R A F T: Notes On Statistical Mechanics
R A F T: Notes On Statistical Mechanics
R A F T: Notes On Statistical Mechanics
Murthy
Notes on
STATISTICAL MECHANICS
August 28, 2017
T
AF
DR
It was certainly not by design that the particles fell into order
They did not work out what they were going to do,
but because many of them by many chances
struck one another in the course of infinite time
and encountered every possible form and movement,
that they found at last the disposition they have,
and that is how the universe was created.
Titus Lucretius Carus (94 BC - 55BC)
de Rerum Natura
Everything existing in the universe is the fruit of chance and ne-
cessity. Democritus (370 BC)
The moving finger writes; and, having writ,
moves on : nor all your piety nor wit
shall lure it back to cancel half a line
nor all your tears wash out a word of it
Omar Khayyam (1048 - 1131)
Whatever happened,
T
happened for good.
Whatever is happening,
AF
is happening for good.
Whatever will happen,
will happen for good.
DR
Bhagavat Gita
”· · · Ludwig Boltzmann, who spent much of his life studying sta-
tistical mechanics, died in 1906, by his own hand. Paul Ehrenfest,
carrying on the work, died similarly in 1933. Now it is our turn ....
to study statistical mechanics. Perhaps it will be wise to approach
the subject rather cautiously.· · · ”
David Goodstein, States Matter, Dover (1975) (opening lines)
All models are wrong, some are useful. George E P Box
DR
AF
T
Contents
Quotes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
1. Micro-Macro Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Aim of Statistical Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Micro - Macro Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Boltzmann Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Boltzmann-Gibbs-Shannon Entropy . . . . . . . . . . . . . . . . 4
1.2.3 Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.4 Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.5 Helmholtz Free Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.6 Energy Fluctuations and Heat Capacity . . . . . . . . . . . . 6
1.3 Micro World : Determinism and T
Time-Reversal Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
AF
1.4 Macro World : Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Books . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6 Extra Reading : Books . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.7 Extra Reading : Papers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
DR
2. Maxwell’s Mischief . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1 Experiment and Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Sample space and events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Probabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Rules of probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Random variable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Maxwell’s mischief : Ensemble . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.7 Calculation of probabilities from an ensemble . . . . . . . . . . . . . 19
2.8 Construction of ensemble from probabilities . . . . . . . . . . . . . . . 19
2.9 Counting of the elements in events of the sample space :
Coin tossing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.10 Gibbs ensemble . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.11 Why should a Gibbs ensemble be large ? . . . . . . . . . . . . . . . . . 22
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
T
AF
DR
DR
AF
T
1. Micro-Macro Synthesis
The first and the most important link, between the micro(scopic) and the
macro(scopic) worlds is given by,
b
S = kB ln Ω(E, V, N ). (1.1)
It was proposed by Boltzmann3 . S stands for entropy and belongs to the
macro world described by thermodynamics. Ω b is the number of micro states
of a macroscopic system . kB is the Boltzmann constant 5 that establishes
4
The micro states are labelled by i and the sum is taken over all the micro
T
states of the macroscopic system; The probability is denoted by pi .
An interesting issue : Normally we resort to the use of probability only when
AF
we have inadequate data about the system or incomplete knowledge of the phe-
nomenon under study. Thus, such an enterprise is sort of tied to our ignorance.
Keeping track of the positions and momenta of some 1030 molecules via Newton’s
equations is undoubtedly an impossible task. It was indeed the genius of Boltzmann
DR
which correctly identified that macroscopic phenomena are tailor-made for a statis-
tical description. It is one thing to employ statistics as a convenient tool to study
macroscopic phenomena but quite another thing to attribute an element of truth to
such a description. But then this is what we are doing precisely in Eq. (1.2) where
we express entropy, which is a property of an object, in terms of probabilities. It is
definitely a bit awkward to think that a property of an object is determined by what
we know or what we do not know about it! But remember in quantum mechanics
3
engraved on the tomb of Ludwig Eduard Boltzmann (1844-1906) in Zentralfried-
hof, Vienna.
4
For example an ordered set of six numbers, three positions and three momenta
specify a single particle. An ordered set of 6N numbers specify a macroscopic
system of N particles. The string labels a micro state.
5
kB = 1.381×10−23 joules (kelvin)−1 . We have kB = R/A where R = 8.314 joules
(kelvin)−1 is the universal gas constant and A = 6.022 × 1023 (mole)−1 is the
Avagadro number.
6
Note that the quantity ln Ωb is a number. Claussius’ entropy is measured in units
of joules/kelvin. Hence the Boltzmann constant is kB and has units joules/kelvin.
It is a conversion factor and helps you go from one unit of energy (joule) to
another unit of energy (kelvin).
1.2 Micro-Macro Connections 5
the observer, the observed, and the observation are all tied together : the act of
measurement places the system, at the time of observation, in one of eigenstates
of the observable; we can make only probabilistic statement about the eigenstate
the wave function of the system would collapse into. For a discussion on this subtle
issue see the beautiful tiny book of Carlo Rovelli7 .
Two important entities we come across in thermodynamics are heat and
work. They are the two means by which a thermodynamic system transacts
energy with its surroundings or with another thermodynamic system. Heat
and work are described microscopically as follows.
1.2.3 Heat
X
q= Ei dpi . (1.3)
i
The sum runs over all the micro states. Ei is the energy of the system when
it is in micro state i. The probability that the system can be found inPmicro
state i given by pi . We need to impose an additional constraint that i dpi
is zero to ensure that the total probability is unity.
1.2.4 Work
The statistical description of work is given by
X
T
AF
W = pi dEi . (1.4)
i
Ei is the energy of the system when it is in the micro state labelled by the
index i.
DR
formation of t → −t.
• The second law tells us that come what may, an engine can not de-
liver work equivalent to the energy it has drawn by heat from a heat
source. However an engine can draw energy by work and deliver exactly
equivalent amount by heat. The second law is a statement of this ba-
sic asymmetry between ”heat→ work” and ”work→ heat” conversions.
The Second law provides a basis for the thermodynamic property called
entropy. The entropy of a system increases by dS = d¯q/T when it ab-
sorbs reversibly d¯q amount of energy by heat at constant temperature.
The second law says that in any process entropy increases or remains
constant : ∆S ≥ 0.
• The third law tells that entropy vanishes at absolute zero tempera-
ture. Notice that in thermodynamics only change in entropy is defined :
dS = d¯q/T . Since dS is an exact differential we can assert S as a ther-
modynamic variable and it describes a property of the system. The third
law demands that this property is zero at absolute zero and hence we can
assign a value S at any temperature. We can say the third law provides
a basis for absolute zero temperature on entropy scale.
Of these, the second law is tricky. It breaks the time-symmetry present
in the microscopic descriptors. Macroscopic behaviour is not time-reversal
invariant. There is a definite direction of time - the direction of increasing
entropy. T
How do we comprehend the time asymmetric macroscopic behaviour
emerging from the time symmetric microscopic laws ?
AF
Let us make life simpler by attributing two aims to statistical mechanics.
The first is to provide a machinery for calculating the thermodynamic prop-
erties of a system on the basis of the properties of its microscopic constituents
e.g. atoms and molecules, and their mutual interactions.
DR
1.5 Books
translation see Sadi carnot, Reflections on the motive power of fire and
on machines fitted to develop that power, in J Kestin (Ed.) The second
law of thermodynamics, Dowden, Hutchinson and Ross, Stroudsburg, PA
(1976)p.16
• J Kestin (Ed.), The second law of thermodynamics, Dowden, Hutchinson
and Ross (1976)
• P Atkin, The Second Law, W H Freeman and Co. (1984)
• G Venkataraman, A hot story, Universities Press (1992)
• Michael Guillen, An unprofitable experience : Rudolf Clausius and the
second law of thermodynamics p.165, in Five Equations that Changed the
World, Hyperion (1995)
• P Atkins, Four Laws that drive the Universe, Oxford university Press
(2007).
• Christopher J T Lewis, Heat and Thermodynamics : A Historical
Perspective, First Indian Edition, Pentagon Press (2009)
• S G Brush, Kinetic theory Vol. 1 : The nature of gases and of heat,
Pergamon (1965) Vol. 2 : Irreversible Processes, Pergamon (1966)
14 1. Micro-Macro Synthesis
• S G Brush, The kind of motion we call heat, Book 1 : Physics and the
Atomists Book 2 : Statistical Physics and Irreversible Processes, North
Holland Pub. (1976)
• I Prigogine, From Being to Becoming, Freeman, San Francisci (1980)
• K P N Murthy, Excursions in thermodynamics and statistical mechan-
ics, Universities Press (2009)
Toss a coin : You get either ”Heads” or ”Tails”. The experiment has two
outcomes. Consider tossing of two coins. Or equivalently toss a coin twice.
There are four outcomes : An outcome is an ordered pair. Each entry in the
pair is drawn from the set {H, T }.
We can consider, in general, tossing of N coins. There are 2N outcomes.
Each outcome is an ordered string of size N with entries drawn from the set
{H, T }.
Select randomly an air molecule in this room and find its position
and momentum : Consider the air molecule to be a point particle. In clas-
sical mechanics, a point particle is completely specified by its three position
(q1 , q2 , q3 ) and three momentum (p1 , p2 , p3 ) coordinates. An ordered set of
six numbers
{q1 , q2 , q3 , p1 , p2 , p3 }
is the outcome of the experiment. A point in the six dimensional phase space
represents an outcome of the experiment or the microstate of the system of
single molecule. We impose certain constraints e.g. the molecule is always
confined to this room. Then all possible strings of six numbers, consistent
with the constrains, are the outcomes of the experiment.
The set of all possible outcomes of an experiment is called the sample space.
Let us denote it by the symbol Ω.
16 2. Maxwell’s mischief
2.3 Probabilities
Probability is defined for an event. What is the probability of the event {H}
in the toss of a coin ? One-half. This would be your immediate response.
The logic is simple. There are two outcomes : ”Heads” and ”Tails”. We have
no reason to believe why should the coin prefer ”Heads” over ”Tails” or
T
vice versa. Hence we say both outcomes are equally probable. What is the
probability of having at least one ”H” in a toss of two coins ? The event
AF
corresponding this statement is {HH, HT, T H} and contains three elements.
The sample size contains four elements. The required probability is thus 3/4.
All the four outcomes are equally probable 1 . Then the probability of an
event is the number of elements in that event divided by the total number of
DR
elements in the sample space. For e.g., the event A of rolling an even number
in a game of dice, P (A) = 3/6 = 0.5. The outcome can be a continuum. For
example, the angle of scattering of a neutron is a real number between zero
and π. We then define an interval (θ1 , θ2 ), where 0 ≤ θ1 ≤ θ2 ≤ π, as an
event. A measurable subset of a sample space is an event.
as Z b
µ= dx x f (x).
a
µ is also called the mean, expectation, first moment etc.
Consider a discrete random variable n, taking values from say 0 to N . Let
P (n) define the discrete probability. We define the average of the random
variable as
2
Maxwell and Boltzmann attached probabilities to events in some way; we got
Maxwell-Boltzmann statistics.
Fermi and Dirac had their own way of assigning probabilities to Fermions e.g.
electrons, occupying quantum states. We got Fermi-Dirac statistics.
Bose and Einstein came up with their scheme of assigning probabilities to
Bosons, populating quantum states; and we got Bose-Einstein statistics.
3
In fact, we stamped dots on the faces of die; this is equivalent to implementing
the idea of an integer random variable : attach an integer between one and six
to each outcome.
For a coin, we stamped ”Heads” on one side and ”Tails” on the other. This is
in the spirit of defining a random variable; we have stamped figures instead of
numbers.
18 2. Maxwell’s mischief
N
X
µ= n P (n).
n=0
But then, we are accustomed to calculating the average in a different
way. For example I am interested in knowing the average marks obtained
by the students in a class, in the last mid-semester examination. How do I
calculate it ? I take the marks obtained by each of you, sum them up and
divide by the total number of students. That is it. I do not need notions
like probabilities, probability density, sum over the product of the random
variable and the corresponding probability, integration of the product of the
continuous random variable and its probability density function etc.
Historically, before Boltzmann and Maxwell, physicists had no use for
probability theory in their work. Newton’s equations are deterministic. There
is nothing chancy about a Newtonian trajectory. We do not need probabilistic
description in the study of electrodynamics described by Maxwell equations;
nor do we need probability to comprehend and work with Einstein’s relativity
- special or general.
However mathematicians had developed the theory of probability as an
important and sophisticated branch of mathematics.
It was Ludwig Eduard Boltzmann who brought, for the first time, the
idea of probability into physical sciences; he was championing the cause of
T
kinetic theory of heat and atomic theory of matter. Boltzmann transport
equation is the first ever equation written for describing the time evolution
AF
of a probability distribution.
We can also do the reverse. Given the outcomes and their probabilities, we
can construct an ensemble. Let ni denote the number of times an outcome
i occurs in an ensemble. Let N denote the total number of elements of the
ensemble. Choose ni such that ni /N equals pi ; note that we have assumed
that pi is already known.
b ) = 2N .
Ω(N
Let Ω(n1 , n2 ; N ) denote a subset of Ω(N ), containing only those outcomes
with n1 ’Heads’ and n2 ’Tails’. Note n1 + n2 = N . How many outcomes are
there in the set Ω(n1 , n2 ; N ) ?
b 1 , n2 ; N ) denote the number of elements in the event Ω(n1 , n2 ; N ).
Let Ω(n
b 1 , n2 ; N ).
In what follows I shall tell you how to derive an expression6 for Ω(n
Take one outcome belonging to the event Ω(n1 , n2 ; N ). There will be
n1 ’Heads’ and n2 ’Tails” in that outcome . Imagine for a moment that
all these ’Heads’ are distinguishable. If you like, you can label them as
H1 , H2 , · · · , Hn1 . Carry out pemutation of all the ’Heads’ and produce
n1 ! new configurations. From each of these new configurations, produce n2 !
configurations by carrying out the permutations of the n2 ’Tails’. Thus from
one outcome belonging to Ω(n1 , n2 ; N ), we have produced n1 ! × n2 ! new con-
figurations. Repeat the above for each element of the set Ω(n1 , n2 ; N ), and
b 1 , n2 ; N )n1 !n2 ! configurations. A moment of thought will tell you
produce Ω(n
that this number should be the same as N !. Therefore
b 1 , n2 ; N ) × n1 ! × n2 ! = N !
Ω(n
Let us work out an example explicitly to illustrate the above. Consider
tossing of five coins. There are 25 = 32 outcomes/microstates listed below.
The number in the brackets against each outcome denotes the number of
”Heads” in that outcome. T
AF
Table 2.1. List of 32 outcomes of tossing five coins. Given in brackets is
the number of ”Heads” in each outcome
DR
The outcomes numbered 4, 6, 7, 10, 11, 13, 18, 19, 20, 25 are the ones
with three ”Heads” and two ”tails”). These are the elements of the event
Ω(n1 = 3; n2 = 2; N = 5).
Take outcome No. 4. Label the three heads as H1 , H2 and H3 . Carry out
permutations of the three ”Heads” and produce 3! = 6 elements. These are
Take an element from the above set. Label the ’Tails’ as T1 and T2 . Carry
6
I remember I have seen this method described in a book on Quantum Mechanics
by Gasiorowicz. Check it out
2.9 Counting of elements of an event 21
Table 2.2. In outcome no. 4, see Table (2.1), the Heads are permuted
(4) H H H T T
H2 H3 H1 T T
H1 H2 H3 T T
H3 H1 H2 T T
H1 H3 H2 T T
H3 H2 H1 T T
H2 H1 H3 T T
out permutation of the ’Tails’ and produce 2! = 2 elements. Do this for each
of the above six elements.
Thus from the outcome No, 4, we have produced 3! × 2! = 12 out-
comes, listed below. Repeat the above exercise on the outcomes numbered
(4) H H H T T
H2 H3 H1 T1 T2
H1 H2 H3 T1 T2
H2 H3 H1 T2 T1
H1 H2 H3 T2 T1
H3 H1 H2 T1 T2
H1 H3 H2 T1 T2
H3 H1 H2 T2 T1
H1 H3 H2 T2 T1
H3 H2 H1 T1 T2
H2 H1 H3 T1 T2
H3 H2 H1 T2 T1
H2 H1 H3 T2 T1
T
6, 7, 10, 11, 13, 18, 19, 20, 25. Table below depicts the results for outcome
no. 6.
AF
Table 2.4. The twelve outcomes produced from outcome No. 6
DR
(6) H H T H T
H2 H3 T1 H1 T2
H1 H2 T1 H3 T2
H2 H3 T2 H1 T1
H1 H2 T2 H3 T1
H3 H1 T1 H2 T2
H1 H3 T1 H2 T2
H3 H1 T2 H2 T1
H1 H3 T2 H2 T1
H3 H2 T1 H1 T2
H2 H1 T1 H3 T2
H3 H2 T2 H1 T1
H2 H1 T2 H3 T1
b 1 = 3, n2 = 2; N = 5) = 5! = 10
Ω(n
3!2!
b 1 , n2 ; N ) is called the binomial coefficient7
Ω(n
7
We have the binomial expansion given by
X⋆ N ! n1 n2
(a + b)N = a b
n1 !n2 !
{n1 ,n2 }
22 2. Maxwell’s mischief
9
This means the values of p and q are the same for all the coins belonging to the
ensemble.
10
If you want to estimate the probability of Heads in the toss of a single coin
experimentally then you have to toss a large number of identical coins. Larger
the size of the ensemble more (statistically) accurate is your estimate .
11
you can find this in several ways. Just guess it. I am sure you would have guessed
the answer as n1 = n2 = N/2. We know that the binomial coefficient is largest
when n1 = n2 = N/2 if N is even, or when n1 and n2 equal the two integers
closest to N/2 for N odd. That is it.
b 1 , n2 ; N ) with respect
If you are more sophisticated, take the derivative of Ω(n
to n1 and n2 with the constraint n1 + n2 = N , set it zero; solve the resulting
equation to get the value of N for which the function is an extremum. Take the
second derivative and show that the extremum is a maximum.
You may find it useful to take the derivative of logarithm of Ω(n b 1 , n2 ; N );
employ Stirling approximation for the factorials : ln(m!) = m ln(m) −
m for large m. Stirling approximation to large factorials is described in the
next section.
You can also employ any other pet method of yours to show that, for n = N/2,
b
the function Ω(n; N ) is maximum.
2.11 Why should Gibbs ensemble be large ? 23
X⋆ N!
b )=
Ω(N = 2N ; (⋆ ⇒ n1 + n2 = N ) (2.3)
n1 ! n2 !
{n1 ,n2 }
b m (N ) = Ω(n
b 1 = n2 = N/2; N ) = N!
