You are on page 1of 14

Statistical Thermodynamics: Molecules to Machines

Venkat Viswanathan
May 5, 2015

Module 4: Electrons, Phonons and Photons


Learning Objectives:
Analyze the statistical thermodynamics of Bose and Fermi particles
Demonstrate consistency with our analysis for the ideal gas of Bose
particles (ideal gas of diatomic molecules)
Analyze the behavior of electrons in a metal
Discuss the thermodynamic behavior of a crystalline solid.
Find the role of lattice vibrations, resulting in the definition of quasiparticles called phonons.
Compare two models for lattice vibrations
Look at photons in the context of Bose-Einstein statistics
Derive Plancks law of radiation and use density of states to derive
thermodynamic properties of a photon gas.

Key Concepts:
Bose and Fermi statistics, bosons and fermions, electron gas, Fermi
energy, Fermi momentum, Fermi temperature, electron pressure, electron heat capacity, crystal lattice, lattice vibrations, phonons, Einstein
model, Debye model, Black Body radiation, statistical mechanics of photons, Bose occupation, Plancks law.

statistical thermodynamics: molecules to machines

Non-interacting particles obeying Bose and Fermi


statistics
Consider a system of non-interacting, indistinguishable particles that
can have energies ( = 1, 2, ...) associated with their quantum mechanical states. The state of the system can be specified by the number
of particles at each energy level, i.e. n is the number of particles at
energy state .
P
Total number of particles is n = N .
P
Total system energy is n = E.
In the canonical ensemble, we write the partition function:
!
!
X
X
X
Q(T , V , N ) =
N
n exp
n
n1 ,n2 ,...

(1)

where we include the delta-function constraint on the summation in


P
order to fix n = N , and we define = 1/(kB T ).
The indistinguishability of the particles is properly accounted for in
this representation since any given set of n contributes a single term
without over-counting indistinguishable states. In the grand canonical
ensemble, we write the grand canonical partition function:
(T , V , ) =

Q(T , V , N ) exp(N )

N =0

!
X

N =0 n1 ,n2 ,...

!
exp

n +

n N

!
n

Y X

n1 ,n2 ,...

en n

The Landau potential is written as:


!
X

pV = kB T log = kB T

log

n n

(2)

If the particles are Bosons, there are no restrictions on the number


of particles in a given state n ( = 1, 2, ...), thus:

X

e +

n

n = 0

1
1 e +

(3)

If the particles are Fermions, any given state can only have either
n = 0 or n = 1 particles, thus:

X
n =0

e +

n

= 1 + e +

(4)

statistical thermodynamics: molecules to machines

Therefore, we generally write:




X
pV = kB T
log 1 e +

(5)

where "-" is for Bosons, and "+" is for Fermions.


From this, we find the average number of particles:


X
X e +
log
hN i =
=
hn i
=

1 e +

(6)

where hn i is the average occupation number in the state.


In the ideal gas limit, is very large and negative. Noting that

e +  1, we have:
X
hN i
e + = e q and pV kB T e q = kB T hN i

where we define the single-particle partition function q =


The grand canonical partition function is:



X
X
1 N N
= epV /(kB T ) = exp e q =
Q(T , V , N )eN
e
q =
N!
N =0

thus Q =

1 N
N! q

N =0

for ideal gas (agrees with previous lecture).

Fermi-Dirac statistics for conducting electrons in a


metal
Electrons in a metal can be modeled as a gas of non-interacting fermions:
Electrons in a metal are at high densities (many atoms per volume
with each contributing to the conducting electrons).
Since no two electrons (fermions) can exist in the same state, a high
density system fills many of the single particle energy levels
The lowest unoccupied state will have a kinetic energy much larger
than kB T , thus thermal excitations result in energetics with large
kinetic energy and comparatively negligible potential energy of interaction.
The large kinetic energy associated with these electrons results in
large conductivity of electrons in metals
Using the results for the thermodynamic behavior of non-interacting
fermions, find the behavior of an ideal gas of electrons in a metal. Electrons in a metal act as non-interacting particles with quantized energies
given by:
E~n =


h2
n2x + n2y + n2z
2
8mL

translational modes

statistical thermodynamics: molecules to machines

which are the energy levels for a particle in a box for a particles with
mass m = 9.10938 1031 kg.
The average number of electrons in the ~n state is:
1
= F (E~n )
(7)
e (E~n ) + 1
h
i1
where we have defined the Fermi function F () = e () + 1
.
The total number of electrons is given by:
hn~n i =

hN i = 2

X
X

X
nx =1 ny =1 nz =1

1
e (E~n ) + 1

=2

X
X

F (E~n )

