172 Sample Chapter
172 Sample Chapter
In 1902 Gibbs, one of the founders of Statistical Mechanics, introduced the concept of ensemble
(a French word meaning assembly of systems). An ensemble is a collection of essentially inde-
pendent systems, which are macroscopically identical but microscopically different.
Three types of ensemble are important in statistical mechanics. The classification depends on
the matter in which the systems interact.
The microcanonical ensemble is a collection of independent systems, having the same number
of particles N, volume V and an energy between E and E + δE. So the individual systems of
a microcanonical ensemble are separated by rigid, impermeable and insulated walls, such that the
values of E,V and N for a particular system are not affected by the presence of other systems. Thus,
in microcanonical ensemble, neither energy nor matter is exchanged.
N,V, E N,V, E
N,V, E N,V, E
57
60 • An Introduction to Equilibrium Statistical Mechanics
Now
N 3
1 N → −p 2 = 1
E= ∑
2m i=1 i ∑ ∑ p2iα
2m i=1 α=1
N 3
⇒ 2mE = ∑ ∑ p2iα (2.1.9)
i=1 α=1
N is of the order of Avogadro’s number and so very large. Thus, Ω (E) is an extremely rapidly
increasing function of the energy E of the system.
The entropy of the ideal gas is
S = k ln ΩE
3
2mπE 2 3N
= Nk ln 2
+ Nk lnV − k ln Γ (2.1.17)
h0 2
3N 3N 3N
Now Γ = −1 ! = !, since N 1.
2 2 2
Again by Stirling’s approximation
3N 3N 3N 3N
ln != ln −
2 2 2 2
23
3N 3N
= N ln −
2 2
23
3N ∼ 3N 3N
∴ ln Γ = N ln − (2.1.18)
2 2 2
Using (2.1.18) in (2.1.17)
3 23
2mπE 2 3N 3
S (E,V ) = Nk ln + Nk lnV − Nk ln + Nk
h20 2 2
3
4mπ E 2 3
⇒ S (E,V ) = Nk ln V 2
+ Nk (2.1.19)
3h0 N 2
Solving for E in terms of S and V , we obtain the internal energy given by
3 h20 N
2 S
E (S,V ) = exp −1 (2.1.20)
4π m V 23 3 Nk
Again from thermodynamics
T dS = dE + PdV (2.1.21)
Then the temperature is given by
∂E 2 E
T= = (2.1.22)
∂S V 3 Nk
Hence, the specific heat at constant volume is given by
∂E 3
CV = = Nk (2.1.23)
∂T V 2
From (2.1.21), the pressure P is given by
∂E 2E
P=− = (2.1.24)
∂V S 3V
62 • An Introduction to Equilibrium Statistical Mechanics
This is the equation of a sphere of radius R in 3N-dimensional space; the surface of the sphere
corresponds to a constant energy E. The number of microstates Ω (N,V, E) is equal to the number
of lattice points on the positive segment of this spherical surface. If the energy of the system is
specified in between E and E + δE, the number of microstates Ω (N,V, E → E + δE) within this
range is equal to the number of lattice points lying within the positive part of the spherical shell
formed between the surfaces of radii R and R + δR in the 3N-dimensional space.
If Φ (N,V, E) is the volume of the positive part of the 3N-dimensional sphere of radius R, then
Ω (N,V, E → E + δE) = Ω = Φ (N,V, E + δE) − Φ (N,V, E)
By Taylor’s expansion,
∂Φ 1 ∂2 Φ
Φ (N,V, E + δE) = Φ (N,V, E) + δE + (δE)2 + . . .
∂E 2! ∂E 2
∼ Φ (N,V, E) + ∂Φ δE
=
∂E
∂Φ
∴ Ω = Φ (N,V, E + δE) − Φ (N,V, E) ∼ = δE (iv)
∂E
Now from appendix II, the volume of the 3N-dimensional sphere of radius R is
3N
π2
V3N (R) = R3N (v)
3N
!
2
Hence,
√ 3N
πL
3N 3N
1 π 2 2LE
3N
hc E 3N
Φ (N,V, E) = = (vi)
3N
2 hc 3N
! !
2 2
√ 3N
πL
∂Φ 3N hc E 3N−1 3N
∴ = = Φ (vii)
∂E 3N E
!
2
Thus,
∂Φ δE
Ω= δE = 3NΦ (N,V, E) (viii)
∂E E
Taking log on both sides
δE
ln Ω = ln Φ + ln 3N + ln (ix)
E
But from (vi), we have
√ √
πLE 3N ∼ πLE 3N 3N 3N
ln Φ = N ln − ln ! = N ln − ln + (x)
hc 2 hc 2 2 2
Elements of Ensemble Theory • 65
P ROBLEM 2.3: The quantity Ω (N,V, E) is called the microcanonical partition function. Show that
= ln Ωmax , where Ωmax is the largest Ωi in the series.
ln Ω (N,V, E) = ln (Ω1 + Ω2 + · · · ) ∼
66 • An Introduction to Equilibrium Statistical Mechanics
S OLUTION
Let us assume that there are as many N systems with Ωi ’s comparable to Ωmax , that is, as many
such systems as there are particles in any one system. Thus,
Ω∼ = N × Ωmax ⇒ ln Ω ∼ = ln Ωmax + ln N (i)
But ln N N.
∴ ln Ω (N,V, E) ∼
= ln Ωmax (Proved) (ii)
P ROBLEM 2.4: A one-dimensional chain is made up of N identical elements of length l. The angle
between successive elements can be either 0◦ or 180◦ , but there is no difference in internal energy
between these two possibilities. For the sake of counting, one can think of each element as either
pointing to the right (+) or to the left (−). Then one has
N = n+ + n−
L = l (n+ − n− ) = l (2n+ − N) .
(a) Use the microcanonical ensemble to find the entropy as a function of N and n+ , S (N, n+ ).
(b) Find an expression for the tension in the chain as a function of T , N and n+ , ℑ (T, N, n+ ).
(c) Rearrange (b) to give the length as a function of N, T and ℑ.
S OLUTION
N!
(a) The number of ways of choosing n+ elements from a total of N is . It follows
(N − n+ )!n+ !
that
N!
Ω (N, n+ ) = (i)
(N − n+ )!n+ !
The entropy of the system is
S (N, n+ ) = k ln Ω = k [ln N! − ln (N − n+ )! − ln n+ !]
Using Stirling’s approximation, the above equation becomes
= k [N ln N − N − (N − n+ ) ln (N − n+ ) − n+ ln n+ + n+ ]
S (N, n+ ) ∼
= k [N ln N − (N − n+ ) ln (N − n+ ) − n+ ln n+ ] (ii)
(b)
ℑ ∂S ∂S ∂n+ ∂S 1
=− =− =− ×
T ∂L N,E ∂n+ ∂L ∂n+ 2l
k N − n+ n+
=− + ln (N − n+ ) − − ln n+
2l N − n+ n+
k N − n+
= − ln
2l n
+
kT N − n+
∴ (N, T, n+ ) = − ln (iii)
2l n+
Elements of Ensemble Theory • 67
Let the number of particles in the ground state of zero energy be N0 and that in the excited state
with energy ε be Nε .
∴ N0 + Nε = N (2.1.25)
The total energy of the system is
E
E = (N0 × 0) + (Nε × ε) ⇒ Nε = (2.1.26)
ε
68 • An Introduction to Equilibrium Statistical Mechanics
If the particles are distinguishable, the number of microstates accessible to the system is equal to
the number of ways choosing Nε particles from N particles and is given by
N! N!
Ω = N CNε = = (2.1.27)
Nε ! (N − Nε )! E E
! N− !
ε ε
Assuming that N, Nε and (N − Nε ) are all large numbers compared to 1, we use Stirling’s approxi-
mation for the factorials of large numbers and then the entropy is
S = k ln Ω = k [N ln N − Nε ln Nε − (N − Nε ) ln (N − Nε )]
E E E E
=k − N ln 1 − − ln (2.1.28)
ε Nε ε Nε
The temperature of the system is given by
1 ∂S k Nε
= = ln −1 (2.1.29)
T ∂E N,V ε E
from which the energy E of the system is given by
Nε
E= ε (2.1.30)
1 + e kT
E N
∴ Nε = = ε (2.1.31)
ε 1 + e kT
Now for T → 0, i.e., kTε 1, Nε → 0, that is, all particles are frozen in the ground state. As T
increases, both E and Nε increase and when T → ∞, i.e., kTε 1, Nε ; N2 , that is, half the particles are
in the ground state and half in the excited state and the energy of the system attains its maximum
= Nε
value of E ∼ 2 . The specific heat at constant volume is given by
ε
∂E ε 2 e kT
CV = = Nk ε (2.1.32)
∂T V kT (1 + e kT )2
CV
kT
ε
0.42
Fig. 2.3. The specific heat at constant volume of a two-state system, showing Schottky hump.
