You are on page 1of 34

1

Solid State Physics Semiclassical motion in a magnetic field 16


Lecture notes by Quantization of the cyclotron orbit: Landau levels 16
Michael Hilke Magneto-oscillations 17
McGill University
(v. 10/25/2006) Phonons: lattice vibrations 17
Mono-atomic phonon dispersion in 1D 17
Optical branch 18
Experimental determination of the phonon
Contents dispersion 18
Origin of the elastic constant 19
Introduction 2 Quantum case 19

The Theory of Everything 3 Transport (Boltzmann theory) 21


Relaxation time approximation 21
H2 O - An example 3 Case 1: F~ = −eE~ 21
Diffusion model of transport (Drude) 22
Binding 3 Case 2: Thermal inequilibrium 22
Van der Waals attraction 3 Physical quantities 23
Derivation of Van der Waals 3
Repulsion 3 Semiconductors 24
Crystals 3 Band Structure 24
Ionic crystals 4 Electron and hole densities in intrinsic (undoped)
Quantum mechanics as a bonder 4 semiconductors 25
Hydrogen-like bonding 4 Doped Semiconductors 26
Covalent bonding 5 Carrier Densities in Doped semiconductor 27
Metals 5 Metal-Insulator transition 27
Binding summary 5 In practice 28
p-n junction 28
Structure 6
Illustrations 6 One dimensional conductance 29
Summary 6
More than one channel, the quantum point
Scattering 6 contact 29
Scattering theory of everything 7
1D scattering pattern 7 Quantum Hall effect 30
Point-like scatterers on a Bravais lattice in 3D 7
General case of a Bravais lattice with basis 8 superconductivity 30
Example: the structure factor of a BCC lattice 8 BCS theory 31
Bragg’s law 9
Summary of scattering 9

Properties of Solids and liquids 10


single electron approximation 10
Properties of the free electron model 10
Periodic potentials 11
Kronig-Penney model 11
Tight binging approximation 12
Combining Bloch’s theorem with the tight binding
approximation 13
Weak potential approximation 14
Localization 14
Electronic properties due to periodic potential 15
Density of states 15
Average velocity 15
Response to an external field and existence of holes
and electrons 15
Bloch oscillations 16
2

INTRODUCTION derived based on a periodic lattice. However it appeared


that most of the properties are very similar independent
What is Solid State Physics? of the presence or not of a lattice. But there are many
Typically properties related to crystals, i.e., period- exceptions, for example, localization due to disorder or
icity. the disappearance of the periodicity.
What is Condensed Matter Physics?
Properties related to solids and liquids including crys- Even in crystals, liquid-like properties can arise, such
tals. For example: liquids, polymers, carbon nanotubes, as a Fermi liquid, which is an interacting electron system.
rubber,... The same is also true the other way around since liquids
Historically, SSP was considers as he basis for the un- can form liquid crystals.
derstanding of solids since many of the properties were Hence, the study of SSP and CMP are strongly interre-
lated and can not be separated. These inter-correlations
are illustrated below.

The Big Bang of Condensed Matter Physics

Crystal
Crystal

Supraconductivity

Opto-effects

Magnetism
Metals Insulators Semiconductors
properties
3

THE THEORY OF EVERYTHING

See ppt notes

H2 O - AN EXAMPLE

See ppt notes

BINDING

What holds atoms together and also keeps them from


collapsing? FIG. 1: Graphical representation of a H2 molecule
We will start with the simplest molecule: H2 to ask
what holds it together.
The total Hamiltonian can then be written as the sum
between the non interacting H atoms and the cross terms
Van der Waals attraction due to Coulomb interactions. Hence,

The Van der Waals force arises simply from the change H = (H0 ) + (Hint ) (1)
in energy due to the cross Coulomb interactions between
atom a and b, which is simply a dipole-dipole interaction. with

µ ¶ µ ¶
~2 ¡ 2 2
¢ e2 e2 2 1 1 1 1
H= − ∇1 + ∇2 − − +e + − − (2)
2m r1a r2b rab r12 r1b r2a

Using standard perturbation theory it is then possible which leads to a total potential of the form
to evaluate the gain in energy due to Hint . This is left µ 6 ¶
as an exercise. The result is σ σ 12
φ(r) = −4² − , (4)
r6 r12
−αa αb
∆E ' 6 , (3) which is usually referred to as the Lennard-Jones po-
rab
tential. The choice of the repulsive term is somewhat
arbitrary but it reflects the short range nature of the
where αx are the atomic polarizabilities.
interaction and represents a good approximation to the
full problem. The parameters ² and σ depend on the
molecule.
Derivation of Van der Waals

Crystals
Problem 1
What about crystals? Let’s first think about what kind
of energy scales are involved in the problem. If we assume
Repulsion that the typical distance between atoms is of the order
of 1Å we have
We just saw that there is an attractive potential of
e2 /1Å ' 14.4eV for the Coulomb energy and (5)
the form −1/r6 . If there were only a Coulomb repul-
sion of strength 1/r this would lead to the collapse of our µ ¶2
molecule. In fact there is a very strong repulsion, which 1
~2 ' 3.8eV for the potential in a 1Å quantum box
comes from the Pauli principle. For the general pur- 1Å
pose this repulsive potential is often taken to be ∼ 1/r12 , (6)
4

