Next: About this document ...
LECTURE 15
Degenerate Electron Gas: Electrons in a Metal
In a metal there are electrons which do
not belong to any particular atom or group of atoms but rather are
delocalized and extended throughout the system. These are conduction
electrons and carry the electric current when an electric field is
applied. The atoms in a metal may also have core electrons which
fill the inner shells and are local to each atom. We will ignore
these core electrons since they don't affect the properties of the metal.
We will focus on the conduction electrons. For simplicity
we can describe them as free electrons which means that we neglect
their Coulomb interactions. So we can regard the electrons as a gas
but because their density is high, their de Broglie wavelength
is much larger than their mean separation (see Reif 7.4). This means
that the classical approximation breaks down and we have to use
quantum statistics. The first thing to understand is that
the electrons fill a Fermi sea. To understand this, recall
the eigenstates for a particle in a box with
periodic boundary conditions. We found that the momentum acquires discrete
values
where the momentum
and the numbers
,
, and
are any set of integers-
positive, negative, or zero. We can imagine setting up coordinate
axes
,
, and
and putting a point everywhere in
space that there is an allowed state. Notice that as the size
of the box gets bigger, the states get closer together in
space.
Since
, the lower energy states are the ones which
are closer to the origin in
space.
Now let's put in our conduction electrons. We will fill the states in
order of increasing energy. Only 2 electrons (spin up and spin down)
can go into each state. After we've finished putting all our electrons
into the states in
space, we have what is called a Fermi sea. In
our case the Fermi sea will be a sphere in
space. The surface
of this sphere is called the Fermi surface. The Fermi energy
is the energy of a state at the Fermi surface, and the Fermi
wavevector
is the radius of the Fermi sea. So in our case
. An electron buried in the
depths of the Fermi sea can't really jump to a nearby state in
space
because the nearby states are occupied. So these electrons don't contribute
to the electric current. It's the electrons near the Fermi surface
that can make transitions to unoccupied states above the Fermi surface
and contribute to the electrical conduction. How near to the Fermi
surface do the electrons have to be? Within
of
.
=3.0 true in
This picture is consistent with the Fermi distribution.
 |
(1) |
At
the Fermi distribution is a step function that tells us that
the states below the Fermi energy are occupied and those above
are unoccupied. The chemical potential at
is called the Fermi
energy,
. The Fermi energy depends on how many electrons
there are; the more electrons, the larger the Fermi energy. More precisely,
the higher the density of electrons, the higher the Fermi energy. We
can calculate this dependence as follows. Since all states with
are occupied at
, we have
 |
(2) |
where the factor of 2 is for the 2 possible spin states of an electron.
Or
 |
(3) |
or
 |
(4) |
We can define a characteristic temperature which is called the Fermi
temperature
 |
(5) |
Typically the Fermi temperature is on the order of
1 eV
10,000 K. So room temperature (300 K) is much less than
. At room temperature the electrons behave a lot like they do
at zero temperature. When we say low temperature for electrons, we
mean
. In this case the Fermi distribution is a little smeared
but still looks step-like. We speak of a ``degenerate'' Fermi gas.
By degenerate, we mean there are a lot of electrons with the same
energy, namely the Fermi energy. If
, then we are in the
classical, nondegenerate limit.
A good example of a degenerate Fermi gas are the electrons in a metal.
At first glance, it would appear that the noninteracting approximation
would be totally inappropriate because of the obvious Coulomb interactions
among the electrons. Two comments should be made. (1) More sophisticated
theory indicates that a noninteracting, single particle approximation
is valid but the particles are not true electrons but rather
``quasiparticles'' with energy
where
is an ``effective mass.'' This approximation improves as the
temperature is lowered. (2) We can compute the consequences of the simple
non-interacting assumption and compare the results to experimental
data. The agreement is excellent.
Consider a typical metal like Copper (Cu). If one plugs the appropriate
numbers into (5), then one finds that
Kelvin
(see page 391 of Reif). This is well above the melting point of copper,
so for any temperature below the melting point, we can use the low
temperature approximation.
One of the many quantities of interest is the contribution of the electrons
in a metal to the total specific heat of the solid. (The other big
contributor are the lattice vibrations or phonons.) Classically, using
the equipartition theorem, we expect
the electrons to contribute
to the molar specific
heat. Yet at room temperature, the observed specific heat of both metals
and insulators is
. This is the classical Dulong-Petit result that
we got by just considering vibrations of the atoms and ignoring the electrons.
So it appears that the electrons don't contribute to the specific heat
at room temperature. In other words, they aren't behaving classically.