(8)

nx =1 ny =1 nz =1

where we include a factor of 2 since the electrons can exist in spin-up


and a spin-down states.
For sufficiently large V , the spectrum of translational wavemodes is
effectively a continuum. Therefore, we can convert this summation to
an integral over ~n, resulting in:
Z
Z
Z
hN i = 2
dnx
dny
dnz F (E~n )
0
0
0
Z
Z
Z
L
L
L
dky
dkz F [(~k )]
=2
dkx
0
0

0
Z
Z
Z
Z
3
2V
L 1
dkx
dky
dkz F [(~k )] =
d~kF [(~k )]
=2 3 3
2
(2 )3

where we have used a coordinate change from ~n to ~k = L


~n, and the
2 k2
}
energy is now written as (~k ) = 2m .
Define the chemical potential at T = 0 to be (T = 0) = F (Fermi
Energy). To proceed, consider the form of the Fermi function at T = 0,
F () = 1 for F and F () = 0 for > F . Define the Fermi

momentum pF according to F =
found to be:
hN i =

p2F
2m

2
}2 k F
2m

. Thus, at T = 0, hN i is

2V 4 3
8 V (2m)3/2 3/2
kF =
F
F =
3
(2 ) 3
3
h3

3
8

2/3

h2
2m

A typical metal (Cu) has a mass density of 9g/cm3 . Assuming each


atom donates a single electron to the conducting electron gas, this density has a Fermi temperature F = kFB 80, 000K. This verifies that
the Fermi energy F is sufficiently large to make the ideal gas approximation valid at room temperature. At room temperature, only states
with energy very near F will be affected by thermal energy kB T .
The spread in the distribution is approximately 2kB T (Fig. 1).
The pressure is found using the relationship for Fermi particles


X
pV = 2kB T
log 1 + e +
(9)

Figure 1: The Fermi function F () =

e () + 1

1

statistical thermodynamics: molecules to machines

Following a similar derivation as before, we write


Z
2V
~
pV =
d~k log{1 + e [(k)] }
(2 )3
Z
i
h
4V (2m)2/3
1/2
()
=
d
log
1
+
e
h3
0

(10)

2 2

k
where we have used = }2m
.
In the limit T 0(or ), we have 1 + e () e (F ) for
< F and 1 + e () 1 for > F .
Therefore, the pressure is written as:
Z
16V (2m)2/3 5/2
4V (2m)3/2 F
1/2
d
(

)
=
F
(11)
pV =
F
h3
15h3
0

This pressure at T = 0 is approximately 106 atm. This large pressure


plays an important role in halting the collapse of a star (white dwarf)
because this enormous pressure offsets the gravitational forces that otherwise drive the collapse.
The average energy hEi is found using:
hEi = 2

e +
1 + e +

(12)

Following a similar derivation as before, we write


Z
1
2V
d~k(~k )
hEi =
3
~
(2 )
e [(k)] + 1
Z
2/3
1
4V (2m)
d3/2 []
=
h3
e
+1
0
Z
2/3
4V (2m)
=
d3/2 F ()
h3
0
where we have used =

}2 k2 .
2m

V (2m)2/3 3/2
F

We apply integration by parts to this equation and use hN i = 83


h3
to get:
Z
3hN i
dF
d5/2
hEi = 3/2
(13)
d
5
0
F

where

dF
d

e ()

= [e () +1]2 .

