Statistical Mechanics
Statistical Mechanics
Theory Of Ensembles
M
X
Hens = Hk + ET h
k=1
where ET h is the thermal interaction energy. We will assume that
ET h << Hk . This is what we mean by a large system. Thus, the total
1
2 1. THEORY OF ENSEMBLES
wavefunction is a product
Y
Ψ= ψk
k
. What, is the probabilty that the systems in the ensemble are dis-
tributed among their various states accoring to the set {Mk } ? By
our fundamental hypothesis, the a-priori probability we are looking
for is ∝ Ω({Mk }) . Which, then, is the most probable distribution ?
This is the distribution for which we have maximum Ω({Mk }) , with
P P
k Mk = M and k Mk Ek = E.
Using lagrange multipliers, the condition we are looking for becomes
∂ ln Ω ∂ X ∂ X
−α Mk − β Mk Ek = 0
∂Mj ∂Mj ∂Mj
where α and β are lagrange multipliers. We will make use of Stirling’s
approximation for the factorial
N
√
N 1
N ! −→N large = 2πN 1 + + ···
e 12N
. In fact for our purposes it will suffice to consider the simpler expres-
N
sion N ! −→N large = Ne since M as well as Mj are large. (In principle
they can be as large as we please, we just have to consider a big enough
ensemble!). Thus we have
X X
ln Ω = M ln M − M − Mk ln Mk + Mk
− ln M̃j − α − βEj = 0
and thus
M̃j = e−(α+βEj )
for the most probable distribution. Here, α and β are to be obtained
from the constraints. For this distribution the average energy per sys-
tem E can be obtained from
X
E= M̃j Ej = M E
and hence X
E= Pj Ej
j
4 1. THEORY OF ENSEMBLES
where
M̃j 1
Pj = = e−βEj
M Q
is the probabilty that a paricular system is in state j. Here we have
defined the normalisation constant Q as
X
Q= e−βEj
j
We will now show that for fixed N and E , the most probable distribu-
tion becomes sharper and sharper as M → ∞ and in the limit becomes
“the distribution”, so to speak. To this end, let us define
X X
f = ln Ω − α Mj − β Mj Ej
∂f
. ∂Mj
= 0 is the extremum condition that we solved to ob-
Mj =M̃j
∂2f
tain the most probable distribution M̃j . The second derivative ∂Mj2
=
− M1j < 0 . So we have really found the maximum.
X X
f˜ = ln Ω̃ − α M̃j − β M̃j Ej
CONNECTION WITH THERMODYNAMICS 5
j
which is the gaussian distribution. From this it is obvious that the
mean value of the number of systems in the state j, Mj equals the
most probable number M̃j . Let us calculate the variance in Mj :
v
u M j − Mj 2
u 2
1
t
2 = √ → 0 as M → ∞
Mj Pj M
F = −kT ln Q
which we will soon identify with the thermodynamic Helmholtz free en-
∂ F
ergy, FTh . the energy E is seen to be related to F by E = −kT 2 ∂T kT V,N
.
Here the differentiation with respect to T is carried out while keep-
ing the energy eigenvalues Ej constant, hence the restriction to fixed
V and N . The thermodynamic free energy is related to the internal
energy by
FTh = E − T S
and hence
(dFTh )V,N = −SdT
and so
∂
2 FTh
−kT = FTh + T S = E
∂T kT V,N
F → E0 − kT ln ω0
FTh → E0 − T S0
(1.2.1) F = FTh
CONNECTION WITH THERMODYNAMICS 7
F = FTh + kT f (V )
dEj = −pj dV
F −FTh
Thus kT
= C is a constant independent of V and T .
F = FTh + CkT
F = −kT ln Tr e−βH
= E 2 − hEi2
APPLICATION TO BLACKBODY RADIATION 9
Thus
∂E ∂E
(∆E)2 = − = kT 2
∂β ∂T
This calculation has been done at fixed Ej , hence we have held N and
V fixed. Thus
2 2 ∂E
(∆E) = kT = kT 2 CV
∂T N,V
where CV is the heat capacity at constant volume. The relative fluctu-
ation is
∆E 1p 2
= kT CV
E E
Now both E and CV are extensive quantities, E ∝ N and CV ∝ N and
this leads to
∆E 1
∝√
E N
Thus, the larger the system, the smaller the relative fluctuation is going
to be. Using the same technique, we may study fluctuations in other
quantities which depend on the energy.
~k = 2π~ (m1 , m2 , m3 )
L
with m1 , m2 , m3 integers. Energy eigenvalues are then E = ~k . To
specify the state of the photon completely, we also need to know the
helicity of the photon, which can take the two values of ±1.
We first find out the energy eigenvalues for the whole system. For
this, we neglect photon-photon interactions, i.e. we assume that the
photons move around independently.
X
E= n~k,λ ωk ~
~k,λ
j {n~k,λ }
X Yh i
= e−βn~k,λ ωk ~
{n~k,λ } ~k,λ
Y X
= e−β~ωk n~k,λ
~k,λ {n~k,λ }
Y 1
=
1 − e−βωk ~
~k,λ
Hence, X
ln 1 − e−βωk ~
ln Q = −
~k,λ
where we have used the fact that the number of states between mo-
L
mentum component ki and ki + dki is given by ∆mi = 2π~ ∆ki . Thus
the energy is given by
Z
2V ωk ~
E= 3
d3 k β~ω
(2π~) e k −1
- this is the Planck formula. Since probabilities of independent events
are multiplicative, the quantity e−nβωk is proprtional to the probability
of having n photons with energy ω and a given helicity. Thus the mean
number of photons with given momentum and helicity is given by
P∞
D E ne−nβ~ωk
n~k,λ = Pn=0 ∞ −nβ~ωk
n=0 e
"∞ #
∂ X
= − ln e−nβ~ωk
∂(β~ωk ) n=0
1
=
eβ~ωk − 1
Othe thermodynamic functions can also be calculated. For example,
the free energy is
Z
2V
d3 k ln 1 − e−β~ωk
F = −kT ln Q = kT 3
(2π~)
and we can obtain the entropy from S = − ∂F
∂T V
and the pressure from
∂F
p = − ∂V T
.
and
X
N Mj(N ) = the total number of particles in the ensemble
En
= Mη
and hence
Mj(N ) = e−α−βEj(N ) −γN
We now define the grand partition function by
XX
QG = e−βEj(N ) −γN
N j(N )
M = e−α QG
and thus the relative frequencies for the most probable distribution are
Mj(N ) 1 −βEj(N ) −γN
= e
M QG
In the limit M → ∞ , the most probable distribution becomes the
distribution. Therefore the probability that a system is in the state
|j(N )i is
Mj(N ) 1 −βEj(N ) −γN
Pj(N ) = lim = e
M →∞ M QG
the values of β and γ are to be determined from the other two con-
straints.
We can now attach one more box to the whole ensemble. The new
entrant can exchange energy only with the members of the initial en-
semble. Thus what we have here is an example of a hybrid ensemble.
We can now deduce that the Lagrange multiplier β, which appears
in the minimization condition via the energy conservation constraint,
must be related to the temperature. Indeed, since the added box par-
ticipates in the energy exchange it will have equal role in determining
β (but not γ, which is fixed by particle number conservation) as the
members of the initial ensemble. So β must be a function of a com-
mon property shared by the new box, hence, of temperature. Thus, we
define the absolute temperature T by
1
β=
kT
The other Lagrange multiplier, γ, is shared by all boxes which can
exchange both energy and particles. It should be a function of the
intensive thermodynamic variables common to these systems, namely
the temperature, T , and the chemical potential, µ
γ = γ(T, µ)
Legendre transform of QG
and
∂ 1 X X
(1.3.2) − ln QG = N e−βEj(N ) −γN = N Pj(N ) = η
∂γ V,β QG En En
ENTROPY 15
Entropy
The summand e−β(Ej(η) −µη) is the relative probability that the system
−βEj(η)
P
has η particles and energy Ej(η) . Since Q(η) = j(η) e is the
EXAMPLE : A SYSTEM OF FREE PARTICLES 17
where nα is the number of particles in the state |αi, which has the
energy εα . For bosons, nα can take all nonnegative integer values,
0, 1, 2, 3, . . . , while for fermions it can only take the values 0 and 1.