Because of the Pauli principle, only those electrons within a few
of the Fermi energy are capable of being excited into higher energy
levels and changing their state. Most of the electrons are far below
the Fermi energy, cannot change their state because all the states near
them in energy are occupied, and therefore do not participate in thermal
processes. We can assume that only a fraction of the electrons can
contribute to the specific heat. This fraction
. Thus we
can approximate the molar specific heat by
 |
(6) |
where
 |
(7) |
This says that the low temperature specific heat of metals should contain
a term proportional to
. This is indeed observed.
Reif does a proper calculation of the
specific heat
for a free electron gas in section 9.17.
One starts with the mean energy
 |
(8) |
where
is the number of states lying between
and
. The thermal occupation
factor is the Fermi function:
 |
(9) |
The specific heat is just the derivative of
:
 |
(10) |
We won't go through the calculation (see Reif 9.17). The answer
for the molar specific heat is
 |
(11) |
The fact that electrons have both spin-up and spin-down states is
included in here. Note that the specific heat is linear in
.
Bose-Einstein Condensation and Superfluidity
(Reference: Robert B. Leighton, Principles of Modern Physics,
McGraw-Hill (1959).)
Interesting things happen at very low temperatures and Bose-Einstein
condensation is one of them. Recall that there is no statistical limit to
the number bosons that can occupy a single state. In a Bose condensed
state, an appreciable fraction of the particles is in the lowest
energy level at temperatures below
. These particles are in
the same state and can be described by the same wavefunction. In other
words a macroscopic number of particles are in one coherent state.
If we write
, then this state is described by
a given phase
.
The oldest known physical manifestation of Bose condensation is superfluid
. A
atom has total angular momentum zero and is therefore
a boson. At
K liquid helium becomes superfluid. The transition
temperature is called the
point because the shape of
the specific heat curve at
is shaped like
. One cools
liquid helium by pumping on it to get rid of the hot atoms (evaporative
cooling). It boils a little. Then at the transition it boils vigorously
and suddenly stops. The reason for this behavior is that the thermal
conductivity increases by a factor of about
at the transition,
so that the superfluid is no longer able to sustain a temperature
gradient. To make a bubble, heat has to locally vaporize the fluid
and make it much hotter than the surrounding fluid. This is no longer
possible in the superfluid state.
Perhaps the hallmark
of a superfluid is that it has no viscosity. As a result the superfluid
can flow through tiny capillary tubes that normal liquid can't get
through. Superfluid
is often described by a two-fluid model, i.e.,
it is thought of as consisting of 2 fluids, one of which is normal
and the other is superfluid. It's the superfluid component which is
able to flow through the capillary tube. So if you use this method to
measure the coefficient of viscosity, you find that it
suddenly drops to zero at the
point.
One can see the effect of
both components by putting a torsional oscillator consisting of a stack
thin, light, closely spaced mica disks immersed in the liquid. If the
liquid has a high viscosity, the liquid between the disks is dragged
along and contributes significantly to the moment of inertia of the disks.
If the viscosity is small, the moment of inertia is more nearly equal to
that of the disks alone. Using this method, no discontinuity is found in the
coefficient of viscosity at the
point.
Another weird thing that superfluid helium does is escape from
a beaker by crawling up the sides, flowing down the outside, and
dripping off the bottom. The helium atoms are attracted by the van
der Waals forces of the walls of the container, and they are
able to flow up the walls because of the lack of viscosity.
The rate of flow can be 30 cm per second or more. The superfluid helium
can surmount quite a high wall, on the order of several meters in height.
(Brief aside to explain the van der Waals force: As an electron moves
in a molecule, there exists at any instant of time a separation of
positive and negative charge in the molecule. The latter has, therefore,
an electric dipole moment
which varies in time. If another molecule
exists nearby, it will have a dipole moment induced by the first molecule.
These two dipoles are attracted to each other. This is the
van der Waals force.)
We can show mathematically that there is
a macrosopic population of the lowest energy state in the following way.
Consider a gas of noninteracting bosons. Let the energy levels be measured
from the lowest energy level, i.e., let the zero point energy be the
zero of energy. Then the chemical potential
must be negative,
otherwise the Bose-Einstein distribution would be negative for some
of the levels. Recall from lecture 13 that the Bose-Einstein distribution
gives the average number of particles in state
:
 |
(12) |
is adjusted so that the total number of particles is
:
 |
(13) |
Let me give a sneak preview: If we assume a continuous distribution of states,
starts out negative and gets bigger
as the temperature decreases.
equals its upper limit of zero
at some temperature
, below
which we can no longer satisfy (13) because the right hand side
can't deliver enough particles. This leads us to
treat the lowest energy level separately and we find that we can satisfy
(13) by keeping the extra particles we need in the lowest
state.