In the limit T 0, the function dF


d becomes peaked near F ,
and we can effectively expand the integrand near = F (Fig. 2).
5 3/2
Expand the integrand about = F to get 5/2 5/2
F + 2 F (
15 1/2
F ) + 8 F ( F )2 + ....
dF ()

Since d is even about F in this limit, the odd-order terms will


integrate to zero, leaving only the even terms. Thus, the final form of
the average energy is going to be:
#
"
 2
T
hEi = hN iF A + B
+ ...
(14)
F

Figure 2:
the Fermi

The derivative
dF ()
function d

e () [e () + 1]

of
=

statistical thermodynamics: molecules to machines

A precise calculation of this low-temperature expansion (outside scope


of this module) gives:
#
"
 
3
5 2 T 2
+ ...
hEi = hN iF 1 +
5
12 F

(15)

Therefore, the heat capacity for a metal in the limit of small temperature is given by:
CV =

3
5 2 T
2
T
hN iF
2 2 =
hN ikB
5
12 F
2
F

(16)

giving a linear temperature dependence in the small-T limit.


The limiting behavior in the small-T limit suggests that a plot of
CV /T approaches a constant as T 0. This proves to be the behavior
of the heat capacity for metals in the limit of small T . As temperature
increases, the fluctuations in the metal nuclei also contribute to the heat
capacity (Fig. 3).

Crystalline Solid
The atoms in a crystal are arranged in a regular array of points in space
with a variety of possible crystalline lattices (Fig. 4). At zero temperature, the atomic coordinates are uniquely locked into spatial positions
that minimize the energy. At finite temperature, the atoms fluctuate
about the energy-minimum positions, leading to lattice vibrations that
govern the thermodynamic behavior (Fig. 5).

Figure 3: The heat capacity CV of


T i3 SiC2 exhibits the temperature dependence CV = 1 T + 3 T 3 + ... in the
low-temperature limit [Ho et al., 1999]

Thermodynamic contribution of lattice vibrations


Consider N atoms with positions {~r}
lattice. Define the potential energy V
for a given system configuration {~r}.
(0)

(0)

(0)

tion {~r(0) } = ~r1 , ~r2 , ..., ~rN

= ~r1 , ~r2 , ..., ~rN in a crystalline


({~r}) that describes the energy
A minimum-energy configura
satisfies the condition ~ri V {~r(0) } =

~ri V |0 = ~0 for i = 1, 2, ..., N . The atomic positions {~r(0) } define the


regular crystalline lattice.
Expand the potential energy about {~r} = {~r(0) }, such that:
N

 X

 
(0)
V ({~r}) V {~r(0) } +
~ri V |0 ~ri ~ri
+
i=1
N N
1 XX

 


(0)
(0)
~ri ~rj V |0 : ~ri ~ri
~rj ~rj
+ ...

i=1 j =1

Figure 4: Crystal lattice structures

statistical thermodynamics: molecules to machines

The linear term is zero (by definition), thus the energy is:
V V0 +

N N
1 XX X
si sj Kij
2
, =x,y,z

(17)

i=1 j =1



(0)
where si = e ~ri ~ri
and Kij = ~ri ~rj V |0 .
The Matrix Kij can be diagonalized into normal modes (eigenvectors) with effective elastic constants (eigenvalues). Since there are
3N 6 3N degrees of freedom, there are 3N normal modes.
The potential energy is written as:
3N

V = V0 +

1X 2
Kl l
2

(18)

l =1

where l is the magnitude of the lth normal mode.


The total energy of the system is then written as:
3N
3N
X
p2l
1X 2
E=
+
Kl l + V0
2m
l
2
l =1

(19)

l =1

where pl and m
l are the effective momentum and mass of the lth normal
mode, respectively.
The total energy E is decomposed into normal modes with an individualmode energy in the form of a harmonic oscillator. These normal modes
are called phonons.
Phonons act as quasi-particles which means they are distinguishable
and independent, i.e. they dont interact between each other.
The Hamiltonian (sum of kinetic and potential energy) of the lth
p2

phonon is given by: Hl = 2ml l + 12 Kl l2 . This harmonic oscillator energy


results in the quantized energy:


1
El = jl +
}l
2

(20)

where jl = 0, 1, 2..., and the phonon frequency is l =


The canonical partition function Q is given by:
Q(T , V , N ) =

...

j1 =0 j2 =0

= eV0

V0

P3N
l=1

l
K
m
l.