Ω (2.4)
(N/2)! (N/2)!
b m (N ) for large values of N . We employ Stirling’s first
Let us evaluate Ω
approximation : N ! = N N exp(−N ) and get
12
b m (N ) = = N N exp(−N ) N
Ω 2 = 2 (2.5)
(N/2)(N/2) exp(−N/2)
The above implies that almost all the outcomes of the experiment belong
to the event with equal number of ’Heads’ and ’Tails’. The outcomes with
unequal number of ’Heads’ and ’Tails’ are so overwhelmingly small that the
difference falls within the small error arising due to the first Stirling approx-
imation.
For estimating the tiny difference between Ω(Nb ) and Ω b m (N ), let us em-
13
ploy the second Stirling approximation :
12
First Stirling Approximation : N ! ≈ N N exp(−N ).
We have, N ! = N × (N − 1) × · · · × 3 × 2 × 1. Therefore,
ln N ! = ln 1 + ln 2 + ln 3 + · + ln N T
XN Z N
N
= ln(k) ≈ ln x dx = (x ln x − x)1 = N ln N − N − 1
AF
k=1 1
≈ N ln N − N
N ! ≈ N N exp(−N )
DR
13
A better approximation√ to large factorials is given by Stirling’s second formula
: N ! ≈ N N exp(−N ) 2πN . We have
Z ∞ Z ∞
Γ (N + 1) = N ! = dx xN e−x = dx eF (x)
0 0
′ N N
F (x) = N ln x − x; F (x) = − 1 ; F ′′ (x) = − 2
x x
Set F ′ (x) = 0; this gives x⋆ = N . At x = x⋆ the function F (x) is maximum.
(Note : F ′′ (x = x⋆ ) is negative). Carrying out Taylor expansion and retaining
only upto quadratic terms, we get,
(x − x⋆ )2 ′′ (x − N )2
F (x) = F (x⋆ ) + F (x = x⋆ ) = N ln N − N −
2 2N
We have,
Z ∞ Z
∞
(x − N )2
N! = dx eF (x) = N N e−N dx exp −
0 0 2N
√ Z ∞ √
= N N e−N N √ dx exp(−x2 /2) ∼
N→∞ N N e−N 2πN
− N
24 2. Maxwell’s mischief
√
N ! = N N exp(−N ) 2πN
and get
b m (N ) = Ω(n
Ω b 1 = n2 = N/2; N )
r
N 2
=2 ×
πN
r
b 2
= Ω(N ) × (2.6)
πN
b 1 , n2 ; N ) as entropy of the event
Let us define S(Ω(n1 , n2 ; N )) = ln Ω(n
Ω(n1 , n2 ; N ). For n1 = n2 = N/2 we get an event with largest entropy. Let
us denote it by the symbol SB (N ).
Let SG = S(Ω(N )) = N ln 2 denote the entropy of the sure event. It
is the logarithm of the number of outcomes of the experiment of tossing N
independent and fair coins.
Taking logarithm of the terms on both sides of Eq. (2.6) we get,
1 1
SB = SG − ln N − ln(2/π) (2.7)
2 2
∼ 1
N →∞ SG − ln N (2.8)
2
T
The entropy of the sure event is of the order of N ; the entropy of the event
with N/2 ”Heads” and N/2 ’Tails” is less by an extremely small quantity
AF
of the order of ln(N ). Hence when you toss a very large number of coins
independently you will get almost always N/2 ’Heads’ and N/2 ’Tails”.
For example take a typical value for N = 1023 . We have SG = 0.69 × 1023
and SB = 0.69 × 1023 − 24.48.
DR
Table 2.5. Probability that in a toss of N coins, the number of Heads deviates by
more than one percent from N/2
ǫ =
1/100; Pout =
N N N N
N 1-P 2
−ǫ 2
≤n≤ 2
+ǫ 2
103 0.752
4
10 0.317
5
10 0.002
10 6
1.5 × 10−23
107 2.7 × 10−2174
{T, H, H, H, T, H, H, T, H, T }.
Notice that the ensemble contains ten elements. Six elements are H and four
are T. This is consistent with the given probabilities: P(H) = 6/10; and
T
P(T ) = 4/10.
However a Gibbs ensemble is constructed by actually tossing N identical
AF
and independent coins. In the limiet N → ∞, sixty percent of the coins shall
be in micro state H and forty, T. To ensure this we need to take the size of
the ensemble N , to be very large. How large ? You will get an answer to this
question in what follows.
DR
0.25
0.3
0.2
0.25
0.15
B(n)
B(n)
0.2
0.1 0.15
0.1
0.05
0.05
0 0
0 5 10 0 5 10
n n
N!
Fig. 3.1. Binomial distribution : B(n; N ) = pn (1 − p)N−n with N = 10;
n!(N − n)!
B(n; N ) versus n; depicted as sticks; (Left) p = 0.5; (Right) p = .35.
N
X N
X N!
hni = n B(n; N ) = n T pn q N −n ,
n=0 n=1
n! (N − n)!
N
AF
X (N − 1)!
= Np pn−1 q N −1−(n−1) ,
n=1
(n − 1)! [N − 1 − (n − 1)]!
N
X −1
(N − 1)!
DR
= Np pn q N −1−n ,
n=0
n!(N − 1 − n)!
= N p(p + q)N −1 = N p.
Second factorial moment of n : The second factorial moment of n
is defined as hn(n − 1)i. It is calculated as follows.
N
X
hn(n − 1)i = n(n − 1)B(n; N ),
n=0
N
X N!
= n(n − 1) pn q N −n ,
n=2
n!(N − n)!
N
X (N − 2)!
= N (N − 1)p2 pn−2 q (N −2)−(n−2) ,
n=2
(n − 2)![(N − 2) − (n − 2)!
N
X −2
(N − 2)!
= N (N − 1)p2 pn q (N −2)−n .
n=0
n![(N − 2) − n]!
3.2 Moment Generating Function 29
We have,
hn(n − 1)i = N (N − 1)p2 ,
hn2 i − hni = N 2 p2 − N p2 ,
hn2 i = N 2 p2 − N p2 + N p,
σn2 = hn2 i − hni2 = N pq. (3.5)
The square-root of variance is called the standard deviation. A relevant quan-
tity is the relative standard deviation. It is given by the ratio of the standard
σn 1 q
T
deviation to the mean. For the Binomial random variable, we have,
r
AF
= √ . (3.6)
hni N p
√
The relative standard deviation is inversely proportional to N . It is small
for large N . It is clear that the number of elements N , in a Gibbs ensemble
DR
Let B(n) denote the probability that n coins are in micro state ”Heads” in
an ensemble of N coins. We have shown that,
N!
B(n) = pn q N −n . (3.7)
n!(N − n)!
The moment generating function is defined as
N
X
e
B(z) = z n B(n). (3.8)
n=0
30 3. Binomial, Poisson, and Gaussian
The first thing we notice is that B(ze = 1) = 1. This guarantees that the
probability distribution B(n) is normalised. The moment generating func-
tion is like a discrete transform of the probability distribution function. We
transform the variable n to z.
Let us now take the first derivative of the moment generating function
with respect to z. We have,
e XN
dB e ′ (z) =
=B n z n−1 B(n),
dz n=0
N
X
e ′ (z) =
zB n z n B(n). (3.9)
n=0
. Substitute in the above z = 1. We get,
e ′ (z = 1) = hni.
B (3.10)
e evaluated at z = 1 generates the first moment.
Thus the first derivative of B
e
Now take the second derivative of B(z) to get
e XN
d2 B
= n(n − 1)z n−2 B(n),
dz 2 n=0
z2
e
d2 B
dz 2
=
XN
z n n(n − 1) B(n).
T (3.11)
AF
n=0
Substitute in the above z = 1 and get,
e
d2 B = hn(n − 1)i, (3.12)
dz 2
DR
z=1
For the Binomial random variable, we can derive the moment generating
function :
XN XN
e n N!
B(z) = z B(n) = (zp)n q N −n , = (q + zp)N . (3.13)
n=0 n=0
n! (N − n)!
The moments are generated as follows.
dBe
= N (q + zp)N −1 p, (3.14)
dz
dBe
hni = = N p, (3.15)
dz z=1
e
d2 B
= N (N − 1)(q + zp)N −2 p2 , (3.16)
dz 2
e
d2 B
hn(n − 1)i = = N (N − 1)p2 . (3.17)
dz 2 z=1
3.3 Binomial → Poisson 31
0 0.0001 6 0.2508
1 0.0016 7 0.2150
2 0.0106 8 0.1209
3 0.0425 9 0.0403
4 0.1115 10 0.0060
5 0.2007 − −
What is the right distribution for this problem and problems of this kind ?
To answer this question, consider what happens to the Binomial distribution
in the limit of N → ∞, p → 0, and N p = µ, a constant1 . Note that
N p = N v/V = ρv = constant.
We shall show below that in this limit, Binomial goes over to Poisson
distribution.
We start with
e
B(z) = (q + zp)N . (3.19)
We can write the above as
N
e zp
B(z) = qN 1 + ,
q
N
zp
= (1 − p)N 1+ , T (3.20)
q
AF
N N
1 zN p 1
= 1 − Np 1+ . (3.21)
N q N
In the above replace N p by µ and q by 1 to get,
DR
µ N zµ N
e
B(z) = 1− 1+ .
N N
In the limit of N → ∞ we have by definition2 ,
e
B(z) ∼ exp(−µ) exp(zµ),
= Pe (z). (3.22)
e
Thus in the limit N → ∞, p → 0 and N p = µ, we find B(z) → Pe (z), given
by
Pe (z) = exp[−µ(1 − z)]. (3.23)
1
Note that for a physicist, large is infinity and small is zero.
2
exponential function is defined as
x N
limit
exp(x) = N→∞ 1+
N
3.4 Poisson Distribution 33
0.35 0.14
0.3 0.12
0.25 0.1
0.2 0.08
0.15 0.06
0.1 0.04
0.05 0.02
0 0
−5 0 5 0 5 T 10 15 20
AF
Fig. 3.2. Poisson distribution with mean µ, depicted as sticks. Gaussian distribu-
tion with mean µ and variance σ 2 = µ depicted by continuous line.
(Left) µ = 1.5; (Right) µ = 9.5. For large µ Poisson and Gaussian coincide
DR
be their sum. Note Y is a random variable and its probabiity density function
is formally given by,
Z +∞ Z +∞
fY (y) = dx1 dx2 · · ·
−∞ −∞
Z N
!
+∞ X
··· dxN f (x1 , x2 , · · · , xN ) δ y− xi (3.37)
−∞ i=1
If the random variables are mutually independent then their joint density can
be written as the product of their individual probability densities, see below.
36 3. Binomial, Poisson, and Gaussian
Z N
!
+∞ X
··· dxN fN (xN ) δ y − xi (3.39)
−∞ i=1
Now, multiply both sides of the equation above by exp(−iky) and integrate
over y from −∞ to +∞. The Left Hand Side gives the characteristic function
of the random variable Y .
Z +∞
dy exp(−iky) fY (y) = φY (k) (3.40)
−∞
Thus we get,
Z +∞ Z +∞ Z +∞
φY (k) = dy exp(−iky) dx1 f1 (x1 ) dx2 f2 (x2 ) · · ·
−∞ −∞ −∞
···
Z +∞
dxN fN (xN ) δ y −
T
N
X
xi
!
(3.41)
AF
−∞ i=1
In the right hand side of the above, the integral over y can be carried out
making use of a property of Dirac-delta function5 , and we get,
DR
5
Z +∞
dx f (x) δ(x − x0 ) = f (x0 )
−∞
3.7 Central Limit Theorem 37
Z +∞ Z +∞
φY (k) = dx1 f1 (x1 ) dx2 f2 (x2 ) · · ·
−∞ −∞
Z +∞
··· dxN fN (xN ) exp[−ik(x1 + x2 + · · · + xN )]
−∞
Z +∞ Z +∞
= dx1 exp(−ikx1 )f1 (x1 ) dx2 exp(−ikx2 )f2 (x2 ) · · ·
−∞ −∞
Z +∞
··· dxN exp(−ikxN )f (xN )
−∞
Carry out the power series expansion of the exponential function and get,
"∞ #
X (−ik)n
Pe (k; µ) = exp µ . (3.47)
n=1
n!
We recognize the above as the cumulant expansion of a distribution for which
T
all the cumulants are the same µ. For large value value of µ, it is adequate to
AF
consider only small values of k. Hence we retain only terms upto quadratic
in k. Thus for k small, we have,
e k2
P (k) = exp −ikµ − µ . (3.48)
2!
DR
3.9 Gaussian
A Gaussian of mean µ and variance σ 2 is given by
1 (x − µ)2
G(x) = √ exp − . (3.49)
σ 2π 2σ 2
The characteristic Rfunction is given by the Fourier transform formally ex-
e +∞
pressed as G(k) = −∞ dx exp(−ikx)G(x). The integral can be worked out,
e
and I leave it as an exercise. We get, G(k) = exp −ikµ − k 2 σ 2 /2 . Consider
a Gaussian of mean zero and variance σ 2 . It is given by
3.9 Gaussian 39
1 1 x2
g(x) = √ exp − 2 . (3.50)
σ 2π 2σ
The width of the Gaussian distribution is 2σ. The Fourier transform of g(x)
is denoted ge(k) and is given by
1
g (k) = exp − k 2 σ 2 .
e (3.51)
2
The Fourier transform is also a Gaussian with zero mean and standard devi-
ation 1/σ 2 . The width of e
g(k) is 2/σ. The product of the width of g(x) and
the width of its Fourier transform eg(k) is 4.
If g(x) is sharply peaked then its Fourier transform ge(k) will be broad
and vice versa.
T
AF
DR
DR
AF
T
4. Isolated System: Micro canonical Ensemble
4.1 Preliminaries
the container. When a billiard ball bounces off the wall, it transmits a certain
momentum to the wall and Bernoulli imagined it as pressure. It makes sense.
First consider air contained in a cube of side one meter. There is a certain amount
of pressure felt by the wall. Now imagine the cube length to be doubled with
out changing the speeds of the molecule. In modern language this assumption
is the same as keeping the temperature constant. The momentum transferred
per collision remains the same. However since each billiard ball molecule has
to travel twice the distance between two successive collisions with the wall, the
force on the wall should be smaller by a factor of two. Also pressure is force per
unit area. The area of the side of the cube is four times more now. Hence the
pressure should be less by a further factor of four. Taking into account both these
factors, we find the pressure should be eight times less. We also find the volume
of cube is eight times more. From these considerations, Bernoulli concluded that
the product of pressure and volume must be a constant when there is no change
in the molecular speeds - a brilliant argument based on simple scaling ideas.
42 4. Isolated System: Micro canonical Ensemble
Fig. 4.1. Two ways of keeping a particle in a box divided into two equal parts.
ǫ = V /2
b V b = kB ln(2).
Ω(V, N = 1, ǫ = V /2) = = 2 and, S = kB ln Ω (4.1)
ǫ
Now consider distributing two distinguishable particles in these two cells each
of volume ǫ = V /2, see Fig. (4.2). We then have
T
AF
DR
Fig. 4.2. Four ways of keeping two distinguishable particles in a box divided into
two equal halves. ǫ = V /2.
2
b V b = 2kB ln(2). (4.2)
Ω(V, N = 2, ǫ = V /2) = = 4 andS = kB ln Ω
ǫ
For N particles we have,
N
b V b = N kB ln(2). (4.3)
Ω(V, N, ǫ = V /2) = = 2N , and S = kB ln Ω
ǫ
Let us now divide the volume equally into V /ǫ parts and count the number of
N
b V
ways or organizing N (distinguishable) particles. We find Ω(V, N) = ,
ǫ
and
4.3 Ideal Gas Law : Derivation 43
b = N kB ln(V /ǫ)
S = kB ln Ω
= N kB ln V − N kB ln ǫ. (4.4)
We have then,
∂U ∂U
T = ; P =− .
∂S V ∂V S
Let us now consider S ≡ S(U, V ), a natural starting point for statistical
mechanics. We have,
∂S ∂S
dS = dU + dV
∂U V ∂V U
To express the partial derivatives in the above in terms of T and P , we rearrange
the terms in the first law equation (4.6) as,
1 P
dS = dU + dV.
T T
Equating the pre-factors of dU and dV in the above two equation, we get,
∂S 1 ∂S P
= ; = .
∂U V T ∂V U T
44 4. Isolated System: Micro canonical Ensemble
b
S(V, N ) = kB ln Ω(V, N ),
= kB [N ln V − N ln N + N − N ln ǫ] ,
V
= N kB ln + N kB − N kB ln ǫ. (4.13)
N
In the above, I have expressed ln N ! = N ln N − N , employing Stirling’s first
formula for a large factorial. We find that this prescription of Boltzmann
restores the extensivity of entropy; i.e. we find
S(λV, λN ) = λS(V, N ). (4.14)
Boltzmann counting, at best, can be considered as a patch work. You don’t
pull down a well-built wall because there is a small crack in it. Instead you
cover the crack by pasting a paper over it. A good formalism is not dismissed
because of a flaw3 . You look for a quick fix. Boltzmann counting provides
one. In fact non extensivity of entropy is a pointer to a deeper malady. The
fault is not with statistical mechanics but with classical formalism employed
to describe ideal gas. For the correct resolution of the Gibbs’ paradox we
have to wait for the arrival of quantum mechanics.
potential energy is zero since the particles do not interact with each other.
The kinetic energy is given by
3N
X p2i
E= . (4.15)
i=1
2m
The system is thus confined to the surface of a 3N dimensional sphere. We
need a formula for the volume of an hyper-sphere in 3N dimensions. To this
end we need to know of Heaviside4 theta function and Dirac5 delta function.
Define a function
ǫ
0 for −∞ ≤ x ≤ − ;
2
1 1 ǫ ǫ
f (x; ǫ) = x + for − ≤ x ≤ + ; (4.16)
ǫ 2 2 2
ǫ
1 for + ≤ x ≤ +∞.
2
T
AF
where ǫ > 0.
Define
lim.
Θ(x) = ǫ → 0 f (x; ǫ).
DR
Θ(x) is called the step function, Heaviside step function, unit step function
or theta function. It is given by,
0 for −∞ ≤ x < 0;
Θ(x) = (4.17)
1 for 0 < x ≤ +∞.
Figure (4.3) depicts f (x; ǫ) for ǫ = 2, 1, 1/2 and the theta function obtained
in the limit of ǫ → 0.
Start with the function f (x; ǫ) defined by Eq. (4.16). Take the derivative of
the function. We find that the derivative is 1/ǫ, when −ǫ/2 < x + ǫ/2 and
zero otherwise, Define,
4
Oliver Heaviside(1850-1925)
5
Paul Adrien Maurice Dirac(1902-1984)
4.9 Dirac and his δ Function 47
1 1
0.8 0.8
0.6 0.6
0.4
0.4
0.2
0.2
0
0
−2 −1 0 1 2 −2 −1 0 1 2
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 T 0
−2 −1 0 1 2 −2 −1 0 1 2
AF
Fig. 4.3. f (x; ǫ) versus x. (Top Left) ǫ = 2; (Top Right) ǫ = 1; (Bottom Left)
ǫ = 1/2; (Bottom Right) Θ(x)
DR
0 for −∞ ≤ x < −ǫ/2;
df
g(x; ǫ) = = 1 for −ǫ/2 < x + ǫ/2; (4.18)
dx
ǫ
0 for +ǫ/2 < x ≤ +∞.