The label α may be collection of quantum numbers, e.g. momentum,
helicity etc. The energy is given by the relativistic formula
p
εα = p~α2 c2 + m2 c4 − mc2
where the upper sign stands for fermions and the lower sign stands for
bosons.
18 1. THEORY OF ENSEMBLES
dE = T dS − pdV + µdη
and
G = µη = E − T S + pV
THE THERMODYNAMIC CONNECTION 19
are both valid, then µ, E/η, S/η are all functions of T and p only (i.e.
they are independent of η)
Proof. We have
dG = dE − T dS − SdT + pdV + V dp
= −SdT + V dp + µdη
ηdµ = −SdT + V dp
and thus
S V
dµ = − dT + dp
η η
Now, since dµ is a perfect differential,
µ must be a function of T and p
∂µ ∂µ
only. Also, Sη = − ∂T p and Vη = ∂p
must be functions of T and p
T
only. Again,
E S V
=µ+T −p
η η η
is also a function of T and p only.
pV = T S − E + ηµ
Since ln QG = pThkT
V
the system obeys the laws of thermodynamics
if and only if ln QG = V f (T, µ) as V → ∞, i.e. if and only if the limit
limV →∞ V1 ln QG exists. The large V behaviour of ln QG depends upon
the nature of the law of force among the particles. We would later
obtain rigorous constraints on the force law, for the thermodynamic
limit to exist.
The thermodynamic limit exists for a system of particles interacting
via short range forces. However, if long range forces (e.g. Coulomb or
gravitational) are present, this limit, in general, does not exist. For
example, in a confined plasma, we can not use thermodynamics, but
statistical mechanics still applies. In other words, in a confined plasma
A SIMPLE EXAMPLE : FREE PARTICLES IN A PERIODIC BOX 21
z = eµ/kT
± ln (1 ± xe−εα/kT )
P
So for our free particle system, for which ln QG = α
the average number of particles is
X (±)2 ze−εα /kT
η =
α
1 ± ze−εα /kT
X 1
= (ε −µ)/kT
α
e α ±1
But η = hnα i where hnα i is the average number of particles in the state
|αi. So, we again find
1
hnα i =
e(εα −µ)/kT ±1
CHAPTER 2
Let us now consider a system of free fermions with the energy mo-
mentum dispersion relation
p
εα = p~α2 c2 + m2 c4 − mc2
Note that 1 and e(µ−εα )/kT represent the relative probabilities that the
single particle state |αi is unoccupied and occupied, respectively. Using
pV P
kT
= ln QG and the sum to integral conversion formula α −→ (2j +
V
R 3
1) (2π)3 d k in the thermodynamic limit, we get
Z
p 2j + 1 3 (µ−ε)/kT
= d k ln 1 + e
kT 8π 3
Since the average number of particles for a energy ε and given helicity,
nε equals the average number of particles with definite momentum ~k
and helicity λ, we have
D E 1
nε = n~k,λ = (ε−µ)/kT
e +1
The total number of particles in the system is
Z
2j + 1
N= V d3 knε
8π 3
and the total energy is
Z
2j + 1
E= V d3 kεnε
8π 3
where ρ = NV
is the particle number density.
For the nonrelativistic case one can easily calculate the energy of
the system to be
3 ~2 kF2 3
E= N = N εF
5 2m 5
and using the pressure - energy density relation derived earlier, we get
2E 2
p= = ρεF
3V 5
For the ultrarelativistic case, the corresponding results are
3
E = N εF
4
and
26 2. SYSTEM OF FREE FERMIONS
1
p = ρεF
4
Thus the free fermion gas exerts a nonzero pressure, even at absolute
zero! This purely quatum mechanical feature is called the electron
degeneracy pressure.
GM 2
In equlibrium, the gravitational force ∼ R2
must balance the hy-
drostatic force ∼ pR2 . Thus,
GM 2
∼ pR2
R2
which leads to
GM 2
∼ pV ∼ E ∼ N εF
R
We could have reached this conclusion via the equipartition theorem,
too.
2
For a nonrelativistic electron gas, εF ∼ ρ 3 . Since N ∼ M we get
5
GM 2 M3
∼ 2
R R
which leads to the mass radius relation
13
1
R∼
M
Thus, in this regime, the radius decreases with increasing mass. This
continues till the elctrons are squeezed to such a high density that
relativistic effects become important.
For very large mass, we can treat the electron gas as ultrarelativis-
1
tic. In this regime εF ∼ ρ 3 . This leads to
4
GM 2 M3
∼
R R
A more careful analysis shows that this actually gives a critical mass
Mc , at which the radius of the white dwarf star vanishes. Beyond this
mass, electron degeneracy pressure can no longer sustain the star and
it would collapse. This is the Chandrashekhar limit.
Reinstating ~ and c, we get
GM 2
≈ N kF ~c
R
1 M
where kF ≈ ρ 3 and N ≈ 2m n
where mn is the nucleon mass. Here,
we have assumed an average nucleus with 2Z ∼ A. For a neutron
star, N ≈ mMn . This leads to an estimate of the critical mass, ignoring
dimensionless factors of the order of unity
4
GMc2
~c Mc 3
≈
R R mn
28 2. SYSTEM OF FREE FERMIONS
ε−µ
where x = kT
. A change of variables leads to
(µ + kT x)l+1
Z ∞
1
Il = dx
−µ/kT l+1 (ex + 1) (e−x + 1)
(µ + kT x)l+1
Z ∞
1
≈ dx
−∞ l+1 (e + 1) (e−x + 1)
x
µ
where the last step is valid since the integrand is nearly zero for x ≤ kT .
We use the binomial theorem
1
(µ + kT x)l+1 = µl+1 + (l + 1)kT µl x + (l + 1)l(kT )2 µl−1 x2 + . . .
2!
kT
where we have to retain the quadratic term in the small parameter µ
since the linear turm vanishes on integration. This gives
Z ∞
µl+1 1 2 l−1 x2
Il = + l(kT ) µ dx x
l+1 2 −∞ (e + 1) (e−x + 1)
30 2. SYSTEM OF FREE FERMIONS
E± = ε ∓ µB H
where the upper (lower) sign refers to electrons with their magnetic
p2
moment parallel (antiparallel) to the magnetic field. Here ε = 2m is
e~
the kinetic energy and µB = 2mc is the Bohr magneton.
The number of electrons with their spin parallel (antiparallel) to
the magnetic field is given by
i−1 Z ∞
X h E± −µ ρ(ε)
n± = e kT +1 → dε E± −µ
ε 0 e kT + 1
where 32
√ 3√
V 2m 2πV
ρ(ε) = 2 ε= (2m) 2 ε
4π ~ h3
32 2. SYSTEM OF FREE FERMIONS
is the density of state for electrons of each kind. To fix µ, note that
the total number of electrons is
Z ∞ " #
1 1
N = n+ + n− = dερ(ε) E+ −µ + E− −µ
0 e kT + 1 e kT + 1
Z ∞
1 1
= dερ(ε) ε−µB H−µ + ε+µB H−µ
0 e kT + 1 e kT +1
This equation has to be solved for µ(H). Changing the sign of H in this
expression, H → −H shows that µ(H) and µ(−H) satisfy the same
equation. Thus µ is an even function of the magnetic field
µ(H) = µ0 + O H 2
M = µB (n+ − n− )
Z ∞
1 1
= µB dερ(ε) ε−µB H−µ − ε+µB H−µ
0 e kT +1 e kT +1
Our aim is to compute the magnetization to O(H). Now
Z ∞
ρ(ε)
n+ = dε ε−µB H−µ
0 e kT +1
Z µB H Z ∞
ρ(ε) ρ(ε)
= dε ε−µB H−µ + dε ε−µB H−µ
0 e kT +1 µB H e kT +1
Consider the first integral. It obviously goes to zero as H → 0. To
obtain the leading term, we can replace H and ε in the denominator
of the integrand by zero, which leads to
Z µB H Z µB H
ρ(ε) 1
dε ε−µB H−µ −→ − µ dερ(ε)
0 e kT +1 e kT + 1 0
Z µB H
√ 3
∝ dε ε ∝ (µB H) 2
0
2.5. PAULI PARAMAGNETISM 33
1The
R∞ 2n−1 R∞ −1 P∞
required integral, 0 dx xex +1 = 0 dxx2n−1 e−x (1 + e−x ) = p=1 (−)p+1
R∞ R∞ P∞ p+1 P∞ p+1
× 0 dxx2n−1 e−px = 0 dxx2n−1 e−x p=1 (−1) p2n = (2n − 1)! p=1 (−1)p2n
P∞ p+1 P∞ 1
Now, 1 (−1) 1 1
P
p2n = 1 p2n − 2 p=even p2n = 1 − 22n−1 ζ(2n) where the Rie-
P∞
mann zeta function ζ(m) = p=1 p1m . The zeta function for even integer argument
is related to the Bernouli numbers Bn through
(2π)2s
ζ(2s) = (−1)s+1 B2s
2(2s)!