In comparison to room temperature (300K' 25meV) the ground state energy is then given by E 0 =-13.6 eV.
these energies are huge. Hence ionic Crystals like NaCl What happens if we add one proton or H + to the system
are extremely stable, with binding energies of the order which is R away. The potential energy for the electron is
of 1eV. then
e2 e2
U (r) = − − (10)
Ionic crystals r |r − R|
The lowest eigenfunction with eigenvalue -13.6eV is
Some solids or crystals are mainly held together by the µ ¶3/2
electrostatic potential and they include the alkali-halides 1 1
ξ(r) = √ e−r/a0 , (11)
like (N aCl −→ N a+ Cl− ). π a0
where a0 is the Bohr radius. But now we have two
d protons. If the protons were infinitely apart then the
general solution to potential 10 is a linear superposi-
tion ,i.e., ψ(r) = αξ(r) + βξ(r − R) with a degener-
ate lowest eigenvalue of E 0 = E 0 =-13.6 eV. When R
is not infinite, the two eigenfunctions corresponding to
the lowest eigenvalues Eb and Ea can be approximated
by ψb = ξ(r) + ξ(r − R) and ψa = ξ(r) − ξ(r − R). See
figure 3.
- +

Ψb(r)

FIG. 2: A simple ionic crystal such as N aCl


V(r) Ψa(r)
0
The energy per ion pair is
Energy e2 C 14.4eV C
= −α + n = −α + with 6 < n ≤ 12, R
Nionpair d d [d/Å] dn
(7) + +
where α is the Madelung constant and can be calcu- 0 r
lated from the crystal structure. C can be extracted ex-
perimentally from the minimum in the potential energy
and typically n = 12 is often used to model the effect
of the Pauli principle. Hence, from the derivative of the FIG. 3: The potential for two protons with the bonding and
potential we obtain: anti-binding wave function of the electron
µ ¶1/11
12C
d0 = (8) The average energy (or expectation value of Eb ) is
e2 α
so that Eb = hψb∗ Hψb i/hψb∗ ψb i (12)
A+B
Energy 11αe2 = E0 − with (13)
= (9) Z 1+∆
Nionpair 12d0
A = e2 drξ 2 (r)/|r − R| (14)
How good is this model? See table below: Z
B = e2 drξ(r)ξ(|r − R|)/r (15)
Quantum mechanics as a bonder Z
∆ = drξ(r)ξ(|r − R|), (16)
Hydrogen-like bonding R
where h·i ≡ ·dr and similarly,
Let’s start with one H atom. We fix the proton at A−B
r = o then we know form basic quantum mechanics that Ea = E 0 − (17)
1−∆
5

Hence, the total energy for state ψb is now


This figure illustrates why they are called bonding and
Ebtotal = Eb + e2 /R (18) anti-bonding, since in the bonding case the energy is low-
ered when the distance between protons is reduced as
When plugging in the numbers Ebtotal has a minimum at
long as R > R0 .
1.5Å. See figure 4

-8 Covalent bonding

tot
Covalent bonds are very similar to Hydrogen bonds,
Ea only that we have to extend the problem to a linear com-
Antibonding
E [eV]

-12 bination of atomic orbitals for every atom. N atoms


Ea would lead to N levels, in which the ground state has a
tot -13.6 eV bonding wave-function.
Eb

-16 Eb Bonding
R0
Metals
R [distance between protons]
In metals bonding is a combination of the effects dis-
FIG. 4: The energies as a function of the distance between cussed above. The idea is to consider a cloud of electrons
them for the bonding and anti-binding wave functions of the only weakly bound to the atomic lattice. The total elec-
electron trostatic energy can then be written as

Z X X Z
e2 e2 e2 n(r1 )n(r2 )
Eel = − drn(r) + 0
+ 1/2 · dr1 dr2 (19)
r−R 0
R−R |r1 − r2 |
R R>R

which corresponds to an ionic contribution of the form This last expressions leads to a minimum at rs /a0 = 1.6.
µ ¶1/3 We can now compare this with experimental values and
αe2 3 the result is off by a factor between 2 and 6. What went
Eel = − where rs = (20)
2rs 4πn wrong. Well, we treated the problem on a semiclassi-
cal level, without incorporating all the electron-electron
and α is the Madelung constant. Deriving this requires
interactions in a quantum theory. This is very very diffi-
quite a bit of effort. On top of this one has to add the
cult, but in the large density case this can be estimated
kinetic energy of the electrons, which is of the form:
and a better agreement with experiments is obtained.
µ ¶2/3
9π 3~2
Ekin = (21)
4 10mrs2
Binding summary
And finally we have to add the exchange energy, with is
a consequence of the Pauli principle. The expression for There are essentially three effects which contribute to
this term is given by the binding of solids:
µ ¶1/3
9π 3 • Van der Waals (a dipole-dipole like interaction)
Eex = − (22)
4 4πrs
• ionic (Coulomb attraction between ions)
Putting all this together we obtain in units of the Bohr
radius: • Quantum mechanics (overlap of the wave-function)
µ ¶
24.35a0 30.1a20 12.5a0 In addition we have two effects which prevent the collapse
E= − + − eV/atom (23)
rS rS2 rs of the solids:
6

• Coulomb Miller indices can be obtained through the construc-


tion illustrated in the figure below:
• Quantum mechanics (Pauli)

STRUCTURE

FIG. 6: Plane intercepts the axes at (3aˆ1 , 2aˆ2 , 2aˆ3 ). The in-
verse of these numbers are (1/3, 1/2, 1/2), hence the smallest
integers having the same ratio are 2,3,3, i.e., the Miller in-
dices are (233). For a negative intercept the convention is 1̄.
FIG. 5: Pyrite, F eS2 crystal with cubic symmetry.
(Picture from Ashcroft and Mermin)

Illustrations SCATTERING

See ppt notes. See supplement on diffraction.