Figure(4.4) depicts g(x; ǫ) for ǫ = 2, 1, 1/2, 1/4.
The Dirac-delta function is defined as,
δ(x) = limit
ǫ→0 g(x; ǫ). (4.19)
Consider the following integral.
Z +∞
I= dx g(x; ǫ). (4.20)
−∞
We find that the integral is the same for all values of ǫ. This gives us an
important property of the Dirac-delta function:
48 4. Isolated System: Micro canonical Ensemble
4 4
3 3
2 2
1 1
0 0
−2 −1 0 1 2 −2 −1 0 1 2
4 4
3 3
2 2
1 1
0 T 0
−2 −1 0 1 2 −2 −1 0 1 2
AF
df
Fig. 4.4. g(x; ǫ) = versus x. (Top Left) ǫ = 2; (Top Right) ǫ = 1; (Bottom
dx
Left) ǫ = 1/2; (Bottom Right) ǫ = 1/4
DR
Z +∞
dx δ(x) = 1. (4.21)
−∞
Let us demonstrate how to use the theta function and delta function to derive
an expression for a circle of radius R. Let us denote the area of a circle by
the symbol V2 (R) - the ’volume’ of a two dimensional ’sphere’ of radius R.
A little thought will tell you,
Z +∞ Z +∞ 2
!
X
2 2
V2 (R) = dx1 dx2 Θ R − xi . (4.22)
−∞ −∞ i=1
Let
yi = xi /R for i = 1, 2.
4.10 Area of a Circle 49
Then,
Z Z 2
!
+∞ +∞ X
2 2
V2 (R) = R dy1 dy2 Θ R (1 − yi2 . (4.23)
−∞ −∞ i=1
We have
Θ(λx) = θ(x) ∀ λ > 0.
Therefore,
Z Z 2
!
+∞ +∞ X
2
V2 (R) = R dy1 dy2 Θ 1 − yi2 (4.24)
−∞ −∞ i=1
= R2 V2 (R = 1). (4.25)
We can now write Eq. (4.22) as
Z +∞ Z 2
!
+∞ X
2
V2 (R = 1)R = dx1 dx2 Θ R2 − x2i . (4.26)
−∞ −∞ i=1
Now differentiate both sides of the above equation with respect to the variable
R. We have already seen that the derivative of a Theta function is the Dirac-
delta function. Therefore
Z +∞ Z +∞ 2
!
X
V2 (R = 1)2R = 2R dx1
−∞
dx2 δ R −2
−∞
2
xi . T (4.27)
i=1
AF
Now multiply both sides of the above equation by exp(−R2 )dR and integrate
over the variable R from 0 to ∞. We get,
Z ∞ Z ∞ Z +∞
2 2
V2 (R = 1) exp(−R )2RdR = exp(−R ) 2RdR dx1
DR
0 0 −∞
Z 2
!
+∞ X
2
dx2 δ R − x2i . (4.28)
−∞ i=1
Z ∞ Z ∞ Z ∞
V2 (R = 1) dt exp(−t) = dx1 dx2 exp(−x21 − x22 ). (4.29)
0 −∞ −∞
Z +∞ 2
V2 (R = 1) × 1 = dx exp(−x2 ) . (4.30)
−∞
Z ∞ 2 Z ∞ 2
V2 (R = 1) = 2 dx exp(−x2 ) = dx x−1/2 exp(−x) . (4.31)
0 0
Z ∞ 2
V2 (R = 1) = x(1/2)−1
exp(−x)dx = [Γ (1/2)]2 = π. (4.32)
0
Thus V2 (R) = V2 (R = 1) × R2 = πR2 , a result we are all familiar with.
50 4. Isolated System: Micro canonical Ensemble
{xi : i = 1, N } to {yi = xi /R : i = 1, N }.
dxi = Rdyi ∀ i = 1, N ;
" N
#! N
!
X X
2
Θ R 1− yi2 = Θ 1− yi2 .
i=1 i=1
We have,
Z Z Z N
!
+∞ +∞ +∞ X
VN (R) = RN dy1 dy2 · · · T dyN Θ 1 − yi2 ; (4.34)
−∞ −∞ −∞ i=1
AF
= VN (R = 1)RN . (4.35)
DR
Differentiate both sides of the above expression with respect to R and get,
Z +∞ Z +∞
N −1
N VN (R = 1)R = dx1 dx2 · · ·
−∞ −∞
Z N
!
+∞ X
··· dxN δ R2 − x2i 2R. (4.37)
−∞ i=1
Now, multiply both sides by exp(−R2 )dR and integrate over R from 0 to ∞.
The Left Hand Side:
Z ∞
LHS = N VN (R = 1) dR exp(−R2 )RN −1 . (4.38)
0
4.11 Volume of an N -Dimensional Sphere 51
N N
= VN (R = 1) Γ ,
2 2
N
=Γ + 1 VN (R = 1). (4.39)
2
The Right Hand Side :
Z ∞ Z +∞ Z +∞
RHS = dR exp(−R2 ) dx1 dx2 · · ·
0 −∞ −∞
Z N
!
+∞ X
2
··· dxN δ R − x2i 2R; (4.40)
−∞ i=1
t = R2 ; dt = 2RdR,
Z ∞ Z Z
T
AF
+∞ +∞
RHS = dt exp(−t) dx1 dx2 · · ·
0 −∞ −∞
Z N
!
+∞ X
dxN δ t− x2i ,
DR
−∞ i=1
Z +∞ Z +∞ Z +∞
= dx1 dx2 · · · dxN exp −(x21 + x22 + · · · x2N ) ,
−∞ −∞ −∞
Z ∞ N
2
= dx exp(−x )
−∞
= π N/2 . (4.41)
Thus we get
π N/2
VN (R = 1) = . (4.42)
Γ N2 + 1
π N/2
VN (R) = RN . (4.43)
Γ N2 + 1
52 4. Isolated System: Micro canonical Ensemble
Let us measure the volume of the phase space in units of h3N , where h is
Planck constant. We have
∆x∆px ≥ h.
Thus h3N is the volume of a ”minimum uncertainty” cube. Thus we have
b V N (2πmE)3N/2
Ω(E, V, N ) = 3N . (4.44)
h Γ 3N
2 +1
T
AF
4.13 Density of States
Let g(E) denote the density of (energy) states. g(E)dE gives the number of
micro states with energy between E and E + dE. In other words,
DR
Z E
b
Ω(E) = g(E ′ )dE ′ . (4.45)
0
From the above, we find
!
b
∂ Ω(E, V, N )
g(E, V, N ) = . (4.46)
∂E
V,N
b
Let us take the partial derivative of Ω(E, V, N ) with respect to E and get,
V N (2πm)3N/2 3N (3N/2)−1
g(E, V, N ) = E . (4.47)
h3N Γ ( 3N
2 + 1)
2
Let us substitute N = 1 in the above and get the single particle density of
states, g(E, V ) as,
V π
g(E, V ) = (8m)3/2 E 1/2 . (4.48)
h3 4
4.14 Entropy of an Isolated System 53
In the limit of N → ∞, the volume of a thin outer shell tends to the volume
of the whole sphere. This intriguing behaviour is a consequence of the power
law behaviour.
VN (R) − VN (R − ∆R) RN − (R − ∆R)N
= ,
VN (R) RN
N
∆R
= 1− 1− = 1 for N → ∞. (4.49)
R
Consider the case with R = 1 and ∆R = 0.1. The percentage of the total
volume contained in the outermost shell of an N dimensional sphere for
N = 1, 2, · · · 10 is given in the table below. . Hence in the limit of N → ∞
VN (R = 1) − VN (R = 0.9) VN (R = 1) − VN (R = 0.9)
N × 100 N × 100
VN (R = 1) VN (R = 1)
1 10.000% 8
T 57.000%
AF
2 19.000% 9 61.000%
3 27.000% 10 65.000%
4 34.000% 20 88.000%
5 41.000% 40 99.000%
6 47.000% 60 99.000%
DR
7 52.000% 80 99.980%
100 99.997%
the number of micro states with energy less than or equal to E is nearly the
same as the number of micro states with energy between E − ∆E and E.
b V N 1 (2πmE)3N/2
Ω(E, V, N ) = 3N . (4.51)
h N ! Γ 3N 2 +1
The corresponding entropy is then,
V 3 E 3 4πm 5
S(E, V, N ) = N kB ln + ln + ln + . (4.52)
N 2 N 2 3h2 2
It is clear that the expression for entropy given above is extensive.
4.15.3 Pressure
The pressure of an isolated system of ideal gas as a function of E, V , and N ,
is given by,
∂S P N kB
= = , (4.55)
∂V E,N T V
N kB T
P = . (4.56)
V
Substituting in the above T as a function of E, V , and N , see Eq. (4.53), we
get,
2E
P = . (4.57)
3V
3N
X
6 p2i
For an ideal gas H = . There are 3N quadratic terms in the Hamiltonian.
i=1
2m
4.15 Properties of an Ideal Gas 55
See the expression for P given in Eq.(4.56). We have the ideal gas law, P V =
N kB T.
N 2 h2
V
= −kB T ln + 3kB T ln(Λ). (4.61)
N
where Λ is the thermal or quantum wavelength9 given by,
7
Chemical potential for an isolated system; it is expressed as a function of E, V ,
and N
8
chemical potential of a closed system : it is expressed as a function of T , V , and
N.
9
Consider a particle with energy kB T . We have
p2 √
E = kB T = ; p2 = 2mkB T ; p = 2mkB T .
2m
The de Broglie wavelength associated a particle having momentum p is
h h
Λ= = √ .
p 2mkB T
56 4. Isolated System: Micro canonical Ensemble
h
Λ= √ . (4.62)
2πmkB T
Let the number density be denoted by ρ. We have thus ρ = N/V . We can
write the chemical potential in a compact form, as follows.
V
µ = −kB T ln + 3kB T ln(Λ),
N
= kB T ln ρ + kB T ln Λ3 ,
= kB T ln(ρΛ3 ). (4.63)
• When the density of particles is small, and temperatures are high, then
ρΛ2 << 1. Note that the thermal wavelength Λ is inversely proportional
to (square root of ) temperature. Hence it is small at high temperatures.
Hence at high temperatures and/or low densities, the chemical potential
is negative and large.
• At ρΛ3 = 1, the chemical potential is zero.
• At low temperatures and high densities, ρΛ3 >> 1. The chemical po-
tential is positive.
Figure (4.5) shows typical classical behaviour of chemical potential with
T
change of temperature.
AF
0.6
DR
0.4
0.2
µ 0
−0.2
−0.4
−0.6
−0.8
−1
0 0.5 1
T 1.5
Fig. 4.5. µ versus T for a classical ideal gas. µ = −(3T /2) ln(T ). We have set
kB = h = ρ/(2πm)3/2 = 1
4.16 Quantum Counting of Micro states 57
ρ
I have set kB = h = = 1 for producing the above graph.
(2πm)3/2
From the figure we observe that at high temperatures µ is negative; at low
temperatures it is positive; as T decreases µ increases, reaches a maximum
and then decreases to zero at zero temperature.
We have done classical counting of micro states and showed that for an iso-
lated system of a single particle confined to a volume V , the number of micro
states with energy less than ǫ, is given by
3/2
b V ) = V (2πmǫ)
Ω(ǫ, , (4.64)
h3 Γ ((3/2) + 1)
We have obtained the above by substituting N = 1 in the equation,
b 1 V N (2πmE)3N/2
Ω(E, V, N ) = .
h3N N! 3N
Γ +1
2
10
10
Now let us do quantum counting of micro states T .
A fundamental entity in quantum mechanics is the wave function ψ(q, t), where
AF
q is the position vector. The wave function is given a physical interpretation that
the point q at time t. Since the system has to be somewhere, for, otherwise we
would not be interested in it, we have the normalization,
Z
ψ ⋆ (q, t)ψ(q, t)dq = 1,
where the integral is taken over the entire coordinate space - each of the x, y
and z coordinates extending from −∞ to +∞.
A central problem in quantum mechanics is the calculation of ψ(q, t) for the
system of interest. We shall be interested in the time independent wave function
ψ(q) describing a stationary states of the system.
How do we get ψ(q) ?
Schrödinger gave a prescription : Solve the equation Hψ(q) = Eψ(q), with ap-
propriate boundary conditions.
We call this the time independent Schrödinger equation.
H is the Hamiltonian operator
~2
H=− ▽2 +U (q). (4.65)
2m
The first operator on the right is kinetic energy and the second the potential
energy.
58 4. Isolated System: Micro canonical Ensemble
Let me tell you how to obtain the energy eigenvalues without invoking
Schrödinger equation. Consider a particle confined to a one dimensional box
of length L. We recognise the segment L must contain integral number of
half wave lengths - so that the wave function vanishes at the boundaries of
the one dimensional box. In other words,
λ
L = n× : n = 1, 2, , · · · (4.68)
2
2L
λ= : n = 1, 2, , · · · (4.69)
n
Substitute the above in the de Broglie relation
h h
p= = n : n = 1, 2, · · · .
λ 2L
This yields
p2 h2
ǫn = = n2 : n = 1, 2, · · · .
2m 8mL2
Consider a particle in an L × L × L cube - a three dimensional infinite well.
The energy of the system is given by T
h2
ǫnx ,ny ,nz = (n2 + n2y + n2z ),
AF
8mL2 x
where nx = 1, 2, · · · , , ny = 1, 2, · · · , and nz = 1, 2 · · · , .
E in the Schrödinger equation is a scalar ... a real number... called energy. It
DR
The ground state is (nx , ny , nz ) = (1, 1, 1); it is non degenerate; the energy
eigenvalue is
3h2
ǫ1,1,1 = . (4.70)
8mL2
The first excited state is three-fold degenerate. The corresponding energy
eigenvalue is
3h2
ǫ2,1,1 = ǫ1,2,1 = ǫ1,1,2 = . (4.71)
4mL2
We start with,
h2
ǫ= (n2 + n2y + n2z ).
8mL2 x
We can write the above as,
8mL2 ǫ
n2x + n2y + n2z = = R2 . (4.72)
h2
(nx , ny , nz ) represents a lattice point in the three dimensional space. The
equation n2x + n2y + n2z = R2 says we need to count the number of lattice
points that are at a distance R from the origin. It is the same as the number
of lattice points that are present on the surface of a sphere of radius R in the
T
positive quadrant; note that the x, y, and z coordinates of the lattice points
are all positive. It is difficult to count the number of lattice points lying on
AF
the surface of a sphere. Instead we count the number of points contained in
a thin spherical shell. To calculate this quantity we first count the number of
points inside a sphere of radius
1/2
DR
8mL2 ǫ
R=
h2
b
and take one-eighth of it. Let us denote this number by Ω(ǫ). We have,
3/2
b 1 4 π 8mL2 ǫ
Ω(ǫ) = π R3 = . (4.73)
8 3 6 h2
We recognize V = L3 and write the above equation as
b V ) = V π (8mǫ)3/2 = V 4π (2mǫ)3/2 ,
Ω(ǫ,
h3 6 h3 3
V 4π (2πmǫ)3/2 V (2πmǫ)3/2
= 3 3/2
= 3 √ ,
h 3 π h (3/2)(1/2) π
V (2πmǫ)3/2 V (2πmǫ)3/2
= 3
= 3 . (4.74)
h (3/2)(1/2)Γ (1/2) h Γ ((3/2) + 1)
60 4. Isolated System: Micro canonical Ensemble
The above is exactly the one we obtained by classical counting, see Eq. (4.64)
Notice that in quantum counting of micro states, the term h3 comes naturally,
while in classical counting it is hand-put11 .
b V ) with
The density of (energy) states is obtained by differentiating Ω(ǫ,
respect to the variable ǫ. We get
V π
g(ǫ, V ) =(8m)3/2 ǫ1/2 . (4.75)
h3 4
The important point is that the density of energy states is proportional to
ǫ1/2 .
b
S(E = 2ǫ, N = 2) = kB ln Ω(E
T
entropy of the two-particle system with energy E = 2ǫ is given by
= 2ǫ, N = 2) = kB ln(3).
AF
Now add a particle, labelled C, such that the energy of the system does
DR
Table 4.2. LEFT : Three micro states of a two-particles system with total energy
2ǫ. RIGHT : Six micro states of a three-particles system with total energy 2ǫ
0 ǫ 2ǫ
0 ǫ 2ǫ A,B - C
B,C - A
A - B
C,A - B
B - A
A B,C -
- A,B -
B C,A -
C A,B -
b
not change. Let Ω(E = 2ǫ, N = 3) denote the number of micro states of the
11
We wanted to count the phase space volume. We took h3 as the volume of a
six-dimensional cube. We considered the six-dimensional phase space (of a single
particle) as filled with non-overlapping exhaustive set of such tiny cubes. We
have to do all these because of Boltzmann ! He told us that entropy is logarithm
of number of micro states. We need to count the number of micro states.
12
See G Cook and R H Dickerson, Understanding the chemical potential, American
Journal of Physics 63(8), 737 (1995)
4.17 Chemical Potential : Toy Model 61
b
S(E = 2ǫ, N = 3) = kB ln Ω(E = 2ǫ, N = 3) = kB ln(6).
In other words, µ is the change in energy of the system when one particle
is added in such a way that the entropy and volume of the system remain
unchanged.
Let us now remove ǫ amount of energy from the three-particles system
and count the number of micro states. Table(4.3) shows that the number of
micro states of the three particles system with energy 2ǫ − ǫ = ǫ is three.
Table 4.3. Three micro states of a three-particles system with total energy ǫ
0 ǫ
A,B C T
B,C A
C,A B
AF
A closed system is one which does not exchange material with the surround-
ings. However, it does exchange energy. It is in thermal contact with the
surroundings. We idealize the surroundings as a ”heat bath”1 .
Thus, a closed system in thermal equilibrium, is characterized by T , V
and N . The system is not isolated. Hence its micro states are not all equi-
probable.
Let us now impose the condition that the three dice add to 6. Under
this condition, let us enquire if the six micro states of the dice are equally
probable.
Let P (k) denote the probability that the red die shows up k given the
three dice add to 6. Because of the condition imposed, the red die can be in
any one of the four states, {1, 2, 3, 4} only; and these four micro states are
not equally probable. The probabilities can be calculated as follows.
We find that for the three-dice system, there are 10 micro states with the
property that the three dice add to 6. These are listed in Table(1).
1
A ”heat bath” transacts energy with the system, but its temperature does not
change.
2
Herbert B Callen, Thermodynamics and an Introduction to Thermostatistics,
Second Edition, Wiley (2006)
3
All micro states are equally probable : Ergodicity.
