Sunteți pe pagina 1din 417

Chapter 1: Chemical Bonding

Linus Pauling (1901–1994)

December 28, 2001

Contents

1 The development of Bands and their filling 4

2 Different 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
Periodic Table
1 18

H 1 He 2
1 Hydrogen 2 13 14 15 16 17 Helium
1.008 4.003

Li 3 Be 4 B 5 C 6 N 7 O 8 F 9 Ne 10
2 Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
6.941 9.012 10.811 12.011 14.007 15.999 18.998 20.180

Na 11 Mg 12 Al 13 Si 14 P 15 S 16 Cl 17 Ar 18
3 Sodium Magnesium 3 4 5 6 7 8 9 10 11 12 Aluminum Silicon Phosphorous Sulfur Chlorine Argon
22.990 24.305 26.982 28.086 30.974 32.066 35.453 39.948

K 19 Ca 20 Sc 21 Ti 22 V 23 Cr 24 Mn 25 Fe 26 Co 27 Ni 28 Cu 29 Zn 30 Ga 31 Ge 32 As 33 Se 34 Br 35 Kr 36
4 Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
39.098 40.078 44.956 47.88 50.942 51.996 54.938 55.847 58.933 58.69 63.546 65.39 69.723 72.61 74.922 78.96 79.904 83.80
2

Rb 37 Sr 38 Y 39 Zr 40 Nb 41 Mo 42 Tc 43 Ru 44 Rh 45 Pd 46 Ag 47 Cd 48 In 49 Sn 50 Sb 51 Te 52 I 53 Xe 54
5 Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
85.468 87.62 88.906 91.224 92.906 95.94 (98) 101.07 102.906 106.42 107.868 112.411 114.82 118.71 121.75 127.60 126.905 131.29

Cs 55 Ba 56 Lu 71 Hf 72 Ta 73 W 74 Re 75 Os 76 Ir 77 Pt 78 Au 79 Hg 80 Tl 81 Pb 82 Bi 83 Po 84 At 85 Rn 86
6 Caesium Barium Lutetium Halfnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury Thallium Lead Bismuth Polonium Astatine Radon
132.905 137.327 174.967 178.49 180.948 183.85 186.207 190.2 192.22 195.08 196.967 200.59 204.383 207.2 208.980 (209) (210) (222)

Fr 87 Ra 88 Lr 103
7 Francium Radium Lawrencium
(223) 226.025 (260)

La 57 Ce 58 Pr 59 Nd 60 Pm 61 Sm 62 Eu 63 Gd 64 Tb 65 Dy 66 Ho 67 Er 68 Tm 69 Yb 70
Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium
138.906 140.115 140.908 144.24 (145) 150.36 151.965 157.25 158.925 162.50 164.93 167.26 168.934 173.04

Ac 89 Th 90 Pa 91 U 92 Np 93 Pu 94 Am 95 Cm 96 Bk 97 Cf 98 Es 99 Fm 100 Md101 No 102


Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium
227.028 232.038 231.036 238.029 237.048 (244) (243) (247) (247) (251) (252) (257) (258) (259)
Solid state physics is the study of mainly periodic systems (or things
that are close to periodic) in the thermodynamic limit ≈ 1021 atoms/cm3 .
At first 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 filling

nl elemental solid
1s H,He
2s Li,Be
2p B→Ne
3s Na,Mg
3p Al→Ar
4s K,Ca
3d transition metals Sc→Zn
4p Ga→Kr
5s Rb,Sr
4d transition metals Y→Cd
5p In-Xe
6s Cs,Ba
4f Rare Earths (Lanthanides) Ce→Lu
5d Transition metals La→Hg
6p Tl→Rn
Table 1: Orbital filling scheme for the first 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
−Znl e2
V (r) = (1)
r
where Znl describes the effective 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 fill 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 filling these
orbitals systematically (as shown in Table 1) starting with the lowest
me4 Znl
2
energy states (where Enl = 2 2
2h̄ n
Note that for large n, the orbitals do not fill up simply as a func-
tion of n as we would expect from a simple Hydrogenic model with
mZ 2 e4
En = 2h̄2 n2
(with all electrons seeing the same nuclear charge Z). For
example, the 5s orbitals fill 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 effective Z5s is larger than Z4d . A
schematic atomic level structure, accounting for screening, is shown in
Fig. 2.
Now let’s 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
5d
V(r) z/r
4f atom
6s
6spdf
5p
4d
d
5s 5spdf
4p
3d
Ce Valence Shell
4s
3p
4spdf
+
3s
3spd 6s
2p
2s 4f 5d s
2sp
1s
1s
2
V(r) + C l (l+1)/r
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 different
principle quantum numbers can be close in energy. I.e., in elemental Ce, (4f 1 5d1 6s2 )
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 off ) with the 5d orbitals
more effectively. 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 modified when brought to-
gether to form a solid, consider a pair of isolated orbitals. If they are far
apart, each orbital has a Hamiltonian H0 = ²n, where n is the orbital
occupancy and we have ignored the effects of electronic correlations
(which would contribute terms proportional to n↑ 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 = ²n,
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 = ²(n1 + n2 ) − t(c†1 c2 + c†2 c1 ) . (2)

where c1 (c†1 ) destroys (creates) an electron on orbital 1. If we rewrite


this in matrix form
  
µ ¶ ² −t c1
  
H = c†1 , c†2 




 (3)
−t ² c2

then it is apparent that system has eigenenergies ² ± t. Thus the two


states split their degeneracy, the splitting is proportional to |t|, and
they remain centered at ²

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 ² (cf. Fig. 5). This, of course, is an oversimplification. Real

+ + + + . . . = Band

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 (t̃) from bringing the atoms closer
together, due to overlap of the cores. Once we have reached the optimal

8
point we fill 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 Different 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 different 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 (À a) closest packed unfilled 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 O2 with a single electron and
∇2 Ze2 Ze2 Z 2 e2
H=− − − + (4)
2m ra rb R
We will search for a variational solution to the the problem of the
molecule (HΨmol = EΨmol ), by constructing a variational wavefunction
from the atomic orbitals ψa and ψb . Consider the variational molecular
wavefunction
Ψ0 = c a ψa + c b ψb (5)
R
0 Ψ0∗ HΨ0
E = R 0∗ 0 ≥ E (6)
Ψ Ψ
The best Ψ0 is that which minimizes E 0 . We now define the quantum
integrals
Z Z Z
S= ψa∗ ψb Haa = Hbb = ψa∗ Hψa Hab = ψa∗ Hψb . (7)

Note that 1 > Sr > 0, and that Habr < 0 since ψa and ψb are bound
states [where Sr = ReS and Habr = ReHab ]b. With these definitions,
(c2a + c2b )Haa + 2ca cb Habr
E0 = (8)
c2a + c2b + 2ca cb Sr

10
∂E 0 ∂E 0 0
and we search for an extremum ∂ca = ∂cb = 0. From the first condition, ∂E
∂ca =

0 and after some simplification, and re-substitution of E 0 into the above


equation, we get the condition

ca (Haa − E 0 ) + cb (Habr − E 0 Sr ) = 0 (9)

∂E 0
The second condition, ∂cb = 0, gives

ca (Habr − E 0 S) + cb (Haa − E 0 ) = 0 . (10)

Together, these form a set of secular equations, with eigenvalues


Haa ± Habr
E0 = . (11)
1 ± Sr
Remember, Habr < 0, so the lowest energy state is the + state. If we
substitute Eq. 10 into Eqs. 8 and 9, we find that the + state corresponds

to the eigenvector ca = cb = 1/ 2; i.e. it is the bonding state.
1 Haa + Habr
Ψ0bonding = √ (ψa + ψb ) Ebonding
0
= . (12)
2 1 + Sr

For the −, or antibonding state, ca = −cb = 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 deficiency 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 configuration!). Energy is lost if there are
more electrons which must fill the antibonding states. Thus the covalent

11
e

ra
rb
R
Ze Ze
Ψanti-bonding
spin triplet

Ψbonding
spin singlet

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 effective 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 Pσ orbital can bond if both are in the same plane;

12
whereas an S and a Pπ orbital cannot. I.e., covalent bonds are di-
rectional! An excellent example of this is diamond (C) in which the

- Pσ
S Pπ S + -
+

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 first sight one might assume that C with a 1s2 2s2 2p2 configuration
could form only 2-bonds with the two electrons in the partially filled
p-shell. However, significant energy is gained from bonding, and 2s and
2p are close in energy (cf. Fig. 2) so that sufficient energy is gained
from the bond to promote one of the 2s electrons. A linear combination
of the 2s 2px , 2py and 2pz orbitals form a sp3 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
x2 − y 2 and 3z 2 − r2 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 x2 − y 2 and 3z 2 − r2
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).
z z z
d y d y d y
xz xy yz

x x x

Face Centered
z z Cubic Structure
d 2 2 y y
x-y d 2 2
3z - r

x x

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 x2 −y 2 and 3z 2 −r2 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
e2 ao
Ecohesive = = 5.19eV . (13)
ao rN a + rCl
Note that this does not agree with the experimental figure given in
the caption of Fig. 9. This is due to uncertainties in the definitions of
the ionic radii, and to oversimplification of the model. However, such
calculations are often sufficient 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, rCs ≈ 1.60 A, rCl ≈

1.81 A) this means the crystal structure will be the closest unfrustrated

15
Na + 5.14 eV e-
Na+ +

e-
Cl + Cl- + 3.61 eV

Na+ + Cl- Na+ Cl- + 7.9 eV

r Cl = 1.81

r Na = 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 figure are experimental. This figure 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 filled cores, a covalent bond including both bonding and anti-

16
bonding states would form. As discussed before, Coulomb repulsion
makes this energetically unfavorable.
Body Centered Face Centered
Cubic
Cubic Cubic

Cl

c
Cl c
b b
a a
Na Cs

Figure 10: Possible salt lattice structures. In the simple cubic and bcc lattices all the
nearest neighbors are of a different 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
e2 B
φij = ± + n (14)
rij rij
where the first 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. 66–71, approximates this heuristic term with an exponen-
tial, B exp (−rij /ρ), also with two free parameters). These are usually
determined from fits to experiment. If a is the separation of nearest
neighbors, rij = apij , and their are N sites in the system, then the total
potential energy may be written as
 
e2 X ± B X 1
Φ = N Φi = N −
 + n . (15)
a i6=j pij a i6=j pnij
P ±
The quantity A = i6=j pij , is known as the Madelung constant. A
depends upon the type of lattice only (not its size). For example
AN aCl = 1.748, and ACsCl = 1.763. Due to the short range of the
potential 1/pn , 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 filled valence bands. From the first 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
filling of space. In addition, since the bands are composed of partially

18
3d x2 - y2
4S

Figure 11: In metallic Ni (fcc, 3d8 4s2 ), 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 different 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.

filled orbitals, it is always possible to supply a small external electric


field 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 filled 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 final 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 fluctuations in the charge which
induce electric dipole moments.

n
P1 P2

x1
+ - - +
R x2

Figure 12: Noble gasses and molecules with saturated orbitals can form short ranged
van der Waals bonds by inducing fluctuating 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 fluctuation on 1 induces a dipole moment p1 . Then
dipole 1 exerts a field
3n(p1 · n) − p1
E1 = (16)
r3
which is felt by 2, which in turn induces a dipole moment p2 ∝ E1 ∝
1/r3 . This in turn, generates a dipole field E2 felt by 1 ∝ p2 /r3 ∝ 1/r6 .
Thus, the energy of the interaction is very small and short ranged.

W = −p1 · E2 ∝ 1/r6 (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 effects 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
x1 and x2 , as shown in Fig. 12. The particles oscillate along the x axis
with frequency ω0 (the strongest optical absorption line of the atom),
and momenta P1 and P2 . If we ignore the interaction between the
charges (other than the self-interaction between the dipole’s charges
which is accounted for in the harmonic oscillator potentials), then the
Hamiltonian of the system is
P12 + P22 1
H0 = + mω02 (x21 + x22 ) . (18)
2m 2

21
If we approximate each pair of charges as point dipoles, then they
interact with a Hamiltonian
−3(p2 · n)(p1 · n) + p1 · p2 −2p1 p2 2e2 x1 x2
H1 ≈ = − = − . (19)
|x1 + R − x2 |3 R3 R3
The total Hamiltonian H0 + H1 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
√ √
xs = (x1 + x2 )/ 2 xa = (x1 − x2 )/ 2 (20)
√ √
Ps = (P1 + P2 )/ 2 Pa = (P1 − P2 )/ 2 (21)

After these substitutions, the total Hamiltonian becomes


   
Ps2 + Pa2 1  2 2e2  2 1  2 2e2  2
H= + mω0 − 3 xs + mω0 + 3 xa (22)
2m 2 R 2 R
The new eigenfrequencies of these two modes are then
 1/2  1/2
2 2e2  2 2e2 
ωs = ω0 −
 
ωa = ω0 + (23)
mR3 mR3
The zero point energy of the system is now
  2 
1  1  2e2 
E0 = h̄(ωs + ωa ) ≈ h̄ω0 1 − 2 3
+ · · ·
 (24)
2 4 mω0 R

or, the zero point energy is lowered by the dipole interaction by an


amount  2
h̄ω0  2e2 
∆U ≈ (25)
4 mω02 R3

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 effect,
in the sense that ∆U → 0 as h̄ → 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 α

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
rn = 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 n1 , n2 , n3 are integers. In this way we may construct any periodic
structure.

1 Lattice Types and Symmetry

1.1 Two-Dimensional Lattices

These structures are classified 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 6= b and α is not a rational fraction of π. Notice that it is invari-
Square Rectangular

b b
a a

|a| = |b|, γ = π/2 |a| = |b|, γ = π/2

Hexangonal Centered

b
a
a
b

|a| = |b|, γ = π/3

Figure 2: Two dimensional lattice types of higher symmetry. These have higher
symmetry since some are invariant under rotations of 2π/3, or 2π/6, or 2π/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 π and 2π. 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
defined 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 Cu O Cu O Cu O Cu O

O O O O O Basis

Cu O Cu O Cu O Cu O
Primitive
O O O O
cell and
lattice
Cu O Cu O Cu O Cu O vectors
b
O O O O

a
Cu O Cu O Cu O Cu O

O O O O

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 CuO2 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

Body Centered Face Centered


Cubic
Cubic Cubic

c
c
b b
a a

a=x a = (x+y-z)/2 a = (x+y)/2


b=y b = (-x+y+z)/2 b = (x+z)/2
c=z c = (x-y+z)/2 c = (y+z)/2

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 different symmetry.
As an example, imagine that a symmetry operation R leaves H
invariant, so that

RHR−1 = H then [H, R] = 0 (2)


P
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
X X
< i|R|k >< k|H|j >= < i|H|k >< k|R|j > . (3)
k k

If we recall that Rik =< i|R|k >= Rii δik since |k > are eigenstates of

6
R, then Eq. 3 becomes

(Rii − Rjj ) Hij = 0 . (4)

So, Hij = 0 if Ri and Rj are different eigenvalues of R. Thus, when the


states are classified 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
X ±
. (5)
ij pij
This may be done by a brute force sum over the lattice, i.e.
X (−1)i+j
lim
n→∞
. (6)
i=−n,n j=−n,n pij
Or, we may realize that the lattice has some well defined operations
which leave it invariant. For example, this lattice in invariant under in-
version (x, y) → (−x, −y), and reflections 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
(a)

(b)

O 4
8
1
4

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 defined 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, reflections,


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!22 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!2D operations). In table. 1,
all of these operations are identified The reflections 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 Identification
(x, y) → (x, y) Identity
(x, y) → (x, −y) reflection about x axis
(x, y) → (−x, y) reflection about y axis
(x, y) → (−x, −y) inversion
(x, y) → (y, x) reflection about x = y
(x, y) → (y, −x) rotation by π/2 about z
(x, y) → (−y, −x) inversion-reflection
(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 Schönflies point group


symbol described in table. 2 As an example, consider the previous ex-

11
Symbol Meaning
Cj (j=2,3,4, 6) j-fold rotation axis
Sj j-fold rotation-inversion axis
Dj 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
Ci a center of inversion
Cs a mirror plane

Table 2: The Schönflies point group symbols. These give the classification according
to rotation axes and principle mirror planes. In addition, their are suffixes 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 π/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 C4v .

12
3 Simple Crystal Structures

3.1 FCC

Face Centered Cubic (FCC)


Close-packed planes Principle lattice vectors
z

b c y
a
x
a = (x+y)/2
b = (x+z)/2 3-fold axes
c = (y+z)/2
4-fold axes

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 Oh .

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 dxy , dyz ,
and dxz orbitals points to a near neighbor. The xz,xy,yz triplet form
rather strong bonds. The dx2 −y2 and d3z 2 −r2 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
z z z
d y d y d y
xz xy yz

x x x

z z
dx2- y2 y d 2 2 y
3z - r

x x

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 Oh . In fact, this same
argument also applies to the bcc and sc lattices, so Oh 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 D3h 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 five 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
3-fold
axis
mirror plane

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
Oh 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
FCC HCP

A A A A A A A A A A
B B B B B B B B B B
C C C C C
A A A A A A A A A A
B B B B B B B B B B
C C C C C
A A A A A A A A A A
B B B B B B B B B B
C C C C C
These spaces unfilled

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 significantly than an fcc (or hcp)
structure would. This increases the effective coordination number by
including the next nearest neighbor shell in the bonding (cf. Fig. 10).

17
2
12 6 24
fcc
1s
R(r) 8 6 12
1 bcc

2s,2p

0 1 2 3
o
r(A)

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 figure was lifted from I&L).

18
Chapter 3: The Classical Theory of Crystal
Diffraction

Bragg

December 29, 2001

Contents

1 Classical Theory of diffraction 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 Diffraction . . . . . . . . . . . . . 22

1
In the last two chapters, we learned that solids generally form pe-
riodic structures of different symmetries and bases. However, given a
solid material, how do we learn what its periodic structure is? Typ-
ically, this is done by diffraction, where we project a beam (of either
particles or radiation) at a solid with a wavelength λ ≈ the characteris-
tic length scale of the lattice ( a ≈ twice the atomic or molecular radii
of the constituents). Diffraction of waves and particles (with de Broglie

incident waves k0
or particles
K
λ ≈ | a 1 | or |a 2 | θ
k0 k
k
K = k - k0

a1
θ d
a2
d sin(θ)

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 θ = λ (the Bragg condition),
and yields a scattering maximum on a distant screen.

wavelength λ = h/p) of λ ≈ a allows us to learn about the periodic


structure of crystals. In a diffraction experiment one identifies 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 λ ≈ 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 4πnq 2 e2  mγ 2 v 3  q 2
≈− ln ∼ 2. (1)
dx mv 2 qeω0 v
As an example, consider a non-relativistic electron scattering into a

solid with a ≈ 2A. If we require that a = λ = h/p = 12.3×10−8 cm/ E
when E is measured in electron volts, then E ≈ 50eV. If we solve
Eq. 1 for the distance δx where the initial energy of the incident is lost
requiring that δE = E, when n ≈ 1023 /cm3 we find that δx ≈ 100A.
Thus, if λ ≈ a, the electrons do not penetrate into the bulk of the
sample (typically the first few hundred A of 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 diffraction, 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 λ ≈ a

1 Classical Theory of diffraction

In this theory of diffraction 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 diffracted 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 diffraction pattern
varies quickly with scattering angle θ so that the angular depen-
dence of a unpolarized dipole, 1 + (cos θ)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 Fermi’s golden rule (first-order Born approxi-
mation).

The basic setup of a scattering experiment is sketched in Fig. 3.


Generally, we will also assume that |R| À |r|, so that we may always
approximate the amplitude of the incident waves on the target as plane
waves.
AP = AO ei(k0 ·(R+r)−ω0 t) . (2)

5
R’ - r
source r B
R
R’ observer
Q P or screen
target

Figure 3: Basic setup of a scattering experiment.

Then, consistent with the second assumption above,


0
0
Z
3 eik·(R −r)
AB (R ) ∝ d rAP ρ(r) 0 , (3)
|R − r|
which, after substitution of Eq.2, becomes
0
Z e−i(k−ko )·r
AB (R0 ) ∝ AO ei(k0 ·R+k·R −ω0 t) d3 rρ(r) . (4)
|R0 − r|
At very large R0 (ie. in the radiation or far zone)
0
0 AO ei(k0 ·R+k·R −ω0 t) Z 3
AB (R ) ∝ 0
d rρ(r)e−i(k−ko )·r . (5)
R
Or, in terms of the scattered intensity IB ∝ |AB |2

|AO |2 ¯¯¯Z 3 ¯2
−i(k−ko )·r ¯¯
IB ∝ ¯ d rρ(r)e ¯ . (6)
R02
The scattering intensity is just the absolute square of the Fourier trans-
form of the density of scatterers. If we let K = k − k0 (cf. Fig. 1), then
we get

|AO |2 ¯¯¯Z 3 ¯2
−iK·r ¯¯ |AO |2
IB (K) ∝ 02 ¯ d rρ(r)e ¯ = 02
|ρ(K)|2 . (7)
R R

6
From the associated Fourier uncertainty principle ∆k∆x ≈ π, 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) ρ(r)

Figure 4: Since the measured scattering intensity I(K) ∝ |ρ(K)|2 the complex phase
information is lost. Thus, a scattering experiment does not provide enough informa-
R d3 r
tion to invert the transform ρ(r) = (2π)3
ρ(K)e+iK·r .

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
Z
ρ(K) = d3 rρ(r)e−iK·r . (8)

Of course, if the experiment just measures I ∝ |ρ(K)|2 , then we lose


the relative phase information (i.e. ρ(K) = ρK eiθK so that I ∝ |ρK |2 ,
and the phase information θK 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 ρ(r) by inverting the Fourier transform. Instead
experimentalists analyze their data by proposing a feasible model struc-
ture (i.e. a ρ(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 fit.

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

ρ(r + rn ) = ρ(r) where rn = n1 a1 + n2 a2 + n3 a3 (9)

for integer n1 , n2 , n2 . This also implies that the Fourier coefficients of ρ


will be chosen from a discrete set. For example, consider a 1-d periodic
structure

ρ(x + na) = ρ(x) . (10)

Then we must choose the Gn


X
ρ(x) = ρn eiGn x , (11)
n

8
so that
X X
ρ(x + ma) = ρn eiGn (x+ma) = ρn eiGn (x) eiGn ma
n n
X iGn (x)
= ρn e = ρ(x) , (12)
n

I.e. eiGn ma = 1, or Gn = 2nπ/a where n is an integer.


This may be easily generalized to three dimensions, for which
X
ρ(r) = ρG eiG·r (13)
G

where the condition of periodicity ρ(r + rn ) = ρ(r) means that

G · rn = 2πm 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 = hg1 + kg2 + lg3 . (15)

The condition of periodicity then requires that

(hg1 + kg2 + lg3 ) · n1 a1 = 2πm m ∈ Z (16)

with similar conditions of the other principle lattice vectors a2 and a3 .


Since g1 , g2 and g3 are not parallel or coplanar, the only way to satisfy
this constraint for arbitrary n1 is for

g1 · a1 = 2π g2 · a1 = g3 · a1 = 0 (17)

9
or some other permutation of 1 2 and 3, which would just amount to
a renaming of g1 , g2 , and g3 . The set (g1 , g2 , g3 ) are called the basis
set for the reciprocal lattice. They may be constructed from
a2 × a 3
g1 = 2π plus cyclic permutations . (18)
a1 · (a2 × a3 )
It is easy to see that this construction satisfies 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 lattice1 .

2.1 The Scattering Intensity for a Crystal

Lets now apply this form for the density


X
ρ(r) = ρG eiG·r (19)
G

to our formula for the scattering intensity


¯ ¯2
|AO |2 ¯¯¯Z 3 X ¯
−i(K−G)·r ¯¯
IB (K) ∝ ¯ d r ρG e ¯ (20)
R02 ¯ G ¯

The integral above is simply





 V if G = K
V δG,K =  , (21)

 0 if G 6= K
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 Oh point group

10
where V is the lattice volume, so
|AO |2
IB (K) ∝ 02
|ρG |2 V 2 δG,K (22)
R
This is called the Laue condition for scattering. The fact that this is
proportional to V 2 rather than V just indicates that the diffractions
spots, in this approximation, are infinitely bright (for a sample in the
thermodynamic limit). Of course, this is because the spots are infinitely
narrow or fine. When real broadening is taken into account, IB (K) ∝ V
as expected.
Then as G = hg1 + kg2 + lg3 , we can label the spots with the three
integers (h, k, l ), or
Ihkl ∝ |ρhkl |2 . (23)

Traditionally, negative integers are cabled with an overbar, so −h → h.


Then as ρ(r) is real, ρG = ρ−G , or

Ihkl = Ihkl Friedel’s rule (24)

Most scattering experiments are done with either a rotating crystal, or


a powder made up of many crystalites. For these experiments, Friedel’s
rule has two main consequences

• For every spot at k − k0 = G, there will be one at k0 − k0 =


−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
|k0 |, if the crystal is rotated by π 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 fixed,
and the joint rotated over all angles), or by holding the crystal fixed
and moving the source and diffraction 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 (a1 , a2 , a3 ), 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

h0 = 1/m k 0 = 1/n l0 = 1/o . (25)

Since these typically are not integers, they are multiplied by p, the

12
h’k’l’ plane

d h’k’l’ Ghkl
γ
a1

a3
a2
O a1
(211) plane

Figure 5: Miller indices identification 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 (a1 , a2 , a3 ), 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 h0 = 1/m k 0 = 1/n l0 = 1/o. Since these typically are not integers,
they are multiplied by p, the smallest integer such that p(h0 , k 0 , l0 ) = (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(h0 , k 0 , l0 ) = (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 Ghkl lies perpendicu-
lar to the (h, k, l) plane, and that the length between adjacent parallel
planes dhkl = 2π/Ghkl . To show this, note that the plane may be defined

13
by two non-parallel vectors v1 and v2 within the plane. Let

v1 = ma1 − na2 = a1 /h0 − a2 /k 0 v2 = oa3 − na2 = a3 /l0 − a2 /k 0 . (27)

Clearly the cross product, v1 × v2 is perpendicular to the (h, k, l) plane


a3 × a 1 a1 × a 2 a2 × a 3
v1 × v 2 = − − − . (28)
h0 l 0 h0 k 0 k 0 l0
If we multiply this by −2πh0 k 0 l0 /a1 · (a2 × a3 ), we get
−2πh0 k 0 l0 v1 × v2
=
a1 · (a2 × a3 )
 
2πp  0 a3 × a1 a 1 × a 2 a 2 × a 3
k + l0 + h0  =
p a1 · (a2 × a3 ) a1 · (a2 × a3 ) a1 · (a2 × a3 )
Ghkl /p (29)

Thus, Ghkl ⊥ to the (h, k, l) plane. Now, if γ is the angle between a1


and Ghkl , then the distance dh0 k0 l0 from the origin to the (h0 , k 0 , l0 ) plane
is given by
|a1 |a1 · Ghkl 2πh 2πp
dh0 k0 l0 = m|a1 | cos γ = = = (30)
h0 |a1 ||Ghkl | h0 Ghkl Ghkl
Then, as there are p planes in this distance (cf. Fig. 5), the distance to
the nearest one is
dhkl = dh0 ,k0 ,l0 /p = 2π/Ghkl (31)

With this information, we can reexamine the Laue scattering con-


dition K = k − ko = Ghkl , and show that it is equivalent to the
more intuitive Bragg condition. Part of the Laue is condition is that
|K| = K = |k − k0 | = Ghkl , now

K = 2k0 sin θ = sin θ and Ghkl = 2π/dhkl (32)
λ
14
thus, the Laue condition implies that

1/dhkl = 2 sin θ/λ or λ = 2dhkl sin θ (33)

which is the Bragg condition. Note that the Laue condition is more

Laue Condition Bragg Condition


(in reciprocal space) (in real space)
K= k - k0 = Ghkl 2dhkl sin(θ) = λ
d hklsin(θ)

Ghkl
k θ
k
d hkl
k θ k0

k0 hkl plane

Figure 6: Comparison of the Bragg λ = 2dhkl sin θ and Laue Ghkl = Khkl 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 defined by k and k0 is perpendicular to the
scattering planes (cf. Fig. 6).

15
O

O
Cu Cu Cu
Body Centered
Cubic
O O O
O

O
Cu Cu Cu
basis
O O O
r
O

O
Cu Cu
rα O
O O O
rn

Figure 7: Examples of lattices with non-trivial bases. The CuO2 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 diffraction pattern for a periodic


lattice ignoring the fine structure of the molecular of the basis. Exam-
ples of non-trivial molecular bases are shown in Fig. 7. Clearly the basis
structure will effect the scattering (Fig. 8). For example, there will be
interference from the scattering off 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 finite 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 different elements of the basis. The structure factors
account for these interference effects. The information about this in-
terference, and the basis structure is contained in the atomic scattering
factor f and the structure factor S.
O

Cu
R
O

Figure 8: Rays scattered from different elements of the basis, and from different 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

Ihkl ∝ |ρhkl |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
1 Z 1 X Z
ρhkl = d3 rρ(r)e−iGhkl ·r = d3 rρ(r)e−iGhkl ·r
V V cells
cell
1 X Z
= d3 rρ(r)e−iGhkl ·(r+rn ) (35)
V n1 ,n2 ,n3 cell

17
where the location of each cell is given by rn = n1 a1 + n2 a2 + a3 a3 .
Then since Ghkl · rn = 2πm, m ∈ Z,
N Z
ρhkl = d3 rρ(r)e−iGhkl ·r (36)
V cell
N
where N is the number of cells and V = 1/Vc , Vc 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 α labeles the different
elements of the basis, each with density ρα (r0 )
1 X −iGhkl ·rα Z 3 0 0
ρhkl = e d r ρα (r0 )e−iGhkl ·r (37)
Vc α
The atomic scattering factor f and the structure factor may then be
defined as parts of this integral
Z
0
fα = d3 r0 ρα (r0 )e−iGhkl ·r (38)

so
1 X −iGhkl ·rα Shkl
e
ρhkl = fα = (39)
Vc α Vc
fα describes the interference of spherical waves emanating from different
points within the atom, and Shkl 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
Z Z
3 0 0 −iGhkl ·r0 0
fα = d r ρα (r )e =− r02 dr0 d(cos θ)dφ ρα (r0 )e−Ghkl r cos θ

18
Z
02 0 sin Ghkl r0
0
= 4π r dr ρ(r ) . (40)
Ghkl r0
As an example, consider a spherical atom of charge Ze− , radius R, and
charge density
3Z
ρα (r0 ) = 3
θ(R − r0 ) (41)
4πR
then
Z R 0
3Z 02 0 sin Ghkl r
fα = r dr
R3 0 Ghkl r0
Z
= − (sin (Ghkl R) − (Ghkl R) cos (Ghkl R)) (42)
(Ghkl R)3
This has zeroes whenever tan (Ghkl R) = Ghkl R and a maximum when
Ghkl = 0, or in terms of the scattering angle 2k0 sin θ = Ghkl , when
θ = 0, π. In fact, we have that fα (θ = 0) = Z. This is true in general,
since
Z
fα (θ = 0) = fα (Ghkl = 0) = d3 r0 ρα (r0 ) = Z . (43)

Thus, for x-ray scattering I ∝ Z 2 . For this reason, it is often difficult


to see small-Z atoms with x-ray scattering.

2.3.1 The Structure Factor of Centered Lattices

Now let’s 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 |a1 | = |a2 | = |a3 |, a1 ⊥ a2 ⊥ a3 and a two-atom basis
rα = 0.5α(a1 + a2 + a3 ), α = 0, 1. Now if both sites in the unit cell

19
(α = 0, 1) contain the same atom with the same scattering factor f ,
then

rα · Ghkl = 0.5α(a1 + a2 + a3 ) · (hg1 + kg2 + lg3 ) = πα(h + k + l) (44)

so that
X
Shkl = f e−iπα(h+k+l)
α=0,1



 0 if h + k + l is odd
= f (1 + e−iπ(h+k+l) ) = 

 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 fCs+ 6= fCl− . In
fact, to a good approximation fCs+ ≈ fXe and fCl− ≈ fAr . 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 significantly 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 (ZCo = 27, and ZFe = 26)
x−ray x−ray
fFe ≈ fCo . (46)

However, since one has a closed nuclear shell and the other doesn’t,
their neutron scattering factors will be quite different

neutron neutron
fFe 6= fCo (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 Diffraction

If you expose a columnated beam of x-rays to a crystal with a single


crystalline domain, you usually will not achieve a diffraction spot. The
reason why can be seen from the Ewald construction, shown in Fig. 10
For any given k0 , the chances of matching up so as to achieve G = k−k0
are remote. For this reason most people use powdered x-ray diffraction
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|k0 | of the origin of
the reciprocal lattice.