Elements of Ensemble Theory • 69
The specific heat CV is zero at both very low and very high temperatures and is maximum at T ∼=
0.42 kε . The hump, observed in the CV vs T plot, is called Schottky anomaly. Such a peak, when
observed experimentally, is therefore an indication of a gap in its energy states.
N,V,T N,V,T
N,V,T N,V,T
A A′
Fig. 2.5. A small system ‘A’ in thermal interaction with a heat reservoir ‘A ’.
We assume weak interaction between A and A , so that their energies are additive. The energy
of A is, of course, not fixed. Only the total energy of the combined system, A0 = A + A (which is
isolated) has a constant value in some range between E 0 and E 0 + δE:
Er + E = E 0 = constant (2.2.1)
70 • An Introduction to Equilibrium Statistical Mechanics
where E denotes the energy of the reservoir A . Thus, when A has an energy Er , the reservoir A
must have an energy E = E 0 − Er .
Let Ω (E ) denote the number of accessible microstates of A , when it has an energy in the
range near E . If the system A is in the particular
state r, the number of states accessible to the
combined system A0 is 1 × Ω (E ) = Ω E 0 − Er .
Then the probability of finding the system A in this state is
Ω (E ) Ω E 0 −Er
Pr = Pr (Er ) = 0 = ,
Ωtotal Ω0total
where Ω0total denotes the total number of states accessible to the combined system A0 .
But Ω0total = constant = c0 (say).
Pr = c0 Ω E 0 − Er
∴ (2.2.2)
0
⇒ ln Pr = ln c + ln Ω (E0 − Er )
(2.2.3)
Since A is very small compared to A , Er E 0 . Expanding ln Ω about E = E 0 ,
∂ ln Ω 0
0 0
ln Ω E − Er = ln Ω E −
Er − . . .
∂E E
Since A acts as a reservoir, Er E 0 and so we can neglect higher order terms.
0 0 ∂ ln Ω
∴ ln Ω E − Er = ln Ω E −
∼
Er (2.2.4)
∂E E 0
∂ ln Ω
The derivative is evaluated at the fixed energy E = E0 and is thus a constant, indepen-
∂E E 0
1
dent of the energy E and ∂ ln Ω =β= , where T is the temperature of the reservoir.
∂E E
0 kT
The equation (2.2.4) then becomes
ln Ω E 0 − Er = ln Ω E 0 − βEr
Ω E 0 − Er = Ω E 0 e−βEr
⇒
Since Ω E 0 is just a constant, independent of r, equation (2.2.2) becomes
Pr = ce−βEr (2.2.5)
where c is a constant, independent of r.
The probability (2.2.5) is a very general result and is of fundamental importance in statisti-
cal mechanics. The exponential factor e−βEr is called the ‘Boltzmann factor’; the corresponding
probability distribution (2.2.5) is known as the ‘canonical distribution’.
Now according to the normalization conditions for probabilities,
∑ Pr = 1 ⇒ c−1 = ∑ e−βE r
r r
Elements of Ensemble Theory • 71
e−βEr e−βEr
∴ Pr = = (2.2.6)
∑ e−βEr Z
r
where
Z = ∑ e−βEr (2.2.6a)
r
‘Z’ is called the canonical partition function. (The letter Z is used because the German name is
Zustandsumme.) It is the sum over all accessible microstates of the system A.
It should be noted that the expression (2.2.6a) is correct, if the energy levels are discrete and
non-degenerate. In the general case, an energy level Er consists of a group of states gr in number all
of the same energy Er and of equal probability. Thus in summing over all accessible microstates, it
is necessary to repeat equal terms (involving Er ) gr times. Instead, it is simpler to use a modified
form of the partition function
Z = ∑ gr e−βEr (2.2.6b)
r
where the label r goes over all energy levels. If the level is non-degenerate, then gr = 1. Equation
(2.2.6b) is the general form of the partition function.
Accordingly,
gr e−βEr
Pr = (2.2.6c)
∑ gr e−βEr
r
Once the probability distribution is known, various mean values can be computed. For example, let
x be any quantity assuming the value xr in state r of the system A. Then the mean value of x is
∑ xr Pr ∑ xr e−βEr
x= r
= ∑ xr Pr = r
(2.2.7)
∑ Pr r ∑ e−βEr
r r
∴ Z = ξN (2.2.19)
where
∞
V βp2
ξ= 3 e− 2m d 3 p (2.2.20)
h0 −∞
Now
∞ ∞
p2x + p2y + p2z
2 β
− βp 3 − 2m
e 2m d p= e d px d py d pz
−∞ −∞
∞ ∞ ∞
3
βp2x βp2y βp2
z 2πm
= e
− 2m d px e − 2m d py e− 2m d pz = ,
β
−∞ −∞ −∞
∞
−αx2 π
where we use the standard integral e dx =
α
−∞
32
V 2πm V 3
∴ ξ= 3 = 3
(2πmkT ) 2 (2.2.21)
h0 β h0
Using (2.2.21) in (2.2.19)
VN 3N
Z (N,V, T ) = 3N
(2πmkT ) 2 (2.2.22)
h0
This is the canonical partition function of a classical ideal gas.
What is going to happen, if we would use the same formula, i.e., equation (2.2.19) to calculate
the partition function for a system of N indistinguishable, identical, non-interacting particles. To
illustrate that this is wrong, we consider two indistinguishable particles:
Z (N = 2,V, T ) = ξ2 = ∑ e−βεr1 × ∑ e−βεr2
r1 r2
= ∑ e −2βεr
+ ∑ ∑ e−β(ε r1 +εr2 )
r=r1 =r2 r1 =r2
particles both particles in different
in the same state states
However, the particles in the second term are counted twice, but since the particles are indistin-
guishable this does not generate a new state, because
1
= ∑
r=r1 =r2 =r3
e−3βεr + · · ·
some particles
+
3! ∑ e−β(ε
r1 =r2 =r3
r1 +εr2 +εr3 )
particles in same state in same state all particles in
different states
UM
r0
0 r
Therefore,
N
3N2 V0 V
1 2πm e−UM β d 3 r + e−UM β d 3 r
Z (N,V, T ) = (2.2.25)
h3N
0 β
0 V0
where V0 is the volume corresponding to r = r0 . Now in the range 0 to V0 ,UM → ∞ and in the range
V0 to V , UM = −U. Therefore, equation (2.2.25) becomes
3N V N
1 2πm 2 Uβ 3
Z (N,V, T ) = 3N e d r
h0 β V0
1 3N
U N
⇒ Z (N,V, T ) = 3N (2πmkT ) 2 (V −V0 ) e kT (2.2.26)
h0
This is the canonical partition function for a classical non-ideal gas in the mean field approximation.
This topic will be further discussed in Chapter 7.
1. Entropy
According to Boltzmann’s entropy relation, the entropy S is related to the thermodynamic proba-
bility Ω through the relation
S = k ln Ω (2.2.27)
For a classical system, the total number of molecules is
N = ∑ ni (2.2.28)
i
and
gni i
Ω = N! ∏ (2.2.29)
i ni !
Taking logarithm and then using Stirling’s approximation
ln Ω = N ln N − N + ∑ (ni ln gi − ni ln ni + ni ) (2.2.30)
This is known as the virial theorem, because ∑ qi ṗi is known as virial in classical mechanics.
i
P ROBLEM 2.5: Find the classical canonical partition function of a one-dimensional linear har-
monic oscillator. Also find its average energy E.
S OLUTION
Since this is a one-dimensional case, the classical canonical partition function is
d p dx
Z= e−βE .
h0
p2 1
where E is the total energy of the oscillator and is given by E = + mω2 x2 .
2m 2
2
∞ − βp 1 2 2
1
∞
− mω x
∴ Z= e 2m d p e 2 dx .
h0 −∞ −∞
∞
−αx2 π
Using the standard integral e dx = ,
−∞ α
1 2πm 2π 2π 2πkT
Z= = = .
h0 β mβω2 h0 βω h0 ω
This is the canonical partition function.
The average energy of the oscillator is
∂ ln Z 1
E = − = = kT.
∂β β
P ROBLEM 2.6: Find the canonical partition function of a relativistic classical ideal gas with
energy-momentum relationship E = pc.
