Anda di halaman 1dari 13

PHY 314: Introduction to Quantum Mechanics, Varsha 2014

Lectures 22, 23 and 24


Anil Shaji
School of Physics, IISER Thiruvananthapuram
Quantum statistical mechanics

Thermodynamics is built on the observation that the behavior of large aggregations of particles
can be described in terms of bulk properties like temperature, pressure, volume, magnetization
etc. The field of statistical mechanics explores the microscopic foundations of thermodynamics
and more. Details of the motion of the individual particles - whether quantum or classical are often not of much concern in statistical mechanics because the energy carried by each of the
particles turns out to be the property that is directly related to the thermodynamical properties of
the bulk. Given the large number of particles that is the topic of study, at the outset itself, we do
away with the idea that we will be able to find exactly the energies of each of them. Instead, one
cares about the probability of each of the particles occupying states with different values of energy.
Note that this probability has nothing to do with the fact that quantum mechanics only makes
probabilistic statements about the results of measurements. Whether we are dealing with the
statistical mechanics of classical particles or with the statistical mechanics of quantum particles,
it is simply the large number of them that makes us choose to describe the whole system in a
probabilistic manner.
The basic assumption in statistical mechanics is that in thermal equilibrium every distinct state
with the same energy E is equally probable. The random motion an collisions of the particle might
move the energy about from particle to particle or transform it from one form to another (kinetic,
rotational, vibrational etc). Whatever might be the details of this process, in equilibrium, this
continuous redistribution of energy does not favor any particular state or the other. The only
constraint in this redistribution process is that the total energy of all the particles taken together
is fixed (due to conservation of energy). The temperature, T is a measure of the total energy of
the system.
Quantum statistical mechanics is different from the statistical mechanics of classical particles in
the way we count the number of distinct states. The symmetrization requirement places restrictions
on which states can be considered distinct. In fact the counting in quantum mechanics is usually
easier because the energy levels are discrete. Of course the counting also changes depending on
whether the individual particles are bosons or fermions. We start off by looking at these counting
differences in a simple example

2
I.

COUNTING STATES: AN EXAMPLE

Consider three noninteracting particles, all of mass m, in a one dimensional box. The total
energy of the system is
E = EA + EB + EC =


2 ~2 2
nA + n2B + n2C .
2
2ma

Now suppose the total energy of the system is fixed at E = 363 2 ~2 /2ma2 so that
n2A + n2B + n2C .
The following combinations of integers all are such that the sum of their squares is 363:
(11, 11, 11)
(13, 13, 5), (13, 5, 13), (5, 13, 13)
(1, 1, 19), (1, 19, 1), (19, 1, 1)
(5, 7, 17), (5, 17, 7), (7, 5, 17), (7, 17, 5), (17, 5, 7), (17, 7, 5)
If these particles are distinguishable, then the fundamental assumption in statistical mechanics
is that they are all equally likely. But, on the other hand, we are not interested in knowing which
particle is which state (whether particle A is in state 7 or particle B is in state 17) but rather we
are interested in how many particles are in each state labeled by 5, 7, 11 etc. The list of occupation
numbers for a given three particle state we refer to as the configuration. If all particles are in 11
then the configuration is
(0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 3, 0, 0, 0, 0, 0, 0, 0, 0, 0 . . .).
In other words N11 = 3 and all other occupation numbers are zero. If two particles are in 13 and
one in 5 , then the configuration is represented by
(0, 0, 0, 0, 1, 0, 0, 0, 0, 0, 0, 0, 2, 0, 0, 0, 0, 0, 0, 0 . . .).
i.e. (N5 = 1, N13 = 2). If two are in 1 and one in 19 , then the configuration is
(2, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 0, 1, 0 . . .).
Finally if one particle is in 5 , another in 7 and the third in 17 , then the configuration is
(0, 0, 0, 0, 1, 0, 1, 0, 0, 0, 0, 0, 0, 0, 0, 0, 1, 0, 0, 0 . . .).
The last configuration is the most probable one because it can be achieved in six different ways.
Now we can look at the following question. Suppose we pick a particle at random assuming that
the three particles are distinguishable, what is the probability that is has a certain energy? For
instance what is the probability that is has energy E1 ? Only the third configuration has particles
in state 1 and the third configuration can occur with a probability 3/13. In the third configuration,