22
Ghkl
Ewald Sphere
k

k0
O

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 k0 to the origin. From the base of k0 , spin k (remember, that for elastic
scattering |k| = |k0 |) 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 − k0 . In the example above we find 8 Bragg peaks. If however, we change k0 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 diffraction 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 first glance, a theory of lattice vibrations would appear impossibly


daunting. We have N ≈ 1023 ions interacting strongly (with energies of
about (e2 /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
¿1 (1)
M
which allows us to derive an accurate theory.
Due to Newton’s third law, the forces on the ions and electrons are
comparable F ∼ e2 /a2 , 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 ∼ e2 /a2 ∼ mωelectron
2 2
a ∼ M ωion a (2)

This means that


à !1/2
ωion m
∼ ∼ 10−3 to 10−2 (3)
ωelectron M
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 R1 , · · · RN and deter-


mine the electronic ground state energy, E(R1 , · · · RN ). 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
n’th unit cell

atom α

rα sα

rn
rnα Rnα = r n + rα + sα
O

Figure 1: Nomenclature for the lattice vibration problem. sn,α is the displacement of
the atom α within the n-th unit cell from its equilibrium position, given by rn,α =
rn + rα , where as usual, rn = n1 a1 + n2 a2 + n3 a3 .

Thus the potential energy for the ions

φ(R1 , · · · RN ) = E(R1 , · · · RN ) + the ion-ion interaction (4)

We will define the zero potential such that when all Rn are at their
equilibrium positions, φ = 0. Then
X Pn2
H= + φ(R1 , · · · RN ) (5)
n 2M
Typical lattice vibrations involve small atomic excursions of the or-
der 0.1A or smaller, thus we may expand about the equilibrium position
of the ions.
∂φ 1 ∂ 2φ
φ({rnαi + snαi }) = φ({rnαi }) + snαi + snαi smβj (6)
∂rnαi 2 ∂rnαi ∂rmβj
The first two terms in the sum are zero; the first by definition, and
the second is zero since it is the first derivative of a potential being

5
evaluated at the equilibrium position. We will define the matrix
∂ 2φ
Φmβj
nαi = (7)
∂rnαi ∂rmβj
From the different conservation laws (related to symmetries) of the
system one may derive some simple relationships for Φ. We will discuss
these in detail later. However, one must be introduced now, that is,

rα sα

Figure 2: Since the coefficients of potential between the atoms linked by the blue lines
(m−n)βj
(or the red lines) must be identical, Φmβj
nαi = Φ0αi .

due to translational invariance.

(m−n)βj ∂ 2φ
Φmβj
nαi = Φ0αi = (8)
∂r0αi ∂r(n−m)βj
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
∂φ({rmβj + smβj })
Fnαi = − (9)
∂snαi
so that the equation of motion is

−Φmβj
nαi smβj = Mα s̈nαi (10)

If there are N unit cells, each with r atoms, then this gives 3N r equa-
tions of motion. We will take advantage of the periodicity of the lattice
by using Fourier transforms to achieve a significant 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
1
snαi = √ uαi (q)ei(q·rn −ωt) (11)

where the first two terms on the rhs serve as the polarization vector
for the oscillation, uαi (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
1
ω 2 uαi (q) = q Φmβj
nαi e
iq·(rm −rn )
uβj (q) sum repeated indices .
Mα Mβ
(12)

7
t=0

t = ∆t

t = 2 ∆t

Figure 3: uαi (q) is independent of n so that a lattice vibration can propagate and
respect the translational invariant of the lattice.

(m−n)βj
Recall that Φmβj
nαi = Φ0αi so that if we identify

βj 1 1
Dαi = q Φmβj
nαi e
iq·(rm −rn )
= q Φpβj
0αi e
iq·(rp )
(13)
Mα Mβ Mα Mβ
where rp = rm − rn , then the equation of motion becomes

βj
ω 2 uαi (q) = Dαi uβj (q) (14)

or
µ ¶
βj βj
Dαi − ω 2 δαi uβj (q) = 0 (15)
³ ´
which only has nontrivial (u 6= 0) solutions if det D(q) − ω 2 I = 0.
For each q there are 3r different solutions (branches) with eigenvalues
ω (n) (q) (or rather ω (n) (q) are the root of the eigenvalues). The de-
pendence of these eigenvalues ω (n) (q) on q is known as the dispersion
relation.

8
basis
M1 M2 α=1 α=2 f

Figure 4: A linear chain of oscillators composed of a two-element basis with different


masses, M1 and M2 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 different masses, M1 and M2 and equal strength springs with
spring constant f . It has the potential energy
1 X
φ = f (sn,1 − sn,2 )2 + (sn,2 − sn+1,1 )2 . (16)
2 n
We may suppress the indices i and j, and search for a solution
1
snα = √ uα (q)ei(q·rn −ωt) (17)

to the equation of motion
1
ω 2 uα (q) = Dαβ uβ (q) where Dαβ = q Φp,β
0α e
iq·(rp )
(18)
Mα Mβ
and,
∂ 2φ
Φm,β
n,α = (19)
∂r0,α ∂r(n−m),β
³ ´
where nontrivial solutions are found by solving det D(q) − ω 2 I = 0.
The potential matrix has the form

Φn,1 n,2
n,1 = Φn,2 = 2f (20)

9
Φn,2 n,1 n−1,2
n,1 = Φn,2 = Φn,1
n+1,1
= Φn,2 = −f . (21)

This may be Fourier transformed on the space index n by inspection,


so that
1
Dαβ = q Φpβ
0α e
iq·(rp )
Mα Mβ
 ³ ´ 
2f
 M1 − √Mf1 M2 1+e −iqa

=   (22)
 ³ ´ 
− √Mf1 M2 1+e +iqa 2f
M2

Note that the matrix D is hermitian, as it must be to yield real, physi-


cal, eigenvalues ω 2 (however, ω can still be imaginary if ω 2 is negative,
³ ´
indicating an unstable mode). The secular equation det D(q) − ω 2 I =
0 becomes
à !
4 1 21 4f 2
ω − ω 2f + + sin2 (qa/2) = 0 , (23)
M1 M2 M1 M2
with solutions
v
à ! uà !2
1 1 u 1 1 4
ω2 = f + ± ft + − sin2 (qa/2) (24)
M1 M2 M1 M2 M1 M2
This equation simplifies significantly in the q → 0 and q/a → π limits.
µ ¶
1 1
In units where a = 1, and where the reduced mass 1/µ = M1 + M2 ,
v v
u u
fµ u
t
u 2f
t
lim ω− (q) = qa lim ω+ (q) = (25)
q→0 2M1 M2 q→0 µ
and
q q
ω− (q = π/a) = 2f /M2 . ω+ (q = π/a) = 2f /M1 (26)

10
2.0
optical mode
1.5
acoustic mode ω+
1.0
ω

ω-
0.5 ω ~ ck

0.0
-4 -2 0 2 4
q
Figure 5: Dispersion of the linear chain of oscillators shown in Fig. 4 when M 1 = 1,
M2 = 2 and f = 1. The upper branch ω+ is called the optical and the lower branch is
the acoustic mode.

As a result, the + mode is quite flat; whereas the − mode varies from
zero at the Brillouin zone center q = 0 to a flat 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. Let’s look at the optical mode at q = 0,
q
ω+ (0) = 2f /µ. Here,
 

 2f /M1 −2f / M1 M2 
D=
 √ 
 . (27)
−2f / M1 M2 2f /M2

11
Eigenvectors are non-trivial solutions to (ω 2 I − D)u = 0, or
  

 2f /µ − 2f /M1 2f / M1 M2  u1 
0=
 √ 


 . (28)
2f / M1 M2 2f /µ − 2f /M2 u2
q
with the solution u1 = − M2 /M1 u2 . In terms of the actual displace-
ments Eqs.11 v
u
sn1 u M 2 u1
=t (29)
sn2 M 1 u2
or sn1 /sn2 = −M2 /M1 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 fields with λ ∼ 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 ω 2 is real. Recall
³ ´
that the dispersion is determined by the secular equation det D(q) − ω 2 I =
0, so if D is hermitian, then its eigenvalues, ω 2 , must be real.

∗βj 1
Dαi = q Φpβj
0αi e
−iq·(rp )
(30)
Mα Mβ
1 −p,β,j iq·(rp )
= q Φ0,α,i e (31)
Mα Mβ
Then, due to the symmetric properties of the second derivative

∗βj 1 0,α,i 1
Dαi = q Φ−p,β,j eiq·(rp ) = q Φp,α,i
0,β,j e
iq·(rp ) αi
= Dβj (32)
Mα Mβ Mα Mβ

13
Thus, DT ∗ = D† = D so D is hermitian and its eigenvalues ω 2 are
real. This means that either ω 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
1
snαi = √ uαi (q)ei(q·rn −ωt) (33)

which is a plane wave. Suppose that the plane wave is moving to the
right so that q = x̂qx , then the plane of stationary phase travels to the
right with
ω
x=t. (34)
qx
Clearly then changing the sign of qx 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

ω(−q) = ω(q) . (35)

αi
Note that this is fully equivalent to the statement that Dβj (q) =
∗αi
Dβj (−q) which is clear from the definition of D.
Now, return to the secular equation, Eq. 15.
µ ¶
βj 2 βj
Dαi (q) −ω (q)δαi ²βj (q) = 0 (36)

14
Lets call the (normalized) eigenvectors of this equation ². They are the
elements of a unitary matrix which diagonalizes D. As a result, they
have orthogonality and completeness relations
X ∗(n) (m)
²α,i (q)²α,i (q) = δm,n orthogonality (37)
α,i
X ∗(n) ∗(n)
²α,i (q)²β,j (q) = δα,β δi,j (38)
n
If we now take the complex conjugate of the secular equation
µ ¶
βj βj
Dαi (−q) − ω 2 (−q)δαi ²∗βj (q) = 0 (39)

Then it must be that


²∗βj (q) ∝ ²βj (−q) . (40)

Since the {²} are normalized the constant of proportionality may be


chosen as one
²∗βj (q) = ²βj (−q) . (41)

Point-Group Symmetry and the Dispersion A point group operation


takes a crystal back to an identical configuration. Both the original
and final 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 definition of D
βj 1
Dαi (q) = q Φpβj
0αi e
iq·(rp )
(42)
Mα Mβ

15
βj βj
it is easy to see that Dαi (q + G) = Dαi (q) (since G · rp = 2πn, where
n is an integer). I.e., D is periodic in k-space, and so its eigenvalues
(and eigenvectors) must also be periodic.

ω (n) (k + G) = ω (n) (k) (43)

²βj (k + G) = ²βj (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 s1,1,i , then the lattice energy is unchanged.

tional invariance of Φ . Suppose we make an overall shift of the lattice


by an arbitrary displacement sn,α,i for all sites n and elements of the
basis α (i.e. sn,α,i = s1,1,i ). Then, since the interaction is only between

16
ions, the energy of the system should remain unchanged.
1 X m−n,β,j
δE = Φ0,α,i sn,α,i sm,β,j = 0 (45)
2 m,n,α,β,i,j
1 X m−n,β,j
= Φ0,α,i s1,1,i s1,1,j (46)
2 mnα,β,i,j
1X X m−n,β,j
= s1,1,i s1,1,j Φ0,α,i (47)
2 i,j mnα,β

Since we know that s1,1,i is finite, it must be that


X m−n,β,j X
Φ0,α,i = Φp,β,j
0,α,i = 0 (48)
m,n,α,β p,α,β

Now consider a strain on the system Vm,β,j , described by the strain


matrix mα,i
β,j
X
Vm,β,j = mα,i
β,j sm,α,i (49)
α,i

After the stress has been applied, the atoms in the bulk of the sample

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
X
0 = F0,α,i = − Φm,β,j γ,k
0,α,i mβ,j sm,γ,k (50)
m,β,j,γ,k

Since this applies for an arbitrary strain matrix mγ,k


β,j , the coefficients

for each mγ,k


β,j must be zero

X
Φm,β,j
0,α,i sm,γ,k = 0 (51)
m

An alternative way (cf. Callaway) to show this is to recall that the


reflection symmetry of the lattice requires that Φm,β,j
0,α,i be even in m;

whereas, sm,γ,k is odd in m. Thus the sum over all m yields zero.
Now let’s apply these constraints to D for an elemental lattice where
r = 1, and we may suppress the basis indices α.
1 X
Dij (q) = Φp,j
0,i e
iq·(rp )
(52)
M p

For small q we may expand D


à !
1 X 1
Dij (q) = Φp,j
0,i 1 + iq · (rp ) − (q · (rp ))2 + · · · (53)
M p 2
We have shown above that the first two terms in this series are zero.
Thus,
1 X 2
Dij (q) ≈ − Φp,j
0,i (iq · (rp )) (54)
2M p

Thus, the leading order (small q) eigenvalues ω 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 ωs (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 finite-sized periodic


system is that we now have a discrete sum rather than an integral over
k. Consider a one-dimensional finite system with N atoms and periodic
boundary conditions. We seek solutions to the phonon problem of the
type
sn = ²(q)ei(qrn −ωt) where rn = na (55)

and we require that


sn+N = sn (56)

or
q(n + N )a = qna + 2πm where m is an integer (57)

Then, the allowed values of q = 2πm/N a. 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-
fine 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 − k0 = Ghkl lie on the perpendicular bisector of Ghkl . The
smallest polyhedron centered at the origin of the reciprocal lattice and enclosed by
perpendicular bisectors of the G’s is called the first Brillouin zone.

over k may then be confined to this region. This region is defined as


the smallest polyhedron centered at the origin of the reciprocal lattice
and enclosed by perpendicular bisectors of the G’s is called the Bril-
louin zone (cf. Fig. 9). Typically, we choose to include only half of the
bounding surface within the first Brillouin zone, so that it can also be
defined 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 k − k0 = Ghkl lie on the perpendicular bisector of Ghkl . 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 first is the use of the point group symmetry of the system. It
is clear from their definition 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 1/23 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 first and second quantize
the system.
Any lattice displacement may be expressed as a sum over the eigen-

22
vectors of the dynamical matrix D.
1 X
sn,α,i = √ Qs (q, t)²sα,i (q)eiq·rn (58)
Mα N q,s

Recall that ²sα,i (q) are distinguished from usα,i (q) only in that they are
normalized. Also since q + G is equivalent to q, we need sum only over
the first Brillouin zone. Finally we will assume that Qs (q, t) contains
the harmonic time dependence and since sn,α,i is real Q∗s (q) = Qs (−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
1 X
T = Mα (ṡnα,i )2 (59)
2 n,α,i
1 X X
= Q̇r (q)²rα,i (q)eiq·rn Q̇s (k)²sα,i (k)eik·rn (60)
2N n,α,i q,k,r,s
Then as
1 X i(k+q)·r X r ∗s
e n
= δk,−q and ²α,i ²α,i = δrs (61)
N n α,i
the kinetic energy may be reduced to
1 X ¯¯ ¯2
¯
T = ¯Q̇r (q)¯ (62)
2 q,r
The potential energy may be rewritten in a similar fashion
1 X
V = Φm,β,j sn,α,i sm,β,j
2 n,m,α,β,i,j n,α,i
m−n,β,j
1 X Φ0,α,i
= q
2 n,m,α,β,i,j N Mα Mβ
X
Qs (q, t)²sα,i (q)eiq·rn Qr (k, t)²rβ,j (k)eik·rm (63)
q,k,s,r

23
Let rl = rm − rn
l,β,j
1 X Φ0,α,i
V = q
2 n,l,α,β,i,j N Mα Mβ
X
Qs (q, t)²sα,i (q)eiq·rn Qr (k, t)²rβ,j (k)eik·(rl +rn ) (64)
q,k,s,r

and sum over n to obtain the delta function δk,−q so that


1 X 1
V = Qs (−k)²sα,i (−k)Qr (k)²rβ,j (k) q Φl,β,j
0,α,i e
ik·rl
. (65)
2 l,α,β,i,j,s,r Mα Mβ
Note that the sum over l on the last three terms yields D, so that
1 X β,j
V = Dα,i (k)Qs (−k)²sα,i (−k)Qr (k)²rβ,j (k) . (66)
2 l,α,β,i,j,s,r
P βj
Then, since β,j Dαi (k)²rβj (k) = ωr2 (k)²rα,i (k) and ²sα,i (−k) = ²∗s
α,i (k),

1 X r
V = ²α,i (k)²∗s 2 ∗
α,i (k)ωr (k)Qs (k)Qr (k) (67)
2 α,i,k,r,s
P r ∗s
Finally, since α,i ²α,i (k)²α,i (k) = δr,s
1X 2
V = ωs (k) |Qs (k)|2 (68)
2 k,s
Thus we may write the Lagrangian of the ionic system as
1 X µ¯¯ ¯2
¯ 2 2

L=T −V = ¯Q̇s (k)¯ − ωs (k) |Qs (k)| , (69)
2 k,s
where the Qs (k) may be regarded as canonical coordinates, and
∂L
Pr∗ (k) = = Q̇∗s (k) (70)
∂Qr (k)
(no factor of 1/2 since Q∗s (k) = Qs (−k)) are the canonically conjugate
momenta.

24
The equations of motion are
 
d  ∂L  ∂L
− ∗
or Q̈s (k) + ωs2 (k)Qs (k) = 0 (71)
dt ∂ Q̇s (k)
∗ ∂Qs (k)
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 {Qs (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


{qi , pi }

2. These have Poisson Brackets


à !
X ∂u ∂v ∂u ∂v
{{u, v}} = − (72)
i ∂qi ∂pi ∂pi ∂qi
{{qi , pj }} = δi,j {{pi , pj }} = {{qi , qj }} = 0 (73)

3. Then define the quantum Poisson Bracket (the commutator)

[u, v] = uv − vu = ih̄{{u, v}} (74)

4. In particular, [qi , pj ] = ih̄δi,j , and [qi , qj ] = [pi , pj ] = 0.

25
Thus, following Dirac, we may now quantize the normal coordinates

[Q∗r (k), Ps (q)] = ih̄δk,q δr,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
 
1 q i
as (k) = √  ωs (k)Qs (k) + q Ps (k) (76)
2h̄ ωs (k)
 
1 q i
a†s (k) = √  ωs (k)Q∗s (k) − q Ps∗ (k) , (77)
2h̄ ωs (k)
or
v
u
u h̄ ³ ´
Qs (k) = t as (k) + a†s (−k) (78)
2ωs (k)
v
u
u h̄ωs (k) ³ ´
t
Ps (k) = −i as (k) − a†s (−k) (79)
2
Where

[as (k), a†r (q)] = δr,s δq,k [as (k), ar (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
à !
X 1
H= h̄ωs (k) a†s (k)as (k) + (81)
k,s 2
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

ns (k) = a†s (k)as (k) (82)

26
and a†s (k) and as (k) create and destroy phonons respectively, in the
state (k, s)
q
a†s (k) |ns (k)i = ns (k) + 1 |ns (k) + 1i (83)
q
as (k) |ns (k)i = ns (k) |ns (k) − 1i (84)

If |0i is the normalized state with no phonons present, then the state
with {ns (k)} phonons in each state (k, s) is
 1 
 Y 1  2  Y ³ † ´ns (k)
|{ns (k)}i =   as (k) |0i (85)
k,s ns (k)! k,s

Finally the lattice point displacement


v
u
1 Xu h̄ ³ ´
sn,α,i =√ t as (q) + a†s (−q) ²sα,i (q)eiq·rn (86)
Mα N q,s 2ωs (q)

will be important in the next section, especially with respect to zero-


D E
point motion (i.e. s2 T =0
6= 0).

4 Theory of Neutron Scattering

To “see” the lattice with neutrons, we want their De Broglie wavelength


λ = h/p
0.29A
λneutron = √ E measured in eV (87)
E
to be of the same length as the intersite distance on the lattice. This
means that their kinetic energy E ≈ 12 M v 2 ≈ 0.1eV, or E/kb ≈ 1000K;
i.e. thermal neutrons.

27
Source of thermal
n λ ≈ | a 1 | or |a 2 |
neutrons a1
a2

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

λ ∼ A À 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
X
V (r) = Vn δ(r − rn ) (89)
rn

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 diffraction. 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 ∝ |ρ(K, Ω)|2 (90)

where K = k0 − k and Ω = ω0 − ω. Furthermore, we take


X
ρ(r(t)) ∝ δ(r − rn (t)) (91)
n

where
1
rn (t) = rn + sn (t) and sn (t) = √ u(q)ei(q·rn −ω(q)t) (92)
M
describes the harmonic motion of the s-mode with wave-vector q.
XZ
ρ(K, Ω) ∝ dtei[K·(rn +sn (t))−Ωt] . (93)
n

For |K| ∼ 2π/A and sn (t) ¿ A we may expand


XZ
ρ(K, Ω) ∝ dtei[K·(rn )−Ωt] (1 + iK · sn (t) + · · ·) (94)
n

The first term yields a finite contribution only when

K = k0 − k = G and Ω = ω0 − ω = 0 (95)

which are the familiar Bragg conditions for elastic scattering.


The second term, however, yields something new. It only yields a
finite result when

K ± q = k0 − k ± q = G and Ω ± ωs (q) = ω0 − ω ± ωs (q) = 0 (96)

29
When multiplied by h̄ , 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 = k 0- q , ω = ω 0 - ω q k = k 0+ q , ω = ω 0 + ω q

q ,ω q q ,ω q
n n
k 0, ω0 k 0, ω0
Stokes Process Anti-Stokes Process
(phonon creation) (phonon absorbtion)

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 finite 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 effects 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

Initial Final
ψ 0 = 1− e i( k 0•⋅ r - ω 0 t ) ψ f = 1− i( k f •⋅ r - ω f t )
V
√ V e

n
k 0 ⋅ω 0
n k ⋅ω

φ0 E0 φf Ef

Figure 12: The initial (left) and final (right) states of the neutron and lattice during
a scattering event. The initial system state is given by Ψ0 = φ0 ψ0 , with energy
²0 = E0 + h̄ω0 where ω0 = k02 /2M . The final system state is given by Ψf = φf ψf ,
with energy ²f = Ef + h̄ωf where ωf = kf2 /2M .

is K = k0 − kf and the energy transfer which is finite for inelastic


scattering is h̄Ω = h̄ (ω0 − ωf ). Again we will take the neutron-lattice

31
interaction to be local
X 1 X Z d3 q X iq·(r−rn )
V (r) = V (r − rn ) = V (q)eiq·(r−rn ) = V0 e (97)
rn N q,n V n

where the locality of the interaction (V (r − rn ) ∝ δ(r − rn )) indicates


that V (q) = V (0) = V0 . Consistent with Aschcroft and Mermin, we
will take
2πh̄2 a 1 XZ
V (r) = d3 qeiq·(r−rn ) (98)
M V rn
2πh̄2 a
where a is the scattering length, and V0 = M is chosen such that the
total cross section σ = 4πa2 .
To formulate our quantum theory, we will use Fermi’s 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 k0 to kf is given by
2π X
P = δ(²0 − ²f ) |hΨ0 |V | Ψf i|2 (99)
h̄ f
2π X
= δ(E0 + h̄ω0 − Ef − h̄ωf )
h̄ f
¯ ¯2
¯1 Z ¯
¯
¯
¯V
3
d re i(k−k0 )·r
hφ0 |V (r)| φf i¯¯¯ (100)

If we now substitute in the ion-neutron potential Eq. 98, then the inte-
gral over r will yield a delta function V δ(q + k − k0 ) which allows the
q integral to be evaluated
¯ ¯
(2πh̄)3 X
2
¯X D
¯
¯ ¯
¯ −iK·rn ¯
E ¯2
¯
P =a 2
δ(E0 − Ef + h̄Ω) ¯¯ φ 0 ¯e ¯ φf ¯
¯
(101)
(M V ) f rn

32
Now, before proceeding to a calculation of the differential cross sec-

tion dΩdE we must be able to convert this probability (rate) for eigen-
states into a flux of neutrons of energy E and momentum p. A differ-
ential volume element of momentum space d3 p contains V d3 p/(2πh̄)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 = L3 with periodic boundary conditions so that for
any state Ψ in V ,

Ψ(x + L, y, z) = Ψ(x, y, z) (102)

1 P iq·r
If we write Ψ(r) = N qe Ψ(q), then it must be that

qx L = 2πm where m is an integer (103)

with similar relations for the y and z components. So for each volume
³ ´
2π 3
element of q-space L there is one such state. In terms of states p =
h̄q, the volume of a state is (2πh̄/L)3 . Thus d3 p contains V d3 p/(2πh̄)3
states.
The incident neutron flux of states (velocity times density) is
¯ ¯
h̄k0 2 h̄k0 ¯¯ 1 ik0 ·r ¯¯2 h̄k0
j= |Ψ0 | = ¯√ e ¯ = (104)
M M ¯ V ¯ MV
Then since the number of neutrons is conserved
dσ h̄k0 dσ d3 p p2 dpdΩ
j dEdΩ = dEdΩ = P V = PV (105)
dEdΩ M V dEdΩ (2πh̄)3 (2πh̄)3

33
And for thermal (non-relativistic) neutrons E = p2 /2M , so dE =
pdp/M , and
h̄k0 dσ h̄kM dEdΩ
dEdΩ = P V (106)
M V dEdΩ (2πh̄)3
or
dσ k (M V )2
=P . (107)
dEdΩ k0 (2πh̄)3
Substituting in the previous result for P
¯ ¯
dσ k (M V )2 2 (2πh̄)3 X ¯X D
¯
¯ ¯
¯ −iK·rn ¯
E ¯2
¯
= 3
a 2
δ(E0 − Ef + h̄Ω) ¯¯ φ 0 ¯e ¯ φf ¯
¯
dEdΩ k0 (2πh̄) (M V ) f rn
(108)
or
dσ k N a2
= S(K, Ω) (109)
dEdΩ k0 h̄
where
¯ ¯
1 X ¯X D ¯ ¯ E ¯2
¯ ¯ −iK·rn ¯ ¯
S(K, Ω) = δ(E0 − Ef + h̄Ω) ¯¯ φ 0 ¯e ¯ φf ¯
¯
. (110)
N f rn

We may deal with the Dirac delta function by substituting


Z ∞ dt iΩt
δ(Ω) = e . (111)
−∞ 2π

so that
1 XZ ∞ dt i((E0 −Ef )/h̄+Ω)t
S(K, Ω) = e
N f
−∞ 2π
X ¯¯D ¯ ¯
¯ iK·rn ¯
E¯ ¯D
¯¯
¯ ¯
¯ −iK·rm ¯

¯
¯ φ 0 ¯e ¯ φ f ¯ ¯ φ f ¯e ¯ φ0 ¯ . (112)
rn ,rm

then as
e−iHt/h̄ |φl i = e−iEl t/h̄ |φl i (113)

34
where H is the lattice Hamiltonian, we can write this as
1 XZ ∞ dt iΩt X D ¯¯ iHt/h̄ iK·rn −iHt/h̄ ¯¯ E
S(K, Ω) = e φ 0 ¯e e e ¯ φf
N f
−∞ 2π rn ,rm
D ¯ ¯ E
φf ¯¯e−iK·rm ¯¯ φ0 , (114)

and the argument in the first expectation value is the time-dependent


operator eiK·rn in the Heisenberg representation

eiK·rn (t) = eiHt/h̄ eiK·rn e−iHt/h̄ . (115)

Thus,
1 XZ ∞ dt iΩt X D ¯¯ iK·rn (t) ¯¯ E D ¯¯ −iK·rm ¯¯ E
S(K, Ω) = e φ 0 ¯e ¯ φf φ f ¯e ¯ φ0
N f
−∞ 2π rn ,rm
1 Z ∞ dt X D ¯ ¯ E
= eiΩt φ0 ¯¯eiK·rn (t) e−iK·rm ¯¯ φ0 . (116)
N −∞ 2π rn ,rm

Now since rn (t) = rn + sn (t) (with rn time independent),


1 XZ ∞ dt i(K·(rn −rm )+Ωt) D ¯¯ iK·sn (t) −iK·sm ¯¯ E
S(K, Ω) = e φ 0 ¯e e ¯ φ0 . (117)
N n,m −∞ 2π

This formula is correct at zero temperature. In order to describe


finite T effects (ie., anti-stokes processes involving phonon absorption)
we must introduce a thermal average over all states
X −βE X −βE
hφ0 |A| φ0 i → hAi = l
e hφl |A| φl i / e l
. (118)
l l

With this substitution,


1 XZ ∞ dt i(K·(rn −rm )+Ωt) D iK·sn (t) −iK·sm E
S(K, Ω) = e e e . (119)
N n,m −∞ 2π

and S(K, Ω) 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.
v
u
1 Xu h̄ ³ ´
sn,α (t) = √ t s
²α (q) as (q)(t) + a†s (−q)(t) eiq·rn (120)
Mα N q,s 2ωs (q)

So that, in particular hsn,α,i (t)i = hsn,α,i (0)i = 0. Then let A = iK ·


sn,α,i (t) and B = iK · sm,α,i (0) and suppose that the expectation values
of A and B are small. Then
* +
D
A B
E 1 1
e e = (1 + A + A2 + · · ·)(1 + B + B 2 + · · ·)
*
2 2 +
1 2 1 2
≈ 1 + A + B + AB + A + B + · · ·
2 2
1 D E
≈ 1 + 2AB + A2 + B 2 + · · ·
2

1
e2 h 2AB+A 2
+B 2 i
(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
D E 2 2
eiK·sn (t) e−iK·sm = e− 2 h(K·sn (t)) i e− 2 h(K·sm ) i ehK·sn (t)K·sm i .
1 1
(122)

Since the Hamiltonian has no time dependence, and the lattice is in-
variant under translations rn
D E 2
eiK·sn (t) e−iK·sm = e−h(K·sn ) i ehK·sn−m (t)K·s0 i , (123)

36
where the first term is called the Debye-Waller factor e−2W .
2
e−2W = e−h(K·sn ) i . (124)

Thus letting l = n − m
Z dt i(K·rl +Ωt) hK·sl (t)K·s0 i
−2W X ∞
S(K, Ω) = e e e . (125)
l
−∞ 2π
Here the Debye-Waller factor contains much of the crucial quantum
physics. It is finite, even at T = 0 due to zero-point fluctuations, and
since hK · sn i2 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 finite.

4.2.2 Zero-phonon Elastic Scattering

One may disentangle the elastic and inelastic processes by expanding


the exponential in the equation above.
X 1
ehK·sl (t)K·s0 i = (hK · sl (t)K · s0 i)m (126)
m m!
If we approximate the exponential by 1, ie. take only the first, m = 0
term, then
XZ ∞
dt i(K·rl −Ωt)
S0 (K, Ω) = e−2W e . (127)
l −∞ 2π
And we recover the lowest order classical result (modified by the Debye-
Waller factor) which gives us the Bragg conditions that S0 (K, Ω) is only
finite when K = G and Ω = ω0 − ωf = 0.
X
S0 (K, Ω) = e−2W δ(Ω)N δK,G , (128)
G

37
dσ0 k N a2 −2W X
= e δ(Ω)N δK,G (129)
dEdΩ k0 h̄ G
However, now the scattering intensity is reduced by the Debye-
Waller factor e−2W , which accounts for zero-point motion and thermal
fluctuations.

4.2.3 One-Phonon Inelastic Scattering

When m = 1, then the scattering involves either the absorption or


creation of a phonon. To evaluate
X Z ∞ dt i(K·r +Ωt)
S1 (K, Ω) = e−2W e l
hK · sl (t)K · s0 (0)i . (130)
l −∞ 2π
we need
v
u
1 Xu h̄ ³ ´
sn,α (t) = √ t ²sα (q) as (q, t) + a†s (−q, t) eiq·rn (131)
Mα N q,s 2ωs (q)
in the Heisenberg representation, and therefore we need,

a(q, t) = eiHt/h̄ a(q)e−iHt/h̄


† †
= ei(ω(q)ta (q)a(q)) a(q)e−i(ω(q)ta (q)a(q)
= a(q)ei(ω(q)t(a (q)a(q)−1)) e−i(ω(q)ta (q)a(q)
† †

= a(q)e−iω(q)t (132)

where we have used the fact that (a† a)n a = (a† a)n−1 a† aa = (a† a)n−1 (aa† −
1)a = (a† a)n−1 a(a† a−1) = a(a† a−1)n . Similarly a† (q, t) = a† (q)eiω(q)t .
Thus,
√ iq·r
1 X h̄e n s ³ ´
sn,α (t) = √ q ²α (q) as (q)e−iωs (q)t + a†s (−q)eiωs (q)t
Mα N q,s 2ωs (q)
(133)

38
and
v
u
1 Xu h̄ ³ ´
s0,α (0) = √ t ²rα (p) ar (p) + a†r (−p) (134)
Mα N p,r 2ωr (p)

Recall, we want to evaluate


Z dt i(K·rl +Ωt)
−2W X ∞
S1 (K, Ω) = e e hK · sl (t)K · s0 (0)i . (135)
l
−∞ 2π
Clearly, the only terms which survive in hK · sl (t)K · s0 (0)i are those
with r = s and p = −q. Furthermore, the sum over l yields a delta
P
function N G δK+q,G . Then as ²(G − k) = ²(−k) = ²∗ (k), and ω(G −
k) = ω(k),

−2W
Z ∞dt iΩt X h̄ |K · ²(K)|2
S1 (K, Ω) = e e δK+q,G (136)
−∞ 2π q,G,s 2ωs (q)M
h D E D Ei
e−iωs (q)t as (−K)a†s (−K) + eiωs (q)t a†s (−K)as (−K)

The occupancy of each mode n(q) is given by the Bose factor


1
hn(q)i = (137)
eβω(q) − 1
So, finally
X h̄
S1 (K, Ω) = e−2W |K · ²s (K)|2 (138)
s 2M ωs (K)
[(1 + ns (K))δ(−Ω + ωs (K)) + ns (K)δ(Ω + ωs (K))] .