64 5. Closed System : Canonical Ensemble
Table 5.1. The ten micro states of Callen’s three-dice system with the constraint
that they add to ten
No. W R W No W R W
1 1 1 4 6 2 2 2
2 2 1 3 7 3 2 1
3 3 1 2 8 1 3 2
4 4 1 1 9 2 3 1
5 1 2 3 10 1 4 1
The ten micro states are equally probable. These are the micro states of the uni-
verse - which consists of the system and its surroundings. The universe constitutes
an isolated system.
Of these ten micro states of the universe, there are four micro states for which
the system die shows up in its microstate 1; therefore P (1) = 0.4. Similarly we can
calculate the other probabilities :
The important point is that the system is not cound in all its microstates with
equal probabiity. T
From the above toy model, we can say that if we consider the system and its
AF
surroundings together to constitute the universe and demand that the universe has
a fixed energy, then the system will not be in its micro states with equal probability.
What is the probability of a micro state of a closed system ? We shall calculate
the probability in the next section employing two different methods. The first in-
volves Taylor expansion of S(E). I learnt of this, from the book of Balescu4 . The
DR
C
X
α exp [−βE(C)] = 1 , (5.6)
C
1 X
Q(T, V, N ) = = exp [−βE(C)] . (5.7)
α C
where, β = 1/[kB T ], and the sum runs over all the micro states of a closed system
at temperature T , volume V and number of particles N . Let Ω(E, b V, N ) denote
b
the density of (energy) states. In other words Ω(E, V, N )dE is the number of micro
states having energy between E and E + dE. The canonical partition function, see
Eq. (5.8), can be written as an integral over energy,
Z ∞
Q(β, V, N ) = b
dE Ω(E, V, N ) exp [−βE(V, N )] . (5.9)
0
We see that the canonical partition function is a ’transform’ of the density of states .
The ”variable” energy is transformed to the ”variable” temperature. The transform
helps us go from a micro canonical (ensemble) description (of an isolated system)
with independent variables E, (V and N ) to a canonical (ensemble) description (of
a closed system) with independent variables T , (V , and N . The density of states is
a steeply increasing function of E. The exponential function exp(−βE) decays with
E for any finite value of β. The decay is steeper at higher value of β or equivalently
at lower temperatures. The product shall be, in general, sharply peaked at a value
of E determined by β.
When β is small (or temperature is large) the integrand would peak at a large
value of E. When β is high (at low temperatures) it would peak at a low value of
E.
molecules. Consider,
Z ∞
Q(β, V, N ) = b
dE Ω(E, V, N ) exp [−βE(V, N )] . (5.10)
0
In the above replace the integral over E by the value of the integrand, evaluated at
E = hEi = U . We get,
b
Q = Ω(E b
= U, V, N ) exp(−βU ) ; ln Q = ln Ω(U, V, N ) − βU ,
b
−kB T ln Q = U − T kB ln Ω(U, V, N )
= U − T S(U, V, N ). (5.11)
7
We identify the right hand side of the above as (Helmholtz) free energy :
1 ∂S
F (T, V, N ) = U − T S(U, V, N ) ; = . (5.12)
T ∂V V,N
7
Legendre Transform : Start with U (S, V, N ). Concentrate on the dependence of
U on S. Can this dependence be described in an equivalent but alternate way ?
Take the slope of the curve U (S) at S. We have
5.4 Helmholtz Free Energy 67
∂U
T (S, V, N ) = .
∂S V,N
We can plot T against S; but then it will not uniquely correspond to the given
curve U (S). All parallel curves in the U -S plane shall lead to the same T (S).
It is the intercept that will tell one curve from the other, in the family. Let us
denote the intercept by the symbol F . We have
U (S, V, N ) − F ∂U
= T ; F (T, V, N ) = U (S, V, N ) − T S; T = .
S ∂S V,N
The equation for micro canonical temperature, is inverted and entropy is ex-
pressed as a function of T , V , and N . Employing the function S(T, V, N ) we get
U (T, V, N ) and F (T, V, N ). The above is called Legendre transform. S transforms
to ’slope’ T and U (S) transforms to the intercept F (T ). We call F (T, V, N ) as
Helmholtz free energy.
Example: Consider U expressed as a function of S, V , and N :
S3
U (S, V, N ) = α .
NV
We get
∂U 3S 2
T = =α .
∂S V,N NV
Inverting, we get, r
S=
NV T
3α
.
T
AF
We then get internal energy as a function of temperature,
r 3/2
NV T
U= .
3α 3
DR
We have,
P
T
AF
i Ei exp(−βEi )
hEi = P ,
i exp(−βEi )
1 X
= Ei exp(−βEi ) ,
Q i
DR
1 ∂Q ∂ ln Q
=− =− . (5.13)
Q ∂β ∂β
We identify hEi with the internal energy, usually denoted by the symbol U in
thermodynamics.
We have,
1 ∂Q
U =− , (5.14)
Q ∂β
2
∂U 1 ∂2Q 1 ∂Q
=− + = − hE 2 i − hEi2 = −σE
2
. (5.15)
∂β V Q ∂β 2 Q ∂β
Now write,
∂U ∂U ∂T
= × = CV (−kB T 2 ) . (5.16)
∂β ∂T ∂β
We get the relation between the fluctuations of energy of an equilibrium system and
the reversible heat required to raise the temperature of the system by one degree
Kelvin :
2
σE = kB T 2 C V . (5.17)
5.6 Canonical Partition Function : Ideal Gas 69
The left hand side of the above equation represents the fluctuations of energy when
the system is in equilibrium. The right hand side is about how the system would
respond when you heat it8 . Note CV is the amount of reversible heat you have to
supply to the system at constant volume to raise its temperature by one degree
Kelvin. The equilibrium fluctuations in energy are related to the linear response;
i.e. the response of the system to small perturbation9 .
N
b= V 1 (2πmE)3N/2
Ω . (5.19)
N ! h3N Γ ( 3N
2
+ 1)
Therefore the density of (energy) states is given by,
T
b
∂Ω V N 1 (2πm)3N/2 3N 3N
g(E) = = E 2 −1 ,
AF
∂E N ! h3N Γ ( 3N
2
+ 1) 2
V N 1 (2πm)3N/2 3N
= E 2 −1 (5.20)
DR
N ! h3N Γ ( 3N
2
)
where we have made use of the relation
3N 3N 3N
Γ +1 = Γ . (5.21)
2 2 2
The partition function is obtained as a ”transform” of the density of states where
the variable E transformed to the variable β.
Z
V N 1 (2πm)3N/2 ∞ 3N
Q(β, V, N ) = dE exp(−β E) E 2 −1 . (5.22)
N ! h3N Γ 3N 2 0
Let,
8
Notice that σ 2 is expressed in units of Joule2 . The quantity kB T 2 is expressed
in units of Joule-Kelvin. CV is in Joule/Kelvin. Thus kB T 2 CV has units of
Joule2 .
9
first order perturbation.
70 5. Closed System : Canonical Ensemble
Γ ( 3N
2
)
= . (5.26)
β 3N/2
Substituting the above in the expression for the partition function we get,
VN 1 VN 1
Q(T, V, N ) = 3N
(2πmkB T )3N/2 = , (5.27)
N! h N ! Λ3N
where
h
Λ= √ . (5.28)
2πmkB T
First check that Λ has the dimension of length. Λ is called thermal wavelength
or quantum wavelength. It is approximately
√ the de Broglie wavelength of a particle
with energy p2 /2m = kB T . We have p = 2mkB T . Therefore
h h h
Λ= = √ ≈ √ .
p 2mkB T 2πmkB T
A wave packet can be constructed by superimposing plane waves of wavelengths
in the neighbourhood of Λ. Thus a wave is spatially localized in a volume of the
T
order of Λ3 . Consider a gas with number density ρ. The inter particle distance is of
the order of ρ1/3 . Consider the situation ρΛ3 << 1. The particles are far apart.
AF
The wave packets do not overlap. Classical description will suffice. Quantum effects
manifest when ρΛ3 >> 1 : when density is large and temperature is low.
DR
mean value; the fluctuations, however, are extremely small. What remains constant
is the chemical potential, µ.
The above observations hold good for energy also. Energy is not a constant. It
fluctuates around a mean value; the fluctuations are small; what remains fixed is
the temperature, T .
Let A denote the number of cubes contained in the big cube.
The universe - the big cube, has a certain amount energy say E and certain
number of molecules N and a certain volume V and these quantities are constants.
You can immediately see that what we have is Gibbs’ grand canonical ensemble
of open systems : each small cube represents an open system. and is a member of a
grand canonical ensemble. All the members are identical as far as their macroscopic
properties are concerned. This is to say the volume V , the temperature T and
chemical potential µ are all the same for all the members.
Now, let us imagine that the walls are made impermeable to movement of
molecules across. A cube can not exchange matter with its neighbouring cubes.
Let us also assume that each cube contains exactly N molecules. However energy
in a cube is not fixed. Energy can flow from one cube to its neighbouring cubes.
This constitutes a canonical ensemble11 .
Aim :
To find the probability for the closed system to be in its micro state i.
First, we list down all the micro states of the system. Let us denote the micro
states as {1, 2, · · · }. Note that the macroscopic properties T , V , and N are the
same for all the micro states. In fact the system switches from one micro state to
another, incessantly. Let Ei denote the energy of the system when it is in micro
state i. The energy can vary from one micro state to another.
T
To each cube, we can attach an index i. The index i denotes the micro state
of the closed system with fixed T , V and N . An ordered set of A indices uniquely
specifies a micro state of the universe.
AF
Let us take an example. Let the micro states of the closed system be denoted by
the indices {1, 2, 3}. There are only three micro states. Let us represent the isolated
system by a big square and construct nine small squares, each of which represents
a member of the ensemble. Each square is attached with an index which can be 1,
DR
2 3 3
2 3 1
11
Usually, canonical ensemble is constructed by taking a system with a fixed value
of V and N and assembling a large number of them in such a way that each is in
thermal contact with its neighbours. Usually these are called mental copies of the
system. The system and its mental copies are then isolated. The isolated system
constitutes the universe. All the mental copies are identical macroscopically in
the sense they all have the same value of T , V and N . Also other macroscopic
properties defined as averages of a stochastic variable e.g. energy are also the
same for all the mental copies. But these mental copies will often differ, one from
the other, in their microscopic properties.
72 5. Closed System : Canonical Ensemble
In the above micro state, there are two squares with index 1, three with index
2 and four with index 3. Let {a1 = 2, a2 = 3, a3 = 4} be the occupation number
string of the micro state. There are several micro states having the same occupation
number string. I have given below a few of them.
Table 5.3. A few micro states with the same occupation number representation of
(2, 3, 4). There are 1260 micro states with the same occupation number representa-
tion
Notice that all the micro states given above have the same occupation number
string {2, 3, 4}. How many micro states are there with this occupation number string
? We have
b 3, 4) = 9! = 1260
Ω(2, (5.29)
2!3!4!
A
X
ai (ã)Ei = E ∀ strings ã (5.32)
i=1
Note that the value of pi varies from one string to another. It is reasonable to
obtain the average value of pi over all possible strings ã. We have
X ai (ã)
Pi = P(ã) (5.33)
ã
A
where
5.8 Lagrange method 73
b
Ω(ã)
P(ã) = P . (5.34)
b
ã Ω(ã)
We are able to write the above because all the elements of an an ensemble have the
same probability given by inverse of the size of the ensemble. In the above,
b A!
Ω(ã) = , (5.35)
a1 !a2 ! · · ·
where we have used a simple notation ai = ai (ã) ∀ i = 1, 2, · · · .
b
Let us take a look at Ω(ã) b
for various strings ã. For large A the number Ω(ã)
will be overwhelmingly large for a particular string, which we shall denote as ã⋆ .
We can ensure this by taking A → ∞. Note that A should be large enough so
that even a micro state of smallest probability is present in the ensemble atleast
once.
Thus we can write
b ⋆)
ai (ã⋆ ) Ω(ã
Pi = ,
A Ω(ã b ⋆)
ai (ã⋆ )
= ,
A
⋆
a
= i (5.36)
A
T b
Thus the problem reduces to finding that string ã⋆ for which Ω(ã) is a maximum.
Of course there are two constraints on the string. They are
X
AF
aj (ã) = A ∀ ã; (5.37)
j
X
aj (ã)Ej = E ∀ ã. (5.38)
j
DR
∂h
=0 (5.40)
∂x
∂h
=0 (5.41)
∂y
We have two equations and two unknowns. In principle we can solve the above two
equations and obtain (x⋆ , y ⋆ ) at which h is maximum.
Now imagine there is a road on the mountain which does not necessarily pass
through the peak of the mountain. If you are travelling on the road, then what is
the highest point you will pass through ? In the equation
∂h ∂h
dh = dx + dy = 0 (5.42)
∂x ∂y
the infinitesimals dx and dy are not independent. You can choose only one of them
independently. The other is determined by the constraint which says that you have
to be on the road.
Let the projection of the mountain-road on the plane be described by the curve
g(x, y) = 0.
∂h ∂h
dh = dx + dy = 0 (5.45)
∂x ∂y
∂g
∂h ∂h
= dx + − ∂x dx = 0, (5.46)
∂x ∂y ∂g
∂y
∂h
∂h ∂y ∂g
= − dx = 0. (5.47)
∂x ∂g ∂x
∂y
We have an equation similar to Eq. (5.48) involving partial derivative with respect
to the variable y, which follows from the definition, see Eq. (5.49), of the Lagrange
undetermined multiplier, λ.
Thus we have two independent equations,
∂h ∂g
−λ = 0, (5.50)
∂x ∂x
∂h ∂g
−λ = 0. (5.51)
∂y ∂y
We can solve and and get x⋆ ≡ x⋆ (λ) and y ⋆ = y ⋆ (λ). The value of x and y at
which h(x, y) is maximum under constraint g(x, y) = 0 can be found in terms of
the unknown Lagrange multiplier λ.
Of course we can determine the value of λ by substituting the solution
(x⋆ (λ), y ⋆ (λ)) in the constraint equation : g(x⋆ (λ), y ⋆ (λ)) = 0.
where
∂f
∂xµ
λ= (5.56)
∂g
∂xµ
There are only N − 1 values of dxi . We have eliminated dxµ . Instead we have the
undetermined multiplier λ. Since dxi : i = 1, N and i 6= µ are all independent of
each other we can set each term in the sum to zero. Therefore
76 5. Closed System : Canonical Ensemble
∂f ∂g
−λ = 0 ∀ i 6= µ (5.57)
∂xi ∂xi
From the definition of λ we get
∂f ∂g
−λ =0 (5.58)
∂xµ ∂xµ
Thus we have a set of N equations
∂f ∂g
−λ = 0 ∀ i = 1, N (5.59)
∂xi ∂xi
There are N equations and N unknowns. In principle we can solve the equation
and get
x⋆i ≡ x⋆i (λ) ∀ i = 1, N,
where the function h is maximum under constraint
g(x1 , x2 , · · · xN ) = 0.
b A!
Ω(ã) = (5.62)
a1 !a2 ! · · ·
The two constraints are
X
aj (ã) = A (5.63)
j
X
aj (ã)Ej = E (5.64)
j
5.11 Mechanical and Thermal Properties 77
b
We extremize entropy given by ln Ω(ã).
X X
b 1 , a2 , · · · ) = ln A! −
ln Ω(a aj ln aj + aj (5.65)
j j
Energy of a closed system varies, in general, from one micro state to another.
An equilibrium system visits a sequence of micro states dictated by Newtonian dy-
namics. Averaging over the visited microstates during the observation time gives
us the time average. Instead, invoking ’ergodicity’, we average over a Gibbs’ en-
semble and equate it to the thermodynamics property. Thus, energy averaged over
a canonical ensemble, gives internal energy of thermodynamics, see below.
exp(−βEi )
pi = P (5.71)
i exp(−βEi )
P
X
Pi Ei exp(−βEi )
hEi = Ei pi = (5.72)
i i exp(−βEi )
earlier lecture we talked of the number of micro states of an isolated system. The
isolated system is constructed as follows : Assemble a large number of closed sys-
tems. Each closed system is in thermal contact with its neighbouring closed systems.
Let η = {i : i = 1, 2, · · · } label the micro states of a closed system. Let A denote
the number of closed systems in the ensemble. The ensemble is isolated. We have
thus an isolated system. We call it the universe. We describe a micro state of the
universe by specifying the index i for each if its members. The index comes from
the set η defied earlier. The micro states of the universe system are all equally
probable; we group them and denote a group by specifying the string {a1 , a2 , · · · }
where ai is the number of elements of the ensemble, having the index i. We have
b 1 , a2 , · · · ) = A!
Ω(a (5.73)
a1 !a2 ! · · ·
b For convenience we
The aim is to find {a⋆i : i = 1, 2, · · · } that maximizes Ω.
b
maximize ln Ω. Let us consider the limit ai → ∞ ∀ i. Also consider the variables
pi = ai /A. Then
12
The probability of a microstate in a micro canonical ensemble is pi =
b
1/Ω(E, b = −kB P pi ln pi
V, N ) ∀ i. Hence kB ln Ω i
5.13 Free Energy to Entropy 79
X X
b = A ln A − A −
ln Ω ai ln ai + ai (5.74)
i i
X X
= ai ln A − ai ln ai (5.75)
i
X a
i
=− ai ln (5.76)
A
X
= −A pi ln pi (5.77)
i
b
ln Ω X
=− pi ln pi (5.78)
A i
1
pi = exp(−βEi ) ;
Q
80 5. Closed System : Canonical Ensemble
F = −kB T ln Q
1 X
U = hEi = Ei exp(−βEi )
Q i
and write,
" #
S 1 X
− = β −kB T ln Q − Ei exp(−βEi )
kB Q i
" #
1 X 1 X
= − ln Q exp(−βEi ) − β Ei exp(−βEi )
Q i Q i
X 1
= exp(−βEi ) − ln Q − Ei β
i
Q
X
1 exp(−βEi )
= exp(−βEi ) ln
i
Q Q
X
= pi ln pi (5.83)
i
canonical ensemble for closed system, and grand canonical ensemble for open sys-
tem. Let pi : i = 1 , 2 · · · denote formally an ensemble. pi is the probability of a
micro state of the system under consideration.
For example if the system is isolated, then all micro states are equally probable.
We simply count the number of micro states of the system; let us say there are Ω b
micro states. Inverse of this shall be the probability of a micro state.
For an isolated system pi = 1/Ω. b
1
For a closed system, pi = exp(−βEi ).
Q
1
For an open system pi = exp(−βEi + βµNi )
Q X
We can write, in general, U = pi Ei . This equation suggests that the inter-
i
nal energy of a closed system can be changed by two ways.
1. change {Ei : i = 1, 2, · · · } keeping {pi : i = 1, 2, · · · } the same. This we call
as work. Note that the energy eigenvalue change when we change the boundary
condition, i.e. when we move the boundary thereby changing the volume.
2. change {pi i = 1, 2, · · · } keeping {Ei P: i = 1, 2, · · · } the same. The changes
in pi should be done in such way that i pi = 1. This we call as heat.
5.14 Microscopic View : Heat and Work 81
Thus we have,
X X⋆
dU = pi dEi + Ei dpi (5.84)
i i
where the superscript ⋆ in the second sum should remind us that all dpi s are not
independent and that they should add to zero.
