You are on page 1of 417

Chapter 1: Chemical Bonding

Linus Pauling (19011994)


December 28, 2001
Contents
1 The development of Bands and their lling 4
2 Dierent Types of Bonds 9
2.1 Covalent Bonding . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Ionic Bonding . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.1 Madelung Sums . . . . . . . . . . . . . . . . . . . 17
2.3 Metallic Bonding . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Van der Waals Bonds . . . . . . . . . . . . . . . . . . . . 20
2.4.1 Van der Waals-London Interaction . . . . . . . . 21
1
Ac 89
Actinium
227.028
Th 90
Thorium
232.038
Pa 91
Protactinium
231.036
U 92
Uranium
238.029
Np 93
Neptunium
237.048
Pu 94
Plutonium
(244)
Am 95
Americium
(243)
Cm 96
Curium
(247)
Bk 97
Berkelium
(247)
Cf 98
Californium
(251)
Es 99
Einsteinium
(252)
Fm100
Fermium
(257)
Md101
Mendelevium
(258)
No 102
Nobelium
(259)
La 57
Lanthanum
138.906
Ce 58
Cerium
140.115
Pr 59
Praseodymium
140.908
Nd 60
Neodymium
144.24
Pm 61
Promethium
(145)
Sm 62
Samarium
150.36
Eu 63
Europium
151.965
Gd 64
Gadolinium
157.25
Tb 65
Terbium
158.925
Dy 66
Dysprosium
162.50
Ho 67
Holmium
164.93
Er 68
Erbium
167.26
Tm 69
Thulium
168.934
Yb 70
Ytterbium
173.04
7
Fr 87
Francium
(223)
Ra 88
Radium
226.025
Lr 103
Lawrencium
(260)
6
Cs 55
Caesium
132.905
Ba 56
Barium
137.327
Lu 71
Lutetium
174.967
Hf 72
Halfnium
178.49
Ta 73
Tantalum
180.948
W 74
Tungsten
183.85
Re 75
Rhenium
186.207
Os 76
Osmium
190.2
Ir 77
Iridium
192.22
Pt 78
Platinum
195.08
Au 79
Gold
196.967
Hg 80
Mercury
200.59
Tl 81
Thallium
204.383
Pb 82
Lead
207.2
Bi 83
Bismuth
208.980
Po 84
Polonium
(209)
At 85
Astatine
(210)
Rn 86
Radon
(222)
5
Rb 37
Rubidium
85.468
Sr 38
Strontium
87.62
Y 39
Yttrium
88.906
Zr 40
Zirconium
91.224
Nb 41
Niobium
92.906
Mo 42
Molybdenum
95.94
Tc 43
Technetium
(98)
Ru 44
Ruthenium
101.07
Rh 45
Rhodium
102.906
Pd 46
Palladium
106.42
Ag 47
Silver
107.868
Cd 48
Cadmium
112.411
In 49
Indium
114.82
Sn 50
Tin
118.71
Sb 51
Antimony
121.75
Te 52
Tellurium
127.60
I 53
Iodine
126.905
Xe 54
Xenon
131.29
4
K 19
Potassium
39.098
Ca 20
Calcium
40.078
Sc 21
Scandium
44.956
Ti 22
Titanium
47.88
V 23
Vanadium
50.942
Cr 24
Chromium
51.996
Mn 25
Manganese
54.938
Fe 26
Iron
55.847
Co 27
Cobalt
58.933
Ni 28
Nickel
58.69
Cu 29
Copper
63.546
Zn 30
Zinc
65.39
Ga 31
Gallium
69.723
Ge 32
Germanium
72.61
As 33
Arsenic
74.922
Se 34
Selenium
78.96
Br 35
Bromine
79.904
Kr 36
Krypton
83.80
3
Na 11
Sodium
22.990
Mg 12
Magnesium
24.305
3 4 5 6 7 8 9 10 11 12
Al 13
Aluminum
26.982
Si 14
Silicon
28.086
P 15
Phosphorous
30.974
S 16
Sulfur
32.066
Cl 17
Chlorine
35.453
Ar 18
Argon
39.948
2
Li 3
Lithium
6.941
Be 4
Beryllium
9.012
B 5
Boron
10.811
C 6
Carbon
12.011
N 7
Nitrogen
14.007
O 8
Oxygen
15.999
F 9
Fluorine
18.998
Ne 10
Neon
20.180
1
H 1
Hydrogen
1.008
2 13 14 15 16 17
He 2
Helium
4.003
1 18
Periodic Table
2
Solid state physics is the study of mainly periodic systems (or things
that are close to periodic) in the thermodynamic limit 10
21
atoms/cm
3
.
At rst this would appear to be a hopeless task, to solve such a large
system.
Figure 1: The simplest model of a solid is a periodic array of valance orbitals embedded
in a matrix of atomic cores.
However, the self-similar, translationally invariant nature of the pe-
riodic solid and the fact that the core electrons are very tightly bound
at each site (so we may ignore their dynamics) makes approximate so-
lutions possible. Thus, the simplest model of a solid is a periodic array
of valance orbitals embedded in a matrix of atomic cores. Solving the
problem in one of the irreducible elements of the periodic solid (cf. one
of the spheres in Fig. 1), is often equivalent to solving the whole sys-
tem. For this reason we must study the periodicity and the mechanism
(chemical bonding) which binds the lattice into a periodic structure.
The latter is the emphasis of this chapter.
3
1 The development of Bands and their lling
nl elemental solid
1s H,He
2s Li,Be
2p BNe
3s Na,Mg
3p AlAr
4s K,Ca
3d transition metals ScZn
4p GaKr
5s Rb,Sr
4d transition metals YCd
5p In-Xe
6s Cs,Ba
4f Rare Earths (Lanthanides) CeLu
5d Transition metals LaHg
6p TlRn
Table 1: Orbital lling scheme for the rst few atomic orbitals
We will imagine that each atom (cf. one of the spheres in Fig. 1)
is composed of Hydrogenic orbitals which we describe by a screened
4
Coulomb potential
V (r) =
Z
nl
e
2
r
(1)
where Z
nl
describes the eective charge seen by each electron (in prin-
ciple, it will then be a function of n and l). As electrons are added to
the solid, they then ll up the one-electron states 1s 2s 3s 3p 3d 4s 4p
4d 4f , where the correspondence between spdf and l is s l = 0,
p l = 1, etc. The elemental solids are then made up by lling these
orbitals systematically (as shown in Table 1) starting with the lowest
energy states (where E
nl
=
me
4
Z
2
nl
2 h
2
n
2
Note that for large n, the orbitals do not ll up simply as a func-
tion of n as we would expect from a simple Hydrogenic model with
E
n
=
mZ
2
e
4
2 h
2
n
2
(with all electrons seeing the same nuclear charge Z). For
example, the 5s orbitals ll before the 4d! This is because the situation
is complicated by atomic screening. I.e. s-electrons can sample the core
and so are not very well screened whereas d and f states face the an-
gular momentum barrier which keeps them away from the atomic core
so that they feel a potential that is screened by the electrons of smaller
n and l. To put is another way, the eective Z
5s
is larger than Z
4d
. A
schematic atomic level structure, accounting for screening, is shown in
Fig. 2.
Now lets consider the process of constructing a periodic solid. The
simplest model of a solid is a periodic array of valence orbitals embed-
ded in a matrix of atomic cores (Fig. 1). As a simple model of how
5
z/r
1s
2sp
2s
3spd
1s
3s
3p
3d
4spdf
4s
4p
5spdf
5s
4d
5p
6spdf
6s
4f
5d
2p
+
s
d
atom
4f
Ce Valence Shell
6s
5d
V(r)
V(r) + C l (l+1)/r
2
Figure 2: Level crossings due to atomic screening. The potential felt by states with
large l are screened since they cannot access the nucleus. Thus, orbitals of dierent
principle quantum numbers can be close in energy. I.e., in elemental Ce, (4f
1
5d
1
6s
2
)
both the 5d and 4f orbitals may be considered to be in the valence shell, and form
metallic bands. However, the 5d orbitals are much larger and of higher symmetry than
the 4f ones. Thus, electrons tend to hybridize (move on or o) with the 5d orbitals
more eectively. The Coulomb repulsion between electrons on the same 4f orbital will
be strong, so these electrons on these orbitals tend to form magnetic moments.
the eigenstates of the individual atoms are modied when brought to-
gether to form a solid, consider a pair of isolated orbitals. If they are far
apart, each orbital has a Hamiltonian H
0
= n, where n is the orbital
occupancy and we have ignored the eects of electronic correlations
(which would contribute terms proportional to n

). If we bring them
together so that they can exchange electrons, i.e. hybridize, then the
degeneracy of the two orbitals is lifted. Suppose the system can gain
6
Figure 3: Two isolated orbitals. If they are far apart, each has a Hamiltonian H
0
= n,
where n is the orbital occupancy.
Figure 4: Two orbitals close enough to gain energy by hybridization. The hybridization
lifts the degeneracy of the orbitals, creating bonding and antibonding states.
an amount of energy t by moving the electrons from site to site (Our
conclusions will not depend upon the sign of t. We will see that t is
proportional to the overlap of the atomic orbitals). Then
H = (n
1
+n
2
) t(c

1
c
2
+c

2
c
1
) . (2)
where c
1
(c

1
) destroys (creates) an electron on orbital 1. If we rewrite
this in matrix form
H =
_
c

1
, c

2
_
_

_
t
t
_

_
_
_
_
_
c
1
c
2
_
_
_
_
(3)
then it is apparent that system has eigenenergies t. Thus the two
states split their degeneracy, the splitting is proportional to |t|, and
they remain centered at
7
If we continue this process of bringing in more isolated orbitals into
the region where they can hybridize with the others, then a band of
states is formed, again with width proportional to t, and centered
around (cf. Fig. 5). This, of course, is an oversimplication. Real
+ + + +
. . .
=
Band
E
Figure 5: If we bring many orbitals into proximity so that they may exchange electrons
(hybridize), then a band is formed centered around the location of the isolated orbital,
and with width proportional to the strength of the hybridization
.
solids are composed of elements with multiple orbitals that produce
multiple bonds. Now imagine what happens if we have several orbitals
on each site (ie s,p, etc.), as we reduce the separation between the
orbitals and increase their overlap, these bonds increase in width and
may eventually overlap, forming bands.
The valance orbitals, which generally have a greater spatial extent,
will overlap more so their bands will broaden more. Of course, even-
tually we will stop gaining energy (

t) from bringing the atoms closer


together, due to overlap of the cores. Once we have reached the optimal
8
point we ll the states 2 particles per, until we run out of electrons.
Electronic correlations complicate this simple picture of band forma-
tion since they tend to try to keep the orbitals from being multiply
occupied.
2 Dierent Types of Bonds
These complications aside, the overlap of the orbitals is bonding. The
type of bonding is determined to a large degree by the amount of over-
lap. Three dierent general categories of bonds form in solids (cf. Ta-
ble 2).
Bond Overlap Lattice constituents
Ionic very small (< a) closest unfrustrated dissimilar
packing
Covalent small ( a) determined by the similar
structure of the orbitals
Metallic very large (a) closest packed unlled valence
orbitals
Table 2: The type of bond that forms between two orbitals is dictated largely by the
amount that these orbitals overlap relative to their separation a.
9
2.1 Covalent Bonding
Covalent bonding is distinguished as being orientationally sensitive. It
is also short ranged so that the interaction between nearest neighbors is
of prime importance and that between more distant neighbors is often
neglected. It is therefore possible to describe many of its properties
using the chemistry of the constituent molecules.
Consider a simple diatomic molecule O
2
with a single electron and
H =

2
2m

Ze
2
r
a

Ze
2
r
b
+
Z
2
e
2
R
(4)
We will search for a variational solution to the the problem of the
molecule (H
mol
= E
mol
), by constructing a variational wavefunction
from the atomic orbitals
a
and
b
. Consider the variational molecular
wavefunction

= c
a

a
+c
b

b
(5)
E

=
_

E (6)
The best

is that which minimizes E

. We now dene the quantum


integrals
S =
_

b
H
aa
= H
bb
=
_

a
H
a
H
ab
=
_

a
H
b
. (7)
Note that 1 > S
r
> 0, and that H
abr
< 0 since
a
and
b
are bound
states [where S
r
= ReS and H
abr
= ReH
ab
]b. With these denitions,
E

=
(c
2
a
+c
2
b
)H
aa
+ 2c
a
c
b
H
abr
c
2
a
+c
2
b
+ 2c
a
c
b
S
r
(8)
10
and we search for an extremum
E

c
a
=
E

c
b
= 0. From the rst condition,
E

c
a
=
0 and after some simplication, and re-substitution of E

into the above


equation, we get the condition
c
a
(H
aa
E

) +c
b
(H
abr
E

S
r
) = 0 (9)
The second condition,
E

c
b
= 0, gives
c
a
(H
abr
E

S) +c
b
(H
aa
E

) = 0 . (10)
Together, these form a set of secular equations, with eigenvalues
E

=
H
aa
H
abr
1 S
r
. (11)
Remember, H
abr
< 0, so the lowest energy state is the + state. If we
substitute Eq. 10 into Eqs. 8 and 9, we nd that the + state corresponds
to the eigenvector c
a
= c
b
= 1/

2; i.e. it is the bonding state.

bonding
=
1

2
(
a
+
b
) E

bonding
=
H
aa
+H
abr
1 +S
r
. (12)
For the , or antibonding state, c
a
= c
b
= 1/

2. Thus, in the bond-


ing state, the wavefunctions add between the atoms, which corresponds
to a build-up of charge between the oxygen molecules (cf. Fig. 6). In the
antibonding state, there is a deciency of charge between the molecules.
Energetically the bonding state is lower and if there are two elec-
trons, both will occupy the lower state (ie., the molecule gains energy
by bonding in a singlet spin conguration!). Energy is lost if there are
more electrons which must ll the antibonding states. Thus the covalent
11
r
a
r
b
R
Ze Ze
e

bonding

anti-bonding
spin singlet
spin triplet
Figure 6: Two oxygen ions, each with charge Ze, bind and electron with charge e.
The electron, which is bound in the oxygen valence orbitals will form a covalent bond
between the oxygens
bond is only eective with partially occupied single-atomic orbitals. If
the orbitals are full, then the energy loss of occupying the antibonding
states would counteract the gain of the occupying the bonding state
and no bond (conventional) would occur. Of course, in reality it is
much worse than this since electronic correlation energies would also
increase.
The pile-up of charge which is inherent to the covalent bond is im-
portant for the lattice symmetry. The reason is that the covalent bond
is sensitive to the orientation of the orbitals. For example, as shown in
Fig. 7 an S and a P

orbital can bond if both are in the same plane;


12
whereas an S and a P

orbital cannot. I.e., covalent bonds are di-


rectional! An excellent example of this is diamond (C) in which the
S P

+
-
S

P
+ -
No bonding Bonding
Figure 7: A bond between an S and a P orbital can only happen if the P-orbital is
oriented with either its plus or minus lobe closer to the S-orbital. I.e., covalent bonds
are directional!
(tetragonal) lattice structure is dictated by bond symmetry. However
at rst sight one might assume that C with a 1s
2
2s
2
2p
2
conguration
could form only 2-bonds with the two electrons in the partially lled
p-shell. However, signicant energy is gained from bonding, and 2s and
2p are close in energy (cf. Fig. 2) so that sucient energy is gained
from the bond to promote one of the 2s electrons. A linear combination
of the 2s 2p
x
, 2p
y
and 2p
z
orbitals form a sp
3
hybridized state, and C
often forms structures (diamond) with tetragonal symmetry.
Another example occurs most often in transition metals where the
d-orbitals try to form covalent bonds (the larger s-orbitals usually form
metallic bonds as described later in this chapter). For example, consider
a set of d-orbitals in a metal with a face-centered cubic (fcc) structure,
13
as shown in Fig. 8. The xy, xz, and yz orbitals all face towards a
neighboring site, and can thus form bonds with these sites; however, the
x
2
y
2
and 3z
2
r
2
orbitals do not point towards neighboring sites and
therefore do not participate in bonding. If the metal had a simple cubic
structure, the situation would be reversed and the x
2
y
2
and 3z
2
r
2
orbitals, but not the xy, xz, and yz orbitals, would participate much
in the bonding. Since energy is gained from bonding, this energetically
favors an fcc lattice in the transition metals (although this may not be
the dominant factor determining lattice structure).
x
y
z
d
xz
d
xy
d
yz
x
y
z
x
y
z
x
y
z
x
y
z
d
x - y
2 2
d
3z - r
2 2
Face Centered
Cubic Structure
Figure 8: In the fcc structure, the xy, xz, and yz orbitals all face towards a neighboring
site, and can thus form bonds with these sites; however, the x
2
y
2
and 3z
2
r
2
orbitals
do not point towards neighboring sites and therefore do not participate in bonding
One can also form covalent bonds from dissimilar atoms, but these
will also have some ionic character, since the bonding electron will no
14
longer be shared equally by the bonding atoms.
2.2 Ionic Bonding
The ionic bond occurs by charge transfer between dissimilar atoms
which initially have open electronic shells and closed shells afterwards.
Bonding then occurs by Coulombic attraction between the ions. The
energy of this attraction is called the cohesive energy. This, when added
to the ionization energies yields the energy released when the solid is
formed from separated neutral atoms (cf. Fig. 9). The cohesive energy
is determined roughly by the ionic radii of the elements. For example,
for NaCl
E
cohesive
=
e
2
a
o
a
o
r
Na
+r
Cl
= 5.19eV . (13)
Note that this does not agree with the experimental gure given in
the caption of Fig. 9. This is due to uncertainties in the denitions of
the ionic radii, and to oversimplication of the model. However, such
calculations are often sucient to determine the energy of the ionic
structure (see below). Clearly, ionic solids are insulators since such a
large amount of energy 10eV is required for an electron to move
freely.
The crystal structure in ionic crystals is determined by balancing
the needs of keeping the unlike charges close while keeping like charges
apart. For systems with like ionic radii (i.e. CsCl, r
Cs
1.60

A
, r
Cl

1.81

A
) this means the crystal structure will be the closest unfrustrated
15
+
e-
Na
+ Na
Cl +
e-
Cl
-
Na
+
Cl
- + Na
+
Cl
- + 7.9 eV
+ 3.61 eV
+ 5.14 eV
r
r
Cl
Na
= 1.81
= 0.97
Figure 9: The energy per molecule of a crystal of sodium chloride is (7.9-
5.1+3.6) eV=6.4eV lower than the energy of the separated neutral atoms. The cohe-
sive energy with respect to separated ions is 7.9eV per molecular unit. All values on
the gure are experimental. This gure is from Kittel.
packing. Since the face-centered cubic (fcc) structure is frustrated (like
charges would be nearest neighbors), this means a body-centered cubic
(bcc) structure is favored for systems with like ionic radii (see Fig. 10).
For systems with dissimilar radii like NaCl (cf. Fig. 9), a simple cubic
structure is favored. This is because the larger Cl atoms requires more
room. If the cores approach closer than their ionic radii, then since
they are lled cores, a covalent bond including both bonding and anti-
16
bonding states would form. As discussed before, Coulomb repulsion
makes this energetically unfavorable.
Cubic
Body Centered
Cubic
Face Centered
Cubic
a
b
c
a
b
c
Na
Cl
Cs
Cl
Figure 10: Possible salt lattice structures. In the simple cubic and bcc lattices all the
nearest neighbors are of a dierent species than the element on the site. These ionic
lattices are unfrustrated. However, it not possible to make an unfrustrated fcc lattice
using like amounts of each element.
2.2.1 Madelung Sums
This repulsive contribution to the total energy requires a fully-quantum
calculation. However, the attractive Coulombic contribution may be
easily calculated, and the repulsive potential modeled by a power-law.
Thus, the potential between any two sites i and j, is approximated by

ij
=
e
2
r
ij
+
B
r
n
ij
(14)
where the rst term describes the Coulombic interaction and the plus
(minus) sign is for the potential between similar (dissimilar) elements.
The second term heuristically describes the repulsion due to the over-
17
lap of the electronic clouds, and contains two free parameters n and B
(Kittel, pp. 6671, approximates this heuristic term with an exponen-
tial, Bexp (r
ij
/), also with two free parameters). These are usually
determined from ts to experiment. If a is the separation of nearest
neighbors, r
ij
= ap
ij
, and their are N sites in the system, then the total
potential energy may be written as
= N
i
= N
_
_

e
2
a

i=j

p
ij
+
B
a
n

i=j
1
p
n
ij
_
_
. (15)
The quantity A =

i=j

p
ij
, is known as the Madelung constant. A
depends upon the type of lattice only (not its size). For example
A
NaCl
= 1.748, and A
CsCl
= 1.763. Due to the short range of the
potential 1/p
n
, the second term may be approximated by its nearest
neighbor sum.
2.3 Metallic Bonding
Metallic bonding is characterized by at least some long ranged and non-
directional bonds (typically between s orbitals), closest packed lattice
structures and partially lled valence bands. From the rst character-
istic, we expect some of the valance orbitals to encompass many other
lattice sites, as discussed in Fig. 11. Thus, metallic bonds lack the
directional sensitivity of the covalent bonds and form non-directional
bonds and closest packed lattice structures determined by an optimal
lling of space. In addition, since the bands are composed of partially
18
3d
x - y
2 2
4S
Figure 11: In metallic Ni (fcc, 3d
8
4s
2
), the 4s and 3d bands (orbitals) are almost
degenerate (cf. Fig. 2) and thus, both participate in the bonding. However, the 4s
orbitals are so large compared to the 3d orbitals that they encompass many other
lattice sites, forming non-directional bonds. In addition, they hybridize weakly with
the d-orbitals (the dierent symmetries of the orbitals causes their overlap to almost
cancel) which in turn hybridize weakly with each other. Thus, whereas the s orbitals
form a broad metallic band, the d orbitals form a narrow one.
lled orbitals, it is always possible to supply a small external electric
eld and move the valence electrons through the lattice. Thus, metal-
lic bonding leads to a relatively high electronic conductivity. In the
transition metals (Ca, Sr, Ba) the d-band is narrow, but the s and p
bonds are extensive and result in conduction. Partially lled bands can
occur by bond overlap too; ie., in Be and Mg since here the full S bonds
overlap with the empty p-bands.
19
2.4 Van der Waals Bonds
As a nal subject involving bonds, consider solids formed of Noble gases
or composed of molecules with saturated orbitals. Here, of course, there
is neither an ionic nor covalent bonding possibility. Furthermore, if the
charge distributions on the atoms were rigid, then the interaction be-
tween atoms would be zero, because the electrostatic potential of a
spherical distribution of electronic charge is canceled outside a neutral
atom by the electrostatic potential of the charge on the nucleus. Bond-
ing can result from small quantum uctuations in the charge which
induce electric dipole moments.
P
1
P
2
n
+
-
+
-
x
1
R
x
2
Figure 12: Noble gasses and molecules with saturated orbitals can form short ranged
van der Waals bonds by inducing uctuating electric dipole moments in each other.
This may be modeled by two harmonic oscillators binding a positive and negative
charge each.
As shown in Fig. 12 we can model the constituents as either induced
dipoles, or more correctly, dipoles formed of harmonic oscillators. Sup-
20
pose a quantum uctuation on 1 induces a dipole moment p
1
. Then
dipole 1 exerts a eld
E
1
=
3n(p
1
n) p
1
r
3
(16)
which is felt by 2, which in turn induces a dipole moment p
2
E
1

1/r
3
. This in turn, generates a dipole eld E
2
felt by 1 p
2
/r
3
1/r
6
.
Thus, the energy of the interaction is very small and short ranged.
W = p
1
E
2
1/r
6
(17)
2.4.1 Van der Waals-London Interaction
Of course, a more proper treatment of the van der Waals interaction
should account for quantum eects in induced dipoles modeled as har-
monic oscillators (here we follow Kittel).
As a model we consider two identical linear harmonic oscillators 1
and 2 separated by R. Each oscillator bears charges e with separations
x
1
and x
2
, as shown in Fig. 12. The particles oscillate along the x axis
with frequency
0
(the strongest optical absorption line of the atom),
and momenta P
1
and P
2
. If we ignore the interaction between the
charges (other than the self-interaction between the dipoles charges
which is accounted for in the harmonic oscillator potentials), then the
Hamiltonian of the system is
H
0
=
P
2
1
+P
2
2
2m
+
1
2
m
2
0
(x
2
1
+x
2
2
) . (18)
21
If we approximate each pair of charges as point dipoles, then they
interact with a Hamiltonian
H
1

3(p
2
n)(p
1
n) +p
1
p
2
|x
1
+Rx
2
|
3
=
2p
1
p
2
R
3
=
2e
2
x
1
x
2
R
3
. (19)
The total Hamiltonian H
0
+H
1
can be diagonalized a normal mode
transformation that isolates the the symmetric mode (where both os-
cillators move together) from the antisymmetric one where they move
in opposition
x
s
= (x
1
+x
2
)/

2 x
a
= (x
1
x
2
)/

2 (20)
P
s
= (P
1
+P
2
)/

2 P
a
= (P
1
P
2
)/

2 (21)
After these substitutions, the total Hamiltonian becomes
H =
P
2
s
+P
2
a
2m
+
1
2
_
_
m
2
0

2e
2
R
3
_
_
x
2
s
+
1
2
_
_
m
2
0
+
2e
2
R
3
_
_
x
2
a
(22)
The new eigenfrequencies of these two modes are then

s
=
_
_

2
0

2e
2
mR
3
_
_
1/2

a
=
_
_

2
0
+
2e
2
mR
3
_
_
1/2
(23)
The zero point energy of the system is now
E
0
=
1
2
h(
s
+
a
) h
0
_
_
_1
1
4
_
_
2e
2
m
2
0
R
3
_
_
2
+
_
_
_ (24)
or, the zero point energy is lowered by the dipole interaction by an
amount
U
h
0
4
_
_
2e
2
m
2
0
R
3
_
_
2
(25)
22
which is typically a small fraction of an electron volt.
This is called the Van der Waals interaction, known also as the
London interaction or the induced dipole-dipole interaction. It is the
principal attractive interaction in crystals of inert gases and also in
crystals of many organic molecules. The interaction is a quantum eect,
in the sense that U 0 as h 0.
23
Chapter 2: Crystal Structures and Symmetry
Laue, Bravais
December 28, 2001
Contents
1 Lattice Types and Symmetry 3
1.1 Two-Dimensional Lattices . . . . . . . . . . . . . . . . . 3
1.2 Three-Dimensional Lattices . . . . . . . . . . . . . . . . 5
2 Point-Group Symmetry 6
2.1 Reduction of Quantum Complexity . . . . . . . . . . . . 6
2.2 Symmetry in Lattice Summations . . . . . . . . . . . . . 7
2.3 Group designations . . . . . . . . . . . . . . . . . . . . . 11
3 Simple Crystal Structures 13
3.1 FCC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2 HCP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3 BCC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1
A theory of the physical properties of solids would be practically
impossible if the most stable elements were not regular crystal lattices.
The N-body problem is reduced to manageable proportions by the ex-
istence of translational symmetry. This means that there exist a set
of basis vectors (a,b,c) such that the atomic structure remains invari-
ant under translations through any vector which is the sum of integral
multiples of these vectors. As shown in Fig. 1 this means that one may
go from any location in the lattice to an identical location by following
path composed of integral multiples of these vectors.
a
b

Figure 1: One may go from any location in the lattice to an identical location by
following path composed of integral multiples of the vectors a and b.
Thus, one may label the locations of the atoms
1
. which compose
the lattice with
r
n
= n
1
a + n
2
b + n
3
c (1)
1
we will see that the basic building blocks of periodic structures can be more complicated than
a single atom. For example in NaCl, the basic building block is composed of one Na and one Cl ion
which is repeated in a cubic pattern to make the NaCl structure
2
where n
1
, n
2
, n
3
are integers. In this way we may construct any periodic
structure.
1 Lattice Types and Symmetry
1.1 Two-Dimensional Lattices
These structures are classied according to their symmetry. For ex-
ample, in 2d there are 5 distinct types. The lowest symmetry is an
oblique lattice, of which the lattice shown in Fig. 1 is an example if
a = b and is not a rational fraction of . Notice that it is invari-
|a| = |b|, = /2
Square
|a| = |b|, = /2
Rectangular
Hexangonal
|a| = |b|, = /3
Centered
a
b
a
b
a
b
a
b
Figure 2: Two dimensional lattice types of higher symmetry. These have higher
symmetry since some are invariant under rotations of 2/3, or 2/6, or 2/4, etc.
The centered lattice is special since it may also be considered as lattice composed of a
two-component basis, and a rectangular unit cell (shown with a dashed rectangle).
3
ant only under rotation of and 2. Four other lattices, shown in
Fig. 2 of higher symmetry are also possible, and called special lattice
types (square, rectangular, centered, hexagonal). A Bravais lattice is
the common name for a distinct lattice type. The primitive cell is the
parallel piped (in 3d) formed by the primitive lattice vectors which are
dened as the lattice vectors which produce the primitive cell with the
smallest volume (a (c c)). Notice that the primitive cell does not
always capture the symmetry as well as a larger cell, as is the case with
the centered lattice type. The centered lattice is special since it may
also be considered as lattice composed of a two-component basis on a
rectangular unit cell (shown with a dashed rectangle).
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Basis
b
a
Primitive
cell and
lattice
vectors
Figure 3: A square lattice with a complex basis composed of one Cu and two O atoms
(c.f. cuprate high-temperature superconductors).
4
To account for more complex structures like molecular solids, salts,
etc., one also allows each lattice point to have structure in the form of
a basis. A good example of this in two dimensions is the CuO
2
planes
which characterize the cuprate high temperature superconductors (cf.
Fig. 3). Here the basis is composed of two oxygens and one copper
atom laid down on a simple square lattice with the Cu atom centered
on the lattice points.
1.2 Three-Dimensional Lattices
a = x
b = y
c = z
a = (x+y-z)/2
b = (-x+y+z)/2
c = (x-y+z)/2
a = (x+y)/2
b = (x+z)/2
c = (y+z)/2
Cubic
Body Centered
Cubic
Face Centered
Cubic
a
b
c
a
b
c
Figure 4: Three-dimensional cubic lattices. The primitive lattice vectors (a,b,c) are
also indicated. Note that the primitive cells of the centered lattice is not the unit cell
commonly drawn.
The situation in three-dimensional lattices can be more complicated.
Here there are 14 lattice types (or Bravais lattices). For example there
5
are 3 cubic structures, shown in Fig. 4. Note that the primitive cells of
the centered lattice is not the unit cell commonly drawn. In addition,
there are triclinic, 2 monoclinic, 4 orthorhombic ... Bravais lattices, for
a total of 14 in three dimensions.
2 Point-Group Symmetry
The use of symmetry can greatly simplify a problem.
2.1 Reduction of Quantum Complexity
If a Hamiltonian is invariant under certain symmetry operations, then
we may choose to classify the eigenstates as states of the symmetry
operation and H will not connect states of dierent symmetry.
As an example, imagine that a symmetry operation R leaves H
invariant, so that
RHR
1
= H then [H, R] = 0 (2)
Then if |j > are the eigenstates of R, then

j
|j >< j| is a repre-
sentation of the identity, and we expand HR = RH, and examine its
elements

k
< i|R|k >< k|H|j >=

k
< i|H|k >< k|R|j > . (3)
If we recall that R
ik
=< i|R|k >= R
ii

ik
since |k > are eigenstates of
6
R, then Eq. 3 becomes
(R
ii
R
jj
) H
ij
= 0 . (4)
So, H
ij
= 0 if R
i
and R
j
are dierent eigenvalues of R. Thus, when the
states are classied by their symmetry, the Hamiltonian matrix becomes
Block diagonal, so that each block may be separately diagonalized.
2.2 Symmetry in Lattice Summations
As another example, consider a Madelung sum in a two-dimensional
square centered lattice (i.e. a 2d analog of NaCl). Here we want to
calculate

ij

p
ij
. (5)
This may be done by a brute force sum over the lattice, i.e.
lim
n

i=n,nj=n,n
(1)
i+j
p
ij
. (6)
Or, we may realize that the lattice has some well dened operations
which leave it invariant. For example, this lattice in invariant under in-
version (x, y) (x, y), and reections about the x (x, y) (x, y)
and y (x, y) (x, y) axes, etc. For these reasons, the eight points
highlighted in Fig. 5(a) all contribute an identical amount to the sum
in Eq. 5. In fact all such interior points have a degeneracy of 8. Only
special points like the point at the origin (which is unique) and points
along the symmetry axes (the xy and x axis, each with a degeneracy of
7
4
4
1
8
(a)
(b)
O
Figure 5: Equivalent points and irreducible wedge for the 2-d square lattice. Due to
the symmetry of the 2-d square lattice, the eight patterned lattice sites all contribute
an identical amount to the Madelung sum calculated around the solid black site. Due
to this symmetry, the sum can be reduced to the irreducible wedge (b) if the result at
each point is multiplied by the degeneracy factors indicated.
four) have lower degeneracies. Thus, the sum may be restricted to the
irreducible wedge, so long as the corresponding terms in the sum are
multiplied by the appropriate degeneracy factors, shown in Fig. 5(b).
An appropriate algorithm to calculate both the degeneracy table, and
the sum 5 itself are:
c First calculate the degeneracy table
c
do i=1,n
do j=0,i
if(i.eq.j.or.j.eq.0) then
8
deg(i,j)=4
else
deg(i,j)=8
end if
end do
end do
deg(0,0)=1
c
c Now calculate the Madelung sum
c
sum=0.0
do i=1,n
do j=0,i
p=sqrt(i**2+j**2)
sum=sum+((-1)**(i+j))*deg(i,j)/p
end do
end do
By performing the sum in this way, we saved a factor of 8! In fact, in
three-dimensions, the savings is much greater, and real band structure
calculations (eg. those of F.J. Pinski) always make use of the point
group symmetry to accelerate the calculations.
The next question is then, could we do the same thing for a more
complicated system (fcc in 3d?). To do this, we need some way of
9
classifying the symmetries of the system that we want to apply. Group
theory allows us to learn the consequences of the symmetry in much
more complicated systems.
A group S is dened as a set {E, A, B, C } which is closed under
a binary operation (ie. A B S) and:
the binary operation is associative (A B) C = A (B C)
there exists an identity E S : E A = A E = A
For each A S, there exists an A
1
S : AA
1
= A
1
A = E
In the point group context, the operations are inversions, reections,
rotations, and improper rotations (inversion rotations). The binary op-
eration is any combination of these; i.e. inversion followed by a rotation.
In the example we just considered we may classify the operations
that we have already used. Clearly we need 2!2
2
of these (ie we can
choose to take (x,y) to any permutation of (x,y) and choose either for
each, in D-dimensions, there would be D!2
D
operations). In table. 1,
all of these operations are identied The reections are self inverting as
is the inversion and one of the rotations and inversion rotations. The
set is clearly also closed. Also, since their are 8 operations, clearly the
interior points in the irreducible wedge are 8-fold degenerate (w.r.t. the
Madelung sum).
This is always the case. Using the group operations one may always
reduce the calculation to an irreducible wedge. They the degeneracy of
10
Operation Identication
(x, y) (x, y) Identity
(x, y) (x, y) reection about x axis
(x, y) (x, y) reection about y axis
(x, y) (x, y) inversion
(x, y) (y, x) reection about x = y
(x, y) (y, x) rotation by /2 about z
(x, y) (y, x) inversion-reection
(x, y) (y, x) inversion-rotation
Table 1: Point group symmetry operations for the two-dimensional square lattice.
All of the group elements are self-inverting except for the sixth and eight, which are
inverses of each other.
each point in the wedge may be determined: Since a group operation
takes a point in the wedge to either itself or an equivalent point in the
lattice, and the former (latter) does (does not) contribute the the de-
generacy, the degeneracy of each point times the number of operations
which leave the point invariant must equal the number of symmetry
operations in the group. Thus, points with the lowest symmetry (in-
variant only under the identity) have a degeneracy of the group size.
2.3 Group designations
Point groups are usually designated by their Sch onies point group
symbol described in table. 2 As an example, consider the previous ex-
11
Symbol Meaning
C
j
(j=2,3,4, 6) j-fold rotation axis
S
j
j-fold rotation-inversion axis
D
j
j 2-fold rotation axes to a j-fold principle rotation axis
T 4 three-and 3 two-fold rotation axes, as in a tetrahedron
O 4 three-and 3 four-fold rotation axes, as in a octahedron
C
i
a center of inversion
C
s
a mirror plane
Table 2: The Sch onies point group symbols. These give the classication according
to rotation axes and principle mirror planes. In addition, their are suxes for mirror
planes (h: horizontal=perpendicular to the rotation axis, v: vertical=parallel to the
main rotation axis in the plane, d: diagonal=parallel to the main rotation axis in the
plane bisecting the two-fold rotation axes).
ample of a square lattice. It is invariant under
rotations to the page by /2
mirror planes in the horizontal and vertical (x and y axes)
mirror planes along the diagonal (x=y, x=-y).
The mirror planes are parallel to the main rotation axis which is itself
a 4-fold axis and thus the group for the square lattice is C
4v
.
12
3 Simple Crystal Structures
3.1 FCC
a = (x+y)/2
b = (x+z)/2
c = (y+z)/2
Face Centered Cubic (FCC)
a
b
c
Principle lattice vectors Close-packed planes
3-fold axes
4-fold axes
x
y
z
Figure 6: The Bravais lattice of a face-centered cubic (FCC) structure. As shown
on the left, the fcc structure is composed of parallel planes of atoms, with each atom
surrounded by 6 others in the plane. The total coordination number (the number of
nearest neighbors) is 12. The principle lattice vectors (center) each have length 1/

2
of the unit cell length. The lattice has four 3-fold axes, and three 4-fold axes as shown
on the right. In addition, each plane shown on the left has the principle 6-fold rotation
axis to it, but since the planes are shifted relative to one another, they do not share
6-fold axes. Thus, four-fold axes are the principle axes, and since they each have a
perpendicular mirror plane, the point group for the fcc lattice is O
h
.
The fcc structure is one of the close packed structures, appropriate
for metals, with 12 nearest neighbors to each site (i.e., a coordination
number of 12). The Bravais lattice for the fcc structure is shown in
Fig. 6 It is composed of parallel planes of nearest neighbors (with six
13
nearest neighbors to each site in the plane)
Metals often form into an fcc structure. There are two reasons for
this. First, as discussed before, the s and p bonding is typically very
long-ranged and therefore rather non-directional. (In fact, when the
p-bonding is short ranged, the bcc structure is favored.) This naturally
leads to a close packed structure. Second, to whatever degree there is a
d-electron overlap in the transition metals, they prefer the fcc structure.
To see this, consider the d-orbitals shown in Fig. 7 centered on one of
the face centers with the face the xy plane. Each lobe of the d
xy
, d
yz
,
and d
xz
orbitals points to a near neighbor. The xz,xy,yz triplet form
rather strong bonds. The d
x
2
y
2 and d
3z
2
r
2 orbitals do not since they
point away from the nearest neighbors. Thus the triplet of states form
strong bonding and anti-bonding bands, while the doublet states do
not split. The system can gain energy by occupying the triplet bonding
states, thus many metals form fcc structures. According to Ashcroft
and Mermin, these include Ca, Sr, Rh, Ir, Ni, Pd, Pt, Cu, Ag, Au, Al,
and Pb.
The fcc structure also explains why metals are ductile since adjacent
planes can slide past one another. In addition each plane has a 6-fold
rotation axis perpendicular to it, but since 2 adjacent planes are shifted
relative to another, the rotation axes perpendicular to the planes are
3-fold, with one along the each main diagonal of the unit cell. There
are also 4-fold axes through each center of the cube with mirror planes
14
x
y
z
d
xz
d
xy
d
yz
x
y
z
x
y
z
x
y
z
x
y
z
d
x - y
2 2
d
3z - r
2 2
Figure 7: The d-orbitals. In an fcc structure, the triplet of orbitals shown on top all
point towards nearest neighbors; whereas, the bottom doublet point away. Thus the
triplet can form bonding and antibonding states.
perpendicular to it. Thus the fcc point group is O
h
. In fact, this same
argument also applies to the bcc and sc lattices, so O
h
is the appropriate
group for all cubic Bravais lattices and is often called the cubic group.
3.2 HCP
As shown in Fig. 8 the Hexagonal Close Packed (HCP) structure is
described by the D
3h
point group. The HCP structure (cf. Fig. 9) is
similar to the FCC structure, but it does not correspond to a Bravais
lattice (in fact there are ve cubic point groups, but only three cubic
Bravais lattices). As with fcc its coordination number is 12. The sim-
plest way to construct it is to form one hexagonal plane and then add
two identical ones top and bottom. Thus its stacking is ABABAB...
15
mirror plane
3-fold
axis
three 2-fold
axes in plane
Figure 8: The symmetry of the HCP lattice. The principle rotation axis is perpen-
dicular to the two-dimensional hexagonal lattices which are stacked to form the hcp
structure. In addition, there is a mirror plane centered within one of these hexagonal
2d structures, which contains three 2-fold axes. Thus the point group is D
3h
.
of the planes. This shifting of the planes clearly disrupts the d-orbital
bonding advantage gained in fcc, nevertheless many metals form this
structure including Be, Mg, Sc, Y, La, Ti, Zr, Hf, Tc, Re, Ru, Os, Co,
Zn, Cd, and Tl.
3.3 BCC
Just like the simple cubic and fcc lattices, the body-centered cubic
(BCC) lattice (cf. Fig. 4) has four 3-fold axes, 3 4-fold axes, with mirror
planes perpendicular to the 4-fold axes, and therefore belongs to the
O
h
point group.
The body centered cubic structure only has a coordination number
of 8. Nevertheless some metals form into a BCC lattice (Ba V Nb, Ta
W M, in addition Cr and Fe have bcc phases.) Bonding of p-orbitals is
16
A A A A A
A A A A A
A A A A A
FCC
HCP
These spaces unfilled
B B B B B
B B B B B
B B B B B
C C C C C
C C C C C
C C C C C
A A A A A
A A A A A
A A A A A
B B B B B
B B B B B
B B B B B
Figure 9: A comparison of the FCC (left) and HCP (right) close packed structures.
The HCP structure does not have a simple Bravais unit cell, but may be constructed
by alternately stacking two-dimensional hexagonal lattices. In contract, the FCC
structure may be constructed by sequentially stacking three shifted hexagonal two-
dimensional lattices.
ideal in a BCC lattice since the nnn lattice is simply composed of two
interpenetrating cubic lattices. This structure allows the next-nearest
neighbor p-orbitals to overlap more signicantly than an fcc (or hcp)
structure would. This increases the eective coordination number by
including the next nearest neighbor shell in the bonding (cf. Fig. 10).
17
2s,2p
R(r)
fcc
bcc
12 6 24
8 6 12
1s
0 1 2 3
1
2
r(A)
o
Figure 10: Absolute square of the radial part of the electronic wavefunction. For the
bcc lattice, both the 8 nearest, and 6 next nearest neighbors lie in a region of relatively
high electronic density. This favors the formation of a bcc over fcc lattice for some
elemental metals (This gure was lifted from I&L).
18
Chapter 3: The Classical Theory of Crystal
Diraction
Bragg
December 29, 2001
Contents
1 Classical Theory of diraction 4
2 Scattering from Periodic Structures 8
2.1 The Scattering Intensity for a Crystal . . . . . . . . . . . 10
2.2 Bragg and Laue Conditions (Miller Indices) . . . . . . . 12
2.3 The Structure Factor . . . . . . . . . . . . . . . . . . . . 16
2.3.1 The Structure Factor of Centered Lattices . . . . 19
2.3.2 Powdered x-ray Diraction . . . . . . . . . . . . . 22
1
In the last two chapters, we learned that solids generally form pe-
riodic structures of dierent symmetries and bases. However, given a
solid material, how do we learn what its periodic structure is? Typ-
ically, this is done by diraction, where we project a beam (of either
particles or radiation) at a solid with a wavelength the characteris-
tic length scale of the lattice ( a twice the atomic or molecular radii
of the constituents). Diraction of waves and particles (with de Broglie
| | or | |
incident waves
or particles
d
d sin()
a
2
a
1
a
1
a
2
k
0
k
k
0
K = k -
k
0
k
K

Figure 1: Scattering of waves or particles with wavelength of roughly the same size
as the lattice repeat distance allows us to learn about the lattice structure. Coherent
addition of two particles or waves requires that 2d sin = (the Bragg condition),
and yields a scattering maximum on a distant screen.
wavelength = h/p) of a allows us to learn about the periodic
structure of crystals. In a diraction experiment one identies Bragg
peaks which originate from a coherent addition of scattering events in
2
multiple planes within the bulk of the solid.
However, not all particles with de Broglie wavelength a will
work for this application. For example, most charged particles cannot
probe the bulk properties of the crystal, since they lose energy to the
scatterer very quickly. Recall, from classical electrodynamics, the rate
at which particles of charge q, mass M, and velocity v lose energy to
the electrons of charge e and mass m in the crystal is given roughly by
dE
dx

