implies z  1, we recover the classical result: γ =


5/3. As T → Tc , z → 1 and the function g1/2 (z) diverges as α−1/2 ; see
eqn. (D.8). Along with it, both γ and CP diverge as (T − Tc )−1 ; see the
relation established in Problem 7.3.
54

7.5. (a) We have to evaluate the quantities


   
1 ∂n 1 ∂n
κT = and κS = ,
n ∂P T n ∂P z

where n = N/V . For N0  N, n(T, z) = aT 3/2 g3/2 (z), where a is a


constant; see eqn. (7.1.8). It follows that
 
3 1/2 3/2 1
dn = aT g3/2 (z)dT + aT g1/2 (z) dz .
2 z

Similarly, since P = cT 5/2 g5/2 (z), where c is a constant,


 
5 3/2 5/2 1
dP = cT g5/2 (z)dT + cT g3/2 (z) dz .
2 z
The quantities κT and κS are then given by
1 a g1/2 (z) 1 3a g3/2 (z)
κT = and κS = .
n cT g3/2 (z) n 5cT g5/2 (z)
Since c = ak , the desired results follow readily.
Note that, as z → 1, κT diverges in the same manner as γ and CP .
(b) Since P = 2U/3V, (∂P/∂T )V = 2CV /3V . It follows that
2 2 g
1 g1/2 (z) 4CV 4CV 1/2 (z)
CP − CV = TV · · = ,
nkT g3/2 (z) 9V 2 9Nk g3/2 (z)
in agreement with eqn. (7.1.48a). The other result follows straight-
forwardly.
7.6. For T > Tc , we employ expression (7.1.37) and write
     
1 ∂CV ∂ CV ∂ ln z
= · .
Nk ∂T V ∂ ln z Nk ∂T v

The first factor turns out to be


15 g3/2 (z)g3/2 (z) − g5/2 (z)g1/2 (z) 9 g1/2 (z)g1/2 (z) − g3/2 (z)g−1/2 (z)

4 {g3/2 (z)}2 4 {g1/2 (z)}2
3 15 g5/2 (z)g1/2 (z) 9 g3/2 (z)g−1/2 (z)
= − + .
2 4 {g3/2 (z)}2 4 {g1/2 (z)}2
The second factor is given by eqn. (7.1.36). Multiplying the two, we obtain
the desired result.
For T < Tc , we employ expression (7.1.31) instead. Since CV is now
proportional to T 3/2 ,
 
1 ∂CV 3 CV
= ,
Nk ∂T V 2T Nk
55

which leads to the result quoted in the problem.


As T → Tc from above, the quantity under study approaches the limiting
value
27 {ζ(3/2)}2 · Γ(3/2)α−3/2
 
1 45 ζ(5/2)
−0− ;
Tc 8 ζ(3/2) 8 {Γ(1/2)α−1/2 }3
on the other hand, as T → Tc from below, we obtain simply

1 45 ζ(5/2)
· .
Tc 8 ζ(3/2)

The discontinuity in the slope of the specific heat curve at T = Tc is,


therefore, given by
  2   2
Nk 27 3 (π 1/2 /2) 27Nk 3
· ζ · = ζ .
Tc 8 2 π 3/2 16πTc 2

7.7. Since P = 2U/3V, (∂ 2 P/∂T 2 )v = (2/3V )(∂CV /∂T )V . An explicit ex-


pression for this quantity can be written down using the result quoted in
Problem 7.6.
Next, since µ = kT ln z, we obtain using eqn. (7.1.36)
   
∂µ kT ∂z 3 g3/2 (z)
= k ln z + · = k ln z − k ,
∂T v z ∂T v 2 g1/2 (z)
 2 
3 g1/2 (z)g1/2 (z) − g3/2 (z)g−1/2 (z)
  
∂ µ ∂ ln z
= k− k
∂T 2 v 2 {g1/2 (z)}2 ∂T v
3k g3/2 (z) 9k {g3/2 (z)}2 g−1/2 (z)
= −
4T g1/2 (z) 4T {g1/2 (z)}3

Similarly, using a result from Problem 7.4, we obtain

15k g5/2 (z) 25k {g5/2 (z)}2 g1/2 (z)


 2 
∂ µ
2
= − .
∂T P 4T g3/2 (z) 4T {g3/2 (z)}3

We also note, see eqns. (7.1.37) and (7.1.48b), that

CP 25 {g5/2 (z)}2 g1/2 (z) 15 g5/2 (z)


= − .
Nk 4 {g3/2 (z)}3 4 g3/2 (z)

It is now straightforward to see that the stated thermodynamic rela-


tions are indeed satisfied. The critical behavior of these quantities is also
straightforward to check.
7.8. One readily sees that
 
∂P 1
w2 = = ,
∂(nm) S mnκS
56

where κS is the adiabatic compressibility of the fluid. Using a result from


Problem 7.5, we get for the ideal Bose gas

5kT g5/2 (z)


w2 = .
3m g3/2 (z)

Next,  
2 2ε 2 U 3kT g5/2 (z)
hu i = = = ;
m mN m g3/2 (z)
see eqns. (7.1.8) and (7.1.11). Clearly, w2 = (5/9) < u2 >.
7.9. We start by calculating the expectation values of the quantities ε1/2 and
ε−1/2 :
R∞ R∞
hnε iε1/2 a(ε)dε hnε iε−1/2 a(ε)dε
0 0
hε1/2 i = R∞ , hε−1/2 i = R∞ .
hnε ia(ε)dε hnε ia(ε)dε
0 0

The integral in the denominator has been evaluated in Section 7.1; those
in the numerator can be evaluated like-wise, with the results
Γ(2)g2 (z) Γ(1)g1 (z)
hε1/2 i = (kT )1/2 , hε−1/2 i = (kT )−1/2 .
Γ(3/2)g3/2 (z) Γ(3/2)g3/2 (z)

It follows that
r r
2 1/2 8kT g2 (z)
hui = hε i = , while
m πm g3/2 (z)
r r
−1 m −1/2 2m g1 (z)
hu i = hε i= .
2 πkT g3/2 (z)

Multiplying the last two expressions, we obtain the desired result.


For z → 0, we recover the classical result stated in Problem 6.6. For
z → 1, we encounter divergence of the quantity hu−1 i, which arises from
the contribution made by the particles in the condensate (for which u = 0).
7.11. Under the conditions of this problem, the summation in eqn. (7.1.2) has
to be carried out over the states of the internal spectrum as well as over
the translational states. Expression (7.1.16) is then replaced by
  
V n  µ o V µ − ε1
Ne = (Ne )0 + (Ne )1 = 3 g3/2 exp + 3 g3/2 exp .
λ kT λ kT

The critical temperature Tc is then determined by the condition


V V
g3/2 (1) + 3 g3/2 (x) = N, where x = e−ε1 /kT c . (1)
λ3c λc
57

For x  1, g3/2 (x) ' x and eqn. (1) gives

λ3c ' (V /N )[ζ(3/2) + x].


3
Comparing this with the standard result λ0c = (V /N )ζ(3/2), we get
2 −2/3
λ0c
 
Tc x 2/3 2/3 −ε1 /kT 0c
≡ ' 1+ '1− x'1− e .
Tc0 λc ζ(3/2) ζ(3/2) ζ(3/2)
For x . 1, on the other hand, g3/2 (x) ' ζ(3/2)−2π 1/2 (− ln x)1/2 ; eqn. (1)
now gives
λ3c ' (2V /N )[ζ(3/2) − π 1/2 (ε1 /kT c )1/2 ], whence
( " 1/2 #)−2/3
π 1/2

Tc ε1
' 2 1−
Tc0 ζ(3/2) kT c
" 1/2 #
1/2

2 π ε 1
' 2−2/3 1 + 21/3 .
3 ζ(3/2) kT 0c

7.12. The relative mean-square fluctuation in N is given by the general for-


mula (4.5.7),
(∆N )2 kT
= κT , (1)
N̄ 2 V
while κT for the ideal Bose gas is given in Problem 7.5. As T → Tc from
above, the function g1/2 (z) and, along with it, both κT and the relative
fluctuation in N diverge!
The mean-square fluctuation in E is given by the general formula (4.5.14),
viz.
(∆E)2 = kT 2 CV + {(∂U/∂N )T,V }2 (∆N )2 . (2)
The first term in (2), for the ideal Bose gas, is determined by eqn. (7.1.37)
and stays finite at all T. The second term can be evaluated with the help
of eqns. (7.1.8 and 11), whereby
   
∂U ∂g5/2 (z) g3/2 (z)
= = . (3)
∂N T,V ∂g3/2 (z) T,V g3/2 (z)

The second term in (2) is, therefore, inversely proportional to g1/2 (z) and
hence vanishes as T → Tc ; this happens because the energy associated
with the Bose condensate (which is, in fact, the component responsible
for the dramatic rise in the fluctuation of N ) is zero. Thus, all in all, the
relative fluctuation in E is negligible at all T.
7.13. It is straightforward to see that for a Bose gas in two dimensions
Z∞ Z∞
1 A · 2πp dp A · 2πmkT dx A2
Ne = = = g1 (z),
z −1 eβε−1 h2 h2 z −1 ex − 1 λ
0 0
58

while
z
N0 = .
1−z
Since Bose-Einstein condensation requires that z → 1, the critical tem-
perature Tc , by the usual argument, is given by
 
N
λ2c = g1 (1) = ∞ [for g1 (z) = − ln(1 − z)].
A
It follows that Tc = 0.
More accurately, the phenomenon of condensation requires that both Ne
and N0 be of order N . This means that, while z ' 1, (1 − z) be of
order N −1 and hence λ2 be of order (A ln N /N ). Since the ratio (A/N ) ∼
`2 , the condition for condensation takes the form (λ2 /`2 ) = O(ln N ). It
follows that
h2 h2 1
T ≡ 2
∼ .
2πmk λ mk `2 ln N
7.14. With energy spectrum ε = Ap s , the density of states in the system is
given by, see formula (C.7b),
V 2π n/2 n−1 V 2π n/2
a(ε)dε = p dp = ε(n/s)−1 dε. (1)
hn Γ(n − 2) hn sAn/s Γ(n/2)

This leads to the expression


Z∞
V 2π n/2 ε(n/2)−1
N − N0 = n n/s dε
h sA Γ(n/2) z −1 eβε
−1
0
V 2π n/2 Γ(n/s)
= n n/s (kT )n/s gn/s (z), (2)
h sA Γ(n/2)

while N0 = z/(1 − z). Similarly,


1 2π n/2 Γ(n/s)
P = (kT )(n/s)+1 g(n+s)+1 (z). (3)
hn sAn/s Γ(n/2)

Next, following the derivation of eqn. (7.1.11), we get


  
2 ∂ PV n
U = kT = PV , (4)
∂T kT z,v s

so that P = sU /nV .
The onset of Bose-Einstein condensation requires that z → 1 at a finite
temperature Tc . A glance at eqn. (2) tells us that this will happen only
if n > s and that the critical temperature Tc will then be determined by
the equation
V 2π n/2 Γ(n/s) n
N = n n/s (kT c )n/s ζ . (5)
h sA Γ(n/2) s
59

For T < Tc , Ne will be equal to N (T /Tc )n/s while N0 will be given by


the balance (N − Ne ).
To study the specific heats we first observe, from eqns. (2)–(4), that for
T > Tc (when N0  N )
n
U= NkT · g(n/s)+1 (z)/gn/s (z) (6)
s
Next, using eqns. (2) and (3), and the recurrence relation (D.10), we get
   
1 ∂z n 1 gn/s(z) 1 ∂z n 1g
(n/s)+1 (z)
=− and =− +1 .
2 ∂T v s T g(n/s)−1 (z) z ∂T P s T gn/s (z)
(7)
It is now straightforward to show that
g
(n/s)+1 (z)
Cv n n  n 2 g (z)
n/s
= +1 − (8)
Nk s s gn/s (z) s g(n/s)−1 (z)
and
2 {g 2
(n/s)+1 (z)} g(n/s−1) (z)
g
CP (n/s)+1 (z)
n n n
= +1 − + 1 . (9)
Nk s {gn/s (z)}3 s s gn/s (z)
The limiting cases suggested in the problem follow quite easily.
7.15. The position and momentum representations of the Schrodinger equation
after the potential is turned off at time t = 0 is
~2 ∂ 2 ψ ∂ψ
− = i~ ,
2m ∂x2 ∂t
p2 ∂ ψ̂
ψ̂ = i~ .
2m ∂t
The momentum representation is easily solved
 2 
p t
ψ̂(p, t) = exp ψ̂(p, 0),
2i~m
where
Z
1
ψ̂(p, 0) = √ eipx/~ ψ(x, 0)dx.
2π~
This leads to (suppressing the normalization factor)
p2
  
2 i~t
ψ̂(p, t) ∼ exp − 2 a + .
2~ m
Inverse Fourier transforming gives

x2
 
a 1
ψ(x, t) = p exp − 2 .
π 1/4 a2 + i~t/m 2 a + i~t/m
60

This solves the Schrodinger equation and leads to the one-dimensional


density
x2
 
2 1 1
|ψ(x, t)| = 1/2 p exp − 2 .
π a 1 + (~t/ma2 )2 a (1 + (~t/ma2 )2 )

This gives the spatial distribution for one cartesian direction once you
note that ~/ma2 = ω0 . At long-time, the width of the distribution grows
linearly in time.
7.16. The one-dimensional normalized joint momentum–position density at time
t = 0 is given by
βp2 βmω 2 x2
 
ω
f (p, x, 0) = exp − − .
2πkT 2m 2
After the potential is turned off at t = 0, the particles move ballistically
so the density becomes
βp2 βmω 2 (x + pt/m)2
 
ω
f (p, x, t) = f (p, x + pt/m, 0) = exp − − .
2πkT 2m 2
The spatial density is then given by
r
1 βmω 2 x2
Z  
ω 2πmkT
n(x, t) = f (p, x + pt/m, 0)dp = exp − .
2πkT 1 + ω 2 t2 2 1 + ω 2 t2
The high-temperature limit of equation (7.2.15) is given by the first term
in the series since at high temperature the chemical potential is large and
negative.
7.17. The ground state density at the center of theptrap is N0 /(π 3 /2a3 ); see
problem 7.15. Using N0 /N = 1−(T /Tc )3 , a = ~/(mω), and kTc /(~ω) =
(N/ζ(3)1/3 ), we get

n(0)λ3 = 7ζ(3)1/2 N 1/2  1.

7.18. Integrating equation (7.2.15) gives


∞ 3 X∞
1 X eβµj (kT )3/2 eβµj
Z 
kT
nex (r)dr = 3 3
= .
λ j=1 j 3 m3/2 ω0 ~ω0 j=1
j3

The excited particles can be counted using the density of states and the
Bose-Einstein factor,
∞ 3 X∞
(kT )3 eβµj
Z Z 
1 2
X
−x βµj kT
Nex = a(ε) β(ε−µ) dε = x e e dx = .
e −1 2(~ω)3 j=1
~ω0 j=1
j3

Above Tc when µ < 0 this counts all of the particles. Below Tc when
µ = 0, this counts the particles that are not in the ground state.
61

7.19. The density of states for a two-dimensional harmonic oscillator is


a(ε) = ε/(~ω0 )2 so the number particles in the trap is given by
Z
ε 1
N (T, µ) = dε .
(~ω)2 eβ(ε−µ) − 1

As T → Tc , µ → 0 so
2 Z 2 2
π 2 kTc
Z   
ε 1 kTc xdx kTc
N = dε = = ζ(2) = .
(~ω)2 eβc (ε) − 1 ~ω0 ex − 1 ~ω0 6 ~ω0
p
so kTc = ~ω 6N/π 2 . The condensate fraction for T ≤ Tc is
N0 /N = 1 − (T /Tc )2 . For this two-dimensional theory to be valid, the
occupancy of √the first excited z-state
√ must be negligible which requires
~ωz  kTc ∼ N ~ω0 , i.e. ωz  N ω0 .

7.20. By eqn. (3.8.14),


−kT ln Q1 = kT ln(eβ~ω/2 − e−β~ω/2 ) = + kT ln(1 − e−β~ω ).
2
Now, concentrating on the thermal part alone and utilizing eqn. (7.3.2),
we get
Z∞
VkT
A(V, T ) ≡ −kT ln Q(V, T ) = 2 3 ln(1 − e−β~ω )ω 2 dω.
π c
0

After an integration by parts, we obtain


Z∞
V~ ω 3 dω π 2 Vk 4 T 4
A(V, T ) = − 2 3 = − ;
3π c eβ~ω−1 45~3 c3
0

cf. eqns. (7.3.17 and 18). We also get


 
∂A 4A
S=− =− and U = A + TS = −3A = 3PV .
∂T V T

Other results of Sec. 7.3 follow straightforwardly.


7.21. Using expressions (7.3.12) and (7.3.23), we readily get

U π4
= kT ' 2.7 kT .
N̄ 30ζ(3)

Note that the numerical factor appearing here is actually Γ(4)ζ(4)/Γ(3)ζ(3).


62

7.22. Since ω = 2πc/λ, the characteristic frequencies of the vibrational modes of


a radiation cavity (and hence the energy eigenvalues of these modes) are
proportional to L−1 , i.e. to V −1/3 . Just as in Problem 1.7, we infer that
the entropy of this system is a function of the combination (V 1/3 U ). It
then follows that during an isentropic process the quantity (V 1/3 U ) stays
constant, i.e.  
1 −2/3
V dV U + V 1/3 dU = 0.
3
Consequently, the pressure of the system is given by
 
∂U 1U
P ≡− = .
∂V S 3V

7.24. The number density of photons in the cosmic microwave background


(CMB) follows from equation (7.3.23)
 3
2ζ(3) kT
n= ' 4.10 × 108 m−3 ' 410. cm−3
π2 ~c
The energy density is
π 2 (kT )4
u= ' 4.17 × 10−14 J/m3 .
15 (~c)3
The entropy density is
3
4π 2 k kT

s= ' 1.48 × 109 k m−3 ' 2.04 × 10−14 J/m3 K.
45 ~c

7.25. According to Sec. 7.4,


Z   Z
∂ ~ω
CV (T ) = g(ω)dω, while CV (∞) = kg(ω)dω.
∂T e~ω/kT − 1
ω ω

It follows that
Z∞ Z  ∞

{CV (∞) − CV (T )}dT = kT − g(ω)dω.
e~ω/kT − 1 0
0 ω

It is easy to show that


~ω 1
lim ≈ kT − ~ω;
T →∞ e~ω/kT − 1 2
see Section 3.8 as well as Fig. 3.4 of the text. The integral on the right-
hand side then becomes
Z
1
~ω · g(ω)dω,
2
ω
63

which is indeed equal to the zero-point energy of the solid.


The physical interpretation of this result lies in noting that the actual
amount of heat required to raise the temperature of a solid is less than
the value predicted classically because the solid already possesses a finite
amount of energy even at T = 0K.

7.26. Using the Debye spectrum (7.4.15), we have for the zero-point energy of
the solid
ZωD
1 9N 9 9
~ω · 3 ω 2 dω = N ~ωD = Nk ΘD .
2 ωD 8 8
0

Indeed,
ω
RD
ω · ω 2 dω
0 3
ω̄ = ω = ωD
RD 4
ω 2 dω
0

and hence the mean energy per mode is equal to 21 ~ω̄ = 38 ~ωD = 38 kΘD .
7.27. We’ll show that if the entropy of a system is given by S = aVT n , where a
is a constant, then the quantity (CP − CV ) of that system is proportional
to T 2n+1 . For the Debye solid, at T  ΘD , this indeed is the case, the
parameter n being equal to 3. Hence the stated result.
We know that
       2
∂P ∂V ∂V ∂P
CP − CV = T = −T .
∂T V ∂T P ∂P T ∂T V

Since  
∂A
S≡− = aVT n ,
∂T V
we must have
A = −aVT n+1 /(n + 1) + f (V ),
where f (V ) is a function of V alone. It follows that

T n+1
 
∂A
P ≡− =a − f 0 (V ),
∂V T n+1

so that    
∂P 00 ∂P
= −f (V ), = aT n ;
∂V T ∂T V

clearly, f 00 (V ) must be non-negative. We thus get

−1 n 2 a2
CP − CV = −T · (aT ) = T 2n+1 .
f 00 (V ) f 00(V )
64

7.33. The specific heat of the system is given by the general expression (7.4.8),
which may in the present case be written as
ZωD
(~ω/kT )2 e~ω/kT
CV (T ) = k g(ω)dω. (1)
(e~ω/kT − 1)2
0

The mode density, g(ω), is given by the relation

g(ω)dω = 3 · V (4πp2 dp)/h3 ,

where p = ~k = ~(A−1 ω)1/s . It follows that

g(ω)dω = Cω (3/s)−1) dω [C = 3V /(2sπ 2 A3/s )]. (2)

Substituting (2) into (1) and introducing the variable x = ~ω/kT , we get
Z x0 (3/s)+1 x  
x e ~ωD
CV (T ) ∼ T 3/s dx x 0 = .
0 (ex − 1)2 kT

At low temperatures, the upper limit of this integral may be replaced by


infinity — making the integral essentially T -independent; this leads to the
desired result CV ∼ T 3/s .
7.34. The mode density in this case is given by, see eqn. (C.7b),

g(ω)dω ∼ k n−1 dk ∼ ω n−1 dω.

The rest of the argument is similar to the one made in the previous prob-
lem; the net result is that the specific heat of the given system, at low
temperatures, is proportional to T n .
It is not difficult to see that if the dispersion relation were ω ∼ k s and the
dimensionality of the system were n, then the low-temperature specific
heat of the system would be proportional to T n/s .
7.35. The Hamiltonian of this system is given by eqn. (7.4.6); the partition
function then turns out to be, see eqn. (3.8.14),

Y  −1
1
Q = e−βΦ0 2 sinh β~ωi ,
i
2

with the result that


X
A = −kT ln Q = Φ0 + kT ln{2 sinh(~ωi /2kT )}, and hence
i
   
∂A ∂Φ0 1 X ~ωi ∂ωi
P =− =− − ~ coth · .
∂V T ∂V 2 i 2kT ∂V
65

0
Recognizing that (i)
P the total
 vibrational energy U of this system is given
1
by the expression 2 ~ωi coth(~ωi /2kT ), see eqn. (3.8.20), and (ii) the
i
coefficient ∂ωi /∂V = −γωi /V , the expression for P may be written as

∂Φ0 U0
P = +γ (U 0 = U − Φ0 ); (1)
∂V V
see eqn. (7.4.7). With Φ0 (V ) = (V − V0 )2 /2κ0 V0 , eqn. (1) takes the form

V − V0 U0
P =− +γ . (2)
κ0 V0 V
Now, the coefficient of thermal expansion of any thermodynamic system
is given by
       
1 ∂V 1 ∂V ∂P ∂P
α≡ =− = κT , (3)
V ∂T P V ∂P T ∂T V ∂T V

where κT is the isothermal compressibility. In the present case, eqn. (2)


gives
U0
   0
−1 ∂P V ∂U
κT ≡ −V = +γ −γ ;
∂V T κ0 V0 V ∂V T
using the thermodynamic formula (∂U/∂V )T = T (∂P/∂T )v − P , where
U = Φ0 + U 0 , we get

U0
 
−1 V ∂P ∂Φ0
κT = +γ − γT + γP + γ .
κ0 V0 V ∂T V ∂V

Next, since  
∂P CV
=γ , (4)
∂V V V
we get
γ 2 TC V
 
V V − V0
κ−1
T = + (1 + γ) P + − (5)
κ0 V0 κ0 V0 V
Under the conditions of the problem, all terms on the right-hand side of
(5), except the first one, can be neglected; the term retained may also be
approximated by κ−1 0 — with the result that κT ≈ κ0 . Equations (3) and
(4) then lead to the desired result for α.
Finally, the quantity (CP − CV ) is given by

γ 2 κ0 TC 2V
     
∂P ∂V ∂P
CP − CV = T =T · αV ≈ .
∂T V ∂T P ∂T V V0
7
Note that, at low temperatures, (Cp − Cv ) ∼ T — as in Problem 7.27.
66

7.36. For rotons, ε = ∆ + (p − p0 )2 /2µ. Therefore, u ≡ dε/dp = (p − p0 )/µ.


Consequently,
1
P = nhp(p − p0 )/µi
3
R −(p−p )2 /2µkT
1 e 0
{p(p − p0 )/µ}p2 dp
= n R .
3 e−(p−p0 )2 /2µkT p2 dp

Substituting p = p0 + 2µkT x, we get
R −x2 √ 3
1 e p0 + 2µkT x (2kT /µ)1/2 x dx
P = n √ 2 .
3
R
e−x2 p0 + 2µkT x dx
As explained in Section 7.6, these integrals are well-approximated by
letting the range of x extend from −∞ to +∞; also remembering that

2µkT  p0 , we get
1/2 √ √
3p20 2µkT · ( π/2)

1 2kT
P ' n √
3 µ p20 · π
= nkT .

7.37. Following Secs. 7.5 and 7.6, the free energy A(v) of a roton gas in mass
motion is given by
Z
V
A(v) = −N̄ kT = −kT · 3 n(ε − v · p)2πp2 · sin θdpd θ.
h
As explained in Section 7.6, though rotons obey Bose-Einstein statistics,
their distribution function is practically Boltzmannian; see eqns. (7.6.6
and 7). We may, therefore, write
Z
V
A(v) = −kT · 3 e−βε+βvp cos θ 2πp2 sin θ dpd θ.
h
Integrating over θ, we get
Z∞
V sinh(βvp)
A(v) = −kT · 3 e−βε 4πp2 dp.
h βvp
0

Integration over p is now carried out the same way as in eqn. (7.6.9); with
appropriate approximation, we end up with the result
A(v) = A(0) sinh(βvp0 )/(βvp0 ).