In the first sum we change Ei by dEi ∀ i keeping pi ∀ i unchanged.
In thePsecond sum we change pi by dpi ∀ i keeping Ei unchanged for all i and
ensuring i dpi = 0.
X
5.14.1 Work in Statistical Mechanics : W = pi dEi
i
We have
X X ∂Ei
pi dEi = pi dV
i i
∂V
P
∂ pi Ei
i
= dV
∂V
∂hEi
= dV
∂V T
AF
∂U
= dV
∂V
= −P dV = W (5.85)
DR
82 5. Closed System : Canonical Ensemble
X
5.14.2 Heat in Statistical Mechanics : q = Ei dpi
i
Method-1
We start with the first term on the right hand side of Eq. (5.84).
X
S = −kB pi ln pi , (5.86)
i
X⋆
dS = −kB [1 + ln pi ] dpi , (5.87)
i
X⋆
T dS = d¯q = −kB T ln pi dpi , (5.88)
i
X⋆ e−βEi
= −kB T [−βEi − ln Q] dpi since pi = (5.89)
i
Q
X⋆
= Ei dpi (5.90)
i
Method-2
Alternately, we recognize that change of entropy is brought about by change
of probabilities of microstates and vice versa, keeping their respective energies the
same. Thus,
X X ∂pi
Ei dpi = Ei dS (5.91)
∂S
i
=
∂
i
P
i Ei pi
dS =
∂hEi
dS =
∂U
T
dS = T dS = q (5.92)
AF
∂S ∂S ∂S
In an adiabatic process, the system does not transact energy by heat. Hence the
probabilities {pi } of the micro states do not change during an adiabatic process.
The only way the internal energy can change is through change of {ǫi }. For a
particle confined to a box, quantum mechanics tells us that the energy eigenvalue
is inversely proportional to the square of the length of the box.
1
ǫi ∼ (5.93)
L2
∼ V −2/3 (5.94)
∂ǫi
∼ V −5/3 (5.95)
∂V
First law of thermodynamics tells us dU = d¯Q + d¯W ; For an adiabatic process
d¯Q − 0; if the process is reversile then d¯W = −P dV . Therefore,
5.16 Q(T, V, N ) for an Ideal Gas 83
X ∂ǫi
dU = pi dV (5.96)
i
∂V
−P dV ∼ V −5/3 dV (5.97)
Z +∞
β
··· dp3N exp − p21 + p22 · · · + p23N (5.99)
−∞ 2m
=
VN 1
Z +∞
dp exp −
T
β 2
p
3N
(5.100)
AF
N ! h3N −∞ 2m
Z +∞ 3N
VN 1 1 p2
DR
= dp exp − (5.101)
N ! h3N −∞ 2 mkB T
Consider the integral
Z +∞
1 p2
I = dp exp − (5.102)
−∞ 2 mkB T
since the integrand is an even function, we can write the above as
Z +∞
1 p2
I =2 dp exp − (5.103)
0 2 mkB T
Let
p2
x= (5.104)
2mkB T
Therefore,
p
dx = dp (5.105)
mkB T
r
mkB T 1
dp = dx (5.106)
2 x1/2
84 5. Closed System : Canonical Ensemble
Z
√ ∞
= 2mkB T dx x(1/2)−1 exp(−x)
0
√ 1
= 2mkB T Γ
2
√ √
= 2πmkB T since Γ (1/2) = π (5.107)
The canonical partition function is thus given by,
VN 1
Q(β, V, N ) = (2πmkB T )3N/2 (5.108)
N ! h3N
We write the above as
VN 1
Q(T, V, N ) = (5.109)
N ! Λ3N
h
Λ(T ) = √ . (5.110)
2πmkB T
Energy
T
AF
V
ln Q = N ln + N − 3N ln(Λ) (5.111)
N
∂ ln Q ∂ ln Λ
DR
3N kB T
U = hEi = (5.114)
2
The above shows that energy is equi-partitioned amongst the 3N degrees of freedom;
each degree of freedom carries kB T /2 of energy.
Heat capacity
The heat capacity is given by
∂U 3N kB
CV = =
∂T 2
We have nR = N kB , where n is the number of moles and R the universal gas
constant. Thus heat capacity per mole of the substance - the so-called molar specific
heat is given by,
CV 3R
= (5.115)
n 2
5.16 Q(T, V, N ) for an Ideal Gas 85
T
AF
DR
DR
AF
T
6. Open System : Grand Canonical Ensemble
Let us take the open system, its boundary and surroundings and construct a
universe, which constitutes an isolated system. We are interested in constructing
an isolated system because, we want to start with the only assumption we make in
statistical mechanics : all the micro states of an isolated system are equally probable.
We have called this the ergodic hypothesis.
Let E denote the total energy of the universe. E >> E, where E is the (average
or typical) energy of the open system under study.
Let N denote the total number of particles in the universe. N >> N, where
N is the (average or typical) number of particles in the open system. V is the total
volume of the universe. The universe is characterized by E , N , and V and these are
held at constant values.
We want to describe the open system in terms of its own micro states. Let C
be a micro state of the open system. Let
• E(C) be the energy of the open system when it is in its micro state C and
• let N (C) be the number of particles in the open system when it is in its micro
state C.
1
equality of temperature signals thermal equilibrium.
2
equality of chemical potential ensures diffusional or material equilibrium. Of
course, equality of pressure shows mechanical equilibrium.
88 6. Open System : Grand Canonical Ensemble
When the open system is in micro state C the universe can be in any one of
its innumerable micro states3 such that the energy of the surroundings is E − E(C)
and the number of particles in the surroundings is N − N (C).
Let ΩC = Ω(E − E(C), N − N (C)), denote the subset of micro states of the
universe having a common property given by the following. When the universe is
in a micro state belonging to ΩC , then the system is in the chosen micro state C
and the surroundings can be in any of its microstates having energy E − E(C) and
number of particles N −N (C). Let Ω bC denote the number of elements in the subset
(event) ΩC . Following Boltzmann we define a statistical entropy associated with
the event ΩC as
bC
SC = kB ln Ω
b − E(C), N − N (C))
= kB ln Ω(E
= S(E − E(C), N − N (C)). (6.1)
Since E(C) << E , and N (C) << N , we can Taylor-expand S retaining only
the first two terms. We have
S E − E(C), N − N (C) = S E , N
∂S
−E(C)
∂E
N E,N
−N (C)
∂S
∂N
T
E E,N
(6.2)
AF
From the first law of thermodynamics we have, for a reversible process,
dE = T dS − P dV + µdN (6.3)
from which it follows,
DR
1 P µ
dS = dE + dV − dN (6.4)
T T T
We have
S ≡ S(E, V, N ) (6.5)
∂S ∂S ∂S
dS = dE + dV + dN (6.6)
∂E V,N ∂V E,N ∂N E,V
3
The picture I have is the following. I am visualizing a micro state of the isolated
system as consisting of two parts. One part holds the signature of the open
system; the other holds the signature of the surroundings. For example a string
of positions and momenta of all the particles in the isolated system defines a
micro state. This string consists of two parts. The first part s contains the string
of positions and momenta of all the particle in the open system and the second
part contains the positions and momenta of all the particles in the surroundings.
Since the system is open the length system-string is a fluctuating quantity and
so is the length of bath-string. However the string of the isolated system is of
fixed length. I am neglecting those micro states of the isolated system which hold
the signature of the interaction between the system and the surroundings at the
boundaries.
6.1 What is an Open System ? 89
Comparing the coefficients of dE, dV and dN , in equations (6.4) and (6.6), we get,
∂S 1 ∂S P ∂S µ
= ; = ; =− (6.7)
∂E V,N T ∂V E,N T ∂N E,V T
Therefore,
1 µ
S E − E(C), N − N (C) = S E , N − E(C) + N (C) (6.8)
T T
The probability of the micro state C is given by
b E − E(C), N − N (C)
Ω
P (C) = (6.9)
ΩbTotal
We are able to write the above because of the postulate of ergodicity : All micro
states of the universe - an isolated system, are equally probable. We have,
1 1
P (C) = exp S E − E(C), N − N (C)
ΩbTotal kB
1 S(E , V, N ) E(C) µN (C)
= exp − +
bTotal
Ω kB kB T kB T
T
AF
= α exp[−β {(E(C) − µN (C)}]. (6.10)
In the above, the constant α can be determined by the normalization condition,
X
P (C) = 1, where the sum runs over all the micro states of the open system. We
DR
c
have,
1
P (C) = exp[−β{E(C) − µN (C)}] (6.11)
Q
where the grand canonical partition function is given by
X
Q(T, V, µ) = exp [−β {E(C) − µN (C)}] (6.12)
C
The fugacity is given by λ = exp(βµ); then we can write the grand canonical
partition function as,
X N(C)
Q(T, V, λ) = λ exp[−βE(C)] (6.13)
C
• Collect those micro states of a grand canonical ensemble with a fixed value
of N . Then these micro states constitute a canonical ensemble described the
canonical partition function, Q(T, V, N ).
• Collect all the micro states of a grand canonical ensemble with a fixed energy.
and fixed number of particles. Then these micro states constitute a microcanon-
ical ensemble.
90 6. Open System : Grand Canonical Ensemble
S → T and N → µ. We have,
G(T, V, µ) = U (S, V, N ) − T S − µN (6.17)
∂U ∂U
T = : µ= (6.18)
∂S V,N
∂N S,V
Some authors e.g. Donald A McQuairrie, Statistical Mechanics, Harper and Row
(1976) would like to identify P V as the thermodynamic counter part of the grand
canonical ensemble. The correspondence is
G = −P V = −kB T ln Q.
4
We shall experience this while trying to evaluate the canonical partition function
for fermions and bosons. We will not be able to carry out the sum over occupation
numbers because of the constraint that they add up to a constant N . Hence we
shall multiply the restricted sum by λN and sum over all possible values of N .
This would remove the restriction and we shall express the partition function as
sum over (micro states) of product (over occupation numbers). We shall interpret
λ as fugacity. The chemical potential and fugacity are related : λ = exp(βµ).
All these can be viewed as mathematical tricks. The language of grand canonical
ensemble gives a physical meaning to these mathematical tricks.
6.2 Micro-Macro Synthesis : Q and G 91
where the sum runs over all the micro states i of the open system. Ei is the energy
of the system when in micro state i, and Ni is the number of particles in the system
when in its micro state i.
We replace the sum over micro states by sum over energy and number of par-
ticles. Let g(E, N ) denote the density of states. We have then,
Z Z
Q(T, V, µ) = dE dN g(E, N ) exp[−β(E − µN )] (6.20)
−kB T ln Q = F − µN = U − T S − µN = G (6.28)
Recall the discussions on Legendre Transform : We start with U ≡ U (S, V, N ).
Transform S in favour of the ”slope” T (partial derivative of U with respect to S).
We get the ”intercept” F (T, V, N ) as U − T S. Let us carry out one more transform
: N → µ. i.e. transform N in favour of the slope µ (partial derivative of U with
respect to N ); µ is the chemical potential. We get the ”intercept” G(T, V, µ) - the
grand potential. We have G(T, V, µ) = U − T S − µN . Thus we have, G(T, V, µ) =
−kB T ln Q(T, V, µ)
92 6. Open System : Grand Canonical Ensemble
∂U ∂U ∂U
S +V +N = U (S, V, N ) (6.30)
∂(λS) ∂(λV ) ∂(λN )
T
The above is true for any value of λ. In particular it is true for λ = 1. Substitute
in the above λ = 1 and get,
AF
∂U ∂U ∂U
S +V +N = U (S, V, N ) (6.31)
∂S ∂V ∂N
T S − P V + µN = U (6.32)
DR
We proceed as follows.
U = T S − P V + µN, (6.34)
dU = T dS − P dV + µdN + SdT − V dP + N dµ (6.35)
From the first law of thermodynamics, we have dU = T dS − P dV + µdN . Hence,
N dµ + SdT − V dP = 0. It follows then,
S V
dµ = − dT + dP = −sdT + vdP (6.36)
N N
where s is the specific entropy - entropy per particle and v is specific volume -
volume per particle.
Q ∂λ ∂λ
N=0 N=0
N!
= exp(λQ1 ) (6.39)
The average of N can be formally expressed as,
∞ N
1 X (λQ1 )N X λN Q N
1
hN i = N = (6.40)
Q N=0 N! N=0
Q N !
ζN
P (N ) thus is a Poisson distribution : P (N ) = exp(−ζ. For Poisson distribution,
N!
the mean equals variance. Therefore
2
σN = hN 2 i − hN i2 = hN i = ζ = λ Q1 . (6.43)
σ 2 = hN 2 i − hN i2 .
In the above
c denotes a micro state of the open system
T
AF
E(c) denotes the energy of the open system when in micro state c
N (c) denotes the number of particles of the open when in micro state c
Let us now take the partial derivative of all the terms in the above equation, with
respect to the variable µ, keeping the temperature and volume constant. We have,
DR
∂Q X
= βN (c) exp − β{E(c) − µN (c)} = βhN iQ(T, V, µ) (6.45)
∂µ T,V c
2 " #
∂ Q ∂Q ∂hN i
= β hN i +Q (6.46)
∂µ2 T,V ∂µ T,V ∂µ T,V
7
The left hand side of the above equation equals β 2 hN 2 iQ.
Substituting this in the above, we get,
7
X
∂Q
= βN (c) exp − β{E(c) − µN (c)}
∂µ T,V c
2
X
∂ Q
= β 2 [N (c)]2 exp − β{E(c) − µN (c)}
∂µ2 T,V c
= β 2 hN 2 iQ.
6.3 Statistics of Number of Particles 95
∂hN i
β 2 hN 2 iQ = β 2 hN i2 Q + βQ (6.47)
∂µ T,V
2 2 2 ∂hN i
σ = hN i − hN i = kB T (6.48)
∂µ T,V
Since hN i = λ Q1 , we have,
∂hN i dλ
= Q1 = βλQ1 (6.49)
∂µ dµ
We have shown
2 ∂hN i
σN = kB T (6.50)
∂µ T,V
∂hN i
Substituting in the above = λβQ1 , we get,
∂µ
2
σN = kB T Q1 βλ (6.51)
= λQ1 = hN i (6.52)
consistent with what we have shown earlier : the distribution of N is Poissonian
with mean=variance=λQ1
In what follows, we shall derive an expression for the fluctuations of N in terms
T
of isothermal compressibility - a property measurable experimentally.
AF
6.3.7 Number Fluctuations and Isothermal Compressibility
Let us express the number fluctuations in terms of experimentally measurable prop-
erties 8 of the open system. Let us define
DR
V
v= .
hN i
It is called specific volume. It is the volume per particle. We have,
8
We shall show that,
∂hN i hN i2
= kT (6.53)
∂µ T,V V
where kT denotes isothermal compressibility - an experimentally measurable
property. Isothermal compressibility is defined as
1 ∂V
kT = − (6.54)
V ∂P T
96 6. Open System : Grand Canonical Ensemble
∂hN i ∂(V /v)
= (6.55)
∂µ T,V ∂µ T,V
V ∂v
=− 2 (6.56)
v ∂µ T,V
hN i2 ∂v
=− (6.57)
V ∂µ T,V
∂hN i
σ 2 = kB T (6.63)
∂µ T,V
hN i2
= kB T kT (6.64)
V
σ2 kB T
= kT (6.65)
hN i2 V
9
Gibbs - Duhem relation reads as hN idµ = V dP − SdT. At constant temperature,
we have,
hN idµ = V dP,
∂P hN i
= (6.59)
∂µ T V
6.3 Statistics of Number of Particles 97
σ 2 = kB T
∂hN i
T
Start with a fluctuation-dissipation relation 12 derived earlier,
AF
(6.66)
∂µ V,T
We consider the reciprocal,
∂µ
DR
(6.67)
∂hN i V,T
V
dµ = dP (6.69)
hN i
∂µ V ∂P
= (6.70)
∂hN i T,V
hN i ∂hN i T,V
∂µ V
= (6.71)
∂P T,V hN i
hN i
Let us now define ρ = , which denotes the particle density : number of particles
V
per unit volume. We have,
∂P ∂P 1 ∂P
= = (6.72)
∂hN i V,T ∂(ρV ) V,T V ∂ρ V,T
The density can be changed either by changing hN i and/or V . Here we change ρ
infinitesimally, keeping V constant. As a result the pressure changes infinitesimally.
Let me repeat : both these changes happen at constant V and T .
As far as P is concerned, it does not care whether ρ has changed by change of
hN i or of V . Hence it is legitimate to write
∂P ∂P
= (6.73)
∂ρ V,T ∂(hN i/V ) hNi,T
=−
V2
∂P
T (6.74)
AF
hN i ∂V hNi,T
Thus we get,
∂µ V ∂P
=
DR
1 ∂P
=
hN i ∂ρ V,T
V2 ∂P
=− (6.75)
hN i2 ∂V hNi,T
2
σN kB T
= kT (6.77)
hN i2 V
6.4 Energy Fluctuations 99
1 ∂2Q
hE 2 i =
Q ∂β 2 λ,V
2
= −σE:OPEN
We proceed in the same fashion and obtain
a similar expression for the fluctuations
∂hEi 2
of energy in a canonical ensemble14 = −σE:CLOSED . Note for an open
∂β N,V
system, the partial derivative of energy with respect to β is taken at constant λ
and V . For a closed system, the partial derivative of energy with respect to beta is
2 2
taken at constant N and V . What is the relation between σE:OPEN and σE:CLOSED ?
The key to answering to this question is in the following entities : • U (T, V, N )
with N constant and
• U (T, V, λ) = U (T, V, N (T, V, λ)) with N depending on T , V , and λ and with λ
kept constant instead of N .
The central relation15 is,
X ∂Q X ∂2Q X 2
14
Q= exp(−βEi ); =− Ei exp(−βEi ); 2
= Ei exp(−βEi );
i
∂β i
∂β i
2
1 ∂Q ∂hEi 1 ∂2Q 1 ∂Q 2
hEi = − ; =− + = −σE:CLOSED
Q ∂β ∂β Q ∂β 2 Q ∂β
15
Equation (6.78) above can be derived from first principles. Formally we start
with U (T, V, λ) ≡ U (T, V, N (T, V, λ)). We have,
100 6. Open System : Grand Canonical Ensemble
∂U ∂U ∂U ∂N
= + × . (6.78)
∂T λ,V ∂T N,V ∂N T,V ∂T λ,V
∂U U (T + ∆T, V, N (T + ∆T, V, λ)) − U (T, V, N (T, V, λ))
= ∆Tlimit
→0
∂T λ,V ∆T
In the above we express
∂N
N (T + ∆T, V, λ) = N (T, V, λ) + ∆T = N + (∆N )λ,V ,
∂T λ,V
∂N
where we have used the short-hand notation (∆N )λ,V = ∆T . There-
∂T λ,V
fore
∂U U (T + ∆T, V, N + (∆N )λ,V ) − U (T, V, N )
= ∆Tlimit
→0
∂T λ,V
∆T
We can write
∂U
U (T + ∆T, V, N + (∆N )λ,V ) = U (T, V, N ) + ∆T
∂T V,N
∂U
+(∆N )λ,V
∂N T,V
Therefore, T
∂U ∂U (∆N )λ,V ∂U
= + ×
AF
∂T λ,V ∂T V,N ∆T ∂N T,V
∂U ∂N ∂U
= + ×
∂T V,N ∂T λ,V ∂N T,V
DR
6.4 Energy Fluctuations 101
The last term in Eq. (6.78) above involving the partial derivative of N with respect
to T can be evaluated, see box below.