For the first term, we get a contribution only when Ω − ωs (K) = ω0 −


ωf − ωs (K) = 0; ie., the final 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 + ns (K)) 6= 0 at any T .
The second term is only finite when Ω + ωs (K) = ω0 − ωf + ωs (K) = 0;
ie., the final 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 finite temperatures, and in fact, the factor
ns (K) = 0 at zero temperature. These terms correspond to the Stokes
and anti-Stokes processes, respectively, illustrated in Fig. 13.
k = k 0- q , ω = ω 0 - ω q k = k 0+ q , ω = ω 0 + ω q

1 + n (K) n (K)
s s
q ,ω q q ,ω q
n n
k 0, ω0 k 0, ω0
Stokes Process Anti-Stokes Process
(phonon creation) (phonon absorbtion)

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 finite-T, when ns (K) 6= 0.

If we were to continue our expansion of the exponential to larger


values of m, we would find 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.
1X
H= |Ps(k)|2 + ωs2(k) |Qs(k)|2 (1)
2 k,s
At a given temperature T, the occupancy of a given mode is
1
hns(k)i = (2)
eβωs(k) − 1
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 different formulas for these quantities,
we will first 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 system H(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 i i
hx · pi = h[H, x · p]i = hHx · p − x · pHi
dt h̄ h̄
iE
= hx · p − x · pi = 0 (3)

where E is the eigenenergy of the stationary state. Let
p2
H= + V (x) , (4)
2m
then
 
* 2 +
 p 
0 =  + V (x), x · p
2m
 
* 2 +
 p 
=  , x · p + [V (x), x · p]
2m
* +
1 · 2 ¸
= p , x · p + x · [V (x), p] (5)
2m
h i
2
Then as p , x = p [p, x] + [p, x] p = −2ih̄p and p = −ih̄∇x,
* +
−ih̄ 2
0= p + ih̄x · ∇xV (x) (6)
m
3
or, the Virial theorem:

2 hT i = hx · ∇xV (x)i (7)

Now, lets apply this to a harmonic oscillator where V = 21 mω 2x2


and T = p2/2m, < H >=< T > + < V >= h̄ω(n + 12 ) and .
We get
1 1
hT i = hV (x)i = h̄ω(n + ) (8)
2 2
Lets now apply this to find the RMS excursion of a lattice
site in an elemental lattice r = 1
¿
2
À 1 X¿ 2 À
s = s
N n,i n,i
1 *X 1 X s iq·rn r ik·rn
+
= Qs(q)²i (q)e Qs(k)²i (k)e
N n,i N M q,s,k,r
1 *X 1 X s iq·rn r −ik·rn
+
= Qs(q)²i (q)e Qs(−k)²i (−k)e
N n,i N M q,s,k,r
1 *X 1 X s iq·rn ∗ ∗r −ik·rn
+
= Qs(q)²i (q)e Qs (k)²i (k)e
N n,i N M q,s,k,r
1 X¿ 2
À
= |Qs(q)|
N M q,s
1 X 2 *1 2 2
+
= ω (q) |Qs(q)|
N M q,s ωs2(q) 2 s
 
1 X 2 1 1
= 2
h̄ωs(q) ns(q) + 
N M q,s ωs (q) 2 2

4
 
¿ À 1 X h̄  1
s2 = ns(q) +  (9)
NM q,s ωs (q) 2
This integral, which must be finite in order for the system to
have long-range order, is still difficult to perform. However, the
integral may be written as a function of ωs(q) only.
 
¿
2
À 1 X h̄  1 1
s = + (10)
NM q,s ωs (q) eβωs (q) − 1 2
It would be convenient therefore, to introduce a density of
phonon states
1 X
Z(ω) = δ (ω − ωs(q)) (11)
N q,s

so that
 
¿ À h̄ Z 1 1
s2 = dωZ(ω) n(ω) +  (12)
M ω 2

1.2 The Phonon Density of States

D E
In addition to the calculation of s2 , the density of states Z(ω)
is also useful in the calculation of E =< H >, the partition
function and the related thermodynamic properties
In order to calculate
1 X
Z(ω) = δ (ω − ωs(q)) (13)
N q,s

5
we must first better define the sum over q. As we discussed last
chapter for a 3-d cubic system, we will assume that we have a
periodic finite lattice of N basis points, or N 1/3 in each of the
principle lattice directions a1, a2, and a3. Then, we want the
Fourier representation to respect the periodic boundary condi-
tions (pbc), so
1/3 (a +a +a )
eiq·(r+N 1 2 3 ) = eiq·r (14)

This means that qi = 2πm/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 qi, so we will choose those within the first
6
Brillouin zone, so that
1
G · q ≤ G2 (15)
2
The size of this region is the same as that of a unit cell of the
reciprocal lattice g1 · (g2 × g3). Since there are N states in this
region, the density of q states is
N N Vc V
= = (16)
g1 · (g2 × g3) (2π)3 (2π)3
where Vc is the volume of a Bravais lattice cell (a1 · a2 × a3),
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 dω is then given by the volume of q-space
between the surfaces defined by ω = ωs(q) and ω = ωs(q)+dω
multiplied by V /(2π)3
V Z ω+dω 3
Z(ω)dω = dq
(2π)3 ω
V XZ 3
= dω d qδ (ω − ωs(q)) (17)
(2π)3 s

7
qy

dq ⊥ Surface
dS ω ω = ωs (q)

qx

Figure 2: States in q-space. Sω is the surface of constant ω = ωs (q), so that d3 q =


dSω dω
dSw dq⊥ = ∇q ωs (q)
.

As shown in Fig. 2, dω = ∇q ωs(q)dq⊥, and


dSω dω
d3q = dSw dq⊥ = , (18)
∇q ωs(q)
where Sω is the surface in q-space of constant ω = ωs(q). Then
V XZ dSω
Z(ω)dω = dω (19)
(2π)3 s ω=ωs (q) ∇q ωs (q)

Thus the density of states is high in regions where the dispersion


is flat so that ∇q ωs(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
M

Dispersion Z(ω) DOS

ω(q) flat
ω0

ω
-π/a π/a q ω0

linear

Figure 3: Linear harmonic chain. The phonon dispersion of this chain must in-
clude an acoustic mode, so ω(q) will be linear near q = 0, and it must be symmetric
about q = 0 and q = π/q due to the point-group symmetry. Thus, the density of
states (DOS) will be flat near ω = 0 corresponding to the acoustic mode (for which
∇q ω(q) =constant), and will be divergent near ω = ω0 corresponding to the peak of
the dispersion (where ∇q ω(q) = 0).

at a zone boundary, with corresponding peaks in the phonon


DOS. However, any point within the Brillouin zone for which
∇q ω(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 h̄ω(q) ∼ kB T which are low frequency modes in gen-
eral. From a very general set of (symmetry) constraints we have
2.0
ω+
ω-
1.5
Debye model
1.0
ω

0.5

0.0
-4 -2 0 2 4
q

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
ωs(q) = cs|q|. Thus, for the thermodynamic properties of the
lattice, we care predominantly about the limit ω(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 ωs(q) = cs|q| for all s and q, then ∇q ωs(q) =
cs for all s and q, and
V X Z dSω
Z(ω) =
(2π)3 s ∇q ωs(q)
V X Z dSω
= (20)
(2π)3 s cs
The surface integral may be evaluated, and yields a constant
R
dSω = Ss for each branch. Typically cs is different for different
modes. However, we will assume that the system is isotropic, so
cs = c. If the dispersion is isotopic, then the surface of constant
ωs(q) is just a sphere, so the surface integral is trivial








2 for d = 1



Ss(ω = ω(q)) = 

2πq = 2πω/c for d = 2 (21)







 4πq 2 = 4πω 2/c2 for d = 3

11
then since the number of modes = d
 

 

 




2/c for d = 1 



 
V 
 

Z(ω) = 3  2πq = 4πω/c2 for d = 2  0 < ω < ωD
(2π) 









 2 2 3 
 4πq = 12πω /c
 for d = 3 

(22)
Note that since the total number of states is finite, we have
introduced a cutoff ωD 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 ω0 and does not interact with its neighbors.
The model is dispersionless ω(q) = ω0, and the DOS for this
system is a delta function Z(ω) = cδ(ω −ω0). 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 specific
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 define long-ranged order (LRO) as a finite value of
 
h̄ X 1  1 1
hs2i = +
M N q,s ωs(q) eβωs(q) − 1 2
h̄ X sinh (βωs(q)/2)
= (23)
2M N q,s ωs(q)cosh (βωs(q)/2)
Since we expect all lattices to melt for some high temperature,
we are interested only in the T → 0 limit. Clearly also given
1
the factor of ωs (q) in the summand, we are most interested in
acoustic modes since they are the ones which will cause a di-
vergence.
 
h̄ X 1  1 1
lim hs2i = lim  +  (24)
β→∞ 2M N β→∞ q,s ωs(q) βωs(q) 2
Clearly the low frequency modes are most important so a Debye
model may be used
 
2 h̄ Z ω
D 1 1 1
lim hs i ≈ 0
dωZ(ω)  + , (25)
β→∞ 2M N ω βω 2

14
where Z(ω) is the same as was defined above in Eq. 22.
 

 

 




2/c for d = 1 



 
V 
 

Z(ω) = 3  2πq = 4πω/c2 for d = 2  0 < ω < ωD
(2π) 









 2 2 3 
 4πq = 12πω /c
 for d = 3 

(26)
Thus in 3-d the DOS always cancels the 1/ω singularity but
in two dimensions the singularity is only cancelled when T = 0
(β = ∞), and in one dimension hs2i = ∞ for all T . This

hs2i d=1 d=2 d=3


T =0 ∞ finite finite
T 6= 0 ∞ ∞ finite

Table 1: hs2 i for lattices of different dimension, assuming the presence of an acoustic
mode.

is a specific case of the Mermin-Wagner Theorem. We should


emphasize that the result hs2i = ∞ 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 fluctuaions 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 fintie 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 ω0 and does not interact with
its neighbors. The DOS for this system is a delta function as
described above. For such a DOS, hs2i is always finite. 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 justified by dividing an infinite system
into a finite 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 ({ns(k)}) ∝ e−βE({ns(k)}) (27)

Thus the partition function is given by


X
Z = e−βE({ns(k)})
{ns (k)}
P 1
=
X
e −β (
k,s h̄ωs (k) ns (k)+ 2 )
{ns (k)}
Y
= Zs(k) (28)
s,k

17
where Zs(k) is the partition function for the mode s, k; i.e. the
modes are independent and decouple.
1
e−βh̄ωs(k)(ns(k)+ 2 )
X
Zs(k) =
n
X
= e−βh̄ωs(k)/2 e−βh̄ωs(k)(ns(k))
n
−βh̄ωs (k)/2
e
=
1 − e−βh̄ωs(k)
1
= (29)
2 sinh (βωs(k)/2)
The free energy is given by
X
F = −kB T ln (Z) = kB T ln (2 sinh (βh̄ωs(k)/2)) (30)
k,s

Since dE = T dS −P dV and dF = T dS −P dV −T dS −SdT ,


the entropy is
 
∂F 
S = − , (31)
∂T V
and system energy is then given by
 
∂F 
E = F + TS = F − T  (32)
∂T V
where constant volume V is guaranteed by the harmonic ap-
proximation (since < s >= 0).
X 1
E= h̄ωs(k)coth (βωs(k)/2) (33)
k,s 2

18
The specific heat is then given by
 
dE X
C =   = kB (βh̄ωs(k))2 csch2 (βωs(k)/2) (34)
dT V k,s

where csch (x) = 1/sinh (x)


Consider the specific heat of our 3-dimensional Debye model.
Z ω
D 2 2
C = kB 0
dωZ(ω) (βω/2) csch (βω/2)
 
2
 12V πω 
Z ω
D 2 2
= kB 0 dω   (βω/2) csch (βω/2) (35)
(2πc)3
Where the Debye frequency ωD is determined by the require-
ment that
 
2
 12V πω 
Z ω Z ω
D D
3rN = dωZ(ω) = dω   , (36)
0 0 (2πc)3
or V /(2πc)3 = 3rN/(4ωD
3
).
Clearly the integral for C is a mess, except in the high and
low T limits. At high temperatures βh̄ωD /2 ¿ 1,
Z ω
D
C ≈ kB 0
dωZ(ω) = 3N rkB (37)

This is the well known classical result (equipartition theorem)


which attributes (1/2)kB of the specific 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, βh̄ω/2 À 1,

csch2 (βω/2) ≈ 2e−βω/2 (38)

Thus, at low T , only the low frequency modes contribute, so


the upper bound of integration may be extended to ∞
 2
3rN 2  βh̄ω 
Z ∞
C ≈ 12πkB 2 0
dωω 2e−βωs(k)/2 . (39)
4ωD 2
If we make the change of variables x = βh̄ω/2, we get
 
9kB rN π  1 3 Z ∞
C ≈ 0
dxx4 e−x (40)
2 ωD βh̄
 
9kB rN π  1 3
≈ 24 (41)
2 ωD βh̄
Then, if we identify the Debye temperature θD = h̄ωD /kB , we
get
3  
T
C ≈ 96πrN kB   (42)
θD
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 define the coefficient of free expansion (P = 0)
as
1 dL 1 dV
αL = or αV = 3αL = . (43)
L dT V dT
Of course, this measurement only makes sense in equilibrium.
 
dF
P = −  = 0 (44)
dV T
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
1
V (x) = bx + cx3 + mω 2x2 (45)
2

21
and let’s 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 first term does not have the desired
effect! It does correspond to a temperature-independent shift
in the oscillator, as can be seen by completing the square
 
1 2 2 1 b 2 b2
mω x + bx = x+ − (46)
2 2 mω 2 2mω 2
Clearly < x >= − mωb 2 , independent of the temperature; that
is, assuming that b is temperature-independent. What we need
is a temperature dependent coefficient b!
The cubic term has the desired effect. 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-field
decomposition.
¿ À
3 2
cx ≈ cηx x + c(1 − η)x2 hxi (47)

and treat these two terms separately (the new parameter η is


22
3
c=0.1
2
V(x)

c=0.0
1
0
-2 -1 0 1 2
x

Figure 7: Plot of the potential V (x) = 12 mx2 + cx3 when m = ω = 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 fixed < x >= 0.

to be determined self-consistently, usually by minimizing the


free energy with respect to η). The first term yields the needed
temperature dependent shift of < x >
 2
1 2 2
¿
2
À 1 2 cη < x2 >  c2 < x 2 > 2
mω x + cηx x = mω x +  −
2 2 mω 2 2mω 2
(48)
2
so that < x >= − cη<x
mω 2
>
. Clearly the renormalization of the
equilibrium position of the harmonic oscillator will be temper-
ature dependent. While the second term, (1 − η)x2 < x >,

23
yields a shift in the frequency ω → ω 0
 1
0  2(1 − η) < x >  2
ω = ω 1 +  (49)
mω 2
which is a function of the equilibrium position. Thus a mean-
field description of the cubic term is consistent with the observed
physics.
In what follows, we will approximate the effect 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 ω to
ω 0; however, we imagine that the energy levels remain of the
form
1
En = h̄ω 0(< x >)(n + ), (50)
2
and that < x > varies with temperature, consistent with the
mean-field approximation just described.
To proceed, imagine the cube of cubic system to be made up
of oscillators which are independent. Since the final result can
be formulated as a sum over these independent modes, consider
µ ¶
dF
only one. In equilibrium, where P = − dV T = 0, the free

24
energy of one of the modes is
1 µ
−βh̄ω

F = Φ + h̄ω + kB T ln 1 − e (51)
2
and (following the notation of Ibach and Lüth), let the lattice
potential
1
Φ = Φ0 + f (a − a0)2 + · · · (52)
2
where f is the spring constant. Then
   
dF 1 ∂ω  1 h̄ω
0 = P =   = f (a − a0) + h̄ω − −βh̄ω
,
da T ω ∂a 2 1−e
(53)
If we identify the last term in parenthesis as ²(ω, T ), and solve
for a, then
1 ∂ω
a = a0 −
²(ω, T ) (54)
ωf ∂a
Since we now know a(T ) for a single mode, we may calculate
the linear expansion coefficient for this mode
1 da 1 ∂ ln w ∂²(ω, T )
αL = =− 2 (55)
a0 dT a0f ∂ ln a ∂T
To generalize this to a solid let αL → αV (as discussed above)
dP
and a20f → V 2 dV = V κ (κ is the bulk modulus) and sum over

25
all modes the modes k, s
1 dV 1 X ∂ ln ωs(k) ∂² (ωs(k), T )
αV = = − . (56)
V dT κV k,s ∂ ln V ∂T
∂²
Clearly (due to the factor of ∂T ), αV will have a behavior sim-
ilar to that of the specific heat (αV ∼ T 3 for low T , and
αV =constant for high T ). In addition, for many lattices, the
Gruneisen number
∂ ln ωs(k)
γ= (57)
∂ ln V
shows a weak dependence upon s, k, and may be replaced by
its average, called the Gruneisen parameter
*
∂ ln ωs(k) +
hγi = , (58)
∂ ln V
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
X 1 X h̄3/2
s3l = r e i(p+q+k−G)·rl
(59)
l (2M N )3/2 l,k,q,p ω(q)ω(k)ω(p)
µ ¶µ ¶µ ¶
† † †
a(k) + a (−k) a(p) + a (−p) a(q) + a (−q) .

The sum over l yields a delta function δp+q+k,G (ie., crystal mo-
mentum conservation). Physically, these processes correspond
26
k (+ G) k (+ G)
q-k q-k

q -q

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, κ, 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 µOhm-cm).

not good thermal conductors. This subsection will be devoted


to understanding what makes stiff crystals like diamond such
good conductors of heat.
The thermal conductivity κ is measured by setting up a small
steady thermal gradient across the material, then

Q = −κ∇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
1 X
Qx = h̄ωs(q)hns(q)ivsx(q) (61)
V q,s
∂ωs (q)
where the group velocity is given by vsx(q) = ∂qx Since we
assume ∇T is small, we will only look at the linear response
of the system where hns(q)i deviates little from its equilibrium
28
value hns(q)i0 . Furthermore since ωs(q) = ωs(−q),
∂ωs(−q)
vsx(−q) = = −vsx(q) (62)
∂ − qx
Thus as hns(−q)i0 = hns(q)i0
1 X
Q0x = h̄ωs(q)hns(q)i0vsx(q) = 0 (63)
V q,s

since the sum is over all q in the B.Z. Thus if we expand

hns(q)i = hns(q)i0 + hns(q)i1 + · · · (64)

we get
1 X
Qx ≈ h̄ωs(q)hns(q)i1vsx(q) (65)
V q,s
since we presumably already know ωs(k), the calculation of Q
and hence κ reduces to the evaluation of the linear change in
< n >.
Within a region, < n > can change in two ways. Either
phonons can diffuse into the region, or they can decay through
an anharmonic (cubic) term into other modes. so
¯ ¯
¯ ¯
dhni ∂hni ¯
¯ ∂hni ¯
¯
= +
¯
¯
¯
¯ (66)
dt ∂t diffusion ¯ ∂t decay ¯

29
decay diffusion
diffusion

Figure 9: Change of phonon density within a trapazoidal region. hns (q)i can change
either by phonon decay or by phonon diffusion into and out of the region.

dhni
However dt = 0 since we are in a steady state. The decay
process is usually described by a relaxation time τ (or a mean-
free path l = vτ )
¯
∂hni ¯¯¯ < n > − < n >0 < n >1
¯ =− =≈ − (67)
∂t ¯¯decay τ τ
dhni
The diffusion part of dt is addressed pictorially in Fig. 10.
Formally,
¯
∂hni ¯¯¯ hn(x − vx∆t)i − hn(x)i ∂hn(x)i
¯ ≈ ≈ − vx (68)
∂t ¯¯diffusion ∆t ∂x
∂hn(x)i ∂T ∂hn(x)i0 + hn(x)i1 ∂T
≈ −vx ≈ −vx
∂T ∂x ∂T ∂x
Keeping only the lowest order term,
¯
∂hni ¯¯¯ ∂hn(x)i0 ∂T
¯ ≈ −vx (69)
∂t ¯¯diffusion ∂T ∂x
30
x
v region of
source region interest

v ∆t
Figure 10: Phonon diffusion. 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 = (nlef t − nright )/∆t.

dhni
Then as dt = 0,
¯ ¯
¯ ¯
∂hni ¯
¯ ∂hni ¯
¯
¯
¯ =− ¯
¯ (70)
∂t diffusion
¯ ∂t decay ¯

or
1 ∂hn(x)i0 ∂T
hni = −vxτs(q) . (71)
∂T ∂x
Thus
1 X 2 ∂hn(x)i0 ∂T
Qx ≈ − h̄ωs(q)vsx(q)τs(q) , (72)
V q,s ∂T ∂x
and since Q = −κ∇T
1 X 2 ∂hn(x)i0
κ≈ h̄ωs(q)vsx(q)τ . (73)
V q,s ∂T
From this relationship we can learn several things. First since
2
κ ∼ vsx (q), phonons near the zone boundary or optical modes
31
with small vs(q) = ∇q ωs(q) contribute little to the thermal
conductivity. Also, stiff materials, with very fast speed of the
acoustic modes vsx(q) ≈ c will have a large κ. Second, since
κ ∼ τs(q), and ls(q) = vsx(q)τs(q), κ will be small for materi-
als with short mean-free paths. The mean-free path is effected
by defects, anharmonic Umklapp processes, etc. We will explore
this effect, especially its temperature dependence, in more de-
tail.
At low T , only low-energy, acoustic, modes can be excited
(those with h̄ωs(q) ∼ kB T ). These modes have

vs(q) = cs (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 ∼ h̄ωc so the thermal current is not disturbed by
anharmonic processes. Thus the anharmonic terms at low T

32
do not affect 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 q1 of an excited
mode is large enough and there exists a sufficiently small G so
that q1 > G/2 (c.f. Fig. 11). This is called an Umklapp process,

q1 = q2 + q3+ G

q3 G q3
Qout
q
1
q2 q2

q1 Q in
first Brillouin zone
Figure 11: Umklapp processes involve a reciprocal lattice vector G in lattice momen-
tum conservation. They are possible whenever q1 > 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 κ are very much

33
smaller for high temperatures where q1 can be larger than half
the smallest G.
So what about diamond? It is very hard and very stiff, so the
sound velocities cs are large, and so thermally excited modes for
which kB T ∼ h̄ω ∼ h̄cq involve small q1 for which Umklapp
processes are irrelevant. Second κ ∼ c2 which is large. Thus κ
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 Infinite-Square-Well Potential . . . . . . . . . . . . . . . . . . . . . . 5
2.2 The Fermi Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 T = 0, The Pauli Principle . . . . . . . . . . . . . . . . . . . . 10
2.2.2 T 6= 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 Effective 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
different time scales involved in these systems.

τion ∼ 1/ωD À τelectron ∼ (1)
EF
where EF 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 ωD /EF as an expansion pa-
rameter); that is if we are capable of solving the problem of the
remaining purely electronic system.
At first 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 justified to ignore the
electron-electron interactions, or treat them as a renormaliza-
tion of the non-interacting problem (effective 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 specific heat.
Except in certain “heavy fermion” metals, the electronic contri-
bution to the specific 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 (Celectron ≈ Cphonon ≈ 3N rkB ).
This puzzle is resolved, at the simplest level: that of the non-
interacting Fermi gas.

2 The Non-Interacting Fermi Gas

2.1 Infinite-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, unaffected 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
P2
H= + V (r) . (2)
2m
5
As shown in Fig. 1, the core potential V (r), like the lattice, is
periodic
V(r)

a 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 ψ ∼ eik·r . Now consider the
core potential as a perturbation. The electrons will be strongly
effected by the periodicity of the potential when λ = 2π/k ∼ a
1
. However, when k is small so that λ À a (or when k is large,
so λ ¿ a) the structure of the potential may be neglected, or
we can assume V (r) = V0 anywhere within the material. The
1
Interestingly, when λ ∼ a, the Bragg condition 2d sin θ ≈ a ≈ λ may easily be satisfied, so the
electrons, which may be though of as DeBroglie waves, scatter off of the lattice. Consequently states
for which λ = 2π/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 confine the electrons (and so maintain
charge neutrality), so V (r) = ∞ anywhere outside the material.

Figure 2: Infinite square-well potential. V (r) = V0 within the well, and V (r) = ∞
outside to confine the electrons and maintain charge neutrality.

Thus we will approximate the potential of a cubic solid with


linear dimension L as an infinite square-well potential.





 V0 0 < r i < L
V (r) = 

(3)


 ∞ otherwise
The electronic wavefunctions in this potential satisfy
h̄2 2
− ∇ ψ(r) = (E 0 − V0) ψ(r) = Eψ(r) (4)
2m

7
The normalize plane wave solution to this model is

1/2 
23
Y
ψ(r) =   sin kixi where i = x, y, or z (5)
i=1 L

and kiL = niπ in order to satisfy the boundary condition that


ψ = 0 on the surface of the cube. Furthermore, solutions with
ni < 0 are not independent of solutions with ni > 0 and may
be excluded. Solutions with ni = 0 cannot be normalized and
are excluded (they correspond to no electron in the state). The
kz

ky

3
(π/L)
kx

π/L
Figure 3: Allowed k-states for an electron confined by a infinite-square potential. Each
state has a volume of (π/L)3 in k-space.

eigenenergies of the wavefunctions are


h̄2∇2 h̄ X 2 h̄2π 2 µ 2 2 2

− ψ= k = n + ny + nz (6)
2m 2m i i 2mL2 x

8
and as a result of these restrictions, states in k-space are con-
fined to the first quadrant (c.f. Fig. 3). Each state has a volume
(π/L)3 of k-space. Thus as L → ∞, the number of states with
energies E(k) < E < E(k) + dE is
(4πk 2 dk)/8
0
dZ = . (7)
(π/L)3
r
h̄2 k 2
Then, since E = 2m ,
2
so k dk = m
h̄2
2mE/h̄2dE
 
0 3 1  2m 3/2 1/2
dZ = dZ /L = 2 E dE . (8)
4π h̄2
or, the density of state per unit volume is
 
dZ 1  2m 3/2 1/2
D(E) = = E . (9)
dE 4π 2 h̄2
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 fill each orbital or k-
state, one of spin up the other spin down. If we account for this
spin degeneracy in D, then
 
1  2m 3/2 1/2
D(E) = 2 E . (10)
2π h̄2

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
Z ∞
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/kB 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 fill up all

10
k-states until we reach some Fermi radius kF , corresponding to
some Fermi Energy EF
kz

ky 2 2
kf h kf
= Ef
2m

occupied
states

kx
kf D(E)

Figure 4: Due to the Pauli principle, all k-states up to kF , and all states with energies
up to Ef are filled at zero temperature.

h̄2kF2
EF = , (13)
2m
thus,
f (E, T = 0) = θ(EF − E) (14)

and
Z ∞ Z E
F
n = 0
D(E)f (E, T )dE = 0
D(E)DE
 
2m 3/2 1 Z EF 1/2
=  E DE
h̄2 2π 2 0
 3/2
2m 1 2 2/3
=  2 E , (15)
h̄ 2π 2 3 F
11
or
h̄2 µ 2 ¶3/2
EF = 3π n = k B TF (16)
2m
which also defines the Fermi temperature TF . Thus for metals,
in which n ≈ 1023/cm3, EF ≈ 10−11 erg ≈ 10eV ≈ kB 105K.
Notice that due to the Pauli principle, the average energy of
the electrons will be finite, even at T = 0!
Z E
F 3
E= 0
D(E)EdE = nEF . (17)
5
However, it is the electrons near EF in energy which may be
excited and are therefore important. These have a DeBroglie
wavelength of roughly

12.3 A ◦
λe = 1/2
≈4A (18)
(E(eV))
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 6= 0, Fermi Statistics

At finite temperatures some of the states will be thermally ex-


cited. The energy available for these excitations is roughly
kB T , and the only possible excitations are from filled to un-
filled electronic states. Therefore, only the states within k B T
(EF − kB T < E < EF + kB T ) of the Fermi surface may be
excited. f (E, T ) must be modified accordingly.
What we need is then f (E, T ) at finite T which also satisfies
the Pauli principle. Lets return to our model of a periodic solid
which is constructed by bringing individual atoms together from
an infinite separation. First, just consider a solid constructed
from only two atoms, each with a single orbital (Fig. 5). For

1 2

δ n1= -1 δ n1= +1
Figure 5: Exchange of electrons in a solid composed of two orbitals.

13
this system, in equilibrium,
X ∂F
0 = δF = δni (19)
i ∂ni
P
electrons are conserved so i δni = 0. Thus, for our two orbital
system
∂F ∂F
δn1 + δn2 = 0 and δn1 + δn2 = 0 (20)
∂n1 ∂n2
or
∂F ∂F
= (21)
∂n1 ∂n2
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
=µ (22)
∂ni
for all i. µ 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 F = E − T S is the appropriate
potential. The system energy E and Entropy S may be written
14
as functions of the orbital energies Ei and occupancies ni and
the degeneracy gi of the state of energy Ei. For example,

E4 g=4 n=2
4 4
E3 g=4 n=4
3 3

E2 g=2 n=1
2 2
E1 g=2 n=2
1 1

Figure 6: states from an ensemble of orbitals.

X
E= ni Ei . (23)
i

The entropy S requires a bit more thought. If P is the number


of ways of distributing the electrons among the states, then

S = kB ln P . (24)

Consider a set of gi states with energy Ei. The number of


ways of distributing the first electron in these states is gi. For
a second electron we then have gi − 1 ways... etc. So for ni
electrons there are
gi !
(25)
ni!(gi − ni)!
15
possible ways of accommodating the ni (indistinguishable) elec-
trons in gi states.
The number of ways of making the whole system (ie, filling
energy levels with Ei 6= Ej ) is then
Y gi !
P = , (26)
i ni !(gi − ni )!
and so, the entropy
X
S = kB ln gi! − ln ni! − ln(gi − ni)! . (27)
i
For large n, ln n! ≈ n ln n − n, so
X
S = kB gi ln gi − ni ln ni − (gi − ni) ln(gi − ni) (28)
i
and
X X
F= ni Ei − k B T gi ln gi − ni ln ni − (gi − ni) ln(gi − ni)
i i
(29)
We will want to use the chemical potential µ in our thermo-
dynamic calculations
∂F
µ= = Ek + kB T (ln nk + 1 − ln(gk − nk ) − 1) , (30)
∂nk
where β = 1/kB T . Solving for nk
gk
nk = . (31)
1 + eβ(Ek −µ)
16
Thus the probability that a quantum state with energy E is
occupied, is (the Fermi function)
1
f (E, T ) = . (32)
1 + eβ(Ek −µ)
At T = 0, β = ∞, and f (E, 0) = θ(µ − E). Thus µ(T = 0) =
EF . However in general µ is temperature dependent, since it
must be adjusted to keep the particle number fixed. In addition,
1.5

1.0
+1)
β(ω−µ)

0.5
1/(e

0.0
0.0 0.5 1.0 1.5 2.0
ω
³ ´
Figure 7: Plot of the Fermi function 1/ e−β(ω−µ) + 1 when β = 1/kB T = 20 and
µ = 1. Not that at energies ω ≈ µ the Fermi function displays a smooth step of width
≈ kB T = 0.05. This allows thermal excitations of particles near the Fermi surface.

when T 6= 0, f becomes less sharp at energies E ≈ µ. This


reflects the fact that particles with energies E − µ ≈ kB T may
be excited to higher energy states.

17
Specific Heat The form of f (E, T ) also clarifies why the elec-
tronic specific heat of metals is so small compared to the clas-
sical result Cclassical = 23 nkB T . The reason is simple: only the
electrons with energies within about kB T of the Fermi surface
kB T
may be excited (about EF of the electron density) each with
excitation energy of about kB T . Therefore,
kB T T
Uexcitation ≈ kB T n = nkB T (33)
EF TF
so
T
C ≈ nkB (34)
TF
Then as T ¿ TF (TF is typically about 105K in most metals2 )
C ≈ nkB TT ¿ Cclassical ≈ nkB . Thus at temperatures where
F

the phonons contribute essentially a classical result to the spe-


cific heat, the electronic contribution is vanishingly small. In
general this holds except at very low T where the phonon con-
tribution Cphonon ∼ T 3 goes to zero faster than the electronic
contribution to the specific heat.
2
Heavy Fermion systems are the exception to this rule. There TF can be as small as a fraction
of a degree Kelvin. As a result, they may have very large electronic specific heats.