Next, the inertial density of the roton gas is given by


Z
1 1
ρ(v) = · 3 n(ε − v · p)p cos θ 2πp2 sin θdpd θ
v h
Z
1
' e−βε+βvp cos θ 2πp3 cos θ sin θdpd θ
vh3
67

Integration over θ now gives


Z∞  
1 −βε cosh(βvp) sinh(βvp)
ρ(v) = e − 4πp3 dp
vh3 βvp (βvp)2
0
Z∞
β (βvp) cosh(βvp) − sinh(βvp)
= e−βε 4πp4 dp.
h3 (βvp)3
0

Finally, integrating over p (under appropriate approximation) and com-


paring the resulting expression with eqn. (7.6.19), we obtain
3{(βvp0 ) cosh(βvp0 ) − sin(βvp0 )}
ρ(v) = ρ(0) .
(βvp0 )3
7.38. We write eqn. (7.6.17) in the form
Z∞  
4π ∂n(p) 4 dp
ρ0 = − 3 p dp
3h ∂p dε
0

and integrate it by parts, to get


∞ Z∞
 
 
4π  dp d dp
ρ0 = − 3 n(p)p4 − n(p) p4

dp  .
3h dε 0 dp dε
0

The integrated part vanishes at both limits, and we are left with
Z∞  
4π d dp
ρ0 = 3 n(p) p4 dp.
3h dp dε
0

Comparing this with the standard result for the equilibrium number of
excitations in the system, viz.
Z∞
4πV
N̄ = 3 n(p)p2 dp,
h
0

we obtain for the effective mass of an excitation


  
ρ0 V 1 1 d 4 dp
meff = = p .
N̄ 3 p2 dp dε

For ideal-gas particles, ε = p2 /2m; the effective mass then turns out to be
precisely equal to m. For phonons, ε = pc; we then get
(meff )ph = 4 < ε > /3c2 ,
in agreement with eqn. (7.5.15). Unfortunately, in the case of rotons this
expression presents certain problems of analyticity at the point p = p0 ;
we then resort to direct calculation — leading to eqn. (7.6.19), whereby
(meff )rot ' p20 /3kT .
Chapter 8

8.1. Referring to Fig. 8.11 and noting that the slope of the tangent at the point
x = ξ is −1/4, the approximate distribution is given by

 1 0 ≤ x ≤ (ξ − 2)
f (x) = (ξ + 2 − x)/4 (ξ − 2) ≤ x ≤ (ξ + 2)
0 (ξ + 0) ≤ x,

where x = ε/kT and ξ = µ/kT . Accordingly,


Z∞ Z∞
2πV
N = g · 3 (2m)3/2 n(ε)ε 1/2
dε = C f (x)x1/2 dx ,
h
0 0

where C = g(2πV /h3 )(2mkT )3/2 . After some algebra, one gets
 
1 5/2 5/2 2 3/2 1 −2
N = C{(ξ +2) −(ξ −2) } = Cξ 1 + ξ + ... (ξ  1).
5 3 2
(1)
Comparing (1) with eqn. (8.1.24), which may be written as
2  εF 3/2
N= C ,
3 kT
we get ( )
 2
εF 1 kT
ξ= 1− + ... . (2)
kT 3 εF
Similarly,
Z∞
1
U = CkT f (x)x3/2 dx =
CkT {(ξ + 2)7/2 − (ξ − 2)7/2 }
35
0
 
2 5/2 5 −2
= CkT ξ 1 + ξ + ... . (3)
5 2
Combining (1) and (3), and then making use of (2), we get
 
3 3 5
U = NkT ξ 1 + 2ξ −2 + . . . = N εF 1 + (kT /εF )2 + . . . .

5 5 3

68
69

It follows that, at temperatures much less than εF /k,


CV = 2Nk (kT /εF ),
which is “correct” insofar as the dependence on T is concerned but is
numerically less than the true value, given by eqn. (8.1.39), by a factor of
4/π 2 .
The reason for the numerical discrepancy lies in the fact that the present
approximation takes into account only a fraction of the particles that are
thermally excited; see Fig. 8.11. In fact, the ones that are not taken into
account have a higher ∆ε than the ones that are, which explains why the
magnitude of the discrepancy is so large.
8.2. By eqns. (8.1.4) and (8.1.5), the temperature T0 is given by
 2/3  2 
N h
T0 = . (1)
gV f3/2 (1) 2πmk
At the same time, the Fermi temperature TF is given by, see eqn. (8.1.24),
2/3
h2

εF 3N
TF ≡ = . (2)
k 4πgV 2mk
It follows that  2/3
T0 4π 1
= . (3)
TF 3 f3/2 (1) π
Now, by eqn. (E. 16), f3/2 (1) = (1–2−1/2 )ζ(3/2) ' 0.765. Substituting
this into (3), we get: T0 /TF ' 0.989.
8.3. This problem is similar to Problem 7.4 of the Bose gas and can be done
the same way — only the functions gv (z) get replaced by fv (z).
To obtain the low-temperature expression for γ, we make use of expansions
(8.1.30–32), with the result
−2
5π 2 π2 π2
  
γ = 1+ (ln z)−2 + . . . 1− (ln z)−2 + . . . 1+ (ln z)−2 + . . .
8 24 8
2 2
 2
π π kT
=1+ (ln z)−2 + . . . ' 1 + .
3 3 εF
8.4. This problem is similar to Problem 7.5 of the Bose gas and can be done the
same way. To obtain the various low-temperature expressions, we make
use of expansions (8.1.30–32). Thus

−1
π2 π2
 
3
κT = 1− (ln z)−2 + . . . 1+ (ln z)−2 + . . .
2n(kT ln z) 24 8
π2
 
3
= 1− (ln z)−2 + . . . .
2n(kT ln z) 6
70

We now employ eqn. (8.1.35) and get


( 2 )−1 ( 2 )
π 2 kT π 2 kT
 
3
κT = 1− + ... 1− + ...
2nεF 12 εF 6 εF
( 2 )
π 2 kT

3
' 1− , (1)
2nεF 12 εF

which is the desired result.


Similarly, using appropriate expansions, we get
π2
 
3 −2
κs = 1− (ln z) + . . .
2n(kT ln z) 2
( 2 )
5π 2 kT

3
' 1− . (2)
2nεF 12 εF

Dividing (1) by (2), we obtain the low-temperature expression for γ, the


same as the one quoted in the previous problem; this also yields the desired
result for (CP − CV )/CV , which is simply (γ − 1).
8.6. This problem is similar to Problem 7.8 of the Bose gas and can be done
the same way. In the limit z → ∞, which corresponds to T → 0K,
w2 ≈ 2kT ln z/3m,
which tends to the limiting value 2εF /3m. Thus
w0 = (2εF /3m)1/2 .
For comparison,
√ the Fermi velocity uF = (2εF /m)1/2 . It follows that
w0 = uF / 3.
8.7. This problem is similar to Problem 7.9 of the Bose gas and can be done
the same way. At low temperatures, using formula (E. 17), we get
−2
π2 π2
 
−1 9 −2 −2
huihu i = 1+ (ln z) + . . . 1+ (ln z) + . . .
8 3 8
( 2 )
π2 2
  
9 9 π kT
= 1+ (ln z)−2 + . . . ' 1+ ;
8 12 8 12 εF

cf. Problem 6.6.


8.8. (i) Refer to eqns. (8.3.1 and 2) of the text and note that for silver ne =
1, na = 4, a = 4.09 Å, while m0 = me — giving εF = 5.49 eV and
TF = 6.37 × 104 K. For lead, ne = 4, na = 4, a = 4.95 Å, while
m0 = 2.1 me — giving εF = 9.45 eV and TF = 10.96 × 104 K. For
aluminum, ne = 3, na = 4, a = 4.05 Å, while m0 = 1.6 me — giving
εF = 11.63 eV and TF = 13.50 × 104 K.
71

(ii) The nuclear radius for 80 Hg200 is about 8.4 × 10−13 cm. Taking all
the nucleons together, this gives a particle density of about 8.06 ×
1037 cm−3 . Substituting this into eqn. (8.1.34), we get: εF = 3.7 ×
107 eV and TF = 4.3 × 1011 K.
(iii) For liquid He3 , the particle density is about 1.59 × 1022 cm−3 . This
yields an εF of about 4.1 × 10−4 eV and a TF of about 4.8 K.
8.9. By eqns. (8.1.4, 5 and 24), the Fermi energy εF is given by
2/3 2/3
h2 3π 1/2
 
3
εF = f3/2 (z) = f3/2 (z) kT .
4π 2mλ2 4

With the help of Sommerfeld’s lemma (E.17), this becomes


2/3
π2 7π 4

εF = kT ln z 1 + (ln z)−2 + −4
(ln z) + . . .
8 640
π2 π4
 
= kT ln z 1 + (ln z)−2 + −4
(ln z) + . . . . (1)
12 180

To invert this series, we write


(  2  4 )
kT kT
kT ln z ≡ µ = εF 1 + a2 + a4 + ... (2)
εF εF

and substitute into (1), to get


2 4 2  4 4
π2 π2
   
kT kT kT π kT
+ a22 − a4

1−a2 +. . . = 1+ + − a2 +. . . .
εF εF 12 εF 180 6 εF

Equating coefficients on the two sides of this equality, we get: a2 =


−π 2 /12, a4 = −π 4 /80, . . .. Equation (2) then gives the desired result
(8.1.35a).
Next, we have from eqns. (8.1.7) and (E.17)

5π 2 7π 4
 
U 3
= kT ln z 1 + (ln z)−2 − (ln z)−4 + . . .
N 5 8 384
 2 4
−1
π −2 7π −4
1+ (ln z) + (ln z) + . . .
8 640
π2 11π 4
 
3 −2 −4
= kT ln z 1 + (ln z) − (ln z) + . . . . (3)
5 2 120
72

Substituting from eqn. (8.1.35a) into (3), we get


( 2 4 )
π 2 kT π 4 kT
 
U 3
= εF 1 − − + ...+
N 5 12 εF 80 εF
( 2 4 )
π 2 kT π 4 kT
 
1+ − + ...
2 εF 120 εF
( 2 4 )
5π 2 kT π 4 kT
 
3
= εF 1 + − + ... . (4)
5 12 εF 16 εF

The specific heat of the gas is then given by


3
π 2 kT 3π 4 kT

CV
= − + .... (5)
Nk 2 εF 20 εF
We note that the ratio of the T 3 -term here to the Debye expression (7.3.23)
is (1/16)(ΘD /TF )3 . For a typical metal, this is O(10−8 –10−9 ).
8.10. This problem is similar to Problem 7.14 of the Bose gas and can be done
the same way.
Parts (i) and (ii) are straightforward. For part (iii), we have to show that
 s 2 C f
CP V (n/s)−1 (z) s  f(n/s)+1 (z)f(n/s)−1 (z)

=1+ = 1+ , (1)
CV n Nk fn/s (z) n {fn/s (z)}2
which can be done quite easily; see eqns. (7)–(9) of the solution to Prob-
lem 7.14. For part (iv), we observe that, since the quantity S/N is a
function of z only, an isentropic process implies that z = const. Accord-
ingly, for such a process,
VT n/s = const. and P/T (n/s)+1 = const.;
see eqns. (2) and (3) of the solution to Problem 7.14. Eliminating T among
these relations, we obtain the desired equation of an adiabat. For part (v),
we proceed as follows.
In this limit z → 0, eqn. (1) gives
CP /CV → 1 + (s/n). (1a)
For z  1, on the other hand, we obtain [see formula (E.17)]
CP
 n  n π2  n  n  π2 
= 1+ +1 (ln z)−2 + . . . 1+ −1 −2 (ln z)−2 + . . .
CV s s 6 s s 6
 2
−2
n n  π
× 1+ −1 (ln z)−2 + . . .
s s 6
π2 π2
=1+ (ln z)−2 + . . . ' 1 + (kT /εF )2 ,
3 3
regardless of the values of s and n.
73

8.11. For T  TF , we get


CV n CP − CV CP n 
' , ' 1, so that ' +1 .
Nk s Nk Nk s

For T  TF , we obtain [see formula (E.17)]


n π2  π2
   
CV n −2 n n
−2
= ln z 1 + (ln z) + . . . − ln z 1 + −1 (ln z) + . . .
Nk s s 3 s s 3
n π2 n π 2 kT
 
= (ln z)−1 + . . . ' .
s 3 s 3 εF

To this order of accuracy, the quantity CP /Nk has the same value as
Cv /Nk . As for the difference between the two, we obtain
3
n π 4 kT

CP − CV
' ,
Nk s 9 εF
consistent with the corresponding value of γ quoted in the previous prob-
lem. The non-relativistic case pertains to s = 2 while the extreme rela-
tivistic one pertains to s = 1.
8.12. For a Fermi gas confined to a two-dimensional region of area A,
A A AkT
N= f1 (zF ) = 2 ln(1 + zF ), EF = 2 f2 (zF ), (1a,b)
λ2 λ λ
while the corresponding results for the Bose gas are
A A AkT
N= g1 (zB ) = 2 ln(1 − zB ), EB = 2 g2 (zB ). (2a,b)
λ2 λ λ
Equating (la) and (2a), we get
1 zB zF
1 + zF = , i.e. zF = or zB = .
1 − zB 1 − zB 1 + zF
Next, since z∂f2 (z)/∂z = f1 (z),
Z zF Z zF  
1 1 1
f2 (zF ) = ln(1 + z)dz = + ln(1 + z)dz .
0 z 0 1+z z(1 + z)
The first part of this integral is readily evaluated; in the second part, we
substitute z = z 0 /(1 − z 0 ), to get
Z zF /(1+zF )
1 1 1
f2 (zF ) = ln2 (1+zF )− 0
ln(1−z 0 )dz 0 = ln2 (1+zF )+g2 (zB ).
2 0 z 2
Equations (1b) and (2b) then yield the desired result, viz.
N 2 h2
EF (N, T ) = + EB (N, T ), whence {CV (N, T )}F = {CV (N, T )}B .
4πmA
74

Letting T → 0, we recognize that the constant appearing in the above


result must be equal to EF (N, 0). To verify this, we note that, since the
Fermi momentum of the gas in two dimensions is given by the equation
N = A · πp2F /h2 , the Fermi energy is given by εF = p2F /2m = Nh 2 /2πmA.
The ground-state energy of the gas then follows readily:
Z pF 2
p A · 2πpdp A · πp4F N 2 h2 1
EF (N, 0) = 2
= 2 = = N εF .
0 2m h 4mh 4πmA 2

8.13. The Fermi energy of the gas is given by the obvious relation
Z εF
N= a(ε)dε. (1)
0

At the same time, the quantities N and U , as functions of µ and T , are


given by the standard integrals
Z ∞ Z ∞
a(ε)dε εa(ε)dε
N= β(ε−µ)
and U = β(ε−µ)
.
0 e +1 0 e +1

At low temperatures we employ formula (E.18), with x = βε and ξ = βµ,


to obtain
Z µ
π2
 
da(ε)
N= a(ε)dε + (kT )2 + ...
0 6 dε ε=µ
Z εF
π2
 
2 da(ε)
' a(ε)dε + (µ − εF )a(εF ) + (kT ) , (2)
0 6 dε ε=εF
Z µ
π2
 
da(ε)
U= εa(ε)dε + (kT )2 a(ε) + ε + ...
0 6 dε ε=µ
Z εF ( )
π2
 
2 da(ε)
' εa(ε)dε + (µ − εF )εF a(εF ) + (kT ) a(εF ) + εF .
0 6 dε ε=εF
(3)

Comparing (1) and (2), we obtain for the chemical potential of the gas

π 2 (kT )2 da(ε)
 
µ ' εF − , (4)
6 a(εF ) dε ε=εF

which leads to the desired result for µ.


Next, substituting (4) into (3), we obtain the remarkably simple expression

U ' U0 + (π 2 /6)k 2 T 2 a(εF ),

whence
CV ' (π 2 /3)k 2 T a(εF ). (5)
75

It follows that Z T
CV dT
S= ' (π 2 /3)k 2 T a(εF ). (6)
0 T
For a gas with energy spectrum ε ∝ ps , confined to a space of n dimensions,

a(ε)dε ∼ pn−1 dp ∼ ε(n/s)−1 dε.

By eqn. (1), the Fermi energy of the gas is given by


Z εF
sA n/s sεF
N= Aε(n/s)−1 dε = εF = a(εF ).
0 n n
Substituting this result into (5), we get

n π2
 
CV kT
' · ; (7)
Nk s 3 εF
cf. eqn. (8.1.39), which pertains to the case n = 3, s = 2. See also
Problem 8.11.
8.14. In the notation of Sec. 3.9, the potential energy of a magnetic dipole in
the presence of a magnetic field B = (0 , 0 , B) is given by the expression
−(gµB m)B, where m = −J, . . . , +J. The total energy ε of the dipole is
then given by ε = (p2 /2m0 ) − gµB mB , m0 being the (effective) mass of
the particle; the momentum of the particle may then be written as

p = {2m0 (ε + gµB mB )}1/2 .

At T = 0, the number of such particles in the gas will be


4πV
Nm = {2m0 (εF + gµB mB )}3/2
3h3
and hence the net magnetic moment of the gas will be given by
X 4πgµB V 0 3/2
X
M= (gµB m)Nm = (2m ) m(εF + gµB mB )3/2 .
m
3h3 m

We thus obtain for the low-field susceptibility (per unit volume) of the
system
  J
M 4πgµB 0 3/2 3 1/2
X
χ0 = Lim = (2m ) · gµB εF m2
B→0 VB 3h3 2
m=−J
2πg 2 µ2B 1/2
= (2m0 )3/2 εF J(J + 1)(2J + 1). (1)
3h3
By eqn. (8.1.24),

h3
 
3/2 3n N
εF = n= . (2)
4π(2J + 1) (2m0 )3/2 V
76

Substituting (2) into (1), we obtain the desired result


1 ∗2
µ∗2 = g 2 µ2B J(J + 1) .

χ0 = nµ /εF
2
With g = 2 and J = 1/2, we obtain: χ0 = (3/2)nµ2B /εF , in agreement
with eqn. (8.2.6).
The corresponding result in the limit T → ∞ is given by
1 ∗2
χ∞ = nµ /kT ;
2
see eqn. (3.9.26). We note that the ratio χ0 /χ∞ = 3kT /2εF , valid for all
J.

8.15. We note that the symbol µ0 (xN ) denotes the chemical potential (≡ kT ln z)
of an ideal gas of xN “spinless” (g = 1) fermions. The corresponding fu-
gacity z is determined by the equation

f3/2 (z) = xN λ3 /V. (1)

Differentiating (1) with respect to x, we get

∂f3/2 (z) ∂ ln z N λ3 1
= = f3/2 (z).
∂ ln z ∂x V x
It follows that
∂µ0 kT f3/2 (z)
= .
∂x x f1/2 (z)
Equation (8.2.20) then assumes the form stated in the problem.
At low temperatures, we get

nµ∗2 π2
 
3
χ= · 1− (ln z)−2 + . . .
kT 2 ln z 6
( 2 )−1 ( 2 )
3nµ∗2 π 2 kT π 2 kT
 
= 1− + ... 1− + ...
2εF 12 εF 6 εF
( 2 )
π 2 kT

' χ0 1 − . (8.2.24)
12 εF

At high temperatures, on the other hand,

nµ∗2 z − 2−1/2 z 2 + . . . nµ∗2


χ= = (1 − 2−3/2 z + . . .)
kT z − 2−3/2 z 2 + . . . kT
' χ∞ (1 − 2−5/2 nλ3 ), (8.2.27)

where use has been made of eqn. (1), with f3/2 (z) ' z and x = 1/2.
77

8.18. The ground-state energy of a relativistic gas of electrons is given by


Z pF
8πV
E0 = 3 mc 2 [{1 + (p/mc)2 }1/2 − 1]p2 dp.
h 0

Making the substitution (8.5.9), we get


Z θF
8πm4 c5 V
E0 = (cosh θ − 1) sinh2 θ cosh θ dθ. (1)
h3 0

Now the integral


Z θF θF Z θF
1 −1
sinh2 θ cosh2 θ dθ = sinh3 θ cosh θ sinh4 θ dθ.

(2)
0 3
0 3 0
Substituting (2) into (1) and making use of eqn. (8.5.12), we get
8πm4 c5 V 8πm4 c5 V
E0 = 3
sinh3 θF cosh θF − P0 V − sinh3 θF ; (3)
3h 3h3
note that the last term is simply Nmc 2 . Finally, using the definition
x = sinh θF , we obtain the desired result.
We observe that eqn. (3) can also be written as

E0 + P0 V = Nmc 2 (cosh θF − 1) = N εF ≡ N µ0 .

To verify that the derivative (∂E0 /∂V )N is equal to −P0 , we have to


show that

[∂{VB (x)}/∂V ](Vx 3 ) = −A(x), i.e. ∂{x−3 B(x)}/∂x−3 = −A(x), i.e.



x4 [8{(x2 + 1)1/2 − 1} − x−3 A(x)] = 3A(x), i.e.
∂x
∂A(x)/∂x = 8x4 (x2 + 1)−1/2 ,

which can be readily verified with the help of expression (8.5.13).


8.19. Utilizing the result obtained in Problem 8.13, we have for a Fermi gas at
low temperatures
CV π 2 a(εF )
= kT . (1)
Nk 3 N
Now, the density of states for the relativistic gas is given by, see eqn. (8.5.7),

8πV dp 8πmV
  p 2 1/2
a(ε) = 3 p2 = p 1 + ,
h dε h3 mc
where p = p(ε). Substituting this result into (1) and making use of
eqn. (8.5.4), we get
CV π2 m n  p o1/2
F
= 2 1+ kT ,
Nk pF mc
78

which leads to the desired result.



In the non-relativistic case pF  mc and εF = p2F /2m , we obtain the fa-
miliar expression (8.1.39); in the extreme relativistic case (pF  mc and ε =
pc), we obtain  
CV kT
= π2 ,
Nk εF
consistent with expression (7) of the solution to Problem 8.13.

8.22. The number of fermions in the trap is


Z εF
dε ε2 dε ε2 ε3F
Z
1
N (T, µ) = 3
= 3
= .
2(~ω) e β(ε−µ) −1 0 2(~ω) 6(~ω)3

Using kTF = εF this gives the following relation for the fugacity z = e−βµ ,
3 Z
x2 dx

T
3 = 1.
TF ex e−βµ + 1

The internal energy is

dε ε3 (kT )4 x3
Z Z
1
U (T, µ) = = .
2(~ω)3 eβ(ε−µ) − 1 2(~ω)3 ex e−βµ − 1

When compared to the ground state energy U0 = (kTF )4 /[8(~ω)3 ], we get


4 Z
x3

U T
=4 .
U0 TF ex e−βµ −1
Chapter 9

9.1. Using the Friedmann equation (9.1.1)


r
da 8πGu
= a,
dt 3c2
and the connection between scale factor a and blackbody temperature T ,
T a = T0 a0 , along with (9.3.4b) we get
r r
dT 8πGu 8π 3 Ggk 4 3
=− 2
T =− T ,
dt 3c 45~3 c5
where g = 43/8 is the effective number of relativistic species from equation
(9.3.6b). The solution of the differential equation is
r
t0
T (t) = T0 ,
T
where
s
1 45~3 c5
t0 = ' 0.99 s
2 8π 3 Gg(kT0 )4

for the case of T0 = 1010 K.


9.2. Just use equations (9.3.4) and (9.3.6) with T = 1010 K. The pressure
3
and energy density are of order 1025 J/m , and the number density and
entropy divided by k are of order 1038 m−3 .
9.3. The average kinetic energy per relativistic electron/positron is of the order
of ue /ne ∼ kT . The Coulomb energy per electron/positron is of the order
of uc ≈ e2 /(4π0 a) where a ≈ (1/ne )1/3 is of the order of the average
distance between the charged particles. Using ne ∼ (kT /~c)3 we get
uc /ue ∼ e2 /(4π0 ~c) ≈ 1/137. This is the justification for treating the
relativistic electrons and positrons as noninteracting.
9.4. Correction to the first printing of third edition: The exponent in the
result should be −3/2. For βmc2  1 but before the time when the

79
80

electron density approaches the protron density, the density of electrons


and positrons are almost identical so µ ≈ 0. Equation (9.5.6) gives
Z ∞ p p
n− n+ 1 x x + βmc2 x − βmc2 dx
≈ ≈
nγ nγ ζ(3) βmc2 ex + 1
2 3/2 Z ∞
e−βmc βmc2 √ −y
≈ ye dy.
2ζ(3) 0

9.5. Correction to the first printing of third edition: The exponent in the
result should be −3/2. After the density of electrons levels off at the
nearly the proton density, you can use equation (9.5.8) to show that the
chemical potential µ− ≈ mc2 . Then the positron number density is given
by equation (9.5.7),
Z ∞ p p
n+ 1 x x + βmc2 x − βmc2 dx

nγ ζ(3) βmc2 ex+βmc2 + 1
2 3/2 Z ∞
e−2βmc βmc2 √ −y
≈ ye dy
2ζ(3) 0
2 3/2 √
e−2βmc βmc2 π
≈ .
4ζ(3)

9.6. Correction to the first printing of third edition: the energy density in the
statement of the problem should read

utotal = (1 + (21/8)(4/11)4/3 )uγ .