S OLUTION
For a classical ideal gas, the canonical partition function is
d 3N pd 3N x
Z= e−βE = ξN ,
h3N
0
where ξ is the single particle partition function and is given by
3 pd 3 x ∞ π 2π
4πV
∞
−βpc d V
−βpc 2
ξ= e = 3 d pe p dθ sin θ dφ = 3 d pe−βpc p2
h30 h0 0 0 0 h0 0
3
4πV 1 4πV
∞
2 kT
= 3 d (βpc) e −βpc
(βpc) = 3 Γ (3)
h0 (βc)3 0 h0 c
Elements of Ensemble Theory • 85
P ROBLEM 2.10: Show that the partition function of an extreme relativistic gas consisting of 3N
indistinguishable particles, with energy-momentum relationship ε = pc, c being the speed of light
1
kT 3N
and moving in one dimension is given by Z (N, L, T ) = (3N)! 2L hc , where L being the ‘length’
of the space available.
S OLUTION
The single particle canonical partition function is
1
∞
L kT
ξ= e −βε
dqd p = e −βpc
dp = L (i)
h h 0 hc
Therefore, the canonical partition function of the system of N non-interacting indistinguishable
particles is
3N
ξ3N 1 kT
Z (N, L, T ) = = L (ii)
(3N)! (3N)! hc
P ROBLEM 2.11: Consider a system in thermal equilibrium at temperature T − its two states with
energy difference 4.8 × 10−14 erg occur with relative probability e2 erg / deg. Calculate the temper-
ature. Boltzmann’s constant k = 1.38 × 10−16 erg / deg.
S OLUTION
Let P1 and P2 be respectively the probabilities of the state of energies E1 and E2 . Then
E1 E2
P1 = ce− kT and P2 = ce− kT , where c is a constant of proportionality.
P1 (E2 −E1 ) (E2 − E1 ) 4.8 × 10−14
∴ = e kT ⇒ T = = = 173.913 K
P2 P1 1.38 × 10−16 × ln e2
k ln
P2
86 • An Introduction to Equilibrium Statistical Mechanics
P ROBLEM 2.12: The first vibrational energy of a diatomic molecule is 600 cm−1 above the ground
state. Calculate the relative population of molecules in these two levels at T = 400 K.
S OLUTION
Let N0 and Eo be respectively the number of molecules and the energy of the ground state and
N1 and E1 be the corresponding values for the first excited state.
(E1 −E0 )
∴ N1 = N0 e− kT (i)
Now
hλ
(E1 − E0 ) = hv = = hcv = 6.6 × 10−27 × 3 × 1010 × 600 = 1.2 × 10−13 erg
c
N1
(E1 −E0 ) 1.2×10−13
−
∴ = e− kT = e 1.38×10−16 ×400 ∼ = 0.1 ⇒ N1 = 0.1 × N0
N0
P ROBLEM 2.13: Consider a system which can take only three different energy states E1 = 0,
E2 = 1.38 × 10−14 ergs, E3 = 2.76 × 10−14 ergs. Find the probability that at temperature 100 K the
system may be (a) in one of the microstates of energy E3 , (b) in the ground state E1 .
S OLUTION
Let P1 , P2 and P3 be respectively the probabilities with which the microstates can occur in three
energy values E1 , E2 and E3 .
E1
Then P1 = ce− kT = c
E2 1.38×10−14
−
P2 = ce− kT = ce 1.38×10−16 ×100 = ce−1
and
E3 2.76×10−14
−
P3 = ce− kT = ce 1.38×10−16 ×100 = ce−2
where c is a constant of proportionality.
Again these three states can occur in 2, 5 and 4 different ways respectively. Then P1 = 2c,
P2 = 5ce−1 and P3 = 4ce−2 . But
3
∑ Pi = 1 ⇒ (P1 + P2 + P3 ) = 1 ⇒
2c + 5ce−1 + 4ce−2 = 1
i=1
1 1
⇒c= ⇒c=
2 + 1.84 + 0.54 4.38
Since e = 2.72.
(a) The probability for the system to be in one of the microstates of energy E3 is
4
P3 = 4ce−2 = = 0.12.
4.38 × (2.72)2
Elements of Ensemble Theory • 89
Again
∂ ln Z
E =− (vi)
∂β
Comparing (v) and (vi)
Fβ = − ln Z
(vii)
⇒ F = −kT ln Z
Finally, using (vii) in (iii)
E F 1 1 S
∑ Pr ln Pr = − kT + kT =
kT
(F − E) =
kT
(−T S) = −
k
r
⇒ S = −k ∑ Pr ln Pr [Proved] (viii)
r
In case of microcanonical ensemble, we have a group of Ω states (which are equally likely to
occur) for each member. The value of Pr is then Ω1 for each of these states and 0 for all others.
Consequently,
Ω
1 1
S = −k ∑ ln = k ln Ω [Proved] (ix)
r=1 Ω Ω
P ROBLEM 2.17: The states of a system are (i) a group of g1 equally likely states with a common
energy ε1 , and (ii) a group of g2 equally likely states with a common energy ε2 = ε1 .
P1 P2
(a) Show that the entropy of the system is given by S = −k P1 ln + P2 ln
g 1 g2
g2 −x x ε2 − ε1
(b) Also show that S = k ln g1 + ln 1 + e + , where x = ,
g1 1 + (g1 /g2 )ex kT
assumed positive.
(c) Check that at T → 0, S → k ln g1 .
S OLUTION
(a) The canonical partition function is
Z = ∑ gr e−βEr (i)
r
and
gr e−βEr
Pr = (ii)
Z
Pr e−βEr Pr
∴ = ⇒ ln = − ln Z − βEr
g Z gr
r
Pr
⇒ ∑ Pr ln gr = − ln Z − βE (iii)
r
90 • An Introduction to Equilibrium Statistical Mechanics
If the temperature T is so high that the thermal energy kT is large compared to the separation hν
hν
between energy levels, i.e., βhν = kT 1,
1 1 1 kT kT
E ≈ hν + hν = hν + ≈ hν × = kT −→ classical result.
2 kT 2 hν hν
P ROBLEM 2.19: A classical point particle is moving in a 3-dimensional harmonic oscillator po-
tential well, V (r) = 1/2Kr2 = 1/2K(x2 + y2 + z2 ) at absolute temperature T . Obtain a formula for
the probability that the particle is in between r and r + dr from the centre of attraction. Also obtain
a formula for the mean square distance of the particle from the centre of attraction and check your
result by comparison with the equipartition principle.
S OLUTION
The probability that the particle will be in phase space volume element d 3 rd 3 p
βp2 1 2 βp2 1 2
e− 2m −β 2 Kr d 3 pd 3 r e− 2m −β 2 Kr d 3 pr2 sin θ dθ dφ dr
= βp2 1
= βp2 1
2 2
e− 2m −β 2 Kr d 3 pd 3 r e− 2m −β 2 Kr d 3 pr2 sin θ dθ dφ dr
Therefore, the probability that the particle will lie in the range r and r + dr is
− βKr2 2 βp2
e 2 r sin θ dθ dφ dr e− 2m d 3 p
θ φ p
P(r)dr = βKr2 βp2
e− 2 r2 sin θ dθ dφ dr e− 2m d 3 p
θ φ p r
βKr2 βKr2
e− 2 r2 dr
e− 2 r2 dr
= ∞ = 23
βKr2
e− 2 r2 dr 1√ 2
4 π Kβ
0
32
4 Kβ 1 2
∴ P(r)dr = √ e− 2 βKr r2 dr (i)
π 2
1. If the magnetic moment of each atom can point either parallel or anti-parallel to the external
→
− →
−
field H , what is the mean magnetic moment µH (in the direction of H , i.e., z-direction) of
such atom ?
2. Calculate the partition function of one atom, the mean magnetic moment along the field
direction and hence the susceptibility of the system, first by treating the system classically
and then quantum mechanically.
S OLUTION
Each atom can be in two possible states: the state (+) where its spin points up (i.e., parallel to
→
− →
−
H ) and the state (–) where its spin points down (i.e. anti-parallel to H ). In the (+) state, →
−µ is
→
−
parallel to H and so µH = µ. The corresponding magnetic energy of the atom is E+ = −µH. The
probability of finding the atom in this state is
P+ = Ce−βE+ = CeβµH (i)
1 →
−
where C is a constant of proportionality and β = kT .
Similarly, in the (–) state, µ is anti-parallel to
→
−
H and so µH = −µ. The corresponding magnetic energy of the atom is E− = +µH. The probability
of finding the atom in this state is
P− = Ce−βE− = Ce−βµH (ii)
∴ The mean value of µH is
eβµH − e−βµH
P+ (µ) + P− (−µ) µH
µH = = µ βµH = µtanh (βµH) = µtanh .
