Next: About this document ...
LECTURE 16
Systems of Interacting Particles
So far we have been considering systems of noninteracting or weakly interacting
particles because these are simple to deal with. However, most systems
in the real world are complicated because they have particles that interact with
each other, e.g., liquids and solids. Our strategy to find the energy eigenvalues,
calculate the partition function , and derive the appropriate thermodynamic
quantities from still holds, but it can be difficult to
find the energy eigenvalues.
However there are cases of interacting systems which we can solve.
For example,
at low temperatures thermodynamic properties are dominated by the low energy
excited states. So we do not need to find all the energy eigenstates; we just
need to describe these low energy states accurately. If the system is ordered
at low temperatures, these low-lying excited states are ``collective modes.''
For example, in a crystal the atoms are located on lattice sites. The collective
modes are normal modes of vibration, i.e., sound waves. Quantized sound waves
are called phonons which are analogous to photons.
Another example is a ferromagnet where all the spins are lined up parallel to
each other in the ground state. The low energy excited states correspond
to spin waves which are called magnons when quantized.
We can also sometimes solve interacting systems in the high temperature limit
where is large compared to the mean energy of interaction. In the infinite
temperature limit the interactions are negligible. At high temperatures the
interactions can be treated as a perturbation and one can do things like
high temperature series expansions.
We will now consider some examples of interacting systems.
Lattice vibrations and normal modes
Consider a solid consisting of atoms, each of mass with position
. The equilibrium position is
.
Each atom can vibrate about its equilibrium position. The displacement
from the equilibrium position is
|
(1) |
The kinetic energy of the solid is given by
|
(2) |
where
is the th component
of the velocity of the th atom.
Since the displacements are small, we can expand the potential energy
in a Taylor series:
|
(3) |
where and go from 1 to ; and and go from 1 to 3.
The derivatives are evaluated at the equilibrium positions
of the atoms. The first term is the potential energy when the atoms
are in their equilibrium configuration. Since this is a minimum of , the
first derivative must vanish:
,
i.e., there is no force on any atom in the equilibrium configuration.
So the first term in the Taylor series vanishes. For the second term, let
|
(4) |
We can neglect higher order terms since the displacements are small.
So we obtain
|
(5) |
and the Hamiltonian becomes
|
(6) |
The kinetic energy is simple since it is just a sum of terms, each of which
just involves one coordinate. But the potential energy is complicated
since it involves cross terms coming from different atoms and different coordinates.
This is the result of interactions. Since the potential energy is
quadratic in the coordinates, we can eliminate this complication by
finding the normal modes of the solid. This amounts to making
a linear transformation to a new set of normal coordinates :
|
(7) |
such that a proper choice of coefficients transforms the
Hamiltonian to the simple (diagonal) form:
|
(8) |
Now we have a sum over independent harmonic oscillators with no cross terms.
Each oscillator has frequency , and its quantum mechanical energy
is given by
|
(9) |
where 0, 1, 2 ... The total energy is the sum of these one dimensional
harmonic oscillator energies:
where
|
(11) |
is a constant independent of .
is the zero point energy.
represents the binding energy per atom in the solid at absolute zero.
It is now straightforward to calculate the partition function which follows
what we did for the Einstein oscillators (see calculation of the Einstein
specific heat in lecture 11).
Notice that each harmonic oscillator partition function is a
geometric series. Recall that for a geometric series
|
(13) |
So we have
|
(14) |
Now we take the logarithm to get :
|
(15) |
We can convert this sum into an integral by defining
to be the number of normal modes with angular frequencies in the range between
and
.
|
(16) |
The mean energy of the solid is
|
(17) |
The heat capacity at constant volume is
The function
is determined by the normal vibrational modes of the
solid. However, regardless of the exact shape of
, we can
make some general statements about the high temperature limit. Let
be the highest frequency of the normal mode spectrum such that:
|
(19) |
If the temperature is high enough such that
,
then
for all relevant
, and we can expand the exponential:
|
(20) |
Then for
,
|
(21) |
since the integral over
is simply the total number of modes:
|
(22) |
Eq. (21) is the Dulong-Petit law that we obtained earlier by
applying the equipartition theorem.
Debye Approximation
The Debye approximation treats the solid as an elastic continuum and ignores
the discreteness of the atoms. This is a good approximation as long as the
wavelength of the elastic vibration is much longer than the mean
atomic lattice spacing , i.e., . Long wavelength corresponds
to low frequency modes. Let
be the function describing
the number of modes in an elastic continuum. Then we expect
at low frequencies. This approximation
will break down for
.
For an elastic continuum, the dispersion relation is where
is the speed of sound. This has the same form as the dispersion
relation for photons. From our previous derivation for the density of
states in lecture 14, we can write
|
(23) |
where the factor of 3 accounts for the 3 phonon polarizations: 1 longitudinal mode and
2 modes transverse to the direction of propagation with wavevector .
The Debye approximation approximates
with
defined by
|
(24) |
where the ``Debye frequency'' is chosen so that
yields the correct total number of normal modes:
|
(25) |
This normalization determines :
|
(26) |
Solving for yields
|
(27) |
For typical numbers:
cm/sec,
,
sec
THz.
The corresponding wavelength
. The corresponding
Debye temperature is defined by
.
Typically the Debye temperature is of order 300 K.
Notice that the Debye density of states is quadratic in .
=3.0 true in
Using the Debye approximation, the heat capacity becomes
where
. Using
and
|
(29) |
we can write
|
(30) |
As we have seen, at high temperatures, this reduces to the Dulong-Petit law.
At low temperatures, and we can replace the upper limit
of the integral with and the integral becomes a constant. As a result,
we can see immediately that
|
(31) |
More precisely, the integral can be evaluated exactly:
|
(32) |
This yields
|
(33) |
Notice that at low temperatures ().
This provides a reasonably
good fit at low temperatures, though it may be necessary to go to temperatures
as low as
. The Debye specific heat certainly gives
better agreement with experiment at low temperature than the exponential temperature
dependence predicted by the Einstein specific heat (see lecture 11). The Einstein
specific heat makes the approximation that all the oscillators have a single
frequency :
|
(34) |
Plugging this into Eq. (18) and using
yields our previous result for the Einstein heat capacity from lecture 11:
|
(35) |
Figure 10.2.2 in Reif
shows a comparison of the Debye and Einstein specific heats. Both have an
S-shape:
=3.0 true in
The Debye approximation is good for the acoustic phonon modes while the
Einstein approximation is good for the high energy optical phonon modes.
=2.5 true in
Next: About this document ...
Clare Yu
2009-03-05