3
a particle is in state 1 with probability 2/3. So the probability of finding the particle we picked in
the state 1 is
P1 =

3 2
2
=
13 3
13

Similarly we find that one can get E5 from either configuration 2 or 4 which appear with respective
probabilities 3/13 and 6/13 and within then E5 appears with probability 1/3 each. So we have
P5 =

3 1
6 1
3
+
= .
13 3 13 3
13

In a similar manner we get


P7 =

2
,
13

P11 =

1
,
13

P13 =

2
,
13

P17 =

2
,
13

P19 =

1
.
13

We can verify that P1 + P5 + P7 + P11 + P13 + P17 + P19 = 1.


Now suppose the particles are identical fermions. They the symmetrization requirement already makes the first second and third configurations impossible because more than one fermion
occupies the same state (Ignore the spin for now by assuming that they are all in the same spin
state). Moreover, there is really only one state corresponding to the fourth configuration because
of indistinguishability. So we have P5 = P7 = P17 = 1/3.
If we have identical bosons, then all the four configuration are allowed but still there is only
one state corresponding to each configuration. So we have
P1 =
P5 =
P7 =
P11 =
P13 =
P17 =
P19 =

1
4
1
4
1
4
1
4
1
4
1
4
1
4

2
1
=
3
6
1 1 1
1
+ =
3 4 3
6
1
1
=
3
12
1
1=
4
2
1
=
3
6
1
1
=
3
12
1
1
=
3
12

This example was to show how the counting of states depend on the type of particles. In fact
this counting is simplified further when the number of particles, N is large because then the most
probable configuration becomes so overwhelmingly so that we really have to only find that. We can
afford to ignore the other configurations altogether. So the distribution of single particle energies
at equilibrium is simply their distribution in the most probable state. So our strategy would be
to first identify the most probable configuration and then find the distribution of particles among
the states within that configuration.

4
II.

COUNTING STATES: THE GENERAL CASE

Let the particles have energies E1 , E2 , E3 , . . . with degeneracies d1 , d2 , d3 , . . . In other words


there are dn one particle states all with the same energy En in the system. Suppose there are
N particles in the system, we are interested in the configuration (N1 , N2 , N3 , . . .) for which there
are N1 particles with energy E2 , N2 with energy E2 and so on. To identify the most probable
configuration we have to first count the number of different ways each configuration can appear.
How many states correspond to this configuration? The answer, Q(N1 , N2 , . . .) depends, as we
know, on whether the particles are distinguishable, fermions or bosons.

1.

Distinguishable particles

If the particles are distinguishable then to populate N1 particles in the level E1 , we must pick
N1 out of the total N particles. This can be done in
N

CN1 =

N!
N1 !(N N1 )!

ways. Now we have to distribute the N1 particles among the d1 degenerate energy levels. Since
there is no restriction of the number of particles in each degenerate level, each particle has d1
choices. So the total number of ways we can distribute N1 particles in the first level is
1
N ! dN
1
.
N1 !(N N1 )!

Out of the remaining N N1 particles we can choose N2 of them and distribute them among
the d2 degenerate energy levels in
2
(N N1 )! dN
2
N2 !(N N1 N2 )!

ways. So it follows that the number of ways in which this configuration can appear is
1
2
3
N ! dN
(N N1 )! dN
(N N1 N2 )! dN
1
2
3

...
N1 !(N N1 )! N2 !(N N1 N2 )! N3 !(N N1 N2 N3 )!

1 N2 N3
Y
n
dN
dN
n
1 d2 d3 . . .
= N!
.
= N!
N1 !N2 !N3 ! . . .
Nn !

Q(N1 , N2 , N3 , . . .) =

(1)

n=1

2.

Indistinguishable fermions

When the particles are indistinguishable there is no question of choosing N1 particles out of the
total N and so on because any set of N1 particles is the same as any other set of N1 particles. In
other words there only one N particle state with the occupation numbers (N1 , N2 , N3 , . . .). So all
we have to worry about is distributing the Nn particles among dn degenerate states. For fermions

5
this problem is very simple since only one particle can occupy a state. The number of ways in
which Nn particles can be distributed among dn states is
dn

CNn =

dn !
.
Nn !(dn Nn )!