P+ + P− e + e−βµH kT
When the system is treated classically
The energy of a magnetic dipole in the presence of the external magnetic field is
→
−
E = −→−µ · H = −µH cos θ (iii)
The canonical partition function of the system is
Z = ξN (iv)
where ξ is the single particle canonical partition function and is given by
2π π
sinh (βµH)
ξ = ∑ exp (βµH cos θ) = eβµH cos θ sin θ dθ dφ = 4π (v)
θ φ=0 θ=0 βµH
sinh (βµH) N
∴ Z = 4π (vi)
βµH
The average magnetic moment along the field direction is given by
∑ µcos θ exp (βµH cos θ)
θ 1 ∂ 1
µH = µcos θ = = ln ξ = µ coth (βµH) −
∑ exp (βµH cos θ) β ∂H βµH
θ
= µL(βµH) (vii)
94 • An Introduction to Equilibrium Statistical Mechanics
gµB H
Let x = βgµB H = −→ a dimensionless parameter. Therefore,
kT
J
Z= ∑ exm
m=−J
= e−xJ + e−x(J−1) + · · · + exJ
1 1
e−xJ − ex(J+1) e−x(J+ 2 ) − ex(J+ 2 ) sinh J + 12 x
= = x x = x
1 − ex e− 2 − e 2 sinh
2
sinh J + 12 βg µB H
∴ Z= (xvi)
sinh 21 βg µB H
Now the probability of finding the atom in a state labelled m is given by
Pm = Ce−βEm (xvii)
where C is a constant of proportionality, determined from the normalization condition, ∑ Pm = 1.
m
Therefore, the mean value of the magnetic moment along the field direction (i.e. z-direction) is
J
∑ µH Pm ∑ µH eβgµB Hm
m m=−J
µH = = (xviii)
∑ Pm J
m ∑ eβgµB Hm
m=−J
→
−
Since the field H is along the z-direction, µH = µz = gµB Jz = gµB m, from (iv).
J J
∑ gµB meβgµB Hm ∑ (βgµB m) eβgµB mH
m=−J 1 m=−J 1 1 ∂Z 1 ∂ ln Z
∴ µH = = = =
J β J β Z ∂H β ∂H
∑ eβgµB Hm ∑ eβgµB mH
m=−J m=−J
1 ∂ ln Z ∂x ∂ ln Z ∂ 1 x
= = gµB = gµB ln sinh J + x − ln sinh
β ∂x ∂H ∂x ∂x 2 2
J + 21 cosh J + 12 x 21 cosh 2x
= gµB −
sinh J + 12 x sinh 2x
P ROBLEM 2.21: Show that entropy increases when two ideal gases at the same temperature and
pressure diffuse into each other. Discuss Gibbs paradox in this connection.
S OLUTION
According to Boltzmann’s relation, the entropy S of a system is related to the thermodynamic
probability Ω and is given by
S = k ln Ω (i)
From Maxwell-Boltzmann statistical count
gni i
Ω = N! ∏
i ni !
⇒ ln Ω = ln N! + ∑ (ni ln gi − ln ni !)
i
= ln N! + ∑ (ni ln gi − ni ln ni + ni )
i
= ln N! + ∑ ni (ln gi − ln ni + 1)
i
gi
Ω = ln N! + ∑ ni ln + 1 (ii)
i ni
Again for most probable state, Ω or ln Ω is maximum and then we have
Ei
ni = e−α gi e− kT (iii)
[see equation (3.1.7)].
ni gi Ei
E
− α+ kTi
∴ =e ⇒ ln = α+ (iv)
gi ni kT
Using (iv) in (ii)
Ei
ln Ω = ln N! + ∑ ni +α+1
i kT
1
= ln N! +
kT ∑ ni Ei + α ∑ ni + ∑ ni
i i i
98 • An Introduction to Equilibrium Statistical Mechanics
3
3 2πmkT 2
SB = Nk + Nk ln V .
2 h20
Hence, the total entropy is simply
3
2πmkT 2
SA + SB = 3Nk + 2Nk ln V (ix)
h20
We now remove the partition very slowly so that the gases diffuse into each other. Therefore, the
total number of molecules 2N of the two gases occupy a volume 2V . Since there is no change in
pressure and temperature, the entropy of the system after the partition is removed is
3
2πmkT 2
S = 3Nk + 2Nk ln (2V ) (x)
h20
Now, according to the additive property of entropy
S = SA + SB (xi)
But from (ix) and (x), we see that
S − [SA + SB ] = 2Nk ln 2 > 0. (xii)
which is not equal to zero as required by (xi).
This paradox was first discussed by Gibbs and is commonly referred to as the ‘Gibbs paradox’.
The root of the difficulty embodied in the Gibbs paradox is that we treated the gas molecules as
distinguishable, as though interchanging the positions of two molecules would lead to a physically
distinct state of the gas. This is not so. Indeed, if we treated the gas by quantum mechanics,
the molecules would have to be regarded as indistinguishable (see Chapter 3). In this case, N!
permutations of the molecules among themselves do not lead to physically distinct situations, so
that we should subtract the term k ln N! from (vi), so that
3
5 V 2πmkT 2
S = Nk + Nk ln (xiii)
2 N h2
This has been verified experimentally at high temperatures and is known as Sackur-Tetrode
equation.
Equation (xii) can be written as
3
V 2πmkT 2 5
S = Nk ln e2 (xiv)
N h2
Using this expression, it can be easily verified that SA + SB − S = 0 and hence the paradox is re-
moved.
P ROBLEM 2.22: Imagine that a system R1 has probability Pr1 of being found in a state of r and
a system R2 has probability Ps2 of being found in a state s. Then one has S1 = −k ∑ Pr1 ln Pr1 and
r
Elements of Ensemble Theory • 99
S2 = −k ∑ Ps2 ln Ps2 . Each state of the composite system R (= R1 + R2 ) can be labelled by a pair of
r
numbers r, s. Let the probability of R being found in this state be denoted by Prs . Then the entropy
is defined by S = −k ∑ ∑ Prs ln Prs . If R1 and R2 are weakly interacting so that they are statistically
r s
independent, then Prs = Pr1 Ps2 . Show that under these circumstances, the entropy is simply additive,
i.e., S = S1 + S2 .
S OLUTION
We have
S = −k ∑ ∑ Prs ln Prs = −k ∑ ∑ Pr1 Ps2 ln Pr1 Ps2 Prs = Pr1 Ps2
∵
r s r s
= ∑ Ps1 −k ∑ Pr1 ln Pr1 + ∑ Pr1 −k ∑ Ps2 ln Ps2
s r r s
= −k ∑ Pr ln Pr + −k ∑ Ps ln Ps
1 1 2 2
∵ ∑ Pr1 =1=∑ Ps2
r s r s
∴ S = S1 + S2 (Proved)
P ROBLEM 2.23: Suppose that a system R1 has probability Pr1 of being found in a state r and a
system R2 has probability Ps2 of being found in a state s. The entropy of R1 is S1 = −k ∑ Pr1 ln Pr1 and
r
that of R2 is S2 = −k ∑ Ps2 ln Ps2 . Each state of the composite system R(= R1 + R2 ) is then labelled by
s
the pair numbers r, s. Let the probability of R being found in this state be Prs ; its entropy is defined
by s = −k ∑ ∑ Prs ln Prs . Assume that R1 and R2 are not weakly interacting, so that Prs = Pr1 Ps2 . Of
r s
course Pr1 = ∑ Prs and Ps2 = ∑ Prs . Moreover, all the probabilities are properly normalized so that
r r
∑ Pr1 = 1, ∑ Ps2 = 1, ∑ ∑ Prs = 1. Show that
r r r s 1 2
(a) S − (s1 + S2 ) = k ∑ Prs ln PPr rsPs , and (b) S ≤ S1 + S2 .
r,s
S OLUTION
(a) We have
Pr1 = ∑ Prs (i)
∴ S ≤ (S1 + S2 ) (x)
S OLUTION
The restoring force is given by
F = −Kx3 (i)
where K is the constant of proportionality.
Hence, the potential energy of a particle is
1
V = Kx3 (ii)
4
Therefore, the total energy of a particle is
p2 1
ε= + kx3 (iii)
2m 4
102 • An Introduction to Equilibrium Statistical Mechanics
L 0 Y
L
Now
L
mgz kT mgL
e− kT dz = 1 − e− kT
mg
0
and
L 2
− mgz kT mgL
kT L mgL
ze kT dz = 1 − e− kT − e− kT
mg mg
0
mgL
kT L mgL
2
kT
1 − e− kT − e− kT mgL
mg mg e− kT
∴ mgz = mg × = kT − mgL
kT − mgL
mgL
1 − e− kT
1−e kT
mg
mgL −1
⇒ mgz = kT + mgL 1 − e kT (Ans.)
(ii)
P ROBLEM 2.26: Calculate the rotational partition function for heteronuclear and homonuclear
molecules separately.
Elements of Ensemble Theory • 103
S OLUTION
The rotation of a diatomic molecule is specified by the angles (θ, φ) and the corresponding
momenta (Pθ , Pφ ) and its kinetic energy assumes the form
Pθ2 Pφ2
εrot = + .