In order to determine the structure of a crystal it


Summary
is possible to observe the interference pattern produced
by scattering particles with wave-length comparable to
Periodic solids can be classified into two main classes: the lattice spacing, i.e., of the order if 1Å. There are
~ can be three main classes of particles, which can be used:
• Bravais lattices: Every point of the lattice R
reached from a a linear combination of the primitive • Photons, in particular X-rays, whose wave lengths
vectors: R~ = n1 aˆ1 + n2 aˆ2 + n3 aˆ3 , where ni are are around 1Å. The probability to scatter off the
integers. crystal is not that large, hence they can penetrate
quite deeply into the crystal. The main scattering
• Lattices with basis: Here every point in a primitive
~ i so that any occurs with the electrons. Hence, what is really
cell is described by a basis vector B
~ = observed with X-rays is the periodic distribution of
point of the lattice can be reached through: R
~ electrons.
n1 aˆ1 + n2 aˆ2 + n3 aˆ3 +Bi .
• Neutrons are also extensively used since they
In 3D there are 14 Bravais lattices and 230 symmetry mainly interact with the nuclei and the magnetic
groups for lattices with basis. In 2D there are 5 Bravais moments.
lattices and in 1D only 1.
This is the zoology of crystals and they all have names. • Electrons, have a very scattering probability with
It is important to remember that many of the physical anything in the crystal, hence they do not penetrate
properties cannot be deduced from the crystal structure very deep, but they are therefore an interesting tool
directly. The same crystal structure could be a metal to probe the surface structure.
(Cu) or an insulator (Ca).
All forms of scattering share very similar basic prin-
ciples and can be applied to the scattering’s theory of
everything described in the next section.
7

Scattering theory of everything Physics is hidden in n(r), which contains the informa-
tion on the position of scatterers and their individual
scattering distribution and probability. In the following
we discuss the most important implications.
dV
← crystal
k r 1D scattering pattern
Incident
k’ Outgoing
beam ϕ
beam

O Let’s suppose we have a 1D crystal along direction ŷ


~
e ik '⋅r with lattice spacing a and that the incoming wave eikin ·~r
e ik⋅r
is perpendicular to the crystal along x̂. We want to cal-
~ . By
culate the scattered amplitude along direction kout
The total scattered wave off V is

A(q ) ∝ ∫∫∫ dr 3 n (r ) ⋅e [iq⋅r ]


~ − k~in , we can write the scattering am-
defining ~q = kout
plitude as
V Z
q = k - k' A(~q) = n(~r)ei~q·~r dr3 (24)
V
where n( r) is the distribution of scatterers
If we assume that the crystal is composed of point-like
scatterers we can write:
N −1
FIG. 7: Diffraction set-up (picture from G. Frossati)
C X
n(x, y, z) = δ(y − an)δ(x)δ(z), (25)
N n=0

This formulation can describe any scattering process where C is simply a constant. Hence, by inserting 25
in terms of the scattering amplitude A(~q). The entire into 24, and defining ~q = (0, q, 0) we have

N −1 ½
C X iqan C 1 − eiaqN C when q = 2πm/a
A(q) = e = =N →∞ = Cδq,G
~ (26)
N n=0 N 1 − eiaq 0 when q 6= 2πm/a

G~ = ŷ2πm/a is the reciprocal lattice and m an integer. and


If we had a screen along ŷ we would see a diffraction
pattern along I(q) = |A(q)|2 , since what is measured N −1 ½
C X i~q·R~n C when ~q ∈ G~
experimentally is the intensity. In this simple case the A(~q) = e =N →∞ ~ = Cδq~,G
~
N n=0 0 when ~q ∈
/G
reciprocal lattice is the same as the real lattice, but with
lattice spacing 2π/a instead. (28)
In this case the Bravais lattice (or Real-Space) is

R~n = n1 a~1 + n2 a~2 + n3 a~3 (29)

and the reciprocal space G~m (or k-space) can be deduced


Point-like scatterers on a Bravais lattice in 3D ~ ~
from eiRn ·Gm = 1, hence

We start again with the general form of A(~q) from 24 G~m = m1 b~1 + m2 b~2 + m3 b~3 , (30)
and assume that our 3D crystal is formed by point-like
~ Hence,
scatterers on a Bravais lattice R.
where

2π a~2 × a~3 2π a~3 × a~1 2π a~1 × a~2


N −1
C X 3 b~1 = , b~2 = and b~3 =
n(~r) = δ (~r − R~n ), (27) |a~1 · a~2 × a~3 | |a~1 · a~2 × a~3 | |a~1 · a~2 × a~3 |
N n=0 (31)
8

As a simplification we suppose that f1 = f2 = f , then

S(~q) = (1 + e(ia/2)~q·(x̂+ŷ+ẑ) )fe(~q), (37)

where fe(~q) is the Fourier transform of f . Hence,


½
e 2δq~,G
~ if q1 + q2 + q3 is even
A(~q) = S(~q) · δq~,G
~ = f (~
q )
0 if q1 + q2 + q3 is odd
General case of a Bravais lattice with basis (38)

The most general form for n(~r), when the atoms sits
on the basis {u~j } along the Bravais lattice R~n is
M,N
X−1
C
n(~r) = fj (~r − R~n − u~j ), (32)
N j=1,n=0

where fj is the scattering amplitude for the atoms on


site u~j and is typically proportional to the number of
electrons centered on u~j . Now
M,N
X−1 Z
C
A(~q) = fj (~r − R~n − u~j )ei~q·~r d3 r
N j=1,n=0
M,N
X−1 Z
C ~
= fj (~r)ei~q·~r d3 r · ei~q·u~j · ei~q·Rn
N j=1,n=0

with ~r → ~r + R~n + u~j


= CS(~q) · δq~,G
~, (33)

where we have defined the structure factor S(~q) as


M Z
X
S(~q) = fj (~r)ei~q·~r d3 r · ei~q·u~j (34)
j=1

This is the most general form. It is interesting to re-


mark that in most case, experiments measure the inten-
sity I(~q) = |A(~q)|2 , rather than the amplitude.