So we have
Q(N1 , N2 , N3 , . . .) =

Y
n=1

3.

dn !
.
Nn !(dn Nn )!

(2)

Indistinguishable bosons

For bosons also the first part of the counting is easy since there is again only one N particle
state with the occupation numbers (N1 , N2 , N3 , . . .). But now any number of particles can go sit
in the same degenerate level. So counting the number of ways of distributing Nn particles among
dn degenerate levels becomes a bit more difficult. The question before us is, how many different
ways can we assign Nn identical particles to dn slots. A particularly clever way of counting is
the following. Let dots represent particles and crosses represent partitions, so that for example if
dn = 5 and Nn = 7,

would indicate that there are two particles in the first degenerate state, one in the second, three
in the third, one in the fourth and none in the fifth. Generalizing this, we can see that in any
particular way of distributing, there will be Nn dots and dn 1 crosses. If all dots and crosses were
labelled (distinguishable), there will be (Nn + dn 1)! ways of permuting the dots and crosses.
But since we do not distinguish between any of the dots (particles), we have to divide this number
by Nn !. Similarly since we do not distinguish between the crosses either, we have to divide by
(dn 1)!. So there are
(Nn + dn 1)! Nn +dn 1
=
CN n
Nn !(dn 1)!
ways of assigning the Nn particles to the dn one-particle states in the nth bin and we conclude that
Q(N1 , N2 , N3 , . . .) =

Y
(Nn + dn 1)!
.
Nn !(dn 1)!

(3)

n=1

A.

The most probable configuration

Now that we have counted the number of ways in which a particular configuration can appear,
we can look for that configuration which appears in the most number of different ways so as to
identify the most probable configuration. Naively one could set

Q(N1 , N2 , N3 , . . .) = 0
Nn

6
to find the equations that will lead to the most probable configuration. However there are a couple
of constraints that one needs to satisfy, namely, the total number of particles must be fixed at N
and the total energy must be fixed at E. In other words

Nn = N

n=1

and

Nn En = E.

n=1

To do the constrained maximization of a function F (x1 , x2 , . . .) of several variables, subject to


the constraints f1 (x1 , x2 , . . .) = 0 and f2 (x1 , x2 , . . .) = 0, etc., is handled by the method of Lagrange
multipliers. We introduce the new function
G(x1 , x2 , x3 , . . .) F + 1 f1 + 2 f2 + . . .
and set all its derivatives equal to zero:
G
= 0;
xn

G
=0
n

In this method 1 and 2 are roughly the cost of violating the constraint. Since we are adding
this cost to the function that we are minimizing, things will work out such that the constraint will
not be violated.
For our case, it is easier to work with ln Q instead of Q to do the minimization. This will work
out since ln Q is a monotonic function function of Q. So we let

G = ln Q + N

Nn + E

n=0


Nn En ,

n=1

where and are the Lagrange multipliers. Setting the derivatives with respect to and to
zero just reproduces the constraints. So we now have to set the derivatives with respect to Nn
equal to zero.

1.

Distinguishable particles

From Eq. (1), we have

Y
n
dN
n
Q(N1 , N2 , N3 , . . .) = N !
Nn !
n=1

for distinguishable particles. So we have


G = ln(N !) +

X
n=1

Nn ln(dn )

X
n=1


ln(Nn !) + N

X
n=0

Nn + E

X
n=1


Nn En

7
To simplify this formula, we can use Stirlings approximation,
ln(z!) ' z ln(z) z
to get
G ' N ln(N ) N

Nn ln(Nn ) +

n=1

Nn +

n=1






X
X
Nn En .
Nn ln(dn ) + N
Nn + E

n=1

n=0

n=1

Simplifying and collecting terms with Nn together, we get


G'

Nn ln(Nn ) + ln(dn ) En + 1 + N ln(N ) N + N + E.

n=1

Setting the derivative of G with respect to Nn to zero, we get




G
Nn
= 0 ln
= ( + En )
Nn
dn
or
Nn = dn e(+En ) .
2.

Fermions

From Eq. (2) we have

Q(N1 , N2 , N3 , . . .) =

n=1

dn !
,
Nn !(dn Nn )!

and
G '

X



dn ln(dn ) dn Nn ln(Nn ) Nn (dn Nn ) ln(dn Nn ) (dn Nn )

n=1


+ N

Nn + E

n=0


Nn En .

n=1

So we have
G
= ln(Nn ) 1 1 + ln(dn Nn ) + 1 + 1 + + En = 0,
Nn
which may be simplified to

ln

Nn
dn Nn


= ( + En )

Nn =

dn
.
1 + e+En

(4)

8
3.