2I 2I sin2 θ
For heteronuclear diatomic molecule:
1
ξrot
heteronuclear = e−3/kT ld pθ d pφ dθdφ
h2
2π ∞ π ∞ p2
1 p20
− φ
= dφ e− 2IkT
d pθ dθ e 2IkT sin2 θ d pφ
h2
φ=0 pφ=0 θ=0 pφ=0
π
1 1√ 1√
= × 2π × 2πIkT × dθ 2πIkT sin θ
h2 2 2
θ=0
2π2 IkT
∴ ξrot
heteronuclear = (i)
h2
For homonuclear diatomic molecule:
For homonuclear diatomic molecule, φ ranges from 0 to π (see Section 4.12.2).
π2 IkT
∴ ξrot
homonuclear =
h2
P ROBLEM 2.27: Study the thermodynamics of a system of N non-interacting diatomic molecules,
each having an electric dipole moment µ, placed in an external electric field E. Assume that (i) the
system is classical and (ii) | µE | kT , where T is the absolute temperature of the system.
S OLUTION
The energy of a diatomic molecule in the presence of the electric field E is given by
P2 Pθ2 Pφ2
ε= + + − µE cos θ (i)
2m 2I 2I sin2 θ
where I is the moment of inertia of the molecule.
Now the single particle canonical partition function is given by
1
ξ= e−βε dr dθ dφ d p d pθ d pφ (ii)
h3
1
V 3
2π ∞
1 2
2
− βP − βPθ
ξ= dr × dφ × d pe ∞d pθ e
2m × 2m
⇒
h3
r=0 φ=0 P=0 Pθ =0
104 • An Introduction to Equilibrium Statistical Mechanics
∞ ∞ βp2φ
dθeβµE cos θ
−
× d pφe 2I sin2 θ
θ=0 Pφ=0
∞
1 1 1 2mπ 1 2Iπ 1 2Iπ
= 3 V 3 × 2π × × × dθeβµE cos θ · sin θ
h 2 β 2 β 2 β
θ=0
√ 5 π
2mIπ 21
= 3
· 3 · dθ sin θeβµE cos θ
2h β2 0
√ 5
2mIπ 2 1 1
βµE βµE
= · · · e − e
2h3 β 2 βµE
3
√ 5
2mIπ 2 1
∴ ξ= sinh (βµE) (iii)
µEh3 β 52
Therefore, the canonical partition function of the system is
√ 5
N
N
2mIπ 2 sinh βµE
Z = ξN = (iv)
µEH 3 β 52
√ 5
2mIπ 2 5
⇒ ln Z = N ln + N ln sinh βµE − ln β (v)
µEh3 2
1 µE 2
2 µE 1 1
cosec h ∼
= − + (ix)
kT µE 3 15 kT
kT
106 • An Introduction to Equilibrium Statistical Mechanics
and
∞ 1
∞
) = e− 2u
∑ 3 = ∑ n2 e−u(n+ 2
∑ n2 e−ν
n=0 n=0
u
= e− 2 e−u + 22 e−2u + 32 e−3u + · · ·
u
= e− 2 · e−u 1 + 22 e−u + 32 e−2u + · · ·
3u (1 + e−u )
= e− 2 · (iv)
(1 − e−u )3
Substituting (ii), (iii) and (iv) in (i)
u
e− 2 e−u (1 + e−u ) e−u 1
ξ= 1 + xu + +
1 − e−u (1 − e−u )2 (1 − e−u ) 4
u (v)
e2 2e−u 1
= 1 + xu +
1 − e−u (1 − e−u ) 4
Therefore, the canonical partition function for the system is given by
Z = ξN
u 2e−u 1
⇒ ln Z = N ln ξ = N − − ln 1 − e−u + ln 1 − xu + (vi)
2 (1 − e−u )2 4
Then the internal energy of the system is
∂ ln Z ∂ ln Z ∂ ln Z
E =− = kT 2 = −hv
∂β ∂T ∂u
2
2
e−u ue−u 2u(e−u )
x 2 1−e−u − −
1 e−u (1−e−u )2 (1−e−u )3 + 41
= −Nhv − −
2 1 − e−u +
2e−u
1 + xu −u 2 1
(1−e ) +4
= E0 + Ecorrection (vii)
where
1 Nhv 1 Nhv
E0 = Nhv + u = Nhv + hv (viii)
2 e −1 2 e kT − 1
The corresponding specific heat at constant volume is
2 hv
∂E0 hv e kT
(CV )0 = = Nk hv 2 (ix)
∂T V kT
e kT 1
2 2
e−u ue−u 2u(e−u )
2 1−e−u − (1−e−u )2 − (1−e−u )3 + 41
Ecorrection = −xNhv (x)
2e−u 1
1 − xu
−u 2
+4
(1−e )
Elements of Ensemble Theory • 107
P ROBLEM 2.29: Calculate the canonical partition function and hence the specific heat at constant
volume for a classical system of N non-interacting diatomic molecules enclosed in a box of volume
V at temperature T .
S OLUTION
The energy of a diatomic molecule can be written as
1 2 1
E (p1 ,p2 ,r1 ,r2 ) = p1 + p22 + K |r1 −r2 |2 (i)
2m 2
where p1 ,p2 ,r1 and r2 are the momenta and position coordinates of the two atoms in a diatomic
molecule.
r1 + r2
Let R = and ρ =r1 −r2 .
2
Then the Jacobian is
∂r1 ∂r2
∂R ∂R 1 1
J = ∂r ∂r = 1 1 = −1 ⇒| J |= 1 (ii)
1
∂ρ ∂ρ 2
2 − 2
∴ d 3 r1 d 3 r2 = d 3 Rd 3 ρ (iii)
Now the single particle canonical partition function is
ξ= e−βH d 3 r1 d 3 r2 d 3 p1 d 3 p2 = e−βH d 3 r1 d 3 r2 d 3 Rd 3 ρ
−βp21 3 −βp22 3 3 −βkp22 3
= e d p1 e d p2 d R e d ρ
3 3 32
2πm 2 2πm 2 2π
= × ×V ×
β β βK
3
8π3 m2 2 1
=V × 9
K β2
108 • An Introduction to Equilibrium Statistical Mechanics
P ROBLEM 2.30: In Problem 2.31, calculate the mean square molecular diameter, |r1 −r2 |2 .
S OLUTION
The mean square molecular diameter is
|r1 −r2 |2 e−βE d 3 r1 d 3 r2 d 3 p1 d 3 p2
2
|r1 −r2 | =
e−βE d 3 r1 d 3 r2 d 3 p1 d 3 p2
2 −β[ 1 ( p2 +p2 )+ 1 Kρ2 ] 3
ρ e 2m 1 2 2 d r1 d 3 r2 d 3 p1 d 3 p2
= 1 2 2 1 2
e−β[ 2m ( p1 +p2 )+ 2 Kρ ] d 3 r1 d 3 r2 d 3 p1 d 3 p2
∞ 4 − 1 βKρ2 ∞ 3 e−x dx
2 − 1 βKρ2 3 ρ e 2 dρ x2
2kT Γ 52
ρ e 2 d ρ 0 2kT 0 3kT
= 1 2 = ∞ = ∞ = 3 =
e 2
− βKρ d3ρ 1 2 K 3 −x K Γ 2 K
ρ2 e− 2 βKρ dρ x 2 e dx
0 0
2
|r1 −r2 | ∞ T.
S OLUTION
The energy of a molecule is
1 2
p1 + p22 + ε|r12 − r0 |
E (p1 ,p2 ,r1 ,r2 ) = (i)
2m
r1 +r2
Let R = and ρ =r1 −r2 .
2
Then the Jacobian is
∂r1 ∂r2
∂R ∂R
1 1
J= = −1 ⇒| J |= 1 (ii)
=
∂r1 ∂r2 21 − 21
∂ρ ∂ρ
∴ d 3 r1 d 3 r2 = d 3 Rd 3 ρ (iii)
The single particle canonical partition function is
−βH 3
ξ= e d r1 d r2 d p1 d p2 = e−βH d 3 r1 d 3 r2 d 3 Rd 3 ρ
3 3 3
βp2 βp2
− 2m1 3 − 2m2 3 3 −βε|ρ−r0 | 3
= e d p1 × e d p2 × d R × ρ d ρ (iv)
∞
3 32
2m 2 1 3 2m 1 3 4π
= 4π τ × 4π τ ×V × y2 e−|y−x| dy
β 2 2 β 2 2 (βε)3
0
Now, when the cylinder is rotated, it becomes energetically favourable for the molecules to move
towards the edge of the cylinder, so we expect the density to increase with radius r.
In a frame of reference rotating with a gas, a molecule of mass m at a distance r from the axis
experiences a centrifugal force
Fc = mω2 r (ii)
away from the axis.