18
Specific Heat Calculation Of course, since we know the free energy
of the non-interacting Fermi gas, we can calculate the form of
the specific heat. Here we will follow Ibach and Lüth 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 finite T , the Fermi gas will gain
energy
Z ∞ Z E
F
U (T ) = 0
dE ED(E)f (E, T ) − 0
dE ED(E) (35)

so
dU Z ∞ df (E, T )
CV = = 0 dE ED(E) . (36)
dT dT
Then since at constant volume the electronic density is constant,
dn R∞
so dT = 0, and n = 0 dED(E)f (E, T ),
dn Z ∞ df (E, T )
0 = EF = 0 dE EF D(E) (37)
dT dT
so we may write
Z ∞ df
CV = 0
dE (E − EF ) D(E) . (38)
dT
In f , the temperature T enters through both β = 1/kB T and
19
µ
df ∂f ∂β ∂f ∂µ
= +
dT ∂β ∂T ∂µ ∂T
 
βeβ(E−µ)  E − µ ∂µ 
= ³ ´2 − (39)
e β(E−µ) +1 T ∂T
∂µ
However, ∂T depends upon the details of the density of states
near the Fermi surface, which can differ greatly from material
∂µ
to material. Furthermore, ∂T < 1 especially in common metals
E−µ
at temperatures T ¿ TF , and the first term T is of order
∂µ
one (c.f. Fig. 7). Thus, for now we will neglect ∂T relative to
E−µ
T (you will explore the validity of this approximation in your
homework), and, consistent with this approximation, replace µ
by EF , so
df βeβ(E−EF ) E − EF
≈³ ´2 , (40)
dT e β(E−E F ) +1 T
and
1 Z∞ (E − EF )2 D(E)βeβ(E−EF )
CV ≈ dE ³ ´2 let x = β(E − EF )
kB T 2 0 e β(E−E F ) +1
  x
Z ∞ x 2 e
≈ kB T −βE dx D  + EF  x x (41)
F β (e + 1)2
x
As shown in Fig. 8, the function x2 (exe+1)2 is only large in the
region −10 < x < 10. In this region, and for temperatures
20
0.5
0.4
2
x e /(e +1)

0.3
x

0.2
2 x

0.1
0.0
-10 -5 0 5 10
x
x
Figure 8: Plot of x2 (exe+1)2 vs. x. Note that this function is only finite for roughly
−10 < x < 10. Thus, at temperatures T ¿ TF ∼ 105 K, we can approximate
³ ´
x
D β
+ EF ≈ D (EF ) in Eq. 42.

µ ¶
x
T ¿ TF , D β + EF ≈ D (EF ), since the density of states
usually does not have features which are sharp on the energy
scale of 10kB T . Thus
Z ∞
2 ex
CV ≈ kB T D (EF ) dx x x
−βEF (e + 1)2
π2 2
≈ k T D(EF ) . (42)
3 B
Note that no assumption about the form of D(E) was made
other than the assumption that it is smooth within kB T of the
Fermi surface. Thus, experimental measurements of the specific
heat at constant volume of the electrons, gives us information
21
about the density of electronic states at the Fermi surface.
Now let’s reconsider the DOS for the 3-D box potential.
  1/2  
1  2m 3/2 1/2 E
D(E) = 2 E = D(EF )   (43)
2π h̄ EF
R EF
For which n = 0 D(E)dE = D(EF ) 23 EF , so
π2 T 3
CV = nkB ¿ nkB (44)
2 TF 2
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 θD ,
3
CV phonon ∼ nkB À CV electron (45)
2
So, the only way to measure the electronic specific heat in most
materials is to go to very low temperatures T ¿ θD , for which
CV phonon ∼ T 3. Here the total specific heat

CV ≈ γT + βT 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 effect of electronic correlations, con-


sider the effect 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 2s22p4 to a 2s22p6 configuration.
The defect will then become a cation, and have a net charge
O

Cu O Cu O Cu O Cu O Cu O Cu O

O O O O O O

Cu O Cu O Cu O Cu O Cu O Cu O

q=2e-
O O O O O O O

Cu O Cu O Cu O Cu O Cu O Cu O

O O O O O O

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 2s2 2p4 to a 2s2 2p6 configuration.

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
near charged Away from
defect charged defect

e
EF

-eδU

Figure 10: The shift in the DOS parabola near a charged defect.

charge to flow away from the defect. Near the defect (since
e < 0 and hence eδU (rnear ) < 0)
Z E +eδU (r
F near )
n(rnear ) ≈ 0
D(E)DE (47)

24
While away from the defect, δU (raway ) = 0, so
Z E
F
n(raway ) ≈ 0
D(E)DE (48)

or
Z E +eδU (r) Z E
F F
δn(r) ≈ 0
D(E)DE − 0
D(E)DE (49)

If |eδU | ¿ EF , then

δn(r) ≈ D(EF ) [EF + eδU − EF ] = eδU D(EF ) . (50)

We can solve for the change in the electrostatic potential by


solving Poisson equation.

∇2δU = 4πδρ = 4πeδn = 4πe2D(EF )δU . (51)

Let λ2 = 4πe2D(EF ), then ∇2δU = λ2δU has the solution3


qe−λr
δU (r) = (52)
r
The length 1/λ = rT F is known as the Thomas-Fermi screening
length.
µ ¶−1/2
2
rT F = 4πe D(EF ) (53)
3
The solution is actually Ce−λr /r, where C is a constant. C may be deterined by letting D(EF ) =
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,
a0 π a0
rT2 F = ≈
3(3π 2n)1/3 4n1/3
 
1  n −1/6
rT F ≈ (54)
2 a30

In Cu, for which n ≈ 1023 cm−3 (and since a0 = 0.53 A)
³ ´−1/6
1 1023 −8 ◦
rT F Cu ≈ ≈ 0.5 × 10 cm = 0.5 A (55)
2 (0.5 × 10−8)−1/2
Thus, if we add a charge defect to Cu metal, the effect of the
1 ◦
defect’s ionic potential is screened away for distances r > 2 A.

r r
/r

/r
-r/rTF

rTF=1/4
-r/rTF

rTF=1
-1/6
-e

rTF= n
-e

bound states
free states

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
rT F ∼ n−1/6 . (56)

This will extend the range of the potential, causing it to trap


or bind more states–making 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 3He, which
due its nuclear spin (the nucleus has two protons, one neutron),
obeys Fermi statistics for sufficiently low energies or temper-
atures. In addition, simple metals, from the first or second
column of the periodic table, for which we may approximate
the ionic potential
V (R) = V0 (57)

are a close approximant to Fermi liquids.


Moreover, Fermi Liquid theory only describes the ”gaseous”
phase of these quantum fermion systems. For example, 3He also
has a superfluid (triplet), and at least in 4He-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 fluidity.
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 À 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 different
single particle states.
1
ψp(r) = √ eip·r/h̄ (58)
V
The occupation of each of these states is given by np = θ(p−pF )
where pF is the radius of the Fermi sphere. The energy of the
system is
X p2
E= np (59)
p 2m
and pF is given by
N 1 Ã p F !3
= 2 (60)
V 3π h̄
Now lets add a particle to the lowest available state p = pF
then, for T = 0,
∂E0 p2F
µ = E0(N + 1) − E0(N ) = = . (61)
∂N 2m
If we now excite the system, we will promote a certain number
of particles across the Fermi surface SF yielding particles above
and an equal number of vacancies or holes below the Fermi
surface. These are our elementary excitations, and they are

30
quantified by δnp = np − n0p





 δp,p0 for a particle p0 > pF
δnp = 

. (62)
 0

 −δp,p0 for a hole p < pF
If we consider excitations created by thermal fluctuations, then
E particle
excitation
δn p’ = 1
EF
δn p= -1
hole
excitation

D(E)

Figure 12: Particle and hole excitations of the Fermi gas.

δnp ∼ 1 only for excitations of energy within kB T of EF . The


energy of the non-interacting system is completely characterized
as a functional of the occupation
X p2 0 X p
2
E − E0 = (np − np) = δnp . (63)
p 2m p 2m

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
2
δF = p’ /2m - µ
2
EF = µ = p /2m
2 F
δF = µ - p /2m

2
δF = | µ - p /2m |

D(E)

Figure 13: Since µ = p2F /2m, the free energy of a particle or a hole is δF =
|p2 /2m − µ| > 0, so the system is stable to these excitations.

which for T = 0, is F = E − µN , and


 
X p2
F − F0 =  − µ
 δnp . (64)
p 2m
The free energy of a particle, with momentum p and δnp0 =
p2
δp,p0 is 2m − µ and it corresponds to an excitation outside SF .
p 2
The free energy of a hole δnp0 = −δp,p0 is µ − 2m , which corre-
sponds to an excitation within SF . However, since µ = p2F /2m,
the free energy of either at p = pF is zero, hence the free energy
of an excitation is
¯ ¯
¯ 2 ¯
p /2m − µ ,
¯
¯
¯
¯ (65)

which is always positive; ie., the system is stable to excitations.


32
3.3.2 Quasiparticles and Quasiholes at T = 0

2 -a/r TF
U ≈ ea e
a

Figure 14: Model for a fermi liquid: a set of interacting particles an average distance
a apart bound within an infinite square-well potential.

Now let’s consider a system with interacting particles an av-


erage distance a apart, so that the characteristic energy of in-
e2 −a/rT F
teraction is ae . We will imagine that this system evolves
slowly from an ideal or noninteracting system in time t (i.e. the
e2 −a/rT F
interaction U ≈ ae 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 n0p,
then the interacting system eigenstate will evolve quasistatis-
tically from n0p to np. In fact if the system is isotropic and

33
remains in its ground state, then n0p = np. However, clearly in
some situations (superconductivity, magnetism) we will neglect
some eigenstates of the interacting system in this way.
Now let’s add a particle of momentum p to the non-interacting
ideal system, and slowly turn on the interaction. As U is

p p p

time = 0 time = t
2
U=0 U = (e /a) exp(-a/rTF)

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 effective 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
difficulties if the lifetime of the quasi-particle τ < t. If so,
then the process is not reversible. If we shorten t so that again
τ À 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
difficulties 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
p2 p2F
Fermi energy ε = 2m − 2m ≈ v(p − pF ).
To estimate this lifetime consider the following argument
from AGD: A particle with momentum p1 above the Fermi
surface (p1 > pF ) interacts with one of the particles below the
Fermi surface with momentum p2. As a result, two new par-
ticles appear above the Fermi surface (all other states are full)

35
p4

p3

p1 p2

Figure 16: A particle with momentum p1 above the Fermi surface (p1 > pF ) interacts
with one of the particles below the Fermi surface with momentum p2 . As a result, two
new particles appear above the Fermi surface (all other states are full) with momenta
p3 and p4 ..

with momenta p3 and p4. This may also be interpreted as a


particle of momentum p1 decaying into particles with momenta
p3 and p4 and a hole with momentum p2. By Fermi’s golden
rule, the total probability of such a process if proportional to
1 Z
∝ δ (ε1 + ε2 − ε3 − ε4) d3p2d3p3 (66)
τ
p21
where ε1 = 2m − EF , and the integral is subject to the con-
straints of energy and momentum conservation and that

p2 < p F , p3 > pF , p4 = |p1 + p2 − p3| > pF (67)


36
It must be that ε1 + ε2 = ε3 + ε4 > 0 since both particles 3
and 4 must be above the Fermi surface. However, since ε2 < 0,
< ε1 is also small, so only of order
if ε1 is small, then |ε2| ∼
ε1 /EF states may scatter with the state k1, conserve energy,
and obey the Pauli principle. Thus, restricting ε2 to a narrow
shell of width ε1 /EF near the Fermi surface, and reducing the
scattering probability 1/τ by the same factor.
Now consider the constraints placed on states k3 and k4 by
momentum conservation

k1 − k 3 = k 4 − k 2 . (68)

Since ε1 and ε2 are confined to a narrow shell around the Fermi


surface, so too are ε3 and ε4 . This can be seen in Fig. 17, where
the requirement that k1 − k3 = k4 − k2 limits the allowed
states for particles 3 and 4. If we take k1 fixed, then the allowed
states for 2 and 3 are obtained by rotating the vectors k1 −k3 =
k4 −k2; 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 final states further reduces the
scattering probability by a factor of ε1/EF .
37
E
ky
k1 k4
k3
EF 3
k2 k 1- k3
2 1
k 4- k2
4 kx

N(E)

Figure 17: A quasiparticle of momentum p1 decays via a particle-hole excitation


into a quasiparticle of momentum p4 . This may also be interpreted as a particle
of momentum p1 decaying into particles with momenta p3 and p4 and a hole with
momentum p2 . Energy conservation requires |ε2 | ∼ < ε1 . Thus, restricting ε2 to a
narrow shell of width ε1 /EF near the Fermi surface. Momentum conservation k1 −
k3 = k4 − k2 further restricts the available states by a factor of about ε1 /EF . Thus
³ ´−2
ε1
the lifetime of a quasiparticle is proportional to EF
.

µ ¶
ε1 2
Thus, the scattering rate 1/τ is proportional to EF so that
excitations of sufficiently small energy will always be sufficiently
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 ≈ RT F ), also significantly reduces the scattering rate.

38
3.4 Energy of Quasiparticles.

As in the non-interacting system, excitations will be quanti-


fied by the deviation of the occupation from the ground state
occupation n0p
δnp = np − n0p . (69)

At low temperatures δnp ∼ 1 only for p ≈ pF where the par-


ticles are sufficiently long lived that τ À t. It is important to
emphasize that only δnp not n0p or np, will be physically rele-
vant. This is important since it does not make much sense to
talk about quasiparticle states, described by np, far from the
Fermi surface since they are not stable.
For the ideal system
X p2
E − E0 = δnp . (70)
p 2m
For the interacting system E[np] becomes much more compli-
cated. If however δnp is small (so that the system is close to
its ground state) then we may expand:
X
E[np] = Eo + ²pδnp + O(δn2p) , (71)
p

39
where ²p = δE/δnp. Note that ²p is intensive (ie. it is indepen-
dent of the system volume). If δnp = δp,p0 , then E ≈ E0 + ²p0 ;
i.e., the energy of the quasiparticle of momentum p0 is ²p0 .
In practice we will only need ²p near the Fermi surface where
δnp is finite. So we may approximate

²p ≈ µ + (p − pF ) · ∇p ²p|pF (72)

where ∇p²p = vp, the group velocity of the quasiparticle. The


ground state of the N + 1 particle system is obtained by adding
∂E0
a particle with ²p = ²F = µ = ∂N (at zero temperature); which
defines the chemical potential µ. We make learn more about
²p by employing the symmetries of our system. If we explicitly
display the spin-dependence,

²p,σ = ²−p,−σ under time-reversal (73)


²p,σ = ²−p,σ under BZ reflection (74)

So ²p,σ = ²−p,σ = ²p,−σ ; i.e. in the absence of an external


magnetic field, ²p,σ does not depend upon σ if. Furthermore,
for an isotropic system ²p depends only upon the magnitude of
p d²p (|p|)
p, |p|, so p and vp = ∇²p(|p|) = |p| d|p| are parallel. Let us

40
define m∗ as the constant of proportionality at the fermi surface

vpF = pF /m∗ (75)

Using m∗ it is useful to define the density of states at the


fermi surface. Recall, that in the non-interacting system,
 
1  2m 3/2 1/2 mpF
D(EF ) = 2 EF = (76)
2π h̄2 πh̄3
where p = h̄k, and E = p2/2m. Thus, for the interacting
system at the Fermi surface
m∗ p F
Dinteracting (EF ) = , (77)
πh̄3
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 effective
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 − F0 = E − E0 − µ(N − N0) . (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,
X
N − N0 = δnp , (79)
p

and since
X
E − E0 ≈ ²pδnp , (80)
p

we get
X
F − F0 ≈ (²p − µ) δnp . (81)
p

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 δ. For our theory/expansion to remain valid, we must
42
δ

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


1 P
that N p |δnp | ¿ 1.

have
1 X
|δnp| ¿ 1 . (82)
N p
6 0, ²p − µ will also be of order δ. Thus,
Where δnp =
X
(²p − µ) δnp ∼ O(δ 2) , (83)
p

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.
X 1 X
F − F0 = (²p − µ) δnp + fp,p0 δnpδnp0 + O(δ 3) (84)
p 2 p,p0
43
where
δE
fp,p0 = (85)
δnpδnp0
The term, proportional to fp,p0 , 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 fp,p0 ∼ 1/V . However,
it is also clear that fp,p0 is an interaction between quasiparticles,
each of which is spread out over the whole volume V , so the
probability that they will interact is ∼ rT3 F /V , thus

fp,p0 ∼ rT3 F /V (86)

In general, since δnp is only of order one near the Fermi


surface, we will only care about fp,p0 on the Fermi surface (as-
suming that it is continuous and changes slowly as we cross the
Fermi surface.
¯
Interested in fp,p0 ¯¯²p=² in only! (87)
p0 =µ

Given this, we can reduce the spin dependence of fp,p0 to


a symmetric and anti symmetric part. First in the absence of
an external field, the system should be invariant under time-

44
reversal, so
fpσ,p0 σ0 = f−p−σ,−p0 −σ0 , (88)

and, in a system with reflection symmetry

fpσ,p0 σ0 = f−pσ,−p0 σ0 . (89)

Then
fpσ,p0 σ0 = fp−σ,p0 −σ0 . (90)

It must be then that f depends only upon the relative orienta-


tions of the spins σ and σ 0, so there are only two independent
components fp↑,p0 ↑ and fp↑,p0 ↓. We can split these into sym-
metric and antisymmetric parts.

a 1³ ´
s 1³ ´
fp,p 0 = fp↑,p0 ↑ − fp↑,p0 ↓ fp,p 0 = fp↑,p0↑ + fp↑,p0 ↓ .
2 2
(91)
a
fp,p 0 may be interpreted as an exchange interaction, or

s 0 a
fpσ,p0 σ0 = fp,p 0 + σ · σ fp,p0 (92)

where σ and σ 0 are the Pauli matrices for the spins.


a
Our ideal system is isotropic in momentum. Thus, fp,p 0 and

s 0
fp,p 0 will only depend upon the angle θ between p and p , and

45
a s
so we may expand either fp,p 0 and fp,p0


X
α
fp,p 0 = flα Pl (cos θ) . (93)
l=0

Conventionally these f parameters are expressed in terms of


reduced units.
V m ∗ pF α
D(EF )flα = 2
α
3 fl = F l . (94)
π h̄

4.2 Local Energy of a Quasiparticle

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 δnp0 . To this, add another quasi-
particle of momentum p (δn0p → δn0p + δp,p0 ). From Eq. 84 the
46
free energy of the additional quasiparticle is
X
²̃p − µ = ²p − µ + fp0,p δnp0 , (95)
p0

(recall that fp,p0 = fp0 ,p ). Both terms here are O(δ). 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 ²̃p plays the part of the local energy of a quasi-
particle. For example, the gradient of ²̃p is the force the system
exerts on the additional quasiparticle. When the quasiparticle
is added to the system, the system is inhomogeneous so that
δnp0 = δnp0 (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
X
−∇r ²̃ = −∇r fp0,p δnp0 (r) (96)
p0

on the added quasiparticle resulting from interactions with other


quasiparticles.
47
4.2.1 Equilibrium Distribution of Quasiparticles at Finite T

²̃p also plays an important role in the finite-temperature prop-


erties of the system. If we write
X 1 X
E − E0 = ²pδnp + fp0 ,pδnp0 δnp (97)
p 2 p,p0
P
Now suppose that p |hδnp i| ¿ N , as indeed it must be for
the expansion above to be valid, so that

δnp = hδnpi + (δnp − hδnpi) (98)

where the first term is O(δ), and the second O(δ 2). Thus,

δnpδnp0 ≈ −hδnpihδnp0 i + hδnpiδnp0 + hδnp0 iδnp (99)

We may use this to rewrite the energy of our interacting system


X 1 X X
E − E0 ≈ ²pδnp − fp0,p hδnpihδnp0 i + fp0,p hδnpiδnp0
p 2 p,p0 p,p0
 
X X 1 X
≈ ²p + fp0,p hδnp0 i δnp − fp0 ,phδnpihδnp0 i
p p 0 2 p,p 0

X 1 X
≈ h²̃piδnp − fp0,p hδnpihδnp0 i + O(δ 4) (100)
p 2 p,p0
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
effect). We will obtain
1
np(T, µ) = , (101)
1 + exp β(h²̃pi − µ)
or
1
δnp(T, µ) = − θ(pf − p) . (102)
1 + exp β(h²̃p i − µ)
However, at least for an isotropic system, this expression bears
closer investigation. Here, the molecular field (evaluated within
kB T of the Fermi surface)
X
h²̃p − ²pi = fp0 ,phδnp0 i (103)
p0

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
X
h²̃p − ²pi = fp0,p hδnp0 i
p0
XZ
∝ d3pfl Pl (cos θ)hδnp0 i
l Z
∝ f0 d3phδnp0 i = 0
(104)

49
In going from the second to the third line above, we made use of
the isotropy of the system, so that hδnp0 i is independent of the
angle θ. The evaluation in the third line, follows from particle
number conservation. Thus, to lowest order in δ
1
np(T, µ) = + O(δ 3) (105)
1 + exp β(²p − µ)

4.3 Effective 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
X Z d3 p
→V . (106)
p (2πh̄)3
The net momentum of the volume V of quasiparticles is
Z d3 p
Pqp = 2V pnp net quasiparticle momentum (107)
(2πh̄)3
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
Z d3 p
Jqp = Jp = 2V vpnp net quasiparticle and particle current
(2πh̄)3
(108)
or, since the momentum is just the particle mass times this
current
d3 p
Z
Pp = 2V m vpnp net quasiparticle and particle current
(2πh̄)3
(109)
where vp = ∇p²̃p, is the velocity of the quasiparticle. So
Z d3 p Z d3 p
pnp = m ∇p²̃pnp (110)
(2πh̄)3 (2πh̄)3
Now make an arbitrary change of np and recall that ²̃p depends
upon np, so that
XZ d3 p
δ²̃p = V fp,p0 δnp0 . (111)
σ0 (2πh̄)3
For Eq. 110, this means that
Z d3 p Z d3 p
pδnp = m ∇p²̃pδnp (112)
(2πh̄)3 (2πh̄)3
Z d 3 p X Z d 3 p0 ³ ´
+mV ∇ f
p p,p 0 δn p np ,
0
(2πh̄)3 σ0 (2πh̄)3

51
or integrating by parts (and renaming p → p0 in the last part),
we get
Z d3 p p Z d3 p
δnp = ∇p²̃pδnp (113)
(2πh̄)3 m (2πh̄)3
XZ d 3 p0 Z d 3 p
−V 3 3
δnpfp,p0 ∇p0 np0 ,
σ 0 (2πh̄) (2πh̄)
Then, since δnp is arbitrary, it must be that the integrands
themselves are equal
p X Z d 3 p0
= ∇p²̃p − V fp,p0 ∇p0 np0 (114)
m σ0 (2πh̄)3
0
The factor ∇p0 np0 = − pp0 δ(p0 − pF ). 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 = pF , let θ
be the angle between p (or the z-axis) and p0, and finally note
¯ ¯
¯ ¯
that on the Fermi surface ∇ ²̃ |
¯
¯
¯
p p p=pF ¯ = vF = pF /m∗ . Thus,

pF pF X Z p02 dpdΩ p0 0
= ∗+ 3
fpσ,p0 σ0 0 δ(p − pF ) (115)
m m σ 0 (2πh̄) p

52
However, since both p and p0 are restricted to the Fermi surface
p0
p0 = cos θ, and evaluating the integral over p, we get
1 1 V pF X Z dΩ
= ∗+ fpσ,p0 σ0 cos θ , (116)
m m 2 σ,σ 0 (2πh̄)3
1
where the additional factor of 2 compensates for the additional
spin sum. If we now sum over both spins, σ and σ 0, only the
symmetric part of f survives (the sum yields 4f s), so
1 1 4πV pF Z
= ∗+ d (cos θ) f s(θ) cos θ , (117)
m m (2πh̄)3
We now expand f in a Legendre polynomial series
X
f α (θ) = flα Pl (cos θ) , (118)
l

and recall that P0(x) = 1, P1(x) = x, .... that


Z 1 2
−1
dxPn(x)Pm(x)dx = δnm (119)
2n + 1
and finally that
V m ∗ pF α
D(0)flα = 2
α
3 fl = F l , (120)
π h̄
we find that
1 1 F1s
= + , (121)
m m∗ 3m∗
53
Quantity Fermi Liquid Fermi Liquid/Fermi Gas
m∗ pF 2 CV m∗
Specific Heat Cv = k T
3h̄3 B CV 0 = m =1+ F1s/3
κ 1+F0s
Compressibility κ0 = 1+F0s /3
µ ¶2
p2F c 1+F0s
Sound Velocity c2 = 3mm∗ (1 + F0s) c0 = 1+F1s /3
m∗ pF β 2 χ 1+F1s /3
Spin Susceptibility χ = π 2 h̄3 1+F0a χ0 = 1+F0a
Table 1: Fermi Liquid relations between the Landau parameters Fnα 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 + F1s/3.
The effective mass cannot be experimentally measured di-
rectly; however, it appears in many physically relevant measur-
able quantities, including the specific heat
 
∂E/V  1 ∂ X
CV = 
  = ²̃pnp. (122)
∂T V N V ∂T p

To lowest order in δ, we may neglect fp,p0 in both ²̃p and np,


so
1 X ∂np
CV = ²p . (123)
V p ∂T
P
Recall that the density of states D(E) = p δ(E−²p), and mak-
ing the same assumption that we made for the non-interacting
54
∂µ
system, that ∂T is negligible, we get,
1 Z ∂ 1
CV = d²D(²)² . (124)
V ∂T exp β(² − µ) + 1
This integral is identical to the one we had to evaluate for the
non-interacting system, and yields the result
π2 2
CV = k T D(EF )
3V B
kB2 T m∗pF
= . (125)
3h̄3
Thus, measuring the electronic contribution to the specific heat
CV yields information about the effective mass m∗, and hence
F1s. 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 ψ(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

V(r)
V(r) = V
0

E E
Ef metal
E insulator
E f
f
Ef "heavy" metal

D(E) D(E)

Figure 1: The additional effects of the lattice potential can have a profound effect 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 effects of a small electronic potential U . In this
chapter we will set U = 0, and consider the effects of the ion po-
tential V (r). As shown in Fig. 1, additional effects of the lattice
potential can have a profound effect 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 ψ(r)

From the symmetry of the electronic potential V (r) one may


infer some of the properties of the electronic wave functions
ψ(r).
Due to the translational symmetry of the lattice V (r) is pe-
riodic

V (r) = V (r + rn), r n = n 1 a1 + n 2 a2 + n 3 a3 (1)

and may then be expanded in a Fourier expansion


X
V (r) = VGeiG·r , G = hg1 + kg2 + lg3 , (2)
G

which, since G · rn = 2πm (m ∈ Z) guarantees V (r) =


P ik·r
V (r + rn). Given this, and letting ψ(r) = k Ck e the

3
Schroedinger equation becomes
 
2
h̄ 2
Hψ(r) = − ∇ + V (r) ψ = Eψ (3)
2m
X h̄2k 2 X 0 X
⇒ Ckeik·r + Ck0 VGei(k +G)·r = E Ckeik·r , k0 → k−G
k 2m k0 G k
(4)
or   
2 2
X h̄ k

ik·r  X


e − E
  Ck + VGCk−G  = 0∀r (5)
k 2m 
G 

Since this is true for any r, it must be that


 
2 2
 h̄ k X
 − E
 Ck + VGCk−G = 0, ∀k (6)
2m G

Thus the potential acts to couple each Ck only with its recip-
rocal space translations Ck+G and the problem decouples in to
N independent problems for each k in the first BZ. Ie., each of
the N problems has a solution which is a sum over plane waves
whos’ wave vectors differ only by G’s. Thus the eigenvalues
may be indexed by k.

Ek = E(k), I.e. k is still a good q.n.! (7)

We may now sum over G to get ψk with the eigenvector sum

4
X X X

X X X

First B.Z.

Figure 2: The potential acts to couple each Ck with its reciprocal space translations
Ck+G (i.e. x → x, • → •, and ° → °) and the problem decouples into N indepen-
dent problems for each k in the first BZ.

restricted to reciprocal lattice sites k, k + G, . . .


 
X X
ψk(r) = Ck−G ei(k−G)·r =  Ck−G e−iG·r  eik·r (8)
G G

ψk(r) = Uk(r)eik·r , where Uk(r) = Uk(r + rn) (9)

Note that if V (r) = 0, U (r) = √1 . This result is called Bloch’s


V
Theorem; ie., that ψ may be resolved into a plane wave and a
periodic function. Its consequences as follows:
 
X −i(G0 −k−G)·r X −iG00 ·r
ψk+G(r) = Ck+G−G0 e = Ck−G00 e  eik·r
G0 G00
= ψk(r), where G00 ≡ G0 − G (10)

5
Ie., ψk+G(r) = ψk(r) and as a result

Hψk = E(k)ψk ⇒ Hψk+G = E(k + G)ψk+G (11)


= Hψk = E(k + G)ψk+G (12)

Thus E(k + G) = E(k) : E(k) is periodic then since both


ψk(r) and E(k) are periodic in reciprocal space, one only needs
knowledge of them in the first BZ to know them everywhere.

2 The nearly free Electron Approximation.

If the potential is weak, VG ≈ 0 ∀G, then we may solve the


VG = 0 problem, subject to our constraints of periodicity, and
treat VG as a perturbation.
When VG = 0, then
h̄2k2
E(k) = free electron (13)
2m
However, we must also have that (if VG 6= 0)
h̄2
E(k) = E(k + G) ≈ |k + G|2 (14)
2m
Ie., the possible electron states are not restricted to a single
parabola, but can be found equally well on paraboli shifted by
6
E

First BZ 2π/a

Figure 3: For small VG , we may approximate the band structure as composed of N


parabolic bands. Of course, it is sufficient to consider this in the first Brillouin zone,
where the parabola centered at finite G cross at high energies. To understand the
effects of the perturbation VG 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 sufficient


to represent this in the first zone only. For example in a 3-D
cubic lattice the energy band structure along kx(ky = kz = 0)
is already rather complicated within the first zone. (See Fig.4.)

The effect of VG can now be discussed. Let’s return to the 1-d


problem and consider the edges of the zone where the [Φparaboli
π
intersect. (See Fig. 3.) An electron state with k = a will
involve at least the two G values G = 0, 2π
a . Of course, the

7
First B.Z.

-π⁄a π⁄a

−π⁄a π⁄a kx

Figure 4: The situation becomes more complicated in three dimensions since there are
many more bands and so they can cross the first zone at lower energies. For example
in a 3-D cubic lattice the energy band structure along kx (ky = kz = 0) is already
rather complicated within the first zone.

exact solution must involve all G since


 
2 2
 h̄ k X
 − Ek 
 Ck + VGCk−G = 0 (15)
2m G

We can generally take V0 = 0 since this just sets a zero for the
h̄2 k2
potential. Then, those G for which Ek = Ek−G ≈ 2m are
going to give the largest contribution since
X Ck−G
Ck = VG h̄2k2 (16)
2m − Ek−G
G
Ck−G1
Ck ∼ VG1 h̄2k2 (17)
2m − Ek−G1

8
X Ck−G1−G
Ck−G1 = VG h̄2k2 (18)
2m − Ek−G−G1
G
Ck
Ck−G1 ∼ V−G1 h̄2k2 (19)
2m − Ek
Thus to a first approximation, we may neglect the other Ck−G,
and since VG = V−G (so that V (r) is real) |Ck| ≈ |Ck−G1 | À
other Gk−G



X


 (eiGx/2 + e−iGx/2 ) ∼ cos πx
a
ψk(r) = Ck−G ei(k−G)·r ∼ 

G 

 (eiGx/2 − e−iGx/2 ) ∼ sin πx
a
(20)
The corresponding electron densities are sketched in Fig. 5.
Clearly ρ+ has higher density near the ionic cores, and will
be more tightly bound, thus E+ < E−. Thus a gap opens in
Ek near k = G2 .

2.1 The Origin of Band Gaps

Now let’s reexamine this gap at k = G1/2 in a quantitative


manner. Start with the eigen value equation shifted by G.
 
h̄2 X X

Ck−G Ek − |k − G|2 
 = VG0 Ck−G−G0 = VG0−GCk−G0
2m G0 G0
(21)
9
ρ+ (x)
E

Gap!

ρ− (x)

V(x)

D(E)

Figure 5: ρ+ ∼ cos2 (πx/a) has higher density near the ionic cores, and will be more
G
tightly bound, thus E+ < E− . Thus a gap opens in Ek near k = 2
.