Example: the structure factor of a BCC lattice

The BCC crystal can be viewed as a cubic crystal with


lattice a and a basis. Therefore, all lattice sites are de-
scribed by

R~n = n1 ax̂ + n2 aŷ + n3 aẑ + ~uj , (35)

where ~u1 = 0(x̂ + ŷ + ẑ) and ~u2 = (a/2)(x̂ + ŷ + ẑ).


The structure factor then becomes:
X Z
S(~q) = fj (~r)ei~q·~r ei~q·u~j d3 r. (36)
j=1,2
9

Bragg’s law nuclear sites. The scattered wave amplitude with wave
number ~k is then
Equivalence between Bragg’s law for Miller planes and Z
~ ~
the reciprocal lattice. A(~k − k~in ) ∼ d3 rn(~r)ei(k−kin )·~r (43)
V

This leads to three typical cases:

• Mono crystal diffraction: Point-like Bragg


peaks
(1,0,0)

(0,1,0) Miller index


in out
q2=k -k2 b
k1out

q1=kin-k1out θ
a
kin
FIG. 8: Point-like Bragg peaks from a single crystal. (Ref:
lassp.cornell.edu/lifshitz)
y z
x
• Powder diffraction:

From Bragg’s law we know that for the planes perpen-


dicular to x̂ or Miller index (1,0,0) the following diffrac-
tion condition applies:

nλ = 2a sin(θ). (39)

hence,


n = 2 sin(θ) · |k~in | = |q~1 |, (40)
a

since k = 2π/λ and we supposed that k~in is perpendicular FIG. 9: Circle-like Bragg peaks from a powder with different
~
to ẑ. This condition is equivalent to q~1 = x̂2π/a ∈ G. grain sizes
The same relation applies for the scattering of the plane
perpendicular to ŷ or Miller index (0,1,0) the following
diffraction condition applies: • Liquid diffraction:
nλ = 2b cos(θ). (41)

hence,


n = 2 cos(θ) · |k~in | = |q~2 |, (42)
b
~
or q~2 = ŷ2π/b ∈ G.

FIG. 10: Pattern evolution for a complex molecule evolving


Summary of scattering from a crystal-like structure to an isotropic liquid

~
We have an incoming wave eikin ·~r diffracting on some
sample with volume V and with a scattering probability The liquid diffraction is essentially the limiting case
n(~r) inside V . For X-ray n is typically given by the elec- of the powder diffraction when the grain size be-
tron distribution whereas for neutrons it is typically the comes comparable to the size of an atom.
10

PROPERTIES OF SOLIDS AND LIQUIDS where the pre-factor 2 comes from the spin degeneracy.
Hence,
The Theory of everything discussed in the first section
kF3 4π 3
can serve as a guideline to illustrate which part of the n3D
e = in D = 3 since VkDF = k (48)
Hamiltonian is important for a given property. For in- 3π 2 3 F
stance, when interested in the mechanical properties, the k2
n2D
e = F in D = 2 since VkDF = πkF2 (49)
terms containing the electrons can be seen as a pertur- 2π
bation. However, when considering thermal conductivity, 2kF
n1D
e = in D = 1 since VkDF = 2kF , (50)
for example, the kinetic terms of the ions and the elec- π
trons are important. where the maximum energy of the electrons is
In the following we will start by considering a few cases
and we will start with the single electron approximation. ~2 kF2
EF = the Fermi energy (51)
2me

single electron approximation


This defines the Fermi energy: it is the highest energy
when all possible states with energy lower than EF are
occupied, which corresponds to the ground state of the
The single electron approximation can be used to de- system. This is one of the most important defi-
rive the energy and density of the electrons. This sim- nitions in condensed matter physics. The electron
plest model will always serve as a reference and in some density can now be rewritten as a function of the Fermi
cases the result is very close to the experimental value. energy, through eqs. (51) and (48) in 3D:
Alkali metals (Li, Na, K,..) are reasonably well described
by this model, when we suppose that the outer shell elec- (2mEF )3/2
tron/atom (there is 1 for Li, Na, K) is free to move inside n(EF ) = . (52)
3π 2 ~3
the metal. This picture leads to the simple example of N
electrons in a box. This box can be viewed as a uniform We now want to define the energy density of states
and positively charge background due to the atomic ions. D(E) as
We suppose that this box with size L×L×L has periodic Z EF

boundary conditions, i.e., n(EF ) ≡ D(E)dE =⇒ D(E) = n(E), (53)
0 ∂E
XN
~2 ∇2n which leads to
H=− with Ψ(0) = Ψ(L) (44)
2me √
n=1 m 2mE
D(E) = in 3D (54)
For one electron the solutions can be written as ~3 π 2
m
D(E) = 2 in 2D (55)
Ψ(~r) = e i~
k·~
r
with E =
~2 |~k|2
and ~k =

(n1 , n2 , n3 ), rπ
~
2me L 2m
D(E) = in 1D (56)
(45) ~2 π 2 E
where ni are the quantum numbers, which are positive
and illustrated below.
or negative integers. Since electrons are Fermions we
cannot have two electrons in the same state, except for 3D

the spin degeneracy. Hence each electron has to have


different quantum numbers. This implies that 2π/L is
D(E)

the minimum difference between two electrons in k-space, 2D

which means that 1 electron uses up 1D

µ ¶D E

(46)
L

of volume in k-space, if D is the dimension of the space. FIG. 11: Density of states in 1D, 2D and 3D
If we now want to compute the electron density (number
of electrons per unit volume ne = N/LD ) in the ground
state, which have k < kF (Fermi sphere with radius kF ),
we obtain:
Properties of the free electron model
µ ¶D
D L
ne = 2 · VkF · /LD , (47)
2π Physical quantities at T=0:
11