(jl + 12 )}l

j3N =0

3N X

e (jl + 2 )}l

l=1 jl =0

= eV0

1
3N
Y
e 2 }l
1 e }l

(21)

l =1

where we use the mathematical property

n=0 z

1
1z .

Figure 5: Schematic showing the atoms


in a crystal in their locked, energyminimum positions at T = 0 and the
lattice vibrations that occur at finite
temperature

statistical thermodynamics: molecules to machines

This gives the Helmholtz free energy:



3N 

 1
X
}l
F = kB T log Q = V0 + kB T
log 1 e
+ }l
2
l =1

The average energy is given by:


3N

X
log Q
hEi =
= V0 +

l =1

1
}l e }l
+ }l
2
1 e }l

The next step is to analyze two competing models for the phonon
frequencies l .

Einstein model
In the Einstein model, we assume there is a single characteristic frequency of the crystal, defined as the Einstein frequency E . The average
energy for this model is given by:
hEi = V0 +

3N }E
3N }E e }l
+
2
1 e }E

For this model, the heat capacity is given by:



CV =

hEi
T


V ,N

2
 2 E

}
k E
E
E
BT
e
e T
3N
k
3N kB k}B
B
T
T
=
=
2


2
}E
E
kB T
T
1

e
1e

where we define E = }kBE .


In the limit T the heat capacity CV 3N kB (Dulong-Petit
Law). As we learned before, the equipartition theorem states that the
energy receives kB T per thermally active degree of freedom for a harmonic oscillator (quantized energy is linear in the quantum index). In
the limit T 0 the heat capacity CV 0. The heat capacity approaches zero exponentially in the small-T limit (Fig. 6).

Debye model
The Debye model treats the solid as an elastic material. Vibrational
modes in an elastic solid correspond to sound waves, thus the frequencies
satisfy = ck, where c is the speed of sound in the solid and k = m/L,
where m = 1, 2, ...

Figure 6: Heat capacity of a crystal


predicted by the Einstein model of lattice vibration

statistical thermodynamics: molecules to machines

Convert the sum over normal modes to a sum over k using:


Z
Z
Z
XXX
X
(...) =
(...) = dm1 dm2 dm3 (...)
l

m1 m2 m3

  3 Z kc
Z kc
Z kc
L
dk1
dk2
dk3 (...)
=

0
0
0
  3 Z kc
 3
Z
L
L
1 D
= 4
dkk 2 (...) = 4
d 2 (...)
3

c
0
0
where kc is a cutoff wavemode (to be determined), and D = Ckc is
called the Debye frequency.
A complete conversion will include one longitudinal mode with cl
and two transverse modes with ct . This gives:
 Z D
 3 
X
L
1
2
+ 3
d 2 (...)
(22)
(...) = 4

c3l
ct
0
l
P
To find D we must enforce that l 1 = 3N , thus we have:
 3 
 Z D
X
L
2
1
1 = 4
+
d 2 (...)

c3l
c3t
0
l

1/3
 3
 3 

1
2 D

9N
L


+ 3
= 3N D =
=

3
L 1
c3l
ct
+ c23
c3
l

For convenience, use D as a parameter, thus:


Z D
X
1
d 2 (...)
(...) = 9N
D 0

(23)

Define the Debye temperature D = }kBD , which defines the temperature scale for vibrational fluctuations. To test this model, we find the
heat capacity:

CV = kB

hEi

T3
= 9N kB 3
D

dx
0

x4 ex

D /T

(1 ex )2

In the limit T the heat capacity approaches:


Z
T 3 D /T
CV 9N kB 3
dxx2 = 3N kB
D 0
which is expected (Dulong-Petit Law).
In the limit T 0 the heat capacity scales as:
Z
T3
x4 ex
T3
CV 9N kB 3
dx
N kB 3
2
D 0
D
(1 ex )
The heat capacity predicted by the Einstein and the Debye models
are very similar. However, the low temperature of the heat capacity
of non-conducting solids matches the Debye model, i.e. CV N T 3
(Fig. 7).