After the electron–positron annihilation, the only relativistic species left


are the photons and the neutrinos. The factor 21/8 = (3)(1)(7/8) in the
energy is because there are three families of neutrinos, the spin degeneracy
factor is 1 (all left handed), and 7/8 is the Fermi-Dirac factor. The factor
(4/11)4/3 is due to the lower temperature of the neutrinos compared to
the photons; see equation (9.6.4). Following the solution to problem 9.1,
we get
v
1u 45~3 c5
t0 = t ' 1.79 s.
u  
2 8π 3 G 1 + 21  4 4/3 (kT )4
8 11 0

9.7. If the current CMB temperature was 27K rather than 2.7K, the baryon-
to-photon ratio would be 103 times smaller. Equation (9.7.8) implies that
the nucleosynthesis temperature would have been about 20% lower which
would have delayed the nucleosynthesis by an extra two minutes. This
would have given the neutrons a longer time to decay leading to q ≈ 0.10
rather than 0.12, leading to a helium content in the universe of about 20%
81

by weight. If the current CMB temperature were 0.27K, that would have
increased the baryon-to-photon ratio by a factor of 103 . Fewer photons per
baryon would have led to an earlier nucleosynthesis, less time for neutrons
to decay and an increase of the neutron fraction to q ≈ 0.135 leading to
about 27% helium content.

9.8. The strong interaction exhibits asymptotic freedom at high energies jus-
tifying treating the quarks an gluons as noninteracting. The effective
number of species in equilibrium in these tiny quark–gluon plasmas is
accounted for using only the up and down quarks and the gluons. Pho-
tons, and leptons, for example, easily escape without interacting with the
plasma.
   
7 uγ 7 uγ
uu = 2 uū = 2 up quarks and antiquarks
8 2 8 2
   
7 uγ 7 uγ
ud = 2 ud¯ = 2 down quarks and antiquarks
8 2 8 2

ug = (8)2 gluons
2
Therefore, the effective number of species is g = 8 + 28/8 = 23/2 and
3 3
uQGP = guγ . The energy density is 4 GeV/fm = 6.4 × 1035 J/m , so
1/4
15(~c)3 GeV
 
kT ' 4 ' 4 × 10−11 J ' 250 MeV,
gπ 2 fm3

and T ' 3 × 1012 K. This is the record hottest temperature for matter
created in the laboratory.

9.9. The strong interaction exhibits asymptotic freedom at high energies jus-
tifying treating the quarks an gluons as noninteracting. The effective
number of species is much larger than during the time near t = 1s due to
82

the muons, quarks and gluons.



uγ = 2 photons
2   
7 uγ 7 uγ
ue− = 2 ue+ = 2 electrons/positrons
8 2 8 2
   
7 uγ 7 uγ
uνe = uν̄e = electron neutrinos/antineutrinos
8 2 8 2
   
7 uγ 7 uγ
uνµ = uν̄µ = muon neutrinos/antineutrinos
8 2 8 2
   
7 uγ 7 uγ
uντ = uν̄τ = tau neutrinos/antineutrinos
8 2 8 2
   
7 uγ 7 uγ
uµ− = 2 uµ+ = 2 muons/antimuons
8 2 8 2
   
7 uγ 7 uγ
uu = 2 uū = 2 up quarks/antiquarks
8 2 8 2
   
7 uγ 7 uγ
ud = 2 ud¯ = 2 down quarks/antiquarks
8 2 8 2

ug = (8)2 gluons
2

The result is u = (149/8)uγ . Proceeding as in problem 9.1 we get


r
10 0.53 s
T (t) = 10 K .
T

Therefore at kT = 300 MeV (T ' 3.5 × 1012 K), the age of the universe
was about 4 × 10−6 s.
Chapter 10

10.1. By eqn. (10.2.3), the second virial coefficient of the gas with the given
interparticle interaction would be
"Z Z ∞ #
D
2π 6
a2 = − 3 −1 · r2 dr + {eε(σ/r) /kT − 1}r2 dr
λ 0 D
 
Z ∞X ∞ j
εσ 6

2π  1 3 1
= 3 D − r2 dr 
λ 3 r0 j=1 j! kT r6
 
∞ j
2πD3  εσ 6

X 1
= 1− ;
3λ3 j=1
(2j − 1)j! kT D 6

cf. eqn. (10.3.6). For the rest of the question, follow the solution to
Problem 10.7.

10.2. For this problem, we integrate (10.2.3) by parts and write


Z ∞
2π ∂u(r) 3
a2 λ3 = − e−u(r)/kT r dr ;
3kT 0 ∂r

cf. eqn. (3.7.17) and Problem 3.23. With the given u(r), we get
Z ∞  
2π m n mA nB
a2 λ3 = e−A/kT r eB/kT r − dr
3kT 0 rm−2 rn−2
Z ∞ ∞  j  
2π −A/kT r m
X 1 B mA nB
= e − n−2+nj dr
3kT 0 j=0
j! kT rm−2+nj r
∞  j (   (m−3+nj )/m   (n−3+nj )/m )
2π X 1 B m − 3 + nj kT n n − 3 + nj kT
= AΓ − BΓ .
3kT j=0 j! kT m A m m A

From the first sum we take the (j = 0)-term out and combine the remain-
ing terms with the second sum (in which the index j is changed to j − 1);

83
84

after considerable simplification, we get


 
 3/m    ∞  "  n/m #j 
2π A m−3 3 X 1 nj − 3 B kT
a2 λ3 = Γ − Γ .
3 kT  m m j=1 j! m kT A 
(1)
For comparison with other cases, we set A = A0 r0m
and B = B 0 r0n
(so
that A0 and B 0 become direct measures of the energy of interaction).
Expression (1) then becomes
 
 0 3/m    ∞  " 0  n/m #j 
2π A m − 3 3 X 1 nj − 3 B kT
a2 λ3 = r3 Γ − Γ .
3 0 kT  m m j=1 j! m kT A0 
(2)
Now, to simulate a hard-core repulsive interaction, we let m → ∞, with
the result that
 
∞  0 j 
2π  X 1 B
a2 λ3 = r3 1 − 3 . (2a)
3 0 (nj − 3)j! kT 
j=1

With n = 6, expression (2a) reduces to the one derived in the preceding


problem. Furthermore, if terms with j > 1 are neglected, we recover the
van der Waals approximation (10.3.8).
For further comparison, we look at the behavior of the coefficient B2 (≡
a2 λ3 ) at high temperatures. While the hard-core expression (2a) predicts
a constant B2 as T → ∞, the soft-core expression (2) predicts a B2 that
ultimately vanishes, as T −3/m , which agrees qualitatively with the data
shown in Fig. 10.2.
10.3. (a) Using the thermodynamic relation
CP − CV = T (∂P/∂T )V (∂V /∂T )P = −T (∂P/∂T )2V /(∂P/∂V )T
and the equation of state (10.3.9), we get
CP − CV T (∂P/∂T )2v T {k/(v − b)}2 1
=− =− = .
Nk k(∂P/∂v)T k{−kT /(v − b2 ) + 2a/v3 } 1 − 2a(v − b)2 /kT v3
(b) In view of the thermodynamic relation
TdS = CV dT + T (∂P/∂T )V dV
and the equation of state (10.3.9), an adiabatic process is character-
ized by the fact that
CV dT + NkT (v − b)−1 dv = 0.
Integrating this result, under the assumption that CV = const., we
get
T CV /Nk (v − b) = const.
85

(c) For this process we evaluate the Joule coefficient

a/v2 N 2a
 
∂T (∂U/∂V )T T (∂P/∂T )V − P
= =− =− =− .
∂V U (∂U/∂T )V CV CV CV V 2

Now integrating from state 1 to state 2, we readily obtain the desired


result.
10.4. Since, by definition,

α = v−1 (∂v/∂T )P and B −1 ≡ κT = −v−1 (∂v/∂P )T ,

we must have:
[∂(αv)/∂P ]T = −[∂(vB −1 )/∂T ]P . (1)

Using the given empirical expressions, we obtain for the left-hand side of
(1)
vB −1 a0
     
∂(αv) 1 ∂v 1
= =− =− v+ 2
∂P T T ∂P T T PT T
and for the right-hand side

∂(vB −1 ) 2a0 2a0 a0


        
1 ∂v 1 1 v
− =− − 3 =− αv − 3 = − + 3 .
∂T P P ∂T P T P T P T T

The compatibility of the given expressions is thus established.


To determine the equation of state of the gas, we note from the given
expression for α that

3a0 3a0
   
∂v v ∂ v
= + 3 , i.e. = 4,
∂T P T T ∂T T P T

whence
v a0 a0
= − 3 + f (P ), i.e. v = − 2 + Tf (P ), (2)
T T T
where f is a function of P only. We then obtain for B

vB −1 = −Tf 0 (P ). (3)

Combining (2) and (3), we get

f 0 (P ) vB −1 1
=− =− .
f (P ) (v + a0 /T )2 P

It follows that f (P ) is proportional to 1 /P and hence, by (2),

P = const. T (v + a0 /T 2 )−1 .
86

10.5. The Joule-Thomson coefficient of a gas is given by


        
∂T (∂H/∂P )T 1 ∂V N ∂v
=− = T −V = −v .
∂P H (∂H/∂T )P CP ∂T P CP ∂T P

By eqn. (10.2.1),

Pv a2 λ3
=1+ + . . . , so that
kT v
a2 λ3 P

kT kT
v= 1+ + ... = + a2 λ3 + . . . .
P kT P
It follows that
∂(a2 λ3 )
   
∂v
T −v= T − a2 λ3 + . . .
∂T P ∂T

and hence the quoted result for (∂T /∂P )H .


With the given interparticle interaction, eqn. (10.2.3) gives
"Z #
D Z r1
3 2 u0 /kT 2
a2 λ = −2π −1 · r dr + (e − 1)r dr
0 D
2π h 3 i
D − r13 − D3 eu0 /kT ,

=
3
whence
∂(a2 λ3 ) 2π h 3  u0  u0 /kT i
T − a2 λ3 = r1 − D3 1 + e − r13 .
∂T 3 kT
The desired result for (∂T /∂P )H now follows readily.
We note that the Joule-Thomson coefficient obtained here vanishes at
a temperature T0 , known as the temperature of inversion, given by the
implicit relationship
r3
 
u0
1+ eu0 /kT 0 = 3 1 3 .
kT 0 r1 − D

For T < T0 , (∂T /∂P )H > 0, which means that the Joule-Thomson ex-
pansion causes a cooling of the gas. For T > T0 , (∂T /∂P )H < 0; the
expansion now causes a heating instead.
10.7. To the desired approximation,
P 1 1 N 1
≡ ln Q = 3 (z − a2 z 2 ), n= = 3 (z − 2a2 z 2 ), (1a,b)
kT V λ V λ
where a2 is the second virial coefficient of the gas. It follows that

z = nλ3 (1 + 2a2 · nλ3 ), whence P = nkT (1 + a2 · nλ3 ). (2a,b)


87

Next

A = NkT ln z − PV = NkT {ln(nλ3 ) − 1 + a2 · nλ3 },


G = NkT ln z = NkT {ln(nλ3 ) + 2a2 · nλ3 },
   
∂A 5 3 ∂ 3
S=− = Nk − ln(nλ ) − n (Ta 2 λ ) ;
∂T N,V 2 ∂T

remember that the coefficient a2 is a function of T . Furthermore,


 
3 ∂
U = A + TS = NkT − nT (a2 λ3 ) ,
2 ∂T
a2 λ3
  
5 ∂
H = U + PV = NkT − nT 2 ,
2 ∂T T
    
∂U 3 ∂ ∂
CV = = Nk −n T2 (a2 λ3 ) , and
∂T N,V 2 ∂T ∂T
(∂P/∂T )2N,V
 

CP − CV = −T = Nk 1 + 2nT (a2 λ3 ) .
(∂P/∂V )N,T ∂T

For the second part, use the expression for a2 λ3 derived in Problem 10.5
and examine the temperature dependence of the various thermodynamic
quantities.

10.8. We consider a volume element dx 1 dy 1 dz 1 around the point P (x1 , 0, 0) in


solid 1 and a volume element dx 2 dy 2 dz 2 around the point Q(x2 , y2 , z2 ) in
solid 2. The force of attraction between these elements will be
`5
−α(n dx 1 dy 1 dz 1 )(n dx 2 dy 2 dz 2 ) 5/2
,
{(x2 − x1 )2 + y22 + z22 }

directed along the line joining the points P and Q. The normal component
of this force will be
5 (x2 − x1 )
−αn2 (dy 1 dz 1 )` 3 dx 1 dx 2 dy 2 dz 2 .
{(x2 − x1 )2 + y22 + z22 }

The net force (per unit area) experienced by solid 1, because of attraction
by all the molecules of solid 2, will thus be
Z 0 Z ∞ Z ∞
(x2 − x1 )
− αn2 `5 dx 1 dx 2 · 2πρdρ
2 2 3
x1 =−∞ x2 =d ρ=0 {(x2 − x1 ) + ρ }
Z ∞
παn2 `5 0 παn2 `5
Z
1
=− 3
dx 1 dx 2 = ,
2 x1 =−∞ x2 =d (x2 − x1 ) 4d

i.e. inversely proportional to d.


88

10.9. For x  1, the spherical Bessel function j` (x) behaves like x` /1.3 . . . (2` +
1) while the spherical Neumann function behaves like −1.3 . . . (2`−1)/x`+1 ;
see Abramowitz and Stegun (1964). Substituting these results into eqn. (10.5.31),
we readily obtain the desired result.
10.10. The symmetrized wave functions for a pair of non-interacting bosons/fermions
are given by
1
Ψα (r1 , r2 ) = √ (eik1 ·r1 eik2 ·r2 ± eik1 ·r2 eik2 ·r1 ).
2V

The probability density operator Ŵ2 of the pair is then given through the
matrix elements
X
h10 , 20 |Ŵ2 |1, 2i = 2λ6 Ψα (10 , 20 )Ψ∗α (1, 2)e−βEα
α
λ6 X ik1 ·r01 ik2 ·r02 0 0
= 2 (e e ± eik1 ·r2 eik2 ·r1 )×
V α
2 2 2
(e−ik1 ·r1 e−ik2 ·r2 ± e−ik1 ·r2 e−ik2 ·r1 )e−β~ (k1 +k2 )/2m
" 0 0 0 0
#
λ6 X X eik1 ·(r1 −r1 ) eik2 ·(r2 −r2 ) + eik1 ·(r2 −r2 ) eik2 ·(r1 −r1 ) ±
= 0 0 0 0
2V 2
k1 k2 eik1 ·(r2 −r1 ) eik2 ·(r1 −r2 ) ± eik1 ·(r1 −r2 ) eik2 ·(r2 −r1 )
2 2 2 2
e−β~ k1 /2m
e−β~ k2
/2m
1 0
= [h1 |Ŵ1 |1ih20 |Ŵ1 |2i + h20 |Ŵ1 |2ih10 |Ŵ1 |1i±
2
h20 |Ŵ1 |1ih10 |Ŵ1 |2i ± h10 |Ŵ1 |2ih20 |Ŵ1 |1i]
= h10 |Ŵ1 |1ih20 |Ŵ1 |2i ± h20 |Ŵ1 |1ih10 |Ŵ1 |2i.

Comparing this with eqn. (10.6.18), we obtain the desired result.


10.11. A particle with spin J can be in any one of the (2J + 1) spin states
characterized by the spin functions χm (m = −J, . . . , J). For a pair of
such particles, we will have (2J + 1)2 spin states characterized by the
symmetrized spin functions

χm1 (1)χm2 (2) ± χm1 (2)χm2 (1) (m1,2 = −J, . . . , J).

Of these, (2J + 1) functions, for which m1 = m2 , can only be symmetric,


for the corresponding antisymmetric combinations vanish identically. The
remaining 2J(2J + 1) functions, for which m1 6= m2 , can be symmetric
or antisymmetric; however, only half of them are linearly independent
functions (because an interchange of the suffices m1 and m2 does not
produce anything new). Thus, in all, we have J(2J + 1) antisymmetric
spin functions, and (J + 1)(2J + 1) symmetric spin functions, that are
linearly independent.
89

Now the total wave function of the pair will be the product of a sym-
metrized space function (like the ones considered in the previous problem)
and a symmetrized spin function (like the ones discussed above). For the
total wave function to be symmetric, as required for a pair of bosons, we
may associate any of the (J + 1)(2J + 1) symmetric spin functions with
a symmetric space function or any of the J(2J + 1) antisymmetric spin
functions with an antisymmetric space function. This will lead to the
quoted expression for the coefficient bs2 . On the other hand, for the total
wave function to be antisymmetric, as required for a pair of fermions, we
may associate any of the J(2J + 1) antisymmetric spin functions with a
symmetric space function or any of the (J + 1)(2J + 1) symmetric spin
functions with an antisymmetric space function. This will lead to the
quoted expression for the coefficient bA
2.

10.12. To derive the desired results, we make the following observations:

(i) Since a pair of particles with spin J has (2J + 1)2 possible spin states
while a pair of spinless particles has only one, we have to divide the
expression for b2 pertaining to the former by (2J + 1)2 so that we are
talking of the average contribution per state.
(ii) To make a transition from discreteness in orientation (that is associ-
ated with a finite value of J) to continuity in orientation, we should
take the limit J → ∞.
(iii) In view of the foregoing, the distinction between the original system
being symmetric or antisymmetric is completely lost, and we are led
to the results quoted in the problem.

Next, using eqns. (10.5.28, 36 and 37), we obtain for the quantum-mechanical
Boltzmannian gas
1 3 5
22π 2
  
D D D
b2 = − − 3π + + ...,
λ λ 3 λ

which differs significantly from the corresponding classical result.


10.13 Expand the definition of the pair density n2 (r, r 0 ) in powers of the fugacity
z using the grand canonical partition function and the Mayer functions
fij = exp (−βu(rij )) − 1.

zN
Z
1 X
n2 (r12 ) = dr 3 · · · dr N exp (−βu(r12 ) − βu(r13 ) − · · · )
Q(µ, V, T ) (N − 2)!
N =2
 Z 
= e−βu(r12 ) z 2 + z 3 (1 + f13 + f23 + f13 f23 ) dr 3 − z 3 Q1 + · · ·
 Z  Z 
= e−βu(r12 ) z 2 + 2z 3 f (r)dr + z 3 f13 f23 dr 3 + · · ·
90

Note every term includes the factor e−βu(r12 ) . The coefficients of those
terms are integrals over the Mayer functions that are continuous functions
of r12 even for the infinite step function potential; see equation (10.3.19)
and discussion in Hansen and McDonald (1986) Chapter 5.
10.14 The pressure is given by
Z
P n du
=1− rg(r) dr,
nkT 2dkT dr
where g(r) = y(r)e−βu(r) . This gives
Z Z
P n du −βu(r) n d  −βu(r) 
=1− ry(r) e dr = 1 + ry(r) e dr,
nkT 2dkT dr 2d dr
For the case of hard spheres,
d  −βu(r) 
e = δ(r − D),
dr
so
P nDd
=1+ Ωd y(D).
nkT 2d
where Ωd is the area of the d-dimensional unit sphere. For hard spheres
y(D) = g(D+ ). In three dimensions η = πnD3 /6 and Ω3 = 4π, so
P/(nkT ) = 1 + 4ηg(D+ ).
10.15 Let P (r) be the cumulative probability that no particles are closer than
r to a given particle. Breaking up the interval between zero and r into
small intervals starting ar rk = k∆r with width ∆r gives
r/∆r
Y
1 − 4πng(rk )rk2 ∆r ,

P (r) =
k=0

since each factor represents the probability there are no neighbors in in-
terval k. This gives
r/∆r r/∆r
X X
ln 1 − 4πng(rk )rk2 ∆r ≈ − 4πng(rk )rk2 ∆r.

ln (P (r)) ≈
k=0 k=0

Therefore
 Z r 
P (r) = exp −4πn r2 g(r)dr .
0

Finally
dP
w(r) = − .
dr
For an ideal gas g(r) = 1, so the integrals are easily evaluated.
91

10.17 & 10.18. For a complete solution to these problems, see Landau and Lifshitz (1958),
sec. 117, pp. 369–74.
10.19. (a) In this problem we are concerned with the integral

Z∞
∂u −βu 3
I= e r dr .
dr
0

Integrating by parts, we get


Z∞
1 ∞ 1
I = − e−βu + c r3 |0 + e−βu + c 3r2 dr .
 
β β
0

An arbitrary constant c has been introduced here to secure “proper


behavior” at r = ∞. Since exp(−βu) → 1 as r → ∞, we choose
c = −1. The integrated part then vanishes [assuming that u(r) → 0
faster than 1/r3 ], and we are left with the result
Z∞
3  −βu
− 1 r2 dr .

I= e
β
0

This reduces eqn. (10.7.11) to the desired form.


(b) In the case of hard-sphere potential, the function f (r) = −1 for r ≤ σ
and 0 for r > σ. We then get

2πnσ 3
 
PV N
'1+ n= .
NkT 3 V

For nσ 3  1, we may write this result in the approximate form

2πnσ 3
 
PV 1 − = NkT .
3

Comparison with Problem 1.4 shows that the parameter b of that prob-
lem is equal to (2π/3)N σ 3 , which is indeed four times the actual space
occupied by the particles.
10.20 Use

−1 ∂p
[κT (n, T )] =n
∂n
 T
∂f
P (n, T ) = n2
∂n T
92

where f = A/N is the Helmholtz free energy per particle. Then


Z n
dn0
P (n, T ) = p(n0 , T ) + 0 0
,
n0 n κ(n , T )
Z n
P (n0 , T ) 0
f (n, T ) = f (n0 , T ) + dn .
n0 (n0 )2

10.21 The most general Gaussian distribution of variables {u1 , · · · , uN } is of the


form
 
1
P (u1 , · · · , uN ) ∼ exp − uT Au
2

where A is a symmetric positive definite matrix. The matrix has only pos-
itive eigenvalues and can be diagonalized into diagonal matrix (Ã = U T =
U −1 using the orthogonal the matrix U . The eigenvalues {λ1 , · · · , λN }
are all positive and detU = 1. The normalization is
Z   Z  
1 1
N = dN u exp − uT Au = dN y exp − y T Ãy
2 2
v
u
u N
Y q
= t(2π)N λi = (2π)N detA .
i=1

The transformed variables are y = U T u so the Jacobian is unity. The


integral of the average of exp aT u can be determined from completing
the square inside the exponential,
Z    
1 1 T −1
dN u exp aT u − uT Au = N exp a A a .
2 2
2
Averaging the quantity aT u is accomplished by transforming to the y
variables which, using
Z
dx x2 exp(−λx2 /2) = (2π)1/2 /λ3/2 ,

gives
Z  
2 1
N T
d u a u exp − uT Au = N aT U Ã−1 U T a = N aT A−1 a.
2

10.22 The pressure is given by


   
∂A 2 ∂A/N
p=− =n ,
∂V N,T ∂n T
93

and the excess pressure is given by

1 + η + η2 − η3 4η − 2η 2
   ex 
ex 2 ∂A /N
P = Pcs − Pideal = nkT − 1 = nkT =n .
(1 − η)3 (1 − η)3 ∂n T

This can be integrated to give


Z η
βAex 4 − 2η 0 0 3 − 2η 4η − 3η 2
= 0 2
dη = − 3 = .
N 0 (1 − η ) (1 − η)2 (1 − η)2

10.23 The simplest rational approximations are

P 1 + η/8
= ≈ 1 + 2η + 3.125η 2 + 4.25η 3 + 5.375η 4 + 6.5η 5
nkT (1 − η)2
+ 7.625η 6 + 8.75η 7 + 9.875η 8 + 11.000η 9 + · · · ,

and
P 1 + 0.128018 η
= ≈ 1 + 2η + 3.128018η 2 + 4.256036η 3 + 5.384054η 4 + 6.512072η 5
nkT (1 − η)2
+ 7.64009η 6 + 8.768108η 7 + 9.896126η 8 + 11.024144η 9 + · · ·

The later gets the first two orders exactly correct, and the third and fourth
order coefficients correct to better that 1%.
Chapter 11

11.4. The relevant results for T < Tc are given in eqns. (11.2.13–15). The
corresponding results for T > Tc follow from eqn. (11.2.10) by neglecting
n0 altogether; we get, to the first order in a,
1 1 4πa~2
A(N, V, T ) = Aid (N, V, T ) + , (13a)
N N mv
4πa~2
P = Pid + , (14a)
mv2
8πa~2
µ = µid + . (15a)
mv
Remembering that vc ∝ T−3/2 , the various quantities of interest turn out
to be
(
0  (T > Tc )
 2 
∂ A 2πa~2 
CV = −T = (CV )id + N 3 6v
∂T 2 N,V mT − 2vc + v2 (T < Tc )
c
(
2πa~2 4/v2 (T > Tc )
 
∂P
K = −v = Kid +
∂v T m 2/v2 (T < Tc )
 2   2  (
∂ P ∂ P 2πa~2 0 (T > Tc )
2
= 2
+ 2 2
∂T v ∂T v,id mT 6/vc (T < Tc )
 2   2  (
∂ µ ∂ µ 4πa~2 0 (T > Tc )
= +
2
∂T v 2
∂T v,id mT 2 3/4vc (T < Tc )
The thermodynamic relationship quoted in part (b) of the problem is
readily verified.
As for the discontinuities at T = Tc , we get (setting v = vc )
9πa~2 1 9aλ2c ζ(3/2) 9a
∆CV = N = Nk 3
= Nk ζ(3/2),
mT c vc 2 λc 2λc
4πa~2 1
∆K = − ,
m vc2
 2 
12πa~2
 2 
∂ P ∂ µ 3πa~2
∆ = , ∆ = .
∂T 2 v mT 2c vc2 ∂T 2 v mT 2c vc

94
95

11.5. (a) We replace the sum over p appearing in eqn. (11.3.14) by an integral,
viz.
Z ∞( 2 )
p2 4πa~2 N 4πa~2 N m V · 4πp2 dp

ε(p) − − + .
0 2m mV mV p2 h3

Substituting p = (8πa~2 N/V )1/2 x, we get


∞ 3/2
4πa~2 N 4πV 8πa~2 N
Z    
2 1/2 2 1
x(x + 2) − x − 1 + 2 x2 dx ,
0 mV 2x h3 V

which readily leads us from eqn. (11.3.14) to (11.3.15). The resulting


integral over x can be done by elementary means, giving
Z ∞  ∞
1 1 1 1 1
x(x2 + 2)1/2 − x2 − 1 + 2 x2 dx = (3x2 − 4)(x2 + 2)3/2 − x5 − x3 + x .
0 2x 15 5 3 2 0

For x  1,
  
1 2 2 3/2 1 5 3 3 3 1
(3x − 4)(x + 2) = (3x − 4x ) 1 + 2 + 4 + O
15 15 x 2x x6
 
1 1 1 1
= x5 + x3 − x + O .
5 3 2 x

The contribution from


√ the upper limit is, therefore, zero. From the
lower limit we get 128/15, which leads to eqn. (11.3.16).
(b) Noting that
3/2 1/2
4πVp 2 dp 8πa~2 N
 
4πV 128
= 3 x2 dx = N (na 3 ) x2 dx ,
h3 h V π

we readily obtain eqn. (11.3.23). Now the integral


Z ∞
x(x2 + 1)

1 1
2 + 2)1/2
− x 2
dx = (x2 − 1)(x2 + 2)1/2 − x3 |∞
0 .
0 (x 3 3

Again, for x  1,
    
1 2 2 1/2 1 3 1 1 1 3 1
(x −1)(x +2) = (x −x) 1 + 2 + O = x +O ,
3 3 x x4 3 x

with the result that the contribution


√ from the upper limit vanishes.
From the lower limit we get 2/3, which leads to eqn. (11.3.24).
11.6. We invert the given equation for n and write

4πa~2 n
 
32
µ0 = 1 + 1/2 (na 3 )1/2 + . . . .
m 3π
96

Substituting this into the given expressions for E0 and P0 , we get


2 
2πa~2 n2
 
E0 32 64
= 1 + 1/2 (na 3 )1/2 + . . . 1 − 1/2 (na 3 )1/2 + . . .
V m 3π 5π
2 2
 
2πa~ n 128
= 1+ (na 3 )1/2 + . . . , and
m 15π 1/2
2 
2πa~2 n2
 
32 3 1/2 128 3 1/2
P0 = 1 + 1/2 (na ) + . . . 1− (na ) + . . .
m 3π 15π 1/2
2πa~2 n2
 
64
= 1 + 1/2 (na 3 )1/2 + . . . ,
m 5π

in complete agreement with eqns. (11.3.16 and 17).