hEi
• Average energy per electron: n = 35 EF

• Pressure: P = − ∂(hEiV )
= 2 1 ∂hEi π2 2
∂V 5 nEF , where hEiV is CV = = k T D(EF ) (64)
the total energy. V ∂T µ,V 3
∂P
• Compressibility κ−1 = −V ∂V = 23 nEF Hence, CV
= γ, which is the Sommerfeld parameter.
T
Case 2: T6= 0: In equilibrium
Periodic potentials
1
fF D = , (57)
e(E−µ)/kT + 1 The periodicity of the underlying lattice has important
consequences for many of the properties. We will walk
where the chemical potential is the energy to add one
through a few of them by starting with the simplest case
electron µ = FN +1 − FN . µ = EF at T=0 and F is the
in 1D.
free energy.
The Sommerfeld expansion is valid for kT << µ and
Z ∞ Kronig-Penney model
hHi = dEH(E)FF D (E) (58)
0
Z µ Let us first consider the following simple periodic po-
π2
' H(E)dE + (kT )2 H 0 (µ) + ... (59) tential in 1D.
0 6
Z ∞ X
~2 ∇2
hEi = dEED(E)FF D (E) (60) H=− −V δ(x − na) (65)
0 2m n
Z EF
π2
' ED(E)dE +(kT )2 D(EF ) + ...
(61) The solutions for na < x < na+a are simply plane waves
0 6
and can be written as
hni = n(T = 0) (62)
π 2 0
D (EF ) ψ(x) = An eikx + Bn e−ikx . (66)
⇒ µ(T ) ' EF − (kT )2 + ... (63)
6 D(EF )
Now the task is to use the boundary conditions in order
Physical quantities at T6=0: Specific heat to determine An and Bn . We have two conditions:

(1) ψ(na − ²) = ψ(na + ²) when ² → 0 and (67)


R na+² ~
(2) na−²
(H − E)ψ(x)dx = 0 =⇒ − 2m (ψ 0 (na 0
+ ²) − ψ (na − ²)) = V ψ(na).

Inserting 66 into condition (1) yields

½
An−1 eikna + Bn−1 e−ikna = An eikna + Bn e−ikna = ψ(na) = ψn
(68)
An−1 eikna e−ika + Bn−1 e−ikna eika = ψn−1

½ ¡ −ika ¢
e−ika ψn − ψn−1 = Bn−1 e−ikan
¡ e − e¢ika
=⇒ (69)
eika ψn − ψn−1 = An−1 eikan eika − e−ika

½ ¡ −ika ¢
e−ika ψn+1 − ψn = Bn e−ikan e−ika
¡ e −¢eika
and (70)
eika ψn+1 − ψn = An eikan eika eika − e−ika

When taking the derivative of 66 then condition (2) implies


 

~ 



− ik An eikan −ik Bn e−ikan −ik An−1 eikan +ik Bn−1 e−ikan  = V ψn , (71)
2m  | {z } | {z } | {z } | {z }
ψn+1 −e−ika ψn ψn+1 −eika ψn eika ψn −ψn−1 e−ika ψn −ψn−1
2i sin(ka) −2i sin(ka) 2i sin(ka) −2i sin(ka)
12

hence,

 
2
~ k  −ika 
− 2ψn+1 + 2ψn−1 + ψn (−e − eika − eika − e−ika ) = V ψn (72)
2m 2 sin(ka) | {z }
−4 cos(ka)

and finally, This equation is often called the tight binding equation
µ ¶ as we will see in the next section. Finding a solution for
2m sin(ka) this equation is trivial since
ψn+1 + ψn−1 = − V + 2 cos(ka) ψn , (73)
~2 k
| {z }
W ψn = eipan (75)
Which can be rewritten as
is a solution. This is easily verified by plugging (75) into
ψn+1 + ψn−1 = W ψn (74) (74), which yields

eipan eipa + eipan e−ipa = W eipan =⇒ W = eipa + e−ipa = 2 cos(pa). (76)

This equation has a solution only if Tight binging approximation

−2 ≤ W ≤ 2 (77)
Let us consider the general potential due to the ar-
We now recall that the Eigenstate of the original Hamil- rangement of the atoms on a lattice:
2 2
tonian is given by E(k) = ~2m k
, where eikx is the plane
~2 ∇2 X
wave between two δ functions, hence the dispersion re- H=− + V0 (~r − R~n ) (78)
2 2 2m
lation is equal to E(k) = ~2mk
as long as −2 ≤ W ≤ 2. |
n
{z }
This condition will create gaps inside the spectrum as V (~
r)
illustrated in the graph below:
This is a very general form for a periodic potential assum-
ing that we only have one type of atoms. The periodicity
is given by the lattice index R~n . For a general periodic
2

∆E potential Bloch’s theorem (see A& M for proof) tells us


<-
W

that a solution to this Hamiltonian can be written as


E(k)

<2
<W
-2

~
-2
ψk (~r) = eik·~r uk (~r), (79)
W<
0 1 2 3

π/a where uk is a periodic function such that uk (~r + R~n ) =


uk (~r). This implies that

~ ~
FIG. 12: Dispersion curve for the Kronig-Penney model ψk (~r + R~n ) = eik·Rn ψk (~r) (80)

Let us now suppose that the solution of the Hamiltonian


2 2
Conclusion: a periodic potential creates gaps, H1atom = − ~2m ∇
+V0 (~r − R~n ) with one atom is φl (~r − R~n )
which leads to the formation of a band structure. corresponding to the energy level ²l , i.e., H1atom φl = ²l φl .
This is a very general statement which is true for almost Now comes the important assumption, which allows us
any periodic potential. to simplify the problem: We suppose that
13