Now let's do the math.
In order to turn the sum over
in (13)
into an integral, let's assume a
continuous density of states. In lecture 14 we found that if
k-space is isotropic, i.e., the same in every direction,
then the number of states in a spherical shell lying between
radii
and
is
 |
(14) |
Now if the energy of the bosons is purely kinetic energy and continuous,
then
 |
(15) |
and
 |
(16) |
or
 |
(17) |
Also (15) implies that
 |
(18) |
Plugging (17) and (18) into (14) yields
 |
(19) |
So we can rewrite (13) as
Now recall that for a geometric series
 |
(20) |
So
 |
(21) |
Plugging this into (20) leads to
 |
(22) |
Let
. Then
 |
(23) |
Let
and
. Then
 |
(24) |
The definition of a gamma function is
 |
(25) |
(Aside:
if
is a positive integer.)
So
 |
(26) |
Thus
 |
(27) |
Now on the right hand side, as the temperature
decreases,
must increase to keep the product constant and equal to
.
can be made as small as we wish, but
, which we said must be
negative, cannot be greater than zero. But the product must be a constant.
So (27) is only valid above a certain critical temperature
.
Below this temperature, our treatment breaks down. Where did we go wrong?
The flaw lies in the fact that we assumed that the states are
continuously distributed. However, since we are interested in very low
temperatures which involves the occupation of the lowest lying energy
levels, we may expect that the actual discrete nature of the level
distribution might play an essential role in the lowest temperature range.
So let us treat the lowest level
separately. We will assume
that it is not degenerate with any other levels and we will assume that the
remaining levels are continuously distributed from
to
as described by (19). So in our summation (13)
we will treat the lowest level separately:
where the second quantity on the right is just the right hand side of
(27), written in terms of the critical temperature
.
because
is defined such that the right hand side of
(27) equals
with
and
.
We see that it is now possible to satisfy this new equation with
negative values of
for all temperatures, since the first
term becomes infinite as
. The inclusion of the lowest
energy level as a separate term in our treatment has thus removed the
previous difficulty of not being able to account for all of the particles
at temparatures below
. If we now inquire into what this equation
means physically, we see that, at temperatures below
, the chemical potential
will take on such values that those
particles which are not included in the continuous distribution will be
found in the lowest level. That is, a kind of condensation occurs;
it is such that an appreciable fraction of the particles is in the
lowest energy level at temperatures below
.
If we write (29) as
 |
(29) |
and realize that
at low temperatures, then we find
the population
of the lowest level to be, approximately,
![\begin{displaymath}
n_1=N\left[1-\left(\frac{T}{T_C}\right)^{3/2}\right]
\end{displaymath}](img97.png) |
(30) |
At
,
while at
,
.
Using
, we find that at low temperatures
 |
(31) |
Notice that
is negative. As
,
:
 |
(32) |
Considering superfluid helium as a 2 component fluid with normal and
superfluid components is consistent with having some of the particles
in the lowest energy level and the rest in higher energy levels. There
is no microscopic theory of superfluid helium, though computer simulations
by Ceperley have been quite successful in reproducing its properties.
One of the complications is that the helium atoms are so closely packed
that they are strongly interacting;
they're in a liquid state. It would be closer to the
ideal case to have a system of bosons which are weakly interacting.
(Reference: H.-J. Miesner and W. Ketterle, ``Bose-Einstein Condensation
in Dilute Atomic Gases,'' Solid State Communications 107, 629 (1998)
and references therein.)
This has recently been achieved in the case of alkali atoms such as rubidium,
sodium, and lithium.
Using a combination of optical and magnetic traps together with laser cooling
and evaporative cooling, several research groups have achieved
Bose condensation in dilute weakly interacting vapors of alkali atoms.
In these systems the thermal deBroglie wavelength exceeds the
mean distance between atoms.
Nanokelvin temperatures and densities of
cm
have
been achieved. (Compare this to a mole of liquid which has a
typical density of
cm
.) At nanokelvin temperatures
the thermal deBroglie wavelength exceeds 1
m which is about 10 times
the average spacing between atoms.
In these experiments they have actually been able to
directly observe the macroscopic population of the zero momentum
eigenstate. In addition the coherence resulting from being in macroscopic
wavefunctions has been demonstrated by observing the interference
between two independent condensates. Two spatially separated condensates
were released from the magnetic trap and allowed to overlap during
ballistic expansion of the gases. Interference patterns were observed
that are
analogous to the pattern produced in a double-slit experiment in optics.
Next: About this document ...
Clare Yu
2008-06-02