4nq
2
e
2
mv
2
ln
_
_
m
2
v
3
qe
0
_
_

q
2
v
2
. (1)
As an example, consider a non-relativistic electron scattering into a
solid with a 2A. If we require that a = = h/p = 12.310
8
cm/

E
when E is measured in electron volts, then E 50eV. If we solve
Eq. 1 for the distance x where the initial energy of the incident is lost
requiring that E = E, when n 10
23
/cm
3
we nd that x 100A.
Thus, if a, the electrons do not penetrate into the bulk of the
sample (typically the rst few hundred Aof most materials are oxidized,
or distorted by surface reconstruction of the dangling bonds at the
surface, etc. See Fig. 2) Thus, electrons do not make a very good probe
of the bulk properties of a crystal (instead in a process call low-energy
electron diraction, LEED, they may be used to study the surface of
especially clean samples. I.e. to study things like surface reconstruction
of the dangling bonds, etc.). Thus although they are obviously easier
to accelerate (electrons or ion beams), they generally do not penetrate
into the bulk and so tell us more about the surface properties of solids
3
v
e
-
Oxygen
Figure 2: An electron about to scatter from a typical material. However, at the surface
of the material, oxidation and surface reconstruction distort the lattice. If the electron
scatters from this region, we cannot learn about the structure of the bulk.
which are often not representative of the bulk.
Thus the particle of the choice to determine bulk properties is the
neutron which is charge neutral and scatters only from the nuclei. Ra-
diation is often also used. Here the choice is only a matter of the
wavelength used. X-rays are chosen since then a
1 Classical Theory of diraction
In this theory of diraction we will be making three basic assumptions.
1. That the operator which describes the coupling of the target to
the scattered object (in this case the operator is the density)
4
commutes with the Hamiltonian. Thus, this will be a classical
theory.
2. We will assume some form of Huygens principle: that every radi-
ated point of the target will serve as a secondary source spherical
wavelets of the same frequency as the source and the amplitude
of the diracted wave is the sum of the wavelengths considering
their amplitudes and relative phases. (For light, this is equivalent
to assuming that it is unpolarized, and that the diraction pattern
varies quickly with scattering angle so that the angular depen-
dence of a unpolarized dipole, 1 + (cos )
2
, may be neglected.)
3. We will assume that resulting spherical waves are not scattered
again. In the fully quantum theory which we will derive later
for neutron scattering, this will correspond to approximating the
scattering rate by Fermis golden rule (rst-order Born approxi-
mation).
The basic setup of a scattering experiment is sketched in Fig. 3.
Generally, we will also assume that |R| |r|, so that we may always
approximate the amplitude of the incident waves on the target as plane
waves.
A
P
= A
O
e
i(k
0
(R+r)
0
t)
. (2)
5
Q
source
P
r
R - r
B
target
observer
or screen
R
R
Figure 3: Basic setup of a scattering experiment.
Then, consistent with the second assumption above,
A
B
(R

)
_
d
3
rA
P
(r)
e
ik(R

r)
|R

r|
, (3)
which, after substitution of Eq.2, becomes
A
B
(R

) A
O
e
i(k
0
R+kR

0
t)
_
d
3
r(r)
e
i(kk
o
)r
|R

r|
. (4)
At very large R

(ie. in the radiation or far zone)


A
B
(R

)
A
O
e
i(k
0
R+kR

0
t)
R

_
d
3
r(r)e
i(kk
o
)r
. (5)
Or, in terms of the scattered intensity I
B
|A
B
|
2
I
B

|A
O
|
2
R
2

_
d
3
r(r)e
i(kk
o
)r

2
. (6)
The scattering intensity is just the absolute square of the Fourier trans-
form of the density of scatterers. If we let K = kk
0
(cf. Fig. 1), then
we get
I
B
(K)
|A
O
|
2
R
2

_
d
3
r(r)e
iKr

2
=
|A
O
|
2
R
2
|(K)|
2
. (7)
6
From the associated Fourier uncertainty principle kx , we can
see that the resolution of smaller structures requires larger values of K
(some combination of large scattering angles and short wavelength of
the incident light), consistent with the discussion at the beginning of
this chapter.
I(K) (r)
Figure 4: Since the measured scattering intensity I(K) |(K)|
2
the complex phase
information is lost. Thus, a scattering experiment does not provide enough informa-
tion to invert the transform (r) =
_
d
3
r
(2)
3
(K)e
+iKr
.
In experiments the intensity I as a function of the scattering angle K
is generally measured. In principle this is under-complete information.
In order to invert the Fourier transform (which is a unitary transfor-
mation) we would need to know both the real and imaginary parts
of
(K) =
_
d
3
r(r)e
iKr
. (8)
Of course, if the experiment just measures I |(K)|
2
, then we lose
the relative phase information (i.e. (K) =
K
e
i
K
so that I |
K
|
2
,
and the phase information
K
is lost). So, from a complete experiment,
measuring I(K) for all scattering angles, we do not have enough infor-
7
mation to get a unique (r) by inverting the Fourier transform. Instead
experimentalists analyze their data by proposing a feasible model struc-
ture (i.e. a (r) corresponding to some guess of which of one the 14 the
Bravais lattice and the basis), Fourier transform this, and compare it to
the experimental data. The parameters of the model are then adjusted
to obtain a best t.
2 Scattering from Periodic Structures
Given this procedure, it is important to study the scattering pattern
that would arise for various periodic structures. The density in a peri-
odic crystal must have the same periodicity of the crystal
(r +r
n
) = (r) where r
n
= n
1
a
1
+ n
2
a
2
+ n
3
a
3
(9)
for integer n
1
, n
2
, n
2
. This also implies that the Fourier coecients of
will be chosen from a discrete set. For example, consider a 1-d periodic
structure
a
(x + na) = (x) . (10)
Then we must choose the G
n
(x) =

n
e
iG
n
x
, (11)
8
so that
(x + ma) =

n
e
iG
n
(x+ma)
=

n
e
iG
n
(x)
e
iG
n
ma
=

n
e
iG
n
(x)
= (x) , (12)
I.e. e
iG
n
ma
= 1, or G
n
= 2n/a where n is an integer.
This may be easily generalized to three dimensions, for which
(r) =

G
e
iGr
(13)
where the condition of periodicity (r +r
n
) = (r) means that
G r
n
= 2m m Z (14)
where Z is the group of integers (under addition). Now, lets consider
G in some three-dimensional space and decompose it in terms of three
independent basis vectors for which any two are not parallel and the
set is not coplanar
G = hg
1
+ kg
2
+ lg
3
. (15)
The condition of periodicity then requires that
(hg
1
+ kg
2
+ lg
3
) n
1
a
1
= 2m m Z (16)
with similar conditions of the other principle lattice vectors a
2
and a
3
.
Since g
1
, g
2
and g
3
are not parallel or coplanar, the only way to satisfy
this constraint for arbitrary n
1
is for
g
1
a
1
= 2 g
2
a
1
= g
3
a
1
= 0 (17)
9
or some other permutation of 1 2 and 3, which would just amount to
a renaming of g
1
, g
2
, and g
3
. The set (g
1
, g
2
, g
3
) are called the basis
set for the reciprocal lattice. They may be constructed from
g
1
= 2
a
2
a
3
a
1
(a
2
a
3
)
plus cyclic permutations . (18)
It is easy to see that this construction satises Eq. 17, and that there
is a one to one correspondence between the lattice and its reciprocal
lattice. So, the reciprocal lattice belongs to the same point group as
the real-space lattice
1
.
2.1 The Scattering Intensity for a Crystal
Lets now apply this form for the density
(r) =

G
e
iGr
(19)
to our formula for the scattering intensity
I
B
(K)
|A
O
|
2
R
2

_
d
3
r

G
e
i(KG)r

2
(20)
The integral above is simply
V
G,K
=
_

_
V if G = K
0 if G = K
, (21)
1
One should note that this does not mean that the reciprocal lattice must have the same Bravais
lattice structure as the real lattice. For example, the reciprocal of a fcc lattic is bcc and vice versa.
This is consistent with the the statement that the reciprocal lattice belongs to the same point group
as the real-space lattice since fcc and bcc share the O
h
point group
10
where V is the lattice volume, so
I
B
(K)
|A
O
|
2
R
2
|
G
|
2
V
2

G,K
(22)
This is called the Laue condition for scattering. The fact that this is
proportional to V
2
rather than V just indicates that the diractions
spots, in this approximation, are innitely bright (for a sample in the
thermodynamic limit). Of course, this is because the spots are innitely
narrow or ne. When real broadening is taken into account, I
B
(K) V
as expected.
Then as G = hg
1
+kg
2
+lg
3
, we can label the spots with the three
integers (h, k, l ), or
I
hkl
|
hkl
|
2
. (23)
Traditionally, negative integers are cabled with an overbar, so h h.
Then as (r) is real,
G
=
G
, or
I
hkl
= I
hkl
Friedels rule (24)
Most scattering experiments are done with either a rotating crystal, or
a powder made up of many crystalites. For these experiments, Friedels
rule has two main consequences
For every spot at k k
0
= G, there will be one at k

k
0
=
G. Thus, for example if we scatter from a crystal with a 3-fold
symmetry axis, we will get a six-fold scattering pattern. Clearly
this can only happen, satisfy the Laue condition, and have |k| =
11
|k
0
|, if the crystal is rotated by in some axis perpendicular to the
three-fold axis. In fact, single-crystal experiments are usually done
either by mounting the crystal on a precession stage (essentially
like an automotive universal joint, with the drive shaft held xed,
and the joint rotated over all angles), or by holding the crystal xed
and moving the source and diraction screen around the crystal.
The scattering pattern always has an inversion center, G G
even when none is present in the target!
2.2 Bragg and Laue Conditions (Miller Indices)
Above, we derived the Laue condition for scattering; however, we began
this chapter by reviewing the Bragg condition for scattering from ad-
jacent planes. In this subsection we will show that, as expected, these
are the some condition.
Consider the real-space lattice shown in Fig. 5. Highlighted by the
solid lines are the parallel planes formed by (1, 2, 2) translations along
the principle lattice vectors (a
1
, a
2
, a
3
), respectively. Typically these
integers are labeled by (m, n, o), however, the plane is not typically
labeled as the (m, n, o) plane. Rather it is labeled with the inverses
h

= 1/m k

= 1/n l

= 1/o . (25)
Since these typically are not integers, they are multiplied by p, the
12
(211) plane
a
1
a
2
a
3
O
a
1

G
hkl
d
hkl
hkl plane
Figure 5: Miller indices identication of planes in a lattice. Highlighted by the solid
lines are the parallel planes formed by (1, 2, 2) translations along the principle lattice
vectors (a
1
, a
2
, a
3
), respectively. Typically these integers are labeled by (m, n, o), how-
ever, the plane is not typically labeled as the (m, n, o) plane. Rather it is labeled with
the inverses h

= 1/m k

= 1/n l

= 1/o. Since these typically are not integers,


they are multiplied by p, the smallest integer such that p(h

, k

, l

) = (h, k, l) Z. In
this case, p = 2, and the plane is labeled as the (2, 1, 1) plane. Note that the plane
formed by (2, 4, 4) translations along the principle lattice vectors is parallel to the
(2, 1, 1) plane.
smallest integer such that
p(h

, k

, l

) = (h, k, l) Z . (26)
In this case, p = 2, and the plane is labeled as the (2, 1, 1) plane.
On may show that the reciprocal lattice vector G
hkl
lies perpendicu-
lar to the (h, k, l) plane, and that the length between adjacent parallel
planes d
hkl
= 2/G
hkl
. To show this, note that the plane may be dened
13
by two non-parallel vectors v
1
and v
2
within the plane. Let
v
1
= ma
1
na
2
= a
1
/h

a
2
/k

v
2
= oa
3
na
2
= a
3
/l

a
2
/k

. (27)
Clearly the cross product, v
1
v
2
is perpendicular to the (h, k, l) plane
v
1
v
2
=
a
3
a
1
h


a
1
a
2
h


a
2
a
3
k

. (28)
If we multiply this by 2h

/a
1
(a
2
a
3
), we get
2h

v
1
v
2
a
1
(a
2
a
3
)
=
2p
p
_
_
k

a
3
a
1
a
1
(a
2
a
3
)
+ l

a
1
a
2
a
1
(a
2
a
3
)
+ h

a
2
a
3
a
1
(a
2
a
3
)
_
_
=
G
hkl
/p (29)
Thus, G
hkl
to the (h, k, l) plane. Now, if is the angle between a
1
and G
hkl
, then the distance d
h

l
from the origin to the (h

, k

, l

) plane
is given by
d
h

l
= m|a
1
| cos =
|a
1
|a
1
G
hkl
h

|a
1
||G
hkl
|
=
2h
h

G
hkl
=
2p
G
hkl
(30)
Then, as there are p planes in this distance (cf. Fig. 5), the distance to
the nearest one is
d
hkl
= d
h

,k

,l
/p = 2/G
hkl
(31)
With this information, we can reexamine the Laue scattering con-
dition K = k k
o
= G
hkl
, and show that it is equivalent to the
more intuitive Bragg condition. Part of the Laue is condition is that
|K| = K = |k k
0
| = G
hkl
, now
K = 2k
0
sin =
4

sin and G
hkl
= 2/d
hkl
(32)
14
thus, the Laue condition implies that
1/d
hkl
= 2 sin / or = 2d
hkl
sin (33)
which is the Bragg condition. Note that the Laue condition is more
G

k
k
k
0
hkl plane
hkl
Laue Condition
(in reciprocal space)
Bragg Condition
(in real space)
k
k
0
d
hkl
d sin()
hkl
2d sin() =
hkl
K= k - k = G
hkl 0
Figure 6: Comparison of the Bragg = 2d
hkl
sin and Laue G
hkl
= K
hkl
conditions
for scattering.
restrictive than the Bragg condition; it requires that both the magnitude
and the direction of G and K be the same. However, there is no
inconsistency here, since whenever we apply the Bragg condition, we
assume that the plane dened by k and k
0
is perpendicular to the
scattering planes (cf. Fig. 6).
15
Cu Cu Cu
Cu Cu Cu
Cu Cu
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
O
Body Centered
Cubic
basis
O
r
n
r
r

Figure 7: Examples of lattices with non-trivial bases. The CuO


2
lattice (left) is char-
acteristic of the cuprate high-temperature superconductors. It has a basis composed of
one Cu and two O atoms imposed on a simple cubic lattice. The BCC lattice(right)
can be considered as a cubic lattice with a basis including an atom at the corner and
one at the center of the cube.
2.3 The Structure Factor
Thus far, we have concentrated on the diraction pattern for a periodic
lattice ignoring the ne structure of the molecular of the basis. Exam-
ples of non-trivial molecular bases are shown in Fig. 7. Clearly the basis
structure will eect the scattering (Fig. 8). For example, there will be
interference from the scattering o of the Cu and two O in each cell. In
fact, even in the simplest case of a single-element basis composed of a
spherical atom of nite extent, scattering from one side of the atom will
interfere with that from the other. In each case, the structure of the
basis will change the scattering pattern due to interference of the waves
16
scattering from dierent elements of the basis. The structure factors
account for these interference eects. The information about this in-
terference, and the basis structure is contained in the atomic scattering
factor f and the structure factor S.
Cu
O
O
R
Figure 8: Rays scattered from dierent elements of the basis, and from dierent places
on the atom, interfere giving the scattered intensity additional structure described by
the form factor S and the atomic form factor f, respectively.
To show this reconsider the scattering formula
I
hkl
|
hkl
|
2
(34)
The Fourier transform of the density may be decomposed into an inte-
gral over the basis cell, and a sum over all such cells

hkl
=
1
V
_
d
3
r(r)e
iG
hkl
r
=
1
V

cells
_
cell
d
3
r(r)e
iG
hkl
r
=
1
V

n
1
,n
2
,n
3
_
cell
d
3
r(r)e
iG
hkl
(r+r
n
)
(35)
17
where the location of each cell is given by r
n
= n
1
a
1
+ n
2
a
2
+ a
3
a
3
.
Then since G
hkl
r
n
= 2m, m Z,

hkl
=
N
V
_
cell
d
3
r(r)e
iG
hkl
r
(36)
where N is the number of cells and
N
V
= 1/V
c
, V
c
the volume of a
cell. This integral may be further subdivided into an integral over the
atomic density of each atom in the unit cell. If labeles the dierent
elements of the basis, each with density

(r

hkl
=
1
V
c

e
iG
hkl
r

_
d
3
r

(r

)e
iG
hkl
r

(37)
The atomic scattering factor f and the structure factor may then be
dened as parts of this integral
f

=
_
d
3
r

(r

)e
iG
hkl
r

(38)
so

hkl
=
1
V
c

e
iG
hkl
r

=
S
hkl
V
c
(39)
f

describes the interference of spherical waves emanating from dierent


points within the atom, and S
hkl
is called the structure factor. Note
that for lattices with an elemental basis S = f.
If we imagine the crystal to be made up of isolated atoms like that
shown on the right in Fig. 8 (which is perhaps accurate for an ionic
crystal) then, since the atomic charge density is spherically symmetric
about the atom
f

=
_
d
3
r

(r

)e
iG
hkl
r

=
_
r
2
dr

d(cos )d

(r

)e
G
hkl
r

cos
18
= 4
_
r
2
dr

(r

)
sin G
hkl
r

G
hkl
r

. (40)
As an example, consider a spherical atom of charge Ze

, radius R, and
charge density

(r

) =
3Z
4R
3
(R r

) (41)
then
f

=
3Z
R
3
_
R
0
r
2
dr

sin G
hkl
r

G
hkl
r

=
Z
(G
hkl
R)
3
(sin (G
hkl
R) (G
hkl
R) cos (G
hkl
R)) (42)
This has zeroes whenever tan (G
hkl
R) = G
hkl
R and a maximum when
G
hkl
= 0, or in terms of the scattering angle 2k
0
sin = G
hkl
, when
= 0, . In fact, we have that f

( = 0) = Z. This is true in general,


since
f

( = 0) = f

(G
hkl
= 0) =
_
d
3
r

(r

) = Z . (43)
Thus, for x-ray scattering I Z
2
. For this reason, it is often dicult
to see small-Z atoms with x-ray scattering.
2.3.1 The Structure Factor of Centered Lattices
Now lets look at the structure factor. An especially interesting situa-
tion occurs for centered lattices. We can consider a BCC lattice as a
cubic unit cell |a
1
| = |a
2
| = |a
3
|, a
1
a
2
a
3
and a two-atom basis
r

= 0.5(a
1
+ a
2
+ a
3
), = 0, 1. Now if both sites in the unit cell
19
( = 0, 1) contain the same atom with the same scattering factor f,
then
r

G
hkl
= 0.5(a
1
+a
2
+a
3
) (hg
1
+kg
2
+lg
3
) = (h +k +l) (44)
so that
S
hkl
=

=0,1
fe
i(h+k+l)
= f(1 + e
i(h+k+l)
) =
_

_
0 if h + k + l is odd
2f if h + k + l is even
(45)
This lattice gives rise to extinctions (lines, which appear in a cubic lat-
tice, but which are missing here)! If both atoms of the basis are identi-
cal (like bcc iron), then the bcc structure leads to extinctions; however,
consider CsCl. It does not have these extinctions since f
Cs
+ = f
Cl
. In
fact, to a good approximation f
Cs
+ f
Xe
and f
Cl
f
Ar
. However CsI
(also a bcc structure) comes pretty close to having complete extinctions
since both the Cs and I ions take on the Xe electronic shell. Thus, in
the scattering pattern of CsI, the odd h+k +l peaks are much smaller
than the even ones. Other centered lattices also lead to extinctions.
In fact, this leads us to a rather general conclusion. The shape
and dimensions of the unit cell determines the location of the Bragg
peaks; however, the content of the unit cell helps determine the relative
intensities of the peaks.
20
Unit Cell of BCC ordered FeCo
Fe
Co
Figure 9: The unit cell of body-centered cubic ordered FeCo.
Extinctions in Binary Alloys Another, and signicantly more interest-
ing, example of extinctions in scattering experiments happens in binary
alloys such as FeCo on a centered BCC lattice. Since Fe and Co are
adjacent to each other on the periodic chart, and the x-ray form factor
is proportional to Z (Z
Co
= 27, and Z
Fe
= 26)
f
xray
Fe
f
xray
Co
. (46)
However, since one has a closed nuclear shell and the other doesnt,
their neutron scattering factors will be quite dierent
f
neutron
Fe
= f
neutron
Co
(47)
Thus, neutron scattering from the ordered FeCo structure shown in
Fig. 9 will not have extinctions; whereas scattering from a disordered
structure (where the distribution of Fe and Co is random, so each site
has a 50% chance of having Fe or Co, independent of the occupation
of the neighboring sites) will have extinctions!
21
2.3.2 Powdered x-ray Diraction
If you expose a columnated beam of x-rays to a crystal with a single
crystalline domain, you usually will not achieve a diraction spot. The
reason why can be seen from the Ewald construction, shown in Fig. 10
For any given k
0
, the chances of matching up so as to achieve G = kk
0
are remote. For this reason most people use powdered x-ray diraction
to characterize their samples. This is done by making a powder with
randomly distributed crystallites. Exposing the powdered sample to
x-rays and recording the pattern. The powdered sample corresponds
to averaging over all orientations of the reciprocal lattice. Thus one
will observe all peaks that lie within a radius of 2|k
0
| of the origin of
the reciprocal lattice.
22
G
k
k
0
hkl
O
Ewald Sphere
Figure 10: The Ewald Construction to determine if the conditions are correct for
obtaining a Bragg peak: Select a point in k-space as the origin. Draw the incident
wave vector k
0
to the origin. From the base of k
0
, spin k (remember, that for elastic
scattering |k| = |k
0
|) in all possible directions to form a sphere. At each point where
this sphere intersects a lattice point in k-space, there will be a Bragg peak with G =
k k
0
. In the example above we nd 8 Bragg peaks. If however, we change k
0
by a
small amount, then we have none!
23
Chapter 4: Crystal Lattice Dynamics
Debye
December 29, 2001
Contents
1 An Adiabatic Theory of Lattice Vibrations 2
1.1 The Equation of Motion . . . . . . . . . . . . . . . . . . 6
1.2 Example, a Linear Chain . . . . . . . . . . . . . . . . . . 8
1.3 The Constraints of Symmetry . . . . . . . . . . . . . . . 11
1.3.1 Symmetry of the Dispersion . . . . . . . . . . . . 12
1.3.2 Symmetry and the Need for Acoustic modes . . . 15
2 The Counting of Modes 18
2.1 Periodicity and the Quantization of States . . . . . . . . 19
2.2 Translational Invariance: First Brillouin Zone . . . . . . 19
2.3 Point Group Symmetry and Density of States . . . . . . 21
3 Normal Modes and Quantization 21
3.1 Quantization and Second Quantization . . . . . . . . . . 24
1
4 Theory of Neutron Scattering 26
4.1 Classical Theory of Neutron Scattering . . . . . . . . . . 27
4.2 Quantum Theory of Neutron Scattering . . . . . . . . . . 30
4.2.1 The Debye-Waller Factor . . . . . . . . . . . . . . 35
4.2.2 Zero-phonon Elastic Scattering . . . . . . . . . . 36
4.2.3 One-Phonon Inelastic Scattering . . . . . . . . . . 37
2
A crystal lattice is special due to its long range order. As you ex-
plored in the homework, this yields a sharp diraction pattern, espe-
cially in 3-d. However, lattice vibrations are important. Among other
things, they contribute to
the thermal conductivity of insulators is due to dispersive lattice
vibrations, and it can be quite large (in fact, diamond has a ther-
mal conductivity which is about 6 times that of metallic copper).
in scattering they reduce of the spot intensities, and also allow
for inelastic scattering where the energy of the scatterer (i.e. a
neutron) changes due to the absorption or creation of a phonon in
the target.
electron-phonon interactions renormalize the properties of elec-
trons (make them heavier).
superconductivity (conventional) comes from multiple electron-
phonon scattering between time-reversed electrons.
1 An Adiabatic Theory of Lattice Vibrations
At rst glance, a theory of lattice vibrations would appear impossibly
daunting. We have N 10
23
ions interacting strongly (with energies of
about (e
2
/A)) with N electrons. However, there is a natural expansion
parameter for this problem, which is the ratio of the electronic to the
3
ionic mass:
m
M
1 (1)
which allows us to derive an accurate theory.
Due to Newtons third law, the forces on the ions and electrons are
comparable F e
2
/a
2
, where a is the lattice constant. If we imagine
that, at least for small excursions, the forces binding the electrons and
the ions to the lattice may be modeled as harmonic oscillators, then
F e
2
/a
2
m
2
electron
a M
2
ion
a (2)
This means that

ion

electron

_
m
M
_
1/2
10
3
to 10
2
(3)
Which means that the ion is essentially stationary during the period
of the electronic motion. For this reason we may make an adiabatic
approximation:
we treat the ions as stationary at locations R
1
, R
N
and deter-
mine the electronic ground state energy, E(R
1
, R
N
). This may
be done using standard ab-initio band structure techniques such
as those used by FJP.
we then use this as a potential for the ions; i.e.. we recalculate
E as a function of the ionic locations, always assuming that the
electrons remain in their ground state.
4
nth unit cell
O
r
n
atom

r
n
r

s
R = + +
n
r
n
r

s
Figure 1: Nomenclature for the lattice vibration problem. s
n,
is the displacement of
the atom within the n-th unit cell from its equilibrium position, given by r
n,
=
r
n
+r

, where as usual, r
n
= n
1
a
1
+n
2
a
2
+n
3
a
3
.
Thus the potential energy for the ions
(R
1
, R
N
) = E(R
1
, R
N
) + the ion-ion interaction (4)
We will dene the zero potential such that when all R
n
are at their
equilibrium positions, = 0. Then
H =

n
P
2
n
2M
+ (R
1
, R
N
) (5)
Typical lattice vibrations involve small atomic excursions of the or-
der 0.1Aor smaller, thus we may expand about the equilibrium position
of the ions.
(r
ni
+ s
ni
) = (r
ni
) +

r
ni
s
ni
+
1
2

r
ni
r
mj
s
ni
s
mj
(6)
The rst two terms in the sum are zero; the rst by denition, and
the second is zero since it is the rst derivative of a potential being
5
evaluated at the equilibrium position. We will dene the matrix

mj
ni
=

2

r
ni
r
mj
(7)
From the dierent conservation laws (related to symmetries) of the
system one may derive some simple relationships for . We will discuss
these in detail later. However, one must be introduced now, that is,

r

s
Figure 2: Since the coecients of potential between the atoms linked by the blue lines
(or the red lines) must be identical,
mj
ni
=
(mn)j
0i
.
due to translational invariance.

mj
ni
=
(mn)j
0i
=

2

r
0i
r
(nm)j
(8)
ie, it can only depend upon the distance. This is important for the next
subsection.
6
1.1 The Equation of Motion
From the derivative of the potential, we can calculate the force on each
site
F
ni
=
(r
mj
+ s
mj
)
s
ni
(9)
so that the equation of motion is

mj
ni
s
mj
= M

s
ni
(10)
If there are N unit cells, each with r atoms, then this gives 3Nr equa-
tions of motion. We will take advantage of the periodicity of the lattice
by using Fourier transforms to achieve a signicant decoupling of these
equations. Imagine that the coordinate s of each site is decomposed
into its Fourier components. Since the equations are linear, we may just
consider one of these components to derive our equations of motion in
Fourier space
s
ni
=
1

u
i
(q)e
i(qr
n
t)
(11)
where the rst two terms on the rhs serve as the polarization vector
for the oscillation, u
i
(q) is independent of n due to the translational
invariance of the system. In a real system the real s would be composed
of a sum over all q and polarizations. With this substitution, the
equations of motion become

2
u
i
(q) =
1
_
M

mj
ni
e
iq(r
m
r
n
)
u
j
(q) sum repeated indices .
(12)
7
t = 0
t = t
t = 2 t
Figure 3: u
i
(q) is independent of n so that a lattice vibration can propagate and
respect the translational invariant of the lattice.
Recall that
mj
ni
=
(mn)j
0i
so that if we identify
D
j
i
=
1
_
M

mj
ni
e
iq(r
m
r
n
)
=
1
_
M

pj
0i
e
iq(r
p
)
(13)
where r
p
= r
m
r
n
, then the equation of motion becomes

2
u
i
(q) = D
j
i
u
j
(q) (14)
or
_
D
j
i

2

j
i
_
u
j
(q) = 0 (15)
which only has nontrivial (u ,= 0) solutions if det
_
D(q)
2
I
_
= 0.
For each q there are 3r dierent solutions (branches) with eigenvalues

(n)
(q) (or rather
(n)
(q) are the root of the eigenvalues). The de-
pendence of these eigenvalues
(n)
(q) on q is known as the dispersion
relation.
8
basis
a
= 1 = 2
M
1
M
2
f
Figure 4: A linear chain of oscillators composed of a two-element basis with dierent
masses, M
1
and M
2
and equal strength springs with spring constant f.
1.2 Example, a Linear Chain
Consider a linear chain of oscillators composed of a two-element ba-
sis with dierent masses, M
1
and M
2
and equal strength springs with
spring constant f. It has the potential energy
=
1
2
f

n
(s
n,1
s
n,2
)
2
+ (s
n,2
s
n+1,1
)
2
. (16)
We may suppress the indices i and j, and search for a solution
s
n
=
1

(q)e
i(qr
n
t)
(17)
to the equation of motion

2
u

(q) = D

(q) where D

=
1
_
M

p,
0
e
iq(r
p
)
(18)
and,

m,
n,
=

2

r
0,
r
(nm),
(19)
where nontrivial solutions are found by solving det
_
D(q)
2
I
_
= 0.
The potential matrix has the form

n,1
n,1
=
n,2
n,2
= 2f (20)
9

n,2
n,1
=
n,1
n,2
=
n1,2
n,1
=
n+1,1
n,2
= f . (21)
This may be Fourier transformed on the space index n by inspection,
so that
D

=
1
_
M

p
0
e
iq(r
p
)
=
_
_
_
_
2f
M
1

M
1
M
2
_
1 + e
iqa
_

M
1
M
2
_
1 + e
+iqa
_
2f
M
2
_
_
_
_
(22)
Note that the matrix D is hermitian, as it must be to yield real, physi-
cal, eigenvalues
2
(however, can still be imaginary if
2
is negative,
indicating an unstable mode). The secular equation det
_
D(q)
2
I
_
=
0 becomes

2
2f
_
1
M
1
+
1
M
2
_
+
4f
2
M
1
M
2
sin
2
(qa/2) = 0 , (23)
with solutions

2
= f
_
1
M
1
+
1
M
2
_
f

_
_
1
M
1
+
1
M
2
_
2

4
M
1
M
2
sin
2
(qa/2) (24)
This equation simplies signicantly in the q 0 and q/a limits.
In units where a = 1, and where the reduced mass 1/ =
_
1
M
1
+
1
M
2
_
,
lim
q0

(q) = qa

_
f
2M
1
M
2
lim
q0

+
(q) =

_
2f

(25)
and

(q = /a) =
_
2f/M
2
.
+
(q = /a) =
_
2f/M
1
(26)
10
-4 -2 0 2 4
q
0.0
0.5
1.0
1.5
2.0

-
optical mode
acoustic mode
~ ck
Figure 5: Dispersion of the linear chain of oscillators shown in Fig. 4 when M
1
= 1,
M
2
= 2 and f = 1. The upper branch
+
is called the optical and the lower branch is
the acoustic mode.
As a result, the + mode is quite at; whereas the mode varies from
zero at the Brillouin zone center q = 0 to a at value at the edge of the
zone. This behavior is plotted in Fig. 5.
It is also instructive to look at the eigenvectors, since they will tell
us how the atoms vibrate. Lets look at the optical mode at q = 0,

+
(0) =
_
2f/. Here,
D =
_
_
_
_
2f/M
1
2f/

M
1
M
2
2f/

M
1
M
2
2f/M
2
_
_
_
_
. (27)
11
Eigenvectors are non-trivial solutions to (
2
I D)u = 0, or
0 =
_
_
_
_
2f/ 2f/M
1
2f/

M
1
M
2
2f/

M
1
M
2
2f/ 2f/M
2
_
_
_
_
_
_
_
_
u
1
u
2
_
_
_
_
. (28)
with the solution u
1
=
_
M
2
/M
1
u
2
. In terms of the actual displace-
ments Eqs.11
s
n1
s
n2
=

_
M
2
M
1
u
1
u
2
(29)
or s
n1
/s
n2
= M
2
/M
1
so that the two atoms in the basis are moving
out of phase with amplitudes of motion inversely proportional to their
masses. These modes are described as optical modes since these atoms,
Figure 6: Optical Mode (bottom) of the linear chain (top).
if oppositely charged, would form an oscillating dipole which would
couple to optical elds with a. Not all optical modes are optically
active.
1.3 The Constraints of Symmetry
We know a great deal about the dispersion of the lattice vibrations
without solving explicitly for them. For example, we know that for each
q, there will be dr modes (where d is the lattice dimension, and r is the
12
number of atoms in the basis). We also expect (and implicitly assumed
above) that the allowed frequencies are real and positive. However,
from simple mathematical identities, the point-group and translational
symmetries of the lattice, and its time-reversal invariance, we can learn
more about the dispersion without solving any particular problem.
The basic symmetries that we will employ are
The translational invariance of the lattice and reciprocal lattice.
The point group symmetries of the lattice and reciprocal lattice.
Time-reversal invariance.
1.3.1 Symmetry of the Dispersion
Complex Properties of the dispersion and Eigenmodes First, from the
symmetry of the second derivative, one may show that
2
is real. Recall
that the dispersion is determined by the secular equation det
_
D(q)
2
I
_
=
0, so if D is hermitian, then its eigenvalues,
2
, must be real.
D
j
i
=
1
_
M

pj
0i
e
iq(r
p
)
(30)
=
1
_
M

p,,j
0,,i
e
iq(r
p
)
(31)
Then, due to the symmetric properties of the second derivative
D
j
i
=
1
_
M

0,,i
p,,j
e
iq(r
p
)
=
1
_
M

p,,i
0,,j
e
iq(r
p
)
= D
i
j
(32)
13
Thus, D
T
= D

= D so D is hermitian and its eigenvalues


2
are
real. This means that either are real or they are pure imaginary. We
will assume the former. The latter yields pure exponential growth of
our Fourier solution, indicating an instability of the lattice to a second-
order structural phase transition.
Time-reversal invariance allows us to show related results. We as-
sume a solution of the form
s
ni
=
1

u
i
(q)e
i(qr
n
t)
(33)
which is a plane wave. Suppose that the plane wave is moving to the
right so that q = xq
x
, then the plane of stationary phase travels to the
right with
x =

q
x
t . (34)
Clearly then changing the sign of q
x
is equivalent to taking t t.
If the system is to display proper time-reversal invariance, so that the
plane wave retraces its path under time-reversal, it must have the same
frequency when time, and hence q, is reversed, so
(q) = (q) . (35)
Note that this is fully equivalent to the statement that D
i
j
(q) =
D
i
j
(q) which is clear from the denition of D.
Now, return to the secular equation, Eq. 15.
_
D
j
i
(q)
2
(q)
j
i
_

j
(q) = 0 (36)
14
Lets call the (normalized) eigenvectors of this equation . They are the
elements of a unitary matrix which diagonalizes D. As a result, they
have orthogonality and completeness relations

,i

(n)
,i
(q)
(m)
,i
(q) =
m,n
orthogonality (37)

(n)
,i
(q)
(n)
,j
(q) =
,

i,j
(38)
If we now take the complex conjugate of the secular equation
_
D
j
i
(q)
2
(q)
j
i
_

j
(q) = 0 (39)
Then it must be that

j
(q)
j
(q) . (40)
Since the are normalized the constant of proportionality may be
chosen as one

j
(q) =
j
(q) . (41)
Point-Group Symmetry and the Dispersion A point group operation
takes a crystal back to an identical conguration. Both the original
and nal lattice must have the same dispersion. Thus, since the recip-
rocal lattice has the same point group as the real lattice, the dispersion
relations have the same point group symmetry as the lattice.
For example, the dispersion must share the periodicity of the Bril-
louin zone. From the denition of D
D
j
i
(q) =
1
_
M

pj
0i
e
iq(r
p
)
(42)
15
it is easy to see that D
j
i
(q +G) = D
j
i
(q) (since G r
p
= 2n, where
n is an integer). I.e., D is periodic in k-space, and so its eigenvalues
(and eigenvectors) must also be periodic.

(n)
(k +G) =
(n)
(k) (43)

j
(k +G) =
j
(k) . (44)
1.3.2 Symmetry and the Need for Acoustic modes
Applying basic symmetries, we can show that an elemental lattice (that
with r = 1) must have an acoustic model. First, look at the transla-
s
11
Figure 7: If each ion is shifted by s
1,1,i
, then the lattice energy is unchanged.
tional invariance of . Suppose we make an overall shift of the lattice
by an arbitrary displacement s
n,,i
for all sites n and elements of the
basis (i.e. s
n,,i
= s
1,1,i
). Then, since the interaction is only between
16
ions, the energy of the system should remain unchanged.
E =
1
2

m,n,,,i,j

mn,,j
0,,i
s
n,,i
s
m,,j
= 0 (45)
=
1
2

mn,,i,j

mn,,j
0,,i
s
1,1,i
s
1,1,j
(46)
=
1
2

i,j
s
1,1,i
s
1,1,j

mn,

mn,,j
0,,i
(47)
Since we know that s
1,1,i
is nite, it must be that

m,n,,

mn,,j
0,,i
=

p,,

p,,j
0,,i
= 0 (48)
Now consider a strain on the system V
m,,j
, described by the strain
matrix m
,i
,j
V
m,,j
=

,i
m
,i
,j
s
m,,i
(49)
After the stress has been applied, the atoms in the bulk of the sample
V
Figure 8: After a stress is applied to a lattice, the movement of each ion (strain) is
not only in the direction of the applied stress. The response of the lattice to an applied
stress is described by the strain matrix.
are again in equilibrium (those on the surface are maintained in equi-
librium by the stress), and so the net force must be zero. Looking at
17
the central (n = 0) atom this means that
0 = F
0,,i
=

m,,j,,k

m,,j
0,,i
m
,k
,j
s
m,,k
(50)
Since this applies for an arbitrary strain matrix m
,k
,j
, the coecients
for each m
,k
,j
must be zero

m,,j
0,,i
s
m,,k
= 0 (51)
An alternative way (cf. Callaway) to show this is to recall that the
reection symmetry of the lattice requires that
m,,j
0,,i
be even in m;
whereas, s
m,,k
is odd in m. Thus the sum over all m yields zero.
Now lets apply these constraints to D for an elemental lattice where
r = 1, and we may suppress the basis indices .
D
j
i
(q) =
1
M

p,j
0,i
e
iq(r
p
)
(52)
For small q we may expand D
D
j
i
(q) =
1
M

p,j
0,i
_
1 + iq (r
p
)
1
2
(q (r
p
))
2
+
_
(53)
We have shown above that the rst two terms in this series are zero.
Thus,
D
j
i
(q)
1
2M

p,j
0,i
(iq (r
p
))
2
(54)
Thus, the leading order (small q) eigenvalues
2
(q) q
2
. I.e. they are
acoustic modes. We have shown that all elemental lattices must have
acoustic modes for small q.
18
In fact, one may show that all harmonic lattices in which the energy
is invariant under a rigid translation of the entire lattice must have at
least one acoustic mode. We will not prove this, but rather make a
simple argument. The rigid translation of the lattice corresponds to a
q = 0 translational mode, since no energy is gained by this translation,
it must be that
s
(q = 0) = 0 for the branch s which contains this
mode. The acoustic mode may be obtained by perturbing (in q) around
this point. Physically this mode corresponds to all of the elements of
the basis moving together so as to emulate the motion in the elemental
basis.
2 The Counting of Modes
In the sections to follow, we need to perform sums (integrals) of func-
tions of the dispersion over the crystal momentum states k within the
reciprocal lattice. However, the translational and point group symme-
tries of the crysal, often greatly reduce the set of points we must sum. In
addition, we often approximate very large systems with hypertoroidal
models with periodic boundary conditions. This latter approximation
becomes valid as the system size diverges so that the surface becomes
of zero measure.
19
2.1 Periodicity and the Quantization of States
A consequence of approximating our system as a nite-sized periodic
system is that we now have a discrete sum rather than an integral over
k. Consider a one-dimensional nite system with N atoms and periodic
boundary conditions. We seek solutions to the phonon problem of the
type
s
n
= (q)e
i(qr
n
t)
where r
n
= na (55)
and we require that
s
n+N
= s
n
(56)
or
q(n + N)a = qna + 2m where m is an integer (57)
Then, the allowed values of q = 2m/Na. This will allow us to convert
the integrals over the Brillouin zone to discrete sums, at least for cubic
systems; however, the method is easily generalized for other Bravais
lattices.
2.2 Translational Invariance: First Brillouin Zone
We can use the translational invariance of the crystal to reduce the
complexity of sums or integrals of functions of the dispersion over the
crystal momentum states. As shown above, translationally invariant
systems have states which are not independent. It is useful then to de-
ne a region of k-space which contains only independent states. Sums
20
G vector
Bisector
First Brillouin Zone
Figure 9: The First Brillouin Zone. The end points of all vector pairs that satisfy
the Bragg condition k k
0
= G
hkl
lie on the perpendicular bisector of G
hkl
. The
smallest polyhedron centered at the origin of the reciprocal lattice and enclosed by
perpendicular bisectors of the Gs is called the rst Brillouin zone.
over k may then be conned to this region. This region is dened as
the smallest polyhedron centered at the origin of the reciprocal lattice
and enclosed by perpendicular bisectors of the Gs is called the Bril-
louin zone (cf. Fig. 9). Typically, we choose to include only half of the
bounding surface within the rst Brillouin zone, so that it can also be
dened as the set of points which contains only independent states.
From the discussion in chapter 3 and in this chapter, it is also clear
that the reciprocal lattice vectors have some interpretation as momen-
tum. For example, the Laue condition requires that the change in
momentum of the scatterer be equal to a reciprocal lattice translation
vector. The end points of all vector pairs that satisfy the Bragg condi-
tion kk
0
= G
hkl
lie on the perpendicular bisector of G
hkl
. Thus, the
FBZ is also the set of points which cannot satisfy the Bragg condition.
21
2.3 Point Group Symmetry and Density of States
Two other tricks to reduce the complexity of these sums are worth
mentioning here although they are discussed in detail elswhere.
The rst is the use of the point group symmetry of the system. It
is clear from their denition in chapter 3, the reciprocal lattice vectors
have the same point group symmetry as the lattice. As we discussed
in chapter 2, the knowledge of the group elements and corresponding
degeneracies may be used to reduce the sums over k to the irreducible
wedge within the the First Brillioun zone. For example, for a cubic
system, this wedge is only 1/2
3
3! or 1/48th of the the FBZ!
The second is to introduce a phonon density of states to reduce the
multidimensional sum over k to a one-dimensional integral over energy.
This will be discussed in chapter 5.
3 Normal Modes and Quantization
In this section we will derive the equations of motion for the lattice, de-
termine the canonically conjugate variables (the the sense of Lagrangian
mechanics), and use this information to both rst and second quantize
the system.
Any lattice displacement may be expressed as a sum over the eigen-
22
vectors of the dynamical matrix D.
s
n,,i
=
1

q,s
Q
s
(q, t)
s
,i
(q)e
iqr
n
(58)
Recall that
s
,i
(q) are distinguished from u
s
,i
(q) only in that they are
normalized. Also since q+G is equivalent to q, we need sum only over
the rst Brillouin zone. Finally we will assume that Q
s
(q, t) contains
the harmonic time dependence and since s
n,,i
is real Q

s
(q) = Q
s
(q).
We may rewrite both the kinetic and potential energy of the system
as sums over Q. For example, the kinetic energy of the lattice
T =
1
2

n,,i
M

( s
n,i
)
2
(59)
=
1
2N

n,,i

q,k,r,s

Q
r
(q)
r
,i
(q)e
iqr
n
Q
s
(k)
s
,i
(k)e
ikr
n
(60)
Then as
1
N

n
e
i(k+q)r
n
=
k,q
and

,i

r
,i

s
,i
=
rs
(61)
the kinetic energy may be reduced to
T =
1
2

q,r


Q
r
(q)