∴ The centrifugal potential energy is
1
Vc = − mω2 r2 (iii)
2
From
the Boltzmann distribution law, we expect the probability of a molecule to be proportional to
Vc
exp − kT and so we can write
mω2 r
σ(r) = A exp (iv)
2kT
where A is a constant of proportionality.
Now the total mass of the gas within the cylinder is given by
a a 2 2
mw r
M = 2πL σ(r)rdr = 2πLA r exp dr (v)
2kT
0 0
The integral can be easily evaluated by noting that
mω2 r2 mω2 r2 mω2
d
exp = exp (vi)
dr 2kT 2kT r2
mω2 a2
2πLAkT
∴ M= exp −1 (vii)
mω2 2kT
Equating (i) and (vii), we have the expression for A:
σ0 mω2 a2 1
A= (viii)
2kT mω2 a2
exp −1
2kT
Using (viii), we get the expression for σ(r) as
mω2 a2
exp
σ0 mω2 a2 2kT
σ(r) = 2 2 2 (ix)
2kT mω exp mω
2kT
a
−1 a2
exp 2kT −1
r mω2 a2
Let R = and Ω = .
a 2kT
σ(R) exp(ΩR2 )
∴ =Ω (x)
σ0 [exp(Ω) − 1]
112 • An Introduction to Equilibrium Statistical Mechanics
P ROBLEM 2.33: Consider an ideal gas at absolute temperature T in a uniform gravitational field
described by acceleration g. By writing the condition of hydrostatic equilibrium for a slice of the
gas located between heights z and z + dz, derive an expression for n(z), the number of molecules per
cm3 at height z. Also show that this result is identical with that derived from statistical mechanics.
S OLUTION
Let us consider a slice between heights z and z + dz. Let the pressure at z be p and that at z + dz
be p + d p.
z + dz, p + dp
z, p
On integration
mg
ln n = − z + constant (iv)
kT
Let at z = 0, n = n(0).
mgz
∴ n = n(0)e− kT (v)
Using ideal gas law: p = nkT, p(0) = n(0)kT
mgz
p = p(0)e− kT (vi)
Now the probability P (r,p) d 3 rd 3 p that the molecule has position lying in the range betweenr and
r + dr and momentum in the range between p and p + dp is given by
d 3 rd 3 p −β 2m
p2
3 3 +mgz
P (r,p) d rd p ∝ 3
e (vii)
h
Therefore, the probability P (z) dz that a molecule is located at a height between z and z + dz is
∞ π 2π
1
P2
P (z) dz ∝ dx dy × e −β 2m
P2 sin θ d p dθ dφ × e−βmgz dz (viii)
×
h3
(x,y) P=0 θ=0 φ=0
But (x,y) dx dt = A, cross-sectional area of the slice.
3/
2
A 2πm mg
∴ P (z) dz = constant × 3 e− kT z (ix)
h β
Let at z = 0, P(z)|z=0 = P(0).
A 2πm
∴ P(0) = constant × 3 (x)
h β
− mg z
⇒ P(z) = P(0)e kT (xi)
Thus, the probability of finding a molecule at height z decreases exponentially with height.
Now mgz
n(z) dz ∝ p(z) dz = constant × P(0)e− kT dz (xii)
At z = 0, n = n(0) ⇒ constant × P(0) = n(0)
mgz
∴ n(z) = n(0)e− kT (xiii)
which is same as (v).
P ROBLEM 2.34: Consider a system consists of N weakly interacting particles, each of which can
be in either of two states with respective energies ε and ε2 (ε2 > ε1 ). Calculate explicitly the mean
energy E and heat capacity CV of this system.
Elements of Ensemble Theory • 115
S OLUTION
Consider a pair of one +ve and one –ve particle.
+ +
a
+ +
a
a
2
+ +
+ +
The adjoining figure shows that there are 4 possible states of this pair. Since the potential energy
of a charged particle of charge in an external electric field is V (x) = −eεx, the average separation
between a +ve ion and a –ve ion is given by
a a
2 · 2a eβ 2 εe − 2 · 2a e−β 2 εe a βeεa
x = a a = tanh (i)
2e β 2 εe + 2e −β 2 εe 2 2
If there are N number of pairs per unit volume, the electric polarization or average dipole moment
per unit volume is given by
a βeεa
µ = Ne x = Ne tanh (ii)
2 2
P ROBLEM 2.36: Calculate the canonical partition function and hence the specific heat of a classical
ideal two-dimensional gas.
S OLUTION
The energy of a single particle is
p2x + p2y
ε= (i)
2m
Then the single particle canonical partition function is
∞ ∞
1 A βp2
β βp2
( p2x +p2y ) − 2mx − 2my
ξ= 2 e − 2m
dx dy d px d py = 2 e d px × e d py
h h −∞ −∞
A 2πm
= 2 (ii)
h β
where A is the area.
116 • An Introduction to Equilibrium Statistical Mechanics
P ROBLEM 2.37: Suppose that a wire of radius r0 is placed along the axis of a metal cylinder of
radius R and length L. The wire is maintained at a positive potential V with respect to the cylinder
and the whole system is at some high absolute temperature T . As a result, electrons emitted from
the hot metals form a dilute gas filling the cylindrical container and in equilibrium with it. The
density of these electrons is so low that their mutual electrostatic interaction can be neglected.
Usine Gauss’ theorem, obtain an expression for the electrostatic potential at a radial distance
r from the wire (r0 < r < R). [Assume the cylinder to be infinitely long so that end effects are
neglected.] Also find the density of the electron gas filling the space between the wire and cylinder,
as a function of the radial distance in equilibrium.
S OLUTION
V V=0
r0
L
R
Elements of Ensemble Theory • 119
S OLUTION
In the rotating frame of reference, the molecules feel a centrifugal force mω2 r radially outward.
This force can be obtained from a potential
σV
Fr = mω2 rr̂ = − r̂ (i)
σr
where r̂ is the unit vector in the radial direction.
Therefore, the potential energy of a molecule in the rotating frame is
1
V = − mω2 r2 (ii)
2
so that the total energy of a molecule in the rotating frame is given by
p2
1 2 2
ε= + − mω r (iii)
2m 2
The probability of finding the molecule in the energy range ε to ε + dε is then given by
1 p2 mω2 r2
P(ε) = e 2mkT e 2kT (iv)
Z
Then the probability of finding the molecule between r and r + dr is obtained by integrating over
all the other coordinates and is given by
mω2 r2
P(r)dr = P(0)e 2kT dr (v)
where P(0) is a constant which correctly normalizes the distribution.
Equation (v) is the radial distribution of molecules. As expected, the molecules crowd towards
the larger radii. It should be noted that we have taken the rotating speed to be very high so that the
effect of the Coriolis force is ignored.
Now from (v), the concentration at r1 is
mω2 r12
σ(r1 ) ∝ e 2kT (vi)
Similarly,
mω2 r22
σ(r2 ) ∝ e 2kT
σ(r1 ) 2 2 2
1 mω (r1 −r2 )
∴ = e 2 kT (vii)
σ(r2 )
2kT σ(r1 )
⇒ m = 2 2 ln
ω r1 − r22 σ(r2 )
P ROBLEM 2.40: In order to measure a certain physical quantity, it is necessary to know the
equilibrium θ0 of a torsional pendulum. θ0 is determined by the minimum in the potential energy
of the pendulum V (θ) = 21 K(θ − θ0 )2 . The pendulum, when rotating, has a kinetic energy T = 12 Iθ,
where θ = dθ/dt. The precision with which θ0 can be found is limited because the pendulum is in
thermal equilibrium with its environment at a temperature T .
Elements of Ensemble Theory • 121
(b) The degeneracy of the nth energy level is n2 . The partition function for a single atom, ne-
glecting the unbound states, is
ε ∞ α
i
Z = ∑ exp − = ∑ n2 exp 2 (iii)
states kT n=1 n
where α = A kT . Since α > 0, it follows that exp α n2 > 1 for all n. Using this, we can set
the lower bound of Z, but Z diverges since the lower bound diverges.
∞
∴ Z> ∑ n2 , which diverges.
n=1
As the separating walls are conducting and permeable, the exchange of heat energy as well as
that of particles between the systems take place in such a way that all the systems arrive at common
temperature T and the chemical potential µ.
Er + E = E 0 = constant
(2.3.1)
Nr + N = N 0 = constant
where E and N denote the energy and number of particles in the reservoir R respectively.