 
X
hφl (r − Rm )|H|φl (r − Rn )i = −tl  δm,~
~ n+~i + δm,~
~ n−~i
 + ²l δm,~
~ n (81)
~i=x̂,ŷ,ẑ

Hence, only the nearest neighbor in every direction is Combining Bloch’s theorem with the tight binding
taken to be non-zero. Further, we assume that there is no approximation
overlap between levels of the one atom potential, which
allows us to look for a general solution of the following The tight binding approximation is very general and
form for each energy level ²l . can be applied to almost any system, including non-
periodic ones, where the tight binding elements can be
X
ψl (r) = c~ln φl (r − R~n ) (82) assembled in an infinite matrix. For the periodic case,
~
n
on the other hand, it is possible to describe the system
with a finite matrix in order to obtain the full disper-
with eigenvalue El (k). To calculate El (k) we plug-in
sion relation. This is obtained by combining the Bloch
this Ansatz into eq. (82) and obtain an equation for
theorem for periodic potentials, where the wave-function
the coefficients c~ln , which leads to the following equation
from (79) is again:
when using the tight binding approximation given in eq.
(81). ~
ψk (~r) = eik·~r uk (~r)
X
clm
~ ²l − tl
l
cm+
~ ~i
= El (k)clm
~ (83) Instead of writing (82) we write the Bloch-tight-
~i=±x̂,ŷ,ẑ binding solution as
The solution of this equation are plane waves, which X ~
i~
k·R~m ψlk (r) = eik·R~n c~ln φl (r − R~n ),
can be verified readily by taking clm
~ = e and plug-
~
n
ging it into the equation to obtain
which now depends explicitly on the wavevector ~k. Using
El (k) = ²l − tl (2 cos(kx ax ) + 2 cos(ky ay ) + 2cos(kz az ), Bloch’s theorem this implies that
(84)
where ~a is the lattice constant in all 3 space directions: c~ln = cm
l
~,
~a = Rm+â
~ − Rm ~ . The energy diagram is illustrated in
fig. . The degeneracy of each original single atomic en- whenever ~n and m~ are related by a linear combination of
ergy level ²l is lifted by the coupling to the neighboring Bravais vectors. Moreover, the tight binding equation in
~
atoms and leads to a dispersion curve or electronic band (83) is the same but with c~ln replaced by c~ln eik·R~n
structure.

E
a

ε3 B

ε2

ε1
-1.0 -0.5 0.0 0.5 1.0

−π/a k π/a FIG. 14: Diatomic square crystal

FIG. 13: Dispersion curve for the tight binding model We apply this to the simple example of a diatomic
square lattice of lattice constant a with alternating atoms
A and B shown in figure . We will further assume that
This tight binding approximation is very successful we have only one band l. Hence, the Bloch-tight-binding
in describing the electrons which are strongly bound to solution is written as
the atoms. In the opposite limit where the electrons or
X ~
more plane-wave like, the weak potential approximation ψ k (r) = eik·R~n c~n φ(r − R~n ),
is more accurate: ~
n
14

~ 2 2
where c~n takes on only to possible values due to Bloch’s ψk0 (~r) = eik·~r , with corresponding energies ²0k = ~2mk
and
theorem: cA or cB . This leads to the following simplified 0 0 0
Hamiltonian H0 , i.e., H0 ψk (~r) = ²k ψk (~r). Since the ori-
tight binding equations (assuming hφl (r − Rn )|H|φl (r − gin of this energy dispersion relation can be chosen from
Rn )i = ²A or ²B and t = −hφl (r − Rn )|H|φl (r − Rm )i any site of the reciprocal lattice, we have ²0k+K = ²0k ,
when n and m are nearest neighbors): hence these energies are degenerate. This implies that
P we have to use a degenerate perturbation theory. The
~~
cA ²A − t ~ cB eiak·i = E(~k)cA mathematical procedure is very similar to the tight bind-
Pi=±x̂,ŷ ~~
cB ²B − t ~i=±x̂,ŷ cA eiak·i = E(~k)cB ing approximation, but we now expand the solution ψ(~r)
of the full Hamiltonian H = H0 + V (~r) in terms of a sum
It is now quite straightforward to rewrite these equa- of plane waves ψk0 (~r). Since ²0k+K = ²0k we will only use
tions in matrix form: two plane waves in this expansion: ψk0 (~r) and ψk+K 0
(~r).
µ ¶µ ¶ µ ¶ Hence,
²A −t · g cA ~ cA
= E(k) ,
−t · g ∗ ²B cB cB
ψ(~r) = αψk0 (~r) + βψk+K
0
(~r), (85)
ia~
k·x̂ −ia~
k·x̂ ia~
k·ŷ −ia~
k·ŷ
with g = e +e +e +e . The disper-
sion relation or band structure is then simply given by where the coefficients α and β have to be determined in
obtaining the eigenvalues of HBT B , where order to solve Schrödinger’s equation:
µ ¶
²A −t · g
HBT B = . (H − E)ψ(~r) = 0. (86)
−t · g ∗ ²B

We can find the solution by first multiplying (86) by


Weak potential approximation (ψk0 )∗ (~r) and then integrating the equation over the whole
space which will lead to one equation, and we obtain a
0
In this case we consider the effect of the periodic po- second equation by multiplying (86) by (ψk+K )∗ (~r) and
tential V (~r) as a perturbation on the plane wave solution then integrating of the whole space. This leads to

½ R R
R α d3 r(ψk0 )∗ (~r)(H − E)ψk0 (~r) +R β d3 r(ψk0 )∗ (~r)(H − E)ψk+K
0
(~r) = 0
3 0 ∗ 0 3 0 ∗ 0 , (87)
α d r(ψk+K ) (~r)(H − E)ψk (~r) + β d r(ψk+K ) (~r)(H − E)ψk+K (~r) = 0

where (assuming normalized plane waves in the integrals)