2
(62)
The potential energy may be rewritten in a similar fashion
V =
1
2

n,m,,,i,j

m,,j
n,,i
s
n,,i
s
m,,j
=
1
2

n,m,,,i,j

mn,,j
0,,i
N
_
M

q,k,s,r
Q
s
(q, t)
s
,i
(q)e
iqr
n
Q
r
(k, t)
r
,j
(k)e
ikr
m
(63)
23
Let r
l
= r
m
r
n
V =
1
2

n,l,,,i,j

l,,j
0,,i
N
_
M

q,k,s,r
Q
s
(q, t)
s
,i
(q)e
iqr
n
Q
r
(k, t)
r
,j
(k)e
ik(r
l
+r
n
)
(64)
and sum over n to obtain the delta function
k,q
so that
V =
1
2

l,,,i,j,s,r
Q
s
(k)
s
,i
(k)Q
r
(k)
r
,j
(k)
1
_
M

l,,j
0,,i
e
ikr
l
. (65)
Note that the sum over l on the last three terms yields D, so that
V =
1
2

l,,,i,j,s,r
D
,j
,i
(k)Q
s
(k)
s
,i
(k)Q
r
(k)
r
,j
(k) . (66)
Then, since

,j
D
j
i
(k)
r
j
(k) =
2
r
(k)
r
,i
(k) and
s
,i
(k) =
s
,i
(k),
V =
1
2

,i,k,r,s

r
,i
(k)
s
,i
(k)
2
r
(k)Q

s
(k)Q
r
(k) (67)
Finally, since

,i

r
,i
(k)
s
,i
(k) =
r,s
V =
1
2

k,s

2
s
(k) [Q
s
(k)[
2
(68)
Thus we may write the Lagrangian of the ionic system as
L = T V =
1
2

k,s
_


Q
s
(k)

2
s
(k) [Q
s
(k)[
2
_
, (69)
where the Q
s
(k) may be regarded as canonical coordinates, and
P

r
(k) =
L
Q
r
(k)
=

Q

s
(k) (70)
(no factor of 1/2 since Q

s
(k) = Q
s
(k)) are the canonically conjugate
momenta.
24
The equations of motion are
d
dt
_
_
L

s
(k)
_
_

L
Q

s
(k)
or

Q
s
(k) +
2
s
(k)Q
s
(k) = 0 (71)
for each k, s. These are the equations of motion for 3rN independent
harmonic oscillators. Since going to the Q-coordinates accomplishes the
decoupling of these equations, the Q
s
(k) are referred to as normal
coordinates.
3.1 Quantization and Second Quantization
P.A.M. Dirac laid down the rules of quantization, from Classical Hamilton-
Jacobi classical mechanics to Hamiltonian-based quantum mechanics
following the path (Dirac p.84-89):
1. First, identify the classical canonically conjugate set of variables
q
i
, p
i

2. These have Poisson Brackets


u, v =

i
_
u
q
i
v
p
i

u
p
i
v
q
i
_
(72)
q
i
, p
j
=
i,j
p
i
, p
j
= q
i
, q
j
= 0 (73)
3. Then dene the quantum Poisson Bracket (the commutator)
[u, v] = uv vu = i hu, v (74)
4. In particular, [q
i
, p
j
] = i h
i,j
, and [q
i
, q
j
] = [p
i
, p
j
] = 0.
25
Thus, following Dirac, we may now quantize the normal coordinates
[Q

r
(k), P
s
(q)] = i h
k,q

r,s
where the other commutators vanish .
(75)
Furthermore, since we have a system of 3rN uncoupled harmonic os-
cillators we may immediately second quantize by introducing
a
s
(k) =
1

2 h
_
_
_

s
(k)Q
s
(k) +
i
_

s
(k)
P
s
(k)
_
_
(76)
a

s
(k) =
1

2 h
_
_
_

s
(k)Q

s
(k)
i
_

s
(k)
P

s
(k)
_
_
, (77)
or
Q
s
(k) =

_
h
2
s
(k)
_
a
s
(k) + a

s
(k)
_
(78)
P
s
(k) = i

_
h
s
(k)
2
_
a
s
(k) a

s
(k)
_
(79)
Where
[a
s
(k), a

r
(q)] =
r,s

q,k
[a
s
(k), a
r
(q)] = [a

s
(k), a

r
(q)] = 0 (80)
This transformation Q, P a, a

is canonical, since is preserves


the commutator algebra Eq. 75, and the Hamiltonian becomes
H =

k,s
h
s
(k)
_
a

s
(k)a
s
(k) +
1
2
_
(81)
which is a sum over 3rN independent quantum oscillators, each one
referred to as a phonon mode!
The number of phonons in state (k, s) is given by the operator
n
s
(k) = a

s
(k)a
s
(k) (82)
26
and a

s
(k) and a
s
(k) create and destroy phonons respectively, in the
state (k, s)
a

s
(k) [n
s
(k)) =
_
n
s
(k) + 1 [n
s
(k) + 1) (83)
a
s
(k) [n
s
(k)) =
_
n
s
(k) [n
s
(k) 1) (84)
If [0) is the normalized state with no phonons present, then the state
with n
s
(k) phonons in each state (k, s) is
[n
s
(k)) =
_

_
_
_

k,s
1
n
s
(k)!
_
_
1
2
_

k,s
_
a

s
(k)
_
n
s
(k)
[0) (85)
Finally the lattice point displacement
s
n,,i
=
1

q,s

_
h
2
s
(q)
_
a
s
(q) + a

s
(q)
_

s
,i
(q)e
iqr
n
(86)
will be important in the next section, especially with respect to zero-
point motion (i.e.
_
s
2
_
T=0
,= 0).
4 Theory of Neutron Scattering
To see the lattice with neutrons, we want their De Broglie wavelength
= h/p

neutron
=
0.29A

E
E measured in eV (87)
to be of the same length as the intersite distance on the lattice. This
means that their kinetic energy E
1
2
Mv
2
0.1eV, or E/k
b
1000K;
i.e. thermal neutrons.
27
| | or | |
a
2
a
1
a
1
a
2
n
Source of thermal
neutrons
Figure 10: Neutron Scattering. The De Broglie wavelength of the neutrons must
be roughly the same size as the lattice constants in order to learn about the lattice
structure and its vibrational modes from the experiment. This dictates the use of
thermal neutrons.
Since the neutron is chargeless, it only interacts with the atomic
nucleus through a short-ranged nuclear interaction (Ignoring any spin-
spin interaction). The range of this interaction is 1 Fermi (10
13
cm.)
or about the radius of the atomic nucleus. Thus
A range of the interaction 10
13
cm. (88)
Thus the neutron cannot see the detailed structure of the nucleus,
and so we may approximate the neutron-ion interaction potential as a
contact interaction
V (r) =

r
n
V
n
(r r
n
) (89)
i.e., we may ignore the angular dependence of the scattering factor f.
4.1 Classical Theory of Neutron Scattering
Due to the importance of lattice vibrations, which are inherently quan-
tum in nature, there is a limit to what we can learn from a classical
28
theory of diraction. Nevertheless it is useful to compare the classical
result to what we will develop for the quantum problems.
For the classical problem we will assume that the lattice is elemental
(r = 1) and start with a generalization of the formalism developed in
the last chapter
I [(K, )[
2
(90)
where K = k
0
k and =
0
. Furthermore, we take
(r(t))

n
(r r
n
(t)) (91)
where
r
n
(t) = r
n
+s
n
(t) and s
n
(t) =
1

M
u(q)e
i(qr
n
(q)t)
(92)
describes the harmonic motion of the s-mode with wave-vector q.
(K, )

n
_
dte
i[K(r
n
+s
n
(t))t]
. (93)
For [K[ 2/A and s
n
(t) A we may expand
(K, )

n
_
dte
i[K(r
n
)t]
(1 + iK s
n
(t) + ) (94)
The rst term yields a nite contribution only when
K = k
0
k = G and =
0
= 0 (95)
which are the familiar Bragg conditions for elastic scattering.
The second term, however, yields something new. It only yields a
nite result when
Kq = k
0
k q = G and
s
(q) =
0

s
(q) = 0 (96)
29
When multiplied by h , these can be interpreted as conditions for
the conservation of (crystal) momentum and energy when the scat-
tering event involves the creation (destruction) of a lattice excitation
(phonon). These processes are called Stokes and antistokes processes,
respectively, and are illustrated in Fig. 11.
k ,
0

0

q
q ,
k = - q , =
n
k ,
0

0

q
q ,
n
Stokes Process
(phonon creation)
Anti-Stokes Process
(phonon absorbtion)

q
-
0
k
0
k = + q , =
q
+
0
k
0
Figure 11: Stokes and antistokes processes in inelastic neutron scattering involving
the creation or absorption of a lattice phonon.
Clearly, the anti-Stokes process can only happen at nite temper-
atures where real (as opposed to virtual) phonons are excited. Thus,
our classical formalism does not correctly describe the temperature de-
pendence of the scattering. Several other things are missing, including:
Security in the validity of the result.
The eects of zero-point motion.
Correct temperature dependence.
30
4.2 Quantum Theory of Neutron Scattering
To address these concerns, we will do a fully quantum calculation.
Several useful references for this calculation include
Ashcroft and Mermin, Appendix N, p. 790)
Callaway, p. 36.
Hook and Hall (for experiment) Ch. 12 p.342-
We will imagine that the scattering shown in Fig. 12 occurs in a box
of volume V . The momentum transfer, from the neutron to the lattice
k
0

0
n
V
1

e
i( r -
0
t )
k
0

n
Initial
k
k
f
V
1

e
i( r -
f
t )

=
0 =
f
0
E
0

f
E
f
Final
Figure 12: The initial (left) and nal (right) states of the neutron and lattice during
a scattering event. The initial system state is given by
0
=
0

0
, with energy

0
= E
0
+ h
0
where
0
= k
2
0
/2M. The nal system state is given by
f
=
f

f
,
with energy
f
= E
f
+ h
f
where
f
= k
2
f
/2M.
is K = k
0
k
f
and the energy transfer which is nite for inelastic
scattering is h = h(
0

f
). Again we will take the neutron-lattice
31
interaction to be local
V (r) =

r
n
V (r r
n
) =
1
N

q,n
V (q)e
iq(rr
n
)
=
_
d
3
q
V
V
0

n
e
iq(rr
n
)
(97)
where the locality of the interaction (V (r r
n
) (r r
n
)) indicates
that V (q) = V (0) = V
0
. Consistent with Aschcroft and Mermin, we
will take
V (r) =
2 h
2
a
M
1
V

r
n
_
d
3
qe
iq(rr
n
)
(98)
where a is the scattering length, and V
0
=
2 h
2
a
M
is chosen such that the
total cross section = 4a
2
.
To formulate our quantum theory, we will use Fermis golden rule
for time dependent perturbation theory. (This is fully equivalent to the
lowest-order Born approximation). The probability per unit time for a
neutron to scatter from state k
0
to k
f
is given by
P =
2
h

f
(
0

f
) [
0
[V [
f
)[
2
(99)
=
2
h

f
(E
0
+ h
0
E
f
h
f
)

1
V
_
d
3
re
i(kk
0
)r

0
[V (r)[
f
)

2
(100)
If we now substitute in the ion-neutron potential Eq. 98, then the inte-
gral over r will yield a delta function V (q +k k
0
) which allows the
q integral to be evaluated
P = a
2
(2 h)
3
(MV )
2

f
(E
0
E
f
+ h)

r
n
_

e
iKr
n


f
_

2
(101)
32
Now, before proceeding to a calculation of the dierential cross sec-
tion
d
ddE
we must be able to convert this probability (rate) for eigen-
states into a ux of neutrons of energy E and momentum p. A dier-
ential volume element of momentum space d
3
p contains V d
3
p/(2 h)
3
neutron states. While this is a natural consequence of the uncertainty
principle, it is useful to show this in a more quantitative sense: Imagine
a cubic volume V = L
3
with periodic boundary conditions so that for
any state in V ,
(x + L, y, z) = (x, y, z) (102)
If we write (r) =
1
N

q
e
iqr
(q), then it must be that
q
x
L = 2m where m is an integer (103)
with similar relations for the y and z components. So for each volume
element of q-space
_
2
L
_
3
there is one such state. In terms of states p =
hq, the volume of a state is (2 h/L)
3
. Thus d
3
p contains V d
3
p/(2 h)
3
states.
The incident neutron ux of states (velocity times density) is
j =
hk
0
M
[
0
[
2
=
hk
0
M

V
e
ik
0
r

2
=
hk
0
MV
(104)
Then since the number of neutrons is conserved
j
d
dEd
dEd =
hk
0
MV
d
dEd
dEd = PV
d
3
p
(2 h)
3
= PV
p
2
dpd
(2 h)
3
(105)
33
And for thermal (non-relativistic) neutrons E = p
2
/2M, so dE =
pdp/M, and
hk
0
MV
d
dEd
dEd = PV
hkMdEd
(2 h)
3
(106)
or
d
dEd
= P
k
k
0
(MV )
2
(2 h)
3
. (107)
Substituting in the previous result for P
d
dEd
=
k
k
0
(MV )
2
(2 h)
3
a
2
(2 h)
3
(MV )
2

f
(E
0
E
f
+ h)

r
n
_

e
iKr
n


f
_

2
(108)
or
d
dEd
=
k
k
0
Na
2
h
S(K, ) (109)
where
S(K, ) =
1
N

f
(E
0
E
f
+ h)

r
n
_

e
iKr
n


f
_

2
. (110)
We may deal with the Dirac delta function by substituting
() =
_

dt
2
e
it
. (111)
so that
S(K, ) =
1
N

f
_

dt
2
e
i((E
0
E
f
)/ h+)t

r
n
,r
m

e
iKr
n


f
_

e
iKr
m


0
_

. (112)
then as
e
iHt/ h
[
l
) = e
iE
l
t/ h
[
l
) (113)
34
where H is the lattice Hamiltonian, we can write this as
S(K, ) =
1
N

f
_

dt
2
e
it

r
n
,r
m
_

e
iHt/ h
e
iKr
n
e
iHt/ h


f
_
_

e
iKr
m


0
_
, (114)
and the argument in the rst expectation value is the time-dependent
operator e
iKr
n
in the Heisenberg representation
e
iKr
n
(t)
= e
iHt/ h
e
iKr
n
e
iHt/ h
. (115)
Thus,
S(K, ) =
1
N

f
_

dt
2
e
it

r
n
,r
m
_

e
iKr
n
(t)


f
_ _

e
iKr
m


0
_
=
1
N
_

dt
2
e
it

r
n
,r
m
_

e
iKr
n
(t)
e
iKr
m


0
_
. (116)
Now since r
n
(t) = r
n
+s
n
(t) (with r
n
time independent),
S(K, ) =
1
N

n,m
_

dt
2
e
i(K(r
n
r
m
)+t)
_

e
iKs
n
(t)
e
iKs
m


0
_
. (117)
This formula is correct at zero temperature. In order to describe
nite T eects (ie., anti-stokes processes involving phonon absorption)
we must introduce a thermal average over all states

0
[A[
0
) A) =

l
e
E
l

l
[A[
l
) /

l
e
E
l
. (118)
With this substitution,
S(K, ) =
1
N

n,m
_

dt
2
e
i(K(r
n
r
m
)+t)
_
e
iKs
n
(t)
e
iKs
m
_
. (119)
and S(K, ) is called the dynamical structure factor.
35
4.2.1 The Debye-Waller Factor
To simplify this relation further, recall that the exponentiated operators
within the brackets are linear functions of the creation and annihilation
operators a

and a.
s
n,
(t) =
1

q,s

_
h
2
s
(q)

(q)
_
a
s
(q)(t) + a

s
(q)(t)
_
e
iqr
n
(120)
So that, in particular s
n,,i
(t)) = s
n,,i
(0)) = 0. Then let A = iK
s
n,,i
(t) and B = iK s
m,,i
(0) and suppose that the expectation values
of A and B are small. Then
_
e
A
e
B
_
=
_
(1 + A +
1
2
A
2
+ )(1 + B +
1
2
B
2
+ )
_

_
1 + A + B + AB +
1
2
A
2
+
1
2
B
2
+
_
1 +
1
2
_
2AB + A
2
+ B
2
_
+
e
1
2

2AB+A
2
+B
2
)
(121)
This relation is in fact true to all orders, as long as A and B are linear
functions of a

and a . (c.f. Ashcroft and Mermin, p. 792, Callaway pp.


41-48). Thus
_
e
iKs
n
(t)
e
iKs
m
_
= e

1
2

(Ks
n
(t))
2
)
e

1
2

(Ks
m
)
2
)
e
Ks
n
(t)Ks
m

. (122)
Since the Hamiltonian has no time dependence, and the lattice is in-
variant under translations r
n
_
e
iKs
n
(t)
e
iKs
m
_
= e

(Ks
n
)
2
)
e
Ks
nm
(t)Ks
0

, (123)
36
where the rst term is called the Debye-Waller factor e
2W
.
e
2W
= e

(Ks
n
)
2
)
. (124)
Thus letting l = n m
S(K, ) = e
2W

l
_

dt
2
e
i(Kr
l
+t)
e
Ks
l
(t)Ks
0

. (125)
Here the Debye-Waller factor contains much of the crucial quantum
physics. It is nite, even at T = 0 due to zero-point uctuations, and
since K s
n
)
2
will increase with temperature, the total strength of the
Bragg peaks will diminish with increasing T. However, as long as a
crystal has long-ranged order, it will remain nite.
4.2.2 Zero-phonon Elastic Scattering
One may disentangle the elastic and inelastic processes by expanding
the exponential in the equation above.
e
Ks
l
(t)Ks
0

m
1
m!
(K s
l
(t)K s
0
))
m
(126)
If we approximate the exponential by 1, ie. take only the rst, m = 0
term, then
S
0
(K, ) = e
2W

l
_

dt
2
e
i(Kr
l
t)
. (127)
And we recover the lowest order classical result (modied by the Debye-
Waller factor) which gives us the Bragg conditions that S
0
(K, ) is only
nite when K = G and =
0

f
= 0.
S
0
(K, ) = e
2W
()N

K,G
, (128)
37
d
0
dEd
=
k
k
0
Na
2
h
e
2W
()N

K,G
(129)
However, now the scattering intensity is reduced by the Debye-
Waller factor e
2W
, which accounts for zero-point motion and thermal
uctuations.
4.2.3 One-Phonon Inelastic Scattering
When m = 1, then the scattering involves either the absorption or
creation of a phonon. To evaluate
S
1
(K, ) = e
2W

l
_

dt
2
e
i(Kr
l
+t)
K s
l
(t)K s
0
(0)) . (130)
we need
s
n,
(t) =
1

q,s

_
h
2
s
(q)

(q)
_
a
s
(q, t) + a

s
(q, t)
_
e
iqr
n
(131)
in the Heisenberg representation, and therefore we need,
a(q, t) = e
iHt/ h
a(q)e
iHt/ h
= e
i((q)ta

(q)a(q))
a(q)e
i((q)ta

(q)a(q)
= a(q)e
i
(
(q)t(a

(q)a(q)1)
)
e
i((q)ta

(q)a(q)
= a(q)e
i(q)t
(132)
where we have used the fact that (a

a)
n
a = (a

a)
n1
a

aa = (a

a)
n1
(aa

1)a = (a

a)
n1
a(a

a1) = a(a

a1)
n
. Similarly a

(q, t) = a

(q)e
i(q)t
.
Thus,
s
n,
(t) =
1

q,s

he
iqr
n
_
2
s
(q)

(q)
_
a
s
(q)e
i
s
(q)t
+ a

s
(q)e
i
s
(q)t
_
(133)
38
and
s
0,
(0) =
1

p,r

_
h
2
r
(p)

(p)
_
a
r
(p) + a

r
(p)
_
(134)
Recall, we want to evaluate
S
1
(K, ) = e
2W

l
_

dt
2
e
i(Kr
l
+t)
K s
l
(t)K s
0
(0)) . (135)
Clearly, the only terms which survive in K s
l
(t)K s
0
(0)) are those
with r = s and p = q. Furthermore, the sum over l yields a delta
function N

G

K+q,G
. Then as (Gk) = (k) =

(k), and (G
k) = (k),
S
1
(K, ) = e
2W
_

dt
2
e
it

q,G,s
h[K (K)[
2
2
s
(q)M

K+q,G
(136)
_
e
i
s
(q)t
_
a
s
(K)a

s
(K)
_
+ e
i
s
(q)t
_
a

s
(K)a
s
(K)
__
The occupancy of each mode n(q) is given by the Bose factor
n(q)) =
1
e
(q)
1
(137)
So, nally
S
1
(K, ) = e
2W

s
h
2M
s
(K)
[K
s
(K)[
2
(138)
[(1 + n
s
(K))( +
s
(K)) + n
s
(K)( +
s
(K))] .
For the rst term, we get a contribution only when
s
(K) =
0

s
(K) = 0; ie., the nal energy of the neutron is smaller than the
initial energy. The energy is lost in the creation of a phonon. Note that
39
this can happen at any temperature, since (1 + n
s
(K)) ,= 0 at any T.
The second term is only nite when +
s
(K) =
0

f
+
s
(K) = 0;
ie., the nal energy of the neutron is larger than the initial energy. The
additional energy comes from the absorption of a phonon. Thus phonon
absorption is only allowed at nite temperatures, and in fact, the factor
n
s
(K) = 0 at zero temperature. These terms correspond to the Stokes
and anti-Stokes processes, respectively, illustrated in Fig. 13.
k ,
0

0

q
q ,
k = - q , =
n
k ,
0

0

q
q ,
n
Stokes Process
(phonon creation)
Anti-Stokes Process
(phonon absorbtion)

q
-
0
k
0
k = + q , =
q
+
0
k
0
1 + n (K)
s
n (K)
s
Figure 13: Stokes and antistokes processes in inelastic neutron scattering involving
the creation or absorption of a lattice phonon. The antistokes process can only occur
at nite-T, when n
s
(K) = 0.
If we were to continue our expansion of the exponential to larger
values of m, we would nd multiple-phonon scattering processes. How-
ever, these terms are usually of minimal contribution to the total cross
section, due to the fact that the average ionic excursion s is small, and
are usually neglected.
40
Chapter 5: Thermal Properties of Crystal Lattices
Debye
December 22, 2000
Contents
1 Formalism 2
1.1 The Virial Theorem . . . . . . . . . . . . . . . . . . . . . 3
1.2 The Phonon Density of States . . . . . . . . . . . . . . . 5
2 Models of Lattice Dispersion 10
2.1 The Debye Model . . . . . . . . . . . . . . . . . . . . . . 10
2.2 The Einstein Model . . . . . . . . . . . . . . . . . . . . . 12
3 Thermodynamics of Crystal Lattices 13
3.1 Long-Range Order . . . . . . . . . . . . . . . . . . . . . 14
3.2 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . 17
3.3 Thermal Expansion, the Gruneisen Parameter . . . . . . 21
3.4 Thermal Conductivity . . . . . . . . . . . . . . . . . . . 27
1
In the previous chapter, we have shown that the motion of
a harmonic crystal can be described by a set of decoupled har-
monic oscillators.
H =
1
2

k,s
[P
s
(k)[
2
+
2
s
(k) [Q
s
(k)[
2
(1)
At a given temperature T, the occupancy of a given mode is
n
s
(k)) =
1
e

s
(k)
1
(2)
In this chapter, we will apply this information to calculate the
thermodynamic properties of the ionic lattice, in addition to
addressing questions regarding its long-range order in the pres-
ence of lattice vibrations (i.e. do phonons destroy the order).
In order to evaluate the dierent formulas for these quantities,
we will rst discuss two matters of formal convenience.
1 Formalism
To evaluate some of these properties we can use the virial the-
orem and, integrals over the density of states.
2
1.1 The Virial Theorem
Consider the Hamiltonian for a quantum systemH(x, p), where
x and p are the canonically conjugate variables. Then the
expectation value of any function of these canonically conjugate
variables f(x, p) in a stationary state (eigenstate) is constant
in time. Consider
d
dt
x p) =
i
h
[H, x p]) =
i
h
Hx p x pH)
=
iE
h
x p x p) = 0 (3)
where E is the eigenenergy of the stationary state. Let
H =
p
2
2m
+ V (x) , (4)
then
0 =
_
_

_
p
2
2m
+ V (x), x p
_

_
_
=
_
_

_
p
2
2m
, x p
_

_ + [V (x), x p]
_
=
_
1
2m
_
p
2
, x
_
p + x [V (x), p]
_
(5)
Then as
_
p
2
, x
_
= p[p, x] + [p, x] p = 2i hp and p = i h
x
,
0 =
_
i h
m
p
2
+ i hx
x
V (x)
_
(6)
3
or, the Virial theorem:
2 T) = x
x
V (x)) (7)
Now, lets apply this to a harmonic oscillator where V =
1
2
m
2
x
2
and T = p
2
/2m, < H >=< T > + < V >= h(n +
1
2
) and .
We get
T) = V (x)) =
1
2
h(n +
1
2
) (8)
Lets now apply this to nd the RMS excursion of a lattice
site in an elemental lattice r = 1
_
s
2
_
=
1
N

n,i
_
s
2
n,i
_
=
1
N
_

n,i
1
NM

q,s,k,r
Q
s
(q)
s
i
(q)e
iqr
n
Q
s
(k)
r
i
(k)e
ikr
n
_
=
1
N
_

n,i
1
NM

q,s,k,r
Q
s
(q)
s
i
(q)e
iqr
n
Q
s
(k)
r
i
(k)e
ikr
n
_
=
1
N
_

n,i
1
NM

q,s,k,r
Q
s
(q)
s
i
(q)e
iqr
n
Q

s
(k)
r
i
(k)e
ikr
n
_
=
1
NM

q,s
_
[Q
s
(q)[
2
_
=
1
NM

q,s
2

2
s
(q)
_
1
2

2
s
(q) [Q
s
(q)[
2
_
=
1
NM

q,s
2

2
s
(q)
1
2
h
s
(q)
_
_
n
s
(q) +
1
2
_
_
4
_
s
2
_
=
1
NM

q,s
h

s
(q)
_
_
n
s
(q) +
1
2
_
_
(9)
This integral, which must be nite in order for the system to
have long-range order, is still dicult to perform. However, the
integral may be written as a function of
s
(q) only.
_
s
2
_
=
1
NM

q,s
h

s
(q)
_
_
1
e

s
(q)
1
+
1
2
_
_
(10)
It would be convenient therefore, to introduce a density of
phonon states
Z() =
1
N

q,s
(
s
(q)) (11)
so that
_
s
2
_
=
h
M
_
dZ()
1

_
_
n() +
1
2
_
_
(12)
1.2 The Phonon Density of States
In addition to the calculation of
_
s
2
_
, the density of states Z()
is also useful in the calculation of E =< H >, the partition
function and the related thermodynamic properties
In order to calculate
Z() =
1
N

q,s
(
s
(q)) (13)
5
we must rst better dene the sum over q. As we discussed last
chapter for a 3-d cubic system, we will assume that we have a
periodic nite lattice of N basis points, or N
1/3
in each of the
principle lattice directions a
1
, a
2
, and a
3
. Then, we want the
Fourier representation to respect the periodic boundary condi-
tions (pbc), so
e
iq
(
r+N
1/3
(a
1
+a
2
+a
3
)
)
= e
iqr
(14)
This means that q
i
= 2m/N
1/3
, where m is an integer, and i
indicates one of the coordinates x, y or z. In addition, we only
G vector
Bisector
First Brillouin Zone
Figure 1: First Brillouin zone of the square lattice
want unique values of q
i
, so we will choose those within the rst
6
Brillouin zone, so that
G q
1
2
G
2
(15)
The size of this region is the same as that of a unit cell of the
reciprocal lattice g
1
(g
2
g
3
). Since there are N states in this
region, the density of q states is
N
g
1
(g
2
g
3
)
=
NV
c
(2)
3
=
V
(2)
3
(16)
where V
c
is the volume of a Bravais lattice cell (a
1
a
2
a
3
),
and V is the lattice volume. Clearly as N the density
increases until a continuum of states is formed (all that we need
here is that the spacing between q-states be much smaller than
any physically relevant value of q). The number of states in
a frequency interval d is then given by the volume of q-space
between the surfaces dened by =
s
(q) and =
s
(q)+d
multiplied by V/(2)
3
Z()d =
V
(2)
3
_
+d

d
3
q
=
V
(2)
3
d

s
_
d
3
q (
s
(q)) (17)
7
dq

dS

Surface
= (q)
s
q
y
q
x
Figure 2: States in q-space. S

is the surface of constant =


s
(q), so that d
3
q =
dS
w
dq

=
dSd
qs(q)
.
As shown in Fig. 2, d =
q

s
(q)dq

, and
d
3
q = dS
w
dq

=
dS

s
(q)
, (18)
where S

is the surface in q-space of constant =


s
(q). Then
Z()d =
V
(2)
3
d

s
_
=
s
(q)
dS

s
(q)
(19)
Thus the density of states is high in regions where the dispersion
is at so that
q

s
(q) is small.
As an example, consider the 1-d Harmonic chain shown in
Fig. 3. Real phonon dispersions have maxima which are not
8
a
M
(q)
q
/a -/a
linear
flat

0
Dispersion DOS

Z()
Figure 3: Linear harmonic chain. The phonon dispersion of this chain must in-
clude an acoustic mode, so (q) will be linear near q = 0, and it must be symmetric
about q = 0 and q = /q due to the point-group symmetry. Thus, the density of
states (DOS) will be at near = 0 corresponding to the acoustic mode (for which

q
(q) =constant), and will be divergent near =
0
corresponding to the peak of
the dispersion (where
q
(q) = 0).
at a zone boundary, with corresponding peaks in the phonon
DOS. However, any point within the Brillouin zone for which

q
(q) = 0 (cusp, maxima, minima) will yield an integrable
singularity in the DOS.
9
2 Models of Lattice Dispersion
2.1 The Debye Model
For most thermodynamic properties, we are interested in the
modes h(q) k
B
T which are low frequency modes in gen-
eral. From a very general set of (symmetry) constraints we have
-4 -2 0 2 4
q
0.0
0.5
1.0
1.5
2.0

-
Debye model
Figure 4: Dispersion for the diatomic linear chain. In the Debye model, we replace
the acoustic mode by a purely linear mode with the same initial dispersion and ignore
any optical modes.
argued that all interacting lattices in which the total energy is
invariant to an overall arbitrary rigid shift in the location of the
lattice must have at least one acoustic mode, where for small
10

s
(q) = c
s
[q[. Thus, for the thermodynamic properties of the
lattice, we care predominantly about the limit (q) 0. This
physics is rather accurately described by the Debye model.
In the Debye model, we will assume that all modes are acous-
tic (elastic), so that
s
(q) = c
s
[q[ for all s and q, then
q

s
(q) =
c
s
for all s and q, and
Z() =
V
(2)
3

s
_
dS

s
(q)
=
V
(2)
3

s
_
dS

c
s
(20)
The surface integral may be evaluated, and yields a constant
_
dS

= S
s
for each branch. Typically c
s
is dierent for dierent
modes. However, we will assume that the system is isotropic, so
c
s
= c. If the dispersion is isotopic, then the surface of constant

s
(q) is just a sphere, so the surface integral is trivial
S
s
( = (q)) =
_

_
2 for d = 1
2q = 2/c for d = 2
4q
2
= 4
2
/c
2
for d = 3
(21)
11
then since the number of modes = d
Z() =
V
(2)
3
_

_
2/c for d = 1
2q = 4/c
2
for d = 2
4q
2
= 12
2
/c
3
for d = 3
_

_
0 < <
D
(22)
Note that since the total number of states is nite, we have
introduced a cuto
D
on the frequency.
2.2 The Einstein Model
Real two dimensional systems, i.e., a monolayer of gas (He)
deposited on an atomically perfect surface (Vycor), may be bet-
ter described by an Einstein model where each atom oscillates
with a frequency
0
and does not interact with its neighbors.
The model is dispersionless (q) =
0
, and the DOS for this
system is a delta function Z() = c(
0
). Note that it does
not have an acoustic mode; however, this is not in violation of
the discussion in the last chapter. Why?
12
Figure 5: Helium adsorbed on a Vycor surface. Each He atom is attracted weakly to
the surface by a van der Waals attraction and sits in a local minimum of the surface
lattice potential.
3 Thermodynamics of Crystal Lattices
We are now in a situation to calculate many of the thermody-
namic properties of crystal lattices. However before addressing
such questions as the lattice energy free energy and specic
heat we should see if our model has long-range order... ie., is it
consistent with our initial assumptions.
13
3.1 Long-Range Order
For simplicity, we will work on an elemental lattice model. We
may dene long-ranged order (LRO) as a nite value of
s
2
) =
h
MN

q,s
1

s
(q)
_
_
1
e

s
(q)
1
+
1
2
_
_
=
h
2MN

q,s
sinh (
s
(q)/2)

s
(q)cosh (
s
(q)/2)
(23)
Since we expect all lattices to melt for some high temperature,
we are interested only in the T 0 limit. Clearly also given
the factor of
1

s
(q)
in the summand, we are most interested in
acoustic modes since they are the ones which will cause a di-
vergence.
lim

s
2
) =
h
2MN
lim

q,s
1

s
(q)
_
_
_
1

s
(q)
+
1
2
_
_
_ (24)
Clearly the low frequency modes are most important so a Debye
model may be used
lim

s
2
)
h
2MN
_

D
0
dZ()
1

_
_
1

+
1
2
_
_
, (25)
14
where Z() is the same as was dened above in Eq. 22.
Z() =
V
(2)
3
_

_
2/c for d = 1
2q = 4/c
2
for d = 2
4q
2
= 12
2
/c
3
for d = 3
_

_
0 < <
D
(26)
Thus in 3-d the DOS always cancels the 1/ singularity but
in two dimensions the singularity is only cancelled when T = 0
( = ), and in one dimension s
2
) = for all T. This
s
2
) d = 1 d = 2 d = 3
T = 0 nite nite
T ,= 0 nite
Table 1: s
2
for lattices of dierent dimension, assuming the presence of an acoustic
mode.
is a specic case of the Mermin-Wagner Theorem. We should
emphasize that the result s
2
) = does not mean that our
theory has failed. The harmonic approximation requires that
the near-neighbor strains must be small, not the displacements.
Physically, it is easy to understand why one-dimensional sys-
15
tems do not have long range order, since as you go along the
chain, the displacements of the atoms can accumulate to pro-
duce a very large rms displacement. In higher dimensional sys-
tems, the displacements in any direction are constrained by
the neighbors in orthogonal directions. Real two dimensional
time
Figure 6: Random uctuaions of atoms in a 1-d lattice may accumulate to produce
a very large average rms displacement of the atoms from small interatomic displace-
ments.
systems, i.e., a monolayer of gas deposited on an atomically
perfect surface, do have long-range order even at ntie temper-
atures due to the surface potential (corrugation of the surface).
These may be better described by an Einstein model where each
16
atom oscillates with a frequency
0
and does not interact with
its neighbors. The DOS for this system is a delta function as
described above. For such a DOS, s
2
) is always nite. You will
explore this physics, in much more detail, in your homework.
3.2 Thermodynamics
We will assume that our system is in equilibrium with a heat
bath at temperature T. This system is described by the canoni-
cal ensemble, and may be justied by dividing an innite system
into a nite number of smaller subsystems. Each subsystem is
expected to interact weakly with the remaining system which
also acts as the subsystems heat bath. The probability that any
state in the subsystem is occupied is given by
P (n
s
(k)) e
E({n
s
(k)})
(27)
Thus the partition function is given by
Z =

{n
s
(k)}
e
E({n
s
(k)})
=

{n
s
(k)}
e

k,s
h
s
(k)
(
n
s
(k)+
1
2
)
=

s,k
Z
s
(k) (28)
17
where Z
s
(k) is the partition function for the mode s, k; i.e. the
modes are independent and decouple.
Z
s
(k) =

n
e
h
s
(k)
(
n
s
(k)+
1
2
)
= e
h
s
(k)/2

n
e
h
s
(k)(n
s
(k))
=
e
h
s
(k)/2
1 e
h
s
(k)
=
1
2 sinh (
s
(k)/2)
(29)
The free energy is given by
T = k
B
T ln (Z) = k
B
T

k,s
ln (2 sinh ( h
s
(k)/2)) (30)
Since dc = TdoPdV and dT = TdoPdV TdoodT,
the entropy is
o =
_
_
T
T
_
_
V
, (31)
and system energy is then given by
c = T + To = T T
_
_
T
T
_
_
V
(32)
where constant volume V is guaranteed by the harmonic ap-
proximation (since < s >= 0).
c =

k,s
1
2
h
s
(k)coth (
s
(k)/2) (33)
18
The specic heat is then given by
C =
_
_
dc
dT
_
_
V
= k
B

k,s
( h
s
(k))
2
csch
2
(
s
(k)/2) (34)
where csch (x) = 1/sinh (x)
Consider the specic heat of our 3-dimensional Debye model.
C = k
B
_

D
0
dZ() (/2)
2
csch
2
(/2)
= k
B
_

D
0
d
_
_
_
12V
2
(2c)
3
_
_
_ (/2)
2
csch
2
(/2) (35)
Where the Debye frequency
D
is determined by the require-
ment that
3rN =
_

D
0
dZ() =
_

D
0
d
_
_
_
12V
2
(2c)
3
_
_
_ , (36)
or V/(2c)
3
= 3rN/(4
3
D
).
Clearly the integral for C is a mess, except in the high and
low T limits. At high temperatures h
D
/2 1,
C k
B
_

D
0
dZ() = 3Nrk
B
(37)
This is the well known classical result (equipartition theorem)
which attributes (1/2)k
B
of the specic heat to each quadratic
degree of freedom. Here for each element of the basis we have
19
6 quadratic degrees of freedom (three translational, and three
momenta).
At low temperatures, h/2 1,
csch
2
(/2) 2e
/2
(38)
Thus, at low T, only the low frequency modes contribute, so
the upper bound of integration may be extended to
C 12k
B
3rN
4
2
D
_

0
d
2
_
_
h
2
_
_
2
2e

s
(k)/2
. (39)
If we make the change of variables x = h/2, we get
C
9k
B
rN
2
_
_
1

D
h
_
_
3
_

0
dxx
4
e
x
(40)

9k
B
rN
2
_
_
1

D
h
_
_
3
24 (41)
Then, if we identify the Debye temperature
D
= h
D
/k
B
, we
get
C 96rNk
B
_
_
T

D
_
_
3
(42)
C T
3
at low temperature is the characteristic signature
of low-energy phonon excitations.
20
3.3 Thermal Expansion, the Gruneisen Parameter
Consider a cubic system of linear dimension L. If unconstrained,
we expect that the volume of this system will change with tem-
perature (generally expand with increasing T, but not always.
cf. ice or Si). We dene the coecient of free expansion (P = 0)
as

L
=
1
L
dL
dT
or
V
= 3
L
=
1
V
dV
dT
. (43)
Of course, this measurement only makes sense in equilibrium.
P =
_
_
dT
dV
_
_
T
= 0 (44)
As mentioned earlier, since < s >= 0 in the harmonic ap-
proximation, a harmonic crystal does not expand when heated.
Of course, real crystals do, so that lack of thermal expansion
of a harmonic crystal can be considered a limitation of the har-
monic theory. To address this limitation, we can make a quasi-
harmonic approximation. Consider a more general potential
between the ions, of the form
V (x) = bx + cx
3
+
1
2
m
2
x
2
(45)
21
and lets see if any of these terms will produce a temperature
dependent displacement. The last term is the usual harmonic
term, which we have already shown does not produce a T-
dependent < x >. Also the rst term does not have the desired
eect! It does correspond to a temperature-independent shift
in the oscillator, as can be seen by completing the square
1
2
m
2
x
2
+ bx =
1
2
_
_
x +
b
m
2
_
_
2

b
2
2m
2
(46)
Clearly < x >=
b
m
2
, independent of the temperature; that
is, assuming that b is temperature-independent. What we need
is a temperature dependent coecient b!
The cubic term has the desired eect. As can be seen in
Fig 7, as the average energy (temperature) of a particle trapped
in a cubic potential increases, the mean position of the particle
shifts. However, it also destroys the solubility of the model. To
get around this, approximate the cubic term with a mean-eld
decomposition.
cx
3
cx
_
x
2
_
+ c(1 )x
2
x) (47)
and treat these two terms separately (the new parameter is
22
-2 -1 0 1 2
x
0
1
2
3
V
(
x
)
c=0.1
c=0.0
Figure 7: Plot of the potential V (x) =
1
2
mx
2
+cx
3
when m = = 1 and c = 0.0, 0.1.
The average position of a particle < x > in the anharmonic potential, c = 0.1, will
shift to the left as the energy (temperature) is increased; whereas, that in the harmonic
potential, c = 0, is xed < x >= 0.
to be determined self-consistently, usually by minimizing the
free energy with respect to ). The rst term yields the needed
temperature dependent shift of < x >
1
2
m
2
x
2
+cx
_
x
2
_
=
1
2
m
2
_
_
_x +
c < x
2
>
m
2
_
_
_
2

c
2
< x
2
>
2
2m
2
(48)
so that < x >=
c<x
2
>
m
2
. Clearly the renormalization of the
equilibrium position of the harmonic oscillator will be temper-
ature dependent. While the second term, (1 )x
2
< x >,
23
yields a shift in the frequency

=
_
_
_1 +
2(1 ) < x >
m
2
_
_
_
1
2
(49)
which is a function of the equilibrium position. Thus a mean-
eld description of the cubic term is consistent with the observed
physics.
In what follows, we will approximate the eect of the anhar-
monic cubic term as a shift in the equilibrium position of the
lattice (and hence the lattice potential) and a change of to

; however, we imagine that the energy levels remain of the


form
E
n
= h

(< x >)(n +
1
2
), (50)
and that < x > varies with temperature, consistent with the
mean-eld approximation just described.
To proceed, imagine the cube of cubic system to be made up
of oscillators which are independent. Since the nal result can
be formulated as a sum over these independent modes, consider
only one. In equilibrium, where P =
_
dF
dV
_
T
= 0, the free
24
energy of one of the modes is
T = +
1
2
h + k
B
T ln
_
1 e
h
_
(51)
and (following the notation of Ibach and L uth), let the lattice
potential
=
0
+
1
2
f(a a
0
)
2
+ (52)
where f is the spring constant. Then
0 = P =
_
_
dT
da
_
_
T
= f(a a
0
) +
1

a
_
_
1
2
h
h
1 e
h
_
_
,
(53)
If we identify the last term in parenthesis as (, T), and solve
for a, then
a = a
0

1
f
(, T)

a
(54)
Since we now know a(T) for a single mode, we may calculate
the linear expansion coecient for this mode

L
=
1
a
0
da
dT
=
1
a
2
0
f
ln w
ln a
(, T)
T
(55)
To generalize this to a solid let
L

V
(as discussed above)
and a
2
0
f V
2 dP
dV
= V ( is the bulk modulus) and sum over
25
all modes the modes k, s

V
=
1
V
dV
dT
=
1
V

k,s

ln
s
(k)
ln V
(
s
(k), T)
T
. (56)
Clearly (due to the factor of

T
),
V
will have a behavior sim-
ilar to that of the specic heat (
V
T
3
for low T, and

V
=constant for high T). In addition, for many lattices, the
Gruneisen number
=
ln
s
(k)
ln V
(57)
shows a weak dependence upon s, k, and may be replaced by
its average, called the Gruneisen parameter
) =
_
ln
s
(k)
ln V
_
, (58)
typically on the order of two.
Before proceeding to the next section, I would like to reex-
amine the cubic term in a crystal where

l
s
3
l
=
1
(2MN)
3/2

l,k,q,p
h
3/2
_
(q)(k)(p)
e
i(p+q+kG)r
l
(59)
_
a(k) + a

(k)
_ _
a(p) + a

(p)
_ _
a(q) + a

(q)
_
.
The sum over l yields a delta function
p+q+k,G
(ie., crystal mo-
mentum conservation). Physically, these processes correspond
26
q - k
q
k (+ G)
q - k
- q
k (+ G)
Figure 8: Three-phonon processes resulting from cubic terms in the inter-ion potential.
Six other three-phonon processes are possible.
to phonon decay in which a phonon can decompose into two
others. As we shall see, these anharmonic processes are crucial
to the calculation of the thermal conductivity, , of crystals.
3.4 Thermal Conductivity
Metals predominately carry heat with free electrons, and are
considered to be good conductors. Insulators, which lack free
electrons, predominantly carry heat with lattice vibrations
phonons. Nevertheless, some very hard insulating crystals have
very high thermal conductivities - diamond C which is often
highly temperature dependent. However, most insulators are
27
material/T 273.2K 298.2K
C 26.2 23.2
Cu 4.03 4.01
Table 2: The thermal conductivities of copper and diamond (CRC) ( in Ohm-cm).
not good thermal conductors. This subsection will be devoted
to understanding what makes sti crystals like diamond such
good conductors of heat.
The thermal conductivity is measured by setting up a small
steady thermal gradient across the material, then
Q = T (60)
where Q is the thermal current density; i.e., the energy density
times the velocity. If the thermal current is in the x-direction,
then
Q
x
=
1
V

q,s
h
s
(q)n
s
(q))v
sx
(q) (61)
where the group velocity is given by v
sx
(q) =

s
(q)
q
x
Since we
assume T is small, we will only look at the linear response
of the system where n
s
(q)) deviates little from its equilibrium
28
value n
s
(q))
0
. Furthermore since
s
(q) =
s
(q),
v
sx
(q) =

s
(q)
q
x
= v
sx
(q) (62)
Thus as n
s
(q))
0
= n
s
(q))
0
Q
0
x
=
1
V

q,s
h
s
(q)n
s
(q))
0
v
sx
(q) = 0 (63)
since the sum is over all q in the B.Z. Thus if we expand
n
s
(q)) = n
s
(q))
0
+ n
s
(q))
1
+ (64)
we get
Q
x