122 • An Introduction to Equilibrium Statistical Mechanics
Let Ω (E N ) denote the number of states accessible to R , when it contains N particles and has
an energy in the range near E . Then if R is in a particular
state r, the number of states accessible
to the combined system R0 is 1 × Ω (E , N ) = Ω E 0 − Er , N 0 − Nr . Hence, the probability of
finding the system R in this state is
1 × Ω E 0 − Er , N 0 − Nr
Pr = (2.3.2)
Ω0total
where Ω0total is the total number of states accessible to the combined system R0 . It is independent of
r and is just a constant. Writing Ω01 = c = constant,
total
Pr = c Ω E 0 − Er , N 0 − Nr
(2.3.3)
Taking logarithms on both sides
ln Pr = ln c + ln Ω E 0 − Er , N 0 − Nr
(2.3.4)
Since R 0is very 0small compared to R , Er E 0 and Nr N 0 , expanding ln Ω (E , N )
= ln Ω E − Er , N − Nr about E = E 0 and N = N 0 , we have
0 0
0 0 ∂ ln Ω ∂ ln Ω
ln Ω E − Er , N − Nr = ln Ω E , N −
Er − Nr . . .
∂E E 0 ∂N N 0
Since Er E 0 and Nr N 0 , we can neglect the higher order terms and so we have
∂ ln Ω ∂ ln Ω
ln Ω E 0 − Er , N 0 − Nr = ln Ω E 0 , N 0 −
E r − Nr (2.3.5)
∂E E 0 ∂N N 0
The derivatives are evaluated
for E = E 0 and 0
N = N and so they are constants, characterizing the
reservoir R . We write ∂ ln Ω = β and ∂ ln Ω = α.
∂E E 0 ∂N
Then equation (2.3.5) becomes
ln Ω E 0 − Er , N 0 − Nr = ln Ω E 0 , N 0 − βEr − αNr
(2.3.6)
0 0
0 0 −βE −αN
⇒ Ω E − Er , N − Nr = Ω E , N e r
r
The normalization condition for probabilities demands ∑ Pr = 1. Applying this to equation (2.3.8),
r
1
we have c = and so
∑ eµβNr −βEr
r
Now
∑ Nr Pr ∑ Nr eµβNr −βEr
N= r
= ∑ Nr Pr = r
.
∑ Pr r ∑ eµβNr −βEr
r r
The grand canonical partition function is
1 ∂Z
Z (µ,V, T ) = ∑ eµβNr −βEr ⇒
β ∂µ ∑
= Nr eµβNr −βEr .
r r
1 1 ∂Z 1 ∂ ln Z
∴ N= =
β Z ∂µ β ∂µ
Again
∑ Nr2 Pr ∑ Nr2 eµβNr −βEr ∑ Nr2 eµβNr −βEr
1 1 ∂2 Z
N2 = r
=∑ Nr2 Pr = r
= r
=
∑ Pr r ∑ eµβNr −βEr Z (µ,V, T ) Z β2 ∂µ2
r r
1 1 ∂2 Z 1 1 ∂Z 2 1 ∂ 1 1 ∂Z
2 ∂N
Hence, (∆N) = 2 − = = kT .
β Z ∂µ2 β Z ∂µ β ∂µ β Z ∂µ ∂µ
∴ The root mean square fluctuation in N is
1/2 ∂N 1/2
∆∗ N = (∆N)2 = kT .
∂µ T,V
The relative fluctuation in N is then
1/2
∆∗ N 1 ∂N
= kT (2.3.10)
N N ∂µ T,V
Elements of Ensemble Theory • 125
and
∑ Er2 eµβNr −βEr ∑ Er2 eµβNr −βEr
2 r r
E = =
∑ eµβNr −βEr Z (µ,V, T )
r
∂2 Z 1 ∂Z 2
1 ∂ 1 ∂Z
= = + 2
Z ∂β2 µ,V ∂β Z ∂β Z ∂β
∂ ln Z 2
∂ ∂ ln Z ∂E 2
= + =− + E
∂β ∂β ∂β ∂β
2 2 2 ∂E 2 ∂E
∴ (∆E) = E − E = − = kT
∂β µ,V ∂T µ,V
and
1 ∂ ln Z∂ ln Z
N= =− (2.3.14)
β V,T ∂µ∂ (−µβ) V,T
Eliminating µ from (2.3.13) and (2.3.14), we see that E = E N,V, T , so that
∂E ∂E ∂E ∂N ∂E ∂N
= + × = CV + ×
∂T µ,V ∂T N,V ∂N T,V ∂T µ,V ∂N T,V ∂T µ,V
where CV = ∂E ∂T N,V , the specific heat of the system at constant volume.
∂E ∂N
∴ (∆E)2 = kT 2 CV +
∂N T,V ∂T µ,V
Again from equation (2.3.13)
∂E ∂N ∂N 1 ∂E
= and hence = .
∂ (−µβ) V,T ∂β µ,V ∂T µ,V T ∂µ T,V
2 2 1 ∂E ∂E
∴ (∆E) = kT CV + ×
T ∂N T,V ∂µ T,V
∂E ∂E
= kT 2CV + kT
∂N T,V ∂µ T,V
2
∂E ∂N
= kT 2CV + kT
∂N T,V ∂µ T,V
2 ∂N
But (∆N) = kT .
∂µ T,V
2 2
2 ∂E 2 2∂E
∴ (∆E) = kT CV +2
(∆N) = (∆E) canonical + (∆N)2 (2.3.15)
∂N T,V ∂N T,V
126 • An Introduction to Equilibrium Statistical Mechanics
This is the desired result, which says that the mean square fluctuation in the energy E of a system
in the grand canonical ensemble is equal to the value it would have in the canonical ensemble plus
a contribution arising due to the fluctuation in the particle number. Under ordinary circumstances,
the relative fluctuation in the energy of the system is negligible. However, in the region of phase
transition, the fluctuation in the energy is large because of the second term.
2. Specific heat
The specific heat at constant volume is given by
∂E
CV = (2.3.17)
∂T N,V
where E is given by (2.3.16).
3. Thermodynamic pressure
The grand canonical partition function is Z = Z (µ,V, T ). For the sake of convenience, let us write
Z as a function of α (= −µβ), V and β (= 1/kT ), i.e., Z = Z (α,V, β).
∂ ln Z ∂ ln Z ∂ ln Z
∴ d ln Z (µ,V, T ) = d ln Z (α,V, β) = dα + dV + dβ (2.3.18)
∂α ∂V ∂β
Now
1 ∂ ln Z ∂ ln Z
N= =− (2.3.19)
β ∂µ V,T ∂α V,T
Using (2.3.16) and (2.3.19) in (2.3.18)
∂ ln Z
d ln Z = −Ndα + dV − Edβ
∂V
∂ ln Z
⇒ d ln Z + Nα + Eβ = αdN + βdE + dV
∂V
Nµ E 1 ∂ ln Z
⇒ d k ln Z − + = dE + dV − µdN
kT kT T ∂V
To interpret the second term on the right-hand side of this equation, we compare the expression
enclosed within the parenthesis with the statement of the first law of thermodynamics, viz.,
δQ = dE + δW − µdN (2.3.20)
Elements of Ensemble Theory • 127
5. Entropy
The entropy of the system is given by
E −F ∂ ln Z (µ,V, T ) Z (µ,V, T )
S≡ = kT + k ln
T ∂T µ,V zN
∂ ln Z (µ,V, T )
∴ S = kT + k ln Z (µ,V, T ) − Nkµβ (2.3.23)
∂T µ,V
P ROBLEM 2.44: Show that the entropy of a grand canonical ensemble can be written as
S = −k ∑ Pr ln Pr .
r
S OLUTION
The grand canonical partition function is
Z(µ,V, T ) = ∑ eµβNr −βEr (i)
r
Now
∂ ln Z ∂ ln Z ∂ ln Z
d ln Z = dα + dV + dβ = Ndα + pβdV + βdE
∂α ∂V ∂β
⇒ d ln Z − Nα + Eβ = −αdN + pβdV + βdE
α
⇒ d ln Z − Nα + Eβ = β dE + pdV − dN = β dE + pdV − µdN
β
Elements of Ensemble Theory • 131
S OLUTION
We know that
S = −k ∑ Pr ln Pr (i)
r
Now
eµβNr −βEr
Pr = ⇒ ln Pr = µβNr − βEr − ln Z (ii)
Z (µ,V, T )
Using (ii) in (i)
S = −k µβ ∑ Pr Nr − β ∑ Pr Er − ln Z ∑ Pr
r r r
∵ ∑ Pr = 1, ∑ Pr Er = E, ∑ Pr Nr = N
= −k µβN − βE − ln Z
r r r
= k ln Z + kβE − kµβN
∂ ln Z ∂ ln Z ∂ ln Z
⇒ dS = k dµ+ k dV + k dT + kβdE + kEdβ − kβµdN − kβNdµ− kµNdβ
∂µ ∂V ∂T
= kβNdµ+ kβpdV + kµNdβ − kEdβ + kβdE + kEdβ − kβµdN − kβNdµ− kµNdβ
∂ ln Z ∂ ln Z ∂ ln Z ∂ ln Z ∂ ln Z
∵ dT = dβ, = βN, = βp, = µN − E
∂T ∂β ∂µ ∂V ∂β
= kβdE + kβpdV − kβµdN
1
⇒ T dS = dE + pdV − µdN ∵ β=
kT
Elements of Ensemble Theory • 133
P ROBLEM 2.47: Show that Z (µ,V, T ) = eN for an ideal gas, if the gas molecules are treated as
indistinguishable.