The Bernouli numbers are defined via their generating function
∞
x X Bn n
= x
ex − 1 n=0 n!
and have the values B1 = − 21 , B2n+1 = 0 for all n = 1, 2, 3, . . . , B2 = 1
6 , B4 =
1 1 1 5 631
− 30 , B6 = 42 , B8 = − 30 , B10 = 66 , B12 = − 2730 etc.
36 2. SYSTEM OF FREE FERMIONS
Solution :
We start from Z
2j + 1
N= V d3 knε
8π 3
which for a nonrelativistic system becomes
Z ∞
2j + 1 3 √
N= 2
V (2m) 2 dε εnε
4π 0
(we have already done this part of the problem - but this time we are
going to carry out the calculation to the fourth order in the parameter
kT
µ
). Using the general result for Fermi integrals in the last section, we
find
Z ∞ " 2 4 #
√ 2 3 3 π 2 kT 9 7π 4 kT
dε ε = µ2 1 + + + ...
0 3 4 6 µ 16 360 µ
" 2 4 #
2 3 π 2 kT 7π 4 kT
= µ2 1 + + + ...
3 8 µ 640 µ
We get
" 2 4 #
2j + 1 4π 3 3 π 2 kT 9 7π 4 kT
N= V (2mµ) 2 1 + + + ...
8π 3 3 4 6 µ 16 360 µ
Using the definition of the fermi energy
2j + 1 4π 3
N= 3
V (2mεF ) 2
8π 3
2.6. EVALUATION OF FERMI-DIRAC INTEGRALS 37
we get " #
2 4
π2 7π 4
3 3 kT kT
εF = µ
2 2 1+ + + ...
8 µ 640 µ
kT
So, to the fourth order in the parameter µ
, we have
" 2 4 # 23
2 4
kT
π 7π kT
εF = µ 1 + + + ...
8
µ 640 µ
" 2 4 #
2 π 2 kT 2 7π 4 1 π 4 kT
= µ 1+ + − + ...
3 8 µ 3 640 9 64 µ
" 2 4 #
π 2 kT π 4 kT
= µ 1+ + + ...
12 µ 180 µ
and we invert this result to get
" 2 4 #−1
2 4
π kT π kT
µ = εF 1 + + + ...
12 µ 180 µ
" 2 4 4 #
π 2 kT π π4 kT
= εF 1 − + − + ...
12 µ 144 180 µ
" 2 4 #
π 2 kT π 4 kT
= εF 1 − + + ...
12 µ 720 µ
2
2 " 2
2 #−2 4
4
π kT π kT π kT
= εF 1 − 1− + + . . .
12 εF 12 εF 720 εF
" 2 4 4 #
π 2 kT π π4 kT
= εF 1 − + − + ...
12 εF 720 72 εF
" 2 4 #
π 2 kT π 4 kT
= εF 1 − − + ...
12 εF 80 εF
and thus
" 2 4 #
π2 π4
kT kT
G = N µ = N εF 1− − + ...
12 εF 80 εF
The total energy is
Z Z ∞
2j + 1 3 2j + 1 3 3
E= V d kεnε = V (2m) 2 dεε 2 nε
8π 3 4π 2
0
38 2. SYSTEM OF FREE FERMIONS
Now,
Z ∞ " 2 4 #
3 2 5 15 π 2 kT 15 7π 4 kT
dεε 2 = µ2 1 + − + ...
0 5 4 6 µ 16 360 µ
" 2 4 #
2 5 5π 2 kT 7π 4 kT
= µ2 1 + − + ...
5 8 µ 328 µ
Hence
R∞ 3
" 2 4 #
dεε 2 4
E 2 3 5π kT 7π kT
= R 0∞ √ = µ 1+ − + ...
N 0
dε ε 5 8 µ 328 µ
" 2 4 #−1
2 4
π kT 7π kT
× 1+ + + ...
8 µ 640 µ
" 2 4 #
3 π 2 kT π 4 kT
= εF 1 − + + ...
5 12 µ 720 µ
" 2 4 #
5π 2 kT 7π 4 kT
× 1+ − + ...
8 µ 328 µ
" 2 4 4 #
π 2 kT 7π π4 kT
× 1− − − + ...
8 µ 640 64 µ
" 2 4 #
3 5π 2 kT 19π 4 kT
= εF 1 + − + ...
5 12 µ 144 µ
2
2 " 2
2 #−2 4
4
3 5π kT π kT 19π kT
= εF 1 + 1− − + . . .
5 12 εF 12 εF 144 εF
" 2 4 #
3 5π 2 kT π 4 kT
= εF 1 + − + ...
5 12 εF 16 εF
and thus
" 2 4 #
3 5π 2 kT π 4 kT
E = N εF 1 + − + ...
5 12 εF 16 εF
2.6. EVALUATION OF FERMI-DIRAC INTEGRALS 39
2We
R∞ 2 R ∞ dz −z Γ( 1 )
start from 0 dqe−λq = 2√1 λ 0 √ e = 2√2λ = 12 πλ and differentiation
p
z
under the integral sign gives
Z ∞ Z ∞ r
2 ∂ 2 1 π
dqq 2 e−λq = − dqe−λq =
0 ∂λ 0 4 λ3
2.7. THE HIGH TEMPERATURE EXPANSION 41
∂ p 1 1 X − Ej(N )
= e kT N z N
∂(ln z) kT V QG
N,j(N )
1
= hN i = ρ
V
So far, we have found the thermodynamic quantities as power series
expansions in the fugacity. What we would like to do is to eliminate
z and write down the expression for, say, pressure as a function of the
density. Since
ρa3 z2
= z − 3 + ...
2j + 1 22
we can solve for z to get
ρa3 ρa3
1
z= + + ...
2j + 1 2 32 2j + 1
42 2. SYSTEM OF FREE FERMIONS
Thus
z2
2j + 1
p = kT z − 5 + . . .
a3 22
3
ρa 1 1
= ρkT 1 + − 5 + ...
2j + 1 2 32 22
3
1 ρa
(2.7.2) = ρkT 1 + 5 + ...
2 2 2j + 1
This expansion is valid for
ρa3 < 1
thus for the region where the interparticle seperation > a1 ∼ thermal de-
Broglie wavelength. Thus at a fixed density we can look upon the above
equation as a high temperature expansion, or, at a fixed temperature
this can be regarded as a low density expansion.
The first term in the expansion for pressure, eq (2.7.2) is ρkT -
which is just the pressure of a classical ideal gas. As can be seen, the
first correction is positive. This is because the pressure in an ideal fermi
gas is always more than that in a classical ideal gas.
The expansion for density, eq (2.7.1) was derived under the condi-
tion z ≤ 1. Let us consider the limiting case z = 1. Then, we have
∞
ρa3 X (−1)l
= 3 ≈ 0.765
2j + 1 l=1 l 2
Finally, let us examine the conditions under which the fermi gas
shows it’s two extremes of behaviour. We have seen that the gas is non-
degenerate at high temperature or low density, while it is degenerate
at low temperature or high density. Thus the controlling factor is the
dimensionless quantity ρa3 . If ρa3 1 the system is nondegenerate,
while for ρa3 1 the system is highly degenerate.
We can understand this condition by a simple physical argument.
When ρa3 1 the wave functions of the particles do not overlap much.
So the symmetry or antisymmetry of wave functions have a very small
role to play in the calculation of the energy. So, the system becomes
almost classical.
2.7. THE HIGH TEMPERATURE EXPANSION 43
Calculations similar to those in the last chapter yield the high tem-
perature expansion for a nonrelativistic gas
∞
p 2j + 1 X z l
=
kT a3 l=1 l 52
and ∞
2j + 1 X z l
ρ=
a3 l=1 l 32
45
46 3. SYSTEM OF FREE BOSONS
Note that the difference in the above expressions from that of the
fermion case lies in the alternating signs in the latter. Eliminating
z from these equations leads to
ρa3 1
p = ρkT 1 − + ...