11.7. By eqns. (11.3.11), the number operator n̂p for the real particles is given by

1  + + +
n̂p = a+ b bp − αp b−p bp + b+ 2

p ap = p b−p + αp b−p b−p .
1 + αp2 p

The terms linear in αp do not contribute to the expectation value of n̂p


(with p 6= 0) because of the absence of the diagonal matrix elements in
+ +
b−p bp and b+ +
p b−p . Further, since bp bp = N̂p and b−p b−p = N̂−p +1, we get

1 
N̄p + αp2 (N̄−p + 1)

np = (p 6= 0).
1 − αp2

Finally, in view of the isotropy of the problem, N −p = N p and we get the


desired result.
11.8. For a solution to this problem, see Feynman (1954).
11.10 and 11. For solutions to these problems, see Fetter (1963, 1965).
11.14. We set x = 1 + ε, where |ε|  1, and find that

x2
 
4 1 + 2ε 3
2x ln 2 ' 2(1 + 4ε) ln  ' 2(1 + 4ε) − ln |ε| − ln 2 + ε ,
|x − 1| 2|ε| 1 + 12 ε 2
 
1 x + 1 2+ε
10 x − ln ' 20ε ln ' 20ε{− ln |ε| + ln 2},
x x − 1 |ε|
p ! √
(2 − x2 )5/2 1 + x (2 − x2 ) 2(2 − x2 )5/2 x + 2 − x2
ln p = ln √
x x

1 − x ((2 − x2 ) x − 2 − x2
2(1 − 2ε)5/2 (1 + ε) + (1 − ε − ε2 )

' ln
1+ε (1 + ε) − (1 − ε − ε2 )
 
2 1
' 2(1 − 6ε) ln
' 2(1 − 6ε) − ln |ε| − ε .
2ε + ε2 2
97

Substituting these results into the square bracket appearing in the formula
for ε(p), we get, to the desired degree of approximation,
11 + {−2 ln |ε| − 2 ln 2 + 3ε − 8ε ln |ε| − 8ε ln 2}
− {−20ε ln |ε| + 20ε ln 2} − {−2 ln |ε| − ε + 12ε ln |ε|}
= (11 − 2 ln 2) − 4(7 ln 2 − 1)ε,
which yields the stated result.
Comparing this result with eqn. (11.8.10), we find that
8 p3 a2
V (p) ' const. − 2
(7 ln 2 − 1) F 2 (p − pF ).
15π m~
Equation (11.8.11) then gives
 
1 1 8 2
' 1− (7 ln 2 − 1)(kF a) ,
m∗ m 15π 2
which leads to the desired result for the ratio m∗ /m.
11.15. At T = 0 K, the chemical potential of a thermodynamic system is given by
 
∂E ∂(E/V )
µ= = .
∂N v ∂(N/V )
It follows that, in the ground state of the given system,
Z n
N n
Z  
N
E=V µ(n)dn = µ(n)dn n= .
0 n 0 V

Now, since pF = (3π 2 n)1/3 ~, the given expression for µ may be written as
~2 ~2 a 2 ~2 a2
µ(n) ' (3π 2 n)2/3 + (2πn) + (3π 2 n)4/3 (11 − 2 ln 2) .
2m m 15π 2 m
It follows that
E 3 ~2 1 ~2 a 3 2 ~2 a2
' (3π 2 n)2/3 + (2πn) + (3π 2 n)4/3 2
(11 − 2 ln 2) ,
N 5 2m 2 m 7 15π m
which agrees with eqn. (11.7.31).
11.16. For a complete solution to this problem, see the first edition of this book —
Sec. 10.3, pages 311-5.
11.17. Correction to the first printing of third edition: In line 3, √
the definition of
3/2
the dimensionless wavefunction
p should read: ψ = aosc Ψ/ N . Using that
substitution and aosc = ~/(mω0 ) gives
1 ˜2 1 4πN a 2
− ∇ ψ + s2 ψ + |ψ| ψ = µ̃ψ,
2 2 aosc
˜ = ∂/∂sx + ... and µ̃ = µ/(~ω0 ).
where s = r/aosc , ∇
98
p
11.18. The solution for the case V = 0 is Ψ = N/V which gives µ = N u0 /V
and E = (2πa~2 N 2 )/(mV ).
11.19. For the case a → 0 the dimensionless G-P equation is
1 ˜2 1
− ∇ ψ + s2 ψ+ = µψ,
2 2
1
exp − 12 s2 with E/(N ~ω0 ) = 3/2, i.e. the

which has solution ψ = π3/4
zero point energy for N particles in the trap.
11.20. Use the dimensionless form from problem 11.17. Ignoring the kinetic en-
ergy term
q
2
µ̃ − s2
ψ=p .
4πN a/a0

The normalization is

2µ̃
s2
Z  
4πa0
1= µ̃ − ds
4πN a 0 2

which gives
a0
N= (2µ̃)5/2 .
15πa
Using the definitions for u0 , µ̃, and a0 gives equations (11.2.25) and
(11.2.26). Equation (11.2.28) follows from the definition of the dimension-
less length scale, and (11.2.27) comes from integrating the dimensionless
energy in problem 11.17, again ignoring the kinetic energy term.
Chapter 12

12.1. We assume the equation of state to be

1 λ3 2 λ6
 
kT
P = 1 − β1 − β2 2 , (1)
v 2 v 3 v

where β1 and β2 are certain functions of T . It follows that

kT λ3 kT λ6
 
∂P kT
= − 2 + β1 3 + 2β2 4 , and
∂v v v v
 2 T 3
∂ P kT kT λ kT λ6
2
= 2 3 − 3β1 4 − 8β2 5 .
∂v T v v v

At the critical point, both these derivatives vanish — with the result that

λ3c λ6 λ3 λ6
(β1 )c + 2(β2 )c 2c = 1 and 3(β1 )c c + 8(β2 )c 2c = 2,
vc vc vc vc
whence
(β1 )c = 2vc /λ3c and (β2 )c = −vc2 /2λ6c . (2)
We infer that, at the critical point, β12 = −8β2 .
Finally, substituting (2) into (1), we get

λ3 λ6
 
Pv 1 2 1 1
= 1 − (β1 )c c − (β2 )c 2c = 1 − 1 + = .
kT c 2 vc 3 vc 3 3

12.2. The given equation of state is


kT −a/kT v
P = e . (1)
v−b
It follows that
     
∂P −a/kT v 1 1 a 1 a
= kT e − + · =P − + ,
∂v (v − b)2 v − b kT v2 (v − b) kT v2
 2 T      
∂ P ∂P 1 a 1 2a
2
= − + 2
+P 2
− .
∂v T ∂v T (v − b) kT v (v − b) kT v3

99
100

At the critical point, both these derivatives vanish — with the result that

a vc2 2a vc3
= and = ,
kT c vc − b kT c (vc − b)2

whence vc = 2b and kT c = a/4b. Equation (1) then gives: Pc = (a/4b2 )e−2


and hence kT c /Pc vc = e2 /2 ' 3.695.

(a) For large v, the given equation of state may be approximated as


 −1  
kT b −a/kT v kT b a
P = 1− e ' 1+ − .
v v v v kT v

Comparing this with eqns. (10.3.7–10), we see that the coefficient B2


in the present case is formally the same as the one for the van der
Waals gas, viz. b − (a/kT ).
(b) We note that the derivative (∂P/∂v)T for the Dietrici gas can be
written as
 
∂P kT −a/kT v
n
2 a o
=− 2 e v − (v − b)
∂v T v (v − b)2 kT
 
kT −a/kT v a 2 ab
=− 2 e v − + (T − Tc ) .
v (v − b)2 2kT kT 2

Clearly, if T > Tc , then (∂P/∂v)T is definitely negative; the same is


true at T = Tc — except for the special case v = a/2kT c = 2b when
(∂P/∂v)T is zero. In any case, for all T ≥ Tc , P is a monotonically
decreasing function of v — with the result that, for any given T and
P , we have a unique v.
(c) For T < Tc , P is a non-monotonic function of v — generally decreas-
ing with v but increasing between the values
r r
a ab a ab
vmin = − (Tc − T ) and vmax = + (Tc − T ).
2kT kT 2 2kT kT 2

For any given T , we now have (for a certain range of P ) three possible
values of v such that

v1 > vmax > v2 > vmin > v3 ;

see Figs. 12.2 and 12.3. We further note that


vmin 1 vmax 1
= 1/2
and = .
vc 1 + (1 − T /Tc ) vc 1 − (1 − T /Tc )1/2

Clearly, vmin < vc < vmax and, hence, v3 < vc < v1 .


101

(d) To examine the critical behavior of the Dietrici gas, we write


a a
P = (1 + π), v = 2b(1 + ψ), T = (1 + t).
4e2 b2 4bk
The equation of state then takes the form
  
1+t 1
1+π = exp 2 1 − .
1 + 2ψ (1 + t)(1 + ψ)
Taking logarithms, carrying out expansions and retaining the most impor-
tant terms, we get
 
8 3 2
π ≈ t− 2ψ − 2ψ + ψ +2{1−(1−t)(1−ψ+ψ 2 −ψ 3 )} ≈ 3t− ψ 3 −2tψ.
2
3 3
We now observe that
(i) at t = 0, π ≈ − 23 ψ 3 , while at ψ = 0, π ≈ 3t,
(ii) for t < 0, we obtain three values of ψ: while |ψ2 |  |ψ1,3 |, implying
once again π ≈ 3t, ψ1,3 ≈ ±(3|t|)1/2 ,
(
 
∂ψ 1 1/2t (t > 0, ψ = 0)
(iii) the quantity − ∂π ≈ 2ψ2 +2t ≈ .
t 1/4|t| (t < 0, ψ = ψ1,3 )

Comparing these results with the ones derived in Sec. 12.2, we infer that
the critical exponents of this gas are precisely the same as those of the
van der Waals gas; the amplitudes, however, are different.
12.3. The given equation of state (for one mole) of the gas is

P = RT /(v − b) − a/vn (n > 1). (1)


Equating (∂P/∂v)T and (∂ 2 P/∂v2 )T to zero, we get
n+1 4n(n − 1)n−1 a
vc = b and Tc = .
n−1 (n + 1)n+1 bn−1 R
Equation (1) then gives
 n+1
n−1 a
Pc = ,
n+1 bn
whence RT c /Pc vc = 4n/(n2 − 1).
To determine the critical behaviour of this gas, we write

P = Pc (1 + π), v = vc (1 + ψ), T = Tc (1 + t).

The equation of state then takes the form


4n(1 + t) n+1
1+π = − .
(n2 − 1)(1 + ψ) − (n − 1)2 (n − 1)(1 + ψ)n
102

Carrying out the usual expansions and retaining only the most important
terms, we get
2n n(n + 1)2 3 n(n + 1)
π≈ t− ψ − tψ.
n−1 12 n−1
It follows that

(i) at t = 0, π ≈ − n(n + 1)2 /12 ψ 3 , while at ψ = 0, π ≈ {2n/(n −
1)}t,
(ii) for t < 0, we obtain three values of ψ; while |ψ2 |  |ψ1,3 |, implying
once again that π ≈ {2n/(n − 1)}t, ψ1,3 ≈ ±{12/(n2 − 1)}1/2 |t|1/2 ,
 
4(n−1)
− ∂ψ
∂π ≈ n(n+1){(n 2 −1)ψ 2 +4t}
t (
(iii) the quantity (n − 1)/n(n + 1)t (t > 0)
≈ .
(n − 1)/2n(n + 1)|t| (t < 0)

Clearly, the critical exponents of this gas are the same as those of the van
der Waals gas — regardless of the value of n. The critical amplitudes (as
well as the critical constants Pc , vc and Tc ), however, do vary with n and
hence are model-dependent.
P
12.4. The partition function of the system may be written as exp f (L), where
L
 
1
f (L) = ln N ! − ln(Np)! − ln(Nq)! + βN qJL2 + µBL .
2
Using the Stirling approximation (B.29), we get
 
1 2
f (L) ≈ −Np ln p − Nq ln q + βN qJL + µBL .
2
With p and q given by eqn. (1) of the problem, the function f (L) is
maximum when
1 1
− N (1 + ln p) + N (1 + ln q) + βN (qJL + µB) = 0.
2 2
Substituting for p and q, the above condition takes the form
1 1+L
ln = β(qJL + µB).
2 1−L
Comparing this with eqn. (12.5.10), we see that the value, L∗ , of L, that
maximizes the function f (L) is identical with L̄.
The free energy and the internal energy of the system are now given by
 
1
A ≈ − kT f (L∗ ) ≈ NkT (p∗ ln p ∗ + q ∗ ln q ∗ ) − N qJL∗2 + µBL∗ ,
2
1
U ≈ − NqJL∗2 − N µBL∗ ,
2
103

whence
S ≈ −Nk (p∗ ln p ∗ + q ∗ ln q ∗ ).

12.5. The relevant results of the preceding problem are

1 + L∗ 1 + L∗ 1 − L∗ 1 − L∗
 
A 1
= kT ln + ln − qJL∗2 − µBL∗ ,
N 2 2 2 2 2
(1)
1 1
N + = N (1 + L∗ ), N − = N (1 − L∗ ), (2)
2 2
where L∗ satisfies the maximization condition
1
{ln(1 + L∗ ) − ln(1 − L∗ )} = β(qJL∗ + µB). (3)
2
Combining (1) and (3), we get

A 1 1 − L∗2 1
= kT ln + qJL∗2 . (4)
N 2 4 2
Now, using the correspondence given in Section 11.4 and remembering
that L∗ is identical with L̄, we obtain from eqns. (2) and (4) the desired
results for the quantities P and v pertaining to a lattice gas.
For the critical constants of the gas, we first note from eqn. (12.5.13) that
Tc = qJ /k, i.e. qε0 /4k; the other constants then follow from the stated
results for P and v, with B = 0 and L̄ = 0.

12.6. The Hamiltonian of this model may be written as


1 X X
H=− c σi σj − µB σi .
2 i
i6=j

σi2 = (NL)2 − N which,


P P P
The double sum here is equal to σi σj −
i j i
for N  1, is essentially equal to N 2 L2 . It follows that asymptotically our
Hamiltonian is of the form
1
H = − (cN )NL2 − µBNL.
2
Now, this is precisely the Hamiltonian of the model studied in Prob-
lem 12.4, except for the fact that the quantity qJ there is replaced by
the quantity cN here. We, therefore, infer that, in the limit N → ∞ and
c → 0 (such that the product cN is held fixed), the mean-field approach of
Problem 12.4 would be exact for the present model — provided that the
fixed value of the product cN is identified with the quantity qJ. It follows
that the critical temperature of this model would be cN/k.
104

12.7 & 8. Let us concentrate on one particular spin, s0 , in the lattice and look at the
Pq
part of the energy E that involves this spin, viz. −2J s0 ·sj −gµB s0 ·H;
j=1
for notation, see Secs. 3.9 and 12.3. In the spirit of the mean field theory,
we replace each of the sj by s̄, which modifies the foregoing expression to
−gµB s0 · Heff , where

Heff = H + H0 (H 0 = 2qJ s̄/gµB ). (1)

We now apply the theory of Sec. 3.9. Taking H (and hence s̄) to be in the
direction of the positive z-axis, we get from eqn. (3.9.22)

µz = gµB s Bs (x) [x = β(gµB s)Heff ], (2)

where Bs (x) is the Brillouin function of order s. At high temperatures


(where x  1), the function Bs (x) may be approximated by {(s+1)/3s}x,
with the result that
g 2 µ2B s(s + 1)
 
2qJ µ
µz ≈ H + 2 2z . (3)
3kT g µB

The net magnetization, per unit volume, of the system is now given by
the formula M = nµ̄z , where n(= N/V ) is the spin density in the lattice.
We thus get from (3)
 
Tc CH CH
M 1− ≈ , i.e. M ≈ , where (4)
T T T − Tc
2s(s + 1)qJ ng 2 µ2B s(s + 1)
Tc = , C= . (5)
3k 3k
The Curie-Weiss law (4) signals the possibility of a phase transition as T →
Tc from above. However, this is only a high-temperature approximation,
so no firm conclusion about a phase transition can be drawn from it. For
that, we must look into the possibility of spontaneous magnetization in
the system.
To study the possibility of spontaneous magnetization, we let H → 0 and
write from (2)

µz = gµB s Bs (x0 ) [x0 = β(gµB s)H 0 = 2sqJ µz /gµB kT ]. (6)

In the close vicinity of the transition temperature, we expect µ̄z to be


much less than the saturation value gµB s, so once again we approximate
the function Bs (x0 ) for x0  1. However, this time we need a better
approximation than the one employed above; this can be obtained by
utilizing the series expansion

1 x x3
coth x = + − + ... (x  1),
x 3 45
105

which yields the desired result:

s+1 (s + 1){s2 + (s + 1)2 } 3


Bs (x) = x− x + ... (x  1). (7)
3s 90s3
Substituting (7) into (6), we get
 3
Tc b Tc
µz = µ − µ + ..., (8)
T z g 2 µ2B s2 T z

where Tc is the same as defined in (5), while b is a positive number given


by
s2
 
3
b= 1+ . (9)
10 (s + 1)2
Clearly, for T < Tc , a non-zero solution for µ̄z is possible; in fact, for
T . Tc , √
µz ≈ (gµB s/ b)(1 − T /Tc )1/2 . (10)
The long-range order L̄0 is then given by

L0 ≡ µz /(gµB s) ≈ (1/ b)(1 − T /Tc )1/2 . (11)

For T  Tc , we employ the approximation

coth x ≈ 1 + 2e−2x , whence Bs (x) ≈ 1 − s−1 e−x/s (x  1). (12)

Equation (6) now gives


   
−1 2sqJ −1 3Tc
L0 ≈ 1 − s exp − = 1 − s exp − . (13)
kT (s + 1)T

For s = 1/2, expressions (11) and (13) reduce precisely to eqns. (12.5.14
and 15) of the Ising model; the expression for Tc is different though.
12.9 & 10. We shall consider only the Heisenberg model; the study of the Ising model
is somewhat simpler. Following the procedure of Problem 12.7, we find
that the “effective field” Ha experienced by any given spin s on the sub-
lattice a would be
2q 0 J 0 2qJ
Ha = H − s̄b − s̄a , (1a)
gµB gµB

where q 0 and q are, respectively, the number of nearest neighbors on the


other and on the same sub-lattice, while J 0 and J are the magnitudes of
the corresponding interaction energies. Similarly,

2q 0 J 0 2qJ
Hb = H − s̄a − s̄b . (1b)
gµB gµB
106

The net magnetization, per unit volume, of the sub-lattices a and b at


high temperatures is then given by, see eqns. (2) and (3) of the preceding
problem,

g 2 µ2B s(s + 1) 4q 0 J 0
 
1 1 4qJ
Ma ≡ n(gµB s̄a ) ≈ n · H − 2 2 Mb − 2 2 Ma ,
2 2 3kT ng µB ng µB
2 2 0 0
 
1 1 g µB s(s + 1) 4q J 4qJ
Mb ≡ n(gµB s̄b ) ≈ n · H − 2 2 Ma − 2 2 Mb .
2 2 3kT ng µB ng µB
Adding these two results, we obtain for the total magnetization of the
lattice
2q 0 J 0
 
C 0 0 2qJ
M ≈ {H − (γ + γ)M} γ = , γ= ; (2)
T ng 2 µ2B ng 2 µ2B

the parameter C here is the same as defined in eqn. (5) of the preceding
problem. Equation (2) may be written in the form
C
M≈ H, where θ = (γ 0 + γ)C; (3)
T +θ
this yields the desired result for the paramagnetic susceptibility of the
lattice. Note that the parameter θ here has no direct bearing on the
onset of a phase transition in the system; for that, we must examine the
possibility of spontaneous magnetization in the two sub-lattices.
To study the possibility of spontaneous magnetization, we let H → 0; in
that limit, the vectors Ma and Mb are equal in magnitude but opposite
in direction. We may then write: Ma = M∗ , Mb = −M∗ , and study
only the former. In analogy with eqn. (6) of the preceding problem, we
now have
4s(q 0 J 0 − qJ )M ∗
 
1
M ∗ = n(gµB s)Bs (x0 ) x0 = . (4)
2 n gµB kT

We now employ expansion (7) of the preceding problem and get


 3
TN ∗ b TN ∗
M∗ = M − 2 M + ..., (5)
T 1 T
2 n gµB s

where
2s(s + 1) 0 0
TN = (q J − qJ ), (6)
3k
while the number b is the same as given by eqn. (9) of the preceding
problem. Clearly, for T < TN , a nonzero solution for M ∗ is possible. Note
that the Néel temperature TN = (γ 0 − γ)C, which should be contrasted
with the parameter θ of eqn. (3); moreover, for the antiferromagnetic
transition to take place in the given system, we must have q 0 J 0 > qJ , the
physical reason for which is not difficult to understand.
107

12.11. To determine the equilibrium distribution f (σ), we minimizePthe free en-


ergy (E − TS ) of the system under the obvious constraint f (σ) = 1.
σ
For this, we vary the function f (σ) to f (σ) + δf (σ) and require that the
resulting variation
1 X
δ(E − TS ) = qN u(σ 0 , σ 00 ){f (σ 0 )δf (σ 00 ) + f (σ 00 )δf (σ 0 )}
2 0 00
σ ,σ
X
+ NkT {1 + ln f (σ 0 )}δf (σ 0 ) = 0,
σ0

δf (σ 0 ) is, of necessity, zero. Introducing the Lagrange multiplier


P
while
σ0
λ and remembering that the function u(σ 0 , σ 00 ) is symmetric in σ 0 and σ 00 ,
our requirement takes the form
( )
X X
qN u(σ , σ )f (σ ) + NkT {1 + ln f (σ )} − λ δf (σ 0 ) = 0.
0 00 00 0

σ0 σ 00

Since the variation δf (σ 0 ) in this expression is arbitrary, the condition for


equilibrium becomes
X
qN u(σ 0 , σ 00 )f (σ 00 ) + NkT ln f (σ 0 ) − N µ = 0,
σ 00

where µ = (λ − NkT )/N . By a change of notation, we get the desired


result " #
X
0 0
f (σ) = C exp −βq u(σ, σ )f (σ ) , (1)
σ0

where
P C is a constant to be determined by the normalization condition
f (σ) = 1.
σ
For the special case u(σ, σ 0 ) = −Jσσ 0 , where the σ’s can be either +1 or
−1, eqn. (1) becomes

f (σ) = C exp[βqJ σ{f (1) − f (−1)}]. (2)