Figure 7: Heat capacity of a crystal


predicted by the Einstein model and
Debye model of lattice vibration

statistical thermodynamics: molecules to machines

Black Body Radiation


We are all familiar with the idea that hot objects emit radiation, a light
bulb, for example. In the hot wire filament, an electron, originally in an
excited state drops to a lower energy state and the energy difference is
given off as a photon, ( = h = 2 1 ). We are also familiar with the
absorption of radiation by surfaces. For example, clothes in the summer
absorb photons from the sun and heat up. Black clothes absorb more
radiation than lighter ones. This means, of course that lighter colored
clothes reflect a larger fraction of the light falling on them.
A black body is defined as one which absorbs all the radiation incident upon it, i.e. a perfect absorber. It also emits the radiation
subsequently. If radiation is falling on a black body, its temperature
rises until it reaches equilibrium with the radiation. At equilibrium, it
re-emits as much radiation as it absorbs so there is no net gain in energy
and the temperature remains constant. In this case, the surface is in
equilibrium with the radiation and the temperature of the surface must
be the same as the temperature of the radiation.
To develop the idea of radiation temperature we construct an enclosure having walls which are perfect absorbers (see Fig. 8). Inside
the enclosure is radiation. Eventually this radiation reaches equilibrium
with the enclosure walls, equal amounts are emitted and absorbed by
the walls. Also, the amount of radiation travelling in each direction becomes equal and is uniform. In this case the radiation may be regarded
as a gas of photons in equilibrium having a uniform temperature. The
enclosure is then called an isothermal enclosure.
An enclosure of this type containing a small hole is itself a black
body. Any radiation passing through the hole will be absorbed. The
radiation emitted from the hole is characteristic of a black body at the
temperature of the photon gas. The properties of the emitted radiation is then independent of the materials of the wall provided they are
sufficiently absorbing that essentially all radiation entering the hole is
absorbed. This universal radiation is called Black Body Radiation.
An everyday example of a photon gas is the background radiation in
the universe. This photon gas is at a temperature of about 5 K. Thus the
earths surface, at a temperature of about 300 K, is not in equilibrium
with this gas. The earth is a net emitter of radiation (excluding the sun)
and this is why it is dark at night and why it is coldest on clear nights
when there is no cloud cover to increase the reflection of the earths
radiation back to the earth. A second example is a Bessemer converter
used in steel manufacture containing molten steel. These vessels actually
contain holes like the isothermal enclosure of Fig. 8. The radiation
emitted from the hole is used in steel making to measure the temperature
in the vessel, by means of an optical pyrometer.

Figure 8: Isothermal Enclosure.

10

statistical thermodynamics: molecules to machines

Statistical Thermodynamics
To derive Plancks radiation law directly from our statistical mechanics,
we note that number of photons in the gas is not fixed. The photons
are absorbed and re-emitted by the enclosure walls. Since the photons
are non-interacting it is by this absorption and re-emission that equilibrium is maintained in the gas. Since, also the free energy F (T , V , N ) is
constant in equilibrium (at constant T and V ) while N varies it follows
that F /N = 0, that is

F
=
=0
N V ,N

(24)

The photon gas is then a Bose gas with = 0 so that the canonical
partition function is given directly as
Q=

r
Y

(1 exp(s ))1

(25)

s=1

And the expected Bose occupation is


ns = (exp(s ) 1)1

(26)

Using  = h and the equation of density of states, we then obtain:


1
( )n ( )g ( )
V

(27)

8
h
2
3
c (exp(h ) 1)

(28)

u( ) =
or
u( ) =

Which is Plancks Radiation Law. We may also recover Wien and


Rayleigh-Jeans laws as limits of Plancks law,
hc
(a) Long wavelength, kT
<< 1. Here
 =

hc/
' kT
(exp(hc/kT ) 1)

(29)

And Eq. 28 becomes


1
8
()g () ' 4 kT
V

Which is Wiens law valid at long wavelengths.