Bosons

The counting of states we did for Bosons we have

Y
(Nn + dn 1)!
Q(N1 , N2 , N3 , . . .) =
.
Nn !(dn 1)!
n=1

Therefore
G '

X


(Nn + dn 1) ln(Nn + dn 1) (Nn + dn 1) Nn ln(Nn ) + Nn

n=1






X
X

Nn En .
(dn 1) ln(dn 1) + (dn 1) + N
Nn + E
n=0

n=1

G
= ln(Nn + dn 1) + 1 1 ln(Nn ) 1 + 1 + + En = 0.
Nn
So we have

ln

Nn
Nn + dn 1


= ( + En ),

and
Nn =

dn
+E
n
e

after setting dn 1 ' dn in order to be consistent with our previous approximations.


B.

Finding and

To evaluate the constants and we have to substitute the expressions we got for Nn into the
two constraint equations,
N=

X
n=1

Nn ,

and E =

Nn En .

n=1

What we substitute for Nn depends on whether we have distinguishable particles, fermions or


bosons. In addition to this, we need to know what En are and their degeneracies. This is fixed by
the model we have for the particles we are considering. If we take the particles to be in a box of
dimensions lx , ly and lz , then we have seen that the available states for the particles are labeled
by the k-space vector,


~k = nx , ny , nz .
lx
ly
lz
The volume occupied by each state in k-space is 3 /V , where V is the volume of the box. We can
take all the states lying within a shell of thickness dk at a radius k within the positive octant as
having the same energy. So the degeneracy of those energy levels is given by
dk =

1 2k 2 , dk
V
= 2 k 2 dk,
3
8 /V
2

9
with the corresponding energy given by
Ek =

~2 k 2
.
2m

Any additional factor that appears due to the spin of the particles will cancel out from the numerator and denominator in the expression for dk and so we need not consider the effect of the
particles having spin here.
Turning the sum into an integral in the first constraint equation and assuming that we are
dealing with distinguishable particles and use the appropriate expression for Nn we get
V
N = 2 e
2

dk e

~2 k2 /2m 2

k =Ve

m
2~2

3/2
,

using
Z

dr rq er/A = Aq+1 (q + 1);

q > 1.

So we have
e

N
=
V

2~2
m

3/2
.

The second constraint equation says


V ~2
E= 2
e
2 2m

dk e


3/2
3V
m
k =
e
.
2
2~2

~2 k2 /2m 4

Substituting the expression for e we get


E=

3N
.
2

We can compare this formula for the average energy of a particle in an ideal gas at temperature,
T,
2
E
= kB T
N
3
and we find that
= kB T
by comparison. To prove this formula in general for all types of particles and with them not
necessarily in a box is beyond the scope of the present discussion but it can be done. The main
thing that has to be done to prove this is that different substances in thermal equilibrium with one
another all have the same value of . Usually we also define the chemical potential, (T ), which is
a function of temperature as
(T ) = kB T.

10
We can now rewrite the formulas we got for the most probable configurations for each type of
particles as the most probable number of particles n() in each state with energy . This is done
simply by dividing both sides by the degeneracy of the corresponding energy level to get

()/kB T

Maxwell Boltzmann

e
1
Fermi Dirac
n() =
(5)
e()/KB T +1

Bose Einstein
e()/KB T 1

The classical result that was known even before the advent quantum mechanics applies to distinguishable particles and is known as the Maxwell-Boltzmann distribution. The result for identical
fermions is the Fermi-Dirac distribution and the one for identical boson is the Bose-Einstein statistics.
At zero temperature, the Fermi-Dirac distribution takes on a particularly simple form,
(
1,
if  < (0)
n() =
0,
if  > (0).
All the states up to an energy (0) are filled and none are filled beyond that. We have already
noted that the energy up to which states are filled in the electron gas model of a solid is the Fermi
energy. The chemical potential at zero temperature is the Fermi-Energy
(0) = EF .
As the temperature increases, the Fermi-Dirac distribution softens the cutoff of the distribution
when T > 0 as shown in Fig. 1.