2j + 1 2 52
In contrast to the fermion case, we find that in a boson gas the first
correction in the pressure over that of the classical ideal gas is negative.
This is because in the absence of degeneracy pressure the bose gas is
easier to compress than the ideal gas is.
For a system of photons the total number of photons is not con-
served1. Thus, in our derivations we do not need the Lagrange mul-
tiplier that maintains the number conservation constraint. Effectively
this means that we can take µ to be zero. We thus get the same answer
as in the canonical ensemble
1
n̄α = εα
e kT − 1
On the other hand, we could derive the same resulta that we have so far
derived, for systems where the total number of particles are conserved,
by using the canonical ensemble instead. This is because the basic
ststistical hypothesis is the same. Using the grand canonical ensemble
just makes life easier!
Thus for N → ∞, the correction is indeed small.Even if we include the effect of all
the low lying states that are macroscopically populated, the net correction due to
these terms is still O N −1/3 .
3.1. BOSE CONDENSATION 49
z is small and we are still far away from the singularity. The system
behaves like an ideal gas. As ρ increases z → 1 and we are approaching
the singularity. At sufficiently high density, we get on the top of the
singularity.
The onset of the singularity can be understood from an estimate
of the critical density. this critical density ρc is the maximum possible
density for which x is still ∼ 0 and z ∼ 1 − δ with δ 0. Thus
∞
3
X 1
ρc a = 3 ≈ 2.612
l=1 l2
and the corresponding critical pressure is given by
∞
p c a3 X 1
= 5 ≈ 1.341
kT l=1 l 2
Both the critical density and the critical pressure are of course functions
of the temperature
3 5
ρc ∝ T 2 p c ∝ T 2
If we increase ρ beyond ρc , p re- Figure
mains constant at pc . This is because 3.1.1. Isotherm
z has already reached it’s maximum for a free bose gas
value of 1 and cannot increase any
further. So the isotherm has the form as shown in the figure (3.1.1).
As ρ increases beyond ρc , more and more particles condense into the
zero momentum ground state. This is known as Bose-Einstein conden-
sation. Since the condensed partickle do not contribute to the energy
density, the pressure remains constant.This is precisely a phase transi-
tion. Since for the phase transition to start, we must have z = 1 which
is possible only when N → ∞4 this occurs only in the thermodynamic
limit. This is a general feature of phase transitions.
The nature of this transition is quite different from the ordinary
gas-liquid phase transition. In this case the isotherm does not have a
branch corresponding to the liquid state. Only analogues of the gas
phase and the two phase region exist.
Thus ∂ ∂p ∂ρ
ln z T −
is finite but ∂ ln z T −
diverges. Hence
∂p
∂p 2 ∂ ln z T −
= −ρ ∂ρ
=0
∂v T− ∂ ln z T −
∂p
Thus, there is no discontinuity in ∂v T
at the bose einstein condensa-
tion point.
3.1. BOSE CONDENSATION 51
2
∂ p
Let us now examine the second derivative ∂v 2 . This is of course
T
again zero in the two phase region. Since, as we have already seen,
∂p
∂ ln z T
= kT ρ we get
2 " #
∂ p ∂ kT ρ3
= − ∂ρ
∂v 2 T ∂v ∂ ln z T T
" #
ρ2 ∂ kT ρ3
= − ∂ρ − ∂ρ
∂ ln z T
∂ ln z ∂ ln z T T
2
3 ∂ ρ
∂ρ kT ρ
2
ρ2 3kT ρ ∂ ln z T ∂ ln z 2
T
= −
∂ρ ∂ρ ∂ρ 2
∂ ln z T ∂ ln z T ∂ ln z T
∂2ρ
3kT ρ4 ∂ ln z 2
= − kT ρ5 T
∂ρ ∂ρ 3
∂ ln z T ∂ ln z T
Since ∂ ∂ρ
ln z T −
diverges, the first term in the above expression vanishes
as
we approach Tc . However, the second term
in2 this
expression has
∂2ρ ∂ ρ 1
P∞ 1 l
∂ ln z 2
in the numerator and this is given by ∂ ln z2 = a3 l=1 l 2 z
T T
1
and as we approach Tc this tends to a−3 ∞
P
l=1 l which is clearly diver-
2
gent. Thus in 2
2 ∂ ρ
∂ p ∂ ln z 2
T−
= −kT ρ5
∂v 2 T − ∂ρ 3
∂ ln z T −
we have the ratio of two divergent quantities. In order to ascertain the
behavioru of the second derivative
near Tc we must carefully examine
∂2ρ ∂ρ
the nature of the divergences in ∂ ln z2 and ∂ ln z T − , respectively.
T−
Thus the problem reduces to finding out the behaviour of the sums
P∞ 1 l P∞ − 1 l
l=1 l z and l=1 l
2 2 z near z = 1.
5This is connected with the integral representation of the Riemann zeta function
∞ Z ∞ 2
X 1 4 x dx
ζ(s) = = √
l s π 0 e 2 −1
x
l=1
52 3. SYSTEM OF FREE BOSONS
Now,
∞ ∞
x2 dx −1
Z Z
2 2
= dx x2 e−x 1 − ze−x
0 e 2 −z
x
0
∞ Z ∞
X 2
= z l−1
dx x2 e−lx
l=1 0
∞ l−1
4 X z
= √ 3
π l2
l=1
Thus we have
∞ ∞
zl x2 dx
Z
X 4z
3 =√
l 2 π 0 ex2 − z
l=1
3.2. ISOCHORES 53
3.2. Isochores
The isochores of a free boson gas are most easily obtained by study-
ing the isothermsfor various values of temperature T , at a fixed specific
volume v (i.e at a fixed density).
3
Recall that the critical density obeys ρc ∝ T 2 . This means that
3
the critical specific volume vc = ρ−1 c ∝ T − 2 . Hence as one goes to
higher temperatures the critical point on the (p-v) diagraqm moves
to the left. In the figure (*), the points α,β and γ are the points of
intersection of a line at constant specific volume with three different
isotherms. As can be seen, α is in the gas phase, β is just on the
critical point, whereas γ is in the two phase region. We obtain the
isochore for that particular density py plotting the pressures at the
points of interaction against corresponding temperature labels of the
isotherms. We also show corresponding points α,β and γ. The point
β corresponds to the critical point. In the region below β, we are in
5
the two phase region where p ∝ T 2 independent of the density. Above
β, we are in the gas phase and here the pressure is a function of both
the temperature and the density. For very high temperatures we have
the classical ideal gas result p = ρkT . If we now consider the isochores
for various values of density, we find that they have identical behaviour
5
in the region below the corresponding critical point ( p ∝ T 2 - with
the constant of proportionality independent of the density). As we
increase the density, of course the critical point moves towards higher
54 3. SYSTEM OF FREE BOSONS
5
temperatures. However, the region lying above the p ∝ T 2 curve is
never reached - it is completely unphysical.6Above the critical point,
5
the isochores deviate from the p ∝ T 2 behaviour.
6Contrast this, once again, with the liquid-gas phase transition. There the region
above the two phase coexistence line corresponds to the liquid state. The unphysical
region that vwe speak of here drives home the point that there is no analogue to
the liquid state in BEC.