P
G0 VG0−GCk−G0
Ck−G = Ã ! (22)
h̄2
Ek − 2m |k − G|2
h̄2 k2
To a first approximation (VG ' 0) let’s set E = 2m (a free-
electron energy) and ignore all but the largest Ck−G ; ie., those
for which the denominator vanishes.

k2 = |k − G|2 , (23)

10
or in 1-d
2π 2 π
k2 = (k −) or k = − (24)
a a
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 first B.Z. Thus the equation above also tells us

highly perturbed

essentially free electrons


Figure 6: We can satisfy the condition Ek ' Ek−G only for k on the edge of the B.Z..
Here the lattice potential strongly perturbs the electronic states (i.e. more than one
Ck−G is finite).

that Ck and Ck−G1 are the most important coefficients (if this
electronic state was unperturbed, only Ck would be important).
Thus approximately for VG ∼ 0, V0 ≡ 0 and for k near the

11
zone boundary
 

 h̄2k 2 

G=0 Ck  E−  = VG1 Ck−G1 (25)
 2m 

 
2 2
h̄ |k − G1|

 
G = G1 Ck−G1  E−  = V−G1 Ck, (26)
 2m 

Again, ignore all other CG. This is a secular equation which


has a nontrivial solution iff
¯ Ã ! ¯
¯ ¯
¯ h̄2 k2 ¯
¯
¯ 2m −E V G1 ¯
¯
¯
¯ à ! ¯¯ =0 (27)
¯ h̄2 |k−G1 |2 ¯
¯
¯ V
−G1 2m −E ¯
¯

or ¯
¯
¯
¯
¯ ¯
¯
¯ Ek0 −E V G1 ¯
¯
¯
¯
¯
¯
¯
¯
=0 (28)
0
¯
¯ V−G1 Ek−G 1
−E ¯
¯

(V−G = VG∗ , so thatV (r) ∈ <)


0
(Ek0 − E)(Ek−G 1
− E) − |V G 1 | 2
=0 (29)
µ ¶
0
Ek0 Ek−G 1
−E Ek0 + 0
Ek−G 1
+ E 2 − |VG1 |2 = 0 (30)
 1
± 1µ 0 ¶ 1 µ ¶
0 2
2
E = Ek−G1 + Ek ±  Ek−G − Ek + |VG1 |2 (31)
0 0
2 4
0
At the zone boundary, where Ek−G 1
= Ek0 , the gap is

∆E = E+ − E− = 2|VG1 | (32)

12
E
k

e−
2 VG k

-π/a 0 π/a k

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 (kf − k0 =
G) is satisfied.
| − 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-

State
Energies

Separation

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
V (r)
A

E i ψi

r r

Atomic cores Valence electrons

Figure 9: In the tight-binding approximation, we generally ignore the core electron


dynamics and consider only the ionic core potential. For now let’s assume that there
is only one valence orbital φi 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 let’s assume that there is only one valence orbital
φi 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 φi, and eigen energies Ei. The
unperturbed Schroedinger equation for the nth atom is

HA(r − rn) · φi(r − rn) = Eiφi(r − rn) (34)


15
There is a weak perturbation v(r − rn) coming from the atomic
potentials of the other atoms rm 6= rn
h̄2∇2
H = HA + v = − + VA(r − rn) + v(r − rn) (35)
2m
X
v(r − rn) = VA(r − rm) (36)
m6=n
We now seek solutions of the Schroedinger equation indexed by
k (Bloch’s theorem)

Hψk(r) = E(k)Ψk(r) (37)


Z hψk|H|ψki
⇒ ψ∗ ⇒ E(k) = (38)
hψk|ψki
where
Z
hψk|ψki ≡ d3rψk∗ (r)ψk(r)
Z
hψk|H|ψki ≡ d3rψk∗ (r)Hψk(r) (39)

Of course, this problem is almost hopelessly complicated. We


cannot solve for ψk. Rather, we will solve for some φk ' ψk
where the parameters of φk are determined by minimizing
hφk|H|φki
≥ E(k). (40)
hφk|φki

16
This is called the Raleigh-Ritz variational principle.
Consistent with our original motivation, we will approximate
ψk with a sum over atomic states.
X X
ψk ' φ k = anφi(r − rn) = eik·rn φi(r − rn) (41)
n n

ψk(r) = Uk(r)eik·r , ψk(r) = ψk+G(r)

Where φk must be a Bloch state φk+G = φk which dictates our


choice an = eik·rn . Thus at this level of approximation we have
no free parameters to vary to minimize hφk|H|φki / hφk|φki ≈
E(k).
Using φk as an approximate state the energy denominator
hφk|φki, becomes
X Z
ik·(rn −rm )
hφk|φki = e d3rφ∗i (r − rm)φi(r − rn) (42)
n,m

Let’s imagine that the valance orbital of interest, φi, has an


very small overlap with adjacent atoms so that
XZ
hφk|φki ' d3rφ∗i (r − rn)φi(r − rn) = N (43)
n

The last identity follows since φi is normalized.

17
φ (r-r )
i 1 φi (r-r2 )

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


1 X ik·(rn −rm )
Z
E(k) ≈ e d3rφ∗i (r−rm) {Ei + v(r − rn)} φi(r−rn) .
N n,m
(44)
Again, in the first part (involving Ei), we may neglect orbital
overlap. For the second term, involving v(r − rn), the over-
lap should be included, but only to the nearest neighbors of
each atom (why?). In the simplest case, where the orbitals φi,
are s-orbitals, then we can use this symmetry to reduce the
complexity of the problem to just two more integrals since the
hybridization (Bi) will be the same in all directions.
Z
Ai = − φ∗i (r − rn)v(r − rn)φi(r − rn)d3r ren. Ei (45)

18
B
i

Bi
B A
i i

B
i

Figure 11: A simple cubic tight binding lattice composed of s-orbitals, with overlap
integral Bi .
Z
Bi = − φ∗i (r − rm)v(r − rn)φi(r − rn)d3 r (46)

Bi describes the hybridization of adjacent orbitals.

Ai; Bi > 0, since v(r − rn) < 0 (47)

Thus
X
E(k) ' Ei − Ai − Bi eik(rn−rm) sum over m n.n. to n
m
(48)
Now, if we have a cubic lattice, then

(rn − rm) = (±a, 0, 0)(0, ±a, 0)(0, 0, ±a) (49)


19
so

E(k) = Ei − Ai − 2Bi{cos kxa + cos ky a + cos kz a} (50)

Thus a band centered about Ei − Ai of width 12Bi is formed.


Near the band center, for k-vectors near the center of the zone
we can expand the cosines cos ka ' 1 − 12 (ka)2 + · · · and let
k 2 = kx2 + ky2 + kz2, so that

E(k) ' Ei − Ai + Bia2k 2 (51)

The electrons near the zone center act as if they were free with
a renormalized mass.
h̄2k 2 2 2 1
= B i a k , i.e. ∝ curvature of band (52)
2m∗ m∗
For this reason, the hybridization term Bi 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 Bi, 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 effective masses (small Bi).
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 filled, the Fermi
surface will be spherical.

The bands are filled 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
1s2 2s22p63s1 atomic configuration the 1s, 2s and 2p orbitals
evolve into (narrow) filled bands, but the 3s1 band will only
be half full, and thus it evolves into a metal. Mg 1s2 2s22p63s2
also metal since the p and s band overlaps the unfilled d-band.
There are exceptions to this rule. Consider C with atomic con-
figuration of 1s2 2s22p2. Its valance s and p states form a strong
sp3 hybrid band which is further split into a bonding and anti-

21
Atomic Potential Tight Binding Bands

E Ek
2 111

E 111
1
A2
k 12 B 2

k=0

E 1k
111
E
2
a A1
12 B1
k
- π/a π/a

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)3d10 4s the d-orbitals are rather small whereas the valence
s-orbitals have a large extent . As a result the s-s hybridization

22
E
sp3antibonding

sp3 bonding

ra a

Figure 14: C (diamond) with atomic configuration of 1s2 2s2 2p2 . Its valance s and p
states form a strong sp3 hybrid band which is split into a bonding and anti-bonding
band.

Biss: is strong and the Bidd is weak.

Bidd ¿ Biss (53)

In addition the s-d hybridization is inhibited by the opposing


symmetry of the s-d orbitals.
Z
Bisd = φsi(r − r1)v(r − r2)φdi(r − r2)d3r ¿ Biss (54)

where φsi is essentially even and φdi is essentially odd. So Bisd ¿


Biss. Thus, to a first 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
D(E)

d
1 +
− −
2
+ + + +
− −


s
+ +

Figure 15: Schematic DOS of Cu 3d10 4s1 . The narrow d-band feature is split due to
crystal fields.

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(EF ) = 0,
we have an insulator (semiconductor). The effective electronic

24
synchrotron

d
hω r
detector

α e

material
sample

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 figure 17.) Of
hω - Eb -φ

X-ray
E
hω - Eb

Eb

Intensity(E)

D(E)
Kinetic Energy of electrons hω - φ

Figure 17: Let the binding energy be defined so that Eb > 0, φ = work function, then
the detected electron intensity I(Ekin − h̄ω − φ) ∝ D(−Eb )f (−Eb )

course, in order for an electron at an energy of Eb below the


Fermi surface to escape the material, the incident photon must
have an energy which exceeds Eb and the work function φ of
the material. If h̄ω > φ, then the emitted electrons will have
a distribution of kinetic energies Ekin , extending from zero to

26
h̄ω −φ. From Fermi’s 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
1
I(Ekin ) = ∝ D(−Eb)f (−Eb)
τ (Ekin )
∝ D(Ekin + φ − h̄ω)f (Ekin + φ − h̄ω) (55)

Thus if we measure the energy and number of ejected particles,


then we know D(−Eb).
Secondary electrons

phonon Coulombic
interation
e

Figure 18: Left: Origin of the background in I(Ekin . 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 off 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 (Ek). Second, due to these

6
background subtracted
background
’’raw’’ data
4
I(Ekin)

0
0 1 2 3 4
Ekin hω-φ

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
off, the subtracted data is proportional to the electronic density of states convolved
with a Fermi function I(Ekin ) ∝ D(Ekin + φ − h̄ω)f (Ekin + φ − h̄ω).

secondary scattering processes, it is very unlikely that an elec-


tron which is excited deep within the bulk, will ever escape from

28
e
measure
E kin

Figure 20: BIS Ekin = h̄ω − Eb − φ, Eb = h̄ω − Ekin − φ

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 > FF , using Inverse Photoemmision. Here,
an electron beam is focussed on the surface and the outgoing
flus of photons are measured.

29
Chapter 8: Magnetism

Holstein & Primakoff

April 13, 2001

Contents
1 Introduction 2
1.1 The Relevance of Magnetostatics . . . . . . . . . . . . . . . . . . . . 3
1.2 Non-interacting Magnetic Systems . . . . . . . . . . . . . . . . . . . . 4

2 Coulombic Correlation Effects 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 χ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 Finite T Behavior of a Band Ferromagnet . . . . . . . . . . . . . . . 28
3.2.1 Effect 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 effects 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 different sites. As we

Figure 1: Both moment formulation and the correlation between these moments (J)
are due to Coulombic effects

will see, systems without these inter-spin correlations (or those


without well defined 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 effects are due to electronic correlations, those
mediated or due to Coulombic effects, 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.
1
U= [m1 · m2 − 3(m1 · n)(m2 · n)] (1)
r3
then,
m1 m2
Udipole−dipole ≈ (2)
r3
If we let,
eh̄
m1 ∼ m2 ∼ gµB ∼
m

r ≈ 2A
then,  2
(gµB )2 e 
2 Ã
a 0 !3 e 2
U∼ ∼   ∼ 10−4eV (3)
r3 h̄c r a0

4
Or roughly one degree Kelvin! Magnetic correlations due to
magnetic interactions would be destroyed by thermal fluctua-
tions at very low temperatures.

1.2 Non-interacting Magnetic Systems

We will define non-interacting magnetic systems as those for


which the independent moments do not interact with each other,
and only interact with the probing field. For the moment let’s
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
≈ µB (L + 2S) (4)

The system energy will change by an amount


eh̄ eV
∆E ∼ gµB B(L + 2S) ∼ µB B, µB = ∼ 5.8 × 10−5
2m Te
(5)
in an external field. The largest field which can regularly be
produced in a lab is ∼ 10T e (100T e or more can be produced

5
at LANL by blowing things up), thus

∆E∼< 10−3 eV or ∼ kB (10oK) (6)

This is a very small energy. Thus magnetic effects are wiped


out by thermal fluctuation for

kB T > µ 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

µ ¶
−β 12 gµB B β 12 gµB B
1 1 e −e
↑s= hmi = gµB µ ¶ (8)
2 2 e−β 12 gµB B + eβ 12 gµB B
µB µB B
'g tanh(βµB B) ≈ µB (g ∼ 2)
2 T
2
∂ hmi µ
χ = ≈ B (9)
∂B kB T
6
Once again, the energy of the moment-field interaction is roughly
µ ¶
−8 eV 2
µ2B 10 Te
E∼ B2 ∼ µ
−4
2
¶2 B , (k = 1) (10)
T T 10◦KeV
When E ∼ T , thermal fluctuations destroy the orientation of
the moments with the external field. If B ∼ 10T e
100 ◦
E∼ K (11)
T
or E ∼ T at 10◦K! However, we know that systems such as
E

∼ kT/EF

D(E)

Figure 2: In a metal, only the electrons near the Fermi surface, which are not paired
2
kT µB
into singlets, contribute to the bulk susceptibility χ ∼ EF k B T
∼ µ2B D(EF )

iron exist for which a small field 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
µ2B
surface can participate, χ = EF . Note that this is even smaller
kB T
than the free electron result by a factor of EF ¿ 1!

2 Coulombic Correlation Effects

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 reflect the atom’s attempt to
lower its Coulombic energy, see Fig. 3. By maximizing the to-
tal spin S, the spin part of the wavefunction becomes symmetric
1
under electron exchange (i.e. for two electrons with s = 2 ↑ ↓
the maximum value of total spin is S = 1 with a wavefunc-
tion |↑↓i + |↓↑i ). 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 L
Max S
ψ(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 ).
µ B U
ε

Figure 4: A simple tight-binding model with a local Coulomb repulsion U . If U = 0,


the rate that electrons hop on and off any site may be approximated using Fermi’s
1
golden rule ∼ π|B|2 D(EF ) ∼ ∆t
. Then by the uncertainty principal ∆E∆t ∼ h̄ so

each site energy acquires an uncertainty or width ∆E ∼ ∆t
∼ π|B|2 D(EF ) ≡ Γ.
The sites will form moments (see Fig. 5) if Γ À U, |²|

To illustrate how band formation modifies this scenerio, lets


9
consider a simple tight binding model (See Fig. 4). By Fermi’s
goldon rule, each level acquires a width (uncertainty in its en-
ergy) Γ = πB 2D(EF ) and each level can be in one of the four
states shown in Fig. 5. Clearly, the states −−−°
−−−− and −−↑−↓−−

2ε+U

moment forms

µ Γ

a
e2 − r T F
Figure 5: A moment forms on an orbital provided that Γ À U, |²|. U ∼ a
e
rT F 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 −² À Γ and ² + U À Γ.
10
L
I

ξ n
e
m

Figure 6: Here ξ = − 1c ∂φ
∂t
, m = −µB L and φ = Bπa2

In this limit U À B, the system will act more like a system


of free moments than a free electron gas. Thus, one might
expect
χinsulator À χmetal (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 effect 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 χ ∝ a2. In the free electron limit (see
J.M. Ziman, )
Ã !2 
2 1 m
χ = µB D(EF ) 1 − ∗
 (13)
3 m
 2
eV
∆E ∼ 10−8   a10−1 eV −1B 2 ∼ 10−9 eV B 2(T ) (14)
T
m∗
For insulators, often with m < 1 the diamagnetism wins;
m∗
whereas for metals, generally with m ≥ 1, the Pauli para-
magnetism wins.

2.2 Magnetism and Intersite Correlations

From both Hund’s 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 fluctuations at very
low temperatures. Consider once again an isolated moment of
magnitude mµB in an external field.
(mµB )2
χ ≈ (15)
kB T

12
ξ

Figure 7: If the magnetic moments in a small volume are correlated, then the magnetic
susceptibility is strongly enhanced.

(mµB B)2
E ≈ (16)
kB T
For magnetism to be significant at room temperature (300K)
we must increase the energy of our system in a field. This
may be accomplished by increasing the effective moment m by
correlating adjacent moments. If the range of this correlation is
4πξ 3 ξ
ξ, so that roughly 3a3
moments are correlated, then let a =3
so that ∼ 102 moments are correlated and m ∼ 102. This
increases E by about 104 , so that E ∼ kbT at T ∼ 103 K.
The observed (measured) susceptibility also then increases by
about 104, 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 effects and are hence usually short ranged due to elec-
1
tronic (Thomas-Fermi) screening. If we consider two s = 2

spins, ↑1 ↓2, then the correlation is usually parameterized by


the Heisenberg exchange Hamiltonian, or

H = −2Jσ1 · σ2 (17)

where J is the exchange splitting between the singlet and triplet


energies.
 

 






|↑ ↑i 





 

 


|↑ ↑i + |↓ ↑i 

Et (18)

 


 


 


 |↓ ↓i 

{|↑ ↓i − |↓ ↑i} Es (19)

Et − Es = −J (20)

The trick then is to calculate J!

14
1 2
r12
e e

r1B r2A
r1A r
2B

+ +
e R e
AB
A B

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 = H1 + H2 + H12 (21)
h̄2 2 e2 e2
H1 = − ∇ − − (22)
2m r1A r1B
e2 e2
H12 = + (23)
r12 RAB
As we did in Chap. 1 to describe binding, we will use the atomic
wave functions to approximate the molecular wavefunction ψ 12.

ψ12 = (φA(1) + φB (1)) (φA(2) + φB (2)) ⊗ spin part


15
= (φA(1)φA (2) + φB (1)φB (2) + φA(1)φB (2) + φA(2)φB (1))
⊗spin part (24)
2
If re12 is strong (it is) then the first 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

ψ12 ' (φA(1)φB (2) + φB (1)φA (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

ψ12 = (φA(1)φB (2) − φB (1)φA (2)) ⊗ spin triplet (26)

or
ψ12 = φA(1)φB (2) ± φB (1)φA (2) ⊗ spin part (27)
The energy of these states may then be calculated by evaluating
hψ12 |H|ψ12 i
hψ12 |ψ12 i .
hψ12|H|ψ12i C± A
E= = 2EI + , + singlet , − triplet
hψ12|ψ12i 1±S
(28)
16
where
 
Z
2 ∗

 h̄2 2 e2 

EI = d r1φA(1)  − ∇ 1 − φA(1) < 0 (29)
 2m r1A 

the Coulomb integral


 
Z  1 1 1 1  2 2
C = e2 3 3
d r1 d r2  + − −  |φA (1)| |φB (2)| < 0
RAB r12 r2A r1B
(30)
the exchange integral
 
Z  1 1 1 1  ∗ ∗
A = e2 d 3 r1 d 3 r2  + − −  φA (1)φA (2)φB (1)φB (2)
RAB r12 r2A r1B
(31)
and finally, the overlap integral is
Z
S= d3r1d3r2φ∗A(1)φA (2)φB (1)φ∗B (2) (0 < S < 1) (32)

All EI , C, A, S ∈ <. So
 
C −A  C + A
−J = Et − Es = 2EI + − 2EI +
1−S 1+S 
C −A C +A
−J = − >0
1−S 1+S
A − SC
J= 2 <0 (33)
1 − S2
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 effective Hamiltonian.

H = −2J σ1 · σ2, J <0 (34)

Clearly this favors an antiparallel or antiferromagnetic align-


ment of the spins (See Fig. 9) since then (classically) σ1 · σ2 < 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.
X
H = −2J σi · σ j , J <0 (35)
hiji

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
ri
e
V r
j

Figure 10: |↑ ↑i triplet-symmetric

function with plane waves, then


1 ½ iki·ri ikj ·rj iki ·rj ikj ·ri
¾
ψij = √ e e −e e
2V
1 iki·ri ikj ·rj ½ i(k −k )·(r −r )
¾
= √ e e 1−e i j i j (36)
2V
19
The probability that the electrons are in volumes d3ri and d3rj
is
1
|ψij |2d3rid3rj = {1 − cos [(k i − k j ) · (r i − r j )]} d 3
r i d 3
rj
V2
(37)
As required by the Pauli principle, this probability vanishes
when ri = rj . 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 ri ≈ rj for triplet spin electrons, but
not singlet spin ones.
Now consider the effects 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
b
e a
e

+ +
Ze Ze
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 effect over the Fermi sea. If one of the electrons
is fixed at the origin (See Figure 12), then the probability that
a second is located a distance r away, in a volume element d3r
is

P↑↑ (r)d3 r = n↑d3r (1 − cos [(ki − kj ) · r]) (38)


| {z }

Fermi sea average

21
i.e. ri ≡ 0
e
O r=rj
e
Figure 12: Geometry to calculate the exchange interaction.

1 1 # electrons
n↑ = n = (39)
2 2 volume
In terms of an electronic charge density, this is
en
ρex (r) = (1 − cos [(ki − kj ) · r])
2  
en 
 1 Z kF 3 3 1 µ ı(ki−kj )·r −ı(ki −kj )·r
¶

= 1 − d k i d k j e + e
2 
 ( 34 kF3 )2 o 2 

 
en  4 3 −2 Z kF 3 ıki·r Z kF 3 ıkj ·r 
= 1 − ( kF ) 0 d ki e d kj e 
2 

3 0

2
en 
 (sin kF r − kF r cos kF r) 
= 1−9 (40)
2 
 (kF r)6 

Note that both of the exponential terms in the second line are
the same, since we integrate over all ki → −ki & kj → −kj .
Since we have only been considering Pauli-principle effects, the
electronic density of spin down electrons remains unchanged.
Thus, the total charge density around the up spin electron fixed
22
ρeff
en

1/2

2 4 kF r

Figure 13: Electron density near an electron fixed at the origin. Coulomb effects would
reduce the density for small r further, but would not significantly effect the size of the
exchange hole or the range of the corresponding potential, both ∼ 1/kF .

at the origin is
 

 9 (sin kF r − kF r cos kF r)2 

ρef f (r) = en  1 −  (41)
 2 (kF r)6 

The size of the exchange hole, and the range of the correspond-
1
ing ferromagnetic exchange potential, is ∼ kF ∼ a which is
rather short.

3 Band Model of Ferromagnetism

Due to the short range of this potential, its Fourier transform is


essentially flat in k. This fact may be used to construct a band

23
theory of FM where the mean effect 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.
IN↑ < 1eV
E↑(k) = E(k) − ; I∼ (42)
N
Likewise for spin down
IN↓
E↓(k) = E(k) − . (43)
N
Where I, the stoner parameter, quantifies the exchange hole
energy. The relative spin occupation R is related to the bulk
moment
 
(N↑ − N↓) N
R= , M = µB   R (44)
N V
Then
I(N↑ +N↓ )
Eσ (k) = E(k) − 2N − σIR
2 , (σ = ±) (45)
≡ Ẽ(k) − σIR
2 . (46)

If R is finite and real, then we have ferromagnetism.


1 X 1
R = ½ ¾
N k exp (Ẽ(k) − IR/2 − EF )/kB T + 1
1
− ½ ¾ (47)
exp (Ẽ(k) + IR/2 − EF )/kB T + 1
24
For small R, we may expand around Ẽ(k) = EF .
2 3 000
f (x − a) − f (x + a) = −2af 0 − af (48)
3!
All derivatives will be evaluated at Ẽ(k) = EF , so f 0 < 0 and
f f′

E
F

E E E
F

f′′ f′′′

E E

Figure 14:

f 000 > 0. Thus,


¯   ¯
IR 1 X ∂f ¯¯¯ 2  IR 3 1 X ∂ 3f ¯¯¯
R = −2 ¯ − ¯ (49)
2 N k ∂ Ẽ(k) ¯¯EF 6 2 N k ∂ Ẽ 3(k) ¯¯EF
This is a quadratic equation in R
¯ ¯
I X ∂f ¯¯¯ 1 3 21 X ∂ 3f ¯¯¯
−1 − ¯ = I R ¯ (50)
N k ∂E(k) ¯¯EF 24 N k ∂E 3(k) ¯¯EF
25
which has a real solution iff
¯
¯
I X ∂f ¯
¯
−1 − >0 ¯
¯ (51)
N k ∂E(k) EF ¯

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 → −δ(Ẽ − EF )
F

1 X ∂f Z V V
T = 0, − = dẼ D(Ẽ)δ(Ẽ−EF ) = D(EF ) = D̃(EF )
N k ∂Ek 2N 2N
(52)
So, the condition for FM at T = 0 is I D̃(EF ) > 1. This is
known as the Stoner criterion. I is essentially flat as a function
I (eV)
Ni

1.0 D(E F) (eV-1)
Fe
Co

1.0 Na
Li

Z 50 Z

Figure 15:

26
of the atomic number, thus materials such as Fe, Co, & Ni with
a large D̃(EF ) are favored to be FM.

3.1 Enhancement of χ

Even those systems without a FM ground state have their sus-


ceptibility strongly enhanced by this mechanism. Let us re-
consider the effect of an external field (gS = 1) on the band
energies.
Inσ
Eσ (k) = E(k) − − µB σB (53)
N
Then
P ∂f
R = − N1 k ∂ Ẽ (IR + 2µB B)
k

= D̃(EF )(IR + 2µB B) (54)

or as M = µB N
V R, we get

N D̃(EF )
M = 2µ2B B (55)
V 1 − I D̃(EF )
or
∂M χ0
χ = = (56)
∂B 1 − I D̃(EF )

27
χ0 = 2µ2B N
V D̃(EF ) (57)
= µ2B D(EF ) (58)

< 1, the susceptibility can be consider-


Thus, when I D̃(EF ) ∼
ably enhanced over the non-interacting result χ0. However, this
approximation usually overestimates χ since it neglects diamag-
netic contributions, and spin fluctuations (at T 6= 0). As we
will see, the latter especially are important for estimating T c.

Figure 16: Spin fluctuations can reduce the total moment within the correlated region,
and even reduce ξ itself. Both effects lead to a reduction in the bulk susceptibility
χ ∼ moment2

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
not correlated
D(E) s electrons (no moments)

d electrons (with moments)

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
1 X IR IR
R= f (Ẽk − −µB B0 −EF )−f (Ẽk + +µB B0 −EF )
N k 2 2
(59)
1
with f (x) = eβx +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 δ 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 D̃(E) should
reflect only the d-electron contribution, we will accommodate
this by setting

D̃(E) ≈ Cδ(E − EF ), (C < 1) (60)

C, an unknown constant, will be determined by the T = 0


behavior. Then
 
IR
 IR 
R = C f (− − µB B0 ) − f ( + µB B0 ) (61)
2 2
R IC
Let C ≡ R̃ and Tc = 4kB , then if B0 = 0

1 1 R̃Tc
R̃ = Ã ! − Ã ! = tanh (62)
exp −2R̃Tc
+1 exp 2R̃Tc
+1 T
T T

R 1 n↑ −n↓
If T = 0, then R̃ = 1 = C = C N . For Ni, the measured
µef f n↑ −n↓
ground state magnetization per Ni atom is µB = 0.54 = N .
µef f
Therefore, C = 0.54 = µB .

30
For small x, tanh x ' x − 31 x3 , and for large x
sinh x ex − e−x 1 − e−2x
tanh x = = =
cosh x ex + e−x 1 + e−2x
= (1 − e−2x )(1 − e−2x ) ' 1 − 2e−2x (63)

Thus,
2Tc
R̃ = 1 − 2e− T , for T ¿ Tc (64)
√ 1
R̃ = 3(1 − TT ) 2 , < Tc
for small R̃ or T ∼ (65)
C

However, neither of the formulas is verified by experiment. The


Eq (64)


R

Eq(65)

T/Tc

Figure 18:

1
critical exponent β = 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 D̃(E) or values of I will not correct these
problems. Clearly something fundamental is missing from this
31
model (spin waves).

spin flip spin wave


Elementary
Excitations
B0

Included Not Included

Figure 19: Local spin-flip excitations, left, due to thermal fluctuations are properly
treated by mean-field like theories such as the one discussed in Secs. 3 and 4. However,
non-local spin fluctuations due to intersite correlations between the spins are neglected
in mean-field theories. These low-energy excitations can fundamentally change the
nature of the transition.

3.2.1 Effect of B

If there is an external field B0 6= 0, then


 

 R̃Tc + µB B0/2k 
B
R̃ = tanh   (66)
 T 

Or for small R and B0, (or rather, large T À Tc.)


µB Tc µB 1
R̃ = B0 + R̃ ⇒ R̃ = B0 (67)
2kT T 2k T − Tc
µB N CµB N ∂M Cµ2B N
Thus since M = V R = V R̃ ⇒ χ = ∂B0 = 2kV T −Tc .

This form for χ


Const
χ= (68)
T − Tc
32
is called the Curie-Weiss form which is qualitatively satisfied
for T À Tc; however, the values of Const and Tc 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)

δ=2

δ=3 i δ=1

δ=4
P P
Figure 20: Terms in the Heisenberg Hamiltonian H = − iδ Jiδ Si · Siδ − gµB B0 i Si
Here i refers to the sites and δ 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 CeSi2−x , with x > 0.2). As we have seen,

33
systems with localized spins are described by the Heisenberg
Hamiltonian.
X X
H=− Jiδ Si · Siδ − gµB B0 Si (69)
iδ i

In general, this Hamiltonian has no solution, and we must


resort to (further) approximation. In this case, we will approx-
imate the field (exchange plus external magnetic) felt by each
spin as the average field due to the neighbors of that spin and
the external field. (See Fig. 21.) Then

Each site υ nearest neighbors


J with exchange interaction J
Si

Figure 21: The mean or average field felt by a spin Si at site i, due to both its neighbors
1 P P
and the external magnetic field, is gµB
h δ Jiδ Siδ i + B0 = Bief f . Where h δ Jiδ Siδ i
is the internal field, due to the neighbors of site i.

 
X X X 
H≈ − gµB Bief f · Si = − Si ·  Jiδ hSiδ i + gµB B0
i i δ
(70)
If Jiδ = J is a constant (independent of i and δ) describing the
34
exchange between the spin at site i and its ν nearest neighbors,
then
P
J iδ hSiδ i + gµB B0 Jν
Bief f = = hSi + B0 (71)
gµB gµB
N
M = gµB hSi ; ν = #nn . (72)
V
For a homogeneous, ordered system,
V
Bef f = νJM + B0 = BM F + B0 (73)
N g 2µ2B
and
X
H ≈ −gµB Bef f · Si (74)
i

ie., a system of independent spins in a field Bef f . The proba-


bility that a particular spin is up, is then
1
P↑ ∝ e−β (−gµB Bef f 2 ) (75)

and
1
P↓ ∝ e−β (+gµB Bef f 2 ) (76)

so, on average
N↓
= e−βgµB Bef f (77)
N↑

35
and, since N↑ + N↓ = N
 
1 N↑ − N ↓ 1 N β
M = gµB = gµB tanh  gµB Bef f  (78)
2 V 2 V 2
Since tanh is odd and Bef f ∝ M , this will only have non-
trivial solutions if J > 0 (if B0 = 0). If we identify
N gµB 1 J
Ms = ; Tc = ν (79)
V 2 4 k
 
M Tc M 
= tanh  (80)
Ms T Ms
< Tc
and again for T = 0, M (T = 0) = Ms, and for T ∼

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 6= 0) solutions for all b > 1.

1
M √  T 2
' 3 1−
  (81)
Ms Tc
36
Again, we get the same (wrong) exponent β = 12 .
When is this approximation good? When each spin really
feels an ”average” field. Suppose we have an ordered solid, so
that

BM F == Breal (82)
2gµB
Now, consider one spin flip excitation adjacent to site i only,

Figure 23: The flip of a single spin adjacent to site i makes a significant change in
the effective exchange field, felt by spin Si , if the site has few nearest neighbors.

Fig. 23.
If there are an infinite # of spins then BM F remains un-
changed but for ν < ∞
Jν ν − 2 Jν
Breal = 6= BM F = . (83)
2gµB ν 2gµB
Clearly, for this approximation to remain valid, we need B real =
37
ν −2
BM F , which will only happen if ν = 1 or ν À 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 flip, 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

J<0

"down" sublattice

"up" sublattice

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 field
from the ν = 8 near-neighbor spins on the “down” sublattice.
so

− 2 2 −
⁄ N g µB )νJM
B = V/(
MF
Figure 25:
 
+
1 N gµB

 V 

M+ = gµB tanh  − 2 2 νJM − (84)
2 V  2kT N g µB 

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
 
+ + 
1 N 
 V νJM 
M + = gµB tanh  −
 2kT N + gµ 
, J < 0 (86)
2 V B

− +
M = −M (87)
39
Again, these equations will saturate at
1 N+
Ms+ = −Ms− = gµB (88)
2 V
so  
M+  TN M + 
 
= tanh  T M 
 (89)
Ms+ s

where TN = − 14 kνJ
B

Now consider the effect of a small external field B0. This


will yield a small increase or decrease in each sublattice’s mag-
netization ∆M ±.
  
1 N+  gµB
 V νJ ³ − − ´


M + + ∆M + = gµB tanh  
B 0 + M + ∆M 
2 V  2kT N −g 2µ2B 

M − + ∆M − = (+ ↔ −) . . . (90)
( )
d 1 ∂M
Or, since dx tanh x = cosh2 x
, then ∆M = ∂Bef f ∆Bef f .
 
+ − 1 N + gµB 1  V νJ 
∆M = ∆M +∆M = gµB 2 B 0 + ∆M 
2 V 2kT cosh x 2N −g 2µ2B
(91)
TN M +
where x = T Ms+ . For T > TN , M + = 0 and so x = 0, and
 
g 2µ2B N  4kB TN V
∆M = 2 B 0 −
2 2 ∆M  (92)
8V kB T N g µB
g 2µ2B N
T ∆M = B0 − ∆M TN (93)
4V kB
40
g 2µ2B N
∆M = B0 (94)
4V kB (T + TN )
g 2µ2B N
χ = (95)
4V kB (T + TN )

T
0 T
−TN N

Figure 26: Sketch of χ = Const/(T + TN ). Unlike the ferromagnetic case, the bulk
susceptibility χ does not diverge at the transition. However, as we will see, this
equation only applies for the paramagnetic state (T > TN ), and even here, there are
important corrections.