1
V

q,s
h
s
(q)n
s
(q))
1
v
sx
(q) (65)
since we presumably already know
s
(k), the calculation of Q
and hence reduces to the evaluation of the linear change in
< n >.
Within a region, < n > can change in two ways. Either
phonons can diuse into the region, or they can decay through
an anharmonic (cubic) term into other modes. so
dn)
dt
=
n)
t

diusion
+
n)
t

decay
(66)
29
decay
diffusion
diffusion
Figure 9: Change of phonon density within a trapazoidal region. n
s
(q) can change
either by phonon decay or by phonon diusion into and out of the region.
However
dn
dt
= 0 since we are in a steady state. The decay
process is usually described by a relaxation time (or a mean-
free path l = v )
n)
t

decay
=
< n > < n >
0

=
< n >
1

(67)
The diusion part of
dn
dt
is addressed pictorially in Fig. 10.
Formally,
n)
t

diusion

n(x v
x
t)) n(x))
t

n(x))
x
v
x
(68)
v
x
n(x))
T
T
x
v
x
n(x))
0
+ n(x))
1
T
T
x
Keeping only the lowest order term,
n)
t

diusion
v
x
n(x))
0
T
T
x
(69)
30
v
x
v t
source region
region of
interest
Figure 10: Phonon diusion. In time t, all the phonons in the left, source region,
will travel into the region of interest on the right, while those on the right region will
all travel out in time t. Thus, n/t = (n
left
n
right
)/t.
Then as
dn
dt
= 0,
n)
t

diusion
=
n)
t

decay
(70)
or
n)
1
= v
x

s
(q)
n(x))
0
T
T
x
. (71)
Thus
Q
x

1
V

q,s
h
s
(q)v
2
sx
(q)
s
(q)
n(x))
0
T
T
x
, (72)
and since Q = T

1
V

q,s
h
s
(q)v
2
sx
(q)
n(x))
0
T
. (73)
From this relationship we can learn several things. First since
v
2
sx
(q), phonons near the zone boundary or optical modes
31
with small v
s
(q) =
q

s
(q) contribute little to the thermal
conductivity. Also, sti materials, with very fast speed of the
acoustic modes v
sx
(q) c will have a large . Second, since

s
(q), and l
s
(q) = v
sx
(q)
s
(q), will be small for materi-
als with short mean-free paths. The mean-free path is eected
by defects, anharmonic Umklapp processes, etc. We will explore
this eect, especially its temperature dependence, in more de-
tail.
At low T, only low-energy, acoustic, modes can be excited
(those with h
s
(q) k
B
T). These modes have
v
s
(q) = c
s
(74)
In addition, since the momentum of these modes q G, we
only have to worry about anharmonic processes which do not
involve a reciprocal lattice vector G in lattice momentum con-
servation. Consider one of the three-phonon anharmonic pro-
cesses of phonon decay shown in Fig. 8 (with G = 0). For these
processes Q hc so the thermal current is not disturbed by
anharmonic processes. Thus the anharmonic terms at low T
32
do not aect the mean-free path, so the thermal resistivity (the
inverse of the conductivity) is dominated by scattering from
impurities in the bulk and surface imperfections at low temper-
atures.
At high T momentum conservation in an anharmonic process
may involve a reciprocal lattice vector G if the q
1
of an excited
mode is large enough and there exists a suciently small G so
that q
1
> G/2 (c.f. Fig. 11). This is called an Umklapp process,
first Brillouin zone
G
q
1
q
2
q
3
q =
1
q + + G
2
q
3
q
1
q
2
q
3
Q
in
Q
out
Figure 11: Umklapp processes involve a reciprocal lattice vector G in lattice momen-
tum conservation. They are possible whenever q
1
> G/2, for some G, and involve a
virtual reversal of the momentum and heat carried by the phonons (far right).
and it involves a very large change in the heat current (almost
a reversal). Thus the mean-free path l and are very much
33
smaller for high temperatures where q
1
can be larger than half
the smallest G.
So what about diamond? It is very hard and very sti, so the
sound velocities c
s
are large, and so thermally excited modes for
which k
B
T h hcq involve small q
1
for which Umklapp
processes are irrelevant. Second c
2
which is large. Thus
for diamond is huge!
34
Chapter 6: The Fermi Liquid
L.D. Landau
December 22, 2000
Contents
1 introduction: The Electronic Fermi Liquid 3
2 The Non-Interacting Fermi Gas 5
2.1 Innite-Square-Well Potential . . . . . . . . . . . . . . . . . . . . . . 5
2.2 The Fermi Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 T = 0, The Pauli Principle . . . . . . . . . . . . . . . . . . . . 10
2.2.2 T = 0, Fermi Statistics . . . . . . . . . . . . . . . . . . . . . . 13
3 The Weakly Correlated Electronic Liquid 23
3.1 Thomas-Fermi Screening . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Fermi liquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 Quasi-particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.1 Particles and Holes . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.2 Quasiparticles and Quasiholes at T = 0 . . . . . . . . . . . . . 33
3.4 Energy of Quasiparticles. . . . . . . . . . . . . . . . . . . . . . . . . . 39
4 Interactions between Particles: Landau Fermi Liquid 42
4.1 The free energy, and interparticle interactions . . . . . . . . . . . . . 42
4.2 Local Energy of a Quasiparticle . . . . . . . . . . . . . . . . . . . . . 46
1
4.2.1 Equilibrium Distribution of Quasiparticles at Finite T . . . . . 48
4.3 Eective Mass m

of Quasiparticles . . . . . . . . . . . . . . . . . . . 50
2
1 introduction: The Electronic Fermi Liquid
As we have seen, the electronic and lattice degrees of freedom
decouple, to a good approximation, in solids. This is due to the
dierent time scales involved in these systems.

ion
1/
D

electron

h
E
F
(1)
where E
F
is the electronic Fermi energy. The electrons may be
thought of as instantly reacting to the (slow) motion of the lat-
tice, while remaining essentially in the electronic ground state.
Thus, to a good approximation the electronic and lattice degrees
of freedom separate, and the small electron-lattice (phonon) in-
teraction (responsible for resistivity, superconductivity etc) may
be treated as a perturbation (with
D
/E
F
as an expansion pa-
rameter); that is if we are capable of solving the problem of the
remaining purely electronic system.
At rst glance the remaining electronic problem would also
appear to be hopeless since the (non-perturbative) electron-
electron interactions are as large as the combined electronic
kinetic energy and the potential energy due to interactions with
3
the static ions (the latter energy, or rather the corresponding
part of the Hamiltonian, composes the solvable portion of the
problem). However, the Pauli principle keeps low-lying orbitals
from being multiply occupied, so is often justied to ignore the
electron-electron interactions, or treat them as a renormaliza-
tion of the non-interacting problem (eective mass) etc. This
will be the initial assumption of this chapter, in which we will
cover
the non-interacting Fermi liquid, and
the renormalized Landau Fermi liquid (Pines Nozieres).
These relatively simple theories resolved some of the most
important puzzles involving metals at the turn of the century.
Perhaps the most intriguing of these is the metallic specic heat.
Except in certain heavy fermion metals, the electronic contri-
bution to the specic heat is always orders of magnitude smaller
than the phonon contribution. However, from the classical theo-
rem of equipartition, if each lattice site contributes just one elec-
tron to the conduction band, one would expect the contributions
4
from these sources to be similar (C
electron
C
phonon
3Nrk
B
).
This puzzle is resolved, at the simplest level: that of the non-
interacting Fermi gas.
2 The Non-Interacting Fermi Gas
2.1 Innite-Square-Well Potential
We will proceed to treat the electronic degrees of freedom, ig-
noring the electron-electron interaction, and even the electron-
lattice interaction. In general, the electronic degrees of freedom
are split into electrons which are bound to their atomic cores
with wavefunctions which are essentially atomic, unaected by
the lattice, and those valence (or near valence) electrons which
react and adapt to their environment. For the most part, we are
only interested in the valence electrons. Their environment de-
scribed by the potential due to the ions and the core electrons
the core potential. Thus, ignoring the electron-electron interac-
tions, the electronic Hamiltonian is
H =
P
2
2m
+ V (r) . (2)
5
As shown in Fig. 1, the core potential V (r), like the lattice, is
periodic
a
V(r)
V(r+a) = V(r)
Figure 1: Schematic core potential (solid line) for a one-dimensional lattice with
lattice constant a.
For the moment, ignore the core potential, then the electronic
wave functions are plane waves e
ikr
. Now consider the
core potential as a perturbation. The electrons will be strongly
eected by the periodicity of the potential when = 2/k a
1
. However, when k is small so that a (or when k is large,
so a) the structure of the potential may be neglected, or
we can assume V (r) = V
0
anywhere within the material. The
1
Interestingly, when a, the Bragg condition 2d sin a may easily be satised, so the
electrons, which may be though of as DeBroglie waves, scatter o of the lattice. Consequently states
for which = 2/k a are often forbidden. This is the source of gaps in the band structure, to be
discussed in the next chapter.
6
potential still acts to conne the electrons (and so maintain
charge neutrality), so V (r) = anywhere outside the material.
Figure 2: Innite square-well potential. V (r) = V
0
within the well, and V (r) =
outside to conne the electrons and maintain charge neutrality.
Thus we will approximate the potential of a cubic solid with
linear dimension L as an innite square-well potential.
V (r) =
_

_
V
0
0 < r
i
< L
otherwise
(3)
The electronic wavefunctions in this potential satisfy

h
2
2m

2
(r) = (E

V
0
) (r) = E(r) (4)
7
The normalize plane wave solution to this model is
(r) =
3

i=1
_
_
2
L
_
_
1/2
sin k
i
x
i
where i = x, y, or z (5)
and k
i
L = n
i
in order to satisfy the boundary condition that
= 0 on the surface of the cube. Furthermore, solutions with
n
i
< 0 are not independent of solutions with n
i
> 0 and may
be excluded. Solutions with n
i
= 0 cannot be normalized and
are excluded (they correspond to no electron in the state). The
/L
k
x
k
y
k
z
(/L)
3
Figure 3: Allowed k-states for an electron conned by a innite-square potential. Each
state has a volume of (/L)
3
in k-space.
eigenenergies of the wavefunctions are

h
2

2
2m
=
h
2m

i
k
2
i
=
h
2

2
2mL
2
_
n
2
x
+ n
2
y
+ n
2
z
_
(6)
8
and as a result of these restrictions, states in k-space are con-
ned to the rst quadrant (c.f. Fig. 3). Each state has a volume
(/L)
3
of k-space. Thus as L , the number of states with
energies E(k) < E < E(k) + dE is
dZ

=
(4k
2
dk)/8
(/L)
3
. (7)
Then, since E =
h
2
k
2
2m
, so k
2
dk =
m
h
2
_
2mE/ h
2
dE
dZ = dZ

/L
3
=
1
4
2
_
_
2m
h
2
_
_
3/2
E
1/2
dE . (8)
or, the density of state per unit volume is
D(E) =
dZ
dE
=
1
4
2
_
_
2m
h
2
_
_
3/2
E
1/2
. (9)
Up until now, we have ignored the properties of electrons. How-
ever, for the DOS, it is useful to recall that the electrons are
spin-1/2 thus 2S + 1 = 2 electrons can ll each orbital or k-
state, one of spin up the other spin down. If we account for this
spin degeneracy in D, then
D(E) =
1
2
2
_
_
2m
h
2
_
_
3/2
E
1/2
. (10)
9
2.2 The Fermi Gas
2.2.1 T = 0, The Pauli Principle
Electrons, as are all half-integer spin particles, are Fermions.
Thus, by the Pauli Principle, no two of them may occupy the
same state. For example, if we calculate the density of electrons
per unit volume
n =
_

0
D(E)f(E, T)dE , (11)
where f(E, T) is the probability that a state of energy E is oc-
cupied, the factor f(E, T) must enforce this restriction. How-
ever, f is just the statistical factor; c.f. for classical particles
f(E, T) = e
E/k
B
T
for classical particles , (12)
which for T = 0 would require all the electrons to go into
the ground state f(0, 0) = 1. Clearly, this violates the Pauli
principle.
At T = 0 we need to put just one particle in each state, start-
ing from the lowest energy state, until we are out of particles.
Since E k
2
in our simple square-well model, will ll up all
10
k-states until we reach some Fermi radius k
F
, corresponding to
some Fermi Energy E
F
k
x
k
y
k
z
k
f
k
f
occupied
states
h k
2 2
f
2m
= E
f
D(E)
Figure 4: Due to the Pauli principle, all k-states up to k
F
, and all states with energies
up to E
f
are lled at zero temperature.
E
F
=
h
2
k
2
F
2m
, (13)
thus,
f(E, T = 0) = (E
F
E) (14)
and
n =
_

0
D(E)f(E, T)dE =
_
E
F
0
D(E)DE
=
_
_
2m
h
2
_
_
3/2
1
2
2
_
E
F
0
E
1/2
DE
=
_
_
2m
h
2
_
_
3/2
1
2
2
2
3
E
2/3
F
, (15)
11
or
E
F
=
h
2
2m
_
3
2
n
_
3/2
= k
B
T
F
(16)
which also denes the Fermi temperature T
F
. Thus for metals,
in which n 10
23
/cm
3
, E
F
10
11
erg 10eV k
B
10
5
K.
Notice that due to the Pauli principle, the average energy of
the electrons will be nite, even at T = 0!
c =
_
E
F
0
D(E)EdE =
3
5
nE
F
. (17)
However, it is the electrons near E
F
in energy which may be
excited and are therefore important. These have a DeBroglie
wavelength of roughly

e
=
12.3

A
(E(eV))
1/2
4

A
(18)
thus our original approximation of a square well potential, ig-
noring the lattice structure, is questionable for electrons near
the Fermi surface, and should be regarded as yielding only qual-
itative results.
12
2.2.2 T = 0, Fermi Statistics
At nite temperatures some of the states will be thermally ex-
cited. The energy available for these excitations is roughly
k
B
T, and the only possible excitations are from lled to un-
lled electronic states. Therefore, only the states within k
B
T
(E
F
k
B
T < E < E
F
+ k
B
T) of the Fermi surface may be
excited. f(E, T) must be modied accordingly.
What we need is then f(E, T) at nite T which also satises
the Pauli principle. Lets return to our model of a periodic solid
which is constructed by bringing individual atoms together from
an innite separation. First, just consider a solid constructed
from only two atoms, each with a single orbital (Fig. 5). For
1 2
n = -1
1
n = +1
1
Figure 5: Exchange of electrons in a solid composed of two orbitals.
13
this system, in equilibrium,
0 = F =

i
F
n
i
n
i
(19)
electrons are conserved so

i
n
i
= 0. Thus, for our two orbital
system
F
n
1
n
1
+
F
n
2
n
2
= 0 and n
1
+ n
2
= 0 (20)
or
F
n
1
=
F
n
2
(21)
A similar relation holds for an arbitrary number of particles.
Apparently this quantity, the increased free energy needed to
add a particle to the system, is a constant
F
n
i
= (22)
for all i. is called the chemical potential.
Now consider an ensemble of orbitals. We will treat the ther-
modynamics of this system within the canonical ensemble (i.e.
the system is in contact with a thermal bath, and the particle
number is conserved) for which T = c To is the appropriate
potential. The system energy c and Entropy o may be written
14
as functions of the orbital energies E
i
and occupancies n
i
and
the degeneracy g
i
of the state of energy E
i
. For example,
g = 2
1
E
1
E
2
E
3
g = 4
3
g = 2
2
E
4
g = 4
4
n = 2
1
n = 4
3
n = 1
2
n = 2
4
Figure 6: states from an ensemble of orbitals.
c =

i
n
i
E
i
. (23)
The entropy o requires a bit more thought. If P is the number
of ways of distributing the electrons among the states, then
o = k
B
ln P . (24)
Consider a set of g
i
states with energy E
i
. The number of
ways of distributing the rst electron in these states is g
i
. For
a second electron we then have g
i
1 ways... etc. So for n
i
electrons there are
g
i
!
n
i
!(g
i
n
i
)!
(25)
15
possible ways of accommodating the n
i
(indistinguishable) elec-
trons in g
i
states.
The number of ways of making the whole system (ie, lling
energy levels with E
i
,= E
j
) is then
P =

i
g
i
!
n
i
!(g
i
n
i
)!
, (26)
and so, the entropy
o = k
B

i
ln g
i
! ln n
i
! ln(g
i
n
i
)! . (27)
For large n, ln n! nln n n, so
o = k
B

i
g
i
ln g
i
n
i
ln n
i
(g
i
n
i
) ln(g
i
n
i
) (28)
and
T =

i
n
i
E
i
k
B
T

i
g
i
ln g
i
n
i
ln n
i
(g
i
n
i
) ln(g
i
n
i
)
(29)
We will want to use the chemical potential in our thermo-
dynamic calculations
=
T
n
k
= E
k
+ k
B
T (ln n
k
+ 1 ln(g
k
n
k
) 1) , (30)
where = 1/k
B
T. Solving for n
k
n
k
=
g
k
1 + e
(E
k
)
. (31)
16
Thus the probability that a quantum state with energy E is
occupied, is (the Fermi function)
f(E, T) =
1
1 + e
(E
k
)
. (32)
At T = 0, = , and f(E, 0) = (E). Thus (T = 0) =
E
F
. However in general is temperature dependent, since it
must be adjusted to keep the particle number xed. In addition,
0.0 0.5 1.0 1.5 2.0

0.0
0.5
1.0
1.5
1
/
(
e

)
+
1
)
Figure 7: Plot of the Fermi function 1/
_
e
()
+ 1
_
when = 1/k
B
T = 20 and
= 1. Not that at energies the Fermi function displays a smooth step of width
k
B
T = 0.05. This allows thermal excitations of particles near the Fermi surface.
when T ,= 0, f becomes less sharp at energies E . This
reects the fact that particles with energies E k
B
T may
be excited to higher energy states.
17
Specic Heat The form of f(E, T) also claries why the elec-
tronic specic heat of metals is so small compared to the clas-
sical result C
classical
=
3
2
nk
B
T. The reason is simple: only the
electrons with energies within about k
B
T of the Fermi surface
may be excited (about
k
B
T
E
F
of the electron density) each with
excitation energy of about k
B
T. Therefore,
U
excitation
k
B
Tn
k
B
T
E
F
= nk
B
T
T
T
F
(33)
so
C nk
B
T
T
F
(34)
Then as T T
F
(T
F
is typically about 10
5
K in most metals
2
)
C nk
B
T
T
F
C
classical
nk
B
. Thus at temperatures where
the phonons contribute essentially a classical result to the spe-
cic heat, the electronic contribution is vanishingly small. In
general this holds except at very low T where the phonon con-
tribution C
phonon
T
3
goes to zero faster than the electronic
contribution to the specic heat.
2
Heavy Fermion systems are the exception to this rule. There T
F
can be as small as a fraction
of a degree Kelvin. As a result, they may have very large electronic specic heats.
18
Specic Heat Calculation Of course, since we know the free energy
of the non-interacting Fermi gas, we can calculate the form of
the specic heat. Here we will follow Ibach and L uth and Kittel;
however, since the chemical potential does depend upon the
temperature, I would like to make the approximations we make
a bit more explicit.
Upon heating from T = 0 to nite T, the Fermi gas will gain
energy
U(T) =
_

0
dE ED(E)f(E, T)
_
E
F
0
dE ED(E) (35)
so
C
V
=
dU
dT
=
_

0
dE ED(E)
df(E, T)
dT
. (36)
Then since at constant volume the electronic density is constant,
so
dn
dT
= 0, and n =
_

0
dED(E)f(E, T),
0 = E
F
dn
dT
=
_

0
dE E
F
D(E)
df(E, T)
dT
(37)
so we may write
C
V
=
_

0
dE (E E
F
) D(E)
df
dT
. (38)
In f, the temperature T enters through both = 1/k
B
T and
19

df
dT
=
f

T
+
f

T
=
e
(E)
_
e
(E)
+ 1
_
2
_
_
E
T


T
_
_
(39)
However,

T
depends upon the details of the density of states
near the Fermi surface, which can dier greatly from material
to material. Furthermore,

T
< 1 especially in common metals
at temperatures T T
F
, and the rst term
E
T
is of order
one (c.f. Fig. 7). Thus, for now we will neglect

T
relative to
E
T
(you will explore the validity of this approximation in your
homework), and, consistent with this approximation, replace
by E
F
, so
df
dT

e
(EE
F
)
_
e
(EE
F
)
+ 1
_
2
E E
F
T
, (40)
and
C
V

1
k
B
T
2
_

0
dE
(E E
F
)
2
D(E)e
(EE
F
)
_
e
(EE
F
)
+ 1
_
2
let x = (E E
F
)
k
B
T
_

E
F
dxD
_
_
x

+ E
F
_
_
x
2
e
x
(e
x
+ 1)
2
(41)
As shown in Fig. 8, the function x
2 e
x
(e
x
+1)
2
is only large in the
region 10 < x < 10. In this region, and for temperatures
20
-10 -5 0 5 10
x
0.0
0.1
0.2
0.3
0.4
0.5
x
2
e
x
/
(
e
x
+
1
)
2
Figure 8: Plot of x
2 e
x
(e
x
+1)
2
vs. x. Note that this function is only nite for roughly
10 < x < 10. Thus, at temperatures T T
F
10
5
K, we can approximate
D
_
x

+ E
F
_
D(E
F
) in Eq. 42.
T T
F
, D
_
x

+ E
F
_
D(E
F
), since the density of states
usually does not have features which are sharp on the energy
scale of 10k
B
T. Thus
C
V
k
B
TD(E
F
)
_

E
F
dxx
2
e
x
(e
x
+ 1)
2


2
3
k
2
B
TD(E
F
) . (42)
Note that no assumption about the form of D(E) was made
other than the assumption that it is smooth within k
B
T of the
Fermi surface. Thus, experimental measurements of the specic
heat at constant volume of the electrons, gives us information
21
about the density of electronic states at the Fermi surface.
Now lets reconsider the DOS for the 3-D box potential.
D(E) =
1
2
2
_
_
2m
h
_
_
3/2
E
1/2
= D(E
F
)
_
_
E
E
F
_
_
1/2
(43)
For which n =
_
E
F
0
D(E)dE = D(E
F
)
2
3
E
F
, so
C
V
=

2
2
nk
B
T
T
F

3
2
nk
B
(44)
where the last term on the right is the classical result. For
room temperatures T 300K, which is also of the same order
of magnitude as the Debye temperatures
D
,
C
V phonon

3
2
nk
B
C
V electron
(45)
So, the only way to measure the electronic specic heat in most
materials is to go to very low temperatures T
D
, for which
C
V phonon
T
3
. Here the total specic heat
C
V
T + T
3
(46)
We will see that gives us some measurement of the electronic ef-
fective mass for our Fermi liquid theory. I.e. it tell us something
about electron- electron interactions.
22
3 The Weakly Correlated Electronic Liquid
3.1 Thomas-Fermi Screening
As an introduction to the eect of electronic correlations, con-
sider the eect of a charged oxygen defect in one of the copper-
oxygen planes of a cuprate superconductor shown in Fig. 9.
Assume that the oxygen defect captures two electrons from the
metallic band, going from a 2s
2
2p
4
to a 2s
2
2p
6
conguration.
The defect will then become a cation, and have a net charge
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
Cu O
O
O
q=2e-
Figure 9: A charged oxygen defect is introduced into one of the copper-oxygen planes of
a cuprate superconductor. The oxygen defect captures two electrons from the metallic
band, going from a 2s
2
2p
4
to a 2s
2
2p
6
conguration.
of two electrons. In the vicinity of this oxygen defect, the elec-
23
trostatic potential and the electronic charge density will be re-
duced.
If we model the electronic density of states in this material
with our box-potential DOS, we can think of this reduction in
the local charge density in terms of raising the DOS parabola
near the defect (cf. Fig. 10). This will cause the free electronic
-eU
E
F
e
near charged
defect
Away from
charged defect
Figure 10: The shift in the DOS parabola near a charged defect.
charge to ow away from the defect. Near the defect (since
e < 0 and hence eU(r
near
) < 0)
n(r
near
)
_
E
F
+eU(r
near
)
0
D(E)DE (47)
24
While away from the defect, U(r
away
) = 0, so
n(r
away
)
_
E
F
0
D(E)DE (48)
or
n(r)
_
E
F
+eU(r)
0
D(E)DE
_
E
F
0
D(E)DE (49)
If [eU[ E
F
, then
n(r) D(E
F
) [E
F
+ eU E
F
] = eUD(E
F
) . (50)
We can solve for the change in the electrostatic potential by
solving Poisson equation.

2
U = 4 = 4en = 4e
2
D(E
F
)U . (51)
Let
2
= 4e
2
D(E
F
), then
2
U =
2
U has the solution
3
U(r) =
qe
r
r
(52)
The length 1/ = r
TF
is known as the Thomas-Fermi screening
length.
r
TF
=
_
4e
2
D(E
F
)
_
1/2
(53)
3
The solution is actually Ce
r
/r, where C is a constant. C may be deterined by letting D(E
F
) =
0, so the medium in which the charge is embedded becomes vacuum. Then the potential of the charge
is q/r, so C = q.
25
Lets estimate this distance for our square-well model,
r
2
TF
=
a
0

3(3
2
n)
1/3

a
0
4n
1/3
r
TF

1
2
_
_
n
a
3
0
_
_
1/6
(54)
In Cu, for which n 10
23
cm
3
(and since a
0
= 0.53

A
)
r
TFCu

1
2
_
10
23
_
1/6
(0.5 10
8
)
1/2
0.5 10
8
cm = 0.5

A
(55)
Thus, if we add a charge defect to Cu metal, the eect of the
defects ionic potential is screened away for distances r >
1
2

A
.
-
e
-
r
/
r
T
F
/
r
r
TF
=1/4
r
r
TF
=1
r
-
e
-
r
/
r
T
F
/
r
bound states
free states
r
TF
= n
-1/6
Figure 11: Screened defect potentials. As the screening length increases, states that
were free, become bound.
Now consider an electron bound to an ion in Cu or some other
metal. As shown in Fig. 11 the screening length decreases, and
bound states rise up in energy. In a weak metal (i.e. something
26
like YBCO), in which the valence state is barely free, a reduction
in the number of carriers (electrons) will increase the screening
length, since
r
TF
n
1/6
. (56)
This will extend the range of the potential, causing it to trap
or bind more statesmaking the one free valance state bound.
Now imagine that instead of a single defect, we have a con-
centrated system of such ions, and suppose that we decrease
the density of carriers (i.e. in Si-based semiconductors, this is
done by doping certain compensating dopants, or even by mod-
ulating the pressure). This will in turn, increase the screening
length, causing some states that were free to become bound,
causing an abrupt transition from a metal to an insulator, and
is believed to explain the MI transition in some transition-metal
oxides, glasses, amorphous semiconductors, etc.
27
3.2 Fermi liquids
The purpose of these next several lectures is to introduce you
to the theory of the Fermi liquid, which is, in its simplest form,
a collection of Fermions in a box plus interactions.
In reality , the only physical analog is a gas of
3
He, which
due its nuclear spin (the nucleus has two protons, one neutron),
obeys Fermi statistics for suciently low energies or temper-
atures. In addition, simple metals, from the rst or second
column of the periodic table, for which we may approximate
the ionic potential
V (R) = V
0
(57)
are a close approximant to Fermi liquids.
Moreover, Fermi Liquid theory only describes the gaseous
phase of these quantum fermion systems. For example,
3
He also
has a superuid (triplet), and at least in
4
He-
3
He mixtures, a
solid phase exists which is not described by Fermi Liquid The-
ory. One should note; however, that the Fermi liquid theory
state does serve as the starting point for the theories of super-
28
conductivity and super uidity.
One may construct Fermi liquid theory either starting from a
many-body diagrammatic or phenomenological viewpoint. We,
as Landau, will choose the latter. Fermi liquid theory has 3
basic tenants:
1. momentum and spin remain good quantum numbers to de-
scribe the (quasi) particles.
2. the interacting system may be obtained by adiabatically
turning on a particle-particle interaction over some time t.
3. the resulting excitations may be described as quasi-particles
with lifetimes t.
3.3 Quasi-particles
The last assumption involves a new concept, that of the quasi-
particles which requires some explanation.
3.3.1 Particles and Holes
Particles and Holes are excitations of the non-interacting system
at zero temperature. Consider a system of N free Fermions
29
each of mass m in a volume V . The eigenstates are the anti-
symmetrized combinations (Slater determinants) of N dierent
single particle states.

p
(r) =
1

V
e
ipr/ h
(58)
The occupation of each of these states is given by n
p
= (pp
F
)
where p
F
is the radius of the Fermi sphere. The energy of the
system is
E =

p
n
p
p
2
2m
(59)
and p
F
is given by
N
V
=
1
3
2
_
p
F
h
_
3
(60)
Now lets add a particle to the lowest available state p = p
F
then, for T = 0,
= E
0
(N + 1) E
0
(N) =
E
0
N
=
p
2
F
2m
. (61)
If we now excite the system, we will promote a certain number
of particles across the Fermi surface S
F
yielding particles above
and an equal number of vacancies or holes below the Fermi
surface. These are our elementary excitations, and they are
30
quantied by n
p
= n
p
n
0
p
n
p
=
_

p,p
for a particle p

> p
F

p,p
for a hole p

< p
F
. (62)
If we consider excitations created by thermal uctuations, then
E
F
E
D(E)
particle
excitation
hole
excitation
n = -1
p
n = 1
p
Figure 12: Particle and hole excitations of the Fermi gas.
n
p
1 only for excitations of energy within k
B
T of E
F
. The
energy of the non-interacting system is completely characterized
as a functional of the occupation
E E
0
=

p
p
2
2m
(n
p
n
0
p
) =

p
p
2
2m
n
p
. (63)
Now lets take our system and place it in contact with a par-
ticle bath. Then the appropriate potential is the free energy,
31
E
D(E)
F = p /2m -
2
F = - p /2m
2
E = = p /2m
F
2
F
F = | - p /2m |
2
Figure 13: Since = p
2
F
/2m, the free energy of a particle or a hole is F =
|p
2
/2m | > 0, so the system is stable to these excitations.
which for T = 0, is F = E N, and
F F
0
=

p
_
_
_
p
2
2m

_
_
_ n
p
. (64)
The free energy of a particle, with momentum p and n
p
=

p,p
is
p
2
2m
and it corresponds to an excitation outside S
F
.
The free energy of a hole n
p
=
p,p
is
p
2
2m
, which corre-
sponds to an excitation within S
F
. However, since = p
2
F
/2m,
the free energy of either at p = p
F
is zero, hence the free energy
of an excitation is

p
2
/2m

, (65)
which is always positive; ie., the system is stable to excitations.
32
3.3.2 Quasiparticles and Quasiholes at T = 0
a
U e
e
2
a

-a/r
TF
Figure 14: Model for a fermi liquid: a set of interacting particles an average distance
a apart bound within an innite square-well potential.
Now lets consider a system with interacting particles an av-
erage distance a apart, so that the characteristic energy of in-
teraction is
e
2
a
e
a/r
TF
. We will imagine that this system evolves
slowly from an ideal or noninteracting system in time t (i.e. the
interaction U
e
2
a
e
a/r
TF
is turned on slowly, so that the non-
interacting system evolves while remaining in the ground state
into an interacting system in time t).
If the eigenstate of the ideal system is characterized by n
0
p
,
then the interacting system eigenstate will evolve quasistatis-
tically from n
0
p
to n
p
. In fact if the system is isotropic and
33
remains in its ground state, then n
0
p
= n
p
. However, clearly in
some situations (superconductivity, magnetism) we will neglect
some eigenstates of the interacting system in this way.
Now lets add a particle of momentump to the non-interacting
ideal system, and slowly turn on the interaction. As U is
p
time = 0
U = 0
p p
time = t
U = (e /a) exp(-a/r )
2
TF
Figure 15: We add a particle with momentum p to our noninteracting (U = 0) Fermi
liquid at time t = 0, and slowly increase the interaction to its full value U at time t.
As the particle and system evolve, the particle becomes dressed by interactions with
the system (shown as a shaded ellipse) which changes the eective mass but not the
momentum of this single-particle excitation (now called a quasi-particle).
switched on, we slowly begin to perturb the particles close to
the additional particle, so the particle becomes dressed by these
interactions. However since momentum is conserved, we have
created an excitation (particle and its cloud) of momentum p.
We call this particle and cloud a quasiparticle. In the same way,
34
if we had introduced a hole of momentum p below the Fermi
surface, and slowly turned on the interaction, we would have
produced a quasihole.
Note that this adiabatic switching on procedure will have
diculties if the lifetime of the quasi-particle < t. If so,
then the process is not reversible. If we shorten t so that again
t, then the switching on of U may not be adiabatic (ie., we
will evolve to a system which is not in its ground state). Such
diculties do not arise so long as the energy of the particle is
close to the Fermi energy. Here there are few states accessible
for creating particle-hole excitations. In fact, one can formulate
a perturbative argument that the lifetime of a quasi-particle is
proportional to the square of its excitation energy above the
Fermi energy =
p
2
2m

p
2
F
2m
v(p p
F
).
To estimate this lifetime consider the following argument
from AGD: A particle with momentum p
1
above the Fermi
surface (p
1
> p
F
) interacts with one of the particles below the
Fermi surface with momentum p
2
. As a result, two new par-
ticles appear above the Fermi surface (all other states are full)
35
p
1
p
4
p
2
p
3
Figure 16: A particle with momentum p
1
above the Fermi surface (p
1
> p
F
) interacts
with one of the particles below the Fermi surface with momentum p
2
. As a result, two
new particles appear above the Fermi surface (all other states are full) with momenta
p
3
and p
4
..
with momenta p
3
and p
4
. This may also be interpreted as a
particle of momentum p
1
decaying into particles with momenta
p
3
and p
4
and a hole with momentum p
2
. By Fermis golden
rule, the total probability of such a process if proportional to
1


_
(
1
+
2

4
) d
3
p
2
d
3
p
3
(66)
where
1
=
p
2
1
2m
E
F
, and the integral is subject to the con-
straints of energy and momentum conservation and that
p
2
< p
F
, p
3
> p
F
, p
4
= [p
1
+ p
2
p
3
[ > p
F
(67)
36
It must be that
1
+
2
=
3
+
4
> 0 since both particles 3
and 4 must be above the Fermi surface. However, since
2
< 0,
if
1
is small, then [
2
[
<


1
is also small, so only of order

1
/E
F
states may scatter with the state k
1
, conserve energy,
and obey the Pauli principle. Thus, restricting
2
to a narrow
shell of width
1
/E
F
near the Fermi surface, and reducing the
scattering probability 1/ by the same factor.
Now consider the constraints placed on states k
3
and k
4
by
momentum conservation
k
1
k
3
= k
4
k
2
. (68)
Since
1
and
2
are conned to a narrow shell around the Fermi
surface, so too are
3
and
4
. This can be seen in Fig. 17, where
the requirement that k
1
k
3
= k
4
k
2
limits the allowed
states for particles 3 and 4. If we take k
1
xed, then the allowed
states for 2 and 3 are obtained by rotating the vectors k
1
k
3
=
k
4
k
2
; however, this rotation is severely limited by the fact that
particle 3 must remain above, and particle 2 below, the Fermi
surface. This restriction on the nal states further reduces the
scattering probability by a factor of
1
/E
F
.
37
k
1 k
4
E
F
3
2 1
4
k - k
1 3
k - k
4 2
E
N(E)
k
y
k
x
k
3
k
2
Figure 17: A quasiparticle of momentum p
1
decays via a particle-hole excitation
into a quasiparticle of momentum p
4
. This may also be interpreted as a particle
of momentum p
1
decaying into particles with momenta p
3
and p
4
and a hole with
momentum p
2
. Energy conservation requires |
2
|
<


1
. Thus, restricting
2
to a
narrow shell of width
1
/E
F
near the Fermi surface. Momentum conservation k
1

k
3
= k
4
k
2
further restricts the available states by a factor of about
1
/E
F
. Thus
the lifetime of a quasiparticle is proportional to
_

1
E
F
_
2
.
Thus, the scattering rate 1/ is proportional to
_

1
E
F
_
2
so that
excitations of suciently small energy will always be suciently
long lived to satisfy the constraints of reversibility. Finally, the
fact that the quasiparticle only interacts with a small number of
other particles due to Thomas-Fermi screening (i.e. those within
a distance R
TF
), also signicantly reduces the scattering rate.
38
3.4 Energy of Quasiparticles.
As in the non-interacting system, excitations will be quanti-
ed by the deviation of the occupation from the ground state
occupation n
0
p
n
p
= n
p
n
0
p
. (69)
At low temperatures n
p
1 only for p p
F
where the par-
ticles are suciently long lived that t. It is important to
emphasize that only n
p
not n
0
p
or n
p
, will be physically rele-
vant. This is important since it does not make much sense to
talk about quasiparticle states, described by n
p
, far from the
Fermi surface since they are not stable.
For the ideal system
E E
0
=

p
p
2
2m
n
p
. (70)
For the interacting system E[n
p
] becomes much more compli-
cated. If however n
p
is small (so that the system is close to
its ground state) then we may expand:
E[n
p
] = E
o
+

p

p
n
p
+ O(n
2
p
) , (71)
39
where
p
= E/n
p
. Note that
p
is intensive (ie. it is indepen-
dent of the system volume). If n
p
=
p,p
, then E E
0
+
p
;
i.e., the energy of the quasiparticle of momentum p

is
p
.
In practice we will only need
p
near the Fermi surface where
n
p
is nite. So we may approximate

p
+ (p p
F
)
p

p
[
p
F
(72)
where
p

p
= v
p
, the group velocity of the quasiparticle. The
ground state of the N +1 particle system is obtained by adding
a particle with
p
=
F
= =
E
0
N
(at zero temperature); which
denes the chemical potential . We make learn more about

p
by employing the symmetries of our system. If we explicitly
display the spin-dependence,

p,
=
p,
under time-reversal (73)

p,
=
p,
under BZ reection (74)
So
p,
=
p,
=
p,
; i.e. in the absence of an external
magnetic eld,
p,
does not depend upon if. Furthermore,
for an isotropic system
p
depends only upon the magnitude of
p, [p[, so p and v
p
=
p
([p[) =
p
|p|
d
p
(|p|)
d|p|
are parallel. Let us
40
dene m

as the constant of proportionality at the fermi surface


v
p
F
= p
F
/m

(75)
Using m

it is useful to dene the density of states at the


fermi surface. Recall, that in the non-interacting system,
D(E
F
) =
1
2
2
_
_
2m
h
2
_
_
3/2
E
1/2
F
=
mp
F
h
3
(76)
where p = hk, and E = p
2
/2m. Thus, for the interacting
system at the Fermi surface
D
interacting
(E
F
) =
m

p
F
h
3
, (77)
where the m

(generally > m, but not always) accounts for the


fact that the quasiparticle may be viewed as a dressed particle,
and must drag this dressing along with it. I.e., the eective
mass to some extent accounts for the interaction between the
particles.
41
4 Interactions between Particles: Landau Fermi
Liquid
4.1 The free energy, and interparticle interactions
The thermodynamics of the system depends upon the free en-
ergy F, which at zero temperature is
F F
0
= E E
0
(N N
0
) . (78)
Since our quasiparticles are formed by adiabatically switching
on the interaction in the N + 1 particle ideal system, adding
one quasiparticle to the system adds one real particle. Thus,
N N
0
=

p
n
p
, (79)
and since
E E
0

p
n
p
, (80)
we get
F F
0

p
(
p
) n
p
. (81)
As shown in Fig. 18, we will be interested in excitations of the
system which distort the Fermi surface by an amount propor-
tional to . For our theory/expansion to remain valid, we must
42

Figure 18: We consider small distortions of the fermi surface, proportional to , so


that
1
N

p
|n
p
| 1.
have
1
N

p
[n
p
[ 1 . (82)
Where n
p
,= 0,
p
will also be of order . Thus,

p
(
p
) n
p
O(
2
) , (83)
so, to be consistent we must add the next term in the Taylor
series expansion of the energy to the expression for the free
energy.
F F
0
=

p
(
p
) n
p
+
1
2

p,p

f
p,p
n
p
n
p
+ O(
3
) (84)
43
where
f
p,p
=
E
n
p
n
p

(85)
The term, proportional to f
p,p
, was added (to the Sommerfeld
theory) by L.D. Landau. Since each sum over p is proportional
to the volume V , as is F, it must be that f
p,p
1/V . However,
it is also clear that f
p,p
is an interaction between quasiparticles,
each of which is spread out over the whole volume V , so the
probability that they will interact is r
3
TF
/V , thus
f
p,p
r
3
TF
/V (86)
In general, since n
p
is only of order one near the Fermi
surface, we will only care about f
p,p
on the Fermi surface (as-
suming that it is continuous and changes slowly as we cross the
Fermi surface.
Interested in f
p,p

p
=
p

=
in only! (87)
Given this, we can reduce the spin dependence of f
p,p
to
a symmetric and anti symmetric part. First in the absence of
an external eld, the system should be invariant under time-
44
reversal, so
f
p,p

= f
p,p

, (88)
and, in a system with reection symmetry
f
p,p

= f
p,p

. (89)
Then
f
p,p

= f
p,p

. (90)
It must be then that f depends only upon the relative orienta-
tions of the spins and

, so there are only two independent


components f
p,p

and f
p,p

. We can split these into sym-


metric and antisymmetric parts.
f
a
p,p
=
1
2
_
f
p,p

f
p,p

_
f
s
p,p
=
1
2
_
f
p,p

+ f
p,p

_
.
(91)
f
a
p,p
may be interpreted as an exchange interaction, or
f
p,p

= f
s
p,p
+

f
a
p,p
(92)
where and

are the Pauli matrices for the spins.