3.2. ISOCHORES 55
Thus, 3 ρ
∂ ln z ∂ ln z
= = − 1 P2 ∞
T
zl
∂T v ∂T ρ a3 l=1 12
l
which vanishes as z → 1− (since the denominator diverges). Thus
∂p 5 pc 5 pc
= + ρkTc × 0 =
∂T v− 2 Tc 2 Tc
∂p ∂p ∂p
and hence ∂T v−
= ∂T v+
and so ∂T v
is continuous across the
critical point. 2
∂ p
We next consider the discontinuity in ∂T 2 :
v
2 2 2
∂ p ∂ p ∂ p
∆ 2
= 2
−
∂T v ∂T v+ ∂T 2 v−
From (3.2.1) we get
2
∂ p ∂ ∂p ∂ ∂p ∂ ln z
= +
∂T 2 v ∂T ∂T z v ∂T ∂ ln z T ∂T v v
Now,
∂ ∂p 5 p 5 ∂p
=− 2 +
∂T ∂T z v 2T 2T ∂T v
∂p
Since both p and ∂T v
are continuous across the critical point, we have
∂ ∂p
∆ =0
∂T ∂T z v
and are thus left with
2
∂ p ∂ ∂p ∂ ln z
∆ = ∆
∂T 2 v ∂T ∂ ln z T ∂T v v
∂ ∂p ∂ ln z
= −
∂T ∂ ln z T ∂T v v−
But,
3
ρa3 ρ 2 a3
∂p ∂ ln z 2 3
= ρkT × − P = − k P∞ zl
∂ ln z T ∂T v T ∞l=1
zl
1
2 l=1 1
l2 l2
56 3. SYSTEM OF FREE BOSONS
and therefore
ρ2
∂ ∂p ∂ ln z 3
= − k P∞ zl
∂T ∂ ln z T ∂T v v 2 l=1 12
l
P∞ l 1
3 2 3 l=1 z l ∂ ln z
2
+ kρ a hP i2
2 ∞ zl ∂T v
l=1 1
l2
As z → 1−,
∞
zl
r
X π
1 →
l=1 l 2 1−z
∞ r
X
l 1 1 π
zl 2 →
l=1
2(1 − z) 1 − z
r
3 ρa3 1 − z
∂ ln z
→ −
∂T v 2 T π
So, in this limit
r
∂ ∂p ∂ ln z 3 2 3 1 π
→ kρ a ×
∂T ∂ ln z T ∂T v v 2 2(1 − z) 1 − z
3r
1−z 3 ρa 1−z
× × −
π 2 T π
9 ρ 3 a6
k= −
8π T
And thus the discontinuity in the second derivative is
2
∂ p 9 ρ3c a6c
∆ = k
∂T 2 v 8π T
3.2.2. Behaviour of the heat capacity at the BEC phase
transition point. The above results derived for the derivatives of
pressure at the BEC transition can be used to deduce the behaviour of
the heat capacity using the fact that for any non-relativistic system of
free particles E = 23 pV . Thus
∂E 3 ∂p
CV = = V
∂T v 2 ∂T v
and so
3 ∂p
∆CV = V ∆ =0
2 ∂T v
3.3. LIQUID He4 57
Classical statistics
59
60 4. CLASSICAL STATISTICS
the coordinate associated with the center of the αth volume. Prop-
erly symmetrized (antisymmetrized) products of these wavefunctions
ψα~k (~r) form a complete set of states for a system of bosons (fermions).
The system wavefunction Ψ (~r1 , . . . , ~rN ) can be expanded in terms of
this complete set.
Upto this point, everything we have done is exact. Now recall that
l a, so that a particle can be localized within a particular small
volume. Again, since d l the chances of two or more particles being
localized within the same volume are extremely remote. So, of all the
configurations, the ones in which each particle occupies a seperate small
volume will be the dominant one. The corresponding wavefunction
(after suitable syymetrization or antisymmetrization) takes the form
1 X Y
Ψ (~r1 , . . . , ~rN ) = √ (±)P ψαi~ki (P ~ri )
N! P i
where the + amd - signs are for bosons and fermions, respectively.
The summation is over all possible perutations of particle coordinates.
Since there is no overlap between particles belonging to different small
volumes, and ψαi~ki (~r) are normalized, the wavefunction Ψ will be nor-
malized. We now calculate hΨ| H |Ψi and replace the partition function
of the system Tr e−βH by a sum over terms of the form e−βhΨ|H|Ψi . This
is our essential approximation1.
To begin with, let us assume that the internal degrees of freedom
are absent, i.e.
X p2
i
H= + U (~r1 , . . . , ~r2 )
i
2m
Since λ a, we can always choose our l so that l λ. In that case
~ ~ ~ ~
hΨ| U (~r1 , . . . , ~rN ) |Ψi ≈ U R1 , . . . , RN hΨ| Ψi = U R1 , . . . , RN
and thus
X k2
hΨ| H |Ψi = i ~ ~
+ U R1 , . . . , RN
i
2m
To calculate the partition function, we will have to sime over all possible
momentum assignments and positions of the small volumes “occupied”
1This does not mean that we are claiming that Ψ is an approximate energy eigen-
state. To calculate Tr e−βH , all we need is a complete set of states which need not
be eigenstates of H. The products of the plane waves localized in small volumes is
a complete set. All we are assuming is that the sum in the trace is dominated by
contribution from one particular state Ψ of this complete set.
4. CLASSICAL STATISTICS 61
by ~
the atoms or molecules. total number
of states
for Ri in the range
~ i, R
R ~ i and ~ki in the range ~ki , ~ki + d~ki is
~ i + dR
QN QN
i=1 d3 Ri i=1 d3 ki
N
N ! (2π~)
Here, the factor of N ! is essential to avoid double counting while sum-
ming over all possible small volumes. Clearly, some of the changes in
position (assuming we have already summed over all possible momen-
tum labels), would result in mere interchange of the small volumes.,
and therefore not produce a new state (recall that the particles are
identical). There are in fact N ! such permutations, hence the factor of
N ! in the denominator.
So the contribution to the grand partition function from all N par-
ticle configurations (apart from the factor of eβµ ) is :
Z Y 3 " #!
1 d Ri
Z Y X k2
QN = d3 ki exp −β i
+U R ~ 1, . . . , R
~N
N! (2π~)N 2m
i i i
p = ρkB T
a3 1
(recall that kB T q
∝ 5 ). Thu8s the total energy is
qT 2
3 ∂
E = G + T S − pV = N kB T + N kB T 2 ln q
2 ∂T p
2We know that ρ = ∂ p
but y = q
implies ln y = ln z + ln aq3 and hence
∂(ln z) kB T a3 z
∂ p
ρ = ∂(ln y) kB T .
EXAMPLE : CLASSICAL IDEAL GAS 63
From the equipartition theorem, the first term is just the average kinetic
energy. As far as the second term is concerned, we see that
∂ 1 X
N kB T 2
ln q = N kB T 2 × 2
ui e−βui
∂T p qkB T int. d.o.f.
P −βui
ui e
= N P −βui = N hui
e
where hui is the average internal energy. So, for a classical ideal gas,
we have
3
E = N kB T + N hui
2
Gibb’s Paradox. Using the equation of state pV = N kB T for the
ideal gas, we get
G 5 ∂
S = − + N kB + N kB T ln q
T 2 ∂T p
3
a 5 ∂
= −N kB ln p − N kB ln + N kB + N kB T ln q
kB T q 2 ∂T p
3
N kB T a 5 ∂
= −N kB ln − N kB ln + N kB + N kB T ln q
V kB T q 2 ∂T p
V
N k ln + N f (T )
N
where
a3
5 ∂
f (T ) = kB + kB T ln q − kB ln kB T − kB ln
2 ∂T p kB T q
is a function of the temperature only.
Now consider two different perfect gases A and B. The inter-
moleculer forces are all equal, UAA = UBB = UAB (In the case of ideal
gases all are identically zero). Their atomic (or molecular) masses are
also the same. Despite this, mthe two gases are differnt because they
are observed to be different - i.e. there is some probe to which they
couple differently. The particles that make up the two gases are thus
distinguishable. Truly identical particles cannot be distinguished and
the total wavefunction must in their case be totally symmetric or anti-
symmetric (no matter what the forces are).
Enclose the gases A and B in boxes of volumes VA and VB , in
equilibrium at a temperature T . Let the number of gas molecules NA
and NB be such that NVAA = NVBB , i.e. the densities in the two boxes are
64 4. CLASSICAL STATISTICS
H = HA + HB
and therefore
QN = Tr e−βHA Tr e−βHB
= QA (NA , V, T )QB (NB , V, T )
F = FA (NA , V, T ) + FB (NB , V, T )
S = SA (NA , V, T ) + SB (NB , V, T )
Sf − Si = N kB ln 2
EXAMPLE : CLASSICAL IDEAL GAS 65
Si = Sf
Classical statistical mechanics did not take into account the indis-
tinguishability of molecules - and hence the factor N ! was missing from
the denominator of the expression for QN . This in turn, means that
the expression for the entropy did not have the −N ln N term. Thus,
when Gibb’s originally calculated the entropy of mixing for two gases,
he had
Si = NA kB ln VA + NB kB ln VB + (NA + NB ) f (T )
Sf = N kB ln V + N f (T )
u1 . Then
q = ω0 + ω1 e−βu1 + . . .