Writing f (σ) = 21 (1 + L̄0 σ), the quantity f (1) − f (−1) becomes precisely
equal to L̄0 , and eqn. (2) takes the form

f (σ) = C exp[βqJ σL0 ]. (3)

From equation (3), we obtain

f (1) + f (−1) = 2C cosh(βqJ L0 ) = 1, while (4)


f (1) − f (−1) = 2C sinh(βqJ L0 ) = L0 . (5)

Dividing (5) by (4), we obtain the Weiss eqn. (12.5.11) for L̄0 .
108

12.12. The configurational energy of the lattice is given by


1
E= qN [ε11 · xA (1 + X) · xA (1 − X) + ε12 {xA (1 + X)(xB + xA X)
2
+ xA (1 − X)(xB − xA X)} + ε22 (xB − xA X)(xB + xA X)]
 
1 1
= qN ε11 x2A + 2ε12 xA xB + ε22 x2B − 2εx2A X 2
  
ε = (ε11 + ε22 ) − ε12 .
2 2
The entropy, on the other hand, is given by
      
1 1 1
S = k ln N ! − ln Nx A (1 + X) ! − ln N (xB − xA X) !+
2 2 2
     
1 1 1
ln N ! − ln Nx A (1 − X) ! − ln N (xB + xA X) !
2 2 2
1
≈ Nk [−xA (1 + X) ln{xA (1 + X)} − (xB − xA X) ln(xB − xA X)
2
− xA (1 − X) ln{xA (1 − X)} − (xB + xA X) ln(xB + xA X)].
To determine the equilibrium value of X, we minimize the free energy of
the system and obtain
 
∂(E − TS ) 1 (1 + X)(xB + xA X)
= −2qN εx2A X + NkTx A ln = 0.
∂X 2 (1 − X)(xB − xA X)
(1)
Now, since xA + xB = 1, the argument of the logarithm can be written
as (1 + z)/(1 − z), where z = X/(xB + xA X 2 ). Equation (1) then takes
the form
2qεxA X 1 1+z
= ln = tanh−1 z, (2)
kT 2 1−z
which is identical with the result quoted in the problem. For xA = xB = 21 ,
eqn. (2) reduces to the more familiar result
qεX 2X
= tanh−1 = 2 tanh−1 X, (2a)
kT 1 + X2
leading to a phase transition at the critical temperature Tc0 = qε/2k.
To determine the transition temperature Tc in the general case when xA 6=
xB , we go back to eqn. (2) and write it in the form
4xA Tc0
 
X
= tanh X . (3)
xB + xA X 2 T
For small X, we get
3
4xA Tc0 1 4xA Tc0

1 xA 3
X − 2 X + ... = X− X 3 + . . . , i.e.
xB xB T 3 T
" 3 #
4xA xB Tc0 4xA xB Tc0
 
1 1
−1 X − 3 − 3xA xB X 3 + . . . = 0.
xB T 3xB T
109

It is now straightforward to see that for T < 4xA xB Tc0 , a non-zero solution
for X is possible whereas for T ≥ 4xA xB Tc0 , X = 0 is the only possibility.
The transition temperature Tc is, therefore, given by 4xA (1 − xA )Tc0 .
12.13. For a complete solution to this problem, see Kubo (1965), problem 5.6,
pp. 335–7.
1
12.14. (a) Setting N ++ + N −− + N +− = 2 qN , we find that, in equilibrium,
2
γ = 1/(1 + sL ). So, in general, it may be written as 1/(1 + sL2 ).
P
(b) As in Problem 12.4, we write Q(B, T ) = exp f (L), where
L

f (L) = ln N ! − ln N+ ! − ln N− ! − βE, with


1 1
N+ = N (1 + L), N− = N (1 − L),
2 2
and E = −J(N++ + N−− − N+− ) − µB(N+ − N− )
1
= −J · qN (L2 + s)/(1 + sL2 ) − µBNL.
2
The condition that maximizes f (L) now reads:

(1 − s2 )L
 
1 1+L
ln = β qJ + µB .
2 1−L (1 + sL2 )2

In the close vicinity of the critical point, L  1 — with the result


that
1 1+L
' β qJ (1 − s2 )L + µB
 
ln (T ' Tc ).
2 1−L
Comparing this with the corresponding equation in the solution to
Problem 12.4, we infer that the critical behavior of this model is qual-
itatively the same as one encounters in the Bragg-Williams approxi-
mation. Quantitatively, though, the effective spin-spin interaction is
reduced by the factor (1 − s2 ) — leading to a critical temperature
Tc = (1 − s2 )qJ /k, instead of qJ /k.
(c) As for the specific-heat singularity, the limit T → Tc− would be iden-
tical with the one obtained in Sec. 12.5; see the derivation leading
to eqn. (12.5.18) and note that the replacement of J by (1 − s2 )J
does not affect the final result 23 Nk . For T > Tc , L0 is identi-
cally zero. We are then left with a finite configurational energy,
− 12 qJNs, that arises from the (assumed) short-range order in the
system; however, unlike in the Bethe approximation, this energy is
temperature-independent and hence does not entail any specific heat.
The singularity in question is, therefore, precisely the same as the one
encountered in Sec. 12.5 and depicted in Fig. 12.8.
110

12.15. Using eqn. (12.6.30), we get


Z ∞ Z γc
S∞ − Sc 1 C(T )dT 1
= = q γ sech 2 γdγ
Nk Nk Tc T 2 0
1
= q(γc tanh γc − ln cosh γc ).
2
Next, we use eqn. (12.6.11) and obtain
"   (  2 )#
S∞ − Sc 1 1 q 1 1 1
= q ln + ln 1 −
Nk 2 2 q−2 q−1 2 q−1
1 q 1
= {ln q − ln(q − 2)} + q{ln q + ln(q − 2) − 2 ln(q − 1)}.
4q−1 4
In view of the fact that S∞ = Nk ln 2, we finally get
Sc 1 q2 1 1 q(q − 2)
= ln 2 − ln q + q ln(q − 1) − ln(q − 2),
Nk 4q−1 2 4 q−1
which leads to the desired result.
For q  1, the stated expression for Sc /Nk reduces to
        
q 1 1 q 1 2 1
ln 2 + − +O 2
− 1 + O − + O
2 q q 4 q q q2
 
1
= ln 2 + O ,
q
which tends to the limit ln 2 as q → ∞.
12.16. Using eqn. (12.6.14), we get
N µ2 ∂ L̄ 2N µ2 1 + exp(−2γ) cosh(2α + 2α0 )
   
∂ M̄
χ≡ = =
∂B T kT ∂α T kT {cosh(2α + 2α0 ) + exp(−2γ)}2
  0 
∂α
1+ .
∂α T
(1)
To determine (∂α0 /∂α)T , we differentiate (12.6.8) logarithmically and ob-
tain after some simplification
 0
∂α (q − 1){tanh(α + α0 + γ) − tanh(α + α0 − γ)}
= . (2)
∂α T 2 − (q − 1){tanh(α + α0 + γ) − tanh(α + α0 + γ)}
Substituting (2) into (1) and letting α → 0, we get
4N µ2 1 + exp(−2γ) cosh(2α0 )
χ0 =
kT {cosh(2α0 ) + exp(−2γ)}2
1
. (3)
2 − (q − 1){tanh(α + γ) − tanh(α0 − γ)}
0
111

To study the critical behavior of χ0 , we let α0 → 0 and γ → γc . Using


eqn. (12.6.11), we see that, while the first two factors of expression (3)
reduce to
4N µ2 1 2N µ2 q
= , (4)
kT c 1 + exp(−2γc ) kT c q − 1
the last factor diverges. To determine the nature of the divergence, we
write γ = γc (1 − t) and carry out expansions in powers of t and α0 . Thus

tanh(γ ± α0 ) = tanh γc + sech 2 γc (−γc t ± α0 )


2
− sech 2 γc tanh γc (−γc t ± α0 ) + . . . ,

so that

tanh(γ + α0 ) + tanh(γ − α0 ) ≈ 2 tanh γc − 2 sech2 γc · γc t


− 2 sech2 γc tanh γc · α02 ;

note that we have dropped terms of order t2 and higher. It now follows
that

2 − (q − 1){tanh(γ + α0 ) + tanh(γ − α0 )}
≈ 2(q − 1)sech2 γc γc t + tanh γc · α02 .

(5)

Substituting (4) and (5) into (3), we finally obtain

N µ2 1
χ0 ≈ . (6)
kT c (q − 2) (γc t + tanh γc · α02 )

For t > 0, α0 = 0; eqn. (6) then gives

N µ2 1
χ0 ≈ . (7a)
kT c (q − 2)γc t

For t < 0, α0 is given by eqn. (12.6.13), whence α02 ' −3(q − 1)γc t; we
now get
N µ2 1
χ0 ≈ . (7b)
kT c 2(q − 2)γc |t|
Note that, for large q, the quantity
   
1 q 1
(q − 2)γc = (q − 2) ln =1+O ;
2 q−2 q

eqns. (7) then reduce to eqns. (12.5.22) of the Bragg-Williams approxima-


tion.
112

12.19. We refer to the solutions to Problems 12.4 and 12.5, whereby


 
A 1 + m0 1 + m0 1 − m0 1 − m0 1 qJ 2
ψ0 ≡ = ln + ln − m
NkT B=0 2 2 2 2 2 kT 0
1 1 − m20 m0 1 + m0 1 Tc 2
= ln + ln − m
2 4 2 1 − m0 2T 0
m4 m3
   
1
= − ln 2 + −m20 − 0 − . . . + m0 m0 + 0 + . . .
2 2 3
1
− (1 − t + . . .)m20
2
1 1
≈ − ln 2 + t m0 + m40 .
2
2 12
Comparing this expression with eqn. (12.9.5), we infer that in the Bragg-
Williams approximation r1 = 1/2 and s0 = 1/12. Now, substituting
these values of r1 and s0 into eqns. (12.9.4, 9–11 and 15), we see that the
corresponding eqns. (12.5.14, 22, 24 and 18) are readily verified.
A similar calculation under the Bethe approximation is somewhat tedious;
the answer, nevertheless, is

q−2 q (q − 1)(q − 2)
r1 = ln , s0 = .
4 q−2 12q 2
12.20. The equilibrium values of m in this case are given by the equation
ψh0 = −h + 2 rm + 4 sm 3 + 6 um 5 = 0 (u > 0). (1)
√ √
With h = 0, we get: m0 = 0, ± A+ or ± A− , where

−s ± s2 − 3 ur
A± = . (2)
3u
First of all, we note that, for A± to be real, s2 must be ≥ 3ur . This
presents
√ no problem
√ if r ≤ 0; however, if r > 0, then s must be either
≥ 3ur or ≤ − 3ur . We also observe that
r 2s
A+ A− = and A+ + A− = − .
3u 3u
It follows that (i) if r < 0, then one of the A’s will be positive, the other
negative (in fact, since A− < A+ , A− will be negative and A+ positive),
(ii) if r = 0, then for s > 0, A− will be negative and A+ = 0, for s = 0
both A− and A+ will be zero whereas for s < 0, A− will be zero while √ A+
will be positive (and equal to 2|s|/3u), (iii) if r >√0, then for s ≥ 3ur
both A+ and A− will be negative whereas for s ≤ − 3ur both A+ and A−
will be positive. We must, in this context, remember that only a positive
A will yield a real m0 . Finally, since
ψ000 = 2r + 12 sm 20 + 30 um 40 , (3)
113

the extremum at m0 = 0 is a maximum if r < 0, a minimum if√r > 0. It


0 the function ψ0 is minimum at√m0 = ± A+ and
follows that for r < √
for r > 0 (and
√ s ≤ − 3ur ) it is maximum at m0 = ± A− and minimum
at m0 = ± A+ . We, therefore, have to contend only with A+ .
The foregoing observations should suffice to prove statements (a), (e), (f)
and (g) of this problem. For the rest, we note that the function ψ0 (m0 )
may be written as
 1
ψ0 (m0 ) = q + rm 20 + sm 40 + um 60 − m0 2rm 0 + 4sm 30 + 6um 50 (4a)

4
1 2 4

= q + m0 r − um 0 ; (4b)
2
note that in writing (4a) we have added an expression which, by the min-
imization condition, is identically zero. It now follows from eqn. (4b) that
ψ0 (m0 = 0) is less than, equal to or greater than p ψ0 (m0 6= 0) accrod-
ing as m20 is less than,pequal to or greater than r/u. The dividing line
corresponds to A+ = r/u, i.e.


r
−s + s2 − 3ur r
= , i.e. s = − 4ur ;
3u u

see the accompanying figure. We also note that, in reference to p the di-
2
viding line,
√ 0m decreases monotonically towards the limiting value √r/3u
as s → − 3ur and increases montonically as s decreases below − 4ur .
These observations should suffice to prove statements (b), (c) and (d).

12.21. With s = 0, the order parameter m is given by the equation

ψh0 = −h + 2rm + 6um 5 = 0.


114

For |t|  1, we set r ≈ r1 t; the equation of state then takes the form

h ≈ 2r1 tm + 6um 5 .

With t < 0 and h → 0, we get: m0 ≈ (r1 /3u)1/4 |t|1/4 , giving β = 1/4.


With t = 0, we get: h ≈ 6um 5 , giving δ = 5. For susceptibility, we have
 
∂m 1
χ∼ ≈ .
∂h t 2r1 t + 30um 4

It follows that 
1/2 r1 t (t > 0, m → 0)
χ0 ≈
1/8 r1 |t| (t < 0, m → m0 ),
giving γ = γ 0 = 1. Finally, using the scaling relation α + 2β + γ = 2, we
get: α = 1/2.
12.22. (a) We introduce the variable ψ[= (vg −vc )/vc ' (vc −v` )/vc ] and obtain

∂G(s)
   
2−α−∆ 0 π
ψ∼ ∼ |t| g . (2)
∂π t |t|∆

With t < 0 and π → 0, we get: ψ ∼ |t|β , where β = 2 − α − ∆. It


follows that the quantities (ρ` − ρc ), (ρc − ρg ) and (ρ` − ρg ) all vary
as |t|β .
(b) Writing g 0 (x) as xβ/∆ f (x), eqn. (1) takes the form

ψ ∼ π β/∆ f (π/|t|∆ ). (2)

It follows that the quantity |t|∆ /π is a universal function of the quan-


tity π β/∆ /ψ, i.e.

|t| ∼ π 1/∆ × a universal function of (π/ψ ∆/β ).

It is now clear that, at t = 0, π ∼ ψ δ , where δ = ∆/β.


(c) For the isothermal compressibility of the system, we have from (1)
     
1 ∂v ∂ψ π
κT ≡ − ∼ ∼ |t|β−∆ g 00 .
v ∂P T ∂π t |t|∆

It follows that, in the limit π → 0, κT ∼ |t|−γ , where γ = ∆−β = β(δ−1).


As for the coefficient of volume expansion, we have the relationship
       
1 ∂v 1 ∂v ∂P ∂P
αP = =− = κT .
v ∂T P v ∂P T ∂T v ∂T v

In the region of phase transition, (∂P/∂T )v is simply (dP/dT ), which is


non-singular. Accordingly, αP ∼ κT ∼ |t|−γ . Similarly, in view of the
115

relation CP = VT (dP /dT )2 κT , established in Problem 13.25, we infer


that CP ∼ κT ∼ |t|−γ .
For CV , we go back to the given expression for G(s) and write
 (s)   
∂G π
S (s) = − ∼ |t|1−α × a universal function of
∂T P |t|∆
 
ψ
∼ |t|1−α × a universal function of .
|t|β

It then follows that


 (s)   
(s) ∂S −α ψ
CV = T ∼ |t| × a universal function of .
∂T V |t|β
(s)
Now, letting ψ → 0, we obtain: CV ∼ |t|−α . And, in view of the relation
CV = VT (dP /dT )2 κS , also established in Problem 13.25, we infer that
κS ∼ CV ∼ |t|−α .
Finally, for the latent heat of vaporization `, we invoke the Clapeyron
equation,
dPσ `
= ,
dT T (vg − v` )
and conclude that ` ∼ |t|β .
12.23. We make the following observations:

(i) With h = 0 and t < 0, m0 = 0 or ±(b−1 |t|)1/2 , giving β = 1/2.


(ii) With t = 0, h = abm 2Θ+1 , giving δ = 2Θ + 1.
(iii) The quantity
 
∂m 1
= .
∂h t a(1 + 3bm 2 )(t + bm 2 )Θ−1

With t > 0 and h → 0, m → 0 and we are left with


 
∂m 1
≈ , giving γ = Θ.
∂h t at Θ

The scaling relation (12.10.22) is readily verified.

12.24. For ri 6= rj , eqns. (12.11.22 and 26) give

(∇2 − ξ −2 )g(r) = 0. (1)


116

Now, if g(r) is a function of r only then


dg dg r dg  x1 xd 
∇g = ∇r = = ,..., , whence
dr dr r dr r r
d d   2
d g 1 dg 1 x2i
  X  
X ∂ dg xi dg 1
∇ · ∇g = = + −
i=1
∂xi dr r i=1
dr r dr 2 r dr r2 r
d2 g dg 1
= + (d − 1). (2)
dr 2 dr r
Substituting (2) into (1), we obtain the desired differential equation — of
which (12.11.26) is the exact solution.
Substituting (12.11.27) into the left-hand side of the given differential
equation, we get

e−r/ξ
  
1 1
const. (d−1)/2 + . . . + (. . .) − ;
r ξ2 ξ2
the ratio of the terms omitted to the ones retained is O(ξ/r). Clearly, the
equation is satisfied for r  ξ. Similarly, substituting (12.11.28) instead,
we get  
1 (2 − d)(1 − d) d − 1 2 − d
const. d−2 + − (. . .) ;
r r2 r r
the ratio of the term omitted to the ones retained is now O(r/ξ)2 . Clearly,
the equation is again satisfied but this time for r  ξ.
12.25. By the scaling hypothesis of Sec. 12.10, we expect that, for h > 0 and
t > 0,

ξ(t, h) = ξ(t, 0) × a universal function of (h/t∆ )


= ξ(t, 0) × a universal function of (t/h1/∆ ).

Now, in view of eqn. (11.12.1), we may write

ξ(t, h) ∼ t−ν (t/h1/∆ )v × a universal function of (t/h∆ )


= h−ν/∆ × a universal function of (t/h1/∆ ).
c
At t = 0, we obtain: ξ(0, h) ∼ h−ν , where ν c = ν/∆.
c
By a similar argument, χ(0, h) ∼ h−γ , where

γ c = γ/∆ = β(δ − 1)/βδ = (δ − 1)/δ.

12.26. Clearly, the ratio (ρ0 /ρs ) ∼ |t|2β−ν . In view of the scaling relations α +
2β + γ = 2, γ = (2 − η)ν and dν = 2 − α, we get

2β − ν = (2 − α − γ) − ν = {dν − (2 − η)ν} − ν = (d − 3 + η)ν.

Setting d = 3, we obtain the desired result.


117

12.27. By definition,
σ ≡ ψA (t) ∼ |t|µ (t . 0). (1)
By an argument similar to the one that led to eqn. (12.12.14), we get

ψA (t) ∼ A−1 ∼ ξ −(d−1) , (2)

where A is the “area of a typical domain in the liquid-vapor interface”.


Now, for a scalar model (n = 1), to which the liquid-vapor transition
belongs, ξ ∼ |t|−ν . Equations (1) and (2) then lead to the desired result

µ = (d − 1)ν = (2 − α)(d − 1)/d.


Chapter 13

13.1. Since M̄ = (N̄+ − N̄− )µ and N̄+ + N̄− = N , we readily see that
 
1 M̄ 1 P (β, B) ± sinh x
N̄± = N 1 ± = N (x = βµB). (1)
2 Nµ 2 P (β, B)

Next, comparing eqns. (12.3.19) and (13.2.12), and keeping in mind that
q = 2, we get

e−4βJ P 2 (β, B) − sinh2 x


N̄+ − N̄++ = N =N . (2)
2D(β, B) 2D(β, B)

It follows from eqns. (1) and (2) that

N̄++ = (N/2D){(P + sinh x)(P + cosh x) − (P 2 − sinh2 x)}


= (N/2D){(P + sinh x)(cosh x + sinh x)}, (3)

which is the desired result. Equation (12.3.17) now gives

N̄+− = 2(N̄+ − N̄++ ) and N̄−− = N − 2N̄+ + N̄++ = N̄++ − (M̄ /µ).

The number N̄+− is just twice the expression (2). The number N̄−− is
given by

N̄−− = (N/2D){(P + sinh x)(cosh x + sinh x) − sinh x · 2(P + cosh x)}


= (N/2D){(P − sinh x)(cosh x − sinh x)}, (4)

which is the desired result.


It is straightforward to check that the sum

N̄++ +N̄−− = (N/D)(P cosh x+sinh2 x) = (N/D){P cosh x+(P 2 −e−4βJ )};

adding N̄+− , one obtains the expected result N . Finally, the product

N̄++ N̄−− = (N/2D)2 {P 2 − sinh2 x} = (N/2D)2 e−4βJ .

Equation (12.6.22) is now readily verified.

118
119

13.2. (a) In view of eqn. (12.3.18), the quantity (N++ + N−− − N+− ) that
appears in the Hamiltonian (12.3.19) of the lattice may be written as
1
2 qN − 2N+− . The partition function (12.3.20) then assumes the
form stated here.
(b) A complete solution to this problem can be found in the first edition
of this book — Sec. 12.9A, pp. 414–8. In any case, this problem is
a special case, q = 2, of the next problem which is treated here at
sufficient length.
13.3. In the notation of Problem 13.2, we now have
 
1 1
ln gN (N+ , N+− ) ≈ qN ln qN − N++ ln N++ − N−− ln N−−
2 2
 
1
− N+− ln N+− + (q − 1){N+ ln N+ + N− ln N− − N ln N },
2
(1)

where, by virtue of eqn. (12.3.17),


1 1 1 1 1
N++ = qN − N+− , N−− = qN − qN + − N+− and N− = N −N+ .
2 + 2 2 2 2
(2)
Now, as usual, the logarithm of the partition function may be approxi-
mated by the logarithm of the largest term in the sum over N+ and N+− ,
with the result that
 
∗ ∗
 1 ∗ ∗

ln QN (B, T ) ≈ ln gN N+ , N+− +βJ qN − 2N+− +βµB 2N+ −N ,
2
(3)
∗ ∗
where N+ and N+− are the values of the variables N+ and N+− that
maximize the summand (or the log of it). The maximizing conditions
turn out to be
   
∂ 1 1
(. . .) = −(1 + ln N++ ) q − (1 + ln N−− ) − q +
∂N+ 2 2
(q − 1){(1 + ln N+ ) + (1 + ln N− )(−1)} + 2βµB
( q/2  q−1 )
N−− N+
= ln + 2βµB = 0, and (4)
N++ N−
   
∂ 1 1
(. . .) = −(1 + ln N++ ) − − (1 + ln N−− ) − −
∂N+− 2 2
  
1
1 + ln N+− − 2βJ
2
( 1/2 1/2 )
N++ N−−
= ln 1
 − 2βJ = 0. (5)
2 N+−
120

Equations (4) and (5), with the help of eqns. (2), determine the equilibrium
values of all the numbers involved in the problem; eqn. (3) then determines
the rest of the properties of the system.
To compare these results with the ones following from the Bethe approx-
imation, we first observe that eqn. (5) here is identical with the corre-
sponding eqn. (12.6.22) of that treatment. As for eqn. (4), we go back to
eqns. (12.6.4 and 8) of the Bethe approximation, whereby
 0
q
N̄+ 2α cosh(α + α + γ) 0
=e 0
= e2α+2α q/(q−1)
N̄− cosh(α + α − γ)

and to eqn. (12.6.21), whereby

N̄−− 0
= e−4(α+α ) .
N̄++

It follows that
 q/2  q−1
N −− N+ 0 0
= e−2q(α+α )+2α(q−1)+2α q = e−2α (α = βµB),
N ++ N−

in complete agreement with eqn. (4). Hence the equivalence of the two
treatments.
13.4. By eqns. (13.2.8 and 37),

cosh x + {e−4βJ + sinh2 x}1/2


 
ξ −1 (B, T ) = ln (x = βµB).
cosh x − {e−4βJ + sinh2 x}1/2

As T → Tc (which, in this case, is 0),


 
cosh x + sinh x
ξ −1 (B, Tc ) ≈ ln = 2x.
cosh x − sinh x

It follows that ξ(B, Tc ) ≈ (1/2x) ∼ B −1 , which means that the critical


exponent ν c = 1. Now, a reference to eqns. (13.2.21 and 35) tells us that
ν = ∆. The relation ν c = ν/∆ is thus verified.
13.5. On integration over B, our partition function takes the form
 !2 
 1/2 X  βµ2 s X
2πs X 
QN (s, T ) ∼ exp σi + βJ σi σi+1 ,
βN  2N
i i

{σi }

which implies an effective Hamiltonian given by the expression


!
1 X
−1
X
Heff = − µ2 s NL2 − J σi σi+1 L=N σi .
2 i i
121

The first term here is equivalent to an infinite-range interaction of the type


considered in Problem 12.6, with µ2 s playing the role of the quantity cN
of that model. This is also equivalent to the Bragg-Williams model, with
µ2 s ↔ qJ ; see Problem 12.4. It follows that, by virtue of this term, the
system will undergo an order-disorder transition at a critical temperature
Tc = µ2 s/k. The second term in the Hamiltonian will contribute towards
the short-range order in the system.
We also note that the root-mean-square value of B in this model is (s/βN )1/2
which, for a given value of s, is negligibly small when N is large. The
order-disorder transition is made possible by the fact that the resulting
interaction is of an infinite range.