Our ideal system is isotropic in momentum. Thus, f
a
p,p
and
f
s
p,p
will only depend upon the angle between p and p

, and
45
so we may expand either f
a
p,p
and f
s
p,p

p,p
=

l=0
f

l
P
l
(cos ) . (93)
Conventionally these f parameters are expressed in terms of
reduced units.
D(E
F
)f

l
=
V m

p
F

2
h
3
f

l
= F

l
. (94)
4.2 Local Energy of a Quasiparticle
p
Figure 19: The addition of another particle to a homogeneous system will yeilds in
forces on the quasiparticle which tend to restore equilibrium.
Now consider an interacting system with a certain distribu-
tion of excited quasiparticles n
p
. To this, add another quasi-
particle of momentum p (n

p
n

p
+
p,p
). From Eq. 84 the
46
free energy of the additional quasiparticle is

p
=
p
+

p

f
p

,p
n
p
, (95)
(recall that f
p,p
= f
p

,p
). Both terms here are O(). The
second term describes the free energy of a quasiparticle due to
the other quasiparticles in the system (some sort of Hartree-like
term).
The term
p
plays the part of the local energy of a quasi-
particle. For example, the gradient of
p
is the force the system
exerts on the additional quasiparticle. When the quasiparticle
is added to the system, the system is inhomogeneous so that
n
p
= n
p
(r). The system will react to this inhomogeneity
by minimizing its free energy so that
r
F = 0. However, only
the additional free energy due the added particle (Eq. 95) is
inhomogeneous, and has a non-zero gradient. Thus, the system
will exert a force

r
=
r

f
p

,p
n
p
(r) (96)
on the added quasiparticle resulting from interactions with other
quasiparticles.
47
4.2.1 Equilibrium Distribution of Quasiparticles at Finite T

p
also plays an important role in the nite-temperature prop-
erties of the system. If we write
E E
0
=

p

p
n
p
+
1
2

p,p

f
p

,p
n
p
n
p
(97)
Now suppose that

p
[n
p
)[ N, as indeed it must be for
the expansion above to be valid, so that
n
p
= n
p
) + (n
p
n
p
)) (98)
where the rst term is O(), and the second O(
2
). Thus,
n
p
n
p
n
p
)n
p
) + n
p
)n
p
+ n
p
)n
p
(99)
We may use this to rewrite the energy of our interacting system
E E
0


p

p
n
p

1
2

p,p

f
p

,p
n
p
)n
p
) +

p,p

f
p

,p
n
p
)n
p


p
_
_
_
p
+

p

f
p

,p
n
p
)
_
_
_ n
p

1
2

p,p

f
p

,p
n
p
)n
p
)


p

p
)n
p

1
2

p,p

f
p

,p
n
p
)n
p
) + O(
4
) (100)
At this point, we may repeat the arguments made earlier to de-
termine the fermion occupation probability for non-interacting
48
Fermions (the constant factor on the right hand-side has no
eect). We will obtain
n
p
(T, ) =
1
1 + exp (
p
) )
, (101)
or
n
p
(T, ) =
1
1 + exp (
p
) )
(p
f
p) . (102)
However, at least for an isotropic system, this expression bears
closer investigation. Here, the molecular eld (evaluated within
k
B
T of the Fermi surface)

p

p
) =

p

f
p

,p
n
p
) (103)
must be independent of the location of p on the Fermi sur-
face (and of course, spin), and is thus constant. To see this,
reconsider the Legendre polynomial expansion discussed earlier

p

p
) =

p

f
p

,p
n
p
)


l
_
d
3
pf
l
P
l
(cos )n
p
)
f
0
_
d
3
pn
p
) = 0
(104)
49
In going from the second to the third line above, we made use of
the isotropy of the system, so that n
p
) is independent of the
angle . The evaluation in the third line, follows from particle
number conservation. Thus, to lowest order in
n
p
(T, ) =
1
1 + exp (
p
)
+ O(
3
) (105)
4.3 Eective Mass m

of Quasiparticles
This argument most closely follows that of AGD, and we will
follow their notation as closely as possible (without introducing
any new symbols). In particular, since an integration by parts
is necessary, we will use a momentum integral (as opposed to a
momentum sum) notation

p
V
_
d
3
p
(2 h)
3
. (106)
The net momentum of the volume V of quasiparticles is
P
qp
= 2V
_
d
3
p
(2 h)
3
pn
p
net quasiparticle momentum (107)
which is also the momentum of the Fermi liquid. On the other
hand since the number of particles equals the number of quasi-
particles, the quasiparticle and particle currents must also be
50
equal
J
qp
= J
p
= 2V
_
d
3
p
(2 h)
3
v
p
n
p
net quasiparticle and particle current
(108)
or, since the momentum is just the particle mass times this
current
P
p
= 2V m
_
d
3
p
(2 h)
3
v
p
n
p
net quasiparticle and particle current
(109)
where v
p
=
p

p
, is the velocity of the quasiparticle. So
_
d
3
p
(2 h)
3
pn
p
= m
_
d
3
p
(2 h)
3

p

p
n
p
(110)
Now make an arbitrary change of n
p
and recall that
p
depends
upon n
p
, so that

p
= V

_
d
3
p
(2 h)
3
f
p,p
n
p
. (111)
For Eq. 110, this means that
_
d
3
p
(2 h)
3
pn
p
= m
_
d
3
p
(2 h)
3

p

p
n
p
(112)
+mV
_
d
3
p
(2 h)
3

_
d
3
p

(2 h)
3

p
_
f
p,p
n
p

_
n
p
,
51
or integrating by parts (and renaming p p

in the last part),


we get
_
d
3
p
(2 h)
3
p
m
n
p
=
_
d
3
p
(2 h)
3

p

p
n
p
(113)
V

_
d
3
p

(2 h)
3
_
d
3
p
(2 h)
3
n
p
f
p,p

p
n
p
,
Then, since n
p
is arbitrary, it must be that the integrands
themselves are equal
p
m
=
p

p

V
_
d
3
p

(2 h)
3
f
p,p

p
n
p
(114)
The factor
p
n
p
=
p

(p

p
F
). The integral may be eval-
uated by taking advantage of the system isotropy, and setting
p parallel to the z-axis, since we mostly interested in the prop-
erties of the system on the Fermi surface we take p = p
F
, let
be the angle between p (or the z-axis) and p

, and nally note


that on the Fermi surface

p

p
[
p=p
F

= v
F
= p
F
/m

. Thus,
p
F
m
=
p
F
m

_
p
2
dpd
(2 h)
3
f
p,p

(p

p
F
) (115)
52
However, since both p and p

are restricted to the Fermi surface


p

= cos , and evaluating the integral over p, we get


1
m
=
1
m

+
V p
F
2

_
d
(2 h)
3
f
p,p

cos , (116)
where the additional factor of
1
2
compensates for the additional
spin sum. If we now sum over both spins, and

, only the
symmetric part of f survives (the sum yields 4f
s
), so
1
m
=
1
m

+
4V p
F
(2 h)
3
_
d (cos ) f
s
() cos , (117)
We now expand f in a Legendre polynomial series
f

() =

l
f

l
P
l
(cos ) , (118)
and recall that P
0
(x) = 1, P
1
(x) = x, .... that
_
1
1
dxP
n
(x)P
m
(x)dx =
2
2n + 1

nm
(119)
and nally that
D(0)f

l
=
V m

p
F

2
h
3
f

l
= F

l
, (120)
we nd that
1
m
=
1
m

+
F
s
1
3m

, (121)
53
Quantity Fermi Liquid Fermi Liquid/Fermi Gas
Specic Heat C
v
=
m

p
F
3 h
3
k
2
B
T
C
V
C
V 0
=
m

m
= 1 + F
s
1
/3
Compressibility

0
=
1+F
s
0
1+F
s
0
/3
Sound Velocity c
2
=
p
2
F
3mm

(1 + F
s
0
)
_
c
c
0
_
2
=
1+F
s
0
1+F
s
1
/3
Spin Susceptibility =
m

p
F

2
h
3

2
1+F
a
0

0
=
1+F
s
1
/3
1+F
a
0
Table 1: Fermi Liquid relations between the Landau parameters F

n
and some exper-
imentally measurable quantities. For the latter, a zero subscript indicates the value
for the non-interacting Fermi gas.
or m

/m = 1 + F
s
1
/3.
The eective mass cannot be experimentally measured di-
rectly; however, it appears in many physically relevant measur-
able quantities, including the specic heat
C
V
=
_
_
_
E/V
T
_
_
_
V N
=
1
V

p

p
n
p
. (122)
To lowest order in , we may neglect f
p,p
in both
p
and n
p
,
so
C
V
=
1
V

p
n
p
T
. (123)
Recall that the density of states D(E) =

p
(E
p
), and mak-
ing the same assumption that we made for the non-interacting
54
system, that

T
is negligible, we get,
C
V
=
1
V
_
dD()

T
1
exp ( ) + 1
. (124)
This integral is identical to the one we had to evaluate for the
non-interacting system, and yields the result
C
V
=

2
3V
k
2
B
TD(E
F
)
=
k
2
B
Tm

p
F
3 h
3
. (125)
Thus, measuring the electronic contribution to the specic heat
C
V
yields information about the eective mass m

, and hence
F
s
1
. Other measurements are related to some of the remaining
Landau parameters, as summarized in table 1.
55
Chapter 7: The Electronic Band Structure of Solids
Bloch & Slater
April 2, 2001
Contents
1 Symmetry of (r) 3
2 The nearly free Electron Approximation. 6
2.1 The Origin of Band Gaps . . . . . . . . . . . . . . . . . . . . . . . . 9
3 Tight Binding Approximation 15
4 Photo-Emission Spectroscopy 24
1
Free electrons -FLT Band Structure
E
E
f
D(E)
V(r)
E
f
E
f
metal
"heavy" metal
insulator
E
D(E)
V(r) = V
0
E
f
Figure 1: The additional eects of the lattice potential can have a profound eect on
the electronic density of states (RIGHT) compared to the free-electron result (LEFT).
In the last chapter, we ignored the lattice potential and con-
sidered the eects of a small electronic potential U. In this
chapter we will set U = 0, and consider the eects of the ion po-
tential V (r). As shown in Fig. 1, additional eects of the lattice
potential can have a profound eect on the electronic density of
2
states compared to the free-electron result, and depending on
the location of the Fermi energy, the resulting system can be
a metal, semimetal, an insulator, or a metal with an enhanced
electronic mass.
1 Symmetry of (r)
From the symmetry of the electronic potential V (r) one may
infer some of the properties of the electronic wave functions
(r).
Due to the translational symmetry of the lattice V (r) is pe-
riodic
V (r) = V (r + r
n
), r
n
= n
1
a
1
+ n
2
a
2
+ n
3
a
3
(1)
and may then be expanded in a Fourier expansion
V (r) =

G
V
G
e
iGr
, G = hg
1
+ kg
2
+ lg
3
, (2)
which, since G r
n
= 2m (m Z) guarantees V (r) =
V (r + r
n
). Given this, and letting (r) =

k
C
k
e
ikr
the
3
Schroedinger equation becomes
H(r) =
_

_
h
2
2m

2
+ V (r)
_

_ = E (3)

k
h
2
k
2
2m
C
k
e
ikr
+

G
C
k
V
G
e
i(k

+G)r
= E

k
C
k
e
ikr
, k

kG
(4)
or

k
e
ikr
_

_
_
_
_
h
2
k
2
2m
E
_
_
_ C
k
+

G
V
G
C
kG
_

_
= 0r (5)
Since this is true for any r, it must be that
_
_
_
h
2
k
2
2m
E
_
_
_ C
k
+

G
V
G
C
kG
= 0, k (6)
Thus the potential acts to couple each C
k
only with its recip-
rocal space translations C
k+G
and the problem decouples in to
N independent problems for each k in the rst BZ. Ie., each of
the N problems has a solution which is a sum over plane waves
whos wave vectors dier only by Gs. Thus the eigenvalues
may be indexed by k.
E
k
= E(k), I.e. k is still a good q.n.! (7)
We may now sum over G to get
k
with the eigenvector sum
4
X
X
X
X
X
X
First B.Z.
Figure 2: The potential acts to couple each C
k
with its reciprocal space translations
C
k+G
(i.e. x x, , and ) and the problem decouples into N indepen-
dent problems for each k in the rst BZ.
restricted to reciprocal lattice sites k, k + G, . . .

k
(r) =

G
C
kG
e
i(kG)r
=
_
_

G
C
kG
e
iGr
_
_
e
ikr
(8)

k
(r) = U
k
(r)e
ikr
, where U
k
(r) = U
k
(r + r
n
) (9)
Note that if V (r) = 0, U(r) =
1

V
. This result is called Blochs
Theorem; ie., that may be resolved into a plane wave and a
periodic function. Its consequences as follows:

k+G
(r) =

G

C
k+GG
e
i(G

kG)r
=
_
_

C
kG
e
iG

r
_
_
e
ikr
=
k
(r), where G

G (10)
5
Ie.,
k+G
(r) =
k
(r) and as a result
H
k
= E(k)
k
H
k+G
= E(k + G)
k+G
(11)
= H
k
= E(k + G)
k+G
(12)
Thus E(k + G) = E(k) : E(k) is periodic then since both

k
(r) and E(k) are periodic in reciprocal space, one only needs
knowledge of them in the rst BZ to know them everywhere.
2 The nearly free Electron Approximation.
If the potential is weak, V
G
0 G, then we may solve the
V
G
= 0 problem, subject to our constraints of periodicity, and
treat V
G
as a perturbation.
When V
G
= 0, then
E(k) =
h
2
k
2
2m
free electron (13)
However, we must also have that (if V
G
,= 0)
E(k) = E(k + G)
h
2
2m
[k + G[
2
(14)
Ie., the possible electron states are not restricted to a single
parabola, but can be found equally well on paraboli shifted by
6
First BZ 2/a
E
Figure 3: For small V
G
, we may approximate the band structure as composed of N
parabolic bands. Of course, it is sucient to consider this in the rst Brillouin zone,
where the parabola centered at nite G cross at high energies. To understand the
eects of the perturbation V
G
consider this special k at the edge of the BZ. where the
paraboli cross.
any G vector. In 1-d Since E(k) = E(k + G), it is sucient
to represent this in the rst zone only. For example in a 3-D
cubic lattice the energy band structure along k
x
(k
y
= k
z
= 0)
is already rather complicated within the rst zone. (See Fig.4.)
The eect of V
G
can now be discussed. Lets return to the 1-d
problem and consider the edges of the zone where the [paraboli
intersect. (See Fig. 3.) An electron state with k =

a
will
involve at least the two G values G = 0,
2
a
. Of course, the
7
a a k
x
First B.Z.
a -a
Figure 4: The situation becomes more complicated in three dimensions since there are
many more bands and so they can cross the rst zone at lower energies. For example
in a 3-D cubic lattice the energy band structure along k
x
(k
y
= k
z
= 0) is already
rather complicated within the rst zone.
exact solution must involve all G since
_
_
_
h
2
k
2
2m
E
k
_
_
_ C
k
+

G
V
G
C
kG
= 0 (15)
We can generally take V
0
= 0 since this just sets a zero for the
potential. Then, those G for which E
k
= E
kG

h
2
k
2
2m
are
going to give the largest contribution since
C
k
=

G
V
G
C
kG
h
2
k
2
2m
E
kG
(16)
C
k
V
G
1
C
kG
1
h
2
k
2
2m
E
kG
1
(17)
8
C
kG
1
=

G
V
G
C
kG
1
G
h
2
k
2
2m
E
kGG
1
(18)
C
kG
1
V
G
1
C
k
h
2
k
2
2m
E
k
(19)
Thus to a rst approximation, we may neglect the other C
kG
,
and since V
G
= V
G
(so that V (r) is real) [C
k
[ [C
kG
1
[
other G
kG

k
(r) =

G
C
kG
e
i(kG)r

_
(e
iGx/2
+ e
iGx/2
) cos
x
a
(e
iGx/2
e
iGx/2
) sin
x
a
(20)
The corresponding electron densities are sketched in Fig. 5.
Clearly
+
has higher density near the ionic cores, and will
be more tightly bound, thus E
+
< E

. Thus a gap opens in


E
k
near k =
G
2
.
2.1 The Origin of Band Gaps
Now lets reexamine this gap at k = G
1
/2 in a quantitative
manner. Start with the eigen value equation shifted by G.
C
kG
_
_
_E
k

h
2
2m
[k G[
2
_
_
_ =

G

V
G
C
kGG
=

G

V
G

G
C
kG

(21)
9
V(x)
(x)

(x)
+
E
k
Gap!
E
D(E)
Figure 5:
+
cos
2
(x/a) has higher density near the ionic cores, and will be more
tightly bound, thus E
+
< E

. Thus a gap opens in E


k
near k =
G
2
.
C
kG
=

G
V
G

G
C
kG

_
E
k

h
2
2m
[k G[
2
_
(22)
To a rst approximation (V
G
0) lets set E =
h
2
k
2
2m
(a free-
electron energy) and ignore all but the largest C
kG
; ie., those
for which the denominator vanishes.
k
2
= [k G[
2
, (23)
10
or in 1-d
k
2
= (k
2
a
)
2
or k =

a
(24)
This is just the Laue condition, which was shown to be equiv-
alent to the Bragg condition. Ie., the strongest perturbation
to the free-electron picture occurs for states with energies at
the edge of the rst B.Z. Thus the equation above also tells us
highly perturbed
essentially free electrons
Figure 6: We can satisfy the condition E
k
E
kG
only for k on the edge of the B.Z..
Here the lattice potential strongly perturbs the electronic states (i.e. more than one
C
kG
is nite).
that C
k
and C
kG
1
are the most important coecients (if this
electronic state was unperturbed, only C
k
would be important).
Thus approximately for V
G
0, V
0
0 and for k near the
11
zone boundary
G = 0 C
k
_

_
E
h
2
k
2
2m
_

_
= V
G
1
C
kG
1
(25)
G = G
1
C
kG
1
_

_
E
h
2
[k G
1
[
2
2m
_

_
= V
G
1
C
k
, (26)
Again, ignore all other C
G
. This is a secular equation which
has a nontrivial solution i

_
h
2
k
2
2m
E
_
V
G
1
V
G
1
_
h
2
|kG
1
|
2
2m
E
_

= 0 (27)
or

E
0
k
E V
G
1
V
G
1
E
0
kG
1
E

= 0 (28)
(V
G
= V

G
, so thatV (r) 1)
(E
0
k
E)(E
0
kG
1
E) [V
G
1
[
2
= 0 (29)
E
0
k
E
0
kG
1
E
_
E
0
k
+ E
0
kG
1
_
+ E
2
[V
G
1
[
2
= 0 (30)
E

=
1
2
_
E
0
kG
1
+ E
0
k
_

_
_
_
1
4
_
E
0
kG
E
0
k
_
2
+ [V
G
1
[
2
_
_
_
1
2
(31)
At the zone boundary, where E
0
kG
1
= E
0
k
, the gap is
E = E
+
E

= 2[V
G
1
[ (32)
12
-/a
/a
0
k
2 V
G
E
k
k
e

Figure 7:
And the band structure looks something like Fig. 7. Within this
approximation, the gap, or forbidden regions in which there are
no electronic states arise when the Bragg condition (k
f
k
0
=
G) is satised.
[ k[ [k + G[ (33)
The interpretation is clear: the high degree of back scattering
for these k-values destroys the electronic states.
Thus, by treating the lattice potential as a perturbation to
the free electron problem, we see that gaps arise due to enhanced
electron-lattice back scattering for k near the zone edge. How-
13
ever, in chapter one, we considered band structure qualitatively
and determined that gaps could arise from perturbing about
the atomic limit. This in fact, is another natural way of con-
Separation
State
Energies
Figure 8: Band gaps in the electronic DOS naturally emerge when perturbing around
the atomic limit. As we bring more atoms together (left) or bring the atoms in the
lattice closer together (right), bands form from mixing of the orbital states. If the
band broadening is small enough, gaps remain between the bands.
structing a band structure theory. It is called the tight-binding
approximation.
14
r r
E
i i
V (r)
A
Valence electrons Atomic cores
Figure 9: In the tight-binding approximation, we generally ignore the core electron
dynamics and consider only the ionic core potential. For now lets assume that there
is only one valence orbital
i
on each atom.
3 Tight Binding Approximation
In the tight-binding approximation, we generally ignore the core
electron dynamics and treat consider only the ionic core poten-
tial. For now lets assume that there is only one valence orbital

i
on each atom. We will also assume that the atomic prob-
lem is solved, and perturb around this solution. The atomic
problem has valence eigenstates
i
, and eigen energies E
i
. The
unperturbed Schroedinger equation for the nth atom is
H
A
(r r
n
)
i
(r r
n
) = E
i

i
(r r
n
) (34)
15
There is a weak perturbation v(rr
n
) coming from the atomic
potentials of the other atoms r
m
,= r
n
H = H
A
+ v =
h
2

2
2m
+ V
A
(r r
n
) + v(r r
n
) (35)
v(r r
n
) =

m=n
V
A
(r r
m
) (36)
We now seek solutions of the Schroedinger equation indexed by
k (Blochs theorem)
H
k
(r) = E(k)
k
(r) (37)

E(k) =

k
[H[
k
)

k
[
k
)
(38)
where

k
[
k
)
_
d
3
r

k
(r)
k
(r)

k
[H[
k
)
_
d
3
r

k
(r)H
k
(r) (39)
Of course, this problem is almost hopelessly complicated. We
cannot solve for
k
. Rather, we will solve for some
k

k
where the parameters of
k
are determined by minimizing

k
[H[
k
)

k
[
k
)
E(k). (40)
16
This is called the Raleigh-Ritz variational principle.
Consistent with our original motivation, we will approximate

k
with a sum over atomic states.

k

k
=

n
a
n

i
(r r
n
) =

n
e
ikr
n

i
(r r
n
) (41)

k
(r) = U
k
(r)e
ikr
,
k
(r) =
k+G
(r)
Where
k
must be a Bloch state
k+G
=
k
which dictates our
choice a
n
= e
ikr
n
. Thus at this level of approximation we have
no free parameters to vary to minimize
k
[H[
k
) /
k
[
k
)
E(k).
Using
k
as an approximate state the energy denominator

k
[
k
), becomes

k
[
k
) =

n,m
e
ik(r
n
r
m
)
_
d
3
r

i
(r r
m
)
i
(r r
n
) (42)
Lets imagine that the valance orbital of interest,
i
, has an
very small overlap with adjacent atoms so that

k
[
k
)

n
_
d
3
r

i
(r r
n
)
i
(r r
n
) = N (43)
The last identity follows since
i
is normalized.
17
(r-r )
i 1 (r-r )
i 2
Figure 10: In the tight binding approximation, we assume that the atomic orbitals
of adjacent sites have a very small overlap with each other.
The energy for our approximate wave function is then
E(k)
1
N

n,m
e
ik(r
n
r
m
)
_
d
3
r

i
(rr
m
) E
i
+ v(r r
n
)
i
(rr
n
) .
(44)
Again, in the rst part (involving E
i
), we may neglect orbital
overlap. For the second term, involving v(r r
n
), the over-
lap should be included, but only to the nearest neighbors of
each atom (why?). In the simplest case, where the orbitals
i
,
are s-orbitals, then we can use this symmetry to reduce the
complexity of the problem to just two more integrals since the
hybridization (B
i
) will be the same in all directions.
A
i
=
_

i
(r r
n
)v(r r
n
)
i
(r r
n
)d
3
r ren. E
i
(45)
18
A
i
B
i
B
i
B
i
B
i
Figure 11: A simple cubic tight binding lattice composed of s-orbitals, with overlap
integral B
i
.
B
i
=
_

i
(r r
m
)v(r r
n
)
i
(r r
n
)d
3
r (46)
B
i
describes the hybridization of adjacent orbitals.
A
i
; B
i
> 0, since v(r r
n
) < 0 (47)
Thus
E(k) E
i
A
i
B
i

m
e
ik(r
n
r
m
)
sum over m n.n. to n
(48)
Now, if we have a cubic lattice, then
(r
n
r
m
) = (a, 0, 0)(0, a, 0)(0, 0, a) (49)
19
so
E(k) = E
i
A
i
2B
i
cos k
x
a + cos k
y
a + cos k
z
a (50)
Thus a band centered about E
i
A
i
of width 12B
i
is formed.
Near the band center, for k-vectors near the center of the zone
we can expand the cosines cos ka 1
1
2
(ka)
2
+ and let
k
2
= k
2
x
+ k
2
y
+ k
2
z
, so that
E(k) E
i
A
i
+ B
i
a
2
k
2
(51)
The electrons near the zone center act as if they were free with
a renormalized mass.
h
2
k
2
2m

= B
i
a
2
k
2
, i.e.
1
m

curvature of band (52)


For this reason, the hybridization term B
i
is often associated
with kinetic energy. This makes sense, from its origins of wave
function overlap and thus electronic transfer.
The width of the band, 12 B
i
, will increase as the electronic
overlap increases and the interatomic orbitals (core orbitals or
valance f and d orbitals) will tend to form narrow bands with
high eective masses (small B
i
).
20
First B.Z.
Fermi surface
Figure 12: Electronic states for a cubic lattice near the center of the B.Z. act like free
electrons with a renormalized mass. Hence, if the band is partially lled, the Fermi
surface will be spherical.
The bands are lled then by placing two electrons in each
band state ( with spins up and down). A metal then forms
when the valence band is partially full. I.e., for Na with a
1s
2
2s
2
2p
6
3s
1
atomic conguration the 1s, 2s and 2p orbitals
evolve into (narrow) lled bands, but the 3s
1
band will only
be half full, and thus it evolves into a metal. Mg 1s
2
2s
2
2p
6
3s
2
also metal since the p and s band overlaps the unlled d-band.
There are exceptions to this rule. Consider C with atomic con-
guration of 1s
2
2s
2
2p
2
. Its valance s and p states form a strong
sp
3
hybrid band which is further split into a bonding and anti-
21
a
E
E
1
E
1
E
2
2
E
1
A
2
A
2
12 B
1
12 B
k = 0
- /a /a
k
k
k
k
111
111
111
Atomic Potential
Tight Binding Bands
Figure 13: In the tight-binding approximation, band form from overlapping orbitals
states (states of the atomic potential). The bandwidth is proportional to the hy-
bridization B (12B for a SC lattice). More localized, compact, atomic states tend to
form narrower bands.
bonding band. (See Fig.14). Here, the gap is not tied to the
periodicity of the lattice, and so an amorphous material of C
may also display a gap.
The tight-binding picture can also explain the variety of fea-
tures seen in the DOS of real materials. For example, in Cu
(Ar)3d
10
4s the d-orbitals are rather small whereas the valence
s-orbitals have a large extent . As a result the s-s hybridization
22
a
P
sp antibonding
3
sp bonding
3
S
a
E
r
Figure 14: C (diamond) with atomic conguration of 1s
2
2s
2
2p
2
. Its valance s and p
states form a strong sp
3
hybrid band which is split into a bonding and anti-bonding
band.
B
ss
i
: is strong and the B
dd
i
is weak.
B
dd
i
B
ss
i
(53)
In addition the s-d hybridization is inhibited by the opposing
symmetry of the s-d orbitals.
B
sd
i
=
_

s
i
(r r
1
)v(r r
2
)
d
i
(r r
2
)d
3
r B
ss
i
(54)
where
s
i
is essentially even and
d
i
is essentially odd. So B
sd
i

B
ss
i
. Thus, to a rst approximation the s-orbitals will form a
very wide band of mostly s-character and the d-orbitals will
form a very narrow band of mostly d-character. Since both the
23
+ +

+
1
2
+ +

+ +

d
s
D(E)
Figure 15: Schematic DOS of Cu 3d
10
4s
1
. The narrow d-band feature is split due to
crystal elds.
s and d bands are valance, they will overlap leading to a DOS
with both d and s features superimposed.
4 Photo-Emission Spectroscopy
The electronic density of electronic states (especially for oc-
cupied states), and to a less extent band structure, are very
important for illuminating the interesting physics of materials.
As we saw in Chap. 6, an enhanced DOS at the Fermi sur-
face indicates an enhanced electronic mass, and if D(E
F
) = 0,
we have an insulator (semiconductor). The eective electronic
24
h
d
detector
synchrotron
r
material
sample
V
e
Figure 16: XPS Experiment: By varying the voltage one may select the kinetic
energy of the electrons reaching the counting detector.
mass also varies inversely with the curvature of the bands. The
density of states away from the Fermi surface can allow us to
predict the properties of the material upon doping, or it can
yield information about core-level states. Thus it is important
to be able to measure D(E). This may be done by x-ray pho-
toemission (XPS), UPS or PS in general. The band dispersion
E(k) may also be measured using angle-resolved photoemission
(ARPES) where angle between the incident radiation and the
25
detector is also measured.
The basic idea is that a photon (usually an x-ray) is used
to knock an electron out of the system (See gure 17.) Of
E
h - E
b
h - E -
b
D(E)
Intensity(E)
Kinetic Energy of electrons
h -
X-ray
E
b
Figure 17: Let the binding energy be dened so that E
b
> 0, = work function, then
the detected electron intensity I(E
kin
h ) D(E
b
)f(E
b
)
course, in order for an electron at an energy of E
b
below the
Fermi surface to escape the material, the incident photon must
have an energy which exceeds E
b
and the work function of
the material. If h > , then the emitted electrons will have
a distribution of kinetic energies E
kin
, extending from zero to
26
h. From Fermis golden rule, we know that the probability
per unit time of an electron being ejected is proportional to
the density of occupied electronic states times the probability
(Fermi function) that the electronic state is occupied
I(E
kin
) =
1
(E
kin
)
D(E
b
)f(E
b
)
D(E
kin
+ h)f(E
kin
+ h) (55)
Thus if we measure the energy and number of ejected particles,
then we know D(E
b
).
phonon Coulombic
interation
Secondary electrons
e
e
Figure 18: Left: Origin of the background in I(E
kin
. Right: Electrons excited deep
within the bulk scatter so often that they rarely escape. Thus, most of the signal I
originates at the surface, which must be clean and representative of the bulk.
There are several problems with this procedure. First some
27
of the photon excited particles will scatter o phonons and elec-
tronic excitations within the material. Since these processes can
occur over a very wide range of energies, they will produce a
broad featureless background in N(E
k
). Second, due to these
0 1 2 3 4
E
kin
0
2
4
6
I
(
E
k
i
n
)
background subtracted
background
raw data
h-
Figure 19: In Photoemission, we measure the rate of ejected electrons as a function
of their kinetic energy. The raw data contains a background. Once this is subtracted
o, the subtracted data is proportional to the electronic density of states convolved
with a Fermi function I(E
kin
) D(E
kin
+ h)f(E
kin
+ h).
secondary scattering processes, it is very unlikely that an elec-
tron which is excited deep within the bulk, will ever escape from
28
measure
h
E
kin
e
Figure 20: BIS E
kin
= h E
b
, E
b
= h E
kin

the material. Thus, we only learn about D(E) near the surface
of the material. Therefore it is important for this surface to be
clean so that it is representative of the bulk. For this reason
these experiments are often carried out in ultra-high vacuum
conditions.
We can also learn about the electronic states D(E) above the
Fermi surface, E > F
F
, using Inverse Photoemmision. Here,
an electron beam is focussed on the surface and the outgoing
us of photons are measured.
29
Chapter 8: Magnetism
Holstein & Primako
April 13, 2001
Contents
1 Introduction 2
1.1 The Relevance of Magnetostatics . . . . . . . . . . . . . . . . . . . . 3
1.2 Non-interacting Magnetic Systems . . . . . . . . . . . . . . . . . . . . 4
2 Coulombic Correlation Eects 7
2.1 Moment Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Magnetism and Intersite Correlations . . . . . . . . . . . . . . . . . 11
2.2.1 The Exchange Interaction Between Localized Spins . . . . . . 14
2.2.2 Exchange Interaction for Delocalized Spins . . . . . . . . . . . 18
3 Band Model of Ferromagnetism 22
3.1 Enhancement of . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 Finite T Behavior of a Band Ferromagnet . . . . . . . . . . . . . . . 28
3.2.1 Eect of B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4 Mean-Field Theory of Magnetism 32
4.1 Ferromagnetism for localized electrons (MFT) . . . . . . . . . . . . . 32
4.2 Mean-Field Theory of Antiferromagnets . . . . . . . . . . . . . . . . 37
1
5 Spin Waves 43
5.1 Quantization of Ferromagnetic Spin Waves . . . . . . . . . . . . . . . 49
5.2 Antiferromagnetic Spin Waves . . . . . . . . . . . . . . . . . . . . . . 56
2
1 Introduction
Magnetism is one of the most interesting subjects in condensed
matter physics. Magnetic eects are responsible for heavy fermion
behavior, ferromagnetism, antiferromagnetism, ferrimagnetism
and probably high temperature superconductivity.
Unlike our previous studies, most magnetic systems are not
well described by simple models which ignore intersite correla-
tions. The reason is simple: magnetism is inherently due to
electronic correlations of moments on dierent sites. As we
J
Figure 1: Both moment formulation and the correlation between these moments (J)
are due to Coulombic eects
will see, systems without these inter-spin correlations (or those
without well dened moments to be correlated) have uninter-
esting and unimportant (energetically) magnetic properties.
3
1.1 The Relevance of Magnetostatics
Perhaps the term magnetism is a misnomer, or rather describes
only the external probe (B) which we use to study magnetic be-
havior. Magnetic eects are due to electronic correlations, those
mediated or due to Coulombic eects, and not due to magnetic
correlations between moments (these are smaller by orders of
v/c, ie., they are relativistic corrections). For example, consider
the magnetic correlation between two moments separated by a
couple of Angstroms.
U =
1
r
3
[m
1
m
2
3(m
1
n)(m
2
n)] (1)
then,
U
dipoledipole

m
1
m
2
r
3
(2)
If we let,
m
1
m
2
g
B

e h
m
r 2

A
then,
U
(g
B
)
2
r
3

_
_
_
e
2
hc
_
_
_
2
_
a
0
r
_
3
e
2
a
0
10
4
eV (3)
4
Or roughly one degree Kelvin! Magnetic correlations due to
magnetic interactions would be destroyed by thermal uctua-
tions at very low temperatures.
1.2 Non-interacting Magnetic Systems
We will dene non-interacting magnetic systems as those for
which the independent moments do not interact with each other,
and only interact with the probing eld. For the moment lets
consider only the magnetic moments due to electrons (as we
will see, since they can interact with each other, they are by
far the most important moments in the system). They have a
moment

B
(L + 2S) (4)
The system energy will change by an amount
E g
B
B(L+2S)
B
B,
B
=
e h
2m
5.810
5
eV
Te
(5)
in an external eld. The largest eld which can regularly be
produced in a lab is 10Te (100Te or more can be produced
5
at LANL by blowing things up), thus
E
<

10
3
eV or k
B
(10
o
K) (6)
This is a very small energy. Thus magnetic eects are wiped
out by thermal uctuation for
k
B
T >
B
B (7)
at about 10 K! Thus experiments which measure the magnetism
of non-interacting systems must be carried out at low temper-
atures. These experiments typically measure the susceptibil-
ity of the system with a Faraday balance or a magnetometer
(SQUID).
For a collection of isolated moments (spin 1/2), the suscep-
tibility may be calculated from the moment
s =
1
2
m) = g
B
1
2
_
e

1
2
g
B
B
e

1
2
g
B
B
_
_
e

1
2
g
B
B
+ e

1
2
g
B
B
_
(8)
g

B
2
tanh(
B
B)
B

B
B
T
(g 2)
=
m)
B


2
B
k
B
T
(9)
6
Once again, the energy of the moment-eld interaction is roughly
E

2
B
T
B
2

10
8
_
eV
Te
_
2
T
_
10
4
eV

K
_
2
B
2
, (k = 1) (10)
When E T, thermal uctuations destroy the orientation of
the moments with the external eld. If B 10Te
E
100
T

K (11)
or E T at 10

K! However, we know that systems such as


kT/E
F
E
D(E)
Figure 2: In a metal, only the electrons near the Fermi surface, which are not paired
into singlets, contribute to the bulk susceptibility
kT
E
F

2
B
k
B
T

2
B
D(E
F
)
iron exist for which a small eld can induce a relatively large
moment at room temperature. This is surprising since for a
metal, or a free electron gas, the susceptibility is much smaller
7
than the free electron result, since only the spins near the Fermi
surface can participate, =

2
B
E
F
. Note that this is even smaller
than the free electron result by a factor of
k
B
T
E
F
1!
2 Coulombic Correlation Eects
2.1 Moment Formation
Of course real materials are not composed of free isolated elec-
trons. Nevertheless some insulators act almost as if they are
composed of non-interacting atoms (ions) with moments given
by Hunds rules: maximum S maximum L which leads to large
atomic moments. Hunds rules reect the atoms attempt to
lower its Coulombic energy, see Fig. 3. By maximizing the to-
tal spin S, the spin part of the wavefunction becomes symmetric
under electron exchange (i.e. for two electrons with s =
1
2

the maximum value of total spin is S = 1 with a wavefunc-
tion [) + [) ). Then, since the total wavefunction must
be antisymmetric under exchange, the spatial part must be an-
tisymmetric requiring it to have a node. The node keeps the
8
Max S
Max L
(x)
x
e- e-
Figure 3: Hunds rules, Maximize S and L, both result from minimizing the Coulomb
energy.
electrons apart, minimizing their Coulomb energy. The second
Hunds rule is also due to Coulombic interactions. Maximizing
L tends to keep the electrons apart, much like a centrifuge. (al-
ternatively, the radial Schroedinger equation obtains an angular
momentum barrier L(L + 1)/r
2
).

B
U
Figure 4: A simple tight-binding model with a local Coulomb repulsion U. If U = 0,
the rate that electrons hop on and o any site may be approximated using Fermis
golden rule [B[
2
D(E
F
)
1
t
. Then by the uncertainty principal Et h so
each site energy acquires an uncertainty or width E
h
t
[B[
2
D(E
F
) .
The sites will form moments (see Fig. 5) if U, [[
To illustrate how band formation modies this scenerio, lets
9
consider a simple tight binding model (See Fig. 4). By Fermis
goldon rule, each level acquires a width (uncertainty in its en-
ergy) = B
2
D(E
F
) and each level can be in one of the four
states shown in Fig. 5. Clearly, the states _ and
moment forms

2+U
Figure 5: A moment forms on an orbital provided that U, [[. U
e
2
a
e

a
r
TF
r
TF
is small for a metal, large for an insulator
do not have a moment, and the states and do. If
these states mix equally a moment will not form. The mixing
between the states with moments is only through one of the
other two states (_ or ) and may be suppressed,
as can the occupancy of the moment-less states, by increasing
the energy of the states without moments. Ie., a moment will
form on each site if and + U .
10
n
m
L
I
e
B

Figure 6: Here =
1
c

t
, m =
B
L and = Ba
2
In this limit U B, the system will act more like a system
of free moments than a free electron gas. Thus, one might
expect

insulator

metal
(12)
for noninteracting systems. However, this is not the case. The
reason is that I have only told you half of the story. A real
atom, or a system composed of such atoms, has a diamagnetic
response due to the angular momentum L of the rotating elec-
trons. This eect is due to Lenz Law. So that any introduced
magnetic induction will induce an EMF and hence a current
that opposes the electron current which reduces the moment.
11
diamagnetism with a
2
. In the free electron limit (see
J.M. Ziman, )
=
2
B
D(E
F
)
_
_
1
1
3
_
m
m

_
2
_
_
(13)
E 10
8
_
_
eV
T
_
_
2
a10
1
eV
1
B
2
10
9
eV B
2
(T) (14)
For insulators, often with
m

m
< 1 the diamagnetism wins;
whereas for metals, generally with
m

m
1, the Pauli para-
magnetism wins.
2.2 Magnetism and Intersite Correlations
From both Hunds rules and a simple tight binding picture,
we argued that moment formation in solids results from local
Coulomb correlations between electrons. We also saw that a
collection of such isolated moments is rather boring since all
magnetic behavior is washed out by thermal uctuations at very
low temperatures. Consider once again an isolated moment of
magnitude m
B
in an external eld.

(m
B
)
2
k
B
T
(15)
12
a

Figure 7: If the magnetic moments in a small volume are correlated, then the magnetic
susceptibility is strongly enhanced.
E
(m
B
B)
2
k
B
T
(16)
For magnetism to be signicant at room temperature (300K)
we must increase the energy of our system in a eld. This
may be accomplished by increasing the eective moment m by
correlating adjacent moments. If the range of this correlation is
, so that roughly
4
3
3a
3
moments are correlated, then let

a
= 3
so that 10
2
moments are correlated and m 10
2
. This
increases E by about 10
4
, so that E k
b
T at T 10
3
K.
The observed (measured) susceptibility also then increases by
about 10
4
, all by only correlating moments in a range of 3 lattice
13
spacings.
Clearly correlations between adjacent spins can make mag-
netism in materials relevant. Such correlations are due to elec-
tronic eects and are hence usually short ranged due to elec-
tronic (Thomas-Fermi) screening. If we consider two s =
1
2
spins,
1

2
, then the correlation is usually parameterized by
the Heisenberg exchange Hamiltonian, or
H = 2J
1

2
(17)
where J is the exchange splitting between the singlet and triplet
energies.
_

_
[ )
[ ) + [ )
[ )
_

_
E
t
(18)
[ ) [ ) E
s
(19)
E
t
E
s
= J (20)
The trick then is to calculate J!
14
1
2
A
r
r
r
r
r
R
e
e
e
B
AB
2B
1B
2A
12
1A
+
e
+
Figure 8: Geometry of two electrons, 1 and 2, bound to two ions A and B.
2.2.1 The Exchange Interaction Between Localized Spins
Imagine that we have two hydrogen atoms A and B which lo-
calize two electrons 1 and 2. As these two electrons approach,
their spins will become correlated.
H = H
1
+ H
2
+ H
12
(21)
H
1
=
h
2
2m

e
2
r
1A

e
2
r
1B
(22)
H
12
=
e
2
r
12
+
e
2
R
AB
(23)
As we did in Chap. 1 to describe binding, we will use the atomic
wave functions to approximate the molecular wavefunction
12
.