hc
(b) Short wavelength, kT
>> 1. Here
u() =

 =

hc
exp(hc/kT )

(30)

(31)

And Eq. 30 becomes


u() =

1
8hc
()g () '
exp(hc/kT )
V
5

(32)

11

statistical thermodynamics: molecules to machines

Which is the Wiens law valid at short wavelength. Employing our


statistical mechanics we readily obtained Plancks radiation law. We
may also derive the Stefan-Boltzmann law for
Z
Z
8
h
u=
du( ) = 3
2
(33)
c 0 (exp(h ) 1)
0
Introducing x = h, this reduces to
u=

8k 4
(hc)3

x3
T 4 = aT 4
exp(x) 1

(34)

8k 4 4
(hc)3 15

(35)

dx
0

Where
a=

In this way we obtain, using statistical mechanics, a law derived


previously using thermodynamics including all the numerical factors.
This gives Stefans constant in
1
caT 4 = T 4
4

(36)

2ck 4 4
erg
= 5.67 105 2
3
(hc) 15
cm Sec K 4

(37)

E=
As
=

A measurement of could then be used, for example, to determine


Plancks constant. Planck, in fact, determined h as the constant needed
in his radiation law to fit the observed spectral distribution law. This
gave him the value h = 6.55 1027 erg.sec which compares with the
present value of h = 6.625 1027 erg.sec

Thermodynamic Properties
We may calculate all the thermodynamic properties of black body radiation using statistical mechanics through the partition function Q, where
F = kT log(Q)

(38)

And Q was derived as shown in Eq. 25. This is the basic method
of statistical thermodynamics. The aim is to reproduce all the thermodynamic properties with all factors and constants evaluated. This
gives

12

statistical thermodynamics: molecules to machines

r
Y

F = kT log

(1 exp(s ))

s=1
r
X

F = kT

log(1 exp(s ))1

s=1

d
log(1 exp(s ))1
h3

F = 2kT

(39)

Where the s are the single photon states and the factor of 2 arises
from the two polarizations available to each photon. This can be integrated in a variety of ways. Perhaps the most direct is to integrate
over phase space (d = dV 4p2 dp) and write  = pc. Introducing the
dimensionless variable x =  = pc, the Helmholtz free energy is:
F =

1
3

8k 4
(hc)3

d(x)3 log(1 exp(x))

V T4

(40)

The dimensionless integral here can be transformed into that appearing in the constant of a of Eq. 35, by an integration by parts, i.e.,
Z

d(x)3 log(1 exp(x))


Z

= (x3 ) log(1 exp(x)) +

I=

x3 d[log(1 exp(x))]

(41)

The first term vanishes because:


(a) limx x3 log(1 exp(x)) ' limx x3 exp(x) = 0
(b) limx0 x3 log(1 exp(x)) ' limx0 x3 log(x) = 0
And
Z

I=

dx
0

x3

=
exp(x) 1
15

(42)

Comparing Eq. 40 and Eq. 35, we get:


1
F = aV T 4
(43)
3
From F we may determine all other thermodynamic properties by
differentiation. For example, the entropy is

S=

F
T



= 4 aV T 3

3
V

(44)

The internal energy is:


U = F + T S = avT 4

(45)

13

statistical thermodynamics: molecules to machines

The pressure is:



p=

F
V



= 1 aT 4

3
T

(46)

Finally, the Gibbs free energy is: s


1
1
G = F + pV = aV T 4 + aV T 4 = 0
3
3

(47)

This is zero as required G = N and the chemical potential = 0.


We may use these expressions to further
verify
thermodynamic consis



U
S
tency, for example that CV = T T
= T .
V
V
In summary, we have obtained the spectral distribution from the
Bose occupation in much the same way as we obtained the MaxwellBoltzmann distribution for a classical gas. The only other required
ingredient was the density of states. We have also obtained all the
thermodynamic properties using the partition function Q.

References
J. C. Ho, H. H. Hamdeh, M. W. Barsoum, and T. El-Raghy. Low
temperature heat capacity of T i3 SiC2 . Journal of Applied Physics, 85
(11):7970, June 1999.

14

You might also like