1
Now, typically u1 ∼ 1 ev and at room temperature kB T ∼ 40
ev. So,
for the first and higher excited states
e−βui 1 for i ≥ 1
Naively, then, one would expect the contributions of the excited states
to the internal partition function to be negligible. As it turns out, this
expectation fails at very low densities!
To see why, let us consider an assemly of Hydrogen atoms. The
spectrum of Hydrogen is well known to be
R
En = −
n2
where R is the Rydberg constant. So the excitation energy to the n-th
level is given by
R
un = R − 2
n
Since the degeneracy of the n-th state is n2 the internal partition func-
tion for a hydrogen atom is given by
∞
2
X
−βR
q=e n2 e−βR/n
n=1
n20 a0 = ρ−1/3
and the total contribution of the excited states to the electronic parti-
tion function is negligible only when
n0
X βR
n2 e n2 e−βR 1
n=2
For this to be valid, n0 must be small. A good criterion for this could
be
n20 e−βR 1
which means
e−βR
1 1
a0 ρ 3
Hence we can ignore excited states only at sufficiently large densities or
low temperatures. At room temperature, this is a valid approximation
for all gases at normal densities. On the other hand, in almost all
stellar configurations the density is so small and the gas so hot that
the excited state contribution is much larger than that of the ground
state. As a result stellar gases are almost always completely ionized.
Diatomic gases. In this case we have to consider rotational and
vibrational motion as well as the electronic degrees of freedom. The
total internal energy, then is
degeneracy and so
∞
X 1
qvib = e−βn~ω =
n=0
1 − e−β~ω
Let us now consider the case of a few typical gases. For H2 the
value of ~ω
kB
is 5798 K while that for HCl is 4031 K. This shows that at
room temperature
• Anharmonicity is negligible and
• Vibrational states are hardly excited!
On the other hand, rotational levels are given by
j(j + 1)~2
urot = , j = 0, 1, 2, . . .
2I
where I is the moment of inertia of the molecule. Corresponding eigen-
functions are the rotational spherical harmonics. Degree of degeneracy
of each rotational state is 2j + 1 and each has a parity of (−1)j . Again,
~2
typical values of 2Ik B
are 85 K for H2 and 15 K for HCl. Thus, unlike
the case with vibrations, rotational levels are easily excited at room
temperature.
Homonuclear molecules. To investigate the effect of the rotational
levels, we first consider molecules with identical atoms (e.g. H2 ). In the
ground state the electron spins are antiparallel4. We assume sufficiently
high densities and low temperatures that the excited electronic states
can be neglected. The electronic ground state is nondegenerate and
antisymmetric. As, we have seen, we can also ignore the excited vibra-
tional levels at room temperature. We now consider the nuclear spin
states. Let’s assume that the nuclear spins are 21 (which is certainly
the case for H2 ). Combined spin states can be
• SN = 1 - triplet (degeneracy = 3) . The spin wave function
is symmetric and hence the space wave function has to be
antisymmetric, i.e. j = 1, 3, 5, . . ..
4This is because in order to bind and hence reduce the electrostatic energy, the
electron wavefunctions must have large amplitudes and overlap at the center of
the molecule. Because of the Pauli exclusion principle, the overlap at the center
is large only when the spins are antiparallel. This argument, admittedly, is rather
heuristic - but is works nonetheless. A more precise argument would involve the
sign of the exchange integral. For diatomic molecules the sign is such that the
antiferromagnetic order is favored.
EXAMPLE : CLASSICAL IDEAL GAS 69
where the factor of 3 comes from the degenercay of the nuclear spin
states. On the other hand, for para-hydrogen
X j(j + 1)
qpara = (2j + 1) exp −
j=even
2IkB T
qpara = 1 + 5e−3/IkB T + . . .
qrot = 4IkB T
where sA and sB are the two nuclear spins. At high temperatures this
tends to
qrot = 2 (2sA + 1) (2sB + 1) IkB T
Corrections to these high temperature results can be calculated by using the Euler Mclurin
formula :
n Z n
X 1
f (j) = f (j)dj − [f (n) − f (m)] −
m m 2
X (−1)l Bl h (2l−1) i
f (n) − f (2l−1) (m)
(2l)!
l≥1
where B1 = 61 , B2 = 1
30
,...are the Bernoulli numbers.
X X
qrot = (2j + 1)e−urot /kB T
j
nuclear
spin
X X
= (2j + 1)e−εj(j+1)
j
nuclear
spin
~2
where ε = 2IkB T
. Consider first the heteronuclear case A 6= B. Total number of nuclear spin
states is ω = (2sA + 1) (2sB + 1). In this case there are no exchange symmetry requirements
on the wavefunction, and so all j values are allowed. Anticipating the homonuclear case,
however, we split the sum over j into one over even j and another over odd j.
X ∞
X ∞
X
(2j + 1) e−εj(j+1) = (4m + 1) e−2εm(2m+1) = φ(m)
j=even m=0 m=0
EXAMPLE : CLASSICAL IDEAL GAS 71
It’s easy to see that the higher order derivatves are at least of the order of ε2 . Hence
Z ∞
X 1
(2j + 1) e−εj(j+1) = dmφ(m) − [φ(∞) − φ(0)]
j=even 0 2
B1 h (1) i B h
2
i
+ φ (∞) − φ(1) (0) − φ(3) (∞) − φ(3) (0) + . . .
2! 4!
Z ∞
1 d −2εm(2m+1) 1 1
= − dm e + − (4 − 2ε)
2ε 0 dm 2 12
1
− 96ε + O(ε2 )
24 × 30
1 1 1
= + + ε + O(ε2 )
2ε 6 30
Similarly, it can be shown that
X 1 1 1
(2j + 1) e−εj(j+1) = + + ε + O(ε2 )
2ε 6 30
j=odd
and so
∞
X 1 1 1
qrot = (2sA + 1) (2sB + 1) (2j + 1) e−εj(j+1) = ω + + ε + O(ε2 )
j=0
ε 3 15
and since
1 1 1 2
qrot = ω + + ε + O(ε )
ε 3 15
ω 1 1 2
= 1+ ε+ ε + O(ε3 )
ε 3 15
we have
1 1 2
ln qrot = ln ω − ln ε + ln 1 + ε + ε + ...
3 15
1 1 2 1 2
= ln ω − ln ε + ε+ ε − ε + O(ε3 )
3 15 18
1 1 2
= ln ω − ln ε + ε + ε + O(ε3 )
3 90
Hence
∂ 1 1 1 dε
ln qrot = − + + ε + O(ε2 )
∂T ε 3 45 dT
2
dε
and since dT
= − 2Ik~ 2 = − Tε we get
BT
1 1 2 3
hEi = kB T 1 − ε − ε + O(ε )
3 45
CHAPTER 5
Imperfect gases
1This is not a serious restriction. In most of the systems we will be interested in,
i.e. collections of atoms and molecules, the basic interaction is always the Couilomb
potential for which u ∼ r−1 . So we maight naively expect that our current analysis
may not be applicable to this case. However, charged particles always tends to form
neutral groups and outsuide these groups the potential would always be screened
and fall off much faster than r−3 (usually the potential is due to mutual polarization
of neutral groups. These induced multipole potentials tend to fall off at least as fast
as r−6 ). So the current analysis is indeed applicable to these systems. However, in a
system interacting gravitationally there is no such shielding and this analysis cannot
be applied to these systems. In fact for such a system the thermoidynamic limit
does not exist (i.e. kBpT = V1 ln Q is not independent of V in the limit V → ∞).
73
74 5. IMPERFECT GASES
N =0
N ! i=1
h
∞ 3N
X eβµN 2πm 2
= 2
q N QV (N )
N =0
N ! βh
∞
X eβµN q N
= QV (N )
N =0
N !a3N
∞
X yN
= QV (N )
N =0
N!
βµ
where y = e a3 q . Here q is the internal partition function and a =
√ h is the thermal wavelength and
2πmkB T
N
Z Y
QV (N ) = d3 ri e−βUN
i=1
7654
0123 7654
0123 7654
0123 7654
0123 7654
0123 7654
0123
1 2 1 2 1 2
7654
0123 7654
0123 7654
0123 7654
0123 7654
0123 7654
0123
3 4 3 4 3 4
=1 = f12 = f13
7654
0123 7654
0123 7654
0123 7654
0123 7654
0123 7654
0123
1 2 1 2 1 ?? 2
??