12
= (
A
(1) +
B
(1)) (
A
(2) +
B
(2)) spin part
15
= (
A
(1)
A
(2) +
B
(1)
B
(2) +
A
(1)
B
(2) +
A
(2)
B
(1))
spin part (24)
If
e
2
r
12
is strong (it is) then the rst two states with both electrons
on the same ion are suppressed, especially if the ions are far
apart. Thus we neglect them, and make the Heitler-London
approximation; for example

12
(
A
(1)
B
(2) +
B
(1)
A
(2)) spin singlet (25)
The spatial wave function is symmetric, and thus appropriate
for the spin singlet state since the total electronic wave function
must be antisymmetric. For the symmetric spin triplet states,
the electronic wave function is

12
= (
A
(1)
B
(2)
B
(1)
A
(2)) spin triplet (26)
or

12
=
A
(1)
B
(2)
B
(1)
A
(2) spin part (27)
The energy of these states may then be calculated by evaluating

12
[H[
12
)

12
[
12
)
.
E =

12
[H[
12
)

12
[
12
)
= 2E
I
+
C A
1 S
, + singlet , triplet
(28)
16
where
E
I
=
_
d
2
r
1

A
(1)
_

h
2
2m

2
1

e
2
r
1A
_

A
(1) < 0 (29)
the Coulomb integral
C = e
2
_
d
3
r
1
d
3
r
2
_
_
_
1
R
AB
+
1
r
12

1
r
2A

1
r
1B
_
_
_
[
A
(1)[
2
[
B
(2)[
2
< 0
(30)
the exchange integral
A = e
2
_
d
3
r
1
d
3
r
2
_
_
_
1
R
AB
+
1
r
12

1
r
2A

1
r
1B
_
_
_

A
(1)
A
(2)
B
(1)

B
(2)
(31)
and nally, the overlap integral is
S =
_
d
3
r
1
d
3
r
2

A
(1)
A
(2)
B
(1)

B
(2) (0 < S < 1) (32)
All E
I
, C, A, S 1. So
J = E
t
E
s
= 2E
I
+
C A
1 S

_
_
_
2E
I
+
C + A
1 + S
_
_
_
J =
C A
1 S

C + A
1 + S
> 0
J = 2
A SC
1 S
2
< 0 (33)
where the inequality follows since the last two terms in the
dominate the integral for A and in the Heitler-London approx-
17
Figure 9:
imation S 1. Or, for the eective Hamiltonian.
H = 2J
1

2
, J < 0 (34)
Clearly this favors an antiparallel or antiferromagnetic align-
ment of the spins (See Fig. 9) since then (classically)
1

2
< 0
and E < 0, so minimizing the energy. This type of interaction
is clearly appropriate for insulators which may be approximate
as a collection of isolated atoms. Indeed antiferromagnets are
generally insulators for this and other reasons.
H = 2J

ij)

i

j
, J < 0 (35)
18
2.2.2 Exchange Interaction for Delocalized Spins
Ferromagnetism, where adjacent spins tend to align forming a
bulk magnetic moment, is most often seen in conducting metals
such as Fe. As we will see in this section, the Pauli principle,
the Coulomb interactions, and the itinerancy of free (metallic)
electrons favors a ferromagnetic (J > 0) exchange interaction.
Consider two like-spin (triplet) free electrons in a volume V
(See Fig. 10). If we describe the spatial part of their wave
e
e
r
i
j
r
V
Figure 10: [ ) triplet-symmetric
function with plane waves, then

ij
=
1

2V
_
e
ik
i
r
i
e
ik
j
r
j
e
ik
i
r
j
e
ik
j
r
i
_
=
1

2V
e
ik
i
r
i
e
ik
j
r
j
_
1 e
i(k
i
k
j
)(r
i
r
j
)
_
(36)
19
The probability that the electrons are in volumes d
3
r
i
and d
3
r
j
is
[
ij
[
2
d
3
r
i
d
3
r
j
=
1
V
2
1 cos [(k
i
k
j
) (r
i
r
j
)] d
3
r
i
d
3
r
j
(37)
As required by the Pauli principle, this probability vanishes
when r
i
= r
j
. This would not be the case for electrons in the
singlet spin state (if the coulomb interaction continues to be
ignored). Thus there is a hole, called the exchange hole, in
the probability density for r
i
r
j
for triplet spin electrons, but
not singlet spin ones.
Now consider the eects of the electron-ion and the electron-
electron coulomb interactions (See Figure 11). If one electron
comes near an ion, it will screen the potential of that ion seen
by other electrons; thereby raising their energy. Thus the ef-
fect of allowing electrons to approach each other, is to increase
the electron-ion coulomb energy, and of course the electron-
electron Coulomb energy. Thus, anything which keeps them
apart without an energy cost, like the exchange hole for triplet
spin electrons, will reduce their energy. As a result, like-spin
20
e
e
Ze
Ze
a
b
+
+
Figure 11: Electron a screens the potential seen by electron b, raising its energy. Any-
thing which keeps pairs of electrons apart, but costs no energy like the exchange hole
for the electronic triplet, will lower the energy of the system. Thus, triplet formation
is favored thermodynamically.
electrons have lower energy and are thermodynamically favored
Ferromagnetism.
To determine the range of this FM exchange interaction, we
must average the eect over the Fermi sea. If one of the electrons
is xed at the origin (See Figure 12), then the probability that
a second is located a distance r away, in a volume element d
3
r
is
P

(r)d
3
r = n

d
3
r (1 cos [(k
i
k
j
) r])
. .
(38)
Fermi sea average
21
e
e
r=
O
r
j
i.e. r 0
i
Figure 12: Geometry to calculate the exchange interaction.
n

=
1
2
n =
1
2
# electrons
volume
(39)
In terms of an electronic charge density, this is

ex
(r) =
en
2
(1 cos [(k
i
k
j
) r])
=
en
2
_

_
1
1
(
4
3
k
3
F
)
2
_
k
F
o
d
3
k
i
d
3
k
j
1
2
_
e
(k
i
k
j
)r
+ e
(k
i
k
j
)r
_
_

_
=
en
2
_
_
_
1 (
4
3
k
3
F
)
2
_
k
F
0
d
3
k
i
e
k
i
r
_
k
F
0
d
3
k
j
e
k
j
r
_
_
_
=
en
2
_

_
1 9
(sin k
F
r k
F
r cos k
F
r)
2
(k
F
r)
6
_

_
(40)
Note that both of the exponential terms in the second line are
the same, since we integrate over all k
i
k
i
& k
j
k
j
.
Since we have only been considering Pauli-principle eects, the
electronic density of spin down electrons remains unchanged.
Thus, the total charge density around the up spin electron xed
22
1/2
2
1
4

en
eff
k r
F
Figure 13: Electron density near an electron xed at the origin. Coulomb eects would
reduce the density for small r further, but would not signicantly eect the size of the
exchange hole or the range of the corresponding potential, both 1/k
F
.
at the origin is

eff
(r) = en
_

_
1
9
2
(sin k
F
r k
F
r cos k
F
r)
2
(k
F
r)
6
_

_
(41)
The size of the exchange hole, and the range of the correspond-
ing ferromagnetic exchange potential, is
1
k
F
a which is
rather short.
3 Band Model of Ferromagnetism
Due to the short range of this potential, its Fourier transform is
essentially at in k. This fact may be used to construct a band
23
theory of FM where the mean eect of a spin-up electron is to
lower the energy of all other band states of spin up electrons by
a small amount, independent of k.
E

(k) = E(k)
IN

N
; I
<

1eV (42)
Likewise for spin down
E

(k) = E(k)
IN

N
. (43)
Where I, the stoner parameter, quanties the exchange hole
energy. The relative spin occupation R is related to the bulk
moment
R =
(N

)
N
, M =
B
_
_
N
V
_
_
R (44)
Then
E

(k) = E(k)
I(N

+N

)
2N

IR
2
, ( = ) (45)


E(k)
IR
2
. (46)
If R is nite and real, then we have ferromagnetism.
R =
1
N

k
1
exp
_
(

E(k) IR/2 E
F
)/k
B
T
_
+ 1

1
exp
_
(

E(k) + IR/2 E
F
)/k
B
T
_
+ 1
(47)
24
For small R, we may expand around

E(k) = E
F
.
f(x a) f(x + a) = 2af
/

2
3!
a
3
f
///
(48)
All derivatives will be evaluated at

E(k) = E
F
, so f
/
< 0 and
E
E
f
E
F
E
f
E
F
E
f f
Figure 14:
f
///
> 0. Thus,
R = 2
IR
2
1
N

k
f


E(k)

E
F

2
6
_
_
IR
2
_
_
3
1
N

3
f


E
3
(k)

E
F
(49)
This is a quadratic equation in R
1
I
N

k
f
E(k)

E
F
=
1
24
I
3
R
2
1
N

3
f
E
3
(k)

E
F
(50)
25
which has a real solution i
1
I
N

k
f
E(k)

E
F
> 0 (51)
Or, the derivative of the Fermi function summed over the BZ
must be enough to overcome the -1 and produce a positive
result. Clearly this is most likely to happen at T = 0, where
f
E(k)

E
F
(

E E
F
)
T = 0,
1
N

k
f
E
k
=
_
d

E
V
2N
D(

E)(

EE
F
) =
V
2N
D(E
F
) =

D(E
F
)
(52)
So, the condition for FM at T = 0 is I

D(E
F
) > 1. This is
known as the Stoner criterion. I is essentially at as a function
1.0
50
Z
I (eV)
Fe
Co
Ni
Na
Li
Z
1.0
D(E ) (eV )
-1
F

Figure 15:
26
of the atomic number, thus materials such as Fe, Co, & Ni with
a large

D(E
F
) are favored to be FM.
3.1 Enhancement of
Even those systems without a FM ground state have their sus-
ceptibility strongly enhanced by this mechanism. Let us re-
consider the eect of an external eld (gS = 1) on the band
energies.
E

(k) = E(k)
In

N

B
B (53)
Then
R =
1
N

k
f


E
k
(IR + 2
B
B)
=

D(E
F
)(IR + 2
B
B) (54)
or as M =
B
N
V
R, we get
M = 2
2
B
N
V

D(E
F
)
1 I

D(E
F
)
B (55)
or
=
M
B
=

0
1 I

D(E
F
)
(56)
27

0
= 2
2
B
N
V

D(E
F
) (57)
=
2
B
D(E
F
) (58)
Thus, when I

D(E
F
)
<

1, the susceptibility can be consider-


ably enhanced over the non-interacting result
0
. However, this
approximation usually overestimates since it neglects diamag-
netic contributions, and spin uctuations (at T ,= 0). As we
will see, the latter especially are important for estimating T
c
.
a

Figure 16: Spin uctuations can reduce the total moment within the correlated region,
and even reduce itself. Both eects lead to a reduction in the bulk susceptibility
moment
2
28
3.2 Finite T Behavior of a Band Ferromagnet
In principle, one could start from an ab-initio calculation of
the electronic band structure of E(k) and I, such as Ni, and
D(E) s electrons (no moments)
d electrons (with moments)
not correlated
correlated
Figure 17: In metallic Ni, the d-orbitals are compact and hybridize weakly due to
low overlap with the s-orbitals (due to symmetry) and with each other (due to low
overlap). Thus, moments tend to form on the d-orbitals and they contribute narrow
features in the electronic density of states. The s-orbitals hybridize strongly and form
a broad metallic band.
calculate the temperature dependence of R (and hence the mag-
netization) using
R =
1
N

k
f(

E
k

IR
2

B
B
0
E
F
)f(

E
k
+
IR
2
+
B
B
0
E
F
)
(59)
with f(x) =
1
e
x
+1
. However, this would be pointless since
all of the approximations made to this point have destroyed the
quantitative validity of the calculation. However, it still retains
29
a qualitative use. For Ni, we can do this by approximating the
very narrow d-electron feature in D(E) as a function and
performing the integral. However, only the d-electrons have
a strong exchange splitting I and hence only they will tend
to contribute to the magnetization. Thus our

D(E) should
reect only the d-electron contribution, we will accommodate
this by setting

D(E) C(E E
F
), (C < 1) (60)
C, an unknown constant, will be determined by the T = 0
behavior. Then
R = C
_
_
_
f(
IR
2

B
B
0
) f(
IR
2
+
B
B
0
)
_
_
_
(61)
Let
R
C


R and T
c
=
IC
4k
B
, then if B
0
= 0

R =
1
exp
_
2

RT
c
T
_
+ 1

1
exp
_
2

RT
c
T
_
+ 1
= tanh

RT
c
T
(62)
If T = 0, then

R = 1 =
R
C
=
1
C
n

N
. For Ni, the measured
ground state magnetization per Ni atom is
eff

B
= 0.54 =
n

N
.
Therefore, C = 0.54 =

eff

B
.
30
For small x, tanh x x
1
3
x
3
, and for large x
tanh x =
sinh x
cosh x
=
e
x
e
x
e
x
+ e
x
=
1 e
2x
1 + e
2x
= (1 e
2x
)(1 e
2x
) 1 2e
2x
(63)
Thus,

R = 1 2e

2T
c
T
, for T T
c
(64)

R =

3(1
T
T
C
)
1
2
, for small

R or T
<

T
c
(65)
However, neither of the formulas is veried by experiment. The
R

T/T
c
Eq (64)
Eq(65)
Figure 18:
critical exponent =
1
2
in Eq. 65 is found to be reduced to

1
3
, and Eq. 64 loses its exponential form, in real systems.
Using more realistic

D(E) or values of I will not correct these
problems. Clearly something fundamental is missing from this
31
model (spin waves).
Elementary
Excitations
spin flip
B
0
Included
spin wave
Not Included
Figure 19: Local spin-ip excitations, left, due to thermal uctuations are properly
treated by mean-eld like theories such as the one discussed in Secs. 3 and 4. However,
non-local spin uctuations due to intersite correlations between the spins are neglected
in mean-eld theories. These low-energy excitations can fundamentally change the
nature of the transition.
3.2.1 Eect of B
If there is an external eld B
0
,= 0, then

R = tanh
_

RT
c
+
B
B
0
/2k
B
T
_

_
(66)
Or for small R and B
0
, (or rather, large T T
c
.)

R =

B
2kT
B
0
+
T
c
T

R

R =

B
2k
1
T T
c
B
0
(67)
Thus since M =

B
N
V
R =
C
B
N
V

R =
M
B
0
=
C
2
B
2kV
N
TT
c
.
This form for
=
Const
T T
c
(68)
32
is called the Curie-Weiss form which is qualitatively satised
for T T
c
; however, the values of Const and T
c
predicted by
band structure are inaccurate. Again, this is due to the neglect
of low-energy excitations.
4 Mean-Field Theory of Magnetism
4.1 Ferromagnetism for localized electrons (MFT)
i = 1
= 2
= 3
= 4
Figure 20: Terms in the Heisenberg Hamiltonian H =

i
J
i
S
i
S
i
g
B
B
0

i
S
i
Here i refers to the sites and refers to the neighbors of site i.
Some of the rare earth metals or ionic materials with valence
d or f electrons are both ferromagnetic and have largely localized
electrons for which the band theory of FM is inappropriate (A
good example is CeSi
2x
, with x > 0.2). As we have seen,
33
systems with localized spins are described by the Heisenberg
Hamiltonian.
H =

i
J
i
S
i
S
i
g
B
B
0

i
S
i
(69)
In general, this Hamiltonian has no solution, and we must
resort to (further) approximation. In this case, we will approx-
imate the eld (exchange plus external magnetic) felt by each
spin as the average eld due to the neighbors of that spin and
the external eld. (See Fig. 21.) Then
i
S
J
Each site nearest neighbors
with exchange interaction J
Figure 21: The mean or average eld felt by a spin S
i
at site i, due to both its neighbors
and the external magnetic eld, is
1
g
B

J
i
S
i
) + B
0
= B
ieff
. Where

J
i
S
i
)
is the internal eld, due to the neighbors of site i.
H

i
g
B
B
ieff
S
i
=

i
S
i

_
_
_

J
i
S
i
) + g
B
B
0
_
_
_
(70)
If J
i
= J is a constant (independent of i and ) describing the
34
exchange between the spin at site i and its nearest neighbors,
then
B
ieff
=
J

i
S
i
) + g
B
B
0
g
B
=
J
g
B
S) + B
0
(71)
M = g
B
N
V
S) ; = #nn. (72)
For a homogeneous, ordered system,
B
eff
=
V
Ng
2

2
B
JM + B
0
= B
MF
+ B
0
(73)
and
H g
B
B
eff

i
S
i
(74)
ie., a system of independent spins in a eld B
eff
. The proba-
bility that a particular spin is up, is then
P

(
g
B
B
eff
1
2
)
(75)
and
P

(
+g
B
B
eff
1
2
)
(76)
so, on average
N

= e
g
B
B
eff
(77)
35
and, since N

+ N

= N
M =
1
2
g
B
N

V
=
1
2
g
B
N
V
tanh
_
_

2
g
B
B
eff
_
_
(78)
Since tanh is odd and B
eff
M, this will only have non-
trivial solutions if J > 0 (if B
0
= 0). If we identify
M
s
=
N
V
g
B
2
; T
c
=
1
4

J
k
(79)
M
M
s
= tanh
_
_
T
c
T
M
M
s
_
_
(80)
and again for T = 0, M(T = 0) = M
s
, and for T
<

T
c
a = tanh (ba)
y = a
y = tanh (ba)
initial slope = b
Figure 22: Equations of the form a = tanh(ba), i.e. Eq. 80, have nontrivial solutions
(a ,= 0) solutions for all b > 1.
M
M
s

3
_
_
1
T
T
c
_
_
1
2
(81)
36
Again, we get the same (wrong) exponent =
1
2
.
When is this approximation good? When each spin really
feels an average eld. Suppose we have an ordered solid, so
that
B
MF
=
J
2g
B
= B
real
(82)
Now, consider one spin ip excitation adjacent to site i only,
i
Figure 23: The ip of a single spin adjacent to site i makes a signicant change in
the eective exchange eld, felt by spin S
i
, if the site has few nearest neighbors.
Fig. 23.
If there are an innite # of spins then B
MF
remains un-
changed but for <
B
real
=
J
2g
B
2

,= B
MF
=
J
2g
B
. (83)
Clearly, for this approximation to remain valid, we need B
real
=
37
B
MF
, which will only happen if
2

= 1 or 2. The
more nearest neighbors to each spin, the better MFT is! (This
remains true even when we consider other lower energy excita-
tions, other than a local spin ip, such as spin waves).
4.2 Mean-Field Theory of Antiferromagnets
Oxides of Fe Co Ni and of course Cu often display antifer-
romagnetic coupling between the transition-metal d orbitals.
Lets assume we have such a magnetic system on a bipartite
lattice composed of two inter-penetrating sublattices, like bcc.
We consider the magnetization of each lattice separately: For
"up" sublattice
"down" sublattice
J < 0
Figure 24: Antiferromagnetism (the Neel state) on a bcc lattice is composed of two
interpenetrating sc sublattices lattices.
38
example, the central site shown in Fig. 24 feels a mean eld
from the = 8 near-neighbor spins on the down sublattice.
so
B = V/( N g )JM
MF

2 2
B
Figure 25:
M
+
=
1
2
g
B
N
+
V
tanh
_

_
g
B
2kT
V
N

g
2

2
B
JM

_
(84)
M

= (+ ) . . . (85)
where M
+
is the magnetization of the up sublattice. These
equations have the same form as that for the ferromagnetic case!
We can make a closer analogy by realizing that N
+
= N

and
M
+
= M

, so that
M
+
=
1
2
g
B
N
+
V
tanh
_

V JM
+
2kTN
+
g
B
_

_
, J < 0 (86)
M

= M
+
(87)
39
Again, these equations will saturate at
M
+
s
= M

s
=
1
2
g
B
N
+
V
(88)
so
M
+
M
+
s
= tanh
_

_
T
N
T
M
+
M
s
_

_
(89)
where T
N
=
1
4
J
k
B
Now consider the eect of a small external eld B
0
. This
will yield a small increase or decrease in each sublattices mag-
netization M

.
M
+
+ M
+
=
1
2
g
B
N
+
V
tanh
_

_
g
B
2kT
_

_B
0
+
V J
N

g
2

2
B
_
M

+ M

_
_

_
_

_
M

+ M

= (+ ) . . . (90)
Or, since
d
dx
tanh x =
1
cosh
2
x
, then
_
M =
M
B
eff
B
eff
_
.
M = M
+
+M

=
1
2
g
B
N
+
V
g
B
2kT
1
cosh
2
x
_

_B
0
+
V J
2N

g
2

2
B
M
_

_
(91)
where x =
T
N
T
M
+
M
+
s
. For T > T
N
, M
+
= 0 and so x = 0, and
M = 2
g
2

2
B
N
8V k
B
T
_

_B
0

4k
B
T
N
V
Ng
2

2
B
M
_

_ (92)
TM =
g
2

2
B
N
4V k
B
B
0
MT
N
(93)
40
M =
g
2

2
B
N
4V k
B
(T + T
N
)
B
0
(94)
=
g
2

2
B
N
4V k
B
(T + T
N
)
(95)
T
T

N N
T
0
Figure 26: Sketch of = Const/(T + T
N
). Unlike the ferromagnetic case, the bulk
susceptibility does not diverge at the transition. However, as we will see, this
equation only applies for the paramagnetic state (T > T
N
), and even here, there are
important corrections.
Below the transition, T < T
N
, the susceptibility displays dif-
ferent behaviors depending upon the orientation of the applied
eld. For T T
N
and a small B
0
parallel to the axis of the
sublattice magnetization, we can approximate M
+
(T) M
+
s
and x
T
N
T
in Eq. 91

g
2

2
B
N
4V k
B
1
T cosh
2
_
T
N
T
_
+ T
N
(96)
41
B
0
Figure 27: When T T
N
, a weak eld applied parallel to the sublattice magnetization
axis only weakly perturbs the spins. Here M
+
(T) M
+
s
and x
T
N
T

g
2

2
B
N
4V k
B
e
2
T
N
T
(97)
Now consider the case where B
0
is perpendicular to the mag-
netic axis. The external eld will cause each spin to rotate a
B
0

B
0
B
MF
Figure 28: When T T
N
, a weak eld B
0
applied perpendicular to the sublattice
magnetization, can still cause a rotation of each spin by an angle proportional to
B
0
/B
MF
.
small angle . (See Fig. 28) The energy of each spin in this
42
external eld and the mean eld
J
g
B
to the rst order in
B
0
is
E =
1
2
g
B
B
0
sin +
1
2
J cos (98)
Equilibrium is obtained when
E

= 0. Since B
0
is taken as
small, 1.
E
1
2
g
B
B
0
+
1
2
J
_
_
1
1
2

2
_
_
(99)
or
E

= 0 =
1
2
g
B
B
0
+
1
2
=
g
B
B
0
J
(100)
The induced magnetization is then
M =
1
2
g
B
B
0
V
=
g
2

2
B
NB
0
2JV
(101)
so

=
g
2

2
B
N
2 [J[ V
= constant (102)
Of course, in general, in a powdered sample, the susceptibility
will reect an average of the two forms, see for example Fig. 30.
43
T

T
N

powdered sample
Figure 29: Below the Neel transition, the lattice responds very dierently to a eld
applied parallel or perpendicular to the sublattice magnetization. However, in a pow-
dered sample, or for a eld applied in an arbitrary direction, the susceptibility looks
something like the sketch on the right.
5 Spin Waves
We have discussed the failings of our mean-eld approaches to
magnetism in terms of their inability to account for low-energy
processes, such as the ipping of spins. (S

) However,
we have yet to discuss the lowest energy spin ip processes which
are spin waves.
We will approach spin-waves two ways. First following Ibach
and Luth we will determine a spin wave in a ferromagnet. Sec-
ond, we will argue that they should be quantized and then
introduce a (canonical) transformation to a Boson representa-
44
95 90 85 80 75 70 65 60 55 50 45 40 35 30 25 20 15 10 5 0
0.0e+0
1.0e-6
2.0e-6
3.0e-6
4.0e-6
5.0e-6
6.0e-6
7.0e-6
8.0e-6
9.0e-6
1.0e-5
1.1e-5
1.2e-5
T (K)
Figure 30: High temperature superconductor Y123 with 25% Fe substituted on Cu,
courtesy W. Joiner, data from a SQUID magnetometer.
tion.
Consider a ferromagnetic Heisenberg Model
H = J

i
S
i
S
i+
(103)
where S
i
= xS
x
i
+ yS
y
i
+ zS
z
i
. If we dene [) =
_
_
_
_
1
0
_
_
_
_
(i.e.
45
[)), =
_
_
_
_
0
1
_
_
_
_
(i.e. [)) so that
S
z
_
_
_
_
0
1
_
_
_
_
=
1
2
_
_
_
_
0
1
_
_
_
_
(104)
and
S
x
=
1
2
_
_
_
_
0 i
i 0
_
_
_
_
S
y
=
1
2
_
_
_
_
0 1
1 0
_
_
_
_
S
z
=
1
2
_
_
_
_
1 0
0 1
_
_
_
_
(105)
_
S

, S

_
= i

(106)
It is often convenient to introduce spin lowering and raising
operators S

and S
+
.
S
+
= S
x
+ iS
y
=
_
_
_
_
0 1
0 0
_
_
_
_
_
S
z
, S

_
= S
+/
(107)
S

= S
x
iS
y
=
_
_
_
_
0 0
1 0
_
_
_
_
_
S
2
, S

_
= 0 (108)
S
+
_
_
_
_
0
1
_
_
_
_
=
_
_
_
_
1
0
_
_
_
_
, S

_
_
_
_
0
1
_
_
_
_
= 0 . . . (109)
They allow is to rewrite H as
H = J

i
S
z
i
S
z
i+
+
1
2
_
S
+
i
S

i+
+ S

S
+
i+
_
(110)
46
Since J > 0, the ground state is composed of all spins oriented,
for example
[0) =
i
[)
i
(i.e. all up) (111)
This is an eigenstate of H, since
S
+
i
S

i+
[0) = 0 (112)
and
S
z
i
S
z
i+
[0) =
1
4
[0) (113)
so
H [0) =
1
4
JN [0) E
0
[0) (114)
where N is the number of spins each with nearest neighbors.
Now consider a spin-ip excitation. (See Fig. 31)
j
Figure 31: A single local spin-ip excitation of a ferromagnetic system. The resulting
state is not an eigenstate of the Heisenberg Hamiltonian.
[
j
) = S

j

n
[)
n
(115)
47
This is not an eigenstate since the Hamiltonian operator S
+
j
S

j+
will move the ipped spin to an adjacent site, and hence create
another state.
However, if we delocalize this spin-ip excitation, then we can
create a lower energy excitation (due to the non-linear nature
of the inter-spin potential) which is an eigenstate. Consider the
j
Figure 32: If we spread out the spin-ip over a wider region then we can create a
lower energy excitation. A spin-wave is the completely delocalized analog of this with
one net spin ip.
state
[k) =
1

j
e
ikr
j
[
j
) . (116)
It is an eigenstate. Consider:
H [k) =
1

j
e
ikr
j
_
_
_

1
4
J(N 2) [
j
)
+
1
2
J [
j
)
1
2
J

([
j+
) + [
j
))
_
_
_
(117)
where the sum in the last two terms on the right is over the near-
48
neighbors to site j. The last two terms may be rewritten:
1

j
e
ikr
j
[
j+
) = 1

N

m
e
ik(r
m
r

)
[
m
) (118)
where
r
j+
= r
j
+ r

= r
m
(119)
so
H [k) =
_
_
_

1
4
JN + J
1
2
J

_
e
ikr

+ e
ikr

_
_
_
_
1

j
e
ikr
j
[
j
)
(120)
Thus [k) is an eigenstate with an eigenvalue
E = E
0
+ J
_
_
_
1
1

cos k r

_
_
_
(121)
Apparently the energy of the excitation described by [k) van-
ishes as k 0.
What is [k)? First, consider
S
z
[k) =

i
S
z
i
[k) =

i
S
z
i
1

j
e
ik r
j
[
j
)
=
1

j
e
ikr
j

i
S
z
i
[
j
) = (SN 1) [k) (122)
I.e., it is an excitation of the ground state with one spin ipped.
Apparently, since E
k=0
= E
0
, the energy to ip a spin in this
way vanishes as k 0.
49
5.1 Quantization of Ferromagnetic Spin Waves
In the ground state all of the spins are up. If we ip a spin,
using a spin-wave excitation, then
S
z
[k) = (SN 1) [k) , S
z
[0) = SN [0) (123)
If we add another spin wave, then S
z
) = (SN2). For spin
1
2
,
S
z
) =
N
2
n, where n is the number of the spin waves. Since
S
z
is quantized, so must be the number of spin waves in each
mode. Thus, we may describe spin waves using Boson creation
and annihilation operators a

and a. By specifying the number


n
k
excitations in each mode k, the corresponding excited spin
state can be described by a Boson state vector [n
1
n
2
. . . n
N
)
We can introduce creation and anhialation operators to de-
scribe the spin excitations on each site. Suppose, in the ground
state, the spin is saturated in the state S
z
= S, then n = 0. If
S
z
= S 1, then n = 1, and so on. Apparently
S
z
i
= S a

i
a
i
S
+
i
a
i
S

i
a

i
(124)
50
S
z
n
3
2
0
1
2
1

1
2
2

3
2
3
Table 1: The correspondence between S
z
and the number of spin-wave excitations on
a site with S = 3/2.
If these excitations are Boselike, then
_
a
i
, a

i
_
= 1 (125)
a
i
[n) =

n
i
[n 1) (126)
a

i
[n) =

n
i
+ 1 [n + 1) (127)
This transformation is faithful (canonical) and will maintain the
dynamical properties of the system (given by

t
= i h[H, ]) if
it preserves the commutator algebra
_
S
+
i
, S

i
_
= 2S
z
i
,
_
S

i
, S
z
i
_
= 2S

i
,
_
S
+
i
, S
z
i
_
= 2S
+
i
(128)
consider
_
S
+
S

S
+
_
[n) = 2S
z
[n) = 2(S n) [n) (129)
51
If S
+
= a, and S

= a

, then the left-hand side of the above


equation would be (n + 1) n [n) = [n) ,= 2(S n) [n).
In order to maintain the commutators, we need
S
+
=

2S na S

= a

2S n (130)
Then
[S
+
, S

] [n) = S
+
S

[n) S

S
+
[n)
=
_
2S a

aaa

_
2S a

a [n) a

(2S a

a)a [n)
= (2S n)(n + 1) [n) n(2S (n 1)) [n)
= (2Sn + 2S n
2
n 2Sn + n
2
n) [n)
= 2(S n) [n) (131)
You can check that this transformation preserves the other
commutators. Of course, we need one other constraint, since
S S
z
S, we also must have
n 2S (132)
This transformation
S
+
i
=
_
2S a

i
a
i
a
i
S

i
= a

i
_
2S a

i
a
i
S
z
i
= S a

i
a
i
(133)
52
is called the Holstein Primako transformation.
If we Fourier transform these operators,
a

i
=
1

k
e
ikR
i
a

k
; a
i
=
1

k
e
ikR
i
a
k
(134)
then (since the Fourier transform is unitary) these new opera-
tors satisfy the same commutation relations
_
a
k
, a

_
=
kk

_
a

k
, a

_
= [a
k
, a
k
] = 0 (135)
To convert the Hamiltonian into this form, assume the number
of magnons in each mode is small and expand
S
+
i
=

2S n
i
a
i

2S(1
n
i
4S
)a
i
(136)

_
1

k
e
ikR
i
a
k

1
4SN
3
2

kpq
e
i(p+qk)R
i
a

k
a
p
a
q
_

_
Of course this is only exact for n
i
2S, i.e. for low T where
there are few spin excitations, and large S (the classical spin
limit).
In this limit
S
+
i

_
2S
N

k
e
ikR
i
a
k
(137)
S

_
2S
N

k
e
ikR
i
a

k
(138)
53
S
z
i
= S
1
N

kk

e
i(kk

)R
i
a

k
a
k
, (139)
the Hamiltonian
H = J

i
_
_
_
S
z
i
S
z
i+
+
1
2
(S
+
i
S

i+
+ S

i
S
+
i+
)
_
_
_
(140)
may be approximated as
H NJS
2
+ 2JS

k
a

k
a
k
2JS

k
_
_
1

e
ik R

_
_
a

k
a
k
+ O(a
4
k
) (141)
H E
0
+

k
2JS(1
k
)a

k
a
k
+ O(a
4
k
) (142)
where
k
=
1

e
ikR

. This is the Hamiltonian of a collection of


k
k
k-q
k+q
k
k
Figure 33: The fourth order correction to Eq. 142 corresponds to interactions between
the spin waves, giving them a nite lifetime
harmonic oscillators plus some other term of order O(a
4
k
) which
54
corresponds to interactions between the spin waves. These in-
teractions are a result of our denition of a spin-wave as an
itinerant spin ip in an otherwise perfect ferromagnet. Once
we have one magnon, another cannot be created in a perfect
ferromagnetic background.
Clearly if the number of such excitations is small (T small)
and S is large, then our approximation should be valid. Fur-
thermore, since these are the lowest energy excitations of our
spin system, they should dominate the low-T thermodynamic
properties of the system such as the specic heat and the mag-
netization. Consider
E) =

k
h
k
e
h
k
1
. (143)
For small k, h
k
= 2JS(1
k
) = 2JSk
2
on a cubic lattice.
Then lets assume that the k-space is isotropic, so that d
3
k
k
2
dk, then

k
=
2

(cos k
x
+ cos k
y
+ ) = 1
k
2

(144)
and
E)

k
2JSk
2
e
2JSk
2
1

_

0
k
4
dk
e
k
2
1
(145)
55
x = k
2
k =
_
_
x

_
_
1
2
dk =
1
2
_
_
1

_
_
1
2
x

1
2
dx (146)
so that
E)
2

1/2
_

0
dx
x
3/2
e
x
1
T
5/2
(147)
Thus, the specic heat at constant volume C
V
T
3/2
, which
is in agreement with experiment.
M/M(0)
T
1 - T
3/2
Figure 34: The magnetization in a ferromagnet versus temperature. At low tempera-
tures, the spin waves reduce the magnetization by a factor proportional to T
3/2
, which
dominates the reduction due to local spin uctuations, derived from our mean-eld
theory. This result is also consistent with experiment.
If we increase the temperature from zero, then the change
in the magnetization is proportional to the number of magnons
generated
M(0) M(T) =
_

k
n
k
_
g
B
V
(148)
56
since each magnon corresponds to spin ip. Thus
M(T) M(0)
_
k
2
dk
e
k
2
1
T
3
2
(149)
which clearly dominates the exponential form found in MFT
(1 2e
2T
c
T
). This is also consistent with experiment!
5.2 Antiferromagnetic Spin Waves
Since the antiferromagnetic ground state is unknown, the spin
wave theory will perturb around the Neel mean-eld state in
which there are both a spin up and down sublattices. Spin
up sublattice
down sublattice
Figure 35: To formulate an antiferromagnetic spin-wave theory, we once again con-
sider a bipartite lattice, which may be decomposed into interpenetrating spin up and
spin down sublattices.
operators can then be written in terms of the Boson creation
57
and annihilation operators as before
up sublattice down sublattice
S
z
i
= S n
i
S
z
i
= S + n
i
(150)
S
+
i
= (S

i
)
+
=

2Sf
i
(S)a
i
S
+
i
=
_
S

i
_
+
=

2Sa

i
f
i
(S)
where
f
i
(S) =

_
1
n
i
2S
and n
i
= a

i
a
i
(151)
Again this transformation is exact (canonical) within the man-
ifold of allowed states
0 n
i
2S S S
z
S . (152)
The Hamiltonian
H = J

i
S
z
i
S
z
i+
=
1
2
_
S
+
i
S

i+
+ S

i
S
+
i+
_
(153)
may be rewritten in terms of Boson operators as
H = +JS
2
N + J

i
a

i
a
i
a

i+
a
i+
JS

i
_
a

i
a
i
+ a

i+
a
i+
+ f
i
(S)a
i
f
i+
(S)a
i+
+ a

i
f
i
(S)a

i+
f
i
(S)
_
(154)
58
Once again, we will expand
f
i
(S) =

_
1
n
i
2S
= 1
n
i
4S

n
2
i
32S
2
(155)
and include terms in H only to O(a
2
)
H JS
2
N JS

i
_
a

i
a
i
+ a

i+
a
i+
+ a
i
a
i+
+ a

i
a

i+
_
(156)
This Hamiltonian may be diagonalized using a Fourier trans-
form
a
i
=
1

k
e
ikR
i
a
k
(157)
and the Bogoliubov transform
a
k
=
k
cosh u
k

k
sinh u
k
(158)
a

k
=

k
cosh u
k

k
sinh u
k
(159)
tanh 2u
k
=
k
(160)

k
=
1

e
ik R

(161)
Here the
k
are also Boson operators
_

k
,

k
_
= 1. To see if this
transform is canonical, we must ensure that the commutators
are preserved.
1 = [a
k
, a

] = [
k
C
k

k
S
k
,

C
k

k
S
k
]
59
=
_
C
2
k
[
k
,

k
] + S
2
k
[

k
,
k
]
_

kk
(162)
=
_
C
2
k
S
2
k
_

kk
=
kk

where C
k
(S
k
) is shorthand for cosh u
k
(sinh u
k
). You should
check that the other relations, [a
k
, a
k
] = [a

k
, a

] = 0, are
preserved.
After this transformation,
H JNS(S + 1) +

k
h
k
_
_

k
+
1
2
_
_
+ O(a
4
) (163)
where h
k
= 2JS
_
1
2
k
. Notice that for small k, h
k

2JSk Ck (C =

2JS is the spin-wave velocity).


The ground state energy of this system (no magnons), is
E
0
= JNS(S + 1) JS

k
_
1
2
k
(164)
If
k
= 0, then each spin decouples from the uctuations of its
neighbors and E
0
= JNS
2
(J < 0) which is the energy of
the Neel state. However, since
k
,= 0, the ground state energy
E
0
< E
N
. Thus the ground state is not the Neel state, and is
thus not composed of perfectly antiparallel aligned spins. Each
sublattice has a small amount of disorder n
i
) in its spin
alignment.
60
E = JNS
N
2
Figure 36: The Neel state of an antiferromagnetic lattice. Due to zero point motion,
this is not the ground state of the Heisenberg Hamiltonian when J < 0 and S is nite.
The linear dispersion of the antiferromagnet means that its
bulk thermodynamic properties will emulate those of a phonon
lattice. For example
E)

k
h
k
e
h
k
1


k
k
e
k
1

_

0
k
3
dk
e
k
1
E) T
4
_

0
x
3
dx
e
x
1
(165)
C =
E)
T
T
3
like phonons! (166)
Which means that a calorimeter experiment cannot distinguish
phonon and magnon excitations of an antiferromagnet.
Therefore, perhaps the most distinctive experiment one may
61
E
f

n
E
i
thermal
spin-
polarized
neutrons
d
dd
S(k, ) I {F(-i[a(t),a (0)])}

n
2 k = k k
i j
h = E E
i f
sample
Figure 37: Polarized neutrons are used for two reasons. First if we look at only spin
ip events, then we can discriminate between phonon and magnon contributions to
S(k, ). Second the dispersion may be anisotropic, so excitations with orthogonal
polarizations may disperse dierently.
perform on an antiferromagnet is inelastic neutron scattering. If
spin-polarized neutrons are scattered from a sample, then only
those with ipped spins have created a magnon. If the neutron
creates a phonon, then its spin remains unchanged. The time
of ight of the neutron allows us to determine the energy loss
or gain of the neutron. Thus, if we plot the dierential cross
section of neutrons with ipped spins, we learn about magnon
dispersion and lifetime. Notice that the peak in S(k, ) has a
width. This is not just due to the instrumental resolution of the
experiment; rather it also reects the fact that magnons have a
62

k

S(k,)
phonon junk
subtracted off
magnon
n
n
2

k
Figure 38: Sketch of neutron structure factor from scattering o of a magnetic system.
The spin-wave peak is centered on the magnon dispersion. It has a width due to the
nite lifetime of magnon excitations.
nite life time t,which broadens their neutron signature by
k
.

k
t h t
1

k
(167)
However in the quadratic spin wave approximation the lifetime
of the modes h
k
is innite. It is the neglected terms in H, of
order O(a
4
) and higher which give the magnons a nite lifetime.
a a a a
i i
i+ i+

Figure 39:
63
Chapter 9: Electronic Transport
Onsager
April 23, 2001
Contents
1 Quasiparticle Propagation 2
1.1 Quasiparticle Equation of Motion and Eective Mass . . . . . . . . . 5
2 Currents in Bands 8
2.1 Current in an Insulator . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Currents in a Metal . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3 Scattering of Electrons in Bands 13
4 The Boltzmann Equation 18
4.1 Relaxation Time Approximation . . . . . . . . . . . . . . . . . . . . . 21
4.2 Linear Boltzmann Equation . . . . . . . . . . . . . . . . . . . . . . . 22
5 Conductivity of Metals 24
5.1 Drude Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.2 Conductivity Using the Linear Boltzmann Equation . . . . . . . . . . 25
6 Thermoelectric Eects 30
6.1 Linearized Boltzmann Equation . . . . . . . . . . . . . . . . . . . . . 31
1
6.2 Electric Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6.3 Thermal and Energy Currents . . . . . . . . . . . . . . . . . . . . . . 34
6.4 Seebeck Eect, Thermocouples . . . . . . . . . . . . . . . . . . . . . 39
6.5 Peltier Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
7 The Wiedemann-Franz Law (for good metals) 42
2
As we have seen, transport in insulators (of heat mostly)
is dominated by phonons. The thermal conductivity of some
insulators can be quite large (cf. diamond). However most
insulators have small and uninteresting transport properties.
Metals, on the other hand, with transport dominated by elec-
trons generally conduct both heat and charge quite well. In
addition the ability to conduct thermal, charge, and entropy
currents leads to interesting phenomena such as thermoelectric
eects.
1 Quasiparticle Propagation
In order to understand the transport of metals, we must un-
derstand how the metallic state propagates electrons: ie., we
must know the electronic dispersion (k). The dispersion is ob-
tained from band structure E(k) = h(k) in which the metal
is approximated as an almost free gas of electrons interacting
weakly with a lattice potential V (r), but not with each other.
3
e
V
V(r)
Figure 1: The dispersion is obtained from band structure E(k) = h(k) in which the
metal is approximated as an almost free gas of electrons interacting weakly with a
lattice potential V (r), but not with each other.
The Bloch states of this system

k
(r) = U
k
(r)e
ikr
, U
k
(r) = U
k
(r + r
n
) (1)
may be approximated as plane waves U
k
(r) = U
k
. Then, the
state describing a single quasiparticle may be expanded.
(x, t) =
1

2
_

dkU(k)e
i(k x (k)t)
(2)
If U(k) = c(k k
0
) then (x, t) e
i(k
0
xt)
and the quasi-
particle is delocalized. On the other hand, if U(k) = constant
then (x, t) (x) and the quasiparticle is perfectly localized.
This is an expression of the uncertainty principle
k x 1 or p x h, (3)
4
Re (x)
Figure 2: If U(k) = c(k k
0
) then (x, t) e
i(k
0
xt)
and the quasiparticle is
delocalized.
so that we cannot know both the momentum and location of
the quasiparticle to arbitrary precision.
(k) ,= constant, so the dierent components propagate with
dierent phase velocities, so the quasiparticle spreads as it prop-
agates. This is also the reason why the group velocity of the
quasiparticle is not the phase velocity. Consider the propaga-
tion of (x, t) which when t = 0.
(x, 0) =
1

2
_

dkU(k)e
ikx
. (4)
Suppose that U(k) has a well-dened dominant peak (See Fig.
3) so that
(k) (k
0
) +
k
(k)[
k
0
(k k
0
)t (5)
then
(x, t)
1

2
_

dkU(k)e
i
(
kx
0
t
k
(k)|
k
0
(kk
0
)t
)
(6)
5
U(k)
k k
0
Figure 3: The distribution of plane wave state that make up a quasiparticle.
(x, t)
e
i
(
k
0

k
(k)|
k
0

0)
t

2
_

dke
ik
(
x
k
(k)|
k
0
t
)
U(k)

_
x
k
(k)[
k
0
t, 0
_
e
i
(
k
0

k
(k)|
k
0

0)
t
(7)
Ie., aside from a phase factor, the quasiparticle travels along
with velocity
k
(k)[
k
0
= v
g
. (If we had considered higher
order terms, we would have seen the quasiparticle distorts as it
propagates. (c.f. Jackson p.305). In general,
v
g
=
k
(k) =
1
h

k
E(k) (8)
1.1 Quasiparticle Equation of Motion and Eective Mass
We are ultimately interested in the transport; i.e. the response
of this quasiparticle to an external electric eld c , from which
6
it gains energy.
E = ec vt (i.e. force distance) (9)
This energy is reected by the quasiparticle ascending to higher
energy k states.
E =
k
E(k) k = hv k (10)
So
hk = ect (11)
h

k = ec E.O.M (12)
This equation of motion is identical to that for free electrons
(c.f. Jackson); however, it may be shown to be applicable to
general Bloch states provided that c is smaller than the atomic
elds, and it must not vary in space or time too fast.
We may put this EOM in a more familiar form
v
i
=
1
h
d
dt
(
k
E(k))
i
=
1
h

2
E
k
i
k
j
dk
j
dt
=
1
h

2
E
k
i
k
j
(ec
j
) (13)
This will have the form F = ma, if we dene the mass tensor
_
_
1
m

_
_
ij
=
_
_
1
m

_
_
ji
=
1
h
2
E
k
i
k
j
symmetric & real (14)
7
which may be diagonalized to dene three principle axes. In
the simple cubic case, the matrix will have the same element
along each principle direction and
m

=
h
2
d
2
E
dk
2
. (15)
In this way the eective mass of electrons on a lattice can vary
strongly, the larger
d
2
E
dk
2
is, the smaller
m

m
is. Consider the simple
1-d case (See Fig. 4)
E(k)
k
k
m*
dk
2
d E
2
= 0
2
dk
2
d E
> 0
2
dk
2
d E
< 0
/a -/a
Figure 4:
8
2 Currents in Bands
Our previous discussion of the motion of an electron (or a quasi-
particle) in a metal under the inuence of an applied eld c,
ignored the presence of other electrons and the Pauli principle.
2.1 Current in an Insulator
The Pauli principle insures that a full band of states is
insulating. Consider the electric current due to d
3
k states
dJ = ev(k)
_
_
L
2
_
_
3
d
3
k (16)
The current density is then
dj =
e
h

k
E(k)
1
(2)
3
d
3
k (17)
ie., dierent occupied states in the Brillouin zone contribute
dierently to the current. The net current density j is then the
integral over all occupied states, which for our full band is the
integral over the rst Brillouin zone
j =
e
8
3
h
_
1st B.Z.

k
E(k)d
3
k. (18)
Thus for each k vector in the integral, there is also k.
9
First B.Z.
Fermi surface
Figure 5: Dierent occupied states make dierent contributions to the current density.
Now consider a lattice with inversion symmetry k k so
that E(k) = E(k). Alternatively, recall that time-reversal
invariance requires that
E

(k) = E

(k) , (19)
but since E

(k) = E

(k) due to spin degeneracy, we must have


that E(k) = E(k). Thus
v
k
=
1
h

k
E(k) =
k
E(k) = v
k
! (20)
i.e., for the insulator
j =
e
8
3
h
_
1st B.Z.
d
3
k
k
E(k) 0 (21)
Now imagine that the band is not full (See Fig. 6, left). Then,
if we apply an external eld c, so that

k =
ec
h
e > 0! (22)
10
occupied states empty
k
x
k
y
empty
occupied states
k
x
k
y
E x
Figure 6: The Fermi sea of a partially lled band will shift under the inuence of an
applied eld E. This destroys the inversion symmetry of the Fermi sea, causing a net
current.
the electrons will redistribute as the Fermi surface shifts (See
Fig. 6 right).
j =
e
8
3
_
koccupied
v(k)d
3
k
=
e
8
3
_
1st B.Z.
d
3
kv(k)
e
8
3
_
empty
d
3
kv(k)
=
e
8
3
_
empty
d
3
kv(k) (23)
Thus the current may be formally described as a current of
positive charge particles (holes) assigned to the unoccupied
states in the band.
11
2.2 Currents in a Metal
Now imagine that the band is almost full. Near the top of the
k
E
k
full states
E
holes
E
E
D(E)
Figure 7: Left: A nearly full simple band. States near the Fermi surface that can be
thermally excited have negative mass. Right: Density of states with holes at the top
which have positive charge and mass.
band
d
2
E
dk
2
< 0, so the mass is negative and the dispersion at the
top of the band is always also parabolic, so
E(k) = E
0

h
2
k
2
2

k = deviation from top! (24)


or
v =
1
h
d
dt

k
E
k
=
h

=
ec

(25)
This is the EOM of a positively charged particle with positive
mass in an electric eld c. I.e., holes at the top of the band
have positive mass.
12
We have just shown that a material with full bands is an
insulator (See Fig. 8 left). Ie., it carries no current, as least at
D(E)
E
Valence band
Conduction band
E
g
E
f
empty
full
D(E)
E
electrons
holes
T=0
T0
Figure 8: An insulator form when the fermi energy falls in a gap of D(E). As the
temperature is raised, electrons are promoted over the gap, and both the electrons and
holes contribute to the conductivity which increases with temperature.
T = 0 and for a small c. However we ignored the presence
of other bands. If there is a conductiong band, for T ,= 0,
and a reasonably small E
g
, there will be conductivity due to
a small number of thermally excited holes and electrons n
exp(E
g
/K
B
T) (See Fig. 8 right). Thus perhaps a better
denition of an insulator is a material for which the conductivity
increases with T.
13
3 Scattering of Electrons in Bands
According to the EOM for hole at the top of a band
v =
e

c (26)
as long as c is nite, these holes will continue to accelerate and
j will increase accordingly. Of course, this does not happen.
Rather the material simply heats up (ie., has a nite R). In
addition, if c is returned to zero, then j likewise returns to
zero. Why?
In 1900 Drude assumed that the electrons scatter from the
lattice yielding resistivity. Of course, as we have seen the quasi-
e
Figure 9: Drude thought that electrons scatter o the lattice yielding resistivity. Bloch
showed this to be wrong.
particle state may be dened from a sum over Bloch waves (de-
scribed by k) each of which is a stationary state and describe
14
the unperturbed propagation of electrons. Thus a perfect lat-
tice yields no resistivity. We can get resistivity in two ways.
1. Deviations from a perfect lattice
(a) Defects (See Fig. 10a)
(b) Lattice vibrations = phonons (See Fig. 10b)
2. Electron - electron interactions (See Fig. 11)
e
e
(a)
k
k + q
q
k
k + q
- q
(b)
or

c c a
k+q k
-q

c c a
k+q k
q
c c (a + a )

k+q k
-q q

E(k) - E(k+q) = h(q)