∂N ∂N ∂µ
=− ×
∂T µ
∂µ T
∂T N
" #
∂N ∂µ ∂µ
= −
DR
" #
∂N µ ∂µ
= − (6.80)
∂µ T,V
T ∂T N,V
∂µ µ
In the last step we have made use of the relation : = , see box
∂T λ,V T
below.
∂µ
λ = exp(βµ); µ = kB T ln λ; = kB ln λ = µ/T
∂T λ
102 6. Open System : Grand Canonical Ensemble
Consider the term within square brackets on the right hand side of Eq. (6.80);
we can show that " #
µ ∂µ 1 ∂U
− = ,,
T ∂T N,V T ∂N T,V
see box below.
In the above, the last step follows from one of Maxwell’s relations, see box below.
Maxwell’s Relation
F = U − TS
dF = −P dV + µdN − SdT
T
= µdN − SdT (at constant V )
AF
We have
∂F ∂F
=µ ; = −S
∂N T ∂T N
∂2F ∂µ ∂2F ∂S
DR
= ; =−
∂T ∂N ∂T N ∂N ∂T ∂N T
2 2
∂ F ∂ F ∂µ ∂S
= ⇒ =−
∂T ∂N ∂N ∂T ∂T N ∂N T
∂N ∂N 1 ∂U
Therefore Eq. (6.80) can be written as, = .
∂T λ,V ∂µ T,V T ∂N T,V
Now we can go back to Eq. (6.78) and write it as
2
∂U ∂U 1 ∂U ∂N
= + (6.81)
∂T λ,V ∂T N,V T ∂N T,V ∂µ T,V
6.4 Energy Fluctuations 103
2
∂U ∂U ∂U
kB T 2 = kB T 2 2
+ σN
∂T λ,V
∂T N,V
∂N T,V
2
2 ∂hEi
σE:OPEN = kB T 2 C V + 2
σN (6.82)
∂hN i T,V
Thus the energy fluctuations in an open system has an extra component given by
the second term in the above, arising due to fluctuations in the number of particles.
T
AF
DR
16
X
Q= exp[−β(E − µN )]
i
∂Q X ∂2Q X 2
=β N exp[−β(E − µN )]; 2
= β2 N exp[−β(E − µN )]
∂µ i
∂µ i
1 ∂Q 1 ∂2Q
= βhN i; = β 2 hN 2 i
Q ∂µ Q ∂µ2
2
∂hN i ∂ 1 ∂Q 1 ∂2Q 1 ∂Q
β = = − = β 2 hN 2 i − hN i2
∂µ ∂µ Q ∂µ Q ∂µ2 Q ∂µ
∂hN i 1 2
= σ
∂µ kB T N
DR
AF
T
7. Quantum Statistics
This provides a description of an open system with its natural variable : temper-
ature, volume and chemical potential. The system can exchange energy as well as
matter with the environment. Thus energy E, and the number of particles N , of
the system would be fluctuating around the means hEi and hN i respectively, where
the angular bracket denotes averaging over a grand canonical ensemble of the open
system.
Thus we have,
∞
X X⋆
Q(T, V, µ) = λN exp [−β(n1 ǫ1 + n2 ǫ2 + · · · )]
N=0 {n1 ,n2 ,··· }
∞
X X⋆
= λn1 +n2 +··· exp [−β(n1 ǫ1 + n2 ǫ2 + · · · )]
N=0 {n1 ,n2 ,··· }
∞
X X⋆
= exp [−β(ǫ1 − µ)n1 − β(ǫ2 − µ)n2 + · · · )]
N=0 {n1 ,n2 ,··· }
∞
X X⋆
= [exp {−β (ǫ1 − µ)}]n1 ×
N=0 {n1 ,n2 ,··· }
∞
X X⋆
Consider evaluating the sum, I1 = xn1 n2
1 x2 , where ni = 0, 1, 2 for
N=0 {n1 ,n2 }
i = 1, 2, and ⋆ ⇒ n1 + n2 = N. Write down all the terms in this sum explicitly and
2 X
Y 2
show it is equivalent to xn
i ..
i=1 n=0
Let n1 and n2 denote two integers. Each of them can take values 0, 1, or 2.
Consider first the restricted sum :
X∞ X⋆ n n
I1 = x1 1 x2 2 (7.4)
N=0 {n1 ,n2 }
where the star over the summation sign reminds us of the restriction n1 + n2 = N .
Thus we get Table (1).
7.2 Open System and Q(T, V, µ) 107
N n1 n2 xn1 n2
1 x2
0 0 0 1
1 1 0 x1
0 1 x2
2 2 0 x21
0 2 x22
1 1 x1 x2
3 3 0 -
2 1 x21 x2
1 2 x1 x22
0 3 -
- - -
2 2 x21 x22
- - -
5 - - -
Table 1. Terms in the restricted sum
T
AF
From the table, we can write down I1 as,
I1 = 1 + x1 + x2 + x21 + x22 + x1 x2 + x21 x2 + x1 x22 + x21 x22 (7.5)
Now consider the product over the unrestricted sums,
DR
2 X
Y 2
I2 = xn
i (7.6)
i=1 n=0
2
Y
= (1 + xi + x2i )
i=1
= (1 + x1 + x21 ) × (1 + x2 + x22 )
= 1 + x2 + x22 + x1 + x1 x2 + x1 x22 + x21 + x21 x2 + x21 x22 = I1 (7.7)
In fact we need not have restricted n1 and n2 to the values 0, 1, and 2. We
could have allowed them to run from zero and infinity, see below.=
108 7. Quantum Statistics
∞ X
X ∞ X
X N
⋆
xn1 n2
1 x2 = xn N−n
1 x2
{n1 ,n1 }
N=0 N=0 n=0
∞
X N
X n
x1
= xN
2
N=0 n=0
x2
X∞
1 − (x1 /x2 )N+1
= xN
2
N=0
1 − (x1 /x2 )
X∞
xN+1
2 − x − 1N+1
=
N=0
x2 − x1
1 x2 x1
= −
x − 2 − x1 1 − x2 1 − x1
1 x2 (1 − x1 ) − x − 1(1 − x2 )
=
x2 − x1 (1 − x1 )(1 − x2 )
1 1
= (7.8)
1 − x1 1 − x2
QED
T
AF
We can now write the grand canonical partition function of N quantum particles
occupying single-particle quantum states determined by volume V .
all
YX all
YX
Q(T, V, µ) = xn
i = [λ exp(−βǫi )]n
DR
i n=0 i n=0
all
YX n
= exp {−β (ǫi − µ)} (7.9)
i n=0
b 1 , n2 , · · · ) = N!
Ω(n (7.12)
n1 !n2 ! · · ·
We then have
∞
X X⋆ N!
QCS (T, V, µ) = λN [exp(−βǫ1 )]n1 ×
N=0
n1 !n2 ! · · ·
{n1 ,n2 ··· }
[exp(−βǫ2 )]n2 × · · ·
∞
" #N ∞
X N
X X
= λ exp(−βǫi ) = λN [Q1 (T, V )]N (7.13)
N=0 i N=0
∞
X X ⋆ [λ exp(−βǫ1 )]n1 [λ exp(−βǫ2 )]n2
= × × ···
N=0 {n1 ,n2 ,··· }
n1 ! n2 !
110 7. Quantum Statistics
∞
X X ⋆ [exp(−β(ǫ1 − µ)]n1 [exp(−β(ǫ2 − µ)]n2
= × × ···
N=0 {n1 ,n2 ,··· }
n1 ! n2 !
∞
X X ⋆ xn1 xn2
1 2
= ···
n1 ! n2 !
N=0 {n1 ,n2 ,··· }
∞
! ∞
!
X xn
1
X xn
2
= ···
n=0
n! n=0
n!
= exp(x1 ) exp(x2 ) · · ·
Y Y
= exp(xi ) = exp[λ exp(−βǫi )]
i i
Y
= exp[exp{−β(ǫi − µ)}] (7.14)
i
We can also express the grand canonical partition function for classical indistin-
guishable ideal gas as,
X
QM B = exp(x1 ) exp(x2 ) · · · = exp xi
X
i
T
AF
= exp exp[−β(ǫi − µ)]
i
X
= exp[λ exp(−βǫi )]
i
DR
QM B (T, V, N = 1)N
QM B (T, V, N ) =
N!
∂U ∂U
T = ; µ= (7.22)
∂S V,N ∂N S,V
number of particles in the system. In fact, in the last section on grand canonical
ensemble, we have derived formal expressions for the mean and fluctuations of
the number of particles in an open system; we have related the fluctuations to
isothermal compressibility - an experimentally measurable property.
The grand potential for the three statistics is given by,
G(T, V, µ) = −kB T ln Q
P
−kB T i exp[−β(ǫi − µ)] Maxwell − Boltzmann
P
= kB T i ln [1 − exp {−β(ǫi − µ)}] Bose − Einstein (7.24)
P
−kB T i ln [1 + exp {−β(ǫi − µ)}] Fermi − Dirac
X
=λ exp(−βǫi ) = λQ1 (T, V, µ) (7.28)
i
In the above Q1 is the single-particle canonical partition function. The above result
on hN i is consistent with formal result from grand canonical formalism :
The grand canonical partition function is formally given by the transform of
canonical partition function. The thermodynamic variable N is transformed to fu-
gacity λ (or to chemical potential µ = kB T ln λ. Thus we have,
∞
X
Q(T, V, µ) = λN Q(T, V, N )
N=0
X∞
QN
= λN , 1
N=0
N!
X∞
QN
1
= exp(βµN )
N=0
N!
= exp(λQ1 ) (7.29)
The grand potential is given by
7.8 Average number of particles, hN i 113
Y 1
Q(T, V, µ) = (7.32)
i
1 − exp[−β(ǫi − µ)]
X
ln Q = − ln [1 − exp {−β(ǫi − µ)}] (7.33)
i
X
∂G exp[−β(ǫi − µ)]
hN i = − = (7.34)
∂µ T,V i
1 − exp[−β(ǫi − µ)]
Q=
Y
1 + exp[−β(ǫi − µ)]
T (7.35)
AF
i
X
G = −kB T ln Q = −kB T ln [1 + exp{−β(ǫi − µ)}] (7.36)
DR
X
∂G exp[−β(ǫi − µ)]
hN i = − = (7.37)
∂µ T,V i
1 + exp[−β(ǫi − µ)]
Let us say we keep µ fixed and change the temperature4 . The mean number
of particles in the system changes when temperature changes. To restore the value
of hN i to the fixed value N , we change the chemical potential. In what follows,
we shall investigate the behaviour of particles under Maxwell-Boltzmann, Bose-
Einstein, and Fermi-Dirac statistics employing grand canonical formalism but with
a fixed N ; This implies µ is not any more an independent property; it is a function
of T .
I must say there is nothing unphysical about this strategy. We are studying
a physical system enclosed by a non-permeable wall - a wall that does not permit
particle exchange. The chemical potential µ, is a well defined property of the system.
It is just that µ is not any more under our control5 . The system automatically selects
the value of µ depending on the temperature.
= kB T ln(ρΛ3 ) (7.50)
µ
= ln(ρΛ3 ) (7.51)
kB T
where ρ = N/V is the number density, i.e. number of particles per unit volume.
Thus we get,
ǫi T
ηi = − ln(ρΛ3 ) (7.52)
kB T
AF
We have earlier shown that classical limit obtains when ρΛ3 → 0. Thus in this limit,
ηi → ∞ or exp(−ηi ) → 0. Let xi = exp(−ηi ). Let us express the grand canonical
partition function for the three statistics in terms of the small parameter xi .
exp(xi ) Maxwell − Boltzmann
DR
Y 1
Bose − Einstein
Q= 1 − xi (7.53)
i
1 + xi Fermi − Dirac
6
In thermodynamics we have
F = U − TS (7.44)
dF = dU − T dS − SdT (7.45)
From the first law of thermodynamics we have dU = T dS − P dV + µdN . Sub-
stituting this in the expression for dF above we get,
dF = −P dV + µdN − SdT (7.46)
Therefore,
∂F
µ= (7.47)
∂N T,V
116 7. Quantum Statistics
Y
∼
λ→0 (1 + λ exp(−βǫi )) (7.58)
i
Y
∼
λ→0 [1 + λ exp(−βǫi )] (7.60)
i
Thus in the limit of high temperatures and low densities Maxwell Boltzmann statis-
tics and Bose Einstein statistics go over to Fermi - Dirac statistics.
b 1 , n2 , · · · ) = 1
7.9.3 Easiest Method Ω(n
We could have shown easily that in the limit of high temperature and low density,
the three statistics are identical by considering the degeneracy factor
1 Bose − Einstein and Fermi − Dirac statistics
b
Ω= (7.62)
1
Maxwell − Boltzmann statistics
n1 !n2 ! · · · T
When the temperature is high the number of quantum states that become available
AF
for occupation is very large; When the density is low the number of particles in the
system is low. Thus we have a very few number of particles occupying a very large
of quantum states. In other words the number of quantum states is very large
compared to the number of particles.
Hence the number of micro states with ni = 0 or 1 ∀ i are overwhelmingly large
DR
Let hnk i denote the average occupation number; the averaging is done over
a grand canonical ensemble of micro states. Formally we have for Bosons and
Fermions,
YX all all
X all
n X
xi nxn
k nxn
k
i6=k n=0 n=0 n=0
hnk i = = all
(7.63)
Y all
X all
X X
xn
i
xn
k
xn
k
i6=k n=0 n=0 n=0
7
Consider
1
S(x) = 1 + x + x2 + · · · = (7.65)
1−x
dS 1
= 1 + 2x + 3x2 + 4x3 + · · · = (7.66)
dx (1 − x)2
dS x
x = x + 2x2 + 3x3 + · · · = (7.67)
dx (1 − x)2
7.11 Mean occupation : Some Remarks 119
In the above the summation in the numerator and the denominator are evaluated
analytically as follows. We start with the definition,
X∞
xn
= exp(x) (7.70)
n=0
n!
Differentiate both sides of the above equation with respect to x. You get
X∞ ∞ ∞
nxn−1 1 X nxn X nxn
exp(x) = = = (7.71)
n=0
n! x n=0 n! n=0
n!
Therefore,
xk exp(xk ) 1
hnk iM B = = xk = exp[−β(ǫk − µ)] = (7.72)
exp(xk ) exp[β(ǫk − µ)]
We see that for Fermi-Dirac statistics the occupation number never exceeds unity.
When ǫk − µ is negative and |ǫk − µ| is large the value of hnk i tends to unity. For
ǫ = µ and T 6= 0, we have hnk i = 1/2.
It is also clear that at high temperature the chemical potential of a Fermion must
be negative, because its behaviour coincides with that of Bosons 8 and classical
indistinguishable particles9 .
1.5
←− M axwell − Boltzmann
1
← Bose − E instein
hnk i
0.5
F ermi − Dirac →
0
−0.5
−1
T
AF
−3 −2 −1 0 1 2 3 4 5
ǫk − µ
kB T
DR
7.11.4 At High T and/or Low ρ all Statistics give the same hnk i
When
ǫk − µ
→ ∞,
kB T
all the three statistics coincide. We have already seen that at high temperatures
classical behaviour obtains. Then, the only way
7.12 Occupation Number : Distribution and Variance 121
ǫk − µ
kB T
can become large at high temperature (note in the expression T is in the denomi-
nator) is when µ is negative and its magnitude also should increase with increase
of temperature.
Thus for all the three statistics, at high temperature, the chemical potential µ
is negative and its magnitude must be large.
But then we know µ
= ln(ρΛ3 ).
kB T
This means that ρΛ3 << 1 for classical behaviour to emerge10 . This is in complete
agreement with our earlier surmise that classical behaviour obtains at low ρ and/or
high T . Hence all the approaches are consistent with each other and all the issues
fall in place.
In the above P (n) ≡ P (nk = n) is the probability that the random variable nk
takes a value n. Comparing the above with the first equation, we find
1
P (n) ≡ P (nk = n) = all
xn
k (7.76)
X
xm
k
m=0
where the denominator ensures that the total probability is normalized to unity.
Here, I shall work out explicitly P (n) for Fermi-Dirac, Bose-Einstein, and
Maxwell-Boltzmann statistics. We shall calculate the variance of the random vari-
able nk ; we shall derive the relative fluctuations : the standard deviation divided
by the mean, for the three statistics.
10
Note that ln(x) = 0 for x = 1 and is negative for x < 1. As x goes from 1 to 0,
the quantity ln(x) goes from 0 to −∞.
122 7. Quantum Statistics
1
X
hn2k i = n2 P (n) = ζ (7.80)
n=0
consistent with the result obtained earlier. Inverting the above, we get,
ζ
xk = (7.85)
1+ζ
Then the probability distribution of the random variable nk can be written in a
convenient form
ζn
P (n) = (7.86)
(1 + ζ)n+1
The distribution is geometric, with a constant common ratio ζ/(ζ + 1). We have
come across geometric distribution earlier11 .
Calculation of variance is best done by first deriving an expression for the
moment generating function given by,
∞
X
P̃ (z) = z n P (n) (7.87)
n=0
∞ n
1 X n ζ
= z (7.88)
1 + ζ n=0 1+ζ
=
1
1
T (7.89)
AF
1+ζ ζ z
1−
1+ζ
1
DR
= (7.90)
1 + ζ(1 − z)
Let us now differentiate P̃ (z) with respect to z and in the resulting expression set
z = 1. We shall get hnk i, see below.
∂ P̃ ζ
= (7.91)
∂z (1 + ζ(1 − z))2
∂ P̃
=ζ (7.92)
∂z z=1
Differentiating twice with respect to z and setting z = 1 in the resulting expression
shall yield the factorial moment hnk (nk − 1)i, see below.
11
Let me recollect : The simplest problem in which geometric distribution arises is
in coin tossing. Take a p-coin. Toss the coin until ”H” appears. The number of
tosses is a random variable with a geometric distribution P (n) = q n−1 p. We can
write this distribution in terms of ζ = hni = 1/p and get P (n) = (ζ − 1)n−1 /ζ n .
124 7. Quantum Statistics
∂2P 2ζ 2
= (7.93)
∂z 2 [1 + ζ(1 − z)]3
∂ 2 P
= hnk (nk − 1)i = 2ζ 2 (7.94)
∂z 2 z=1
hn2k i = 2ζ 2 + ζ (7.95)
2
σ = hn2k i − hnk i2 = ζ 2 + ζ (7.96)
r
σ 1
ηBE = = +1 (7.97)
ζ ζ
For doing the problem , You will need the following tricks.