S OLUTION
The grand canonical partition function is Z (µ,V, T ) = ∑ eµβNr −βEr . We can write this partition
r
function in another form:
∞
Z (µ,V, T ) = ∑ eµβN−βENr = ∑ eµβN e−βENr = ∑ eµβN Z (N,V, T )
Nr Nr N=0
S OLUTION
The probability that the region of volume v exactly contains n molecules is given by
∑ eµβn−βEnr eµβn ∑ e−βEnr
r r eµβn Z (n, v, T )
Pn = = = (i)
Z (µ, v, T ) Z (µ, v, T ) Z (µ, v, T )
where Z(n, v, T ) = canonical partition function = ξn /n!, ξ being the single particle canonical parti-
tion function. n! appears because of the essential indistinguishability of the particles. Again, from
problem 2.50, n = eµβ ξ and Z (µ, v, T ) = en .
(eµβ ξ)n / n! (n)n −n
∴ Pn = en
= n! e . This is known as Poisson’s distribution.
P ROBLEM 2.49: Consider a system of N molecules forming a substance which consists of two
phases. The whole system is isolated so that its total energy E and its total volume V are both fixed.
Find the necessary conditions for equilibrium between the two phases.
S OLUTION
Let the two phases be 1 and 2. The whole system is isolated, so that its total energy E, its total
volume V and its total number of particles N are fixed. Let Ni be the number of molecules of the
substance in phase i and denote the energy of this phase by Ei and its volume by Vi .
Then we must have the conservation conditions:
E1 + E2 = E = constant
⇒ dE1 = −dE2 (i)
V1 +V2 = V = constant
⇒ dV1 = −dV2 (ii)
N1 + N2 = N = constant
⇒ dN1 = −dN2 (iii)
The equilibrium condition corresponds to the most probable situation, i.e., when the thermodynamic
probability Ω is maximum. But S = k ln Ω. Hence, the equilibrium condition is that the entropy is
maximum, i.e.,
S = S (E1 ,V1 , N1 ; E2 ,V2 , N2 ) = maximum (iv)
Again S = S1 (E1 ,V1 , N1 ) + S2 (E2 ,V2 , N2 ).
Thus, the maximum condition (iv) yields
dS = dS1 + dS2 = 0 (v)
Now S1 = S1 (E1 ,V1 , N1 ).
∂S1 ∂S1 ∂S1
∴ dS1 = dE1 + dV1 + dN1 (vi)
∂E1 V1 ,N1 ∂V1 E1 ,N1 ∂N1 E1 ,V1
In a simple case, when the number N1 is fixed, the fundamental thermodynamic relation asserts that
1
dS1 = (dE1 + p1 dV1 ) (vii)
T1
Elements of Ensemble Theory • 135
where T1 and p1 are respectively the temperature and mean pressure of phase 1. Under this circum-
stance, dN1 = 0. A comparison of the coefficients of dE and dV in (vi) and (vii) yields
∂S1 1
= (viii)
∂E1 V,N1 T1
∂S1 p1
= (ix)
∂V1 E1 ,N T1
∂S1
Let µ1 = −T1 ∂N1 . The quantity µ1 is the chemical potential per molecule in phase 1.
E1 ,V1
Then (vi) can be written in the form
1 p1 µ1
dS1 = dE1 + dV1 − dN1 (x)
T1 T1 T1
Similarly,
1 p2 µ2 1 p2 µ2
dS2 = dE2 + dV2 − dN2 = − dE1 − dV1 + dN1 (xi)
T2 T2 T2 T2 T2 T2
Using (x) and (xi) in (v)
1 1 p1 p2 µ1 µ2
− dE1 + − dV1 − − dN1 = 0 (xii)
T1 T2 T1 T2 T1 T2
Since (xii) is to be valid for arbitrary variations dE1 , dV1 , dN1 , it follows that the coefficients of all
these differentials must vanish separately. Thus, one obtains
1 1
− =0
T1 T2
p1 p2
− =0
T1 T2
µ1 µ2
− =0
T1 T2
whence
T1 = T2 (xiii)
p1 = p2 (xiv)
µ1 = µ2 (xv)
These are the necessary conditions for equilibrium between two phases. (xiii) is the condition for
thermal equilibrium, (xiv) for mechanical equilibrium and (xv) for diffusive equilibrium.
Elements of Ensemble Theory • 137
where v is 2f-dimensional velocity of the representative points and n̂ is a unit vector, outward
normal to the point of location of the surface element ds of the surface s, enclosing the volume τ
and dτ = dq1 . . . dq f d p1 . . . d p f .
Applying Gauss’ divergence theorem
ρv · n̂ds = ∇ · (ρv) dτ (2.5.3)
s τ
Using (2.5.3) in (2.5.2)
∂ρ
+ ∇ · (ρv) dτ = 0. Since τ is arbitrary,
τ ∂t
∂ρ
+ ∇ · (ρv) = 0 (2.5.4)
∂t
This is the well-known equation of continuity and is a statement of the conservation of density in
phase space. Here ∇ is a 2f-dimensional del operator.
The second term of (2.5.4) can be written as
f
∇ · (ρv) = ∑ ∂ (ρq̇i ) + ∂ (ρ ṗi )
i=1 ∂qi ∂pi
(2.5.5)
f f
∂q̇i ∂ ṗi ∂ρ ∂ρ
= ρ∑ + +∑ q̇i + ṗi
i=1 ∂qi ∂pi i=1 ∂qi ∂pi
In view of the Hamilton’s equation (2.5.1), the first term is
f 2
∂2 H
∂ H
∑ − =0
i=1 ∂qi ∂pi ∂pi ∂qi
∴ Equation (2.5.4) takes the final form
f
∂ρ ∂ρ ∂ρ
+∑ q̇i + ṗi = 0 (2.5.6)
∂t i=1 ∂qi ∂pi
Again using Hamilton’s equation (2.5.1), we can put (2.5.6) in another form:
f
∂ρ ∂ρ ∂H ∂ρ ∂H
+∑ − =0
∂t i=1 ∂qi ∂pi ∂pi ∂qi
∂ρ
⇒ + [ρ, H] = 0 (2.5.7)
∂t
f
∂ρ ∂H ∂ρ ∂H
where [ρ, H] = ∑ − , the Poisson bracket.
i=1 ∂qi ∂pi ∂pi ∂qi
dρ
Since ρ is a function of q’s, p’s and t, the left-hand side is the total time derivative, .
dt
∂ρ ∂ρ
∴ = + [ρ, H] = 0 (2.5.8)
∂t ∂t
138 • An Introduction to Equilibrium Statistical Mechanics
QUESTIONS
1. Give the concept of ensemble.
2. What do you mean by microcanonical ensemble? For what type of system is it suitable?
3. Show that Ω(E) ∼ E f , i.e., Ω(E) is a rapidly increasing function of the energy E of the
system.
4. Discuss the specific heat behaviour of a two-level system. What is Schottky anomaly?
5. What do you mean by canonical ensemble? For what type of system is it suitable?
6. For a canonical system, derive the Boltzmann’s canonical probability distribution law,
pr = e−βEr 1
where β = kT .
∑ e−βEr
r
7. What is partition function? What is its importance in statistical mechanics?
8. Show that the energy fluctuation in a canonical system is proportional to √1N
9. How can you justify the equivalence between a microcanonical ensemble and a canonical
ensemble?
140 • An Introduction to Equilibrium Statistical Mechanics
3. Consider a system of N particles at a temperature T . Each particle has mass m and free to ex-
ecute one-dimensional simple harmonic motion about its mean position. If the restoring force
is proportional to x, calculate the specific heat of the system. Assume that the temperature T
is high enough so that classical statistics is applicable.
4. Find the canonical distribution function for a body rotating with angular velocity.
5. Suppose that two atoms interact with the mutual potential energy of the form
a 12 a 6
U (x) = U0 −2 , where x is the separation between the two particles. The
x x
particles are in contact with a heat reservoir at an absolute temperature T such that kT U0 ,
but high enough so that classical statistics is applicable. Calculate the mean separation x (T )
1 ∂x
of the particles and use it to compute the expansion coefficient α = .
x ∂T