Figure 10: Electrons do scatter from defects in the lattice or lattice vibrations. They
contribute to the resistivity, with the phonon contribution increasing with temperature,
and the defect contribution more-or-less constant.
15
Due to the strength of the electron-electron interaction and the
density of electrons, (2) should dominate. However, it is easy
to show, using the Pauli principle, that eect of (2) is quite
often negligible, so that we may return to regarding the pure
electronic system as a (perhaps renormalized) non-interacting
Fermi gas.
According to momentum and energy conservation Fig 11
E
1
+ E
2
= E
3
+ E
4
k
1
+ k
2
= k
3
+ k
4
. (27)
(Of course, momentum conservation is only up to a recipro-
1
2
3
4
E
1
k
1
E
2
k
2
E
3
k
3
E
4
k
4
Figure 11: E
1
+ E
2
= E
3
+ E
4
and k
1
+k
2
= k
3
+k
4
. Electron-electron interactions
also contribute to the resistivity (from simple order of magnitude arguments based
on relative strengths of the interactions, their contribution should dominatebut due
to the Pauli principle, it does not).
cal lattice vector G, k
1
+ k
2
= k
3
+ k
4
= G; however, as with
phonon conductivity, these processes with nite Ginvolve much
16
higher energies, and may be neglected near T = 0.) Further-
more, since all states up to E
F
are occupied, E
3
; E
4
> E
F
!
Suppose E
1
is (thermally) excited, so E
1
> E
F
and it collides
with an occupied state E
2
< E
F
. Then
(E
1
E
F
) + (E
2
E
F
) = (E
3
E
F
) + (E
4
E
F
) > 0 (28)

1
+
2
=
3
+
4
> 0,
3
;
4
> 0 (29)
or
1
+
2
> 0, However, since
2
< 0, if
1
is small, then
[
2
[
1
, is also small, so only states with

2
E
F


1
E
F
states
may scatter with the state k
1
conserve energy and obey the
Pauli principle, thus restricting
2
to a narrow shell of width
1
around the Fermi surface.
Now consider the restrictions placed on the states 3 and 4 by
momentum conservation.
k
1
k
3
= k
4
k
2
(30)
I.e. k
1
k
3
and k
4
k
2
must remain parallel, and since k
1
is xed, this restriction on the nal states further reduces the
scattering probability by a factor of

1
E
F
.
17
3
k - k
1 3
k
1
k
4
E
F
2
1
4
k - k
4 2
E
D(E)
k
y
k
x
k
3
k
2
Figure 12: Momentum and energy conservation severley restrict the states that can
an electron can scatter with and into.
Thus the total scattering cross section is reduced from the
classical result
0
by
_

1
E
F
_
2
. If the initial excitation
1
is due to
thermal eects, then
1
k
B
T and

_
_
k
B
T
E
F
_
_
2
1! (31)
The total scattering due to electron - electron repulsion is very
small. Therefore, unless E
F
can be made small, the dominant
contribution to a materials resistivity is due to defects and
phonons.
18
4 The Boltzmann Equation
The nonequilibrium (but steady-state) situation of an electronic
current in a metal driven by an external eld is described by
the Boltzmann equation.
L
V
e j
E
defect
Figure 13: Electronic transport due to an applied eld E, is limited by inelastic colli-
sions with lattice defects and phonons.
This diers from the situation of a system in equilibrium in
that a constant deterministic current diers from random par-
ticle number uctuations due to coupling to a heat and particle
bath. Away from equilibrium (c ,= 0) the distribution function
may depend upon r and t as well as k (or E(k)). Nevertheless,
when c = 0 we expect the distribution function of the particles
19
in V to return to
f
0
(k) = f(r, k, t)[
E=0
=
1
e
(E(k)E
F
)
+ 1
(32)
As indicated,
To derive a form for f(r, k, t), we will consider length scales
larger than atomic distances

A
, but smaller than distances in
which the eld changes signicantly. In this way the system
is considered essentially homogeneous with any inhomogeneity
driven by the external eld. Now imagine that there is no scat-
tering (no defects, phonons), then since electrons are conserved
t - dt
r - dt
dr
dt
k - dt
dk
dt
r
k
t
hk = -eE

Figure 14: In lieu of scattering, particles ow without decay.


f(r, k, t) = f
_
_
r vdt, k +
ec
h
dt, t dt
_
_
(33)
Now consider defects and phonons (See Fig. 15) which can
scatter a qauasiparticle in one state at r v dt and time tdt,
to another at r and time t, so that f(r, k, t) ,= f
_
r vdt, k +
eE
h
dt, tdt
_
.
20
We will express this scattering by adding a term.
Figure 15: Scattering leads to quasiparticle decay.
f(r, k, t) = f
_
_
r v dt, k + ec
dt
h
, t dt
_
_
+
_
_
f
t
_
_
S
dt
(34)
For small dt we may expand
f(r, k, t) = f(r, k, t)v
r
f+ec
k
f
h

f
t
+
_
_
f
t
_
_
S
(35)
or
f
t
+v
r
f
e
h
c
k
f =
_
_
f
t
_
_
S
Boltzmann Equation
(36)
If the phonon and defect perturbations are small, time-independent,
and described by H, then the scattering rate from a Bloch state
k to k

(occupied to unoccupied) is w
k

k
=
2
h
[k

[H[ k)[
2
.
Then
_
_
_
f(k)
t
_
_
_
S
=
_
_
L
2
_
_
3
_
d
3
k

(1 f(k)) w
kk
f(k

)
(1 f(k

)) w
k

k
f(k) (37)
21
Needless to say it is extremely dicult to solve these last two
coupled equations.
4.1 Relaxation Time Approximation
As a result we make a series of approximations and ansatz. The
rst of these is the relaxation time approximation that the rate
at which a system returns to equilibrium f
0
is proportional to
its deviation from equilibrium
_
_
f
t
_
_
S
=
f(k) f
0
(k)
(k)
. (38)
Here (k) is called the relaxation time (for a spatially inhomo-
geneous system will also depend upon r). Ie., we make the
assumption that scattering merely acts to drive a nonequilib-
rium system back to equilibrium.
If c ,= 0 for t < 0 and then at t = 0 it is switched o so
that for t > 0 c = 0, then for a homogeneous system
f
t
=
_
_
f
t
_
_
S
=
f f
0

(39)
so that
f f
0
= (f(t = 0) f
0
) e

(40)
22
ie., is the time constant at which the system returns to equi-
librium.
Now consider the steady-state situation of a metallic system
in a time-independent external eld c = c x. Then
f
t
= 0 (41)
Furthermore since the system is homogeneous

r
f = 0 (42)
then

e
h
c
k
f(k) =
_
_
f
t
_
_
S
=
f(k) f
0
(k)
(k)
(43)
ie
f(k) = f
0
(k) +
e
h
(k)c
k
f(k) (44)
which may be solved iteratively, generating a power series in c
(or c
x
).
4.2 Linear Boltzmann Equation
For small c (Ohmic conditions)
f(k) f
0
(k) +
e
h
(k)c
k
f
0
(k) linear Boltzmann Eqn.
(45)
23
I.e. the lowest order Taylor series of f(k). Or equivalently, if
c = c
x
x
f(k) f
0
_
k +
e
h
(k)c
_
(46)
Ie., the eect is to shift the Fermi surface from its equilibrium
position by an amount
k
x
1ST BZ
k
x
k
y
Figure 16: According to the linear Boltzmann equation, the eect of a eld E
x
is to
shift the Fermi surface by k
x
= e
Ex
h
k
x
= e
c
x
h
(47)
From the discussion in Sec.??, it is clear that a nite current
results.
Interesting! Note that elastic scattering [k[ = [k

[ cannot
restore equilibrium. Rather they would only cause the Fermi
24
surface to expand. Inelastic scattering (i.e. from phonons) is
needed to explain relaxation.
A
B
Figure 17: Note that elastic scattering |k| = |k

| cannot restore equilibrium. Rather


they would only cause the Fermi surface to expand.
5 Conductivity of Metals
5.1 Drude Approximation
As mentioned above, Drude calculated the conductivity of met-
als assuming that
all free electrons participate, and
electron-lattice scattering yields a scattering rate 1/.
Under these assumptions, the EOM is
m v +
m

_
v v
therm
_
= ec (48)
25
where v v
therm
= v
D
, the drift velocity, and
m

v
D
is friction.
Again when c = 0, we again have an exponential decay of v
so is again the relaxation time.
In steady-state v = 0
v
D
=
e
m
c (49)
so that
j = env
D
=
ne
2

m
c (50)
or dening j = c, and = ne,
=
ne
2

m
=
e
m
(51)
5.2 Conductivity Using the Linear Boltzmann Equation
Of course, this is wrong since all free electrons do not par-
ticipate in due to the Pauli principle. And a more careful
derivation, using the Boltzmann Equation, is required. Again,
the relationship between j and f(k) is
j =
e
8
3
_
d
3
k v(k)f(k) (52)

e
8
3
_
d
3
k v(k)
_

_
f
0
(k) +
e(k)
h
c
x
f
0
k
x
_

_
(53)
26
V
E = E x
Figure 18: To calculate the conductivity, we apply a eld in the x-direction only and
use the linearized Boltzmann Eqn.
For an isotopic material j
z
= j
y
= 0, and the equation becomes
scalar. Furthermore, again
_
v(k)f
0
(k)d
3
k = 0 (54)
since v
k
= v
k
. Then as
f
0
k
x
=
f
0
E
E
k
x
=
f
0
E
hv
x
(55)
j
x

e
2
8
3
c
x
_
d
3
kv
2
x
(k)
f
0
E
(56)
Then as
f
0
E
(E E
F
) for T E
F
the integral in k is
conned to the surface of constant E, and
d
3
k = dS
E
dk

= dS
E
dE
hv(k)
(57)
then
27
constant energy
surface
k
x
k
y
k
z
k E
k
dk =

E
k
dE
Figure 19:
=
j
x
c
x
=
e
2
8
3
h
_
dS
E
dE
v
2
x
(k)
v(k)
(k)(E E
F
) (58)
=
e
2
8
3
h
_
E=E
F
dS
E
v
2
x
(k)
v(k)
(k) . (59)
As expected only the properties of the electrons on the Fermi
surface are relevant.
E
E = 0
E 0
Figure 20: Only the electrons near the fermi surface participate in the transport. Far
below the Fermi surface, pairs of states k and k are occupied. Their contribution
to the conductivity cancels, leaving contributions to only the occupied states near the
fermi surface.
We can now calculate the conductivity of a metal by aver-
28
aging
v
2
x
v
(k) over the Fermi surface. Consider a simple system
with a spherical Fermi surface, then
_
dS
E
(k)v
2
x
(k)
v(k)
=
4
3
k
2
F
(E
F
)v(E
F
)
=
4
3
k
2
F
(E
F
)
hk
F
m

(60)
or
=
e
2
8
3
h
4
3
k
2
F
(E
F
)
hk
F
m

(61)
then as k
B
T E
F
, N = 2
4
3

k
3
F
(
2
L
)
3
k
3
F
= 3
2
n we nd
that
=
e
2
(E
F
)
m

n =
e(E
F
)
m

(62)
For semiconductors where n is T dependent, and for more re-
alistic material where the Fermi surface ,= sphere, the formula
is more complicated.
However, for metals the temperature dependence of is dom-
inated by that of ; ie., by the temperature dependence of
phonons. However, before we can calculate (T), we must rst
disentangle the phonon from the defect scattering. Assuming
29
that the two mechanisms are independent, they must add
1

=
1

ph
+
1

defect
(63)
i.e.
=
ph
+
defect
Matthiesens Rule (64)
The defect contribution is proportional to the defect cross sec-
tion
defect
and the current, or v(E
F
),
1

defect

defect
v(E
F
).
Is is roughly temperature independent, since the cross section

defect
and v(E
F
) are.
The phonon contribution, on the other hand, is highly tem-
perature dependent since at zero temperature, there are no
phonons. The scattering cross section is roughly proportional to
the rms phonon excursion
_
S
2
(q)
_
. However, from the equipar-
tition theorem
1
2
M
2
q
_
S
2
(q)
_
=
k
B
T
2
T
D
. (65)
Thus
1

ph

_
S
2
(q)
_

k
B
T
m
2
q
(66)
Ie., at high temperatures, all modes contribute a linear in T
30
scattering to
1

ph
. Therefore, at T
D
(
D
= debye temper-
ature)
= aT +
defect
(67)
C
u
+
3
.3
2
%
N
i
C
u
+
2
.1
6
%
N
i
C
u
+
1
.1
2
%
N
i

T
Cu
R
T
D
T
T
5
0.1
Figure 21: The phonon and defect contributions to the resistivity add (left), and the
phonon contribution is linear at high temperatures T
D
.
6 Thermoelectric Eects
Until now, we have assumed that the transport system is ther-
mally homogeneous. Of course this need not be the case since we
can obviously maintain both an electrical and a thermal current.
Here, each electron can carry a charge current ev e
2
c
and a thermal current kTkT. In fact, a heat current can be
31
T
1
T T
1 2
T
2
x

Figure 22: Thermoelctric eects are important in systems with both electric potential
and thermal gradients. We will assume both are in the x-direction
used to induce an electrical potential (Seebeck or thermoelectric
eect) and, conversely, an electric current can be used to move
heat (Peltier eect) which makes the solid state refrigeration
possible.
6.1 Linearized Boltzmann Equation
To allow for a thermal gradient T, our formalism must be
modied. Imagine that T and c are xed in time, then the
Boltzmann equation becomes
f
t
+v
r
f
e
h
c
k
f =
_
_
f
t
_
_
S
=
f(k) f
0
(k)
(k)
(68)
32
where in steady state
f
t
0. After linearizing (replacing f
by f
0
in the left-hand side), we get
f(k) f
0
(k) (k)
_
v
r
f
0

e
h
c
k
f
0
_
(69)
Then, as before
e
h
c
k
f
0
=
e
h
c
f
0
E
hv =
e
h
c
x
f
0
E
hv
x
(70)
The spatial inhomogeneity is through T, and in a semicon-
ductor for which E
F
depends strongly upon T, through E
F
v
r
f
0
= v
_
_
_
f
0
T
T +
f
0
E
F
E
F
_
_
_
= v
_
_
_
f
0
T
T
f
0
E
E
F
_
_
_
(71)
Apparently E
F
only contributes a term which modies the
electric eld dependence
v
x
f
0
E
ec
x
+ (E
F
)
x
v
x
f
0
E
ec

x
(72)
Of course, in a metal c

= c.
6.2 Electric Current
Thus, we now have
f(k) = f
0
(k)
f
0
T
v
x
T
x
+
e
h
c

x
hv
x
f
0
E
(73)
33
Then for
j
x
=
e
8
3
_
d
3
kv
x
(k)f(k) (74)
j
x
=
e
8
3
_
d
3
kv
x
(k)
_
_
_
f
0
(k)
f
0
T
v
x
T
x
+
e
h
c

x
f
0
E
hv
x
_
_
_
recall that the last term yielded last time (and still will)
j
x
= c

x
+
e
8
3
_
d
3
kv
2
x

f
0
T
T
x
(75)
Again we will calculate the second term assuming a spherical
Fermi surface. The term
f
0
T
connes the integral to the Fermi
E
T 0
T = 0
E
f
f
0
Figure 23: The derivative
f
0
E
is only signicant near the fermi surface.
sphere and so again eectively it amounts to a Fermi-surface
average, so

v
2
x

1
3
v
2

2
3m

1
2
m

v
2
=
2
3m

E or changing to an
integral over the DOS
j
x
= c

x
+
2
3
e
m

_
dE(E)ED(E)
f
0
T
T
x
(76)
34
Assuming that (E) (E
F
), we get
j
x
= c

x
+
2
3
e
m

(E
F
)
T
x
_
dEE
f
0
T
D(E) (77)
j
x
= c

x
+
2
3
e
m

(E
F
)
T
x
c
v
(T) c
v
m

(78)
In general, this intuitive form is rewritten as
j
x
= c

x
+ /
12
xx
_
_

T
x
_
_
(79)
and from it we see that both an electric eld (or the generalized
eld strength c

), and the thermal gradient contribute to the


electron current, j
x
.
6.3 Thermal and Energy Currents
Of course one can have a thermal current without having an
electric current (same number of electrons moving right and
left, but more of the hot ones moving right). Thermodynamics
is needed to quantify this though since these electrons will also
carry entropy as well as energy and heat.
Imagine that a small subsection of our material is in thermal
equilibrium and then some electrons are introduced/taken away
35
so that
dQ = TdS = dU dN First Law of Thermodynamics
(80)
in terms of particle ow
j
Q
= j
E
E
F
j
n
ej
n
= j (81)
where this equation denes j
Q
, the thermal current, and
j
E
=
_
d
3
k
8
3
E(k)v(k)f(k, r) . (82)
Again one could work out the form of j
Q
for the spherical Fermi
surface using the linearized Boltzmann equation. However one
must obtain a form like
j = /
11
c

+ /
12
(T) (83)
j
Q
= /
21
c

+ /
22
(T) (84)
(The fact that /
12
= /
21
is referred as the Onsaser relation.)
These relationship between the /s and the transport co-
ecients depends upon what experiment is being done. For
example in Fig. 24 there is a potential gradient (V ,= 0) but no
36
T
1
T = T
1 2
T
1
I 0
I
V 0
V
j = L E
11
j = L E
Q
12
11
L =
j = E
=
-1
Figure 24: Here, there is a potential gradient (V = 0) but no thermal gradient since
T
1
= T
2
. The electric eld drives both electric and thermal currents. Thus, a heat
bath is required to keep both sides of the sample at the same temperature.
thermal gradient since T
1
= T
2
. The electric eld drives both
electric and thermal currents.
j = /
11
c
j
Q
= /
12
c (85)
Thus, we may identify
=
ne
m

= /
11
(86)
Note that since there is a thermal current induced by the po-
tential gradient, a heat bath is required to keep both sides of
the sample at the same temperature.
37
In Fig. 25 we maintain a thermal gradient, but turn o the
electric current. Here,
T
1
T T
1 2
T
2
I = 0
I
V
Figure 25: Here j = 0 = L
11
E

+ L
12
(T) and j
Q
=
_
L
12
_
L
12
L
11
_
+ L
22
_
(T)
where L
12
_
L
12
L
11
_
+ L
22
= T
j = 0 = /
11
c

+ /
12
(T) (87)
and
j
Q
= /
21
c

+ /
22
(T) =
_
_
_/
21
_
_
_
/
12
/
11
_
_
_ + /
22
_
_
_ (T)
(88)
and since (you will show)
/
12
=
2e
3m

c
v
= /
21
(89)
= /
22

_
/
12
_
2
/
11
(90)
38
We could also measure the thermal conductivity by driving
a heat current through the sample, maintaining the ends at the
same potential (see Fig. 26 right). Here, we would nd
T T
1 2
= L
22
j = L (-T)
12
V V
1 2
V
1
V
2
j
Q
V
1
V
2
I
j
Q
V = V
1 2
j = L (-T)
Q
22
11
L
= L -
22
12
(L )
Figure 26: Two methods for measuring .
= /
22
. (91)
Thus, we can identify
= /
22
or /
22

_
/
12
_
2
/
11
(92)
depending upon the experiment. These are the same if the
sample is a good metal where /
11
= is large (Young Kim).
39
David Mast measures by the method of the left of Fig. 26.
This yields the more conventional denition of .
6.4 Seebeck Eect, Thermocouples
These relations result in some interesting physical eects. Con-
sider a bimetallic conducting loop with two junctions main-
tained at temperatures T
1
T
2
. Let metal A be dierent than B,
so that /
ij
A
,= /
ij
B
. If no current ows around the loop, then
V
metal A
metal B
metal B
T
0
T
1
T
2
2 1
0
Figure 27: A bimetallic conducting loop with junctions maintained at T
1
and T
2
. If
T
1
= T
2
, and L
12
A
/L
11
A
= L
12
B
/L
11
B
, then the heat current induces a potential V T
2
T
1
j = 0 = /
11
c
x
+ /
22
(T) c
x
=
_
_
_
/
12
/
11
_
_
_
dT
dx
(93)
40
where S =
L
12
L
11
is called the thermopower and is a property of
a material.
The potential measured around the loop is given by
V =
_
1
0
c
B
dx +
_
2
1
c
A
dx +
_
0
2
c
B
dx (94)
or
V = S
B
_
_
_
_
1
0
T
x
dx +
_
0
2
T
x
dx
_
_
_
+ S
A
_
2
1
T
x
dx (95)
= S
B
_
1
2
T
x
dx + S
A
_
2
1
T
x
dx (96)
= (S
A
S
B
)
_
T
2
T
1
dT = (S
A
S
B
) (T
2
T
1
) (97)
Or if S
A
and S
B
are not T-independent V =
_
T
2
T
1
dT (S
A
S
B
).
So if T
1
,= T
2
and S
A
,= S
B
, then the heat current in-
duces an emf! This is called the Seebeck eect (solid state
thermometer with ice H
2
O as a reference).
6.5 Peltier Eect
Now consider the inverse situation where an electrical current j
is driven through the loop which is held at a xed temperature
41
metal A
metal B
metal B
T
0
2 1
0
j
Figure 28: An electrical current j is driven through the loop which is held at a xed
temperature
_
T
x
= 0
_
. Then
j
Q
= /
21
c j = /
11
c (98)
j
Q
=
_
_
_
/
21
/
11
_
_
_ j = j (99)
This is known as the Peltier eect whereby heat is carried
from one junction to the other or an electric current is accom-
panied by a heat current. One may use this eect to create an
extremely simple (and similarly inecient) refrigerator.
42
T
0
0
A
j
B
j
B
j
j
Q
j
Q
=
A
B B
=
( - )j = j
B A Q
j = ( - )j
A
B Q
Figure 29: If dT/dx = 0 and
A
= L
21
A
/L
11
A
= L
21
B
/L
11
B
=
B
then the electric current
also induces a heat from one junction to another.
7 The Wiedemann-Franz Law (for good metals)
One may independently measure the thermal and electrical
conductivities. However, in general one expects that T
j = E
V V
1 2
V
1
V
2
j
T
1
T
2
T = T
1 2
T T
1 2
V
1
V
2
j
Q
V = V
1 2
j = (-T)
Q
T
1
T
2
thermal field -
T
x
E
electric field -e/E
Figure 30: The thermal and electrical conductivities may be measured independently.
since in electrical conduction each election carries a charge e
43
and is acted on by a force ec. The current per unit electric
eld proportional to e
2
. In thermal conduction each electron
carries a thermal energy k
B
T and is acted on by a thermal
force k
B
T. The heat current per unit thermal gradient is
proportional to k
2
B
T, thus one expects


k
2
B
e
2
T (100)
Due to the simplicity of these arguments, our formalism
should reproduce this relationship. As we discussed before
j
Q
= j
E
E
F
j
n
(101)
=
1
8
3
_
d
3
k (E E
F
)v(k)f(k) (102)
In the linear approximation to the Boltzmann equation for c

x
=
0, we get
j
Q
=
1
8
3
_
d
3
k (E E
F
)
f
0
T
v
2
x

_
_

T
x
_
_
(103)
where EE
F
and
f
0
T
are odd in (EE
F
), for the Fermi liquid
j
Q

_
_
T
x
_
_
1
3
v
2
F

F
_
d
3
k
3
3
(E E
F
)
f
0
T
(104)
j
Q
=
_
_

T
x
_
_
1
3
v
3
F

F
c
v
(105)
44
E E
F
F
(E - E )
T
f
0
Figure 31: The function (EE
F
)
f
0
T
is sharply peaked at the fermi surface and even.
in E E
F
.
=
1
3
v
3
F

F
c
v
(106)
Now recall that for the Fermi liquid c
v
= k
B

2
2
nk
B
T
k
B
T
F
so that
=
1
3
m

v
2
F
m


F
k
B

2
2
nk
B
T
E
F
=

2
3

F
n
k
2
B
T
m

(107)
However, also for the Fermi liquid, we found that = e
2

F
n
m

,
so

=

2
3
_
_
k
B
e
_
_
2
T (108)
Of course this relationship only holds in a good metal. There
are two reasons for this. First we are neglecting terms like
(
L
12
)
2

in which are unimportant for a good metal (or if we


electrically short the sample.) Second, we are assuming that
is dominated by electronic transport.
45
Chapter 10: Superconductivity
Bardeen, Cooper, & Schrieer
May 9, 2001
Contents
1 Introduction 2
1.1 Evidence of a Phase Transition . . . . . . . . . . . . . . . . . . . . . 2
1.2 Meissner Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 The London Equations 7
3 Cooper Pairing 10
3.1 The Retarded Pairing Potential . . . . . . . . . . . . . . . . . . . . . 11
3.2 Scattering of Cooper Pairs . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 The Cooper Instability of the Fermi Sea . . . . . . . . . . . . . . . . 14
4 The BCS Ground State 17
4.1 The Energy of the BCS Ground State . . . . . . . . . . . . . . . . . . 18
4.2 The BCS Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5 Consequences of BCS and Experiment 28
5.1 Specic Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2 Microwave Absorption and Reection . . . . . . . . . . . . . . . . . . 28
5.3 The Isotope Eect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1
6 BCS Superconducting Phenomenology 32
7 Coherence of the Superconductor Meisner eects 37
8 Quantization of Magnetic Flux 41
9 Tunnel Junctions 43
2
1 Introduction
From what we have learned about transport, we know that there
is no such thing as an ideal ( = 0) conventional conductor.
All materials have defects and phonons (and to a lessor degree
of importance, electron-electron interactions). As a result, from
our basic understanding of metallic conduction must be nite,
even at T = 0. Nevertheless many superconductors, for which
= 0, exist. The rst one Hg was discovered by Onnes in
1911. It becomes superconducting for T < 4.2

K. Clearly this
superconducting state must be fundamentally dierent than the
normal metallic state. Ie., the superconducting state must
be a dierent phase, separated by a phase transition, from the
normal state.
1.1 Evidence of a Phase Transition
Evidence of the phase transition can be seen in the specic heat
(See Fig. 1). The jump in the superconducting specic heat C
s
indicates that there is a phase transition without a latent heat
3
C (J/molK)
T
T
c
C
S
C T
n
Figure 1: The specic heat of a superconductor C
S
and and normal metal C
n
. Below
the transition, the superconductor specic heat shows activated behavior, as if there is
a minimum energy for thermal excitations.
(i.e. the transition is continuous or second order). Furthermore,
the activated nature of C for T < T
c
C
s
e

(1)
gives us a clue to the nature of the superconducting state. It is
as if excitations require a minimum energy .
1.2 Meissner Eect
There is another, much more fundamental characteristic which
distinguishes the superconductor from a normal, but ideal, con-
4
ductor. The superconductor expels magnetic ux, ie., B = 0
within the bulk of a superconductor. This is fundamentally dif-
ferent than an ideal conductor, for which

B = 0 since for any
closed path
C
S
Superconductor
Figure 2: A closed path and the surface it contains within a superconductor.
0 = IR = V =
_
c dl =
_
S
c dS =
1
c
_
S
B
t
dS, (2)
or, since S and C are arbitrary
0 =
1
c

B S

B = 0 (3)
Thus, for an ideal conductor, it matters if it is eld cooled or
zero eld cooled. Where as for a superconductor, regardless
of the external eld and its history, if T < T
c
, then B = 0
inside the bulk. This eect, which uniquely distinguishes an
5
Zero-Field Cooled Field Cooled
B = 0
B = 0
B = 0
B 0
B 0
B 0
T < T
c
T < T
c
T < T
c
T < T
c
T > T
c
T > T
c
Ideal Conductor
Figure 3: For an ideal conductor, ux penetration in the ground state depends on
whether the sample was cooled in a eld through the transition.
ideal conductor from a superconductor, is called the Meissner
eect.
For this reason a superconductor is an ideal diamagnet. I.e.
B = H = 0 = 0 M = H =
1
4
H (4)
6

SC
=
1
4
(5)
Ie., the measured , Fig. 4, in a superconducting metal is very
large and negative (diamagnetic). This can also be interpreted

0
T

T
c
D(E )
Pauli
F
4
-1
H
M
j
s
Figure 4: LEFT: A sketch of the magnetic susceptibility versus temperature of a su-
perconductor. RIGHT: Surface currents on a superconductor are induced to expel the
external ux. The diamagnetic response of a superconductor is orders of magnitude
larger than the Pauli paramagnetic response of the normal metal at T > T
C
as the presence of persistent surface currents which maintain a
magnetization of
M =
1
4
H
ext
(6)
in the interior of the superconductor in a direction opposite
to the applied eld. The energy associated with this currents
7
increases with H
ext
. At some point it is then more favorable
(ie., a lower free energy is obtained) if the system returns to a
normal metallic state and these screening currents abate. Thus
there exists an upper critical eld H
c
Normal
S.C.
T
c
H
c
T
H
Figure 5: Superconductivity is destroyed by either raising the temperature or by ap-
plying a magnetic eld.
2 The London Equations
London and London derived a phenomenological theory of su-
perconductivity which correctly describes the Meissner eect.
They assumed that the electrons move in a frictionless state, so
that
8
m v = ec (7)
or, since
j
t
= en
s
v,
j
s
t
=
e
2
n
s
m
c (First London Eqn.) (8)
Then, using the Maxwell equation
c =
1
c
B
t

m
n
s
e
2

j
s
t
+
1
c
B
t
= 0 (9)
or

t
_
_
m
n
s
e
2
j
s
+
1
c
B
_
_
= 0 (10)
This described the behavior of an ideal conductor (for which
= 0), but not the Meissner eect. To describe this, the
constant of integration must be chosen to be zero. Then
j
s
=
n
s
e
2
mc
B (Second London Eqn.) (11)
or dening
L
=
m
n
s
e
2
, the London Equations become
B
c
=
L
j
s
c =
L
j
s
t
(12)
9
If we now apply the Maxwell equation H =
4
c
j B =
4
c
j then we get
(B) =
4
c
j =
4
c
2

L
B (13)
and
(j) =
1

L
c
B =
4
c
2

L
j (14)
or since B = 0, j =
1
c

t
= 0 and ( a) =
( a)
2
a we get

2
B
4
c
2

L
B = 0
2
j
4
c
2

L
j = 0 (15)
B
SC
j B z x
s

z
B
x
x
y
z

j
^ ^
Figure 6: A superconducting slab in an external eld. The eld penetrates into the
slab a distance
L
=
_
mc
2
4ne
2

.
10
Now consider a the superconductor in an external eld shown
in Fig. 6. The eld is only in the x-direction, and can vary in
space only in the z-direction, then since B =
4
c
j, the
current is in the y-direction, so

2
B
x
z
2

4
c
2

L
B
x
= 0

2
j
sy
z
2

4
c
2

L
j
sy
= 0 (16)
with the solutions
B
x
= B
0
x
e

L
j
sy
= j
sy
e

L
(17)

L
=

c
2

L
4
=

mc
2
4ne
2

is the penetration depth.


3 Cooper Pairing
The superconducting state is fundamentally dierent than any
possible normal metallic state (ie a perfect metal at T = 0).
Thus, the transition from the normal metal state to the super-
conducting state must be a phase transition. A phase transition
is accompanied by an instability of the normal state. Cooper
rst quantied this instability as due to a small attractive(!?)
interaction between two electrons above the Fermi surface.
11
3.1 The Retarded Pairing Potential
The attraction comes from the exchange of phonons. The lat-
v 10 cm/s
F
8
+
+ +
+ +
+
+
+
+
+
+
region of
positive charge
attracts a second
electron
ions
+ +
+
+
e
-
e
-
Figure 7: Origin of the retarded attractive potential. Electrons at the Fermi surface
travel with a high velocity v
F
. As they pass through the lattice (left), the positive ions
respond slowly. By the time they have reached their maximum excursion, the rst
electron is far away, leaving behind a region of positive charge which attracts a second
electron.
tice deforms slowly in the time scale of the electron. It reaches
its maximum deformation at a time
2

D
10
13
s after the
electron has passed. In this time the rst electron has traveled
v
F
10
8cm
s
10
13
s 1000

A
. The positive charge of
the lattice deformation can then attract another electron with-
out feeling the Coulomb repulsion of the rst electron. Due
to retardation, the electron-electron Coulomb repulsion may be
neglected!
12
The net eect of the phonons is then to create an attrac-
tive interaction which tends to pair time-reversed quasiparticle
states. They form an antisymmetric spin singlet so that the
e
e
1000

k
- k
Figure 8: To take full advantage of the attractive potential illustrated in Fig. 7, the
spatial part of the electronic pair wave function is symmetric and hence nodeless. To
obey the Pauli principle, the spin part must then be antisymmetric or a singlet.
spatial part of the wave function can be symmetric and nodeless
and so take advantage of the attractive interaction. Further-
more they tend to pair in a zero center of mass (cm) state so
that the two electrons can chase each other around the lattice.
3.2 Scattering of Cooper Pairs
This latter point may be quantied a bit better by considering
two electrons above a lled Fermi sphere. These two electrons
13
are attracted by the exchange of phonons. However, the max-
imum energy which may be exchanged in this way is h
D
.
Thus the scattering in phase space is restricted to a narrow
shell of energy width h
D
. Furthermore, the momentum in
k
2
k
1
k
2
k
1
k
1
k
1
k
2
k
2

D
E k
2
k
Figure 9: Pair states scattered by the exchange of phonons are restricted to a narrow
scattering shell of width h
D
around the Fermi surface.
this scattering process is also conserved
k
1
+ k
2
= k

1
+ k

2
= K (18)
Thus the scattering of k
1
and k
2
into k

1
and k

2
is restricted to
the overlap of the two scattering shells, Clearly this is negligible
unless K 0. Thus the interaction is strongest (most likely)
if k
1
= k
2
and
1
=
2
; ie., pairing is primarily between
14
time-reversed eigenstates.
k
1
-k
2
K
scattering shell
Figure 10: If the pair has a nite center of mass momentum, so that k
1
+ k
2
= K,
then there are few states which it can scatter into through the exchange of a phonon.
3.3 The Cooper Instability of the Fermi Sea
Now consider these two electrons above the Fermi surface. They
will obey the Schroedinger equation.

h
2
2m
(
2
1
+
2
2
)(r
1
r
2
) +V (r
1
r
2
)(r
1
r
2
) = ( +2E
F
)(r
1
r
2
)
(19)
If V = 0, then = 0, and

V =0
=
1
L
3/2
e
ik
1
r
1
1
L
3/2
e
ik
2
r
2
=
1
L
3
e
ik(r
1
r
2
)
, (20)
15
where we assume that k
1
= k
2
= k. For small V, we will
perturb around the V = 0 state, so that
(r
1
r
2
) =
1
L
3

k
g(k)e
ik(r
1
r
2
)
(21)
The sum must be restricted so that
E
F
<
h
2
k
2
2m
< E
F
+ h
D
(22)
this may be imposed by g(k), since [g(k)[
2
is the probability of
nding an electron in a state k and the other in k. Thus we
take
g(k) = 0 for
_

_
k < k
F
k >

2m(E
F
+ h
D
)
h
(23)
The Schroedinger equations may be converted to a k-space
equation by multiplying it by
1
L
3
_
d
3
r e
ik

r
S.E. (24)
so that
h
2
k
2
m
g(k) +
1
L
3

g(k

)V
kk
= ( + 2E
F
)g(k) (25)
where
V
kk
=
_
V (r)e
i(k k

)r
d
3
r (26)
16
now describes the scattering from (k, k) to (k

, k

). It is
usually approximated as a constant for all k and k

which obey
the Pauli-principle and scattering shell restrictions
V
kk
=
_

_
V
0
E
F
<
h
2
k
2
2m
,
h
2
k

2
2m
< E
F
+ h
D
0 otherwise
. (27)
so
_
_
_
h
2
k
2
m
+ + 2E
F
_
_
_ g(k) =
V
0
L
3

g(k

) A (28)
or
g(k) =
A

h
2
k
2
m
+ + 2E
F
(i.e. for E
F
<
h
2
k
2
2m
< E
F
+ h
D
)
(29)
Summing over k
V
0
L
3

k
A
h
2
k
2
m
2E
F
= +A (30)
or
1 =
V
0
L
3

k
1
h
2
k
2
m
2E
F
(31)
This may be converted to a density of states integral on E =
h
2
k
2
2m
17
1 = V
0
_
E
F
+ h
D
E
F
Z(E
F
)
dE
2E 2E
F
(32)
1 =
1
2
V
0
Z(E
F
) ln
_
_
2 h
D

_
_
(33)
=
2 h
D
1 e
2/(V
0
Z(E
F
))
2 h
D
e
2/(V
0
Z(E
F
))
< 0, as
V
0
E
F
0
(34)
4 The BCS Ground State
In the preceding section, we saw that the weak phonon-mediated
attractive interaction was sucient to destabilize the Fermi sea,
and promote the formation of a Cooper pair (k , k ). The
scattering
(k , k ) (k

, k

) (35)
yields an energy V
0
if k and k

are in the scattering shell E


F
<
E
k
, E
k
< E
F
+ h
D
. Many electrons can participate in this
process and many Cooper pairs are formed, yielding a new state
(phase) of the system. The energy of this new state is not just
18
N
2
less than that of the old state, since the Fermi surface is
renormalized by the formation of each Cooper pair.
4.1 The Energy of the BCS Ground State
Of course, to study the thermodynamics of this new phase, it is
necessary to determine its energy. It will have both kinetic and
potential contributions. Since pairing only occurs for electrons
above the Fermi surface, the kinetic energy actually increases:
if w
k
is the probability that a pair state (k , k ) is occupied
then
E
kin
= 2

k
w
k

k
,
k
=
h
2
k
2
2m
E
F
(36)
The potential energy requires a bit more thought. It may be
written in terms of annihilation and creation operators for the
pair states labeled by k
[1)
k
(k , k )occupied (37)
[0)
k
(k , k )unoccupied (38)
or
[
k
) = u
k
[0)
k
+ v
k
[1)
k
(39)
19
where v
2
k
= w
k
and u
2
k
= 1 w
k
. Then the BCS state, which
is a collection of these pairs, may be written as
[
BCS
)

k
u
k
[0)
k
+ v
k
[1)
k
. (40)
We will assume that u
k
, v
k
1. Physically this amounts to
taking the phase of the order parameter to be zero (or ), so
that it is real. However the validity of this assumption can only
be veried for a more microscopically based theory.
By the Pauli principle, the state (k , k ) can be, at most,
singly occupied, thus a (s =
1
2
) Pauli representation is possible
[1)
k
=
_
_
_
_
1
0
_
_
_
_
k
[0)
k
=
_
_
_
_
0
1
_
_
_
_
k
(41)
Where
+
k
and

k
, describe the creation and anhialation of the
state (k , k )

+
k
=
1
2
(
1
k
+ i
2
k
) =
_
_
_
_
0 1
0 0
_
_
_
_
(42)

k
=
1
2
(
1
k
i
2
k
) =
_
_
_
_
0 0
1 0
_
_
_
_
(43)
Of course
+
k
_
_
_
_
0
1
_
_
_
_
k
=
_
_
_
_
1
0
_
_
_
_

+
k
[1)
k
= 0
+
k
[0)
k
= [1)
k
(44)
20

k
[1)
k
= [0)
k

+
k
[0)
k
= 0 (45)
The process (k , k ) (k

, k

), if allowed, is
associated with an energy reduction V
0
. In our Pauli matrix
representation this process is represented by operators
+
k

k
,
so
V =
V
0
L
3

kk

+
k

k
(Note that this is Hermitian) (46)
Thus the reduction of the potential energy is given by
BCS
[V [
BCS
)

BCS
[V [
BCS
) =
V
0
L
3
_
_
_

p
(u
p
0[ + v
p
1[)

kk

+
k

_
u
p
[0)
p

+ v
p
[1)
p

_
_

_
(47)
Then as
k
1[1)
k

=
kk
,
k
0[0)
k

=
kk
and
k
0[1)
k

= 0

BCS
[V [
BCS
) =
V
0
L
3

kk

v
k
u
k
u
k
v
k
(48)
Thus, the total energy (kinetic plus potential) of the system of
Cooper pairs is
W
BCS
= 2

k
v
2
k

V
0
L
3

kk

v
k
u
k
u
k
v
k
(49)
As yet v
k
and u
k
are unknown. They may be treated as
variational parameters. Since w
k
= v
2
k
and 1 w
k
= u
2
k
, we
21
may impose this constraint by choosing
v
k
= cos
k
, u
k
= sin
k
(50)
At T = 0, we require W
BCS
to be a minimum.
W
BCS
=

k
2
k
cos
2

V
0
L
3

kk
cos
k
sin
k
cos
k
sin
k
=

k
2
k
cos
2

V
0
L
3

kk

1
4
sin 2
k
sin 2
k
(51)
W
BCS

k
= 0 = 4
k
cos
k
sin
k

V
0
L
3

cos 2
k
sin 2
k
(52)

k
tan 2
k
=
1
2
V
0
L
3

sin 2
k
(53)
Conventionally, one introduces the parameters E
k
=
_

2
k
+
2
, =
V
0
L
3

k
u
k
v
k
=
V
0
L
3

k
cos
k
sin
k
. Then we get

k
tan 2
k
= 2u
k
v
k
= sin 2
k
=

E
k
(54)
cos 2
k
=

k
E
k
= cos
2

k
sin
2

k
= v
2
k
u
2
k
= 2v
2
k
1 (55)
w
k
= v
2
k
=
1
2
_
_
1

k
E
k
_
_
=
1
2
_
_
_
_
1

k
_

2
k
+
2
_
_
_
_
(56)
If we now make these substitutions
_
2u
k
v
k
=

E
k
, v
2
k
=
1
2
_
1

k
E
k
__
into W
BCS
, then we get
W
BCS
=

k

k
_
_
1

k
E
k
_
_

L
3
V
0

2
. (57)
22
0
w = v
k k
2
T = 0
k F
= -E +
h k
2 2
2m
clearly kinetic
energy increases
1
Figure 11: Sketch of the ground state pair distribution function.
Compare this to the normal state energy, again measured
relative to E
F
W
n
=

k<k
F
2
k
(58)
or
W
BCS
W
n
L
3
=
1
L
3

k
_
_
1 +

k
E
k
_
_

2
V
0
(59)

1
2
Z(E
F
)
2
< 0. (60)
So the formation of superconductivity reduces the ground state
energy. This can also be interpreted as Z(E
F
) electrons pairs
per and volume condensed into a state below E
F
. The aver-
age energy gain per electron is

2
.
23
4.2 The BCS Gap
The gap parameter is fundamental to the BCS theory. It tells
us both the energy gain of the BCS state, and about its excita-
tions. Thus is usually what is measured by experiments. To
see this consider
W
BCS
=

k
2
k
1
2
_
_
1

k
E
k
_
_

L
3

2
V
0
(61)
Lots of algebra (See I&L)
W
BCS
=

2E
k
v
4
k
(62)
Now recall that the probability that the Cooper state (k , k )
was occupied, is given by w
k
= v
2
k
. Thus the rst pair breaking
excitation takes v
2
k
= 1 to v
2
k
= 0, for a change in energy
E =

k=k

2v
4
k
E
k
+

k
2v
4
k
E
k
= 2E
k
= 2
_

2
k

+
2
(63)
Then since
k
=
h
2
k
2
2m
E
F
, the smallest such excitation is just
E
min
= 2 (64)
This is the minimum energy required to break a pair, or create
an excitation in the BCS ground state. It is what is measured
by the specic heat C e
2
for T < T
c
.
24
k
-k
k k
w = v = 1
2
k
2
v = 0
e
e
Figure 12: Breaking a pair requires an energy 2
_

2
k
+
2
2
Now consider some experiment which adds a single electron,
or perhaps a few unpaired electrons, to a superconductor (ie
tunneling). This additional electron cannot nd a partner for
superconductor
normal
metal
Figure 13:
pairing. Thus it must enter one of the excited states discussed
25
above. Since it is a single electron, its energy will be
E
k
=
_