∞
X
S(x) = xn = 1 + x + x2 + x3 + · · ·
n=0
1
= (7.98)
1−x
X∞
dS
= nxn−1 = 1 + 2x + 3x2 + 4x3 + · · ·
dx n=1
1
= (7.99)
(1 − x)2
X∞
dS
x = nxn = x + 2x2 + 3x3 + 4x4 + · · ·
dx
=
n=1
x
T (7.100)
AF
(1 − x)2
X∞
d dS
x = n2 xn−1 = 1 + 22 x + 32 x2 + 42 x3 + · · ·
dx dx n=1
DR
2x 1
= + (7.101)
(1 − x)3 (1 − x)2
X∞
d dS
x x = n2 xn = x + 22 x2 + 32 x3 + 42 x4 + · · ·
dx dx n=1
2x2 x
= + (7.102)
(1 − x)3 (1 − x)2
You can employ the above trick to derive power series for ln(1 ± x), see below.
Z
dx x2 x3 x4
= − ln(1 − x) = x + + + ··· (7.103)
1−x 2 3 4
hnk i = xk (7.104)
xn
k 1
P (nk = n) ≡ P (n) = (7.105)
n! exp(xk )
ζn
= exp(−ζ) (7.106)
n!
The random variable nk has Poisson distribution. The variance equals the mean.
Thus the relative standard deviation is given by
σ 1
ηM B = = √ (7.107)
ζ ζ
We can now write the relative fluctuations for the three statistics in one single
formula as,
+1 for Fermi − Dirac Statistics
r
1
η= − a with a = 0 for Maxwell − Boltzmann Statistics (7.108)
ζ
−1 for Bose − Einstein Statistics
Let us look at the ratio,
P (n)
r= T.
P (n − 1)
For the Maxwell-Boltzmann statistics, r = ζ/n. The ratio r is inversely proportional
AF
to n. This is the normal behaviour; inverse dependence of r on n is what we should
expect.
On the other hand, for Bose-Einstein statistics, the ratio is given by
P (n) ζ
DR
r= =
P (n − 1) ζ+1
r is independent of n. This means, a new particle will get into any of the quantum
states, with equal probability irrespective of how abundantly or how sparsely that
particular quantum state is already populated. An empty quantum state has the
same probability of acquiring an extra particle as an abundantly populated quantum
state.
Thus, compared to classical particles obeying Maxwell-Boltzmann statistics,
Bosons exhibit a tendency to bunch together. By nature, Bosons like to be together.
Note that this ”bunching-tendency” is not due to interaction between Bosons. We
are considering ideal Bosons. This bunching is purely a quantum mechanical effect;
it arises due to symmetry property of the wave function.
For Fermions, the situation is quite the opposite. There is what we may call
an aversion to bunching; call it anti-bunching if you like. No Fermion would like to
have another Fermion in its quantum state.
DR
AF
T
8. Bose-Einstein Condensation
8.1 Introduction
For bosons we found that the grand canonical partition function is given by,
Y 1
Q(T, V, µ) = (8.1)
i
1 − exp[−β(ǫi − µ)]
X
= kB T ln [1 − exp {−β(ǫi − µ)}] (8.2)
i
T
Recall, from thermodynamics, that G is obtained by Legendre transform of
AF
U (S, V, N ) : S → T ; N → µ; and U → G.
∂U ∂U
G(T, V, µ) = U − T S − µN ; T = ; µ= (8.3)
∂S V,N ∂N S,V
DR
X
8.2 hN i = hnk i
k
λ exp(−βǫk )
hnk i = (8.5)
1 − λ exp(−βǫk )
where λ is fugacity. We have
λ = exp(βµ). (8.6)
In the above µ is the chemical potential and equals the energy change due to
addition of a single particle under constant entropy and volume :
∂U
µ= . (8.7)
∂N S,V
The average number of particles is given by,
X X λ exp(−βǫk )
hN i = hnk i = (8.8)
1 − λ exp(−βǫk )
k k
We would like to study Bosonic system with a fixed number of bosons at var-
ious temperatures i.e. a closed system of bosons described by canonical ensemble.
However we would like to employ grand canonical ensemble formalism in our study.
In grand canonical ensemble formalism we can vary T and µ independently by
choosing appropriate heat bath and particle bath. If T is changed at a fixed value
of µ the average number of particles in the system changes.
However if we want to keep N constant, we lose control over µ. When we
change the temperature, the chemical potential should change in such a way that
the average number of particles remains at the value chosen by us. In other words
in a closed system, µ is a function of temperature.
T
In what follows we shall study a closed system of ideal bosons employing the
grand canonical ensemble formalism in which the chemical potential depends on
AF
temperature; the dependence is such that the average number of bosons in the
system remains the same at all temperatures
DR
This gives us
∞
X ∞
X
λ exp(−βǫ)
= λk+1 exp[−β(k + 1)ǫ] = λk exp[−βkǫ]. (8.13)
1 − λ exp(−βǫ)
k=0 k=1
3/2 X
∞ Z ∞
2m exp(−kβǫ)(kβǫ)1/2 d(kβǫ)
= V 2π λk ,
h2 k=1 0 β 3/2 k3/2
3/2 X
∞ Z ∞
2mkB T λk
= V 2π exp(−x)x1/2 dx,
h2 k3/2 0
2mkB T
3/2 X
∞
k=1
λk
T
AF
= V 2π Γ (3/2),
h2 k3/2
k=1
3/2 X∞
2mkB T 1 λk
= V 2π Γ (1/2) ,
DR
h2 2 k=1
k3/2
3/2 ∞
2mkB T √ X λk
=V π π ,
h2 k3/2
k=1
3/2 X
∞
2πmkB T λk
=V . (8.14)
h2 k=1
k3/2
We have earlier defined a thermal wave length denoted by the symbol Λ. This is
the de Broglie wave length associated with a particle having thermal energy of the
order of kB T . It is also called quantum wavelength. It is given by, see earlier notes,
h
Λ= √ . (8.15)
2πmkB T
The sum over k, in the expression for N given by Eq. (8.14) is usually denoted by
the symbol g3/2 (λ):
X∞
λk λ2 λ3
g3/2 (λ) = 3/2
= λ+ √ + √ +··· (8.16)
k=1
k 2 2 3 3
130 8. Bose-Einstein Condensation
V
Thus we get N = g3/2 (λ) We can write it as,
Λ3
N Λ3
= ρΛ3 = g3/2 (λ).
V
(8.17)
It is easily verified, see below, that at high temperature we get results consistent
with Maxwell Boltzmann statistics :
The fugacity λ is small at high temperature. For small λ we can replace g3/2 (λ)
by λ. We get N = λV /Λ3 . This result is consistent with Maxwell-Boltzmann statis-
tics, as shown below.
For Maxwell-Boltzmann statistics, hnk i = λ exp(−βǫk ). Therefore,
X
N = hnk i
k
X
=λ exp(−βǫk )
k
3/2 Z ∞
2m
= λ 2πV dǫ ǫ1/2 exp(−βǫ)
h2 0
3/2 Z ∞
2m (βǫ)1/2 exp(−βǫ) d(βǫ)
= λ 2πV
h2 0 β 3/2
3/2
2mkB T
= λ 2πV Γ (3/2)
h2 T
3/2
2mkB T 1
= λ 2πV Γ (1/2)
AF
h2 2
3/2
2mkB T √
= λ πV π
h2
DR
3/2
2πmkB T
=λV
h2
V
= λ (8.18)
Λ3
The occupancy of all the excited states is given by the second term, where the sum
is taken only over the indices k representing the excited states. Let Ne denote the
occupancy of excited states. It is given by,
X λ exp(−βǫk
Ne = (8.21)
1 − λ exp(−βǫk )
k
In the above, the sum over k can be replaced by an integral over energy. In the
integral over energy, we can still keep the lower limit of integration as zero, since
the density of states giving weight factors for occupancy of states is zero at zero
energy. Accordingly we write
N = N0 + Ne (8.22)
λ V
= + 3 g3/2 (λ) (8.23)
1−λ Λ
We thus have,
1
The chemical potential approaches the energy of the ground state. With out loss
of generality, we can set the ground state at zero energy; i.e. ǫ0 = 0.
132 8. Bose-Einstein Condensation
N Λ3 Λ3 λ
= + g3/2 (λ) (8.24)
V V 1−λ
Let us define the number of density - number of particles per unit volume, denoted
by the symbol ρ. It is given by
N
ρ= (8.25)
V
The function λ/(1 − λ) diverges at λ = 1, as you can see from Figure (8.2).
Hence the relevant curve for carrying out graphical inversion should be the one
that depicts the sum of the singular part (that takes care of the occupancy of the
ground state) and the regular part (that takes care of the occupancy of the excited
states). For a value of Λ3 /V = .05 we have plotted both the curves and their sum
in Figure (8.3). Thus for any value of ρΛ3 we can now determine the fugacity by
graphical inversion.
We carry out such an exercise and obtain the values of λ for various values
of ρΛ3 and Fig. (8.4) depicts the results. It is clear from Fig. (8.4) that when
ρΛ3 > 2.612, the fugacity λ is close unity.
Let us postulate2
a
λ=1− .
N
where a is a number. To determine a we proceed as follows.
We have,
λ
=
N
−1
T (8.26)
AF
1−λ a
N
≈ if N >> a (8.27)
a
DR
We start with,
Λ3 λ
ρΛ3 = + g3/2 (λ) (8.28)
V 1−λ
Sunstitute λ = 1 − a/N in the above and get3 ,
ρΛ3
ρΛ3 = + g3/2 (1) (8.29)
a
Thus we get,
ρΛ3
a= (8.30)
ρΛ3 − g3/2 (1)
2 a
We have reasons to postulate λ = 1− . This is related to the mechanism under-
N
lying Bose-Einstein condensation; we shall discuss the details later. In fact, fol-
lowing Donald A McQuarrie, Statistical Mechanics, Harper and Row (1976)p.173
a
we can make a postulate λ = 1 − . This should also lead to the same conclu-
V
sions.
3
g3/2 (1 − a/N ) ≈ g3/2 (1),
8.6 Bose-Einstein Condensation Temperature 133
Thus λ is less than unity and can be very close to unity; the value of 1 − λ can
be as small as the inverse of the total number of particles in the system. Precisely
1 − λ can be as small as a/N .
The point ρΛ3 = g3/2 (1) = 2.612 is a special point indeed. What is the physical
significance of this point ? To answer this question, consider the quantity ρΛ3
as a function of temperature with ρ kept at a constant value. The temperature
dependence of this quantity is shown below.
3
h
ρΛ3 = ρ √ (8.31)
2πmkB T
At high temperature for which ρΛ3 < g3/2 (1) = 2.612, we can determine the
value of λ from the equation g3/2 (λ) = ρΛ3 by graphical or numerical inversion.
At low temperatures for which ρΛ3 > 2.612, we have λ = 1 − a/N where
ρΛ3
a= (8.32)
ρΛ3 − g3/2 (1)
The quantity λ/(1 − λ) is the number of particles in the ground state. At
temperatures for which ρΛ3 > 2.612, we have,
λ
N0 = (8.33)
1−λ
N
= (8.34)
a T
AF
N0 1
= (8.35)
N a
1
= 1− g3/2 (1) (8.36)
DR
ρΛ3
We can write the above in a more suggestive form by defining a temperature TBEC
by
ρΛ3BEC = g3/2 (1) (8.37)
Therefore,
3 √ 3
N0 1 ρΛ3BEC ΛBEC T
= =1− = 1 − = 1 − √ (8.38)
N a ρΛ3 Λ TBEC
3/2
T
= 1− fer T < TBEC (8.39)
TBEC
We have depicted the behaviour of the fractional number of particles in the ground
state as a function of temperature in Figure (8.5).
Now we shall be careful and separate the singular part and regular part to get,
X
G = kB T ln(1 − λ) + kB T ln[1 − λ exp(−βǫk )] (8.43)
k
In the above, convert the sum over k by an integral over dǫ by the prescription
below,
X 3/2 Z ∞
2m T
(·) −→ V 2π 2
(·) ǫ1/2 dǫ, (8.44)
k
h 0
AF
We get,
G = kB T ln(1 − λ)
DR
3/2 Z ∞
2m
−kB T V 2π dǫ ln (1 − λ exp(−βǫ)) ǫ1/2 (8.45)
h2 0
We have
∞
X
ln[1 − λ exp(−βǫ)] = − λk exp(−kβǫ) (8.46)
k=1
Then we have,
8.8 Energy of Bosonic System 135
G = kB T ln(1 − λ)
3/2 X
∞ Z ∞
2m
−kB T V 2π λk dǫ ǫ1/2 exp(−kβǫ)
h2 0
k=1
= kB T ln(1 − λ) −
3/2 X
∞ Z ∞
2m (kβǫ)1/2 exp(−kβǫ)
kB T V 2π λk d(kβǫ)
h2 0 k3/2 β 3/2
k=1
3/2 X∞
2mkB T λk
= kB T ln(1 − λ) − kB T V 2π Γ (3/2)
h2 k3/2
k=1
3/2 X
∞
2πmkB T λk
= kB T ln(1 − λ) − kB T V
h2 k=1
k3/2
V
= kB T ln(1 − λ) − kB T g3/2 (λ) (8.47)
Λ3
Thus we have,
T
AF
V
G(T, V, λ) = kB T ln(1 − λ) − kB T g3/2 (λ) (8.48)
Λ3
We can also derive4 the above relation employing Grand canonical formalism.
4
An open system is described by a grand canonical partition function. It is for-
mally given by,
X
Q(β, V, µ) = exp[−β(Ei − µNi )] (8.51)
i
In the above Ei is the energy of the open system when in micro state i; Ni is
the number of particles in the open system when in micro state i. Let γ = βµ.
136 8. Bose-Einstein Condensation
Let us now go to continuum limit by converting the sum over micro states by
an integral over energy and get,
3 1
U = V kB T 3 g5/2 (λ) (8.56)
2 Λ
8.8.2 T ≤ TBEC T
For temperatures less that TBEC , the ground state gets populated anomalously. The
AF
bosons in the ground state do not contribute to the energy. For T ≤ TBEC , we have
µ = 0. This means λ = 15 . Substituting λ = 1 in Eq. ((8.56) we get,
3 1
U = V kB T 3 ζ(5/2) (8.59)
2 Λ
DR
Then we get,
X
Q(β, V, µ) = exp(−βEi ) exp(+γNi ) (8.52)
i
1 ∂Q ∂ ln Q
− =− = hEi = U (8.54)
Q ∂β ∂β
For bosons, we have,
Y 1 X X ǫi λ exp(−βǫi )
Q= ln Q = − ln[1 − λ exp(−βǫi )]U = (8.55).
i
1 − λ exp(−βǫi ) i i
1 − λ exp(−βǫi )
5
we have postulated that λ = 1 − O(1/N ) for T ≤ TBEC .
8.9 Specific Heat Capacity of bosons 137
We also have,
N (ΛBEC )3
V = (8.60)
ζ(3/2)
Hence for T < TBEC , we have
3 (ΛBEC )3 ζ(5/2)
U = N kB T
2 Λ3 ζ(3/2)
3/2
3 ζ(5/2) T
= N kB T (8.61)
2 ζ(3/2) TBEC
Thus we have,
3 g5/2 (λ)
N k B T for T > TBEC
2 g3/2 (λ)
U = (8.62)
3/2
3 g5/2 (1) T
N kB T for T < TBEC
2 g3/2 (1) TBEC
8.9.1
CV
N kB
for T > TBEC
T
AF
Let us consider first the case with T > TBEC . We have
U 3 g5/2 (λ)
= T (8.63)
N kB 2 g3/2 (λ)
DR
1 ∂U CV ∂ 3T g5/2 (λ)
= = (8.64)
N kB ∂T N kB ∂T 2 g3/2 (λ)
To work out the derivative in the above, we need the following :
∂ 3
First Relation : g3/2 (λ) = − g3/2 (λ).
∂T 2T
Proof :
We start with ρΛ3 = g3/2 (λ). Therefore,
138 8. Bose-Einstein Condensation
∂ ∂Λ ∂ h
[g3/2 (λ)] = 3 ρΛ2 = 3 ρΛ3 √ (8.65)
∂T ∂T ∂T 2πmkB T
3 h 1
=− ρΛ2 √
2 2πmkB T 3/2
3 h
=− ρΛ2 √
2T 2πmkB T
3 3
=− ρΛ3 = − g3/2 (λ) (8.66)
2T 2T
————–Q.E.D
∂ 1
Second Relation : [gn/2 (λ)] = g(n/2)−1 (λ).
∂λ λ
Proof : ∞
X λk
We have by definition, gn/2 (λ) = . Therefore,
k=1
kn/2
"∞ # ∞ ∞
∂ ∂ X λk X kλk−1 1X λk
[gn/2 (λ)] = n/2
= n/2
= (n/2)−1
∂λ ∂λ k=1 k k=1
k λ k=1 k
T =
1
λ
g(n/2)−1 (λ) (8.67)
AF
— Q.E.D
1 dλ 3 g3/2 (λ)
8.9.2 Third Relation : =−
DR
λ dT 2T g1/2 (λ)
Proof :
We proceed as follows :
∂ ∂ dλ 3 1 dλ
[g3/2 (λ)] = [g3/2 (λ)] : − g3/2 (λ) = g1/2 (λ) (8.68)
∂T ∂λ dT 2T λ dT
1 dλ 3 g3/2 (λ)
From the above we get, =− –Q.E.D
λ dT 2T g1/2 (λ)
We have,
8.9 Heat capacity below TC 139
CV ∂ 3T g5/2 (λ)
=
N kB ∂T 2 g3/2 (λ)
3 g5/2 (λ) 3T ∂ g5/2 (λ)
= +
2 g3/2 (λ) 2 ∂T g3/2 (λ)
" #
3 g5/2 (λ) 3T g5/2 (λ) ∂g3/2 (λ) 1 ∂g5/2 (λ) dλ
= − 2
−
2 g3/2 (λ) 2 g3/2 (λ) ∂T g3/2 (λ) ∂λ dT
" #
3 g5/2 (λ) 3T g5/2 (λ) 3 1 1 dλ
= − 2
− g3/2 (λ) − g3/2 (λ)
2 g3/2 (λ) 2 g3/2 (λ) 2T g3/2 (λ) λ dT
" #
3 g5/2 (λ) 3T g5/2 (λ) 3 1 dλ
= − 2
− g3/2 (λ) −
2 g3/2 (λ) 2 g3/2 (λ) 2T λ dT
" #
3 g5/2 (λ) 3T g5/2 (λ) 3 3 g3/2 (λ)
= − 2
− g3/2 (λ) +
2 g3/2 (λ) 2 g3/2 (λ) 2T 2T g1/2 (λ)
3 g5/2 (λ) 9 g5/2 (λ) 9 g3/2 (λ)
= + −
2 g3/2 (λ) 4 g3/2 (λ) 4 g1/2 (λ)
15 g5/2 (λ) 9 g3/2 (λ)
= − (8.69)
4 g3/2 (λ) 4 g1/2 (λ)
CV
8.9.3 for T < TBEC T
N kB
AF
Now, let us consider the case with T < TBEC . We have,
3/2
U 3 g5/2 (1) T
= T (8.70)
N kB 2 g3/2 (1) TBEC
DR
3/2
1 3 g5/2 (1) 5 T
CV = (8.71)
N kB 2 g3/2 (1) 2 TBEC
Thus we have,
15 g5/2 (λ) 9 g3/2 (λ)
− for T > TBEC
4 g3/2 (λ) 4 g1/2 (λ)
1
CV = (8.72)
N kB
3/2
15 g5/2 (1) T
for T < TBEC
4 g3/2 (1) TBEC
The specific heat is plotted against temperature in Figure (8.6). The cusp in the
heat capacity at T = TBEC is the signature of Bose-Einstein condensation. Asymp-
totically T → ∞, the heat capacity tends to the classical value consistent with
equi-partition.