13.6. Since the Hamiltonian H{τi } of the present model is formally similar to
the Hamiltonian H{σi } of Sec. 13.2, the partition function of this system
can be written down in analogy with eqn. (13.2.10):
1 h 1/2 i
ln Q ≈ ln eβJ2 cosh(βJ1 ) + e−2βJ2 + e2βJ2 sinh2 (βJ1 )

,
N
from which the various thermodynamic properties of the system can be
derived. In particular, we get

1 ∂ sinh(βJ1 )
σi σi+1 = ln Q =  1/2 , and
βN ∂J1 e −4βJ 2
2 + sinh (βJ )
1
1 ∂
σi σi+2 = ln Q = 1−
βN ∂J2
2e−4βJ2
1/2
h 1/2 i .
+ sinh2 (βJ1 ) cosh(βJ1 ) + e−4βJ2 + sinh2 (βJ1 )
 
e−4βJ2

As a check, we see that in the limit J2 → 0 these expressions reduce to


tanh(βJ1 ) and tanh2 (βJ1 ), respectively; on the other hand, if J1 → 0,
they reduce to 0 and tanh(βJ2 ) instead.
13.7. In the symmetrized version of this problem, the transfer matrix P is
given by
 

0 0
0 0
 1 0 0

σi , σi |P|σi+1 , σi+1 = exp K1 σi σi+1 + σi σi+1 + K2 σi σi + σi+1 σi+1 ,
2

where K1 = βJ1 and K2 = βJ2 . Since (σi , σi0 ) = (1, 1), (1, −1), (−1, 1) or
(−1, −1), we get

e−2K1 +K2
 2K +K 
e 1 2 1 1
 1 e2K1 −K2 e−2K1 −K2 1 
(P) = 
e−2K1 −K2 e2K1 −K2

 1 1 
e−2K1 +K2 1 1 e2K1 +K2
122

The eigenvalues of this matrix are


 
λ1 1h i
= (A + B + C + D) ± {(A − B + C − D)2 + 16}1/2 ,
λ2 2
λ3 = A − C, λ4 = B − D,

where

A = e2K1 +K2 , B = e2K1 −K2 , C = e−2K1 +K2 , D = e−2K1 −K2 .

Since λ1 is the largest eigenvalue of P,


 h i
1 1
ln Q ≈ ln λ1 = ln (A + B + C + D) + {(A − B + C − D)2 + 16}1/2 .
N 2
Substituting for A, B, C and D, we obtain the quoted result. The study
of the various thermodynamic properties of the system is now straightfor-
ward.
13.8. In the notation of Sec. 13.2, the transfer matrix of this model is

hσi |P|σi+1 i = exp(βJ σi σi+1 ) (σi = −1, 0, 1).

It follows that
eK 1 e−K
 

(P) =  1 1 1  (K = βJ),
e−K 1 eK

with eigenvalues
 
λ1 1h i
= (1 + 2 cosh K) ± {8 + (2 cosh K − 1)2 }1/2 , λ3 = 2 sinh K.
λ2 2
Since λ1 is the largest eigenvalue of P,
 h i
1 1
ln Q ≈ ln λ1 = ln (1 + 2 cosh K) + {8 + (2 cosh K − 1)2 }1/2 ,
N 2
which leads to the quoted expression for the free energy A.
In the limit T → 0, K → ∞, with the result that cosh K ≈ 12 eK and hence
A ≈ −NJ ; this corresponds to a state of perfect order in the system, with
U = −NJ and S = 0. On the other hand, when T → ∞, cosh K → 1 and
hence A → −NkT ln 3; this corresponds to a state of complete randomness
in a system with 3N microstates.
13.9. (a) Making use of the correspondence established in Sec. 12.4, we obtain
for a one-dimensional lattice gas (q = 2).
(i) The fugacity z = e−4βJ+2βµB = η 2 y, where η = e−2βJ and
y ↔ e2βµB .
123

(ii) The pressure P = −(A/N ) − J + µB; using eqn. (13.2.11), this


becomes
h i
P = kT ln cosh(βµB) + {e−4βJ + sinh2 (βµB)}1/2 + µB. (1)

(iii) The density (1/v) = 12 {1 + (M /N µ)}; using eqn. (13.2.13), this


becomes
 
1 1 sinh(βµB)
= 1 + −βJ . (2)
v 2 {e + sinh2 (βµB)}1/2

Our next step consists in eliminating the magnetic variable (βµB) in


favor of the fluid variable y. For this, we note that
1 1/2
cosh(βµB) = (y + y −1/2 ) = (y + 1)/2y 1/2 ,
2
1
sinh(βµB) = (y 1/2 − y −1/2 ) = (y − 1)/2y 1/2 ,
2
while βµB = 21 ln y. Substituting these results into eqns. (1) and (2),
we obtain the quoted expressions for P/kT and 1/v. It may be
verified that these expressions satisfy the thermodynamic relation
     
1 ∂ P ∂ P
=z =y
v ∂z kT T ∂y kT η

At high temperatures, η → 1 and we get


P 1 y
≈ ln(y + 1), ≈ .
kT v y+1
Moreover, in this limit y ' z  1. One may, therefore, write
P 1 Pv
≈ y, ≈ y and hence ≈ 1.
kT v kT
At low temperatures, η becomes very small and y very large — with
the result that
P η2 1 η2
≈ ln y + , ≈1− .
kT y v y

(b) For a hard-core lattice gas (y → 0, η → ∞), we get


 √  √
P 1 + 1 + 4z 1 + 4z − 1
= ln , ρ= √ .
kT 2 2 1 + 4z

From the second equation, it follows that 1 + 4z = 1/(1 − 2ρ). Sub-
stituting this into the first equation, we obtain the quoted expression for
P (ρ).
124

13.10. Applying eqn. (12.11.11) to a one-dimensional system, we get


∞ ∞
χ0 X
−(a/ξ)|x|
X
= e = 1 + 2 e−(a/ξ)x
N βµ2 x=−∞ x=1

e−a/ξ 1 + e−a/ξ
 
a
=1+2 = = coth .
1 − e−a/ξ 1 − e−a/ξ 2ξ

For ξ  a, coth(a/2ξ) ≈ (2ξ/a), making χ0 ∝ ξ 1 — consistent with the


fact that for this system (2 − η) = 1. For ξ  a, we recover the familiar
result: χ0 ≈ N µ2 /kT .
For an n-vector model, eqn. (13.3.17) leads to the result

χ0 1 + In/2 (βJ)/I(n−2)/2 (βJ)


2
= ,
N βµ 1 − In/2 (βJ)/I(n−2)/2 (βJ)

which agrees with the expression quoted in the problem. For the special
case n = 1, the ratio I1/2 (x)/I−1/2 (x) = tanh x, whence

χ0 1 + tanh(βJ) cosh(βJ) + sinh(βJ)


2
= = = e2βJ ,
N βµ 1 − tanh(βJ) cosh(βJ) − sinh(βJ)

in agreement with eqn. (13.2.14).


2 2 2 2
13.11. We introduce an extra factor σk+1 . . . σ`−1 σm+1 . . . σn−1 , which is identi-
cally equal to 1, and write

σk σ` σm σn = (σk σk+1 ) . . . (σ`−1 σ` )(σm σm+1 ) . . . (σn−1 σn ).

Following the same procedure that led to eqn. (13.2.31), we now get
`−1
Y n−1
Y
σk σ` σm σn = tanh(βJi ) tanh(βJi ).
i=k i=m

Employing a common J, we obtain the desired result

σk σ` σm σn = {tanh(βJ)}`−k {tanh(βJ)}n−m = {tanh(βJ)}n−m+`−k

13.12. For a complete solution to this problem, see Thompson (1972b), sec. 6.1,
pp. 147–9.

13.13. With J 0 = 0, we get a much simpler result, viz.

Zπ Zπ
1 1
ln Q = ln 2 + 2 ln{cosh(2γ) − sinh(2γ) − cos ω}dωdω 0 .
N 2π
0 0
125

The integration over ω 0 is straightforward; the one over ω can be done


with the help of the formula
Zπ " √ #
a + a2 − b2
ln(a − b cos ω)dω = π ln (|b| < a), (1)
2
0

which yields the expected result


 
1 1 cosh(2γ) + 1
ln Q = ln 2 + ln = ln(2 cosh γ).
N 2 2

With J 0 = J, we have
Zπ Zπ
1 1
ln Q = ln 2 + 2 ln{cosh2 (2γ) − sinh 2γ(cos ω + cos ω 0 )}dωdω 0 .
N 2π
0 0
(2)
We substitute ω = θ + ϕ and ω 0 = θ − ϕ; this will replace the sum
(cos ω + cos ω 0 ) by the product 2 cos θ cos ϕ and the element dωdω 0 by
2dθdϕ. As for the limits of integration, the periodicity of the integrand
allows us to choose the rectangle [0 ≤ θ ≤ π, 0 ≤ ϕ ≤ π/2] without
affecting the value of the integral. We thus have

Zπ Zπ/2
1 1
ln Q = ln 2 + 2 ln{cosh2 (2γ) − 2 sinh(2γ) cos ϕ cos θ}dθdϕ.
N π
0 0

Integration over θ may now be carried out using formula (1), with the
result
Zπ/2   
1 1 1
q
2 4 2 2
ln Q = ln 2 + ln cosh (2γ) + cosh (2γ) − 4 sinh (2γ) cos ϕ dϕ
N π 2
0
  Zπ/2 n
1 1 p o
= ln 2 · √ cosh(2γ) + ln 1 + 1 − κ2 cos2 ϕ dϕ,
2 π
0

where κ is given by eqn. (13.4.23). Finally, we replace cos2 ϕ by sin2 ϕ


(without affecting the value of the integral) and recover eqn. (13.4.22).
We know that this model is singular at κ = 1. A close look at the integral
in (2) shows that the singularity arises when contributions in the neigh-
borhood of the point ω = ω 0 = 0 pile up. Since cos ω and cos ω 0 are almost
unity there, the situation becomes catastrophic when cosh2 2γ = 2 sinh 2γ,
i.e. when κ = 1. In the anisotropic case, a similar observation suggests
that the singularity will arise when

cosh(2γ) cosh(2γ 0 ) = sinh(2γ) + sinh(2γ 0 ).


126

Squaring both sides of this equation and simplifying, we get


sinh(2γ) sinh(2γ 0 ) = 1 (3)
as the criterion for the onset of phase transition in this system; cf. eqn. (13.4.18).
The study of the thermodynamic behavior of the system in the neighbor-
hood of the critical point in the general case, when J 0 6= J, is highly
complicated; the fact, however, remains that the internal energy U0 is
continuous and the specific heat C0 displays a logarithmic divergence at
a critical temperature Tc given by eqn. (3).
13.14. As κ → 1, the first integral tends to the limit
Zπ/2 Zπ/2
1 − sin ϕ cos ϕ π/2
dϕ = dϕ = ln(1 + sin ϕ)|0 = ln 2.
cos ϕ 1 + sin ϕ
0 0

The second integral diverges as κ → 1, so we need to examine it carefully.


Setting cos ϕ = x, this integral takes the form
Z1 Z1 √ o 1
κ dx dx n
' √ = ln x + κ02 + x2 .

p
(1 − κ2 ) + κ2 x2 κ02 + x2 0
0 0
02
Since κ is close to 1, κ  1; we, therefore, get for this integral the
asymptotic result ln 2 − ln |κ0 |. It follows that K1 (κ) ≈ 2 ln 2 − ln |κ0 | =
ln(4/|κ0 |).
13.15. The quantity to be evaluated here is
   
Sc U A
= − .
Nk NkT c NkT c

The first term, by eqn. (13.4.28), is −Kc coth(2Kc ) = − 2Kc . The sec-
ond, by eqn. (13.4.22), is
Zπ/2
1
ln 2 + ln(1 + cos ϕ)dϕ
π
0
 
Zπ/2
1 π/2 ϕ sin ϕ
= ln 2 + ϕ ln(1 + cos ϕ)|0 + dϕ .

π 1 + cos ϕ
0

The integrated part vanishes while the remaining integral has the value
−(π/2) ln 2+2G; see Gradshteyn and Ryzhik (1965), p. 435. Thus, finally,
Sc √ 1
= − 2Kc + ln 2 + 2G/π ' 0.3065.
Nk 2
The corresponding result under the Bethe approximation is 2 ln 3−(7/3) ln 2 '
0.5799 and that under the Bragg-Williams approximation is ln 2 ' 0.6931.
127

13.16. Expanding around K = Kc , we get

sinh(2K) = sinh(2Kc ) + 2 cosh(2Kc )(K − Kc ) + 2 sinh(2Kc )(K − Kc )2 + . . .



= 1 + 2 2(K − Kc ) + 2(K − Kc )2 + . . . . (1)

Since (K − Kc ) = Kc {(1 + t)−1 − 1} = Kc (−t + t2 − . . .), eqn. (1) becomes


√  √ 
sinh(2K) = 1 − 2 2Kc t + 2 2Kc + 2Kc2 t2 + . . . . (2)

Raising expression (2) to the power −4, we get


√  √ 
{sinh(2K)}−4 = 1 + 8 2Kc t − 8 2Kc + 8Kc2 − 80Kc2 t2 + . . . ,

so that
√  √ 
1 − {sinh(2K)}−4 = −8 2Kc t + 8 2Kc − 72Kc2 t2 + . . .

   
9
= 8 2Kc |t| 1 + 1 − √ Kc |t| + . . . . (3)
2

Taking the eighth-root of (3), we obtain the desired result.


13.17. Making use of the correspondence established in Sec. 12.4 and utilizing a
result obtained in Problem 13.15, we have for a two-dimensional lattice
gas (q = 4)
   
Pc A 2J 1 2G
=− − = ln 2 + − 2Kc .
kT c NkT c kT c 2 π

We also have: vc = 2 (because, at T = Tc , the spontaneous magnetization


of the corresponding ferromagnet is zero). It follows that

Pc vc 4G
= ln 2 + − 4Kc ' 0.09659.
kT c π
Taking the reciprocal of this result, we obtain the one stated in the prob-
lem.
13.18. In one dimension, expression (13.5.31) assumes the form
Z∞
1 1
W1 (ϕ) = e−(1+ϕ)x I0 (x)dx = 2 1/2
= 2 .
{(1 + ϕ) − 1} (λ − J 2 )1/2
0

The constraint equation (13.5.19) then becomes

N 1 N µ2 β 2
+ = N, (1)
2β (λ2 − J 2 )1/2 4(λ − J)2
128

which agrees with the quoted result. Comparing (1) with the formal con-
straint equation (13.5.13), we conclude that

∂Aλ N 1 N µ2 B 2
= + , (2)
∂λ 2β (λ2 − J 2 )1/2 4(λ − J)2

It follows that
N n o N µ2 B 2
Aλ = ln λ + (λ2 − J 2 )1/2 + + C, (2)
2β 4(λ − J)

where C is a constant of integration. To determine C, we observe from


eqn. (13.5.12a) that, for B = 0 and J = 0, the partition function QN =
(π/βλ)N/2 — with the result that Aλ = (N/2β) ln(βλ/π). It follows that
C = (N/2β) ln(β/2π), which leads to the quoted result for Aλ .
With B = 0 but J 6= 0, eqn. (1) gives: (λ2 − J 2 )1/2 = 1/2β, and hence
λ = (1 + 4β 2 J 2 )1/2 /2β. Equation (13.5.15) then gives

(1 + 4β 2 J 2 )1/2 + 1
 
βAS βAλ 1 1
= − βλ = ln − (1 + 4β 2 J 2 )1/2 .
N N 2 4π 2

13.19. With βJi = β · nJ 0 = nK , eqn. (13.3.8) becomes


N −1
Γ(n/2)
QN (nK ) = I (nK ) .

(n−2)/2 (n−2)/2
1

2 nK

For n, N  1, we get
 n n 1  
1 1 n
ln QN (nK ) ≈ ln Γ − ln nK + ln In/2 · 2K
nN n 2 2 2 2
  
1 n n n n 1
= ln − + . . . − ln nK
n 2 2 2 2 2
(4K 2 + 1)1/2 + 1
   
n 2 1/2
+ (4K + 1) − ln + ...
2 2K
(4K 2 + 1)1/2 + 1
  
1
≈ (4K 2 + 1)1/2 − 1 − ln .
2 2

13.20. By eqn. (13.5.69), we have for the spherical model at T < Tc

N µ2 N 2 µ2 (K − Kc )
χ0 = ≈ .
2 Jϕ J

Replacing (K − Kc ) by m20 K, see eqn. (13.5.44), and remembering that


K = J/kB T , we obtain the desired result.
129

13.21. Before employing the suggested approximation, we observe that the major
contribution to the integral over θj in expression (13.5.58) comes from
those values of θj that are either close to the lower limit 0 or close to the
upper limit 2π. We, therefore, write this integral in the form

Nj
2 cos(Rj θj / a)e−x+x cos θj dθj ,

0

so that the major contribution now comes only from those values of θ that
are close to 0. We may, therefore, replace (1 − cos θj ) by θj2 /2 and, at the
same time, replace the upper limit of the integral by ∞. This yields the
asymptotic result

2 Nj Nj −Rj2 /2a2 x
2 cos(Rj θj / a)e−xθj /2 dθj = √ e ,
2π 2πx
0

where use has been made of formula (B.41). Equation (13.5.57) then
becomes
Z∞
1 N 2 2
G(R) ≈ d/2
e−ϕx−R /2a x dx ,
2N βJ (2πx)
0

which is precisely the expression that led to eqns. (13.5.61 and 62).
In the case of the Bose gas, we are concerned with the expression

exp(ik · R)
Z
1
dd k,
(2π)d eα+β~2 k2 /2m − 1

see Section 13.6, which may now be approximated by

exp(ik · R) d
Z
1
d
d k.
(2π) α + β~2 k 2 /2m

Using the representation (13.5.27), this may be written as

Z∞ d Z∞
 
Y 1 2 2
e−αx  eik n Rn −β~ kn x/2m dk n  dx
n=1

0 −∞
Z∞ " d   # "  1/2 #
−αx
Y 1 2 2 2πβ
= e √ e−πRn /λ x dx λ=~
n=1
λ x m
0
Z∞
1 2
/λ2 x 1
= d e−αx−πR dx ,
λ xd/2
0

which is precisely the expression that led to eqns. (13.6.35 and 36).
130

13.22. The constraint equation (13.5.21) now takes the form


X  −1
1
2N β(1 − m2 ) = J ϕ + k σ aσ ,
2
k

which may as well be written as


X −1
2 −1 1
2K(1−m ) = F (ϕ), where K = βJ and F (ϕ) = N ϕ + k σ aσ .
2
k

To determine the behavior of the function F (ϕ) at small ϕ, we look at the


derivative −2
X 1 σ σ
0 −1
F (ϕ) = −N ϕ+ k a .
2
k

In view of eqns. (13.5.17) and (C.7b), we obtain


d   Z∞  −2
0 −1
Y Nj a 1 σ σ 2π d/2 d−1
F (ϕ) ≈ −N ϕ+ k a k dk
j=1
2π 2 Γ(d/2)
0

d Z∞  −2
a 1 σ σ
=− 1+ k a k d−1 dk .
2d−1 π d/2 Γ(d/2)ϕ2 2ϕ
0

σ σ
We substitute k = (2ϕ/a )x and get
Z∞
0 (2ϕ)d/σ x(d/σ)−1 dx
F (ϕ) ≈ −
2d−1 π d/2 Γ(d/2)σϕ2 (1 + x)2
0
2d/σ Γ(d/σ)Γ(2 − d / σ) (d−2σ)/σ
=− ϕ (d < 2σ).
2d−1 π d/2 Γ(d/2)σ
Integrating over ϕ, we obtain

2d/σ Γ{(d/σ)/σ}Γ{(2σ − d)/σ} (d−2σ)/σ


F (ϕ) ≈ F (0) − ϕ ;
2d−1 π d/2 Γ(d/2)σ
cf. eqn. (G.7c). The function F (0) exists for all d > σ and may be
identified with 2Kc , leading to the constraint equation

2K(1 − m2 ) = 2Kc − const. ϕ(d−σ)/σ (σ < d < 2σ).

The critical exponents of the model follow straightforwardly from this


equation. The first one to emerge is β = 1/2, as before. Next, γ =
σ/(d−σ), whence α = 2−2β −γ = (d−2σ)/(d−σ), while δ = 1+(γ/β) =
(d + σ)/(d − σ). Next, from the very starting form of the function F (ϕ),
we infer that the correlation length ξ ∼ ϕ−1/σ and hence ∼ t−1/(d−σ) ; it
follows that ν = 1/(d − σ). We then get: η = 2 − (γ/ν) = 2 − σ.
131

For d > 2σ, the derivative F 0 (0) exists — with the result that

F (ϕ) ≈ F (0) − |F 0 (0)|ϕ.

This leads to mean-field results for the exponents β, γ, α and δ. The


correlation length is, once again, given by ξ ∼ ϕ−1/σ but now this is
∼ t−1/σ , so that ν = 1/σ (and not 1/2); accordingly, η is, once again,
2 − σ (and not 0).
13.23. The derivation of eqns. (13.6.9 and 11) proceeds exactly as of eqns. (7.1.36
and 37). The derivation of eqns. (13.6.10 and 13) proceeds exactly as
in Problem 7.4; eqn. (13.6.12) then follows as a product of expressions
(13.6.11 and 13): The derivation of eqns. (13.6.14 and 15) proceeds exactly
as in Problem 7.5; note that one may first obtain here

1 g(d−2)/2 (z) d gd/2 (z)


κT = , κS = ,
nk B T gd/2 (z) (d + 2)nk B T g(d+2)/2 (z)

and then use eqn. (13.6.7) to express these quantities in terms of P . Fi-
nally, the derivation of eqn. (13.6.23) proceeds exactly as in Problem 7.6.
13.24. The derivations here proceed exactly as in Problem 7.7. The singularity
of these quantities arises from the last term of the two expressions, and
is qualitatively similar to the singularity of the quantity (∂CV /∂T ). Note

that the singularity of the combination v(∂ 2 P/∂T 2 )v − (∂ 2 µ/∂T 2 )v is
relatively mild.
13.25. By definition,
     
∂S ∂S ∂V
CP ≡ T =T .
∂T P ∂V P ∂T P

Using the Maxwell relation (∂S/∂V )P = (∂P/∂T )S , we get


           
∂P ∂V ∂P ∂V ∂P
CP = T =T − =
∂T S ∂T P ∂T S ∂P T ∂T V
     
∂P ∂P
T V κT .
∂T S ∂T V
Next      
∂S ∂S ∂P
CV ≡ T =T .
∂T V ∂P V ∂T V
Now, using the Maxwell relation (∂S/∂P )V = −(∂V /∂T )S , we get
         
∂V ∂P ∂V ∂P ∂P
CV = −T = −T =
∂T S ∂T V ∂P S ∂T S ∂T V
   
∂P ∂P
T (V κS ) .
∂T S ∂T V
132

In the two-phase region, (∂P/∂T )S = (∂P/∂T )V = dP /dT , with the


result that

CP = VT (dP / dT )2 κT , CV = VT (dP / dT )2 κS .

Now, by eqn. (13.6.28), dP/dT, at T < Tc , = (d + 2)P/2T . Using this


result and eqn. (13.6.15), we get
 2
(d + 2)P d d(d + 2) PV
CV = VT = .
2T (d + 2)P 4 T

Substituting for P from eqn. (13.6.28), we recover expression (13.6.30) for


CV .
13.26. As shown in Problem 1.16, dµ = −sdT + vdP . It follows that
   
1 ∂v ∂v
κT ≡ − =− .
v ∂P T ∂µ T

Since v = 1/ρ, we readily obtain the desired result for κT .


For the ideal Bose gas at T < Tc , the particle numbers N0 and Ne are
given by eqns. (13.6.24) and (13.6.27). Since α ≡ −µ/kB T , we have

N N0 Ne kB T ζ(d / 2)
ρ= = + =− + .
V V V Vµ λd
It follows that
2
N02 V ρ20
  
1 kB T 1 kB T N0
κT = 2 = 2 − = = .
ρ V µ2 ρ V kB T ρ2 Vk B T ρ2 kB T

Incidently, using the relationship between CP and κT , as developed in


Problem 13.25, we can show that, in the two-phase region of the Bose gas,
 2  d  2
CP d + 2 ζ{(d + 2)/2} T ρ0
=N .
N kB 2 ζ(d/2) Tc ρ

Comparing this with eqn. (13.6.30), we find that, in this region, the ratio
CP /CV = O(N ).