R 3 R

 R d r(ψk0 )∗ (~r)(H − E)ψk0 (~r) = dR3 re−ikr (H0 + V (r) − E)eikr = ²0k + 0 −R E
 3 0 ∗ 0
R d3 r(ψk0 ) (~r)(H − E)ψk+K (~r) = d3 re−ikr (H0 + V (r) − E)ei(k+K)r = d3 reiKr (²0k+K + V (r) − E) = VK
∗ 0

 d r(ψ ) (~
r )(H − E)ψk+K (~ r) = ²0k+K − E
R 3 k+K
0
d r(ψk+K )∗ (~r)(H − E)ψk0 (~r) = V−K ,
(88)
R 3 iK·~~ R R
and VK ~ = d re r V (~r) (the Fourier transform), d3 reiKr = 0 (for K 6= 0), and d3 rV (~r) = 0. With coefficients
(88), equation (87) leads to the following couple of equations:
½ ¯ 0 ¯ q 0 0
α(²0k − E) + βVK = 0 ¯ ² −E VK ¯ 0 0 2
⇒¯¯ k ¯ = 0 ⇒ E = ²k +²k+K ± (²k −²k+K ) + VK V−K . (89)
0
αV−K + β(²k+K − E) = 0 V−K ²k+K − E ¯
0 2 4

Finally if ²0k = ²0k+K , we have E = ²0k ± |VK |, which the wavevector which corresponds to the two dispersion
leads to a splitting 2|VK | of the energy levels at these curves which led to the degenerate energy level.
degenerate energies. For the example in figure , this
weak potential approximation would give us a splitting of
2|VK=2π/a | at k = −π/a and k = K − π/a = π/a. This Localization
implies that the first order calculation of the energy split-
ting due to the weak periodic potential V (r) is equal to
When, instead of having a purely periodic potential
twice the fourier transform of this potential evaluated at
disorder is included into the system, we no more have
15

Bloch wave solutions but localization of the wave func- equation Hψ = E(k)ψ it follows that uk is a solution of
tions occur. This is particularly important in low dimen- µ 2 ¶
sional systems and tends to suppress transport. ~
(k − i∇)2 + V − E(k) uk = 0. (93)
2m
| {z }
Hk −E(k)
Electronic properties due to periodic potential
In order to calculate hvi we use first a first order pertur-
Density of states bation in k + q, where q is very small. Therefore, the
eigenvalue corresponding to Hk+q is E(k + q), which is
to first order
Density of states in 1D:
E(k + q) = E(k) + huk |Hk+q − Hk |uk i
∂n δn ~2
D(E) = ' = E(k) + hψeikr | ( q 2 +2q(k − i∇))|e−ikr ψi
∂E δE 2m |{z}
¯ ¯−1 →0
δn ¯¯ ∂E(k) ¯¯
= × |{z}
2 × |{z}
2 ~2
δk ¯ ∂k ¯ P = E(k) + hψeikr | q(k − i∇))|e−ikr ψi
spin ±k
m
¯ ¯−1 ~
1 δN ¯¯ ∂E(k) ¯¯ = E(k) + hψ|q~ (−i∇)|ψi
= ×4 m
L δk ¯ ∂k ¯ = E(k) + q~hψ|v|ψi
¯ ¯−1
2 ¯¯ ∂E(k) ¯¯ = E(k) + q∇k E(k) (Taylor expansion)
D(E) = , (90)
π ¯ ∂k ¯ 1
⇒ hvi = ∇k E(k). (94)
~
where we used that δE = (∂E/∂k)δk and δk = 2π/L for
one electron, i.e., δN = 1. Hence, the expectation value of the velocity is determined
In three dimensions (D=3) we have: by the slope of the dispersion relation. This also implies
that the sign of the average velocity depends on the sign
of ∂E/∂k.
∂n δn
D(E) = '
∂E δE
X δn −1
Response to an external field and existence of holes and
= |∇k E(k)| × |{z}2 electrons
δk
E(k)=E spin
Z The idea is to describe the average motion of an elec-
2 −1
= d2 k |∇k E(k)| , (91) tron in the presence of an external field (electric, Eel , or
(2π)3 E(k)=E
magnetic, B) in a semiclassical way. Hence, we want
P 2
where we used δE = (∇k E(k)) · δk, E(k)=E (δk) → d
R 2 3 3 m∗ hvi = F = (qEel or qhvi × B), (95)
E(k)=E
d k, and δn/(δk) = 1/(2π) . This result shows dt
that the density of state in the presence of a periodic
potential, i.e., for a crystal depends only on the slope of where m∗ is an affective mass and q the charge. Using
the dispersion relation or the band structure. (95) and (94) we have

∂hvi
m∗ k̇ = qEel
Average velocity ∂k
1 ∗ ∂2E
m k̇ = qEel
The average velocity of an electron in a periodic po- ~ ∂k 2
tential is given by the expectation value of the velocity, ⇒ ~k̇ = qEel = F, (96)
i.e.,
Where we defined the effective mass m∗ as
hvi = hψ|v|ψi, (92) ¯ 2 ¯−1

¯ ¯
2 ¯∂ E ¯
m =~ ¯ 2¯ . (97)
where ψ is the wavefunction from the Hamiltonian with ∂k
2 2
periodic potential V , i.e., H = ~2m