140 8. Bose-Einstein Condensation
3h2
ǫ1,1,1 = ǫ0 = .
8mL2
The chemical potential is always less than or equal to ǫ0 . As temperature de-
creases, the chemical potential increases and comes closer and closer to the ground
state energy ǫ0 ≥ 0. Let us estimate how close µ can get to ǫ0 . In other words, we
want to estimate the smallest possible value of (ǫ0 − µ)/[kB T ]. To this end, consider
the expression for the average number of bosons in the ground state. Let us denote
this by N0 . It is given by,
1
N0 = (8.73)
ǫ0 − µ
exp −1
kB T
As temperature goes to zero, the chemical potential goes toward the ground
ǫ0 − µ
state energy. For a non-zero value of T , when is small, we can write
kB T
exp
ǫ0 − µ
kB T
=1+
ǫ0 − µ
kB T
T (8.74)
AF
Substituting this in the expression for N0 , given above, we get,
kB T
N0 = (8.75)
ǫ0 − µ(T )
DR
Therefore the chemical potential can never take a value close to any of the
excited states, since all of them invariably lie above the ground state. In a sense,
the ground state forbids the chemical potential to come close to any energy level
other than the ground state energy. It sort of guards all the excited states from a
close visit of µ. As T → 0, the number of bosons in the ground state increases.
This precisely is the subtle mechanism underlying Bose-Einstein con-
densation.
T
AF
DR
5
4.5
3.5
1.5
0.5
0
0 0.2 0.4 0.6 0.8 1
λ
Fig. 8.1. g3/2 (λ) versus λ. Graphical inversion to determine fugacity
T
AF
DR
100
90
80
70
60
50
40
λ
30 →
1 −λ
20
10
λ
0
0 0.2 0.4 0.6 0.8 1
0.9
0.8
0.7
0.6
λ
0.5
0.4
3 Λ3 λ
0.3 ρΛ = + g3/2 (λ)
3
V 1 −λ
0.2 Λ
= .05
0.1 V
2.612
0
0 0.5 1 1.5 2 2.5 3 3.5 4
ρΛ3
Fig. 8.4. Fugacity λ versus ρΛ3
T
AF
1.5
3 43/2
N0 T
=1−
DR
1
N TBEC
N0
N
0.5
0
0 0.5 1 1.5
T
TBEC
Fig. 8.5. Ground state occupation as a function of temperature
2
1.8
1.6
Classical : 3NkB /2
1.4
1.2
Nk B
CV
0.8
0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
T
TBEC
Fig. 8.6. Heat capacity in the neighbourhood of Bose - Einstein condensation
temperature
T
AF
DR
DR
AF
T
9. Elements of Phase Transition
T
AF
DR
148 9. Elements of Phase Transition
T
AF
DR
9. Elements of Phase Transition 149
T
AF
DR
150 9. Elements of Phase Transition
T
AF
DR
9. Elements of Phase Transition 151
T
AF
DR
152 9. Elements of Phase Transition
T
AF
DR
9. Elements of Phase Transition 153
T
AF
DR
154 9. Elements of Phase Transition
T
AF
DR
9. Elements of Phase Transition 155
T
AF
DR
DR
AF
T
10. Statistical Mechanics of Harmonic
Oscillators
Z
+∞
1 1 q2
Q1 (T ) = dq exp − r !2 ×
h −∞ 2 kB T
mω 2
Z " #
+∞
1 p2
dp exp − √ 3 (10.3)
−∞ 2 mkB T
Let σ1 and σ2 denote the standard deviations of of the two zero-mean Gaussian
distributions. These are given by,
r
kB T
σ1 = (10.4)
mω 2
√
σ2 = mkB T (10.5)
kB T
σ1 σ2 = (10.6)
ω
We have normalization identity for a Gaussian
158 10. Statistical Mechanics of Harmonic Oscillators
Z
+∞
1 x2 √
dx exp − 2
= σ 2π (10.7)
−∞ 2 σ
Therefore,
1 √ √ kB T
Q1 (T ) = (σ1 2π)(σ2 2π) = (10.8)
h ~ω
If all the oscillators are identical i.e. they all have the same characteristic frequency
of oscillations, then
3N
kB T
Q3N (T ) = (10.9)
~ω
On the other hand if the oscillators have distinct characteristic frequency {ωi :
i = 1, 2, · · · , 3N }, then
3N
Y kB T
Q3N (T ) = (10.10)
i=1
~ωi
F (T, V, N ) = U − T S (10.15)
dF = dU − T dS − SdT
∂F ~ω
µ= = kB T ln (10.18)
∂N T,V
kB T
∂F kB T
S =− = N kB ln +1 (10.19)
∂T V,N ~ω
We also have,
∂ ln Q
U =−
∂β V,N
= 3N kB T, T (10.20)
AF
consistent with equipartition theorem which says each quadratic term in the Hamil-
tonian carries kB T /2 of energy. The Hamiltonian of a single harmonic oscillator has
two quadratic terms - one in position q and the other in momentum p.
We also find that the results are consistent with the Dulong and Petit’s law
which says that the heat capacity at constant volume is independent of temperature:
DR
∂U
CV = = 3N kB = 3nR (10.21)
∂T V
CV
= 3R ≈ 6 calories (mole)−1 (kelvin)−1 (10.22)
n
More importantly, the heat capacity is the same for all the materials; it depends
only on the number of molecules or the number of moles of the substance and not
on what the substance is. The heat capacity per mole is approximately 6 calories
per Kelvin.
exp(−β~ω/2)
= (10.24)
1 − exp(−β~ω)
The partition function of a collection of 3N non-interacting quantum harmonic
oscillators is then given by
exp(−3N β~ω/2)
QN (T ) = (10.25)
[1 − exp(−β~ω)]3N
If the harmonic oscillators are all of different frequencies, the partition function is
given by
3N
Y exp(−β~ωi /2)
Q(T ) = (10.26)
i=1
1 − exp(−β~ωi )
The free energy is given by,
F (T, V, N ) = −kB T ln Q3N (T )
1
= 3N ~ω + kB T ln {1 − exp(−β~ω)} (10.27)
2
For 3N independent harmonic oscillators with different frequencies we have
F =
3N
X ~ωi
+ kB T ln {1 − exp(−β~ωi )}
T (10.28)
AF
i=1
2
Z ∞
~ω
= dω + kB T ln {1 − exp(−β~ω)} g(ω) (10.29)
2
DR
Z ∞
dω g(ω) = 3N (10.30)
0
We can obtain the thermodynamic properties of the system from the free energy.
We get,
∂F 1
µ= = ~ω + kB T ln [1 − exp(−β~ω)] (10.31)
∂N T,V 2
∂F
P =− =0 (10.32)
∂V T,N
∂F β~ω
S=− = 3N kB − ln {1 − exp(−β~ω)} (10.33)
∂T V,N exp(β~ω) − 1
∂ ln Q ~ω ~ω
U =− = 3N + (10.34)
∂β 2 exp(β~ω) − 1
10.1 Classical Harmonic Oscillators 161
The expression for U tells that the equipartition theorem is the first victim of quan-
tum mechanics : Quantum harmonic oscillators do not obey equipartition theorem.
The average energy per oscillator is higher than the classical value of kB T . Only for
T → ∞, we have kB T >> ~ω, the ”quantum” results coincide with the ”classical”
results. Let x = k~ω
BT
. Then Eq. (10.34) can be written as
U x x
= + (10.35)
3N kB T 2 exp(x) − 1
We have plotted U measured in units of 3N kB T as a function of oscillator energy
~ω measured in units of kB T . It is seen that when x → 0 (T → ∞), this quantity
tends to unity, see also below, implying that classical result of equipartition of
energy obtains at high temperature. Quantum effects start showing up only at low
temperatures.
In the limit x → 0 (T → ∞), we have,
U ∼ x x
x→0 + x2 x3
(10.36)
3N kB T 2 x + 2 + 6 + O(x4 )
x 1
= + 2 (10.37)
2 1 + x2 + x6 + O(x3 )
x x x2 x2
= +1− − + + O(x3 ) (10.38)
2 2 6 4
x2 x2
= 1− + + O(x3 ) (10.39)
6 4
= 1+
x2
12
+ O(x3 )
T (10.40)
AF
= 1 (10.41)
The heat capacity at constant volume is given by
∂U
CV = (10.42)
DR
∂T V,N
2
3N ~ω exp[β~ω]
= (10.43)
kB T (exp[β~ω] − 1)2
The second victim of quantum mechanics is the law of Dulong and Petit. The heat
capacity depends on temperature and on the oscillator frequency. The heat capacity
per mole will change from substance to substance because of its dependence on the
oscillator frequency. Only in the limit of T → ∞ (the same as β → 0), do we get
the classical results.
The temperature dependence of heat capacity is an important ”quantum” out-
come. We find that the heat capacity goes to zero exponentially as T → 0. However
experiments suggest that the fall is algebraic and not exponential. The heat capacity
is found to go to zero as T 3 . This is called T 3 law.
3N
X ∂V
∂xi
(xi − xi,0 ) +
T
AF
i=1 x1,0 ,x2,0 ,··· ,x3N,0
3N X
X 3N 2
1 ∂ V
(xi − xi,0 )(xj − xj,0 ) (10.44)
2 ∂xi ∂xj x1,0 ,x2,0 ,···x3N,0
DR
i=1 j=1
The first term gives the minimum energy of the solid when all its atoms are in their
equilibrium positions. We can denote this energy by V0 .
The second set of terms involving the first order partial derivatives of the poten-
tial are all identically zero by definition : V has a minimum at {xi = xi,0 ∀i = 1, 3N }
1
To appreciate the above statement, consider a class room wherein the chairs
are already arranged with constant spacing along the length and breadth of the
class room. The students occupy these chairs and form a regular structure. This
corresponds to a situation wherein each student is bound independently to his
chair.
Now consider a situation wherein the students are mutually bound to each
other. Let us say that the students interact with each other in the following
way : Each is required to keep an arm’s length from his four neighbours. If the
distance between two neighbouring students is less, they are pushed outward;
if more, they are pulled inward. Such mutual interactions lead to the student
organizing themselves in a two dimensional regular array
I shall leave it to you to visualize how such mutual nearest neighbour inter-
actions can give rise to three dimensional arrays.
10.1 Classical Harmonic Oscillators 163
The third set of terms involving second order partial derivatives describe har-
monic vibrations. We neglect the terms involving higher order derivatives and this
is justified if only small oscillations are present in the crystalline .
Thus under harmonic approximations we can write the Hamiltonian as,
X3N 2 X 3N X3N
1 dξi
H = V0 + + αi,j ξi ξj (10.45)
i=1
2 dt i=1 j=1
where
ξ = xi − x̄i (10.46)
1 ∂2V
αi,j = (10.47)
2 ∂xi ∂xj x̄1 ,x̄2 ,··· ,x̄3N
decoupled. If there are N atoms in the crystals there are 3N degrees of freedom.
Three of the degrees of freedom are associated with the translation of the whole
crystal; and three with rotation. Thus, there are strictly 3N − 6 normal mode
oscillations. If N is of the order of 1025 or so, it doesn’t matter if the number of
normal modes is 3N − 6 and not 3N .
We can write the canonical partition function as
3N
Y exp(−β~ωi /2)
Q= (10.49)
i=1
1 − exp(−β~ωi )
We have,
164 10. Statistical Mechanics of Harmonic Oscillators
3N
X
β~ωi
− ln Q = + ln {1 − exp(−β~ωi )} (10.51)
i=1
2
Z ∞
β~ω
= + ln {1 − exp(−β~ω)} g(ω)dω (10.52)
0 2
The problem reduces to finding the function g(ω). Once we know g(ω), we
can calculate the thermodynamic properties of the crystal. In particular we can
calculate the internal energy U and heat capacity, see below.
Z ∞
~ω ~ω exp(−β~ω)
U = + g(ω)dω
0 2 1 − exp(−β~ω)
Z ∞
~ω ~ω
= + g(ω)dω (10.53)
0 2 exp(β~ω) − 1
Z ∞
(β~ω)2 exp(β~ω)
C V = kB g(ω)dω (10.54)
0 [exp(β~ω) − 1]2
T
The problem of determining the function g(ω) is a non-trivial task. It is precisely
here that the difficulties lie. However, there are two well known approximations to
AF
g(ω). One of them is due to Einstein and the other due to Debye.
Einstein assumed all the 3N harmonic oscillators to have the same frequency. In
other words,
g(ω) = 3N δ(ω − ωE ) (10.55)
where ωE is the Einstein frequency or the frequency of the Einstein oscillator. The
Einstein formula for the heat capacity is then given by
2
~ω exp(~ωE /[kB T ])
CV = 3N kB (10.56)
kB T (exp(~ωE /[kB T ]) − 1)2
Let us define
~ωE
ΘE = (10.57)
kB
and call ΘE as Einstein temperature. Verify that this quantity has the unit of
temperature. In terms of Einstein temperature we have,
2
ΘE exp(ΘE /T )
CV = 3N kB (10.58)
T [exp(ΘE /T ) − 1]2
10.1 Classical Harmonic Oscillators 165
0 for ω > ωD
Let us now calculate CV under Debye’s theory. We start with
Z ∞
(β~ω)2 exp(β~ω)
CV (T ) = kB g(ω)dω (10.62)
0 [exp(β~ω) − 1]2
Let
x = β~ω (10.63)
~ωD
ΘD = (10.64)
kB
ΘD is called the Debye temperature. Then we have,
3 Z ΘD /T
T x4 exp(x)
CV = (3N kB ) × 3 dx (10.65)
ΘD 0 [exp(x) − 1]2
2
transverse or longitudinal
3
transverse mode is doubly degenerate and longitudinal mode is non-degenerate,
etc..
166 10. Statistical Mechanics of Harmonic Oscillators
= 3N kB (−3 + 4) T (10.70)
AF
= 3N kB (10.71)
In the low temperature limit we have T << ΘD . We start with,
ΘD 1
CV = 3N kB −3
T exp(ΘD /T ) − 1
DR
3 Z ΘD /T #
T x3
+ 12 dx (10.72)
ΘD 0 exp(x) − 1
In the limit T → 0, the first term inside the square bracket goes to zero like
exp(−ΘD /T ). The upper limit of the integral in the second term inside the square
bracket can be set to ∞. From standard integral tables5 , we have,
Z ∞
x3 π4
dx = (10.73)
0 exp(x) − 1 15
Thus we have in the low temperature limit,
3
∼ 12π 4 T
CV T →0 N kB (10.74)
5 ΘD
4
Take u(x) = x4 and dv(x) = exp(x)dx/[exp(x) − 1]2
5
The integral equals Γ (4)ζ(4), where Γ (·) is the gamma function and ζ(·) is the
Riemann zeta function. Γ (4) = 3! = 6 and ζ(4) = π 4 /90. See e.g. G B Arfken
and H J Weber, Mathematical Methods for Physicists, Fourth Edition, Academic
Press, INC, Prism Books PVT LTD (1995).
10.1 Classical Harmonic Oscillators 167
and ζ(·) is the Riemann zeta function, see below. Note ζ(2) = π 2 /6 and ζ(4) =
π 4 /90, etc..
Riemann zeta function is defined as
∞
X
ζ(p) = n−p (10.78)
n=1
x−p dx = (10.79)
1
∞
ln x for p = 1
DR
The integral and hence the series is divergent for p ≤ 1 and convergent for p > 1.
2(2n)!
= (−1)n−1 ζ(2n), n = 1, 2, 3, · · · (10.83)
(2π)2n
π6 π8
ζ(6) = ζ(8) = (10.85)
945 9450
T
AF
DR
Index
canonical ensemble 44, 63, 66, 71, 79, factorial moment 28, 123
80, 89, 90, 97, 128 Feller, Willam 33
canonical partition function 64–66, Fermi, Enrico 12
68–70, 84, 89–91, 105, 163 Fermi-Dirac statistics 108, 111, 113,
Carlo Rovelli 5 114, 116, 117, 119, 122
Carnot, Sadi 3 Fermions 109, 116, 118
Central Limit Theorem 38 fermions 90
Chandler, D 9 Finn, C. B. P. 12
characteristic function 34, 37, 38 first law of thermodynamics 88, 115
chemical potential 56, 71, 90, 97, 105, Fourier transform 34, 38
111–115, 119, 120, 127, 128, 131, 140, fugacity 89, 90, 105, 128, 130–132
141
Chowdhury, Debashish 9 Gabriel, Weinreich 12
Clausius 3, 79 Gaussian distribution 33, 34, 38
closed system 63–66, 71, 72, 77, 79, 80, geometric distribution 122, 123
157 Gibbs ensemble 27
cumulant 35, 38 Gibbs ensemble 18, 22, 24
cumulant generating function 35 Gibbs, J. W. 3, 10, 22
Gibbs-Duhem relation 92, 96, 97
Debye frequency 165 Glazer, M. 11
Debye temperature 165 Goodstein, David 9
Debye’s theory of specific heat 165 grand canonica partition function 114
170 Index
grand canonical ensemble 71, 80, 87, moment generating function 29, 30,
90, 128, 131 38, 123
grand canonical partition function 89, moments 35
90, 92, 99, 105, 108, 109, 115, 116, 127,
135 Newton 3
grand potential 90–92, 111, 112, 127
open system 8, 71, 80, 87, 88, 91, 94,
harmonic oscillators 157–159, 163 95, 111, 114, 127, 131, 135
Helmholtz 3
Helmholtz free energy 66, 68, 90, 91, Palash, B. Pal 9
158 Pathria, R. K. 9, 97
Huang, Kerson 10 Pauli’s exclusion principle 108
Planck, Max 12
ideal gas 83 Poisson distribution 32, 33, 38, 94, 125
ideal gas law 41, 43 Prigogine, Ilya 12
isolated system 64, 69–71, 78, 80 probability density function 17, 34
isothermal compressibility 95, 97, 112,
127 quantum statistics 105