7654
0123 7654
0123 7654
0123 7654
0123 7654
0123 7654
0123
?
3 4 3 4 3 4
7654
0123 7654
0123 7654
0123 7654
0123
1 2 1 ?? 2
??
7654 7654 7654 7654
0123 0123 0123 0123
?
3 4 3 4
= f12 f13 f24 f23 = f12 f13 f14 f24 f23 f34
7654
0123 7654
0123 7654
0123 7654
0123
1 2 1 2
7654 7654 7654 7654
0123 0123 0123 0123
is not a cluster but this one is
3 4 3 4
Since the two body potentials fall off at infinite seperation, a cluster
contributes only when all the points are close together and then this
integral ∝ V where V is the volume of the system. Similarly integral
76 5. IMPERFECT GASES
7654
0123 7654
0123
1 2
7654
0123 7654
0123
of a product of two clusters e.g. ∝ V 2 , the integral of
7654
0123 7654
0123
3 4
1 2
7654
0123 7654
0123
the product of four clusters e.g. ∝ V 4 , and so on.
3 4
7654 7654
0123 0123
A few examples may make this point clear. Let us consider the
1 2
7654
0123 7654
0123
integral of cluster = f12 f13 f34 .
3 4
Z Z
3 3 3 3
d r1 d r2 d r3 d r4 f12 f13 f34 = d3 r1 d3 r12 d3 r13 d3 r34 f12 (r12 ) f13 (r13 ) f34 (r34 )
Z
∝ d3 r1 = V
where in the first step we have substituted the relative vectors ~r12 =
~r2 −~r1 , ~r13 = ~r3 −~r1 and ~r34 = ~r4 −~r3 in place of the original vectors ~r2 ,~r3
and ~r4 (it is easy to verify theat the Jacobian of this transformation
is 1). Due to the short range nature of the functions fij , integrals like
R 3
d r12 f12 (r12 ) are not proprtional to the volume. In particular they
7654
0123 7654
0123
stay finite in the thermodynamic limit V → ∞2.
1 2
7654 7654
0123 0123
As another example consider the cluster = f12 f13 f23 f24 .
3 4
It’s integral is
Z Z
3 3 3 3
d r1 d r2 d r3 d r4 f12 f13 f23 f24 = d3 r1 d3 r12 d3 r13 d3 r24
2It is to achieve precisely this that the function uij has to fall off faster than r−3 .
5.1. CLUSTER EXPANSIONS 77
7654
0123 7654
0123
1 2
7654
0123 7654
0123
Again consider a product of two clusters, = f12 f34 .
3 4
Z Z
3 3 3 3
d r1 d r2 d r3 d r4 f12 f34 = d3 r1 d3 r12 d3 r3 d3 r34 f12 (r12 ) f34 (r34 )
Z Z
∝ d r1 d3 r3 = V 2
3
b1 = 1
2
Z Y
1 1
b2 = d3 ri f12
2! V i=1
Z Y3
1 1
b3 = d3 ri [f12 f13 + f12 f23 + f13 f23 + f12 f13 f23 ]
3! V i=1
..
.
Z Yl
11 X Y
bl = d3 ri fij
l! V i=1 All l clusters (i,j)
Note for example that the integrand in b3 is the sum of all possible
three clusters
7654
0123 7654
0123 + 7654
0123 7654
0123 + 7654
0123 7654
0123 + 7654
0123 7654
0123
1 2 1 2 1 2 1 2
7654 7654 7654 7654
0123 0123 0123 0123
3 3 3 3
The cluster integrals, bl are functions of V and T in general. How-
ever, provided the two body potential falls of faster than r−3 at large
separations, the limit limV →∞ bl (V, T ) exists.
∞
p X
= bl y l
kB T l=1
∞
X
and ρ = lbl y l
l=1
4
7654
0123 7654
0123 + 7654
0123 7654
0123 + 7654
0123 7654
0123 + 7654
0123 7654
0123
1 2 1 2 1 2 1 2
7654 7654 7654
0123 7654 7654
0123 7654 7654
0123 7654
0123 0123 0123 0123 0123
3 4 3 4 3 4 3 4
Now, integrations over different groups are independent, since they
correspond to different clusters. For example, the grouping {1, 2, 3} {4}
in our last example integrates to 3!V b3 × 1!V b1 = 6V 2 b3 . It is easy to
see that each decomposition contributes a factor of
Y
(l!bl V )nl to QV (N )
l
and thus
X N! Y
QV (N ) = Q nl n !]
(l!bl V )nl
l [(l!) l
{nl } l
4In general, different terms in the expansion of QN and hence Q will contain differ-
ent powers of V , depending on the number of associated clusters. We will show that
the series is going to exponentiate in a remarkable way, and in the limit V → ∞ the
exponent is proportional to V (provided uij falls off faster than r−3 at large sepa-
rations). Hence kBpT will be independent of V as V → ∞, and the thermodynamic
limit will exist.
5There are in general N ! ways of permuting the N points. Out of these, the l!
ways of permuting the elements within each of the nl groups do not lead to a new
grouping. Again, the permutations of the nl groups among themselves, in nl ! ways,
do not count as a new decomposition.
5.1. CLUSTER EXPANSIONS 79
P
with the sum running over all {nl } satisfying l nl l = N . When we
calculate the grand partition function Q, the sum runs over all values
of N from 0 to ∞. Hence, we can relax this restriction over the possible
values of nl and sum over all nl from 0 to ∞.
∞
X yN
Q = QN (V )
N =0
N!
∞ Y ∞ n
X V bl y l l
=
nl =0 l=1
nl !
∞ X ∞ n
Y V bl y l l
=
l=1 n =0
nl !
l
∞ ∞
!
V bl y l
Y X
l
= e = exp V bl y
l=1 l=1
and hence ∞
p 1 X
= ln Q = bl y l
kB T V l=1
and ∞
∂ p X
ρ= = lbl y l
∂ ln y kB T V,T l=1
Q.E.D.
p
For very large volumes V → ∞ , both and ρ become volume
kB T
independent intensive quantities, since limV →∞ bl (V, T ) exists.
Example 5.1.1. Consider the perfect gas for which uij = 0 and
hence fij = 0. We have b1 = 1 and bl = 0 for all l > 1. Hence
p
=y=ρ
kB T
which is the equation of state of the ideal gas.
and hence
p
= y + b2 y 2 + . . .
kB T
= ρ − 2b2 ρ2 + b2 ρ2 + . . .
= ρ − b 2 ρ2 + . . .
which is the virial expansion upto the second virial coefficient. Hence
we see that the second virial coefficient is
Z
1
b2 = d3 r1 d3 r2 f12 (r12 )
2V
Z
1
d3 r1 d3 r2 e−βU (|~r1 −~r2 | − 1
=
2V
Z
1
d3 r e−βU (r) − 1
=
2
Let us compare this with the well known Van der Waals equation
of state for one mole of gas
a
p + 2 (v − b) = RT
v
NA
which can be rewritten in terms of the particle density ρ = v
(where
NA is the Avogadro number) in the form
p + a0 ρ2 (1 − b0 ρ) = ρkB T
where a0 = a
NA2 and b0 = b
NA
. Thus
p ρ a0 2
= − ρ + ...
kB T 1 − b0 ρ kB T
a0
0
= ρ+ b − ρ2 + O(ρ2 )
kB T
Comparing this with the virial expansion we have just obtained we get
Z
0 1
a = − d3 r u(r)
2 r>d
1 4π 3
b0 = d
2 3
Or, in terms of the original Van der Waals coefficients
Z
1 2
a = − NA d3 r u(r)
2 r>d
1 4π
b = NA d3
2 3
For attractive forces, u(r) < 0 and hence a > 0, while for repulsive
forces u(r) > 0 and a < 0. The positive constant b can be seen to be
related to the volume occupied by the hard spheres, thus confiming our
intuitive notions about it’s origin.
In the above we have calculated upto the 2nd virial coefficient. If
we retain upto the 3rd virial coefficient, we will see deviations from the
Van der Waals equation. So the Van der Waals equation is correct only
at low density and weal long-range intermolecular potentials.
ρ = y + 2b2 y 2 + 3b3 y 3 + . . .