Below the transition, T < TN , the susceptibility displays dif-


ferent behaviors depending upon the orientation of the applied
field. For T ¿ TN and a small B0 parallel to the axis of the
sublattice magnetization, we can approximate M +(T ) ≈ Ms+
TN
and x ≈ T in Eq. 91
g 2µ2B N 1
χ' µ ¶ (96)
4V kB T cosh2 TTN + TN

41
B0

Figure 27: When T ¿ TN , a weak field applied parallel to the sublattice magnetization
TN
axis only weakly perturbs the spins. Here M + (T ) ≈ Ms+ and x ≈ T

g 2µ2B N −2 TN
χ' e T (97)
4V kB
Now consider the case where B0 is perpendicular to the mag-
netic axis. The external field will cause each spin to rotate a

B
0

B0
BMF

Figure 28: When T ¿ TN , a weak field B0 applied perpendicular to the sublattice


magnetization, can still cause a rotation of each spin by an angle proportional to
B0 /BM F .

small angle α. (See Fig. 28) The energy of each spin in this

42
external field and the mean field ∝ ν gµJ to the first order in
B

B0 is
1 1
E = − gµB B0 sin ∝ + νJ cos α (98)
2 2
Equilibrium is obtained when ∂E
∂α = 0. Since B0 is taken as

small, α ¿ 1.
 
1 1 1
E ∼ − gµB B0α + νJ 1 − α2 (99)
2 2 2
or
∂E 1 1 gµB B0
= 0 = gµB B0 + να ⇒ α = − (100)
∂α 2 2 νJ
The induced magnetization is then
1 gµB B0 g 2µ2B N B0
∆M = α =− (101)
2 V 2νJV
so
g 2µ2B N
χ⊥ = = constant (102)
2ν |J| V
Of course, in general, in a powdered sample, the susceptibility
will reflect an average of the two forms, see for example Fig. 30.

43
χ
χ powdered sample
χ⊥
χ
χ


T T T
N

Figure 29: Below the Neel transition, the lattice responds very differently to a field
applied parallel or perpendicular to the sublattice magnetization. However, in a pow-
dered sample, or for a field 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-field approaches to


magnetism in terms of their inability to account for low-energy
processes, such as the flipping of spins. (S α → −S α ) However,
we have yet to discuss the lowest energy spin flip 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
1.2e-5

1.1e-5

1.0e-5

9.0e-6

8.0e-6

7.0e-6

6.0e-6

5.0e-6

4.0e-6

3.0e-6

2.0e-6

1.0e-6

0.0e+0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95

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
X
H = −J Si · Si+δ (103)

 
1
 
where Si = x̂Six + ŷSiy + ẑSiz . If we define |αi = 


 (i.e.
0

45
 
0
 
|↑i), β = 


 (i.e. |↓i) so that
1
   
0
z  1 0
S 


 =− 


 ··· (104)
1 2 1
and
     
1  0 −i  1 0 1 1 1 0 
Sx = 


 Sy = 


 Sz = 



2 i 0 2 1 0 2 0 −1
(105)
· ¸
α β
S ,S = i²αβγ S γ (106)

It is often convenient to introduce spin lowering and raising


operators S − and S +.
 

0 1 h i
S + = S x + iS y = 


 S z , S ± = ±S +/−(107)
0 0
 

− x y 
0 0 
·
2 ±
¸
S = S − iS = 


 S ,S =0 (108)
1 0
     

0 
1 
0
S+ 


 =


 , S− 


 = 0... (109)
1 0 1
They allow is to rewrite H as
X 1³ + − + ´
H = −J z
Siz Si+δ + Si Si+δ + S−Si+δ (110)
iδ 2

46
Since J > 0, the ground state is composed of all spins oriented,
for example
|0i = Πi |αii (i.e. all up) (111)

This is an eigenstate of H, since


Si+Si+δ |0i = 0 (112)

and
1
Siz Si+δ
z
|0i = |0i (113)
4
so
1
H |0i = − JνN |0i ≡ E0 |0i (114)
4
where N is the number of spins each with ν nearest neighbors.
Now consider a spin-flip excitation. (See Fig. 31)

j
Figure 31: A single local spin-flip excitation of a ferromagnetic system. The resulting
state is not an eigenstate of the Heisenberg Hamiltonian.

|↓j i = Sj−Πn |αin (115)

47

This is not an eigenstate since the Hamiltonian operator Sj+Sj+δ
will move the flipped spin to an adjacent site, and hence create
another state.
However, if we delocalize this spin-flip 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-flip 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 flip.

state
1 X
|ki = √ eik·rj |↓j i . (116)
N j

It is an eigenstate. Consider:

1 X ik·rj  1
H |ki = √ e − νJ(N − 2) |↓j i
N j 4

1 1 X 
+ νJ |↓j i − J (|↓j+δ i + |↓j−δ i) (117)
2 2 δ
where the sum in the last two terms on the right is over the near-

48
neighbors δ to site j. The last two terms may be rewritten:
1 X ik·rj √ X ik·(r −r )
√ e |↓j+δ i = 1 N e m δ |↓mi (118)
N j m

where
rj+δ = rj + rδ = rm (119)

so
 
1 1 X µ ik·rδ −ik·rδ
¶ 1 X
H |ki = − νJN + νJ − J e +e 
√ eik·rj |↓j i
4 2 δ N j
(120)
Thus |ki is an eigenstate with an eigenvalue
 
1X
 
E = E0 + Jν 1 − cos k · rδ  (121)
ν δ
Apparently the energy of the excitation described by |ki van-
ishes as k → 0.
What is |ki? First, consider
X X 1 X
S z |ki = Siz |ki = Siz √ eik· rj
|↓j i
i i N j
1 X ik·rj X z
=√ e Si |↓j i = (SN − 1) |ki (122)
N j i

I.e., it is an excitation of the ground state with one spin flipped.


Apparently, since Ek=0 = E0 , the energy to flip 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 flip a spin,
using a spin-wave excitation, then

S z |ki = (SN − 1) |ki , S z |0i = SN |0i (123)

If we add another spin wave, then hS z i = (SN −2). For spin 21 ,


N
hS z i = 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
nk excitations in each mode k, the corresponding excited spin
state can be described by a Boson state vector |n1n2 . . . nN i
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

Siz = S − a†i ai
Si+ ∝ ai
Si− ∝ a†i (124)
50
Sz n
3
2 0
1
2 1
− 12 2
− 32 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


· ¸
ai, a†i =1 (125)


ai |ni = ni |n − 1i (126)

a†i |ni = ni + 1 |n + 1i (127)

This transformation is faithful (canonical) and will maintain the



dynamical properties of the system (given by ∂t θ = ih̄ [H, θ]) if
it preserves the commutator algebra
h i h i h i
Si+, Si− = 2Siz , Si−, Siz = 2Si−, Si+, Siz = −2Si+ (128)

consider
n
+ − − +o
S S −S S |ni = 2S z |ni = 2(S − n) |ni (129)
51
If S + = a, and S − = a†, then the left-hand side of the above
equation would be {(n + 1) − n} |ni = |ni 6= 2(S − n) |ni.
In order to maintain the commutators, we need
+
√ − †

S = 2S − n a S = a 2S − n (130)

Then

[S +, S −] |ni = S +S − |ni − S −S + |ni


q q
= 2S − a†aaa† 2S − a†a |ni − a†(2S − a†a)a |ni
= (2S − n)(n + 1) |ni − n(2S − (n − 1)) |ni
= (2Sn + 2S − n2 − n − 2Sn + n2 − n) |ni
= 2(S − n) |ni (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
r r
Si+ = 2S − a†i aiai Si− = a†i 2S − a†i ai Siz = S − a†i ai
(133)
52
is called the Holstein Primakoff transformation.
If we Fourier transform these operators,
1 X 1 X
a†i = √ eik·Ri a†k ; ai = √ e−ik·Ri ak (134)
N k N k

then (since the Fourier transform is unitary) these new opera-


tors satisfy the same commutation relations
· ¸ · ¸
ak , a†k0 = δkk0 a†k , a†k0 = [ak , ak0 ] = 0 (135)

To convert the Hamiltonian into this form, assume the number


of magnons in each mode is small and expand
+
√ √ ni
Si = 2S − niai ' 2S(1 − )ai (136)
 4S 
 1

X ik·R 1 X i(p+q−k)·R †


≈  √ e i a k − 3 e i a a a
k p q

N k 4SN 2 kpq 

Of course this is only exact for ni ¿ 2S, i.e. for low T where
there are few spin excitations, and large S (the classical spin
limit).
In this limit
v
u
2S X ik·Ri
u
u
Si+ ' t e ak (137)
v
N k
u
u 2S X
u
Si− ' t e−ik·Ri a†k (138)
N k
53
1 X 0
Siz = S − ei(k−k )·Ri a†k ak0 , (139)
N kk 0

the Hamiltonian
 
X z z 1 + − − +

H = −J S S
 i i+δ + (S S + S S ) (140)
iδ 2 i i+δ i i+δ 

may be approximated as
X
H ' −N JνS 2 + 2JνS a†k ak
 
k
X 1 X ik· Rδ  †
−2JνS  e ak ak + O(a4k ) (141)
k ν δ
X
H ' E0 + 2JνS(1 − γk )a†k ak + O(a4k ) (142)
k
1 P ikRδ
where γk = ν δe . This is the Hamiltonian of a collection of
k-q
k′+q k k′

k k′

Figure 33: The fourth order correction to Eq. 142 corresponds to interactions between
the spin waves, giving them a finite lifetime

harmonic oscillators plus some other term of order O(a4k ) which

54
corresponds to interactions between the spin waves. These in-
teractions are a result of our definition of a spin-wave as an
itinerant spin flip 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 specific heat and the mag-
netization. Consider
X h̄ωk
hEi = . (143)
k eβh̄ωk − 1
For small k, h̄ωk = 2JνS(1 − γk ) = 2JνSk 2 on a cubic lattice.
Then let’s assume that the k-space is isotropic, so that d3k ∼
k 2dk, then
2 k2
γk = (cos kx + cos ky + · · ·) = 1 − (144)
ν ν
and
X 2JνSk 2 Z ∞ k 4dk
hEi ≈ ∝ (145)
k eβ2JνSk2 − 1 0 eβαk2 − 1
55
 1  1
x 2 1 1 2 1
x = βαk 2 k=  dk =   x− 2 dx (146)
βα 2 βα
so that
−2 −1/2 x3/2Z ∞
hEi ∝ β β 0
dx x ∼ T 5/2 (147)
e −1
Thus, the specific heat at constant volume CV ∼ T 3/2, which
is in agreement with experiment.

M/M(0)
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 fluctuations, derived from our mean-field
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
* +
X gµB
M (0) − M (T ) = nk (148)
k V

56
since each magnon corresponds to spin flip. Thus
k 2dk Z 3
M (T ) − M (0) ∼ − βαk2 ∼ T2 (149)
e −1
which clearly dominates the exponential form found in MFT
2Tc
(1 − 2e 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-field state in
which there are both a spin up and down sublattices. Spin

down sublattice

up 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


Siz = S − ni Siz = −S + ni (150)
√ ³
− ´+

− +
+
Si = (Si ) = 2Sfi(S)ai Si+ = Si = 2Sa†i fi(S)

where v
u
u ni
fi(S) = 1 − t and ni = a†i ai (151)
2S
Again this transformation is exact (canonical) within the man-
ifold of allowed states

0 ≤ ni ≤ 2S ⇔ −S ≤ Sz ≤ S . (152)

The Hamiltonian
X 1³ + − + ´
H = −J z
Siz Si+δ = Si Si+δ + Si−Si+δ (153)
iδ 2
may be rewritten in terms of Boson operators as
X
H = +JS 2N ν + J a†i aia†i+δ ai+δ

X½ †
− JS ai ai + a†i+δ ai+δ
iδ ¾
+ fi(S)aifi+δ (S)ai+δ + a†i fi(S)a†i+δ fi(S) (154)

58
Once again, we will expand
v
u
u ni ni n2i
fi(S) = 1 −
t =1− − − ··· (155)
2S 4S 32S 2
and include terms in H only to O(a2)
X½ † ¾
2
H ' JS N ν − JS ai ai + a†i+δ ai+δ + aiai+δ + a†i a†i+δ

(156)
This Hamiltonian may be diagonalized using a Fourier trans-
form
1 X −ik·Ri
ai = √ e ak (157)
N k
and the Bogoliubov transform

ak = αk cosh uk − α−k sinh uk (158)
a†k = αk† cosh uk − α−k sinh uk (159)

tanh 2uk = −γk (160)


1 X ik· Rδ
γk = e (161)
ν δ
· ¸

Here the αk are also Boson operators αk , αk = 1. To see if this
transform is canonical, we must ensure that the commutators
are preserved.

1 = [ak , a†k0 ] = [αk Ck − α−k Sk , αk† 0 Ck0 − α−k0 Sk0 ]
59
½ ¾
= Ck2[αk , αk† ] + †
Sk2[α−k , α−k ] δkk0 (162)
½ ¾
= Ck2 − Sk2 δkk0 = δkk0

where Ck (Sk ) is shorthand for cosh uk (sinh uk ). You should


check that the other relations, [ak , ak0 ] = [a†k , a†k0 ] = 0, are
preserved.
After this transformation,
 
X 1
H ≈ JN νS(S + 1) + h̄ωk αk† αk +  + O(a4) (163)
k 2
r
where h̄ωk = −2JSν 1 − γk2. Notice that for small k, h̄ωk ∼
√ √
2JSνk ≡ Ck (C = − 2JSν is the spin-wave velocity).
The ground state energy of this system (no magnons), is
r
X
E0 = JN νS(S + 1) − JSν 1 − γk2 (164)
k
If γk = 0, then each spin decouples from the fluctuations of its
neighbors and E0 = JN νS 2 (J < 0) which is the energy of
the Neel state. However, since γk 6= 0, the ground state energy
E0 < EN . 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 ∝ hnii in its spin
alignment.
60
2
E N = JNνS

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 finite.

The linear dispersion of the antiferromagnet means that its


bulk thermodynamic properties will emulate those of a phonon
lattice. For example
Xh̄ωk X αk
hEi ' βh̄ωk − 1
'
k e k eβαk − 1
Z ∞ αk 3 dk
∼ 0 βαk
e −1
3
4 ∞ x dx
Z
hEi ∼ T 0 x (165)
e −1
∂ hEi
C= ∼ T3 like phonons! (166)
∂T
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 dθ ∝ S(k, ω) ∝ I {F(-i〈[a(t),a†(0)]〉)}
dΩdω ∼ n
n
Ei θ
sample
2θ ∝ k = k i − k j
thermal
spin- h ω = E i − Ef
polarized
neutrons

Figure 37: Polarized neutrons are used for two reasons. First if we look at only spin
flip events, then we can discriminate between phonon and magnon contributions to
S(k, ω). Second the dispersion may be anisotropic, so excitations with orthogonal
polarizations may disperse differently.

perform on an antiferromagnet is inelastic neutron scattering. If


spin-polarized neutrons are scattered from a sample, then only
those with flipped spins have created a magnon. If the neutron
creates a phonon, then its spin remains unchanged. The time
of flight of the neutron allows us to determine the energy loss
or gain of the neutron. Thus, if we plot the differential cross
section of neutrons with flipped spins, we learn about magnon
dispersion and lifetime. Notice that the peak in S(k, ω) has a
width. This is not just due to the instrumental resolution of the
experiment; rather it also reflects the fact that magnons have a

62
S(k,ω)

n↓ γ 〈ω k 〉
magnon k

n↑
ω ω
phonon junk k 2θ
subtracted off

Figure 38: Sketch of neutron structure factor from scattering off of a magnetic system.
The spin-wave peak is centered on the magnon dispersion. It has a width due to the
finite lifetime of magnon excitations.

finite life time δt,which broadens their neutron signature by γ k .


1
γk δt ∼ h̄ δt ∼ (167)
γk
However in the quadratic spin wave approximation the lifetime
of the modes h̄ωk is infinite. It is the neglected terms in H, of
order O(a4) and higher which give the magnons a finite lifetime.


a i a i a†i+δ a ⇒
i+δ

Figure 39:

63
Chapter 9: Electronic Transport

Onsager

April 23, 2001

Contents
1 Quasiparticle Propagation 2
1.1 Quasiparticle Equation of Motion and Effective 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 Effects 30
6.1 Linearized Boltzmann Equation . . . . . . . . . . . . . . . . . . . . . 31

1
6.2 Electric Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6.3 Thermal and Energy Currents . . . . . . . . . . . . . . . . . . . . . . 34
6.4 Seebeck Effect, Thermocouples . . . . . . . . . . . . . . . . . . . . . 39
6.5 Peltier Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 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
effects.

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 ω(k). The dispersion is ob-
tained from band structure E(k) = h̄ω(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
V
e

V(r)

Figure 1: The dispersion is obtained from band structure E(k) = h̄ω(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

φk (r) = Uk (r)eik·r , Uk (r) = Uk (r + rn) (1)

may be approximated as plane waves Uk (r) = Uk . Then, the


state describing a single quasiparticle may be expanded.
1 Z∞ ·x −ω(k)t)
ψ(x, t) = √ −∞
dkU (k)ei(k (2)

If U (k) = cδ(k − k0) then ψ(x, t) ∝ ei(k0·x−ωt) and the quasi-
particle is delocalized. On the other hand, if U (k) = constant
then ψ(x, t) ∝ δ(x) and the quasiparticle is perfectly localized.
This is an expression of the uncertainty principle

∆k ∆x ∼ 1 or ∆p ∆x ∼ h̄ , (3)
4
Re ψ(x)

Figure 2: If U (k) = cδ(k − k0 ) then ψ(x, t) ∝ ei(k0 ·x−ωt) and the quasiparticle is
delocalized.

so that we cannot know both the momentum and location of


the quasiparticle to arbitrary precision.
ω(k) 6= constant, so the different components propagate with
different 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 ψ(x, t) which when t = 0.
1 Z∞ ik·x
ψ(x, 0) = √ dkU (k)e . (4)
2π −∞
Suppose that U (k) has a well-defined dominant peak (See Fig.
3) so that

ω(k) ' ω(k0 ) + ∇kω(k)|k0 · (k − k0)t (5)

then
1 Z∞ i(k·x−ω0 t− ∇k ω(k)|k (k−k0 )t)
ψ(x, t) ' √ dkU (k)e 0 (6)
2π −∞
5
U(k)

k0 k

Figure 3: The distribution of plane wave state that make up a quasiparticle.

ei(k0· ∇kω(k)|k0 −ω0)t Z ∞ ik·(x− ∇k ω(k)|k t)


ψ(x, t) ' √ dke 0 U (k)
2π −∞
µ ¶
' ψ x − ∇kω(k)|k0 t, 0 ei(k0 ∇kω(k)|k0 −ω0)t (7)

Ie., aside from a phase factor, the quasiparticle travels along


with velocity ∇kω(k)|k0 = vg . (If we had considered higher
order terms, we would have seen the quasiparticle distorts as it
propagates. (c.f. Jackson p.305). In general,
1
vg = ∇k ω(k) = ∇k E(k) (8)

1.1 Quasiparticle Equation of Motion and Effective Mass

We are ultimately interested in the transport; i.e. the response


of this quasiparticle to an external electric field E , from which

6
it gains energy.

δE = −eE · vδt (i.e. force × distance) (9)

This energy is reflected by the quasiparticle ascending to higher


energy k states.

δE = ∇k E(k) · δk = h̄v · δk (10)

So
h̄δk = −eEδt (11)
h̄k̇ = −eE 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 E is smaller than the atomic
fields, and it must not vary in space or time too fast.
We may put this EOM in a more familiar form
1d 1 X ∂ 2E dkj
v̇i = (∇k E(k))i =
h̄ dt h̄ j ∂ki∂kj dt
1 X ∂ 2E
= (−eEj ) (13)
h̄ j ∂ki∂kj
This will have the form F = ma, if we define the mass tensor
   
1  = 
1  =
1 ∂E
 symmetric & real (14)
m∗ ij m∗ ji h̄2 ∂ki∂kj
7
which may be diagonalized to define three principle axes. In
the simple cubic case, the matrix will have the same element
along each principle direction and

∗ h̄2
m = d2 E
. (15)
dk 2

In this way the effective mass of electrons on a lattice can vary


2 ∗
strongly, the larger ddkE2 is, the smaller mm is. Consider the simple
1-d case (See Fig. 4)
E(k)

2
d E <0
dk2

2
d E =0
dk2
k
2
-π/a d E >0 π/a
dk2
m*

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 influence of an applied field E,
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 d3k states

3 
L
dJ = −ev(k)   d3k (16)

The current density is then
−e 1 3
dj = ∇k E(k) dk (17)
h̄ (2π)3
ie., different occupied states in the Brillouin zone contribute
differently 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 first Brillouin zone
e Z
j = − 3 1st B.Z. ∇k E(k)d3 k . (18)
8π h̄
Thus for each k vector in the integral, there is also −k.
9
First B.Z.

Fermi surface

Figure 5: Different occupied states make different 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
1
v−k = ∇−kE(−k) = −∇kE(k) = −vk! (20)

i.e., for the insulator
−e Z
j = 3 1st B.Z. d3k∇kE(k) ≡ 0 (21)
8π h̄
Now imagine that the band is not full (See Fig. 6, left). Then,
if we apply an external field E, so that
eE
k̇ = − e > 0! (22)

10
ky ky

E  x

kx kx

empty occupied states empty occupied states

Figure 6: The Fermi sea of a partially filled band will shift under the influence of an
applied field 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).
−e Z 3
j = v(k)d k
8π koccupied
3
−e Z 3 −e Z 3
= d kv(k) − d kv(k)
8π 3 Z1st B.Z. 8π 3 empty
e 3
= d kv(k) (23)
8π empty
3

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
Ek E
E
E
holes

k D(E)
full states

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.

d2 E
band dk 2
< 0, so the mass is negative and the dispersion at the
top of the band is always also parabolic, so
h̄2k 2
E(k) = E0 − ¯¯ ∗ ¯¯ k = deviation from top! (24)
2 ¯¯mˆ¯¯
or
1d h̄k̇ eE
v̇ =∇k Ek = − ¯¯ ∗ ¯¯ = ¯¯ ∗ ¯¯ (25)
h̄ dt ¯m ¯
¯
¯m ¯
ˆ¯ ¯ ˆ¯
This is the EOM of a positively charged particle with positive
mass in an electric field E. 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

E E
T=0 T≠0

empty Conduction band


electrons
Ef Eg
Valence band holes
full

D(E) D(E)

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 E. However we ignored the presence


of other bands. If there is a conductiong band, for T 6= 0,
and a reasonably small Eg , there will be conductivity due to
a small number of thermally excited holes and electrons n ∼
exp(−Eg /KB T ) (See Fig. 8 right). Thus perhaps a better
definition 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


e
v̇ = ¯ ¯E (26)
¯ ∗ ¯
¯
¯ mˆ¯¯
as long as E is finite, 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 finite R). In
addition, if E 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 off the lattice yielding resistivity. Bloch
showed this to be wrong.

particle state may be defined 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)

(a) (b)
e

e

E(k) - E(k+q) = hω(q)

† k+q k+q
c † c a -q or c † c a q
k+q k k+q k -q
q

c †
c (a q + a†-q )
k+q k k k

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 effect 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

E1 + 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-


k3 k4
3 4
E E4
3

E E2
1
1 2
k1 k2

Figure 11: E1 + E2 = E3 + E4 and k1 + k2 = k3 + k4 . Electron-electron interactions


also contribute to the resistivity (from simple order of magnitude arguments based
on relative strengths of the interactions, their contribution should dominate–but due
to the Pauli principle, it does not).

cal lattice vector G, k1 + k2 = k3 + k4 = G; however, as with


phonon conductivity, these processes with finite G involve much
16
higher energies, and may be neglected near T = 0.) Further-
more, since all states up to EF are occupied, E3; E4 > EF !
Suppose E1 is (thermally) excited, so E1 > EF and it collides
with an occupied state E2 < EF . Then

(E1 − EF ) + (E2 − EF ) = (E3 − EF ) + (E4 − EF ) > 0 (28)

²1 + ²2 = ²3 + ²4 > 0, ² 3 ; ²4 > 0 (29)

or ²1 + ²2 > 0, However, since ²2 < 0, if ²1 is small, then


²2 ²1
|²2| ≤ ²1, is also small, so only states with EF ≤ EF states
may scatter with the state k1 conserve energy and obey the
Pauli principle, thus restricting ²2 to a narrow shell of width ²1
around the Fermi surface.
Now consider the restrictions placed on the states 3 and 4 by
momentum conservation.

k1 − k 3 = k 4 − k 2 (30)

I.e. k1 − k3 and k4 − k2 must remain parallel, and since k1


is fixed, this restriction on the final states further reduces the
²1
scattering probability by a factor of EF .

17
E
k1 ky
k4
k3
EF 3
k2 k 1- k3
2 1
k 4- k2
4 k
x

D(E)

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 σ is reduced from the


µ ¶
²1 2
classical result σ0 by EF . If the initial excitation ²1 is due to
thermal effects, then ²1 ∼ kB T and
 2 
σ k B T
∼  ¿ 1! (31)
σ0 EF
The total scattering due to electron - electron repulsion is very
small. Therefore, unless EF can be made small, the dominant
contribution to a material’s 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 field is described by
the Boltzmann equation.
L

V
e j defect

Figure 13: Electronic transport due to an applied field E, is limited by inelastic colli-
sions with lattice defects and phonons.

This differs from the situation of a system in equilibrium in


that a constant deterministic current differs from random par-
ticle number fluctuations due to coupling to a heat and particle
bath. Away from equilibrium (E 6= 0) the distribution function
may depend upon r and t as well as k (or E(k)). Nevertheless,
when E = 0 we expect the distribution function of the particles

19
in V to return to
1
f0(k) = f (r, k, t)|E=0 = (32)
eβ(E(k)−EF ) + 1
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 field changes significantly. In this way the system
is considered essentially homogeneous with any inhomogeneity
driven by the external field. Now imagine that there is no scat-
tering (no defects, phonons), then since electrons are conserved

dr
r- dt
dt r
·
k - dk dt k hk = -eE
dt
t - dt t

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


 
eE
f (r, k, t) = f r − vdt, k + 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 t−dt,
µ ¶
eE
to another at r and time t, so that f (r, k, t) 6= f r − vdt, k + h̄ dt, tdt .
20
We will express this scattering by adding a term.

Figure 15: Scattering leads to quasiparticle decay.

   
dt ∂f
f (r, k, t) = f r − v dt, k + eE , t − dt +   dt
h̄ ∂t S
(34)
For small dt we may expand
 
f ∂f ∂f
f (r, k, t) = f (r, k, t)−v · ∇r f +eE · ∇k − +  (35)
h̄ ∂t ∂t S
or
 
∂f e ∂f
+ v · ∇ r f − E · ∇k f =   Boltzmann Equation
∂t h̄ ∂t S
(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 k0 (occupied to unoccupied) is wk0k = 2π
h̄ |hk0 |H| ki|2.
Then
   
3Z
 ∂f (k)  L
  =   d3 k 0 {(1 − f (k)) wkk0 f (k0)
∂t S 2π
− (1 − f (k0)) wk0k f (k)} (37)
21
Needless to say it is extremely difficult to solve these last two
coupled equations.

4.1 Relaxation Time Approximation

As a result we make a series of approximations and ansatz. The


first of these is the relaxation time approximation that the rate
at which a system returns to equilibrium f0 is proportional to
its deviation from equilibrium
 
∂f  f (k) − f0(k)
 =− . (38)
∂t S τ (k)
Here τ (k) is called the relaxation time (for a spatially inhomo-
geneous system τ will also depend upon r). Ie., we make the
assumption that scattering merely acts to drive a nonequilib-
rium system back to equilibrium.
If E 6= 0 for t < 0 and then at t = 0 it is switched off so
that for t > 0 E = 0, then for a homogeneous system
 
∂f  ∂f  f − f0
= =− (39)
∂t ∂t S τ
so that
t
f − f0 = (f (t = 0) − f0) e− τ (40)
22
ie., τ 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 field E = E x̂. Then
∂f
=0 (41)
∂t
Furthermore since the system is homogeneous

∇r f = 0 (42)

then
 
e ∂f f (k) − f0(k)
− E · ∇k f (k) =   = − (43)
h̄ ∂t S τ (k)
ie
e
f (k) = f0(k) + τ (k)E · ∇k f (k) (44)

which may be solved iteratively, generating a power series in E
(or Ex).

4.2 Linear Boltzmann Equation

For small E (Ohmic conditions)


e
f (k) ' f0(k) + τ (k)E · ∇k f0(k) linear Boltzmann Eqn.

(45)
23
I.e. the lowest order Taylor series of f (k). Or equivalently, if
E = Exx̂
à !
e
f (k) ' f0 k + τ (k)E (46)

Ie., the effect is to shift the Fermi surface from its equilibrium
position by an amount

k
1ST BZ y

δk x
k
x

Figure 16: According to the linear Boltzmann equation, the effect of a field E x is to
shift the Fermi surface by δkx = −eτ Eh̄x

Ex
δkx = −eτ (47)

From the discussion in Sec.??, it is clear that a finite current
results.
Interesting! Note that elastic scattering |k| = |k0| 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.

Figure 17: Note that elastic scattering |k| = |k0 | 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/τ .

Under these assumptions, the EOM is


mµ ¶
mv̇ + v − vtherm = −eE (48)
τ
25
m
where v − vtherm = vD , the drift velocity, and τ vD is friction.
Again when E = 0, we again have an exponential decay of v
so τ is again the relaxation time.
In steady-state v̇ = 0
−eτ
vD = E (49)
m
so that
ne2τ
j = −envD = E (50)
m
or defining j = σE, and σ = µne,
ne2τ eτ
σ = µ = (51)
m m

5.2 Conductivity Using the Linear Boltzmann Equation

Of course, this is wrong since all free electrons do not par-


ticipate in σ 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
−e Z 3
j = d k v(k)f (k) (52)
8π 3  
−e Z
3

 eτ (k) ∂f0  
' d k v(k)  f0(k) + Ex (53)
8π 3  h̄ ∂kx 

26
V

E = E x̂

Figure 18: To calculate the conductivity, we apply a field in the x-direction only and
use the linearized Boltzmann Eqn.

For an isotopic material jz = jy = 0, and the equation becomes


scalar. Furthermore, again
Z
v(k)f0(k)d3k = 0 (54)

since v−k = −vk . Then as


∂f0 ∂f0 ∂E ∂f0
= = h̄vx (55)
∂kx ∂E ∂kx ∂E
e2 Z ∂f0
jx ' − 3 Ex d3kvx2 τ (k) (56)
8π ∂E
∂f0
Then as ∂E ' −δ(E − EF ) for T ¿ EF the integral in k is
confined to the surface of constant E, and
dE
d3k = dSE dk⊥ = dSE (57)
h̄v(k)
then
27
kz

k⊥ ∇k E

dk⊥ = dE
∇k E 

ky

constant energy
surface
kx

Figure 19:

jx e2 Z vx2 (k)
σ = = dSE dE τ (k)δ(E − EF ) (58)
Ex 8π 3h̄ v(k)
e2 Z vx2 (k)
= dSE τ (k) . (59)
8π 3h̄ E=EF v(k)
As expected only the properties of the electrons on the Fermi
surface are relevant.
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
vx2
aging v τ (k) over the Fermi surface. Consider a simple system
with a spherical Fermi surface, then
Z τ (k)vx2 (k) 4π 2
dSE = kF τ (EF )v(EF )
v(k) 3
4π h̄kF
= kF2 τ (EF ) ∗ (60)
3 m
or
e2 4π 2 h̄kF
σ = 3 kF τ (EF ) ∗ (61)
8π h̄ 3 m
3
kF
then as kB T ¿ EF , N = 2 43 π 2π 3
⇒ kF3 = 3π 2n we find
( )
L
that
e2τ (EF ) eτ (EF )
σ = n µ = (62)
m∗ m∗
For semiconductors where n is T dependent, and for more re-
alistic material where the Fermi surface 6= sphere, the formula
is more complicated.
However, for metals the temperature dependence of σ is dom-
inated by that of τ ; ie., by the temperature dependence of
phonons. However, before we can calculate σ(T ), we must first
disentangle the phonon from the defect scattering. Assuming

29
that the two mechanisms are independent, they must add
1 1 1
= + (63)
τ τph τdefect
i.e.
ρ = ρph + ρdefect Matthiesen’s Rule (64)

The defect contribution is proportional to the defect cross sec-


1
tion Σdefect and the current, or v(EF ), τ ∝ Σdefect v(EF ).
defect
Is is roughly temperature independent, since the cross section
Σdefect and v(EF ) 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
D E
2
the rms phonon excursion S (q) . However, from the equipar-
tition theorem
1 2
¿
2
À kB T
M ωq S (q) = T À θD . (65)
2 2
Thus
1 ¿
2
À kB T
∝ S (q) ∝ (66)
τph mωq2
Ie., at high temperatures, all modes contribute a linear in T

30
1
scattering to τ . Therefore, at T À θD (θD = debye temper-
ph
ature)
ρ = aT + ρdefect (67)

Ni
. 3 2%
+3 i
ρ Cu %N R
2 .16 i αT
Cu
+ %N
1 .12
+
Cu
Cu
5
αT

T
0.1
T Θ
D

Figure 21: The phonon and defect contributions to the resistivity add (left), and the
phonon contribution is linear at high temperatures T À θD .