13.28. Use the relations N/(L − N D) = βP and


 z n
lim 1 + = exp(z),
n→∞ n
and collect factors.
133

13.29. Using equation (10.7.20a) and ignoring the delta–function contribution to


S(k) gives
 
∞ Z ∞
n X 1
S(k) = 1 +  (βP (x − jD))j exp (−βP (x − jD) + ikx) + c.c. .
βP (1 − nD) j=1 (j − 1)! jD

Using n/(βP (1 − nD)) = 1 we get

βP eikD
 
S(k) = 1 + + c.c.
βP − ik − βP eikD
k2
= 2
k + 2(βP )2 (1 − cos(kD)) + 2βP k sin(kD)

13.30. The isobaric partition function is


 Z ∞
βmω 2
  
1 2
YN (P, T ) = exp −βP y − (y − a) dx ,
λ −∞ 2
NP2
 
kT
G(N, P, T ) = N kT ln + NPa − ,
~ω 2mω 2

which gives L = (∂G/∂P )T = N a − P/mω 2 . As P → 0 the length
goes to N a, i.e. N times the equilibrium length of one spring. However,
the masses and springs do not form a long-range-ordered lattice since the
variance of the neighbor distances grows with n.
P
hxn − x0 i = nhyi = a −
mω 2
kT
h(xn − x0 )2 i − hxn − x0 i2 = n hy 2 i − hyi2 = na2

.
mω 2 a2
The heat capacity is CP = N k.

13.31. Here is a C code snippet that performs the calculation.

int L=4;
int n=L*L;
int* s=new int[n]; //spins: 0 or 1
int* i1 = new int[n]; //neighbors to the right
int* i3 = new int[n]; //neighbors above
for (int i=0;i<n;i++)
{
i1[i] = i+1; // site to right
i3[i] = i+L; // site above
if ((i1[i] % L) == 0) i1[i] -= L; //implement periodic boundary conditions
if (i3[i] >= n) i3[i] -= n; //implement periodic boundary conditions
134

}
double* hist=new double[n+1]; // histogram: double since would overflow integers for L>5;
for (int e=0;e<=n;e++) hist[e]=0.0;
double nconfig=pow(2.0,n); // number of configurations is 2 ∧ n
for (double iconfig=0.0; iconfig < nconfig; iconfig += 1.0)
{
double state = iconfig;
for (int i=0;i<n;i++) // determine spins (0 or 1) for each site
{
s[i]=(int) fmod(state,2.0);
state = floor(0.5*state);
}
int e=0;
for (int i=0;i<n;i++) // count energy above ground state (unequal neighbors)
{
e += (s[i] != s[i1[i]]);
e += (s[i] != s[i3[i]]);
}
hist[e/2] += 1.0; //increment histogram of energies, only even values needed
}
for (int e=0;e<=n;e++) cout << e << ” ” << hist[e] << endl; //output the results

This code gives the coefficients in the problem. Since the L = 6 case
involves 236 configurations, you might try a bitwise calculation with each
row represented by an integer between 0 and 2L − 1 and the spins repre-
sented by the bits. This is computationally much more efficient since the
energy can be determined easily using simple bit rotations and exclusive
ors.

13.32. Separating off the ground state energy gives


 
1 x
T =
x 1

which has eigenvalues λ = 1 ± x. Therefore

QN = (1 + x)N + (1 − x)N ,

which can be expanded

N  N/2 
N !xj N !(−x)j
 
X X n!
QN = + = 2 x2k
j=0
j!(N − j)! j!(N − j)! (2k)!(N − 2k)!
k=0
135

13.33. The 8 × 8 partition function coefficients are

g = {2, 0, 128, 256, 4672, 17920, 145408, 712960, 4274576, 22128384, 118551552, 610683392,
3150447680, 16043381504, 80748258688, 396915938304, 1887270677624, 8582140066816,
36967268348032, 149536933509376, 564033837424064, 1971511029384704,
6350698012553216, 18752030727310592, 50483110303426544, 123229776338119424,
271209458049836032, 535138987032308224, 941564975390477248, 1469940812209435392,
2027486077172296064, 2462494093546483712, 2627978003957146636,
2462494093546483712, 2027486077172296064, 1469940812209435392,
941564975390477248, 535138987032308224, 271209458049836032,
123229776338119424, 50483110303426544, 18752030727310592, 6350698012553216,
1971511029384704, 564033837424064, 149536933509376, 36967268348032,
8582140066816, 1887270677624, 396915938304, 80748258688, 16043381504, 3150447680,
610683392, 118551552, 22128384, 4274576, 712960, 145408, 17920, 4672, 256, 128, 0, 2}

The specific heats for the 8 × 8, 16 × 16 and 32 × 32 Ising model are shown
below.
C
Nk
3.0

2.5

2.0

1.5

1.0

0.5

kT
0.0
0 1 2 3 4 J

Notice that the height of the peak grows linearly with ln(L).
136

13.34. The specific heat for the 64 × 64 Ising model is shown below.
C
Nk
3.0

2.5

2.0

1.5

1.0

0.5

kT
0.0
0 1 2 3 4 J
Chapter 14

14.1. We start with expression (13.2.3) for the partition function QN and carry
out summation over σ2 , σ4 , . . .. The resulting expression will consist of
1
2 N factors such as
X
hσ1 |P|σ2 ihσ2 |P|σ3 i = hσ1 |P2 |σ3 i,
σ2

and will be formally similar to the expression we started with. Calling the
new transfer operator P0 {K0 }, we clearly get eqn. (1) of the problem.
From the given expression for P{K}, we readily get
+ e−2K1 eK2 + e−K2
 2(K +K ) 
2K0 e
1 2
0 0
P {K } = e .
eK2 + e−K2 e2(K1 −K2 ) + e−2K1
Expressing this in a form similar to eqn. (2), we obtain
0 0 0
eK0 +K1 +K2 = e2K0 {e2(K1 +K2 ) + e−2K1 }
0 0 0
eK0 +K1 −K2 = e2K0 {e2(K1 −K2 ) + e−2K1 }, and
0 0
eK0 −K1 = e2K0 {eK2 + e−K2 }2
which are identical with eqns. (14.2.7) and will lead precisely to eqns. (14.2.8).
14.2. We’ll do the second part only, for it includes the first as a special case.
For this, we have to show that the given function f (K1 , K2 ) satisfies the
functional equation (14.2.11). Now, the right-hand side of this equation is
  
0 0 0 0

0 0 0 0
!2 1/2 
K +K K −K K +K K −K

1  0 e 1 2 +e 1 2
 0 e 1 2 −e 1 2
 
− ln eK0  + e−2K1 +  .

2   2  2  

Substituting from eqns. (14.2.7) with K0 = 0, this becomes


1
ln[eK2 cosh(2K1 + K2 ) + e−K2 cosh(2K1 − K2 ) + 4 cosh2 K2 +


2
1/2
(eK2 cosh(2K1 + K2 ) − e−K2 cosh(2K1 − K2 ))2 ]
1
= − ln[e2K1 cosh(2K2 ) + e−2K1 + {4 cosh2 K2 + (e2K1 sinh(2K2 ))2 }1/2 ].
2

137
138

The left-hand side of the same equation is


− ln[eK1 cosh K2 + {e−2K1 + e2K1 sinh2 K2 }1/2 ]
1 
= − ln e2K1 cosh2 K2 + e−2K1 + e2K1 sinh2 K2
2 i
+2eK1 cosh K2 {e−2K1 + e2K1 sinh2 K2 }1/2
1
= − ln[e2K1 cosh(2K2 ) + e−2K1 + {4 cosh2 K2 + 4e4K1 cosh2 K2 sinh2 K2 }1/2 ],
2
which is precisely the same as the right-hand side.
14.3. We’ll do the second part only, for it includes the first as a special case. For
this, we have to show that the given function f (K1 , K2 , Λ) satisfies the
functional equation (14.2.27). Now, the right-hand side of this equation is
" p #
1 0 1 Λ0 + Λ02 − K102 K202
− K0 + ln − .
2 4 2π 8 (Λ0 − K10 )

Substituting from eqns. (14.2.25 and 28) with K0 = 0, this becomes


"  p #
1  π  K22 1 Λ − K12 /2Λ + Λ2 − K12
− ln − + ln
4 Λ 8Λ 4 2π
2
 
K2 K1
− 1+
8(Λ − K1 ) Λ
"  p #
2
1 2Λ Λ − K1 /2Λ + Λ2 − K12
= ln .
4 2π 2π
K22
 
Λ − K1 Λ + K1
− +
8(Λ − K1 ) Λ Λ
" p #
1 Λ + Λ2 − K12 K22
= ln − .
2 2π 4(Λ − K1 )

which is precisely f (K1 , K2 , Λ).


14.4. Making the suggested substitution into eqn. (14.2.24), we get
 0 
Z Z N    N 0 /2
X
0 0 2Λ 0 0 0 2Λ 02 2Λ
QN = · · · exp  K0 + K1 · s s −Λ · s 
j=1
K1 j j+1 K1 j K1

ds 01 . . . ds 0N .
In view of eqns. (14.2.25), with K0 = K2 = 0, we now have
 1/2    1/2  1/2
0 2Λ π 1/2 2Λ 2π 00
eK0 · = · = = eK0 , say,
K1 Λ K1 K1
2Λ 2Λ 2Λ2
K10 · = K1 , and Λ0 · = − K1 = Λ00 , say.
K1 K1 K1
139

The resulting expression for QN , when compared with eqn. (14.2.19), leads
to the functional equation
1 1
f (K1 , Λ) = − K000 + f (K1 , Λ00 ), (1)
2 2
where
2Λ2
 
1 2π
K000 = ln and Λ00 = − K1 . (2)
2 K1 K1
To verify that the function (14.2.32) satisfies the functional equation (1),
we note that the right-hand side of this equation is
" p #
Λ00 + Λ002 − K12
 
1 2π 1
− ln + ln
4 K1 4 2π
" p #
1 K1 (2Λ2 /K1 − K1 ) + 4Λ4 /K12 − 4Λ2
= ln ·
4 2π 2π
" p # " p #
1 2Λ2 − K12 + 2Λ Λ2 − K12 1 Λ + Λ2 − K12
= ln = ln ,
4 4π 2 2 2π

which is precisely f (K1 , Λ).


14.5. For a solution to this problem, see Kadanoff (1976a).
14.6. The eigenvalues λ1 and λ2 of the matrix A∗` are determined by the equation

a11 − λ a12
= 0.
a21 a22 − λ

Clearly,

λ1 + λ2 = a11 + a22 , while λ1 λ2 = a11 a22 − a12 a21 . (1)

The eigenfunctions ϕ1 and ϕ2 are given by


 
x1 y1 λ1 − a11 a21
ϕ1 = const , where = = , and (2)
y1 x1 a12 λ1 − a22
 
x2 y2 λ2 − a11 a21
ϕ2 = const , where = = , (3)
y2 x2 a12 λ2 − a22

(a) Now, by eqn. (14.3.13a),

k1 = u1 x1 + u2 x2 , k2 = u1 y1 + u2 y2 , where
k1 y2 − k2 x2 k1 y1 − k2 x1
u1 = , u2 = .
x1 y2 − y1 x2 x2 y1 − y2 x1
It follows that the slope of the line u1 = 0 in the (k1 , k2 )-plane is
m1 = y2 /x2 , which is given by eqn. (3), while the slope of the line
140

u2 = 0 is m2 = y1 /x1 , which is given by eqn. (2). We readily see


that the product
λ2 − a11 λ1 − a11 λ2 λ1 − a11 (λ2 + λ1 ) + a211
m1 m2 = · =
a12 a12 a212
Substituting from eqns. (1), we get
(a11 a22 − a12 a21 ) − a11 (a11 + a22 ) + a211 a21
m1 m2 = =− .
a212 a12
It follows that the two lines will be mutually perpendicular if and
only if a12 = a21 .
(b) If a12 or a21 = 0, then by eqns. (1), λ1 = a11 and λ2 = a22 . The
stated results then follow straightforwardly.
(c) If a11 = 0, then m1 = −a21 /λ1 = λ2 /a12 and m2 = −a21 /λ2 =
λ1 /a12 . On the other hand, if a22 = 0, then m1 = a21 /λ2 = −λ1 /a12
and m2 = a21 /λ1 = −λ2 /a12 .
14.7. In the limit n → ∞, we obtain from eqns. (14.4.38–40)
1 1 3 1 1
ν≈ + ε, ∆ ≈ + ε, α ≈ − ε,
2 4 2 2 2
1 1
β ≈ , γ ≈ 1 + ε, δ ≈ 3 + ε, η ≈ 0.
2 2
At the same time, we obtain directly from eqns. (13.5.47, 66 and 67), with
d = 4 − ε where 0 < ε  1,
−ε 1 2 1
α' , β= , γ= ' 1 + ε,
2 2 2−ε 2
1 − 16 ε
 
6−ε 1
δ= =3 ' 3 1 + ε = 3 + ε, η = 0,
2−ε 1 − 12 ε 3
 
1 1 1 1 1
ν= ' 1 + ε = + ε.
2−ε 2 2 2 4
To the given order in ε, the two sets of results are in complete agreement.
14.8 & 9. For d = 4 − ε where 0 < ε  1, eqn. (14.4.46) gives

sin(π − πε/2)Γ(3) 1
Sd ≈ ≈ ε.
2π{Γ(2)}2 2
Equations (14.4.43–45) then give
4ε · 21 ε 1 ε2
η' = ,
4 n 2n     
2 3ε 1 3ε 1 3
γ' 1− ' 1+ ε 1− '1+ − ε,
2−ε n 2 n 2 n
   
ε 12ε 1 ε 12
α'− 1− '− 1− .
2−ε ε n 2 n
141

Next, we obtain
 
1 1 2(2d − 5)Sd 1 1
β = (2 − α − γ) = − +O
2 2 d−2 n n2
1 3
' − ε,
2 2n
γ 4 1 − 6Sd /n + O(1/n2 )
δ =1+ =1+
β d − 2 1 − 4(2d − 5)/(d − 2) · (Sd /n) + O(1/n2 )
    
4 1 1
=1+ 1+O ' 1 + 2 1 + ε = 3 + ε,
2−ε n2 2
γ 1 1 − 6Sd /n + O(1 + n2 )
ν= =
2−η d − 2 1 − 2(4 − d)/d · (Sd /n) + O(1/n2 )
        
1 3ε 1 1 3ε 1 1 3
' 1− ' 1+ ε 1− ' 1+ − ε
2−ε n 2 2 n 2 2 n
 
1 1 6
= + 1− ε.
2 4 n

All these results agree with the corresponding ones following from eqns. (14.4.38–
40).
Chapter 15

15.1. (i) We multiply expression (15.1.11) by ∆T , take its average and utilize
relations (15.1.14), to obtain (∆T ∆S) = kT .
(ii) We multiply expression (15.1.12) by ∆V , take its average and utilize
relations (15.1.14), to obtain (∆P ∆V ) = −kT .
(iii) We multiply expression (15.1.11) by ∆V , take its average and utilize
relations (15.1.14), to obtain
       
∂P 1 ∂V ∂V
∆S∆V = kT − V = kT .
∂T V V ∂P T ∂T P

(iv) We multiply expression (15.1.12) by ∆T , take its average and utilize


−1
relations (15.1.14), to obtain (∆P ∆T ) = kT 2 CV (∂P/∂T )V .
15.2. If we choose ∆S and ∆P as our independent variables, then
     
∂T ∂T T ∂T
∆T = ∆S + ∆P = ∆S + ∆P, and
∂S P ∂P S CP ∂P S
     
∂V ∂V ∂T
∆V = ∆S + ∆P = ∆S − V κS ∆P.
∂S P ∂P S ∂P S

It follows that
T
−∆T ∆S + ∆P ∆V = − (∆S)2 − V κS (∆P )2 ,
CP
which converts expression (15.1.8) into (15.1.15), leading directly to ex-
pressions (15.1.16) for (∆S)2 , (∆P )2 and (∆S∆P ).
For an independent evaluation of these averages, we proceed as in Prob-

142
143

lem 15.1. From eqns. (15.1.11, 12 and 14), we readily obtain


2 " 2 #
2
 
C ∂P ∂P
(∆S)2 = V2 (∆T )2 + (∆V )2 = k CV + TV κT = kC P ,
T ∂T V ∂T V
 2 "  2 #
2
∂P 2
1 2
kT ∂P CV
(∆P ) = (∆T ) + 2 2 (∆V ) = T +
∂T V κT V CV ∂T V κT V
kT CP kT
= · = , and
CV κT V κS V
   
CV ∂P ∂P 1
(∆S∆P ) = (∆T )2 − (∆V )2 = 0.
T ∂T V ∂T V κT V

15.3. We start with expression (15.1.6) and eliminate ∆S by writing


     2 
∂S ∂S 1 ∂ S
∆S = ∆E + ∆V + (∆E)2
∂E 0 ∂V 0 2 ∂E 2 0
 2   2  
∂ S ∂ S 2
+2 ∆E∆V + (∆V ) + . . . .
∂E∂V 0 ∂V 2 0

Replacing (∂S/∂E)0 by 1/T and (∂S/∂V )0 by P/T, and retaining terms


up to second order only, expression (15.1.6) takes the form
       
1 ∂θ 2 ∂θ ∂π ∂π 2
p ∝ exp (∆E) + 2 or ∆E∆V + (∆V ) ,
2k ∂E 0 ∂V ∂E 0 ∂V 0

where θ = 1/T and π = P/T . The covariance matrix of this distribution


is given by
   ∂θ ∂θ
−1
(∆E)2 (∆E∆V ) − ∂E − ∂V
= k ∂π ∂π .
(∆V ∆E) (∆V )2 − ∂E − ∂V

The evaluation of the inverse here is rather tricky; the interested reader
may consult Kubo (1965), problem 6.2, pp. 382–5, where a complete so-
lution, along with the desired results for (∆E)2 , (∆V )2 and (∆E∆V ), is
given.
In passing, we note that two of the aforementioned results are also given
in eqns. (15.1.14 and 18); the third may be obtained as follows: multiply
(15.1.17) by ∆V , take its average and utilize relations (15.1.14), to get
     
∂E ∂P
(∆E∆V ) = (∆V )2 = T − P kT κT V
∂V T ∂T V
     
∂V ∂V
= kT T +P .
∂T P ∂P T
144

15.4. With a given displacement y(x), the overall shape of the string would, on
an average, be as shown in Fig. 1. This amounts to a strain, ∆`, in the
string given by the expression
p p
∆` = x2 + y 2 + (` − x)2 + y 2 − `;

the energy Φ associated with this strain is obviously F ∆`. For small y,
 2
y2

y F`
Φ(y) ≈ F + = y2 ,
2x 2(` − x) 2x(` − x)

which leads to a probability distribution for y that is Gaussian, with vari-


ance
kT
(∆y)2 = x(` − x).
F`

For the second part, we refer to Fig. 2 for which


q p q
∆` = x21 + y12 + (x2 − x1 )2 + (y1 − y2 )2 + (` − x2 )2 + y22 − `
y12 (y1 − y2 )2 y22
≈ + + , and hence
2x1 2(x2 − x1 ) 2(` − x2 )
F
x2 (` − x2 )y12 − 2x1 (` − x2 )y1 y2 + x1 (` − x1 )y22 .
 
Φ(y1 , y2 ) ≈
2x1 (x2 − x1 )(` − x2 )

This leads to a bivariate Gaussian distribution in the variables y1 and y2 ,


with the covariance matrix
!  −1
y12 y1 y2 kT x1 (x2 − x1 )(` − x2 ) x2 (` − x2 ) −x1 (` − x2 )
=
y2 y1 y22 F −x1 (` − x2 ) x1 (` − x1 )
 
kT x1 (` − x1 ) x1 (` − x2 )
= .
F ` x1 (` − x2 ) x2 (` − x2 )

15.5. The quantity in question here is

{(∆NA )2 }1/2 /N̄A = (kT κT /VA )1/2 ; (1)

see eqn. (15.1.20). Assuming the gas to be an ideal one, the compressibility
κT may be taken as 1/(nkT ), where n is the particle density in the system;
see eqn. (15.2.12). This reduces (1) to the simple expression (1/N̄A )1/2 .
145

For this fraction to be 1 per cent, the volume VA of the subsystem must be
such that it contains, on an average, 104 particles. At normal temperature
and pressure, this volume would be about 3.7 × 10−22 m3 — for instance,
a cube of side 7.2 × 10−8 m.
15.6. By eqns. (15.2.23) and (15.3.11), and by Note 5, we have

x2 = 2Dt = 2BkTt = kTt/3πηa.

It follows that
k = 3πηax2 /Tt.
Substituting the given data, we get: k = 1.18 × 10−16 erg K −1 , which may
be compared with the accepted value of 1.38 × 10−16 erg K −1 .
15.7. By eqn. (15.3.2), we have
 
dv 1 1 d 1
hv · Fi = M v · + hv · vi = M hv2 i + hv2 i.
dt B 2 dt B

Substituting for hv2 i from eqn. (15.3.29) and remembering that B = τ /M ,


we get
     
M 3kT 1 3kT 3kT
hv · Fi = − v2 (0) e−2t/τ + − − v2 (0) e−2t/τ
τ M B M M
3kT 3kT
= = ,
BM τ
which holds at all t. By tacit assumption, the statement hr · Fi = 0 also
holds at all t. On the other hand, the quantities hv · F i and hr · F i behave
somewhat differently.
First of all,  
dv 1 d
hv · F i = M v· = M hv2 i
dt 2 dt
If the Brownian particle has already attained thermal equilibrium, then
hv2 i = 3kT /M and hence hv · F i = 0; if it hasn’t, then
 
M 3kT
hv · F i = − v (0) e−2t/τ ,
2
τ M
which decays exponentially with t. Next, by eqns. (15.3.1 and 5),
 
dv M M d 2
hr · F i = M r · = − hr · vi = − hr i.
dt τ 2τ dt
Once again, if the particle has already attained thermal equilibrium, then,
by eqn. (15.3.7),

hr · F i = −3kT (1 − e−t/τ ) −→ −3kT ;


tτ
146

if it hasn’t, then, by eqn. (15.3.31),


hr · F i = [−3kT + {3kT − M v2 (0)}e−t/τ ](1 − e−t/τ )
which too approaches −3kT when t becomes much larger than τ .
15.8. Integrating eqn. (15.3.14) over t, we get
 0 
Zt h it Z t Zt
0 0
r(t) = v(t0 )dt 0 = v(0) −τ e−t /τ + e−t /τ eu/τ A(u)du  dt 0 .
0
0 0 0
(1)
The remaining integration may be carried out by parts, with the result
 0 t
Zt Zt
0 0 0
(−τ e −t /τ
) r A(u)du  − (−τ e−t /τ ){et /τ A(t0 )}dt 0
u/τ

0 0 0
Zt Zt
= −τ e−t/τ eu/τ A(u)du + τ A(t0 )dt 0 . (2)
0 0

Substituting (2) into (1), we obtain the desired result


Zt
−t/τ
r(t) = v(0)τ (1 − e )+τ {1 − e(u−t)/τ }A(u)du. (3)
0
2
To obtain an expression for hr (t)i, we take the square of (3) and average
it over an ensemble. The cross-term vanishes on averaging, and we are
left with
Zt Zt
−t/τ 2
2 2 2
hr (t)i = v (0)τ (1 − e ) +τ 2
{1 − e(u1 −t)/τ }{1 − e(u2 −t)/τ }
0 0
hA(u1 ) · A(u2 )idu 1 du 2 . (4)
Noting that the autocorrelation function hA(u1 ) · A(u2 )i, which is the
same as the function K(s) of Sec. 15.3, may be treated as a delta function,
see the passage from eqn. (15.3.24) to (15.3.25) along with eqns. (15.3.26
and 28), we may write
hA(u1 ) · A(u2 )i = Cδ(u2 − u1 ), where C = 6kT /M τ.
The second term in (4) then takes the form
Zt
6kT τ
{1 − e(u−t)/τ }2 du
M
0
 
6kT τ 1 −t/τ −t/τ
= t − τ (1 − e )(3 − e ) . (5)
M 2
Substituting (5) into (4), we obtain eqn. (15.3.31).
147

15.13. By eqn. (15.5.14), we have in the first case


Z∞
2
w(f ) = 4 K(0)e−αs cos(2πf ∗ s) cos(2πfs)ds
0
Z∞
2
= 2K(0) e−αs [cos{2π(f − f ∗ )s} + cos{2π(f + f ∗ )s}]ds.
0

Using formula (B.41), we get the desired result


 π 1/2 2
(f −f ∗ )2 /α 2
(f +f ∗ )2 /α
w(f ) = K(0) [e−π + e−π ].
α
In the limit α → 0 (with f ∗ > 0), w(f ) → K(0)δ(f − f ∗ ); see eqn. (B.43).
In the limit f ∗ → 0 (with α > 0), w(f ) → 2K(0)(π/α)1/2 exp{−(π 2 f 2 /α)}.
On the other hand, if both α and f ∗ → 0, w(f ) tends to be 2K(0)δ(f ).
In either case, eqn. (15.5.16) is satisfied.
In the second case, we get
 
α α
w(f ) = 2K(0) 2 + 2 .
α + 4π 2 (f − f ∗ )2 α + 4π 2 (f + f ∗ )2

Now, in the limit α → 0 (with f ∗ > 0), w(f ) → 2πK(0)δ{2π(f − f ∗ )} =


K(0)δ(f − f ∗ ); see eqn. (B.36). In the limit f ∗ → 0 (with α > 0), w(f ) →
4K(0)α/(α2 + 4π 2 f 2 ). On the other hand, if both α and f ∗ → 0, w(f )
again tends to be 2K(0)δ(f ).