82 5. IMPERFECT GASES
which leads to
And so the third virial coefficient is 4b22 − 2b3 . Let us observe what this
means in terms of the cluster integrals :
d3 r1 d3 r2 7654
0123 7654
0123
Z
1 h i
b2 = 1 2
7654 7654
2!V
0123 0123
2 2
7654 7654
0123 0123
Z
1
b3 = d3 r1 d3 r2 d3 r3
3 1 ??? + 1
7654 7654
???
0123 0123
3!V
3 3
Now,
7654
0123
2
7654
0123
Z
Z
3 3 3
d r1 d r 2 d r3
1 ???
= d3 r1 d3 r2 d3 r3 f12 (r12 ) f13 (r13 )
7654
0123
3
Z
= d3 r1 d3 r12 d3 r13 f12 (r12 ) f13 (r13 )
Z 2
3
= V d r12 f12 (r12 ) = V (2b2 )2
So,
7654
0123
2
7654
0123
Z
2 1
b3 = 2(b2 ) + d3 r1 d3 r2 d3 r3
1 ???
7654
0123
3!V
3
7654
and hence we find that the third virial coefficient is
0123
2
7654
0123
Z
2
4(b2 )2 − 2b3 = − d3 r1 d3 r2 d3 r3
1 ???
7654
0123
3!V
3
5.1. CLUSTER EXPANSIONS 83
7654
0123
A cluster is called irreducible if it is not reducible6.
2
For example, the cluster 7654
0123
1? is reducible since it becomes discon-
7654
0123
??
7654
0123
3
2
nected if we remove the vertex labelled 1. On the other hand, 7654
0123
1 ??
7654
0123
?
3
is irreducible.
We will see that the virial coefficients involve only irreducible clus-
ters. This result is Mayer’s second theorem. We define
Z k+1 !
1 Y X Y
βk = d3 ri fij
k!V i=1 (i,j)
irreducible
cluster of
k + 1 points
Note that the sum is over irreducible clusters of k + 1 particles. Then
we have
Theorem. Mayer’s second theorem
" ∞
#
p X k
=ρ 1− βk ρk
kB T k=1
k + 1
Mayer’s second theorem gives a direct virial expansion in terms of
the βk ’s. The original proof is due to Mayer (Mayer & Mayer ). There
is an elegant derivation due to Ford and Uhlenbeck (Studies in Stat.
Mech 1, 123 (1962)). We will present a proof due to R. Friedberg (J.
Math. Phys. 16 20 (1975)) in the appendix. This proof has a general
framework that can be used in field theory as well.
6These definitions have analogues in field theory. There, of course reducibility
or irreducibility is defined in terms of cutting through lines rather than through
vertices.
84 5. IMPERFECT GASES
Not only the virial equation of state, but all thermodyanamic prop-
erties are available in terms of the irreducible cluster integrals βk . Con-
sider a monoatomic gas for which the electronic excitations can be
ignored. Then the internal partition function q = ω where ω is the
degeneracy of the ground state. Then
∂
E = − ln Q
∂β V,βµ
2 ∂ p
= kB T V
∂T kB T z
From Mayer’s second theorem
" ∞
#
p X k k
=ρ 1− βk ρ
kB T k=1
k+1
we get
" ∞
# ∞
∂ p ∂ρ X
k
X k ∂βk
= 1− kβk ρ − ρk+1
∂T kB T z ∂T z k=1 k=1
k + 1 ∂T z
Again
" ∞
#
∂ p ∂ρ X
ρ= = 1− kβk ρk
∂ ln y kB T V,T ∂ ln y V,T k=1
which integrates to
∞
X
ln y = ln ρ − βk ρk + f (V, T )
k=1
ln z = ln y + ln a3 − ln q
∞
3 X
= − ln T + ln ρ − βk ρk + constant
2 k=1
and so
" ∞
# " ∞ #
∂ρ X 3 X ∂βk
1− kβk ρk = ρ + ρk
∂T z k=1
2T k=1
∂T z
5.2. A SOLUBLE MODEL : 1 DIMENSIONAL GAS OF HARD RODS 85
and thus
" ∞ # X ∞
∂ p 3 X ∂βk k k ∂βk
z = ρ + ρ − ρk+1
∂T kB T 2T k=1
∂T z k=1
k + 1 ∂T z
" ∞ #
3 X 1 ∂βk
= ρ + ρk
2T k=1
k + 1 ∂T z
and so
" ∞ #
3 X 1 ∂β k
E = kB T 2 V ρ + ρk
2T k=1
k + 1 ∂T z
" ∞ #
3 X 1 ∂βk
= N kB T +T ρk
2 k=1
k + 1 ∂T z
Again
∞
!
a3 a3
X
k
G = N µ = N kB T ln z = N kB T ln y + ln = N kB T ln ρ − βk ρ + ln
ω k=1
ω
Since G = E + pV − T S we have
" ∞
#
ω X
k E + pV
S = N kB ln 3 + βk ρ +
a ρ k=1 T
" ∞
# " ∞ #
ω X 3 X 1 ∂β k
= N kB ln 3 + βk ρk + N kB +T ρk
a ρ k=1 2 k=1
k + 1 ∂T z
" ∞
#
X k
+N kB 1 − βk ρk
k=1
k + 1
" ∞ #
ω 5 X 1 ∂βk
= N kB ln 3 + + βk + T ρk
a ρ 2 k=1 k + 1 ∂T z
" 3 ∞ #
mkB T 2 ω 5 X 1 ∂
= N kB ln + + (T βk ) ρk
2π~2 ρ 2 k=1 k + 1 ∂T z
Thus
1
QN = (L − N d)N
aN N !
and so the free energy is given by
7Of course this can also be obtained from the grand canonical partition function.
5.2. A SOLUBLE MODEL : 1 DIMENSIONAL GAS OF HARD RODS 87
We will now show how one can obtain the same result using Mayer’s
theorem. First note that in this case
(
−1 for |xi − xj | < d
fij = e−βuij − 1 =
0 for |xi − xj | > d
So
Again,
7654
0123
2
7654
0123
Z
1
β2 = dx1 dx2 dx3 1
7654
???
0123
2!L
3
Z
1
= dx12 dx13 f (x12 ) f (x13 ) f (x23 )
2
where x12 = x2 − x1 , x13 = x3 − x1 and x23 = x3 − x2 = x13 − x12 . Here
the integrand is nonvanishing and equal to -1 for
−d < x12 < d, −d < x13 < d, and − d < x13 − x12 < d
Thus for d > x12 ≥ 0, the integrand is -1 for −d + x12 < x13 < d and
for −d < x12 ≤ 0 for −d < x13 < d + x12 . So, we have
Z d Z d Z 0 Z d+x12
1
β2 = − dx12 dx13 + dx12 dx13
2 0 −d+x12 −d −d
Z d Z 0
1
= − dx12 (2d − x12 ) + dx12 (2d + x12 )
2 0 −d
3
= − d2
2
So upto O(ρ3 ) we have from Mayer’s theorem
p 1 2 2
= ρ 1 − β1 ρ − β2 ρ − . . .
kB T 2 3
= ρ 1 + dρ + d2 ρ2 + . . .
r3
2 r 1 3
= π d d− − d −
2 3 8
2 1 1
= π d3 − rd2 + r3
3 2 24
and hence
4 1
V (r) = π d3 − rd2 + r3
3 12
Hence
d
5π 2 6
Z
1
β2 = − 4πr2 V (r) dr = − d
2 0 12
5.3. HARD SPHERE GAS IN THREE DIMENSIONS 89
also, all the circles are numbered 1, 2, . . . , nc while the squares are num-
bered 1, 2, . . . , ns . For example
3
7654
0123 7654
0123 7654
0123 7654
0123
1 1 1 2 2 1 1 2 4
are all trees while
3?
??
7654 7654
??
0123 0123
1 1 2 2 4
is not (it is a loop which can be changed into a tree by cutting a line).
Two trees will be considered identical if they are topologically identical
including the numbers associated with the circles and the squares.
We can represent a cluster by a dual tree. For example
7654
0123
7654
0123 1
2
7654
0123
1 −→ 1?
7654
???
0123
??
7654 7654
??
0123 0123
3
2 3
7654
0123 7654
0123 −→ 7654
0123 7654
0123
1 2 1 1 2
91
92 A. FRIEDBERG’S PROOF OF MAYER’S SECOND THEOREM
7654
0123 7654
0123
1 2
7654
0123 7654
0123
3 4
−→
7654
0123 7654
0123
1 ?? 2
??
7654 7654
0123 0123
?
3 4