6 Thermoelectric Effects

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 ∼ e2E
and a thermal current kT k∇T . In fact, a heat current can be

31
T1 T2
T1 ≠ T
x̂ 2

Figure 22: Thermoelctric effects 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


effect) and, conversely, an electric current can be used to move
heat (Peltier effect) which makes the solid state refrigeration
possible.

6.1 Linearized Boltzmann Equation

To allow for a thermal gradient ∇T , our formalism must be


modified. Imagine that ∇T and E are fixed in time, then the
Boltzmann equation becomes
 
∂f e ∂f f (k) − f0(k)
+ v · ∇ r f − E · ∇k f =   = − (68)
∂t h̄ ∂t S τ (k)

32
∂f
where in steady state ∂t → 0. After linearizing (replacing f
by f0 in the left-hand side), we get
( )
e
f (k) ' f0(k) − τ (k) v · ∇r f0 − E · ∇k f0 (69)

Then, as before
e e ∂f0 e ∂f0
E · ∇ k f0 = E · h̄v = Ex h̄vx (70)
h̄ h̄ ∂E h̄ ∂E
The spatial inhomogeneity is through ∇T , and in a semicon-
ductor for which EF depends strongly upon T , through ∇EF
   
 ∂f0 ∂f0   ∂f0 ∂f0 
v· ∇r f0 = v·  ∇T + ∇EF  = v·  ∇T − ∇EF 
∂T ∂EF ∂T ∂E
(71)
Apparently ∇EF only contributes a term which modifies the
electric field dependence
∂f0 ∂f0 0
vx {eEx + (∇EF )x} ≡ vx eE (72)
∂E ∂E x
Of course, in a metal E 0 = E.

6.2 Electric Current

Thus, we now have


∂f0 ∂T e 0 ∂f0
f (k) = f0(k) − τ vx + τ Exh̄vx (73)
∂T ∂x h̄ ∂E
33
Then for
e Z 3
jx = − 3 d kvx(k)f (k) (74)
8π  
e Z 3  ∂f0 ∂T e 0 ∂f0 
jx = − 3 d kvx(k) f0(k) − τ vx + τE h̄vx
8π ∂T ∂x h̄ x ∂E
recall that the last term yielded σ last time (and still will)
e Z 3 2 ∂f0 ∂T
jx = σEx0 + 3 d kvxτ (75)
8π ∂T ∂x
Again we will calculate the second term assuming a spherical
∂f0
Fermi surface. The term ∂T confines the integral to the Fermi

f
0

T=0

T≠0

E E
f

∂f0
Figure 23: The derivative ∂E
is only significant near the fermi surface.

sphere and so again effectively it amounts to a Fermi-surface


average, so v¯2 → 1 v2 ≈ 2 ∗ 1 m∗v2 = 2 ∗ E or changing to an
x 3 3m 2 3m

integral over the DOS


2 e Z ∂f0 ∂T
jx = σEx0 + dEτ (E)ED(E) (76)
3 m∗ ∂T ∂x

34
Assuming that τ (E) ∼ τ (EF ), we get
2 e ∂T Z ∂f0
jx = σEx0+ τ (E F ) dEE D(E) (77)
3 m∗ ∂x ∂T
2 e ∂T
jx = σEx0 + τ (E F ) c v (T ) c v ∝ m ∗
(78)
3 m∗ ∂x
In general, this intuitive form is rewritten as
 
∂T 
jx = σEx0 + L12
xx
− (79)
∂x
and from it we see that both an electric field (or the generalized
field strength E 0), and the thermal gradient contribute to the
electron current, jx.

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 = T dS = dU − µdN First Law of Thermodynamics


(80)
in terms of particle flow

jQ = j E − E F jn − ejn = j (81)

where this equation defines jQ, the thermal current, and


Z d3 k
jE = E(k)v(k)f (k, r) . (82)
8π 3
Again one could work out the form of jQ for the spherical Fermi
surface using the linearized Boltzmann equation. However one
must obtain a form like

j = L11E 0 + L12(−∇T ) (83)


jQ = L21E 0 + L22(−∇T ) (84)

(The fact that L12 = L21 is referred as the Onsaser relation.)


These relationship between the L’s and the transport co-
efficients depends upon what experiment is being done. For
example in Fig. 24 there is a potential gradient (V 6= 0) but no

36
T1 T1 ρ = σ-1
T1 = T
2 11
I≠0 j = L E′ j=σE
V≠0 12
j = L E′ 11
Q L =σ

V
I

Figure 24: Here, there is a potential gradient (V 6= 0) but no thermal gradient since
T1 = T2 . The electric field 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 T1 = T2. The electric field drives both


electric and thermal currents.

j = L11E
jQ = L12E (85)

Thus, we may identify


neτ 11
σ = = L (86)
m∗
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 off the
electric current. Here,
T1 T2
T1 ≠ T
2
I=0

V
I

³ ³ ´ ´
L12
Figure 25: Here j = 0 = L11 E 0 + L12 (−∇T ) and jQ = −L12 L11
+ L22 (−∇T )
³ ´
L12
where −L12 L11
+ L22 = κT

j = 0 = L11 E 0 + L12(−∇T ) (87)

and
   
12
21 0 21  L 
jQ = L E + L (−∇T ) = −L  11  + L22 
22 
 (−∇T )
L
(88)
and since (you will show)
2eτ
L12 = − ∗
cv = L21 (89)
3m
³ ´
12 2
L
κ = L22 − (90)
L11
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 find
T1 ≠ T
2
V1 V2 V2
V1
j
j Q
Q

V1 ≠ V V1 = V2 I
2

12
j = L (-∇T)
12 22
22 (L ) j = L (-∇T)
κ=L - 11
Q
L

22
κ=L

Figure 26: Two methods for measuring κ.

κ = L22 . (91)

Thus, we can identify


³ ´2
L12
κ = L22 22
or L − (92)
L11
depending upon the experiment. These are the same if the
sample is a good metal where L11 = σ is large (Young Kim).

39
David Mast measures κ by the method of the left of Fig. 26.
This yields the more conventional definition of κ.

6.4 Seebeck Effect, Thermocouples

These relations result in some interesting physical effects. Con-


sider a bimetallic conducting loop with two junctions main-
tained at temperatures T1 T2. Let metal A be different than B,
so that Lij ij
A 6= LB . If no current flows around the loop, then

metal A

T1 1 2 T
2

metal B T0 metal B

Figure 27: A bimetallic conducting loop with junctions maintained at T1 and T2 . If


T1 6= T2 , and L12 11 12 11
A /LA 6= LB /LB , then the heat current induces a potential V ∝ T2 −T1

 
12
L  dT
j = 0 = L11Ex + L22 (−∇T ) ⇒ Ex = 
  (93)
L11 dx

40
L12
where S = L11
is called the thermopower and is a property of
a material.
The potential measured around the loop is given by
Z 1 Z 2 Z 0
V = 0
EB dx + 1
EAdx + 2
EB dx (94)

or
 
∂T
Z 1 Z 0 ∂T  Z 2 ∂T
V = SB  0 dx + 2 dx + SA 1 dx (95)
∂x ∂x  ∂x
Z 1 ∂T Z 2 ∂T
= SB 2 dx + SA 1 dx (96)
∂x Z ∂x
T
= (SA − SB ) T 2 dT = (SA − SB ) (T2 − T1) (97)
1

R T2
Or if SA and SB are not T -independent V = T1 dT (SA − SB ).
So if T1 6= T2 and SA 6= SB , then the heat current in-
duces an emf! This is called the Seebeck effect ⇒ (solid state
thermometer with ice H2O as a reference).

6.5 Peltier Effect

Now consider the inverse situation where an electrical current j


is driven through the loop which is held at a fixed temperature

41
metal A

1 2

j T0 metal B
metal B
0

Figure 28: An electrical current j is driven through the loop which is held at a fixed
temperature
µ ¶
∂T
∂x = 0 . Then

jQ = L21E j = L11 E (98)


 
21
L 
jQ = 
11
 j = πj (99)
L
This is known as the Peltier effect 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 effect to create an
extremely simple (and similarly inefficient) refrigerator.

42
π j
j = A
Q A

j = (π A - π )j (π - π )j = j
Q B j A B Q
Q

=
π j
π Bj B
B T0 B
0

Figure 29: If dT /dx = 0 and πA = L21 11 21 11


A /LA 6= LB /LB = πB 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 κ and electrical σ


conductivities. However, in general one expects that κ ∝ σT
j = κ(-∇T)
Q j = σE
V1 V2 T T T = T
T1 ≠ T 1 2 1 2
2
T1 j T
2 V1 = V2 E V1 ≠ V
Q j 2

V1 V2

electric field -e/E

∂T
thermal field -
∂x

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 −eE. The current per unit electric
field proportional to e2. In thermal conduction each electron
carries a thermal energy kB T and is acted on by a thermal
force −kB ∇T . The heat current per unit thermal gradient is
proportional to kB2 T , thus one expects
κ kB2
∝ 2T (100)
σ e
Due to the simplicity of these arguments, our formalism
should reproduce this relationship. As we discussed before

jQ = j E − E F jn (101)
1 Z 3
= d k (E − EF )v(k)f (k) (102)
8π 3
In the linear approximation to the Boltzmann equation for E x0 =
0, we get
 
1 Z 3 ∂f0 2  ∂T 
jQ = 3 d k (E − EF ) v τ − (103)
8π ∂T x ∂x
∂f0
where E −EF and ∂T are odd in (E −EF ), for the Fermi liquid

 
∂T 1 Z d3 k ∂f0
2
j Q ' −   v F τF (E − E F ) (104)
∂x 3 3π 3 ∂T
 
∂T 1
jQ = −  vF3 τF cv (105)
∂x 3
44
∂f 0
(E - E )
F ∂T

EF E

Figure 31: The function (E − EF ) ∂f


∂T
0
is sharply peaked at the fermi surface and even.
in E − EF .

1
κ = vF3 τF cv (106)
3
2
Now recall that for the Fermi liquid cv = kB π2 nkB k TT so that
B F

1 m∗vF2 π2 T π2 kB2 T
κ= τF kB nkB = τF n ∗ (107)
3 m∗ 2 EF 3 m
However, also for the Fermi liquid, we found that σ = e2τF mn∗ ,
so
 
κ π 2  kB 2
= T (108)
σ 3 e
Of course this relationship only holds in a good metal. There
are two reasons for this. First we are neglecting terms like
2
(L12)
σ in κ which are unimportant for a good metal (or if we
electrically short the sample.) Second, we are assuming that κ
is dominated by electronic transport.

45
Chapter 10: Superconductivity

Bardeen, Cooper, & Schrieffer

May 9, 2001

Contents
1 Introduction 2
1.1 Evidence of a Phase Transition . . . . . . . . . . . . . . . . . . . . . 2
1.2 Meissner Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 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 Specific Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2 Microwave Absorption and Reflection . . . . . . . . . . . . . . . . . . 28
5.3 The Isotope Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

1
6 BCS ⇒ Superconducting Phenomenology 32

7 Coherence of the Superconductor ⇒ Meisner effects 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 (ρ = 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 ρ must be finite,
even at T = 0. Nevertheless many superconductors, for which
ρ = 0, exist. The first one Hg was discovered by Onnes in
1911. It becomes superconducting for T < 4.2◦K. Clearly this
superconducting state must be fundamentally different than the
”normal” metallic state. Ie., the superconducting state must
be a different 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 specific heat


(See Fig. 1). The jump in the superconducting specific heat Cs
indicates that there is a phase transition without a latent heat

3
C (J/mol°K)

Cn ∼ γT

CS

T T
c

Figure 1: The specific heat of a superconductor CS and and normal metal Cn . Below
the transition, the superconductor specific 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 < Tc

Cs ∼ e−β∆ (1)

gives us a clue to the nature of the superconducting state. It is


as if excitations require a minimum energy ∆.

1.2 Meissner Effect

There is another, much more fundamental characteristic which


distinguishes the superconductor from a normal, but ideal, con-
4
ductor. The superconductor expels magnetic flux, ie., B = 0
within the bulk of a superconductor. This is fundamentally dif-
ferent than an ideal conductor, for which Ḃ = 0 since for any
closed path
Superconductor

Figure 2: A closed path and the surface it contains within a superconductor.

I Z 1 Z ∂B
0 = IR = V = E · dl = S ∇ × E · dS = − S · dS , (2)
c ∂t
or, since S and C are arbitrary
1
0 = − Ḃ · S ⇒ Ḃ = 0 (3)
c
Thus, for an ideal conductor, it matters if it is field cooled or
zero field cooled. Where as for a superconductor, regardless
of the external field and its history, if T < Tc, then B = 0
inside the bulk. This effect, which uniquely distinguishes an

5
Ideal Conductor
Zero-Field Cooled Field Cooled

T > Tc T > Tc
B=0 B≠0

T < Tc T < Tc
B=0 B≠0

T<T T < Tc
c
B≠0 B=0

Figure 3: For an ideal conductor, flux penetration in the ground state depends on
whether the sample was cooled in a field through the transition.

ideal conductor from a superconductor, is called the Meissner


effect.
For this reason a superconductor is an ideal diamagnet. I.e.

µ−1
B = µH = 0 ⇒ µ = 0 M = χH = H (4)

6
1
χSC = − (5)

Ie., the measured χ, Fig. 4, in a superconducting metal is very
large and negative (diamagnetic). This can also be interpreted
χ ∝ D(E )
Pauli F
Tc
0
T

js
χ M


∼ ∼

H

-1

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 flux. 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
1
M=− H (6)
4π ext
in the interior of the superconductor in a direction opposite
to the applied field. The energy associated with this currents
7
increases with Hext . 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 field Hc

Normal
Hc

S.C.

Tc T

Figure 5: Superconductivity is destroyed by either raising the temperature or by ap-


plying a magnetic field.

2 The London Equations

London and London derived a phenomenological theory of su-


perconductivity which correctly describes the Meissner effect.
They assumed that the electrons move in a frictionless state, so
that

8
mv̇ = −eE (7)
∂j
or, since ∂t = −ensv̇,

∂js e2ns
= E (First London Eqn.) (8)
∂t m
Then, using the Maxwell equation
1 ∂B m ∂js 1 ∂B
∇×E =− ⇒ ∇ × + =0 (9)
c ∂t ns e 2 ∂t c ∂t
or
 
∂  m 1 
∇ × j s + B =0 (10)
∂t nse2 c
This described the behavior of an ideal conductor (for which
ρ = 0), but not the Meissner effect. To describe this, the
constant of integration must be chosen to be zero. Then
ns e 2
∇ × js = − B (Second London Eqn.) (11)
mc
m
or defining λL = ns e 2
, the London Equations become
B ∂js
= −λL∇ × js E = λL (12)
c ∂t

9

If we now apply the Maxwell equation ∇×H = c j ⇒ ∇×B =

c µj then we get
4π 4πµ
∇ × (∇ × B) = µ∇ × j = − 2 B (13)
c c λL
and
1 4πµ
∇ × (∇ × j) = − ∇×B=− 2 j (14)
λL c c λL
or since ∇ · B = 0, ∇ · j = 1c ∂ρ
∂t = 0 and ∇ × (∇ × a) =

∇(∇ · a) − ∇2a we get

4πµ 4πµ
∇2 B − B=0 ∇2 j − j=0 (15)
c2 λL c2 λL
x

SC

^ ^
∂Bx
B j ∝∇×B∝z×x
s ∂z
y z
j

Figure 6: A superconducting slab in an external field. The field penetrates into the
q
mc2
slab a distance ΛL = 4πne2 µ
.

10
Now consider a the superconductor in an external field shown
in Fig. 6. The field is only in the x-direction, and can vary in

space only in the z-direction, then since ∇ × B = c µj, the
current is in the y-direction, so
∂ 2Bx 4πµ ∂ 2jsy 4πµ
− 2 Bx = 0 − jsy = 0 (16)
∂z 2 c λL ∂z 2 c2 λL
with the solutions
− z − Λz
Bx = B0xe ΛL jsy = jsy e L (17)
s s
c 2 λL mc2
ΛL = 4πµ = 4πne2 µ
is the penetration depth.

3 Cooper Pairing

The superconducting state is fundamentally different 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
first quantified 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-

+
e-
ions
e- + + + + + +

region of + +
8 + + + +
vF ∼ 10 cm/s positive charge
attracts a second
+
electron

Figure 7: Origin of the retarded attractive potential. Electrons at the Fermi surface
travel with a high velocity vF . As they pass through the lattice (left), the positive ions
respond slowly. By the time they have reached their maximum excursion, the first
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 τ ∼ ωD ∼ 10−13 s after the
electron has passed. In this time the first electron has traveled

∼ vF τ ∼ 108 cm 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 first electron. Due
to retardation, the electron-electron Coulomb repulsion may be
neglected!
12
The net effect 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
k↑
e

ξ ∼ 1000Α°

e

- 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 quantified a bit better by considering


two electrons above a filled 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 ∼ h̄ωD .
Thus the scattering in phase space is restricted to a narrow
shell of energy width h̄ωD . Furthermore, the momentum in
k1

2
Ek ∼ k
k’1
k’ k’
1
2
ω
D
k’ k2
2 k2 k1

Figure 9: Pair states scattered by the exchange of phonons are restricted to a narrow
scattering shell of width h̄ωD around the Fermi surface.

this scattering process is also conserved

k1 + k2 = k01 + k02 = K (18)

Thus the scattering of k1 and k2 into k01 and k02 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 k1 = −k2 and σ1 = −σ2; ie., pairing is primarily between
14
time-reversed eigenstates.

scattering shell

k1 -k
2
K

Figure 10: If the pair has a finite center of mass momentum, so that k1 + k2 = 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.
h̄2 2
− (∇1 + ∇22)ψ(r1r2) + V (r1r2)ψ(r1r2) = (² + 2EF )ψ(r1r2)
2m
(19)
If V = 0, then ² = 0, and
1 ik1·r1 1 ik2·r2 1 ik(r1−r2)
ψV =0 = e e = e , (20)
L3/2 L3/2 L3

15
where we assume that k1 = −k2 = k. For small V, we will
perturb around the V = 0 state, so that
1 X ik·(r1 −r2 )
ψ(r1r2) = g(k)e (21)
L3 k
The sum must be restricted so that
h̄2k2
EF < < EF + h̄ωD (22)
2m
this may be imposed by g(k), since |g(k)|2 is the probability of
finding an electron in a state k and the other in −k. Thus we
take 




 k < kF
g(k) = 0 for 
√ (23)

 2m(EF +h̄ωD )
 k >


The Schroedinger equations may be converted to a k-space
equation by multiplying it by
1 Z 3 −ik0· r
dre ⇒ S.E. (24)
L3
so that
h̄2k 2 1 X
g(k) + 3 g(k0)Vkk0 = (² + 2EF )g(k) (25)
m L k0
where
Z
−k0 )·r 3
Vkk0 = V (r)e−i(k dr (26)
16
now describes the scattering from (k, −k) to (k0, −k0). It is
usually approximated as a constant for all k and k0 which obey
the Pauli-principle and scattering shell restrictions

 2

 h̄2 k2 h̄2 k0

 −V0 EF < 2m , 2m < EF + h̄ωD
Vkk0 = 

. (27)


 0 otherwise
so
 
2 2
 h̄ k V0 X
− + ² + 2EF 
 g(k) = −
3
g(k0) ≡ −A (28)
m L k0
or
−A h̄2 k2
g(k) = 2 2 (i.e. for EF < 2m < EF + h̄ωD )
− h̄ mk + ² + 2EF
(29)
Summing over k
V0 X A
= +A (30)
L3 k h̄2k2 − ² − 2EF
m
or
V0 X 1
1= (31)
L3 k h̄2k2 − ² − 2EF
m
This may be converted to a density of states integral on E =
h̄2 k2
2m

17
Z E +h̄ω
F D dE
1 = V0 EF
Z(EF ) (32)
2E − ² − 2EF
 
1 ² − 2h̄ω D
1 = V0Z(EF ) ln   (33)
2 ²
2h̄ωD −2/(V0 Z(EF )) V0
² = ' −2h̄ω D e < 0, as → 0
1 − e2/(V0 Z(EF )) EF
(34)

4 The BCS Ground State

In the preceding section, we saw that the weak phonon-mediated


attractive interaction was sufficient to destabilize the Fermi sea,
and promote the formation of a Cooper pair (k ↑, −k ↓). The
scattering

(k ↑, −k ↓) → (k0 ↑, −k0 ↓) (35)

yields an energy V0 if k and k0 are in the scattering shell EF <


Ek, Ek0 < EF + h̄ωD . 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
2² 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 wk is the probability that a pair state (k ↑, −k ↓) is occupied
then
X h̄2k2
Ekin = 2 wk ξk , ξk = − EF (36)
k 2m
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

|1ik (k ↑, −k ↓)occupied (37)


|0ik (k ↑, −k ↓)unoccupied (38)

or
|ψk i = uk |0ik + vk |1ik (39)

19
where vk2 = wk and u2k = 1 − wk . Then the BCS state, which
is a collection of these pairs, may be written as
Y
|φBCS i ' {uk |0ik + vk |1ik } . (40)
k
We will assume that uk , vk ∈ <. Physically this amounts to
taking the phase of the order parameter to be zero (or π), so
that it is real. However the validity of this assumption can only
be verified for a more microscopically based theory.
By the Pauli principle, the state (k ↑, −k ↓) can be, at most,
singly occupied, thus a (s = 21 ) Pauli representation is possible
   

1 
0
|1ik = 


 |0ik = 


 (41)
0 k 1 k
Where σk+ and σk−, describe the creation and anhialation of the
state (k ↑, −k ↓)
 

0 1
σk+ = 12 (σk1 + iσk2 ) = 


 (42)
0 0
 

0 0
σk− = 12 (σk1 − iσk2 ) = 


 (43)
1 0
   

0 
1
Of course σk+ 


 = 



1 k 0
σk+ |1ik = 0 σk+ |0ik = |1ik (44)
20
σk− |1ik = |0ik σk+ |0ik = 0 (45)

The process (k ↑, −k ↓) → (k0 ↑, −k0 ↓), if allowed, is


associated with an energy reduction V0. In our Pauli matrix
representation this process is represented by operators σk+0 σk−,
so
V0 X + −
V =− σ 0σ (Note that this is Hermitian) (46)
L3 kk0 k k
Thus the reduction of the potential energy is given by hφBCS |V | φBCS i

V 0 Y X
hφBCS |V | φBCS i = − 3  (up h0| + vp h1|) σk+σk−0
L p 
kk0
Yµ ¶

up0 |0ip0 + vp0 |1ip0 

(47)
p0

Then as k h1|1ik0 = δkk0 , k h0|0ik0 = δkk0 and k h0|1ik0 = 0


V0 X
hφBCS |V | φBCS i = − vk u k 0 u k vk 0 (48)
L3 kk0
Thus, the total energy (kinetic plus potential) of the system of
Cooper pairs is
X V0 X
WBCS = 2 vk2 ξk − vk u k 0 u k vk 0 (49)
k L3 kk0
As yet vk and uk are unknown. They may be treated as
variational parameters. Since wk = vk2 and 1 − wk = u2k , we
21
may impose this constraint by choosing

vk = cos θk , uk = sin θk (50)

At T = 0, we require WBCS to be a minimum.


P P
WBCS = k 2ξk cos2 θk − LV03 kk 0 cos θk sin θk0 cos θk0 sin θk
P P
= k 2ξk cos2 θk − LV03 1
kk 0 4 sin 2θk sin 2θk0 (51)
∂WBCS V0 X
= 0 = −4ξk cos θk sin θk − 3 cos 2θk sin 2θk0 (52)
∂θk L k0
1 V0 X
ξk tan 2θk = − 3 sin 2θk0 (53)
2 L k0
r
Conventionally, one introduces the parameters Ek = ξk2 + ∆2, ∆ =
V0 P V0 P
L3 k
u k vk = L3 k
cos θk sin θk . Then we get

ξk tan 2θk = −∆ ⇒ 2uk vk = sin 2θk = (54)
Ek
−ξk
cos 2θk = = cos2 θk − sin2 θk = vk2 − u2k = 2vk2 − 1 (55)
Ek
 
 
1 −ξk  1 ξk 
wk = vk2 = 1− = 1 − r

 (56)
2 Ek 2 2
ξk + ∆ 2
µ µ ¶¶
∆ 2 1 ξk
If we now make these substitutions 2uk vk = Ek , v k = 2 1− Ek

into WBCS , then we get


 
X ξ k  L3 2
WBCS = ξk 1 −
 − ∆. (57)
k Ek V0
22
wk = v 2
k T=0 clearly kinetic
energy increases
1

2 2
h k
0 ξ k = -E +
F 2m

Figure 11: Sketch of the ground state pair distribution function.

Compare this to the normal state energy, again measured


relative to EF
X
Wn = 2ξk (58)
k<kF
or
 
WBCS − Wn 1 X  ξk  ∆2
= − 3 ξk 1 + − (59)
L3 L k Ek V0
1
≈ − Z(EF )∆2 < 0. (60)
2
So the formation of superconductivity reduces the ground state
energy. This can also be interpreted as ∆Z(EF ) electrons pairs
per and volume condensed into a state ∆ below EF . 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
 
1 ξ k  L3 ∆ 2
X
WBCS = 2ξk 1 − − (61)
k 2 E k V0
↓ Lots of algebra (See I&L)
X
WBCS = − 2Ek vk4 (62)

Now recall that the probability that the Cooper state (k ↑, k ↓)


was occupied, is given by wk = vk2 . Thus the first pair breaking
excitation takes vk20 = 1 to vk20 = 0, for a change in energy
r
X X
∆E = − 2vk4 Ek + 2vk4 Ek = 2Ek0 = 2 ξk20 + ∆2 (63)
k6=k 0 k

h̄2 k 02
Then since ξk0 = 2m − EF , the smallest such excitation is just

∆Emin = 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 specific heat C ∼ e−β2∆ for T < Tc.
24
k′↑ e

e
-k′↓

2
w = v2 = 1 vk′ =0
k′ k′
q
Figure 12: Breaking a pair requires an energy 2 ξk2 + ∆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 find 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
r
Ek = ξk2 + ∆2 (65)
h̄2 k 02
For ξk2 À ∆, Ek = ξk = 2m − EF , 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 unaffected by any phase transition. Thus it must be
that the number of states in d3k is equal.

3
kz d k

ky
3
π
L

k
x

Figure 14: The number of k-states within a volume d3 k of k-space is unaffected by


any phase transition.

Ds(Ek )dEk = Dn(ξk )dξk (66)

In the vicinity of ∆ ∼ ξk , Dn(ξk ) ≈ Dn(EF ) since |∆| ¿ EF


26
(we shall see that ∆ ≤ 2wD ). Thus for ξk ∼ ∆
Ds(Ek ) dξx d r 2 Ek
= = Ek − ∆ 2 = r Ek > ∆
Dn(EF ) dEk dEk Ek2 − ∆2
(67)

Density of additional
∆ electron states only!

1 Ds Dn

Figure 15:

Given the experimental and theoretical importance of ∆, it


should be calculated.
V0 X V0 X V0 X ∆
∆ = sin θ k cos θ k = u v
k k = (68)
L3 k L3 k L3 k 2Ek
1 V0 X ∆
∆ = r (69)
2 L3 k ξk2 + ∆2
Convert this to sum over energy states (at T = 0 all states with

27
h̄2 k2
ξ < 0 are occupied since ξk = 2m − EF ).
V0 Z h̄ω
D Z(EF + ξ)dξ
∆ = ∆ √ 2 (70)
2 −h̄ωD ξ + ∆2
1 Z h̄ω dξ
= 0 D√ 2 (71)
V0Z(EF ) ξ + ∆2
 
1 h̄ω D
= sinh−1  (72)
V0Z(EF ) ∆
For small ∆,
h̄ωD 1
∼ e V0 Z(EF )
(73)

1
− V Z(E
∆ ' h̄ωD e 0 F) (74)

sinh x
∼ ex

Figure 16:

28
5 Consequences of BCS and Experiment

5.1 Specific 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 ∼ 2∆e−β2∆ T ¿ Tc (75)
∂∆E ∂β ∆2 −β2∆
C∼ ∼ 2e (76)
∂β ∂T T

5.2 Microwave Absorption and Reflection

Another direct measurement of the gap is reflectivity/absorption.


A phonon impacting a superconductor can either be reflected
or absorbed. Unless h̄ω > 2∆, the phonon cannot create an ex-
citation and is reflected. Only if h̄ω > 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 h̄ω < 2∆ and if B < Bc, then the microwave in-
tensity is high I = Is. On the other hand, if h̄ω > 2∆ ,or

29
superconductor I s - In
In B=0
cavity

10

microwave


hω = 2∆
B

Figure 17: If B > Bc or h̄ω > 2∆, then absorption reduces the intensity to the
normal-state value I = In . For B = 0 the microwave intensity within the cavity is
large so long as h̄ω < 2∆

B > Bc, then the intensity falls in the cavity I = In due to


absorbs ion by the walls.
Note that this also allows us to measure ∆ as a function of
T. At T = Tc, ∆ = 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 effected by the presence of a Cooper pair . The proba-
µ√ ¶
bility that an electron is unpaired is f 2 2
ξ + ∆ + EF , T =
√1 so, the probability that a Cooper pair exists is
exp β ξ 2 +∆2 +1

30
k′↑ e
kT ∼ 2∆

e
-k′↓

Figure 18:

√ 2µ
2

1 − 2f ξ + ∆ + EF , T . Thus for T 6= 0
( Ãr !)
1 Z h̄ω
D dξ
= 0 √ 2 2 2
1 − 2f ξ + ∆ + EF , T
V0Z(EF ) ξ + ∆2
(77)

Note that as ξ 2 + ∆2 ≥ 0, when β → ∞ we recover the
T = 0 result.
This equation may be solved for ∆(T ) and for Tc. To find Tc
In Pb
∆(T) Sn
∆(0)
Real SC data (reflectivity)

T/Tc
1

Figure 19: The evolution of the gap (as measured by reflectivity) as a function of tem-
perature. The BCS approximation is in reasonably good agreement with experiment.

31
T
consider this equation as Tc → 1, the first solution to the gap
equation, with ∆ = 0+, occurs at T = Tc. Here
 
1 Z h̄ω dξ ξ 
= 0 D tanh  (78)
V0Z(EF ) ξ 2kB Tc
which may be solved numerically to yield
1.14h̄ωD
1 = V0Z(EF ) ln (79)
k B Tc
kB Tc = 1.14h̄ωD e−1/{V0Z(EF )} (80)

but recall that ∆ = 2h̄ωD e−1/{V0Z(EF )} , so


∆(0) 2
= = 1.764 (81)
kB Tc 1.14

metal Tc◦K Z(EF )V0 ∆(0)/kB Tc


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)/kB Tc 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 Effect

Finally, one should discuss the isotope effect. We know that


Vkk0 , results from phonon exchange. If we change the mass of
one of the vibrating members but not its charge, then V0N (EF )
etc are unchanged but
v
u
u
u k 1
ωD ∼ t ∼ M −2 . (82)
M
1
Thus Tc ∼ M − 2 . This has been confirmed for most normal
superconductors, and is considered a ”smoking gun” for phonon
mediated superconductivity.

6 BCS ⇒ Superconducting Phenomenology

Using Maxwell’s equations, we may establish a relation between


the critical current and the critical field necessary to destroy the
superconducting state. Consider a long thick wire (with radius
r0 À ΛL) and integrate the equation

∇×H= j (83)
c

33
(r - r0 )/ΛL
j = j0 e

H H
• ⊗
S r0 j
0
dl

Λ
L

Figure 20: Integration contour within a long thick superconducting wire perpendicular
to a circulating magnetic field. The field only penetrates into the wire a distance Λ L .

along the contour shown in Fig. 20.