15.14. By eqn. (15.5.14), we get


Z∞
sin(as) sin(bs)
w(f ) = 4 K(0) cos(2πfs)ds
abs 2
0
Z∞
2K(0) ds
= sin(as)[sin{(b − 2πf )s} + sin{(b + 2πf )s}] . (1)
ab s2
0

To evaluate the integral in (1), we use the formula, see Gradshteyn and
Ryzhik (1965),
Z∞ (
dx pπ/2 if p ≤ q
sin(px ) sin(qx ) 2 = .
x qπ/2 if q ≤ p
0

It follows that if 0 < f ≤ (a − b)/2π, then the integral in (1) is equal to

(b − 2πf )π/2 + (b + 2πf )π/2 = bπ. (2)


148

If (a − b)/2π ≤ f ≤ (a + b)/2π, then our integral is equal to

(b − 2πf )π/2 + aπ/2 = (a + b − 2πf )π/2. (3)

If f ≥ (a + b)/2π, then we have

−aπ/2 + aπ/2 = 0. (4)

Substituting (2)–(4) into (1), we obtain the desired result for w(f ).
It is quite straightforward to check that the function w(f ) obtained here
satisfies eqn. (15.5.16).

15.15. (a) From the defining equation of the variable Y (t), we get
u+t Z
Z u+t

hY 2 (t)i = hy(u1 )y(u2 )idu 1 du 2 (1)


u u

Since y(u) is statistically stationary, we may write


Z∞
hy(u1 )y(u2 )i = w(f ) cos(2πfs)df (s = u2 − u1 ); (2)
0

see eqn. (15.5.15). Substituting (2) into (1), we get


Z∞
2
hY (t)i = w(f )I(f, t)df , where (3)
0
u+t Z
Z u+t

I(f, t) = {cos(2πfu 2 ) cos(2πfu 1 ) + sin(2πfu 2 ) sin(2πfu 1 )}du 1 du2


u u
 u+t 2  u+t 2
Z Z
= cos(2πfu)du  +  sin(2πfu)du 
u u
1 
= [sin{2πf (u + t)} − sin(2πfu)]2 +
4π 2 f 2
[cos(2πfu) − cos{2πf (u + t)}]2


1
= [1 − cos(2πft)], regardess of the initial instant u.
2π 2 f 2
(4)

Substituting (4) into (3), we obtain the desired result for hY 2 (t)i.
149

Next, it follows that


Z∞
∂ 1 w(f )
hY 2 (t)i = sin(2πft)df , and (5)
∂t π f
0
Z∞
∂2
hY 2 (t)i = 2 w(f ) cos(2πft)df . (6)
∂t2
0

Taking the sine transform of (5) and the cosine transform of (6), we
obtain the other quoted results. Finally, a comparison of eqns. (2)
and (6) shows that

1 ∂2
Ky (s) = hY 2 (s)i. (7)
2 ds 2
(b) If the variable y(u) is the x-component of the velocity of a Brownian
particle, with power spectrum (15.5.21), then eqns. (3) and (4) give
Z∞
2 2kT τ 1 − cos(2πft)
hx (t)i = 2 df
π M f 2 {1 + (2πf τ )2 }
0
Z∞
4kT τ 2 1 − cos(xt/τ )
= dx
πM x2 (1 + x2 )
0
2kT τ 2 t
 
= − (1 − e−t/τ ) , (8)
M τ

in complete agreement with eqn. (15.3.7) for the quantity hr2 (t)i. We
also note that
1 ∂2 2 kT −s/τ
2
hx (s)i = e (s > 0), (9)
2 ∂s M
which indeed is equal to the autocorrelation function K(s) of the
variable vx ; see eqn. (15.6.20).
15.16. First we’ll prove the following lemma.
Lemma:
For a given variable x(t), define a complementary function
T /2
Z
1
yx (f, T ) = √ x(t)e−2πift dt. (1)
T
−T /2

The power spectrum of the variable x(t) is then given by

wx (f ) = 2 lim |yx (f, T )|2 . (2)


T →∞
150

Proof:
From (1), it readily follows that
T /2 Z
Z T /2
1
2
|yx (f, T )| = − x(t1 )x(t2 )e2πif (t2 −t1 ) dt 1 dt 2 .
T
−T /2 −T /2

Changing over to the variables S and s, as defined in eqns. (15.3.23), we


get
Z Z    
1 1 1
|yx (f, T )|2 = x S − s x S + s cos(2πfs)dSds.
T 2 2
Integrating over S and letting T → ∞ amounts to taking an ensemble
average of the quantity x S − 21 s x S + 12 s ; this reduces the above ex-


pression to
Z∞
Kx (s) cos(2πfs)ds
−∞

which, by eqn. (15.5.14), is equal to 12 wx (f ). Hence the lemma.


We now proceed to establish the stated relation between the power spectra
wv (f ) and wA (f ). For this we refer to eqn. (15.3.5) for the variable A(t)
and construct its complementary function
T /2 
Z 
1 dv v
yA (f, T ) = √ + e−2πift dt. (3)
T dt τ
−T /2

The first part here gives


 
T /2
Z
1  T /2
√ ve−2πift − v(−2πif )e−2πift dt  .

T −T /2
−T /2

Equation (3) then becomes


       
1 T −πift T πifT 1
yA (f, T ) = √ v e −v − e + + 2πif yv (f, T ).
T 2 2 τ
Since the variable v(t) is bounded, the limit T → ∞ gives
2
1
|yA (f, T )|2 ≈ + 2πif |yv (f, T )|2 .

(4)
τ
Using lemma (2), we finally get
 
1
wA (f ) = 2 + (2πf )2 wv (f ), (5)
τ
151

which is the desired result.


Now, by eqn. (15.5.21),

12kT τ 1
wv (f ) = . (6)
M 1 + (2πf τ )2

Substituting (6) into (5), we readily obtain the stated result for wA (f ).
Note that this result is consistent with the assertion that, for most practi-
cal purposes, the autocorrelation function KA (s) may be taken as Cδ(s),
with C = 6kT /M τ ; see eqns. (15.3.26 and 28).
15.17. (a) Using eqn. (15.3.14), we construct the quantity v(t) · v(t + s) and
average it over an ensemble. The cross-term vanishes on averaging,
and we are left with

 
Zt Zt+s
hv(t)·v(t+s)i = e−2(t+s)/τ v2 (0) + e(u1 +u2 )τ hA(u1 ) · A(u2 )idu 1 du 2  .
0 0
(1)
In view of the argument leading from eqn. (15.3.24) to (15.3.25), we
may replace the function hA(u1 ) · A(u2 )i by the singular expression
Cδ(u2 − u1 ), where C = 6kT /M τ . At the same time, we observe
that the integral

Zt (
(u1 +u2 )/τ e2u2 /τ if 0 < u2 < t
e δ(u1 − u2 )du 1 =
0 otherwise.
0

The double integral in (1) is then equal to

Zt
τ
Ce 2u2 /τ du 2 = C (e2t/τ − 1) if s > 0, and
2
0
Zt+s
τ
Ce 2u2 /τ du 2 = C (e2(t+s)/τ − 1) if s < 0.
2
0

Substituting these results into (1), we obtain eqns. (15.6.7) and (15.6.8).
Equation (15.6.9) follows straightforwardly.
(b) To evaluate hr2 (t)i, we write eqn. (15.6.9) in the form
 
3kT 3kT −|s|/τ
Kv (s) = v2 (0) − e−2S/τ + e , (2)
M M

1
where S = 2 (u1 + u2 ) and s = (u2 − u1 ). Substituting (2) into
152

eqn. (15.6.6), we get


 
  Zt/2 Zt
3kT  −2S/τ
hr2 (t)i = v2 (0) −  e · 4SdS + e−2S/τ · 4(t − S)dS 

M
0 t/2
 
Zt/2 Z2S Zt 2(t−S)
Z
3kT  −2s/τ
+  2 e dsdS + 2 e−2s/τ dsdS 

M
0 0 t/2 0
 
3kT 3kT
= v2 (0) − · τ 2 (1 − e−t/τ )2 + · 2τ {t − τ (1 − e−t/r )},
M M

which is the same as expression (15.3.31). Note that the second part
of this result is identical with expression (15.3.7) that pertains to a
stationary ensemble.
15.18. Using equation (15.3.37), the response function is
Z ∞
χ̃vx (ω) = χvx (s)eiωs ds
0

which has imaginary part



1 ω ω 2 − ω02
χ00vx (ω) = .
M (ω 2 − ω02 )2 + γ 2 ω 2

The correlation function is


Z t Z t0
Gvx (t − t0 ) = ds ds0 χvx (t − s)χxx (t0 − s0 ) hF (s)F (s0 )i
−∞ −∞

Using hF (s)F (s0 )i = 2γM kT δ(s − s0 ) and Fourier transforming gives



2kT ω 2 − ω02
Svx (ω) = .
M (ω 2 − ω02 )2 + γ 2 ω 2

15.19. Just differentiate equation (15.6.29) with respect to t and equation (15.6.28)
drops out.

Correction to the first printing of third edition: 15.20 and 15.21: The
correlation function relation should read:

GAB (t) = GBA (−t − iβ~).


153

15.20. Since we need to evaluate h[A(t), B(0)]i, we need to relate hB(0)A(t)i to


hA(t)B(0)i.
1  
hB(0)A(t)i = Tr BeiHt/~ Ae−iHt/~ e−βH
Q
1  
= Tr eiHt/~+βH Ae−iHt/~−βH Be−βH
Q
= hA(t − iβ~)B(0)i
Now equation (15.6.34) can be evaluated as
Z
1
χ̂00AB (ω) = (hA(t)B(0)i − hA(t − iβ~)B(0)i) eiωt dt
2~
Z 
1 
= hA(t)B(0)ieiωt − hA(t − iβ~)B(0)ieiω(t−iβ~) e−β~ω dt
2~
1
1 − e−β~ω SAB (ω)

=
2~
15.21. Since hB(0)A(t)i = hA(t − iβ~)B(0)i,
 
dA
hA(t)B(0) − B(0)A(t)i ≈ iβ~ B
dt
as ~ → 0. Therefore,
Z Z  
1 iβ~ dA
χ00AB (ω) = hA(t)B(0) − B(0)A(t)ieiωt dt ≈ B(0) eiωt dt
2~ 2~ dt
Z
βω ω
= hA(t)B(0)i eiωt dt = SAB (ω)
2 2kT
15.22. The self–diffusion term can be written
Z
Sself (ω) = he−ik·(r(t)−r(0)) ieiωt dt
Z
2
= eiωt−Dk |t| dt,

using the diffusion relation h(x(t) − x(0))2 i = 2D|t|. Integrating gives


2Dk 2
Sself (ω) = ;
ω 2 + (Dk 2 )2
compare to the heat diffusion term in equation (15.6.45).

15.23. The magnitude of the wavevector transfer is k = 2k0 = 7 × 106 m−1 and
the width of the Rayleigh peak is ∆ω = DT k 2 = 7 × 106 s−1 . The location
of the sound peak is at ω = ck = 2.4 × 109 m−1 is well separated from the
Rayleigh peak.
15.24. The Raman peak has ~ω = 0.05 eV ' 2kT at room temperature so the
peaks are not symmetric. Since Γ ∼ 1012 s−1 and ω ∼ 8 × 1013 s−1 , the
Raman peak is well resolved.
Chapter 16

16.1. Here is a C code snippet for a pseudorandom number generator based


on the L’Ecuyer prime number linear congruential generator discussed in
Appendix I.

double rand(double seed[])


{
seed[0] = fmod(seed[0] * 40014., 2147483563.);
seed[1] = fmod(seed[1] * 40692., 2147483399.);
double r=seed[0]- seed[1];
if (r<= 0.0) r += 2147483562.;
return r/2147483563.;


For a sequence of N numbers, one should test that hxi ≈ 0.5 ± 1/(12 N )
and hx2 i − hxi2 ≈ 1/12.
16.2 Here is a code snippet for generating gaussian pseudorandom numbers
based on the Box-Muller algorithm in Appendix I.

double s,w;
do
{
double x=2.0*rand(seed)-1.0;
double y=2.0*rand(seed)-1.0;
s=x*x+y*y;
} while( s >= 1.0);
w=sqrt(-2.0*log(s)/s);
gaussrand = x*w;

For efficiency, one can also use y*w as an independent gaussian pseudo-
random number. √ For a sequence of N numbers, one should test that
hxi ≈ 0.0 ± 1/ N and hx2 i ≈ 1.0. The reader should also determine the
expected uncertainty in the value of the variance for N numbers. The his-
togram of points for pairs of gaussian random numbers should be centered
at 0, be isotropic, and have variance hx2 + y 2 i = 2.

154
155

16.3 First note that the sum of two gaussian random distrubutions is also
gaussian,
s2 s2
   
exp − 2σ12 exp − 2σ22
1 2
P1 (s1 ) = p , P2 (s2 ) = p ,
2πσ12 2πσ22
Z Z
P (S = s1 + s2 ) = δ(S − s1 − s2 )P1 (s1 )P2 (s2 )ds1 ds2 ,
 2

exp − 2(σ2S+σ2 )
1 2
P (S) = p .
2π(σ12 + σ22 )
Iterating the equation defining the correlated random numbers gives

X
sk = (1 − α) αj rk−j .
j=0

This implies that the s’s are also gaussian. The averages hsk i are clearly
zero and the variance is given by
∞ X
X ∞
hs2k i = (1 − α)2 αi+j hrk−i rk−j i,
i=0 j=0

which is easily evaluated to give



X (1 − α)2 1−α
hs2k i = (1 − α)2 α2j = = .
j=0
1 − α2 1+α

The correlations are then given by


1−α
hsk sk−l i = α|l| .
1+α

16.4 A Monte Carlo Sweep of an ordered list of N particles


x0 < x2 < x3 < · · · < xN −1 is done with the following C code snippet.

xtrial = x[0] + dx*(rand(seed)-0.5);


if (xtrial - (x[n-1] - L) > 1.0 && x[1] - xtrial > 1.0) x[i]=xtrial;
for (int i=1;i<n-1;i++)
{
xtrial=x[i] + dx*(rand(seed)-0.5);
if (xtrial - x[i-1] > 1.0 && x[i+1]-xtrial > 1.0) x[i]=xtrial;
}
xtrial = x[n-1] + dx*(rand(seed)-0.5);
if (xtrial - x[n-2] > 1.0 && (x[0] + L) - xtrial > 1.0) x[i]=xtrial;

Note the periodic boundary conditions treat particle to the left of particle
particle 0 as particle N − 1 shifted left by L, and particle to the right of
156

particle N − 1 as particle 0 shifted to the right by L. The random step


size dx is typically chosen on the order of L/(DN ) − 1 but must be less
than 2 to avoid particles getting out of order.
16.5 Here is a C code snippet for a Monte Carlo of a two-dimensional system
of hard spheres in a LX × LY periodic box.
int i = n*rand(seed); // choose a particle randomly
double xtrial = x[i] + dx*(rand(seed) -0.5);
double ytrial = y[i] + dx*(rand(seed) -0.5);
int collision = 0;
for (int j=0;j<n;j++)
{
if( j != i)
{
double dx = fabs(x[j]-x[i]);
double dy = fabs(y[j]-y[i]);
if (dx > halfLX) dx = LX-dx; // use periodic boundary conditions
if (dy > halfLY) dy = LY-dy; // halfLX=0.5*LX and halfLY=0.5*LY
if (dx*dx+dy*dy < 1.0) collision=1; // test for collision
}
if (collision == 1) break;
}
if (collision == 0) //accept trial position if no collision
{
x[i] = xtrial;
y[i] = ytrial;
if (x[i] > LX) x[i] -= LX; // impose periodic boundary conditions
if (y[i] > LY) y[i] -= LY;
if (x[i] < 0.0) x[i] += LX;
if (y[i] < 0.0) y[i] += LY;
}

Here is a C code snippet to collect correlation function information.


for (int i=0;i<n;i++) for (int j=i+1;j<n;j++)
{
double dx = fabs(x[j]-x[i]);
double dy = fabs(y[j]-y[i]);
if (dx > halfLX) dx = LX-dx; // use periodic boundary conditions
if (dy > halfLY) dy = LY-dy; // halfLX=0.5*LX and halfLY=0.5*LY
int ir=sqrt(dx*dx+dy*dy)/dr; // dr is the binsize
histogram[ir] ++; //increment the histogram
}

16.6 The new additions to the code accept moves ∆y > 0 with probability
exp(−βmg∆y), and to reject moves that go outside the vertical bound-
aries. At low density and small βmgLy , the density will be proportional to
exp(−βmgy). Large βmgLy will result in an interface with a low density
phase above a high density phase.
157

16.7 Here is a C code snippet for one time step for Lennard-Jones particles in
a two dimensional LX×LY box with periodic boundary conditions. The
arrays x1 and x0 store the current and previous positions of the n particles
respectively.

// calculate forces
for (int i=0;i<n;i++)
{
fx[i]=0.0;
fy[i]=0.0;
}
for (int i=0;i<n;i++) for (int j=i+1;j<n;j++)
{
double dx=x1[j]-x1[i];
double dy=y1[j]-y1[i];
if (dx > halfLX) dx -= LX; // use periodic boundary conditions
if (dy > halfLY) dy -= LY; // halfLX=0.5*LX and halfLY=0.5*LY
if (dx < -halfLX) dx += LX;
if (dy < -halfLY) dy += LY;
double r2=1.0/(dx*dx+dy*dy);
double r4=r2*r2;
double r6=r2*r4;
double r8=r4*r4;
double r14=r8*r6;
double f0=48.0*r14 - 24.0*r8; // see equation (16.3.5)
double fx0=f0*dx;
double fy0=f0*dy;
fx[j] += fx0; // Use Newtons’s third law to update forces on each particle
fy[j] += fy0;
fx[i] -= fx0;
fy[i] -= fy0;
}
//update positions using Verlet
for (int i=0;i<n;i++)
{
double xnew=2*x1[i]-x0[i]+dtsqr*fx[i];
double ynew=2*y1[i]-y0[i]+dtsqr*fy[i];
x0[i]=x1[i];
x1[i]=xnew;
y0[i]=y1[i];
y1[i]=ynew;
}
// impose periodic boundary conditions
for (int i=0;i<n;i++)
{
if (x1[i] > LX) { x1[i] -= LX; x0[i] -= LX; }
158

if (y1[i] > LY) { y1[i] -= LY; y0[i] -= LY; }


if (x1[i] < 0.0) { x1[i] += LX; x0[i] += LX; }
if (y1[i] < 0.0) { y1[i] += LY; y0[i] += LY; }
}

16.8 The new additions are to generate a one–body force Fy = −mg and re-
pulsive forces with the top and bottom walls. The average kinetic energy
per particle will be independent of the position in the box. At low density
and small βmgLy , the density will be proportional to exp(−βmgy). Large
βmgLy will result in an interface with a low density phase above a high
density phase.

16.9 Each Monte Carlo step involves determining the energy change of a spin
flip with is proportional to ∆ = si (si+1 + si−1 ) using periodic boundary
conditions. If ∆ ≤ 0 flip the spin. Otherwise ∆ = +2, so flip the spin
with probability exp(−4K). Due to the periodic boundary conditions
the correlation function will also be periodic. You can generalize the
calculation of the correlation function in section 13.2 for a finite periodic
lattice to show that the zero field correlation function is of form
"   N −|i−j| #
|i−j|
1 λ2 λ2
hsi sj i =  N + ,
λ1 λ1
1 + λλ12

so the correlations are minimized halfway across the lattice.


159

16.10 Here is a C code snippet for the two dimensional Ising model

int L=32; // size of lattice


int n=L*L; // number of sites
double K=-0.5*log(sqrt(2.0)-1); // K=critical value
int nstat = 1000000; // number of Monte Carlo Sweeps
int neq=100000; // number of equilibration sweeps
int* s=new int[n]; // spins: +1 or -1
for (int i=0;i<n;i++) s[i]=1; // all spins initially up (+1)
int* i1 = new int[n]; // arrays i1[], i2[], i3[], i4[]: neighbor sites
int* i2 = new int[n];
int* i3 = new int[n];
int* i4 = new int[n];
for (int i=0;i<n;i++)
{
i1[i] = i+1; // site to right
i2[i] = i-1; // site to left
i3[i] = i+L; // site above
i4[i] = i-L; // site below
if ((i1[i] % L) == 0) i1[i] -= L; //implement periodic boundary conditions
if (((i2[i]+L) % L) == L-1) i2[i] += L; //implement periodic boundary conditions
if (i3[i] >= n) i3[i] -= n; //implement periodic boundary conditions
if (i4[i] < 0) i4[i] += n; //implement periodic boundary conditions
}
double* boltz=new double[5]; //precompute spin flip Boltzmann factors for efficiency
boltz[2]=exp(-4.0*K); // energy increase = 4
boltz[4]=exp(-8.0*K); // energy increase = 8
int* e=new int[nstat]; //stored energy after each pass
int* m=new int[nstat]; //stored magnetization after each pass
int energy;
int mag;
for (int iter=0; iter<(nstat+neq);iter++) // perform nstat Monte Carlo Sweeps
//after neq equilibration steps
{
for (int ii=0;ii<n;ii++)
{
int i=n*rand(seed); //choose a random site
int neighborsum=s[i1[i]]+s[i2[i]]+s[i3[i]]+s[i4[i]]; //sum of spins on
//neighboring sites
int de = s[i]*neighborsum; // energy change of spin flip is 2*de
if (de <= 0) s[i]=-s[i]; // accept if energy change is not positive
else if (rand(seed) < boltz[de]) s[i]=-s[i]; //if energy increase,
//accept with Boltzmann factor probability
}
if (iter >= neq)
{
160

mag=0;
energy=0;
for (int i=0;i<n;i++)
{
mag += s[i];
energy -= s[i]*(s[i1[i]]+s[i3[i]]);
}
e[iter-neq] = energy; //store energy for later analysis
m[iter-neq] = mag; //store magnetization for later analysis
// collect other statistics here, especially for correlations
}
}

16.11 Use the code snippet to collect a histogram of energies. Use the code
posted at www.elsevierdirect.com to calculate the energy distribution at
K = 0.4, Kc = 0.4406868, 0.5. Here is a plot. The horizontal axis is the
energy above the ground state in units of 4J and the vertical axis is the
probability for each energy.

0.030

0.025

0.020

0.015

0.010

0.005

0.000
0 100 200 300 400 500

16.12 Each Monte Carlo step will involve involve creating a trial state
(θtrial = θi + ∆θ(rand(seed) − 0.5)) and calculating the change in energy,

∆ε = cos(θi + θi+1 ) − cos(θi − θi−1 )


− cos(θtrial − θi+1 ) − cos(θtrial − θi−1 ).

Accept (i.e. set θi = θtrial ) if ∆ε < 0 or rand(seed) < exp(−β∆ε).


161

16.13 Here is is C code snippet for the two dimensional XY model

int L=32; // size of lattice


int n=L*L; // number of sites
double K=1.12; // K=critical value
int nstat = 1000000; // number of Monte Carlo Sweeps
int neq=100000; // number of equilibration sweeps
double* theta=new double[n]; // angles of spins
double dtheta=1.0; // range for random angle changes
double twopi=8.0*atan(1.0);
for (int i=0;i<n;i++) theta[i]=0.0; // all spins initially along x direction
int* i1 = new int[n]; // arrays i1[], i2[], i3[], i4[]: neighbor sites
int* i2 = new int[n];
int* i3 = new int[n];
int* i4 = new int[n];
for (int i=0;i<n;i++)
{
i1[i] = i+1; // site to right
i2[i] = i-1; // site to left
i3[i] = i+L; // site above
i4[i] = i-L; // site below
if ((i1[i] % L) == 0) i1[i] -= L; //implement periodic boundary conditions
if (((i2[i]+L) % L) == L-1) i2[i] += L; //implement periodic boundary conditions
if (i3[i] >= n) i3[i] -= n; //implement periodic boundary conditions
if (i4[i] < 0) i4[i] += n; //implement periodic boundary conditions
}
double* e=new double[nstat]; //stored energy after each pass
double* mx=new double[nstat]; //stored x component of magnetization after each pass
for (int iter=0; iter<(nstat+neq);iter++) // perform nstat Monte Carlo Sweeps
//after neq equilibration steps
{
for (int ii=0;ii<n;ii++)
{
int i=n*rand(seed); //choose a random site
double thetatrial = fmod(theta[i] + dtheta*(rand(seed)-0.5) + twopi , twopi);
double de=cos(theta[i]-theta[i1[i]]) - cos(thetatrial-theta[i1[i]])
+ cos(theta[i]-theta[i2[i]]) - cos(thetatrial-theta[i2[i]])
+ cos(theta[i]-theta[i3[i]]) - cos(thetatrial-theta[i3[i]])
+ cos(theta[i]-theta[i4[i]]) - cos(thetatrial-theta[i4[i]]);
if (de<0.0) theta[i]=thetatrial; //accept if energy decreases
else if (rand(seed) < exp(K*de) theta[i]=thetatrial; //or with Boltzmann factor
}
if (iter >= neq)
{
double magx=0.0;
double magy=0.0;
162

double energy=0.0;
for (int i=0;i¡n;i++)
{
magx += cos(theta[i]);
magy += sin(theta[i]);
energy -= (cos(theta[i]-theta[i1[i]])+cos(theta[i]-theta[i3[i]]));
}
e[iter-neq] = energy; //store energy for later analysis
mx[iter-neq] = magx; //store x magnetization for later analysis
my[iter-neq] = magy; //store y magnetization for later analysis
// collect other statistics here, especially for correlations
}
}

You might also like