2
k
+
2
(65)
For
2
k
, E
k
=
k
=
h
2
k
2
2m
E
F
, which is just the energy of
a normal metal state. Thus for energies well above the gap, the
normal metal continuum is recovered for unpaired electrons.
To calculate the density of unpaired electron states, recall
that the density of states was determined by counting k-states.
These are unaected by any phase transition. Thus it must be
that the number of states in d
3
k is equal.
d k
3
k
x
k
y
k
z

L
3
Figure 14: The number of k-states within a volume d
3
k of k-space is unaected by
any phase transition.
D
s
(E
k
)dE
k
= D
n
(
k
)d
k
(66)
In the vicinity of
k
, D
n
(
k
) D
n
(E
F
) since [[ E
F
26
(we shall see that 2w
D
). Thus for
k

D
s
(E
k
)
D
n
(E
F
)
=
d
x
dE
k
=
d
dE
k
_
E
2
k

2
=
E
k
_
E
2
k

2
E
k
>
(67)

1
E
D
s D
n
Density of additional
electron states only!
Figure 15:
Given the experimental and theoretical importance of , it
should be calculated.
=
V
0
L
3

k
sin
k
cos
k
=
V
0
L
3

k
u
k
v
k
=
V
0
L
3

2E
k
(68)
=
1
2
V
0
L
3

2
k
+
2
(69)
Convert this to sum over energy states (at T = 0 all states with
27
< 0 are occupied since
k
=
h
2
k
2
2m
E
F
).
=
V
0
2

_
h
D
h
D
Z(E
F
+ )d

2
+
2
(70)
1
V
0
Z(E
F
)
=
_
h
D
0
d

2
+
2
(71)
1
V
0
Z(E
F
)
= sinh
1
_
_
h
D

_
_
(72)
For small ,
h
D

e
1
V
0
Z(E
F
)
(73)
h
D
e

1
V
0
Z(E
F
)
(74)
sinh x
x
e
x
Figure 16:
28
5 Consequences of BCS and Experiment
5.1 Specic Heat
As mentioned before, the gap is fundamental to experiment.
The simplest excitation which can be induced in a supercon-
ductor has energy 2. Thus
E 2e
2
T T
c
(75)
C
E

T


2
T
2
e
2
(76)
5.2 Microwave Absorption and Reection
Another direct measurement of the gap is reectivity/absorption.
A phonon impacting a superconductor can either be reected
or absorbed. Unless h > 2, the phonon cannot create an ex-
citation and is reected. Only if h > 2 is there absorption.
Consider a small cavity within a superconductor. The cavity
has a small hole which allows microwave radiation to enter the
cavity. If h < 2 and if B < B
c
, then the microwave in-
tensity is high I = I
s
. On the other hand, if h > 2 ,or
29
h
cavity
B
10
h
h = 2
I
n
I - I
s n
microwave
superconductor
B=0
Figure 17: If B > B
c
or h > 2, then absorption reduces the intensity to the
normal-state value I = I
n
. For B = 0 the microwave intensity within the cavity is
large so long as h < 2
B > B
c
, then the intensity falls in the cavity I = I
n
due to
absorbs ion by the walls.
Note that this also allows us to measure as a function of
T. At T = T
c
, = 0, since thermal excitations reduce the
number of Cooper pairs and increase the number of unpaired
electrons, which obey Fermi-statistics. The size of (Eqn. 71) is
only eected by the presence of a Cooper pair . The proba-
bility that an electron is unpaired is f
_

2
+
2
+ E
F
, T
_
=
1
exp

2
+
2
+1
so, the probability that a Cooper pair exists is
30
k
-k
e
e
kT 2
Figure 18:
1 2f
_

2
+
2
+ E
F
, T
_
. Thus for T ,= 0
1
V
0
Z(E
F
)
=
_
h
D
0
d

2
+
2
_
1 2f
_
_

2
+
2
+ E
F
, T
__
(77)
Note that as

2
+
2
0, when we recover the
T = 0 result.
This equation may be solved for (T) and for T
c
. To nd T
c
(T)
(0)
T/T
c
1
In
Sn
Pb
Real SC data (reflectivity)
Figure 19: The evolution of the gap (as measured by reectivity) as a function of tem-
perature. The BCS approximation is in reasonably good agreement with experiment.
31
consider this equation as
T
T
c
1, the rst solution to the gap
equation, with = 0
+
, occurs at T = T
c
. Here
1
V
0
Z(E
F
)
=
_
h
D
0
d

tanh
_
_

2k
B
T
c
_
_
(78)
which may be solved numerically to yield
1 = V
0
Z(E
F
) ln
1.14 h
D
k
B
T
c
(79)
k
B
T
c
= 1.14 h
D
e
1/{V
0
Z(E
F
)}
(80)
but recall that = 2 h
D
e
1/{V
0
Z(E
F
)}
, so
(0)
k
B
T
c
=
2
1.14
= 1.764 (81)
metal T
c

K Z(E
F
)V
0
(0)/k
B
T
c
Zn 0.9 0.18 1.6
Al 1.2 0.18 1.7
Pb 7.22 0.39 2.15
Table 1: Note that the value 2.15 for (0)/k
B
T
c
for Pb is higher than BCS predicts.
Such systems are labeled strong coupling superconductors and are better described by
the Eliashberg-Migdal theory.
32
5.3 The Isotope Eect
Finally, one should discuss the isotope eect. We know that
V
kk
, results from phonon exchange. If we change the mass of
one of the vibrating members but not its charge, then V
0
N(E
F
)
etc are unchanged but

_
k
M
M

1
2
. (82)
Thus T
c
M

1
2
. This has been conrmed for most normal
superconductors, and is considered a smoking gun for phonon
mediated superconductivity.
6 BCS Superconducting Phenomenology
Using Maxwells equations, we may establish a relation between
the critical current and the critical eld necessary to destroy the
superconducting state. Consider a long thick wire (with radius
r
0

L
) and integrate the equation
H =
4
c
j (83)
33
d l

L
S
r
0 j
0
j = j e
0
0
(r - r )/
L
H

H
Figure 20: Integration contour within a long thick superconducting wire perpendicular
to a circulating magnetic eld. The eld only penetrates into the wire a distance
L
.
along the contour shown in Fig. 20.
_
HdS =
_
H dl =
4
c
_
j ds (84)
2r
0
H =
4
c
2r
0

L
j
0
(85)
If j
0
= j
c
(j
c
is the critical current), then
H
c
=
4
c

L
j
c
(86)
Since both H
c
and j
c
, they will share the temperature-
dependence of .
At T = 0, we could also get an expression for H
c
by noting
34
that, since the superconducting state excludes all ux,
1
L
3
(W
n
W
BCS
) =
1
8
H
2
c
(87)
However, since we have earlier
1
L
3
(W
n
W
BCS
) =
1
2
N(0)
2
, (88)
we get
H
c
= 2
_
N(0) (89)
We can use this, and the relation derived above j
c
=
c
4
L
H
c
,
to get a (properly derived) relationship for j
c
.
j
c
=
c
4
L
2
_
N(0) (90)
However, for most metals
N(0)
n
E
F
(91)

L
=

_
mc
2
4ne
2

(92)
taking = 1
j
c
=
c
4

_
4ne
2
mc
2
2

_
n2m
h
2
k
2
F
=

2
ne
hk
F
(93)
35
This gives a similar result to what Ibach and L uth get, but
for a completely dierent reason. Their argument is similar to
one originally proposed by Landau. Imagine that you have a
uid which must ow around an obstacle of mass M. From the
perspective of the uid, this is the same as an obstacle moving
in it. Suppose the obstacle makes an excitation of energy and
M
v
M
v
P
E
Figure 21: A superconducting uid which must ow around an obstacle of mass M.
From the perspective of the uid, this is the same as an obstacle, with a velocity equal
and opposite the uids, moving in it.
momentum p in the uid, then
E

= E P

= Pp (94)
or from squaring the second equation and dividing by 2M
M
P
E
M
P
E
p
(a)
(b)

Figure 22: A large mass M moving with momentum P in a superuid (a), creates an
excitation (b) of the uid of energy and momentum p
36
P
2
2M

P
2
2M
=
P p
M
+
p
2
2M
= E

E = (95)
P
P
p
v = P/M

Figure 23:
=
pP cos
M

p
2
2M
(96)
= pv cos
p
2
2M
(97)
If M (a defect in the tube which carries the uid could
have essentially an innite mass) then

p
= v cos (98)
Then since cos 1
v

p
(99)
Thus, if there is some minimum ,then there is also a mini-
mum velocity below which such excitations of the uid cannot
37
happen. For the superconductor
v
c
=

min
p
=
2
2 hk
F
(100)
Or
j
c
= env
c
=
ne
hk
F
(101)
This is the same relation as we obtained with the previous
thermodynamic argument (within a factor

2). However, the


former argument is more proper, since it would apply even for
gapless superconductors, and it takes into account the fact that
the S.C. state is a collective phenomena ie., a minuet, not a
waltz of electric pairs.
7 Coherence of the Superconductor Meisner
eects
Superconductivity is the Meissner eect, but thus far, we have
not yet shown that the BCS theory leads to the second London
equation which describes ux exclusion. In this subsection, we
will see that this requires an additional assumption: the rigidity
of the BCS wave function.
38
In the BCS approximation, the superconducting wave func-
tion is taken to be composed of products of Cooper pairs. One
can estimate the size of the pairs from the uncertainty principle
2 =
_
_
_
p
2
2m
_
_
_
p
F
m
p p 2m

p
F
(102)

cp
x
h
p

hp
F
2m
=
h
2
k
F
2m
=
E
F
k
F

(103)

cp
10
3
10
4

A
size of Cooper pair wave function (104)
Thus in the radius of the Cooper pair, about
4n
3
_
_

cp
2
_
_
3
10
8
(105)
other pairs have their center of mass.
Figure 24: Many electron pairs fall within the volume of a Cooper wavefunction.
This leads to a degree of correlation between the pairs and to rigidity of the pair
wavefunction.
39
The pairs are thus not independent of each other (regardless
of the BCS wave function approximation). In fact they are
specically anchored to each other; ie., they maintain coherence
over a length scale of at least
cp
.
SC
Normal Metal

BCS
2
>
cp coh
Figure 25:
In light of this coherence, lets reconsider the supercurrent
j =
2e
4m
p

p (106)
where pair mass = 2m and pair charge = 2e.
p = i h
2e
c
A (107)
A current, or a CM momentumK, modies the single pair state
(r
1
, r
2
) =
1
L
3

k
g(k)e
iK (r
1
+r
2
)/2
e
ik (r
1
r
2
)
(108)
(K, r
1
, r
2
) = (K = 0, r
1
, r
2
)e
iKR
(109)
40
where R =
r
1
+r
2
2
is the cm coordinate and hK is the cm mo-
mentum. Thus

BCS
e
i

BCS
(K = 0) = e
i
(0) (110)
= K (R
1
+ R
2
+ ) (111)
(In principle, we should also antisymmetrize this wave function;
however, we will see soon that this eect is negligible). Due to
the rigidity of the BCS state it is valid to approximate
=
R
+
r

R
(112)
Thus
j
s

2e
4m

_
_
_

BCS
_
_
i h
R

+
2eA
c
_
_

BCS
+
BCS
_
_
i h
R

+
2eA
c
_
_

BCS
_

_
(113)
or
j
s
=
2e
2m
_
_
_
[(0)[
2
4eA
c
+ 2 h[(0)[
2

_
_
_
(114)
Then since for any , = 0
j
s
=
2e
2
mc
[(0)[
2
A (115)
41
or since [(0)[
2
=
n
s
2
j =
ne
2
mc
B (116)
which is the second London equation which as we saw in Sec.??
leads to the Meissner eect. Thus the second London equation
can only be derived from the BCS theory by assuming that the
BCS state is spatially homogeneous.
8 Quantization of Magnetic Flux
The rigidity of the wave function (superconducting coherence)
also guarantees that the ux penetrating a superconducting
loop is quantized. This may be seen by integrating Eq. 114
along a contour within the superconducting bulk (at least a
distance
L
from the surface).
j
s
=
e
2
n
s
mc
A
e hn
s
2m

(117)
_
j
s
dl =
e
2
n
s
ms
_
A dl
e hn
s
2m

dl (118)
Presumably the phase of the BCS state
BCS
= e
i
(0) is
42
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
B

L
C
superconducting loop
Figure 26: Magnetic ux penetrating a superconducting loop is quantized. This may
be seen by integrating Eq. 114 along a contour within the superconducting bulk (a
distance
L
from the surface).
single valued, so

_

R

dl = 2N N Z (119)
Also since the path l may be taken inside the superconductor
by a depth of more than
L
, where j
s
= 0, we have that
_
j
s
dl = 0 (120)
so

e
2
n
s
ms
_
A dl =
e
2
n
s
ms
_
B ds = 2N
e hn
s
2m
(121)
Ie., the ux in the loop is quantized.
43
9 Tunnel Junctions
Imagine that we have an insulating gap between two metals,
and that a plane wave (electronic Block State) is propagating
towards this barrier from the left
a b c
V
0
metal
insulator
metal
d
dx
2
2
2m
2
h
+ E = 0
dx
2
d
2
2m
2
h
+ (E - V )
0
d
dx
2
2
2m
2
h
+ E = 0
V
x
0
d
Figure 27:

a
= A
1
e
ikx
+ B
1
e
ikx

b
= A
2
e
ik

x
+ B
2
e
ik

c
= B
3
e
ikx
(122)
These are solutions to the S.E. if
k =

2mE
h
in a & c (123)
k

=
_
2m(E V
0
)
h
in b (124)
44
The coecients are determined by the BC of continuity of
and

at the barriers x = 0 and x = d. If we take B


3
= 1
and E < V
0
, so that
k

= i =
_
2m(E V
0
)
h
(125)
then, the probability of having a particle tunnel from left to
right is
P
lr

[B
3
[
2
[B
1
[
2
=
1
[B
1
[
2
=
_

_
1
2

1
8
_
_
k

k
_
_
2
+
1
8
_
_
k

+

k
_
_
2
cosh 2d
_

_
1
(126)
For large d
P
lr
8
_
_
k

+

k
_
_
2
e
2d
(127)
8
_
_
k

+

k
_
_
2
exp
_

2d
_
2m(V
0
E)
h
_

_
(128)
Ie, the tunneling probability falls exponentially with distance.
Of course, this explains the physics of a single electron tun-
neling across a barrier, assuming that an appropriate state is
45
lled on the left-hand side and available on the right-hand side.
This, as can be seen in Fig. 28, is not always the case, es-
pecially in a conductor. Here, we must take into account the
densities of states and their occupation probabilities f. We will
be interested in applied voltages V which will shift the chemical
potential eV . To study the gap we will apply
X
eV
S I
N
N(E)
E
Figure 28: Electrons cannot tunnel accross the barrier since no unoccupied states are
available on the left with correspond in energy to occupied states on the right (and
vice-versa). However, the application of an appropriate bias voltage will promote the
state on the right in energy, inducing a current.
eV (129)
We know that
2
k
B
T
c
4,
4k
B
T
c
2
10

K. However typical
metallic densities of states have features on the scale of electron-
46
volts 10
4
K. Thus, on this energy scale we may approximate
the metallic density of states as featureless.
N
r
() = N
metal
() N
metal
(E
F
) (130)
The tunneling current is then, roughly,
I P
_
df( eV )N
r
(E
F
)N
l
()(1 f())
P
_
df()N
l
()N
r
(E
F
)(1 f( eV )) (131)
For eV = 0, clearly I = 0 i.e. a balance is achieved. For
E
F
Figure 29: If eV= 0, but there is a small overlap of occupied and unoccupied states on
the left and right sides, then there still will be no current due to a balance of particle
hopping.
eV ,= 0 a current may occur. Lets assume that eV > 0
and k
B
T . Then the rightward motion of electrons is
47
suppressed. Then
I PN
r
(E
F
)
_
df( eV )N
l
() (132)
and
dI
dV
PN
r
(E
F
)
_
d
f( eV )
V
N
l
() (133)
f
V
e( eV E
F
) (T E
F
) (134)
dI
dV
PN
r
(E
F
)N
l
(eV + E
F
) (135)
Thus the low temperature dierential conductance
dI
dV
is a mea-
sure of the superconducting density of states.
/e
I
V
V
/e
dI
dV
Figure 30: At low temperatures, the dierential conductance in a normal metal
superconductor tunnel junction is a measure of the quasiparticle density of states.
48
Chapter 11: Dielectric Properties of Materials
Lindhardt
May 8, 2002
Contents
1 Classical Dielectric Response of Materials 2
1.1 Conditions on . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Kramers Kronig Relations . . . . . . . . . . . . . . . . . . . . . . . 6
2 Absorption of E and M radiation 8
2.1 Transmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Reectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Model Dielectric Response . . . . . . . . . . . . . . . . . . . . . . . . 13
3 The Free-electron gas 17
4 Excitons 19
1
Electromagnetic elds are essential probes of material prop-
erties
IR absorption
Spectroscopy
The interaction of the eld and material may be described ei-
ther classically or Quantum mechanically. We will rst do the
former.
1 Classical Dielectric Response of Materials
Classically, materials are characterized by their dielectric re-
sponse of either the bound or free charge. Both are described
by Maxwells equations
E =
1
c
B
t
, H =
4
c
j +
1
c
D
t
(1)
and Ohms law
j = E . (2)
Both eects may be combined into an eective dielectric con-
stant , which we will now show. For an isotropic medium, we
2
x <<
< G (k,) () k <
E(x,t) E(x ,t)
0

x
0
e
x
Figure 1: If the average excursion of the electron is small compared to the wavelength
of the radiation < x > , then we may ignore the wave-vector dependence of the
radiation so that (k, ) ().
have
D() = ()E() (3)
where
E(t) =
_
de
it
E() H(t) =
_
de
it
H() (4)
D(t) =
_
de
it
D() B(t) =
_
de
it
B() (5)
E() = E

() E(t) (6)
Then H =
4
c
j +
1
c
D
t

3

_
de
it
H() =
4
c
_
de
it
j() +
1
c

t
_
de
it
D()
(7)
_
de
it
_
_
_
H()
4
c
j()
1
c
(i)D()
_
_
_
= 0 (8)
H =
4
c
j i

c
D()
=
4
c
E i

c
E
=
_

_
i

c
E
_

c

4
c
_
=
i
c
E
4
c
E
_

c
4
i
c

_
=
4
c
E
(9)
Thus we could either dene an eective conductivity =
i
4
which takes into account dielectric eects, or an eective
dielectric constant = +i
4

, which accounts for conduction.


1.1 Conditions on
From the reality of D(t) and E(t), one has that E(+) =
E

() and D() = D

(), hence for D = E,


() =

() (10)
4
Additional constraints are obtained from causality
D(t) =
_
d()E()e
it
=
_
d()e
it
_
dt

2
e
it

E(t

)
=
1
2
_
dtd (() 1 + 1) E(t

)e
i(tt

)
(11)
then we make the substitution
()
1
4
(12)
D(t) = 2
_
dtd
_
() +
1
4
_
E(t

)e
i(tt

)
D(t) = E(t) + 2
_
dtd()E(t

)e
(tt

)
(13)
Dene G(t) 2
_
d()e
it
, then
D(t) = E(t) +
_

dt

G(t t

)E(t

) (14)
Thus, the electric displacement at time t depends upon the eld
at other times; however, it cannot depend upon times t

> t by
causality. Hence
G() 0 < 0 (15)
5
< 0
no poles
poles
X X
Figure 2: If () is analytic in the upper half plane, then causality is assured.
This can be enforced if () is analytic in the upper half
plane, then we may close
G() = 2
_

d()e
i
= 2
_
d()e
i
0 (16)
contour in the upper half plane and obtain zero for the integral
since is analytic within and on the contour.
1.2 Kramers Kronig Relations
One may also derive an important relation between the real
and imaginary parts of the dielectric function () using this
analytic property. From the Cauchy integral formula, if is
6
analytic inside and on the contour C, then
(z) =
1
2i
_

C
(

)
z

(17)
Then taking the contour shown in Fig. 3 and assuming that

.
Figure 3: The contour (left) used to demonstrate the Kramers Kronig Relations.
Since the contour must contain , we must deform the contour so that it avoids the
pole 1/(

).
()
<

for large (in fact Re


<

and Im
<

2
) we
may ignore the large semicircle. Let z = + i0
+
, so that
( + i0
+
) =
1
2i
_

)d

i0
+
(18)
2i( + i0
+
) = P
_

)d

+ i() (19)
() =
1
i
P
_

(20)
Then let 4() = () 1 =
1
+ i
2
1 and we get

1
() + i
2
() 1 =
2

P
_

i
1
(

) +
2
(

) + 1

(21)
7
or

1
() 1 =
1

P
_

2
(

(22)

2
() =
1

P
_

1
(

) 1

(23)
which are known as the Kramers Kronig relations. Many ex-
periments measure
2
and from Eq. 22 we can calculate
1
!
2 Absorption of E and M radiation
2.1 Transmission
= e
0
-i(t-nx/c)
~
n = 1
~
n = 1
~

~
n = n + i = ()
= n -
1
2
2
~
n = n +2in - = + i
1
2 2
2
2
= 2n
Figure 4: In transmission experiments a laser beam is focused on a thin slab of some
material we wish to study.
8
Imagine that a laser beam of known frequency is normally
incident upon a thin slab of some material (see Fig. 4), and we
are able to measure its transmitted intensity.
Upon passing through a boundary, part of the beam is re-
ected and part is transmitted (see Fig. 5). In the case of normal
incidence, it is easy to calculate the related coecients from the
conditions of continuity of E

and H

= B

( = 1)
X
X
XX
= 1
= 1

B
B
B
0

0
Figure 5: The assumed orientation of electromagnetic elds incident on a surface.
E =
1
c
B
t
, ik E =
i
c
B (24)
|nE

| = |B

| (25)
E
0
+ E

= E

(E
0
E

) = nE

_
t =
2
n + 1
(26)
9
In this way the other coecients may be calculated (see Fig. 6).
Accounting for multiple reections the total transmitted eld is
t
1
r
1
r
2
t
2
t =
1
2
n + 1
~
t =
2
2n
n + 1
~
~
r =
2 n + 1
n - 1
~
~
Figure 6: Multiple events contribute the the radiation transmitted through and reected
from a thin slab.
E = E
0
t
1
t
2
e
ikd
+ E
0
t
1
r
2
r
2
t
2
e
3ikd
+ , where k =
n
c
(27)
E =
E
0
t
1
t
2
e
id n/c
1 r
2
2
e
2id n/c
(28)
If n 1 and d is small, then
E E
0
e
id n/c
(29)
I I
0
e
id( n n

)/c
= I
0
e
d2/c
(30)
=

2
2

1
(31)
I I
0
e
(
2
/c)d
I
0
_
1

c
d
_
(32)
10
If the thickness d is known, then the quantity

2
()
c
=
4
1
()
c
(33)
may be measured (the absorption coecient). The real part
of the dielectric response,
1
() may be calculated with the
Kramers Kronig relation

1
() = 1 +
1

P
_

2
(

(34)
According to Young Kim, this analysis works for suciently
thin samples in the optical regime, but typically fails in the IR
where
2
becomes large.
2.2 Reectivity
Of course, we could also have performed the experiment on a
very thick slab of the material, and measured the reectivity R
However, this is a much more complicated experiment since
R
I
depends upon both
1
and
2
(), and is hence much more dif-
cult to analyze [See Frederick Wooten, Optical Properties
of Solids, (Academic Press, San Diego, 1972)].
11
n
~
n = 1
~
R
I
2
n - 1
~
n +1
~
=
Figure 7: The coecient of reectivity (the ratio of the reected to the incident inten-
sities) R/I depends upon both
1
() and
2
(), making the analysis more complicated
than in the transmission experiment.
To analyze these experiments, one must rst measure the
reectivity, R() over the entire frequency range. We then
write
R() = r()r

() =
(n 1)
2
+
2
(n + 1)
2
+
2
(35)
where n = n + i and
r() = ()e
i
(36)
so that R() = ()
2
. If () 1 and 0 fast enough
as the frequency increases, then we may employ the Kramers
Kronig relations replacing with ln r() = ln () +i(), so
that
(

) =
2

P
_

0
ln
_
R()

2
d (37)
12
Thus, if R() is measured over the entire frequency range
where it is nite, then we can calculate (). This complete
knowledge of R is generally not available, and various extrapo-
lation and tting schemes are used on R() so that the integral
above may be completed.
We may then use Eq. 35 above to relate R and to the real
and imaginary parts of the refractive index,
n() =
1 R()
1 + R() 2
_
R() cos ()
(38)
() =
2
_
R() sin ()
1 + R() 2
_
R() cos ()
, (39)
and therefore the dielectric response () = (n() + ())
2
.
2.3 Model Dielectric Response
We have seen that EM radiation is a sensitive probe of the
dielectric properties of materials. Absorption and reectivity
experiments allow us to measure some combination of
1
or

2
, with the remainder reconstructed by the Kramers-Kronig
relations.
13
In order to learn more from such measurements, we need to
have detailed models of the materials and their corresponding
dielectric properties. For example, the electric eld will interact
with the moving charges associated with lattice vibrations. At
the simplest level, we can model this as the interaction of iso-
lated dipoles composed of bound damped charge e

and length
s
p = e

s (40)
The equation of motion for this system is
s = s
2
0
s + e

E (41)
where the rst term on the right-hand side is the damping force,
the second term is the restoring force, and the third term is the
external eld. Furthermore, the polarization P is
P =
N
V
e

s +
N
V
E (42)
where is the polarizability of the dierent molecules which
make up the material. It is to represent the polarizability of
rigid bodies. For example, (see Fig. 8) a metallic sphere has
= a
3
(43)
14
a
E = 0
Figure 8: We may model our material as a system harmonically bound charge and of
metallic spheres with polarizablity = a
3
.
If we F.T. these two equations, we get

2
s = irs
2
0
s + e

E() (44)
s()
_

2
0

2
i
_
=
e

E()

(45)
and
P() = ne

s() + nE() (46)


or
P() = E()
_

_
ne
2
/

2
0

2
i
+ n
_

_
(47)
The term in brackets is the complex electric susceptibility of
the system = E/P.
() = n +
ne
2
/

2
0

2
i
=
1
4
(() 1) (48)
() = 1 + 4n +
4ne
2
/

2
0

2
i
, where
2
p
=
4ne
2

(49)
15
Or, introducing the high and zero frequency limits

= 1 + 4n (50)

0
= 1 + 4n +

2
p

2
0
=

+

2
p

2
0
(51)
() =

+

2
0
(
0

2
0

2
i
(52)
For our causality arguments we must have no poles in the upper
complex half plane. This is satised since > 0. In addition,
we need
=
1
4
(() 1) 0, as (53)
This means that

= 1 + 4n = 1. Of course is nite.
The problem is that in making = constant, we neglected the
electron mass. Ie., for our example of a metallic sphere, < a
3
for very high ! (See Fig. 9) due to the nite electronic mass.
To analyze experiments is separated into real
1
and imag-
inary parts
2

1
() =
4


2
=

+
(
0

)
2
0
(
2
0

2
)
(
2
0

2
)
2
+
2

2
(54)

2
() =
4


1
=
(
0

)
2
0

(
2
0

2
)
2
+
2

2
(55)
16
a
()
-
-
-
-
-
Figure 9: For very high , < a
3
since the electrons have a nite mass and hence
cannot respond instantaneously to changes in the eld.
Thus a phonon mode will give a roughly Lorentzian-like line
shape in the optical conductivity
1
centered roughly at the
phonon frequency. In addition, this form may be used to con-
struct a model reectivity R = | n 1|
2
/| n + 1|
2
, with n =

which is often appended to the high end of the reectivity data,


so that the Kramers-Kronig integrals may be completed.
3 The Free-electron gas
Metals have a distinct feature in their optical conductivity
1
()
which may be emulated by the free-electron gas. The equation
of motion of the free-electron gas is
nm s = s neE (56)
17

0
0
()
1
()
2

Figure 10: Sketch of the real and imaginary parts of the dielectric response of the
harmonically bound charge model.
In steady state s = neE; however
ne s = j = E =
ne
2

m
E (57)
in the relaxation time approximation, so
s =
e
m
E =
ne

E or =
nm

. (58)
Thus, the equation of motion is
nm s =
nm

s neE (59)
If we work in a Fourier representation, then
m
2
s() =
im

s() eE() (60)


18
However, since P = ens
_
_
m
2
+
im

_
_
P = ne
2
E() (61)
or
P =
ne
2
/m

2
+ i/
E = E =
1
4
( 1)E (62)
= 1
4n
2
e
2
m
1
+ i/
= 1

2
p

_
i/

2
+ 1/
2
_

_
(63)

1
= 1

2
p

2
+ 1/
2
_

_
,
2
=

2
p

2
+ 1/
2
(64)
However, recall that
1
=
4
2

and
2
=
4
1

, so that

1
() =

2
p
4
1/

2
+ 1/
2
. (65)
4 Excitons
One of the most dramatic eects of the dielectric properties
of semiconductors are excitons. Put simply, an exciton is a
hydrogenic bound state made up of a hole and an electron.
The Hamiltonian for such a system is
H =
1
2
m

h
v
2
h
+
1
2
m

e
v
2
e

e
2
r
(66)
19
Drude peak
(electronic)
1/
phonons
()
1
() d = /4
2
1
p
Figure 11: A sketch of the optical conductivity of a metal at nite temperatures. The
low-frequency Drude peak is due to the coherent transport of electrons. The higher
frequency peak is due to incoherent scattering of electrons from phonons or electron-
electron interactions.
As usual, we will work in the center of mass, so
1
2
m

h
v
2
h
+
1
2
m

e
v
2
e
=
1
2
(m

h
+ m

e
)

R
2
+ r
2
(67)

R =
m

e
x
e
+ m

h
x
h
m

h
+ m

e
, r = x
e
x
h
(68)
The eigenenergies may be obtained from the Bohr atom solution
E
n
= (Ze
2
)
2
/(2 h
2
n
2
) by making the substitutions
Ze
2

e
2

(69)

=
m

h
m

e
m

h
+ m

e
(70)
In addition there is a cm kinetic energy, let hk = P, then
E
nK
=
h
2
K
2
2(m

h
+ m

e
)

e
4

2
2 h
2
n
2
+ E
g
(71)
20
electron
hole
E
g
O
R
K
r
Figure 12: Excitons are hydrogenic bound states of an electron and a hole. In semi-
conductors with an indirect gap, such as Si, they can be very long lived, since a
phonon or defect must be involved in the recombination process to conserve momen-
tum. For example, in ultra-pure Si ( 10
12
impurities per cc) the lifetime can exceed

Si
10
5
s; whereas in direct-gap GaAs, the lifetime is much shorter
GaAs
10
9
s
and is generally limited by surface states. On the right, the coordinates of the exciton
in the center of mass are shown.
The binding energy is strongly reduced by the dielectric eects,
since 10. (it is also reduced by the eective masses, since
typically m

< m). Thus, typically the binding energy is a


small fraction of a Ryberg. Similarly, the size of the exciton is
much larger than the hydrogen atom
a

m
e
a
0
(72)
This in fact is the justication for the hydrogenic approxima-
tion. Since the orbit contains many sites the lattice may be
approximated as a continuum and thus the exciton is well ap-
21
proximated as a hydrogenic atom.
10A

e
e
+
Figure 13: An exciton may be approximated as a hydrogenic atom if its radius is large
compared to the lattice spacing. In Si, the radius is large due to the reduced hole and
electron masses and the enhanced dielectric constant 10.
22
Chapter 12: Semiconductors
Bardeen & Shottky
May 18, 2001
Contents
1 Band Structure 4
2 Charge Carrier Density in Intrinsic Semiconductors. 7
3 Doping of Semiconductors 12
4 Carrier Densities in Doped semiconductor 15
1
Semiconductors are of obvious technological importance - so
much so, that a whole chapter will be dedicated to them.
Semiconductors are distinguished from metals in that they
have a gap at the Fermi surface, and are distinguished from
insulators in that the gap is small
<

1eV . Most condensed


metal insulator semiconductor
Figure 1: There is no band gap at the Fermi energy in a metal, while there is a band
gap in an insulator. Semiconductors on the other hand have a band gap, but it is
much smaller than those found in insulators.
matter physicists make the distinction on the basis of the con-
ductivity and its temperature dependence. In the Drude model
(parabolic band)
=
ne
2

, =
e
m

, = ne (1)
Almost always
1

increases as T increases, ie the thermal exci-


tations increase the scattering rate and decrease the lifetime of
2
the quasiparticle.
For example, we have seen that
1

T at high temperatures
due to electron-phonon interactions. In metals, n is about con-
stant, so the temperature dependence of metals is dictated by
.
metals
1
T
, as T (2)
However, in semiconductors, the population of free carriers n is
temperature dependent. The exponential always will dominate
E
g
e n
-E /2kT
g
Figure 2: This shows the temperature dependence for the excitation of electrons, thus
allowing the number of free carriers to vary with changes in temperature.
the power law dependence of .
n
1
T
e
E
g
/kT
(3)
as T (4)
The same is true for insulators, of course, except here n is so
3
small that for all realistic purposes 0.
1 Band Structure
Clearly the band structure of the semiconductors is crucial then
for their device applications. Semiconductors fall into several
categories, depending upon their composition, the simplest,
type IV include silicon and germanium. The type refers to
their valence.
r
0
P
S
r
E
conduction band
valence band
B
AB
E
g
E T
g
sp 4 electrons
per band
3
Figure 3: Sketch of the sp
3
bands in Si vs. Si-Si separation.
Recall that Si and Ge have a s
2
p
2
atomic shell, which forms
highly directional sp
3
hybrid bonds in the solid state (with
tetragonal symmetry). It is the covalent bonding, or rather
the splitting between the bonding and antibonding bands, that
4
forms the gap. The band structure is also quite rich

K
L
X
000
100
111
110
K
X K L
Si
E
g
Figure 4: Sketch of quasiparticle bands in Si (right) along the high symmetry directions
(left).
1
m

ij
=
1
h
2

2
E(k)
k
i
k
j
(5)
The situation in III-V semiconductors such as GaAs is sim-
ilar, in that covalent sp
3
bands still form. However, the gap
is direct. For this reason GaAs makes more ecient optical
devices than does either Si or Ge. A particle-hole excitation
across the gap can readily recombine, emit a photon (which has
essentially no momentum) and conserve momentum in GaAs;
whereas, in an indirect gap semiconductor, this recombination
requires the addition creation or absorption of a phonon or some
5
X K L
Ge
E
g
Figure 5: Sketch of quasiparticle bands in Ge along the high symmetry directions.
Note the indirect, roughly L, minimum gap energy.
other lattice excitation to conserve momentum. For the same
reason, excitons live much longer in Si and especially Ge than
they do in GaAs.
material
exciton
GaAs 1ns(10
9
s)
Si 19s(10
5
s)
Ge 1ms(10
3
s)
Table 1:
6

X K L
GaAs
E 1.5eV
g
Figure 6: Sketch of quasiparticle bands in GaAs along the high symmetry directions
of the Brillouin zone. Note the direct, , minimum gap energy. The nature of
the gap can be tuned with Al doping.
2 Charge Carrier Density in Intrinsic Semicon-
ductors.
Both electrons and holes contribute to the conductivity with the
same sign. Here the mobilities are assumed to be constant. This
is valid since for semiconductors all of the conducting carriers
full near the top or bottom of bands, where E
k

h
2
k
2
2m

p
and
the eective mass approximation is valid. Here, we found that
e/m
However, as mentioned before, the carrier concentrations are
7
conduction
valence
h E
g
phonon
photon
v k
s
ck E / h
g
Figure 7: A particle-hole excitation across the gap can readily recombine, emit a
photon (which has essentially no momentum) and conserve momentum in a direct
gap semiconductor (left) such as GaAs. Whereas, in an indirect gap semiconductor
(right), this recombination requires the addition creation or absorption of a phonon
or some other lattice excitation to conserve momentum.
E
j
v
j
v
e
+
e
-
n =
# electrons
volume
p =
# holes
volume
= e(n + p )
n r
Figure 8: The contribution of the electrons and holes to the conductivity.
highly T-dependent since all of the carriers in an intrinsic (un-
doped) semiconductor are thermally induced (i.e. n = p = 0 at
T = 0).
n =
_
E
top
E
c
D
C
(E)f(E, T)dE
_

E
c
D
C
(E)f(E, T)dE (6)
p =
_
E
v
E
bottom
D
V
(E) {1 f(E, T)} dE
_
E
v

D
V
(E) {1 f(E, T)} dE
(7)
To proceed further we need forms for D
C
and D
V
. Recall that in
8

e
m
*
E
g
E
c
E
v
E
k
h k
2 2
2m
n
*
E
k
h k
2 2
p
2m
*
Figure 9:
the parabolic approximation E
k

h
2
k
2
2m

we found that D(E) =


(2m

)
3
2
2
2
h
3

E. Thus,
D
C
(E) =
(2m

n
)
3
2
2
2
h
3

E E
C
(8)
D
V
(E) =
_
2m

p
_3
2
2
2
h
3

E
V
E (9)
for E > E
C
and E < E
V
respectively, and zero otherwise
E
V
< E < E
C
.
In an intrinsic (undoped) semiconductor n = p, and so E
F
must lie in the band gap. However, if m

n
= m

p
(ie. D
C
=
D
V
), then the chemical potential, E
F
, must be adjusted up or
down from the center of the gap so that n = p.
Furthermore, the carriers which are induced across the gap
are relatively high in energy, compared to k
B
T, since typically
9
E
g
= E
C
E
V
k
B
T.
E
g
(eV ) n
i
(cm
3
)(300

K)
Ge 0.67 2.4 10
13
Si 1.1 1.5 10
10
GaAs 1.43 5 10
7
Table 2:
1eV
k
B
10000

K 300

K T (10)
Thus, assuming that E E
F
>

E
g
2
k
B
T
1
e
(EE
F
)/k
B
T
+ 1

1
e
(EE
F
)/k
B
T
= e
(EE
F
)/k
B
T
(11)
ie., Boltzmann statistics. A similar relationship holds for holes
where (E E
F
)
>

E
g
2
k
B
T
1
1
e
(EE
F
)/k
B
T
+ 1
1
_
1 e
(EE
F
)/k
B
T
_
= e
(EE
F
)/k
B
T
(12)
since (1 f(E)) = f(E) and e
(EE
F
)/k
B
T
is small. Thus, the
concentration of electrons n
n
(2m

n
)
3
2
2
2
h
3
e
E
F
/k
B
T
_

E
C

E E
C
e
E/k
B
T
dE (13)
10
=
(2m

n
)
3
2
2
2
h
3
(k
B
T)
3
2
e
(E
C
E
F
)
_

0
x
1
2
e
x
dx (14)
= 2
_
_
2m

n
k
B
T
h
2
_
_
3
2
e
(E
C
E
F
)
= N
C
eff
e
(E
C
E
F
)
(15)
Similarly
p = 2
_
_
_
2m

p
k
B
T
h
2
_
_
_
3
2
e
(E
V
E
F
)
= N
V
eff
e
(E
V
E
F
)
(16)
where N
C
eff
and N
V
eff
are the partition functions for a classical
gas in 3-d and can be regarded as eective densities of states
which are temperature-dependent. Within this interpretation,
we can regard the holes and electrons statistics as classical. This
holds so long as n and p are small, so that the Pauli principle
may be ignored - the so called nondegenerate limit.
In general, in the nondegenerate limit,
np = 4
_
_
k
B
T
2 h
2
_
_
3
_
m

n
m

p
_3
2
e
E
g
(17)
this, the law of mass action, holds for both doped and intrinsic
semiconductor so long as we remain in the nondegenerate limit.
However, for an intrinsic semiconductor, where n = p, it
11
gives us further information.
n
i
= p
i
= 2
_
_
k
B
T
2 h
2
_
_
3
2 _
m

n
m

p
_3
4
e
E
g
/2
(18)
However, we already have relationships for n and p involving
E
C
and E
V
n = p = N
C
eff
e
(E
C
E
F
)
= N
V
eff
e
(E
V
E
F
)
(19)
e
2E
F
=
N
V
eff
N
C
eff
e
(E
V
+E
C
)
(20)
or
E
F
=
1
2
(E
V
+ E
C
) +
1
2
k
B
T ln
_
_
_
N
V
eff
N
C
eff
_
_
_ (21)
E
F
=
1
2
(E
V
+ E
C
) +
3
4
k
B
T ln
_
_
_
m

p
m

n
_
_
_ (22)
Thus if m

p
= m

n
, the chemical potential E
F
in a semiconductor
is temperature dependent. Recall that this T-dependence was
important for the transport of a semiconductor in the presence
of a thermal gradient T.
3 Doping of Semiconductors
= ne, so the conductivity depends linearly upon the doping
(it may also eect in some materials, leading to a non-linear
12
doping dependence). A typical metal has
n
metal
10
23
/(cm)
3
(23)
whereas we have seen that a typical semiconductor has
n
i
SeC

10
10
cm
3
atT 300

K (24)
Thus the conductivity of an intrinsic semiconductor is quite
small!
To increase n (or p) to 10
18
or more, dopants are used.
For example, in Si the elements used as dopants have either a
s
2
p
1
or s
2
p
3
atomic valence. Thus, in the tetrahedral bonding
of Si there is either an extra electron (half bond) or an un-
satised bond or a hole. Thus P or B will either donate or
Si
P
B
Si Si
Si
Si
Si Si
Si
Si
Si
Si Si
Si
Si
Si Si
Si
Si
Si
Si
Si Si Si Si
Si
Si
Si
Si Si
Si
r
e
+
e
-
Si 3s 3p
2 2
B 3s 3p
2 1
P 3s 3p
2 3
r =
2
h
2
*
m e
big!
Figure 10:
absorb additional electron (with the latter called the creation
13
of a hole). As in an exciton, these additional charges will be
localized around the donor or acceptor ion. The dierence is
that here the donor/acceptor is xed and may be treated as
having innite mass, thus the binding energy is given by
E =
m

e
4
2
2
h
2
n
2
(25)
m

=
_

_
hole mass acceptor(B)
electron mass donor(P)
(26)
Again, since
m

m
< 1 and 10 these energies are often much
less than 13.6eV c.f. in Si E 30MeV 300

K or in Ge
E GMeV 60

K. Thus thermal excitations will often


ionize these dopant sites. In terms of energy levels
E
F
E
V
E
C
E
V
E
C
E
F
E
D
E
A
n - SeC p - SeC
occupied at T= 0
unoccupied at T= 0
(occupied by holes at T = 0)
Figure 11:
14
4 Carrier Densities in Doped semiconductor
The law of mass action is valid so long as the use of Boltzmann
statistics is valid i.e., if the degeneracy is small. Thus, even for
doped semiconductor
np = N
C
eff
N
V
eff
e
E
g
= n
2
i
= p
2
i
(27)
Now imagine the temperature is nite so that some of the donors
or acceptors are ionized. Furthermore, in equilibrium, the semi-
E
V
E
C
E
D
E
F
E
A
D
N = N + N
D D
0 +
N = N + N
A A A
0 +
# un-ionized
# ionized
Figure 12:
conductor is charge neutral so that
n + N

A
= p + N
+
D
(28)
15
The probability that a donor/acceptor is occupied by an elec-
tron is determined by Fermi statistics
n
D
= N
0
D
= N
D
1
1 + e
(E
D
E
F
)
(29)
p
A
= N
0
A
= N
A
(1 f(E
A
)) = N
A
1
1 + e
(E
F
E
A
)
(30)
To provide a solvable example, imagine that we have an n-type
semiconductor (no p-type dopants) so that N
A
= N
0
A
= N
+
A
=
0, then
n = N
C
eff
e
(E
C
E
F
)
(31)
N
D
= N
0
D
+ N
+
D
(32)
N
0
D
= N
D
1
e
(E
D
E
F
)
+ 1
(33)
Furthermore, charge neutrality requires that
n = p + N
+
D
(34)
An excellent approximation is to assume that for a (commer-
cially) doped semiconductor
N
+
D
n
i
(35)
ie., many more carriers are provided by dopants than are ther-
mally excited over the entire gap, then as np = n
2
i
, it must be
16
that N
+
D
p so that
n N
+
D
= N
D
N
0
D
(36)
n N
D
_
_
1
1
e
(E
D
E
F
)
+ 1
_
_
(37)
If we recall that thermally induced carriers satisfy the Boltz-
mann equation,
n = N
C
eff
e
(E
F
E
C
)
(38)
we can eliminate E
F
in n (where E
d
= E
c
E
D
)
n =
N
D
1 + e
E
d
n/(N
C
eff
)
(39)
This quadratic equation has only one meaningful solution
n =
2N
D
1 +

1 + 4
_
N
D
/N
C
eff
_
e
E
d
(40)
At low T
E
d
k
B
n
_
N
D
N
C
eff
e
E
d
(41)
at higher T
E
d
k
B
n = N
D
(42)
At still higher T our approximation breaks down that N
+
D
n
since thermally excited carriers will dominate.
17

You might also like