FIG. 1: The Fermi-Dirac distribution for T = 0 and its softening for T > 0.

For distinguishable particles, we found that the total energy of all the particles is
3
E = N kB T,
2

11
and the chemical potential is

(T ) = kB T = kB T ln

N
V

3
+ ln
2

2~2
mkB T


.

We can work out similar expressions for the total energy of Bosons and Fermions but the actual
integrals that appear are rather unwieldy. These will be dealt with in the course on statistical
mechanics.

C.

The Blackbody Spectrum

We now go back to the beginnings of quantum mechanics to the problem that Max Planck
pondered about. The question was to come up with a theoretical analysis that will match the
observed intensities coming out at various wavelengths/frequencies from a black body. A black
body is an object that absorbs all the radiation incident on it. An example of a black body which
is not even black, typically is a cavity in a piece of metal connected by a small hole to the outside.
The hole (and not the cavity) is the black body because any radiation falling on the hole goes into
the cavity and effectively gets trapped in there. The radiation comes to equilibrium with the walls
of the cavity and whatever radiation that comes out of the hole can be treated as a small amount
that leaks out of the equilibrium cavity field. So in that sense the hole is the black body.
The radiation coming out of a black body depends on its temperature. What is measured is the
intensity of radiation coming out at each frequency (or equivalently wavelength). The spectrum
for different temperatures is shown in Fig. (2)

FIG. 2: Black body spectrum

As shown in the figure, calculations based on classical theories gave an answer that was in stark
disagreement with the observed spectrum. This was one of the two dark-clouds hanging over
the firmament of Physics at the beginning of the last century (the other being the negative result
of the Michaelson-Morely experiment on luminiferous ether).

12
The classical theory had two results regarding black body spectra. The Raleigh-Jeans law that
worked at low frequencies and the Wiens displacement law that did well at high frequencies. Max
Planck was trying to put the two together. The starting point was to model the field inside the
cavity as a collection of standing waves of light. The standing waves are nothing but the various
stationary states for a particle in a box like problem. For light, the frequency is proportional to
|~k| since = c/ and 1/ = |~k|. We know that the number of states with the same energy as
characterized by k (or ) is the number of lattice points in the shell of thickness dk and radius k
in k-space. In other words,
dk = d k 2 ' 2 .
Coupled with the equipartition theorem that had worked so well in classical statistical mechanics
and thermodynamics, this posed a problem because there are far more available states at higher
frequencies than at lower frequencies. So one would expect almost all the energy available inside
the cavity to be distributed into high frequency modes rather than into the fewer number of low
frequency modes. This meant that the predicted spectrum diverged as a function of . Planck
needed to understand why these high frequency modes were not actually excited in a real black
body.
Plancks insight was to suggest that the energy of each mode of frequency comes in discrete
packages of size ~. This means that no modes of frequency max such that ~max > Etot will not
be excited at all where Etot is the total energy of the black body. The number of modes with lower
frequencies that come into play also are bounded by n~ < Etot .
Instead of taking Plancks route, we can use what we know now coupled with Plancks hypothesis
to come up with the result that Planck got. The assumptions that go in are
1. The energy of a photon (mode of the electromagnetic field) is related to its frequency by
E = h = ~.
2. The wave number k is related to the frequency by k = 2/ = /c, where c is the speed of
light.
3. Only two spin states occur for the photon m = 1 only (m = 0 is not allowed).
4. The number of photons is not a conserved quantity (photons have no chemical potential).
When the temperature increases, the number of photons per unit volume also increases.
The photon is about as relativistic a particle as you can get since it is moving at c but we are
applying non-relativistic quantum mechanics to it! The additional assumptions above (especially
the one about only two spin states) cannot be justified within non-relativistic quantum mechanics
so one has to accept the four points above either simply as assertions or as assumptions. Having
assumed these we have
N =

dk
~/k
BT
e

13
For free photons in a box,
dk = 2

V 2
V
k dk = 2 3 2 d,
2
2
c

the factor of two coming in because of the two possible spin states. So the energy density, N ~/V
in a frequency range d is given by
() =

~ 3

.
2 c3 e~/kB T 1

This is Plancks formula that agreed well with experiments and which started off the development
of quantum mechanics.

Anda mungkin juga menyukai