Z Z 4π Z
∇ × HdS = H · dl = j · ds (84)
c

2πr0H = 2πr0ΛLj0 (85)
c
If j0 = jc (jc is the critical current), then


ΛL jc Hc = (86)
c
Since both Hc and jc ∝ ∆, they will share the temperature-
dependence of ∆.
At T = 0, we could also get an expression for Hc by noting

34
that, since the superconducting state excludes all flux,
1 1 2
(W n − W BCS ) = H (87)
L3 8π c
However, since we have earlier
1 1 2
(W n − W BCS ) = N (0)∆ , (88)
L3 2
we get
r
Hc = 2∆ πN (0) (89)
c
We can use this, and the relation derived above jc = 4πΛL Hc ,

to get a (properly derived) relationship for jc.


r
c
jc = 2∆ πN (0) (90)
4πΛL
However, for most metals
n
N (0) ' (91)
EF
v
u
mc2
u
u
u
ΛL = t (92)
4πne2µ
taking µ = 1
v v
u
c 4πne2 u
u
u u πn2m
u
√ ne
jc = t 2∆ t = 2∆ (93)
4π mc2 h̄2kF2 h̄kF

35
This gives a similar result to what Ibach and Lüth get, but
for a completely different reason. Their argument is similar to
one originally proposed by Landau. Imagine that you have a
fluid which must flow around an obstacle of mass M . From the
perspective of the fluid, this is the same as an obstacle moving
in it. Suppose the obstacle makes an excitation of energy ² and

v M vP M
E

Figure 21: A superconducting fluid which must flow around an obstacle of mass M .
From the perspective of the fluid, this is the same as an obstacle, with a velocity equal
and opposite the fluids, moving in it.

momentum p in the fluid, then

E0 = E − ² P0 = P − p (94)

or from squaring the second equation and dividing by 2M


E′
(a) (b)
P P′
M M
E p

Figure 22: A large mass M moving with momentum P in a superfluid (a), creates an
excitation (b) of the fluid of energy ² and momentum p

36
P 02 P2 P·p p2
− =− + = E0 − E = ² (95)
2M 2M M 2M

P θ

v = P/M

P′

Figure 23:

pP cos θ p2
² = − (96)
M 2M
p2
² = pv cos θ − (97)
2M
If M → ∞ (a defect in the tube which carries the fluid could
have essentially an infinite mass) then
²
= v cos θ (98)
p
Then since cos θ ≤ 1
²
v ≥ (99)
p
Thus, if there is some minimum ²,then there is also a mini-
mum velocity below which such excitations of the fluid cannot

37
happen. For the superconductor
²min 2∆
vc = = (100)
p 2h̄kF
Or
ne
jc = envc = ∆ (101)
h̄kF
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


effects

Superconductivity is the Meissner effect, but thus far, we have


not yet shown that the BCS theory leads to the second London
equation which describes flux 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  pF ∆
2∆ = δ   ∼ δp ⇒ δp ∼ 2m (102)
2m m pF
h̄ h̄pF h̄2kF EF
ξcp ∼ δx ∼ ∼ = = (103)
δp 2m∆ 2m∆ kF ∆

ξcp ∼ 103 − 104 A∼ size of Cooper pair wave function (104)

Thus in the radius of the Cooper pair, about

 
4πn  ξcp 3
∼ 108 (105)
3 2
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
specifically anchored to each other; ie., they maintain coherence
over a length scale of at least ξcp.
Normal Metal
SC 2
φBCS

ξ coh > ξ cp

Figure 25:

In light of this coherence, lets reconsider the supercurrent

2e
j=− {ψp∗ψ ∗ + ψ ∗pψ} (106)
4m
where pair mass = 2m and pair charge = −2e.
2e
p = −ih̄∇ − A (107)
c
A current, or a CM momentum K, modifies the single pair state
1 X (r1 +r2 )/2 ik· (r1 −r2 )
ψ(r1, r2) = 3
g(k)eiK· e (108)
L k
ψ(K, r1, r2) = ψ(K = 0, r1, r2)eiK·R (109)
40
r1 +r2
where R = 2 is the cm coordinate and h̄K is the cm mo-
mentum. Thus

ΦBCS ' eiφΦBCS (K = 0) = eiφΦ(0) (110)

φ = K · (R1 + R2 + · · ·) (111)

(In principle, we should also antisymmetrize this wave function;


however, we will see soon that this effect is negligible). Due to
the rigidity of the BCS state it is valid to approximate

∇ = ∇R + ∇r ≈ ∇R (112)

Thus
  
2e X  ∗  2eA 
js ≈ Φ −ih̄∇ R + ΦBCS
4m ν  BCS ν
c 
 ∗
2eA ∗


+ΦBCS ih̄∇Rν +
  ΦBCS  (113)
c 

or
 
2e  2 4eA 2X 
js = − |Φ(0)| + 2h̄ |Φ(0)| ∇ Rν φ (114)
2m  c ν

Then since for any ψ, ∇ × ∇ψ = 0


2e2
∇ × js = − |Φ(0)|2 ∇ × A (115)
mc
41
or since |Φ(0)|2 = ns
2

ne2
∇×j=− B (116)
mc
which is the second London equation which as we saw in Sec.??
leads to the Meissner effect. 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 flux 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 ΛL from the surface).
e 2 ns eh̄ns X
js = − A− ∇ Rν φ (117)
mc 2m ν
Z e 2 ns Z eh̄ns X Z
◦js · dl = − ◦A · dl − ◦∇Rν φ · dl (118)
ms 2m ν
Presumably the phase of the BCS state ΦBCS = eiφΦ(0) is

42
superconducting loop

C
X
X
X
X
X
X X X
B
X X X X ΛL
X X
X
X X

Figure 26: Magnetic flux penetrating a superconducting loop is quantized. This may
be seen by integrating Eq. 114 along a contour within the superconducting bulk (a
distance ΛL from the surface).

single valued, so
XZ
∇Rν φ · dl = 2πN N ∈Z (119)
ν

Also since the path l may be taken inside the superconductor


by a depth of more than ΛL, where js = 0, we have that
Z
js · dl = 0 (120)

so
e 2 ns Z e 2 ns Z eh̄ns
− A · dl = − B · ds = 2N π (121)
ms ms 2m
Ie., the flux 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
V
a b c

V0

metal insulator metal


x
0 d d2ψ 2m
d 2ψ 2m + 2 Eψ = 0
+ 2 Eψ = 0 dx 2 h
dx 2 h 2
d ψ + 2m (E - V )ψ
h2 0
dx2

Figure 27:

0 0
ψa = A1eikx + B1e−ikx ψb = A2eik x + B2e−ik x
ψc = B3e−ikx (122)

These are solutions to the S.E. if


2mE
k = in a & c (123)
r

2m(E − V0)
k0 = in b (124)

44
The coefficients are determined by the BC of continuity of ψ
and ψ 0 at the barriers x = 0 and x = d. If we take B3 = 1
and E < V0, so that
r

0 2m(E − V0)
k = iκ = (125)

then, the probability of having a particle tunnel from left to
right is

     −1
|B3|2 1 1
 1  k κ 2 1  k κ 2 

Pl→r ∝ = = − − + + cosh 2κd
|B1|2 |B1|2 2 8 κ k 8 κ k 

(126)
For large κd

 −2
k κ
Pl→r ∝ 8  +  e−2κd (127)
κ k  r 
 −2  
k κ 
 2d 2m(V 0 − E) 

∝ 8  +  exp  −  (128)
κ k 
 h̄ 

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
filled 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

S I N
E

eV
X N(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)
2∆ 4kB Tc
We know that k B Tc ∼ 4, ∆ ∼ 2 ∼ 10◦K. However typical
metallic densities of states have features on the scale of electron-
46
volts ∼ 104◦K. Thus, on this energy scale we may approximate
the metallic density of states as featureless.

Nr (²) = Nmetal (²) ≈ Nmetal (EF ) (130)

The tunneling current is then, roughly,


Z
I∝ P d²f (² − eV )Nr (EF )Nl (²)(1 − f (²))
Z
−P d²f (²)Nl (²)Nr (EF )(1 − f (² − eV )) (131)

For eV = 0, clearly I = 0 i.e. a balance is achieved. For

EF

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 6= 0 a current may occur. Let’s assume that eV > 0


and kB T ¿ ∆. Then the rightward motion of electrons is
47
suppressed. Then
Z
I ∼ P Nr (EF ) d²f (² − eV )Nl (²) (132)

and
dI Z ∂f (² − eV )
∼ P Nr (EF ) d² Nl (²) (133)
dV ∂V
∂f
∼ eδ(² − eV − EF ) (T ¿ EF ) (134)
∂V
dI
' P Nr (EF )Nl (eV + EF ) (135)
dV
dI
Thus the low temperature differential conductance dV is a mea-
sure of the superconducting density of states.
dI
I
dV

V
∆/e V ∆/e

Figure 30: At low temperatures, the differential 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 ² . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Kramer’s Kronig Relations . . . . . . . . . . . . . . . . . . . . . . . 6

2 Absorption of E and M radiation 8


2.1 Transmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Reflectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Model Dielectric Response . . . . . . . . . . . . . . . . . . . . . . . . 13

3 The Free-electron gas 17

4 Excitons 19

1
Electromagnetic fields are essential probes of material prop-
erties

• IR absorption

• Spectroscopy

The interaction of the field and material may be described ei-


ther classically or Quantum mechanically. We will first 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
1 ∂B 4π 1 ∂D
∇×E =− , ∇×H= j+ (1)
c ∂t c c ∂t
and Ohm’s law
j = σE . (2)

Both effects may be combined into an effective dielectric con-


stant ²̃, which we will now show. For an isotropic medium, we
2
x e

x << λ
x
0
E(x,t) ≅ E(x0 ,t)

k << G ε(k,ω) ≅ ε(ω)

Figure 1: If the average excursion of the electron is small compared to the wavelength
of the radiation < x >¿ λ, then we may ignore the wave-vector dependence of the
radiation so that ²(k, ω) ≈ ²(ω).

have
D(ω) = ²(ω)E(ω) (3)

where
Z
R
E(t) = dωe −iωt
E(ω) H(t) = dωe−iωt H(ω) (4)
Z
R
D(t) = dωe−iωt D(ω) B(t) = dωe−iωt B(ω) (5)

E(ω) = E ∗(−ω) ⇒ E(t) ∈ < (6)



Then ∇ × H = c j + 1c ∂D
∂t ⇒

3
Z
−iωt 4π Z −iωt 1∂ Z
∇ × dωe H(ω) = dωe j(ω) + dωe−iωt D(ω)
c c ∂t
 
(7)
Z  4π 1 
dωe−iωt ∇ × H(ω) − j(ω) − (−iω)D(ω) = 0 (8)
c c
4π ω
∇×H = j − i D(ω)
c c
4πσ ω
= E − i ²E
c µ
c


 ω c 4πσ

 −i E ² −

c ω c = − iωc E ²̃
= 

µ ¶ (9)
 4π c iω 4π
c E σ − 4π c ² = c E σ̃

Thus we could either define an effective conductivity σ̃ = σ −


iω²
4π which takes into account dielectric effects, or an effective
dielectric constant ²̃ = ² + i 4πσ
ω , which accounts for conduction.

1.1 Conditions on ²

From the reality of D(t) and E(t), one has that E(+ω) =
E ∗(−ω) and D(ω) = D∗(−ω), hence for D = ²E,

²(ω) = ²∗(−ω) (10)

4
Additional constraints are obtained from causality
Z
D(t) = dω²(ω)E(ω)e−iωt
0
−iωt dt iωt0
Z Z
= dω²(ω)e e E(t0)

1 Z 0
= dtdω (²(ω) − 1 + 1) E(t0)e−iω(t−t ) (11)

then we make the substitution
²−1
χ(ω) ≡ (12)

µ ¶
R 1 0
D(t) = 2 dtdω χ(ω) + 4π E(t0)e−iω(t−t )
R 0
D(t) = E(t) + 2 dtdωχ(ω)E(t0 )e−ω(t−t ) (13)
R
Define G(t) ≡ 2 dωχ(ω)e−iωt , then
Z ∞
D(t) = E(t) + −∞
dt0G(t − t0)E(t0 ) (14)

Thus, the electric displacement at time t depends upon the field


at other times; however, it cannot depend upon times t0 > t by
causality. Hence

G(τ ) ≡ 0 τ <0 (15)

5
no poles
τ<0

X X
poles

Figure 2: If χ(ω) is analytic in the upper half plane, then causality is assured.

This can be enforced if χ(ω) is analytic in the upper half


plane, then we may close
Z ∞
G(τ ) = 2 −∞
dωχ(ω)e−iωτ
Z
= 2 ◦dωχ(ω)e−iωτ ≡ 0 (16)

contour in the upper half plane and obtain zero for the integral
since χ is analytic within and on the contour.

1.2 Kramer’s Kronig Relations

One may also derive an important relation between the real


and imaginary parts of the dielectric function ²(ω) using this
analytic property. From the Cauchy integral formula, if χ is

6
analytic inside and on the contour C, then
1 Z χ(ω 0) 0
χ(z) = ◦ dω (17)
2πi C z − ω 0
Then taking the contour shown in Fig. 3 and assuming that
ω’

ω.

Figure 3: The contour (left) used to demonstrate the Kramer’s Kronig Relations.
Since the contour must contain ω, we must deform the contour so that it avoids the
pole 1/(ω − ω 0 ).

<
χ(ω) ∼ 1 <
for large ω (in fact Reχ ∼ 1 <
and Imχ ∼ 1
) we
ω ω ω2

may ignore the large semicircle. Let z = ω + i0+, so that


+ 1 Z ∞ χ(ω 0)dω 0
χ(ω + i0 ) = (18)
2πi −∞ ω − ω 0 − i0+
Z ∞ χ(ω 0 )dω 0
+
2πiχ(ω + i0 ) = P −∞ + iπχ(ω) (19)
ω − ω0
0
1 Z∞ 0 χ(ω )
χ(ω) = P −∞ dω (20)
iπ ω − ω0
Then let 4πχ(ω) = ²(ω) − 1 = ²1 + i²2 − 1 and we get
2 Z ∞ −i²1(ω 0) + ²2(ω 0) + 1 0
²1(ω) + i²2(ω) − 1 = P −∞ dω (21)
π ω − ω0
7
or
1 Z ∞ ²2(ω 0) 0
²1(ω) − 1 = P −∞ dω (22)
π ω − ω0
1 Z ∞ ²1(ω 0) − 1 0
²2(ω) = − P −∞ dω (23)
π ω − ω0
which are known as the Kramer’s Kronig relations. Many ex-
periments measure ²2 and from Eq. 22 we can calculate ²1!

2 Absorption of E and M radiation

2.1 Transmission

~
ξ = ξ e-iω(t-nx/c)
0

~=1 ~=1
n
n
~n = n + iκ = √ε(ω)
~n = n2 +2inκ - κ 2 = ε + iε
1 2
2
ε = n - κ2
ε 1
ε = 2nκ
2

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-
flected and part is transmitted (see Fig. 5). In the case of normal
incidence, it is easy to calculate the related coefficients from the
conditions of continuity of E⊥ and H⊥ = B⊥(µ = 1)

µ=1 µ=1
B0
B′
X
ξ
0 X
ξ″ ξ′
X

B″

Figure 5: The assumed orientation of electromagnetic fields incident on a surface.

1 ∂B iω
∇×E =− , ik × E = B (24)
c ∂t c
|nE⊥| = |B⊥| (25)



00 0 
E0 + E = E 
 2
 t= (26)
(E0 − E ) = nE 00 0 


 n+1

9
In this way the other coefficients may be calculated (see Fig. 6).
Accounting for multiple reflections the total transmitted field is

t = ~n 2+ 1
1

r ~
2n
t 1
1
t =
2 ~
n+1

r ~n - 1

t 2
2
r = 2
~n + 1

Figure 6: Multiple events contribute the the radiation transmitted through and reflected
from a thin slab.

ñω
E = E0t1t2eikd + E0t1r2r2t2e3ikd + · · · , where k =
c
(27)
E0t1t2eidñω/c
E= (28)
1 − r22e2idñω/c
If ñ ∼ 1 and d is small, then

E ' E0eidñω/c (29)


∗ )/c
I ' I0eidω(ñ−ñ = I0e−dω2κ/c (30)
²2
κ = √ (31)
2 ²1
Ã
−(²2 ω/c)d ²2 ω !
I ' I0 e ' I0 1 − d (32)
c
10
If the thickness d is known, then the quantity
ω²2(ω) 4πσ1(ω)
= (33)
c c
may be measured (the absorption coefficient). The real part
of the dielectric response, ²1(ω) may be calculated with the
Kramers Kronig relation
1 Z ∞²2(ω 0) 0
²1(ω) = 1 + P −∞ ω 0 − ω
dω (34)
π
According to Young Kim, this analysis works for sufficiently
thin samples in the optical regime, but typically fails in the IR
where ²2 becomes large.

2.2 Reflectivity

Of course, we could also have performed the experiment on a


very thick slab of the material, and measured the reflectivity R
R
However, this is a much more complicated experiment since I

depends upon both ²1 and ²2(ω), and is hence much more dif-
ficult to analyze [See Frederick Wooten, Optical Properties
of Solids, (Academic Press, San Diego, 1972)].

11
2
~
n=1 R= n~ - 1
I ~
n +1

~
n

Figure 7: The coefficient of reflectivity (the ratio of the reflected to the incident inten-
sities) R/I depends upon both ²1 (ω) and ²2 (ω), making the analysis more complicated
than in the transmission experiment.

To analyze these experiments, one must first measure the


reflectivity, R(ω) over the entire frequency range. We then
write
(n − 1)2 + κ2

R(ω) = r(ω)r (ω) = (35)
(n + 1)2 + κ2
where ñ = n + iκ and

r(ω) = ρ(ω)eiθ (36)

so that R(ω) = ρ(ω)2. If ρ(ω) → 1 and θ → 0 fast enough


as the frequency increases, then we may employ the Kramers
Kronig relations replacing χ with ln r(ω) = ln ρ(ω) + iθ(ω), so
that r
0
0 2ω Z ∞ ln R(ω)
θ(ω ) = − P 0 dω (37)
π ω 2 − ω 02
12
Thus, if R(ω) is measured over the entire frequency range
where it is finite, then we can calculate θ(ω). This complete
knowledge of R is generally not available, and various extrapo-
lation and fitting schemes are used on R(ω) so that the integral
above may be completed.
We may then use Eq. 35 above to relate R and θ to the real
and imaginary parts of the refractive index,
1 − R(ω)
n(ω) = r (38)
1 + R(ω) − 2 R(ω) cos θ(ω)
r
2 R(ω) sin θ(ω)
κ(ω) = r , (39)
1 + R(ω) − 2 R(ω) cos θ(ω)
and therefore the dielectric response ²(ω) = (n(ω) + κ(ω)) 2.

2.3 Model Dielectric Response

We have seen that EM radiation is a sensitive probe of the


dielectric properties of materials. Absorption and reflectivity
experiments allow us to measure some combination of ²1 or
²2, 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 field 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

µs̈ = −µγ ṡ − µω02s + e∗E (41)

where the first term on the right-hand side is the damping force,
the second term is the restoring force, and the third term is the
external field. Furthermore, the polarization P is
N ∗ N
P = e s + αE (42)
V V
where α is the polarizability of the different molecules which
make up the material. It is to represent the polarizability of
rigid bodies. For example, (see Fig. 8) a metallic sphere has

α = a3 (43)
14
a

E=0

Figure 8: We may model our material as a system harmonically bound charge and of
metallic spheres with polarizablity α = a3 .

If we F.T. these two equations, we get

−µω 2s = iωµrs − µω02s + e∗E(ω) (44)


½ ¾ e∗E(ω)
⇒ s(ω) ω02 2
− ω − iωγ = (45)
µ
and
P (ω) = ne∗s(ω) + nαE(ω) (46)

or  
∗2
ne /µ 
 

P (ω) = E(ω)  + nα (47)
ω02 − ω 2 − iωγ
 

The term in brackets is the complex electric susceptibility of


the system χ = E/P .
ne∗2 /µ 1
χ(ω) = nα + 2 = (²(ω) − 1) (48)
ω0 − ω 2 − iωγ 4π
4πne∗2 /µ 4πne∗2
²(ω) = 1 + 4πnα + 2 , where ωp2 =
ω0 − ω 2 − iγω µ
(49)
15
Or, introducing the high and zero frequency limits

²∞ = 1 + 4πnα (50)
ωp2 ωp2
²0 = 1 + 4πnα + 2 = ²∞ + 2 (51)
ω0 ω0
ω02(²0 − ²∞)
²(ω) = ²∞ + 2 (52)
ω0 − ω 2 − iγω
For our causality arguments we must have no poles in the upper
complex half plane. This is satisfied since γ > 0. In addition,
we need
1
χ = (²(ω) − 1) → 0, as ω → ∞ (53)

This means that ²∞ = 1 + 4πnα = 1. Of course α is finite.
The problem is that in making α = constant, we neglected the
electron mass. Ie., for our example of a metallic sphere, α < a3
for very high ω! (See Fig. 9) due to the finite electronic mass.
To analyze experiments ² is separated into real ²1 and imag-
inary parts ²2
4π (²0 − ²∞)ω02(ω02 − ω 2)
²1(ω) = − σ2 = ²∞ + (54)
ω (ω02 − ω 2)2 + γ 2ω 2
4π (²0 − ²∞)ω02γω
²2(ω) = σ1 = 2 (55)
ω (ω0 − ω 2)2 + γ 2ω 2
16
a

--
- - -
↑ ξ(ω)

Figure 9: For very high ω, α < a3 since the electrons have a finite mass and hence
cannot respond instantaneously to changes in the field.

Thus a phonon mode will give a roughly Lorentzian-like line


shape in the optical conductivity σ1 centered roughly at the
phonon frequency. In addition, this form may be used to con-

struct a model reflectivity R = |ñ − 1|2/|ñ + 1|2 , with ñ = ²
which is often appended to the high end of the reflectivity data,
so that the Kramers-Kronig integrals may be completed.

3 The Free-electron gas

Metals have a distinct feature in their optical conductivity σ 1(ω)


which may be emulated by the free-electron gas. The equation
of motion of the free-electron gas is

nms̈ = −γ ṡ − neE (56)


17
ε2 (ω)

ε1 (ω)

ε γ
0

ε∝
ω0 ω0 ω

Figure 10: Sketch of the real and imaginary parts of the dielectric response of the
harmonically bound charge model.

In steady state γ ṡ = −neE; however


ne2τ
−neṡ = j = σE = E (57)
m
in the relaxation time approximation, so
eτ ne nm
ṡ = − E =− E or γ = . (58)
m γ τ
Thus, the equation of motion is
nm
nms̈ = − ṡ − neE (59)
τ
If we work in a Fourier representation, then
iωm
−mω 2s(ω) = s(ω) − eE(ω) (60)
τ
18
However, since P = −ens
 
mω 2 +
iωm 
P = −ne2E(ω) (61)
τ
or
ne2/m 1
P =− 2 E = χE = (² − 1)E (62)
ω + iω/τ 4π
 
4πn2e2 1 ωp2 
 ω − i/τ  
² =1− =1− (63)
mω ω + i/τ ω  ω 2 + 1/τ 2 

 
ωp2 
 ω 
 ωp2 1
²1 = 1 − , ² 2 = (64)
ω   ω 2 + 1/τ 2 
 τ ω ω 2 + 1/τ 2
However, recall that ²1 = − 4πσ 4πσ1
ω and ²2 = ω , so that
2

ωp2 1/τ π
σ1(ω) = . (65)
4 ω 2 + 1/τ 2

4 Excitons

One of the most dramatic effects 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
1 ∗ 2 1 ∗ 2 e2
H = m h vh + m e ve − (66)
2 2 ²r
19
Drude peak
σ1 (ω) (electronic)

∫ σ (ω) dω = ω2p /4
1

1/τ
phonons

Figure 11: A sketch of the optical conductivity of a metal at finite 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 1 ∗ 2 1 ∗
mhvh + me ve = (mh + m∗e )Ṙ2 + µṙ2 (67)
2 2 2
m∗e x̄e + m∗hx̄h
R̄ = , r = x̄e − x̄h (68)
m∗h + m∗e
The eigenenergies may be obtained from the Bohr atom solution
En = −µ(Ze2)2/(2h̄2n2) by making the substitutions
e2 2
Ze → (69)
²
∗ m∗hm∗e
µ → µ = ∗ (70)
mh + m∗e
In addition there is a cm kinetic energy, let h̄k = P , then
h̄2K 2 µ∗ e 4
EnK = − + Eg (71)
2(m∗h + m∗e ) ²22h̄2n2
20
K

electron
Eg r
hole 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 (≈ 1012 impurities per cc) the lifetime can exceed
τSi ≈ 10−5 s; whereas in direct-gap GaAs, the lifetime is much shorter τGaAs ≈ 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 effects,


since ² ' 10. (it is also reduced by the effective 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
²µ
a' a0 (72)
me
This in fact is the justification 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.
+
e

°
∼ 10A

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 ² ≈ 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
< 1eV . Most condensed
insulators in that the gap is small ∼
metal semiconductor insulator

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)
ne2τ eτ
σ = , µ = , σ = neτ (1)
m∗ m∗
1
Almost always τ increases as T increases, ie the thermal exci-
tations increase the scattering rate and decrease the lifetime of
2
the quasiparticle.
1
For example, we have seen that τ ∼ 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
τ.
1
metals σ ∼ τ∼
, σ ↓ as T ↑ (2)
T
However, in semiconductors, the population of free carriers n is
temperature dependent. The exponential always will dominate

-Eg /2kT
Eg ∼e ∼n

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 τ .


1 −Eg /kT
σ ∼ τn ∼ e (3)
T
σ ↑ as T ↑ (4)

The same is true for insulators, of course, except here n is so


3
small that for all realistic purposes σ ∼ 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.
E
conduction band

AB

Eg P
Eg ↓ T↓
S
sp3 ⇒ 4 electrons
per band
B valence band

r0 r

Figure 3: Sketch of the sp3 bands in Si vs. Si-Si separation.

Recall that Si and Ge have a s2p2 atomic shell, which forms


highly directional sp3 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
111
L Eg

110
K

000
Γ X
100
Si

L Γ X K Γ

Figure 4: Sketch of quasiparticle bands in Si (right) along the high symmetry directions
(left).

1 1 ∂ 2E(k)
= (5)
m∗ij h̄2 ∂ki∂kj
The situation in III-V semiconductors such as GaAs is sim-
ilar, in that covalent sp3 bands still form. However, the gap
is direct. For this reason GaAs makes more efficient 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
Eg

Ge

L Γ X K Γ

Figure 5: Sketch of quasiparticle bands in Ge along the high symmetry directions.


Note the indirect, roughly Γ → 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 τexciton
GaAs 1ns(10−9 s)
Si 19µs(10−5 s)
Ge 1ms(10−3 s)
Table 1:

6
E g ∼ 1.5eV

GaAs

L Γ X K Γ

Figure 6: Sketch of quasiparticle bands in GaAs along the high symmetry directions
of the Brillouin zone. Note the direct, Γ → Γ, 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
h̄2 k2
full near the top or bottom of bands, where Ek ∼ 2m∗p and
the effective mass approximation is valid. Here, we found that
µ ∼ eτ /m∗
However, as mentioned before, the carrier concentrations are

7
conduction
hω ∼ Eg
phonon vs k ≅ ω

photon ck ≅ ω ≅ E g/ h
valence

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+ v j n = # electrons
volume
E e-
p = # holes
v j volume

σ = e(nµ n + pµ 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).
Z E Z ∞
top
n= Ec
DC (E)f (E, T )dE → Ec
DC (E)f (E, T )dE (6)
Z E Z E
v v
p= Ebottom
DV (E) {1 − f (E, T )} dE → −∞
DV (E) {1 − f (E, T )} dE
(7)
To proceed further we need forms for DC and DV . Recall that in

8
2 2
h k
∼E k ∼
Ec 2m*
µ ∼ eτ* Eg n
m
E 2 2
v ∼E k ∼ h k*
2m p

Figure 9:

h̄2 k2
the parabolic approximation Ek ' 2m∗ we found that D(E) =
3√
(2m∗ ) 2
2π 2 h̄3
E. Thus,
3
(2m∗n) 2 √
DC (E) = 2 3
E − EC (8)
2π h̄
µ ¶3
2m∗p 2 √
DV (E) = EV −E (9)
2π 2h̄3
for E > EC and E < EV respectively, and zero otherwise
EV < E < E C .
In an intrinsic (undoped) semiconductor n = p, and so EF
must lie in the band gap. However, if m∗n 6= m∗p (ie. DC 6=
DV ), then the chemical potential, EF , 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 kB T , since typically
9
E g = E C − E V À kB T .

Eg (eV ) ni(cm−3 )(300◦ K)


Ge 0.67 2.4 × 1013
Si 1.1 1.5 × 1010
GaAs 1.43 5 × 107
Table 2:

1eV
∼ 10000◦ K À 300◦K ∼ T (10)
kB
>
Thus, assuming that E − EF ∼
Eg
À kB T
2

1 1
' = e−(E−EF )/kB T (11)
e(E−EF )/kB T + 1 e(E−EF )/kB T
ie., Boltzmann statistics. A similar relationship holds for holes
where −(E − EF ) ∼ > E g À kB T
2

1 ½
−(E−EF )/kB T
¾
1− ' 1− 1 − e = e−(E−EF )/kB T
e(E−EF )/kB T + 1
(12)
since (1 − f (E)) = f (−E) and e(E−EF )/kB T is small. Thus, the
concentration of electrons n
3
(2m∗n) 2 EF /kB T Z ∞ √ −E/kB T
n ' 3 e E − E C e dE (13)
2π 2h̄ EC

10
3
(2m∗n) 2 3 Z
−β(EC −EF ) ∞ 21 −x
= (kB T ) e
2 x e dx (14)
2π 2h̄3 0
 ∗ 3
2πm k T
n B 
2
= 2 2
e−β(EC −EF ) = Nef
C −β(EC −EF )
fe (15)
h
Similarly
 3

2πm∗p kB T  2 −β(EV −EF ) V −β(EV −EF )
p = 2  e = N ef f e (16)
h2
C V
where Nef f and Nef f are the partition functions for a classical

gas in 3-d and can be regarded as ”effective 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,
 3µ
k B T ¶3
∗ ∗ 2 −βEg
np = 4   mn mp e (17)
2πh̄2
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.
 3
kB T  2 µ ∗ ∗¶ 43 −βEg /2
ni = p i = 2  mn mp e (18)
2πh̄2
However, we already have relationships for n and p involving
EC and EV
C −β(EC −EF ) V β(EV −EF )
n = p = Nef fe = Nef fe (19)
V
Nef
e 2βEF
= C f eβ(EV +EC ) (20)
Nef f
or  
1 1 NV
 ef f 
EF = (EV + EC ) + kB T ln  C  (21)
2 2 Nef f
 

1 3 m p
EF = (EV + EC ) + kB T ln  

 (22)
2 4 mn
Thus if m∗p 6= m∗n, the chemical potential EF 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

σ = neµ, so the conductivity depends linearly upon the doping


(it may also effect µ in some materials, leading to a non-linear
12
doping dependence). A typical metal has

nmetal ∼ 1023/(cm)3 (23)

whereas we have seen that a typical semiconductor has


1010
niSeC ∼ atT ' 300◦K (24)
cm3
Thus the conductivity of an intrinsic semiconductor is quite
small!
To increase n (or p) to ∼ 1018 or more, dopants are used.
For example, in Si the elements used as dopants have either a
s2p1 or s2p3 atomic valence. Thus, in the tetrahedral bonding
of Si there is either an extra electron (half bond) or an un-
satisfied bond or a hole. Thus P or B will either donate or
2 2
Si 3s 3p
Si Si Si Si Si Si Si Si
e+ 2 3
P 3s 3p
Si Si Si Si Si B Si Si 1
e- B 3s2 3p
Si P Si Si Si Si Si Si
2
r r = h ε 2 big!
Si Si Si Si Si Si Si Si m* e

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 difference is
that here the donor/acceptor is fixed and may be treated as
having infinite mass, thus the binding energy is given by
m∗ e 4
E= 2 2 2 (25)
2² h̄ n





 hole mass acceptor(B)
m∗ = 

(26)


 electron mass donor(P )
m∗
Again, since m < 1 and ² ∼ 10 these energies are often much
less than 13.6eV c.f. in Si E ∼ 30M eV ∼ 300◦K or in Ge
E ∼ GM eV ∼ 60◦K. Thus thermal excitations will often
ionize these dopant sites. In terms of energy levels
n - SeC p - SeC

EC EF
EC
ED unoccupied at T= 0
(occupied by holes at T = 0)
occupied at T= 0 EA
EV EV EF

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

C V −βEg
np = Nef f Nef f e = n2i = p2i (27)

Now imagine the temperature is finite so that some of the donors


or acceptors are ionized. Furthermore, in equilibrium, the semi-

# ionized
EC
+ 0
ED ND = N D + ND

EF
# un-ionized
+ 0
EA NA = NA + NA
EV

Figure 12:

conductor is charge neutral so that

n + NA− = p + ND+ (28)

15
The probability that a donor/acceptor is occupied by an elec-
tron is determined by Fermi statistics
1
nD = ND0 = ND (29)
1 + eβ(ED −EF )
1
pA = NA0 = NA(1 − f (EA )) = NA (30)
1 + eβ(EF −EA)
To provide a solvable example, imagine that we have an n-type
semiconductor (no p-type dopants) so that NA = NA0 = NA+ =
0, then
C −β(EC −EF )
n = Nef fe (31)
ND = ND0 + ND+ (32)
1
ND0 = ND (33)
eβ(ED −EF ) + 1
Furthermore, charge neutrality requires that

n = p + ND+ (34)

An excellent approximation is to assume that for a (commer-


cially) doped semiconductor

ND+ À ni (35)

ie., many more carriers are provided by dopants than are ther-
mally excited over the entire gap, then as np = n2i , it must be
16
that ND+ À p so that

n ≈ ND+ = ND − ND0 (36)


 
1
n ≈ ND 1 − β(E −E )  (37)
e D F +1
If we recall that thermally induced carriers satisfy the Boltz-
mann equation,
C β(EF −EC )
n = Nef fe (38)

we can eliminate EF in n (where Ed = Ec − ED )


ND
n= C ) (39)
1 + eβEd n/(Nef f
This quadratic equation has only one meaningful solution
2ND
n= s µ ¶ (40)
1+ 1+4 C
ND /Nef eβEd
f

Ed
At low T ¿ kB
r
n' C e−βEd
ND Nef (41)
f

Ed
at higher T À kB

n = ND (42)

At still higher T our approximation breaks down that ND+ À n


since thermally excited carriers will dominate.
17

S-ar putea să vă placă și