+ V . From Bloch’s 2
theorem (79) we can write ψ(r) = eikr uk (r), where uk (r) If ∂∂kE2 is negative we need to change the sign of q in order
2
has the same periodicity as V (r). From the Schrödinger to remain consistent. Hence, when ∂∂kE2 > 0, the charge
16
2
of an electron is q = −e but if ∂∂kE2 < 0 then q is positive Using (101) and (94) and since k̇ is perpendicular to hvi
(+e). In this case we describe the particles as holes. ˙ ∼ ∂E/∂kk we obtain
and B, (k)
They represent missing electrons. With these definitions
Z k2
of q and m∗ , which is also called the band mass, the ~2 dk⊥
semiclassical equations of motion of single electrons in a δt =
qB k1 ∂E/∂kk
periodic potential are simply given by eqs. (95) and (96). 2 Z k2
~ d
In general, the effective mass is given by a tensor de- = kk dk⊥ . (103)
fined as qB dE k1
¯ 2 ¯−1

¯
2¯ ∂ E ¯
¯ For a complete turn this leads to
mαβ = ~ ¯ , (98)
∂kα ∂kβ ¯ I
~2 d
where α and β are the spatial directions. An important T = kk dk⊥ . (104)
qB dE
consequence of this semiclassical description of the mo- | {z }
S
tion of electrons is the dependence of the effective mass
on the energy and the band structure. In some cases the Here S is the area enclosed by an orbit in k-space. This
2
effective mass can even diverge (when ∂∂kE2 = 0). Sim- orbit corresponds to an equipotential line perpendicular
ilarly the sign of the carriers also depends on the band to the magnetic field.
structure and the energy of the carriers. By definition we Let’s suppose for simplicity that the effective mass ten-
call the bottom of an energy band and electronic band sor m∗ is diagonal and given by
2
when ∂∂kE2 > 0 and a hole band when at the top of the  
2 mx 0 0
energy band ∂∂kE2 < 0.
m∗ =  0 my 0  (105)
0 0 mz
Bloch oscillations
and that the energy dispersion is harmonic (which is usu-
ally true at a band extremum, i.e.,
In the presence of an electric field and a periodic po-
tential we can use the equation of motion (96), i.e., ~2 kx2 ~2 ky2 ~2 kz2
E(k) = + + . (106)
eEel 2mx 2my 2mz
~k̇ = −eEel ⇒ k = − t, (99)
~
If we assume that the magnetic field is along z and that
but in a periodic potential and in the tight binding ap- the average effective mass perpendicular to B is given by
proximation the energy is given by E(k) = −2t0 cos(ka), m⊥ , we can rewrite (106) as
where a is the lattice constant and t0 the nearest neighbor
overlap integral. Hence, since v = ṙ and hvi = ~−1 ∂E/∂k ~2 k⊥
2 ~2 kk2
we have E(k) = + , (107)
2m⊥ 2mk
2t0 aeEel t
hri = cos( ). (100) 2
eEel ~ where k⊥ = kx2 +ky2 . Using (104) and (107 we then obtain
This means that the average position of the electrons 2
S = πk⊥ = π(2m⊥ E/~2 ) − πkk2 m⊥ /mk
oscillates in time (Bloch oscillations). In artificial struc-
tures these Bloch oscillations are typically of the order of dS π2m⊥
⇒ =
1THz. dE ~2
2π qB
⇒ ωc = = , (108)
T m⊥
Semiclassical motion in a magnetic field
which is the cyclotron frequency. Hence, the cyclotron
In the presence of a magnetic field (B), we can describe frequency depends on the average effective mass perpen-
the semiclassical trajectories in k-space using (96), i.e., dicular to the magnetic field. This allows us to measure
the effective mass along different directions, simply by
~k̇ = qhvi × B. (101) changing the direction of the magnetic field and by mea-
Hence, only the values of k perpendicular to the magnetic suring the cyclotron frequency.
field will change, which we denote by k⊥ . The component
parallel to the field, kk is not affected by B. During a Quantization of the cyclotron orbit: Landau levels
small time difference
Z t2 Z k2 =k(t2 )
In quantum mechanics the energies of these cyclotron
δt = t2 − t1 = dt = dk⊥ /|k̇|. (102)
t1 k1 =k(t1 ) orbits become quantized. To see this we can write the
17
500
Hamiltonian of an electron in a magnetic field in the har- 450 300

Rxx [Ω]
200
monic approximation (107) as 400
100
350
0
300

Rxx [Ω]
1 1
1.1 1.3 1.5 1.8 2.0 2.2 2.4 2.6 2.8

Pk2 + (P⊥ + qA)2 ,


-1
250
H= (109) 1/B [T ]

2mk 2m⊥ 200


150
where B = ∇ × A ⇒ A = −Byx̂ in the Landau Gauge if 100

B is along ẑ. In analogy to the harmonic oscillator, the 50

eigenvalues of (109) are then given by 0


0.0 0.5 1.0
B [T]
~2 kk2 qB
En,kk = + (n + 1/2)~ . (110)
2mk m⊥ FIG. 15: Magneto-resistance oscillations in a GaAs/AlGaAs
|{z}
ωc quantum well.

These eigenvalues can be found by writing


the wavefunction as ψ = eikx x φn (y − y0 )eikk z
with y0 = − ~k
qB ,
x
which leads to Hψ =
µ 2 2 ³ ´2 ¶ PHONONS: LATTICE VIBRATIONS
2
~ kk Py 1 qB
2mk + 2m⊥ + 2 m⊥ m⊥ (y − y0 )2 φn (y − y0 ).
In the y direction this is simply the harmonic oscillator In general:
with eigenvalues (n + 1/2)~ωc and in the direction
parallel to the field we have a plane wave so that the X
M ül = − φlm um , (112)
total energy is given by (110). The quantized levels m
(n + 1/2)~ωc due to the magnetic field are called the
Landau levels.
where ul are the deviations from the original lattice sites
and φlm are
P the elastic constants who have to obey this
Magneto-oscillations sum rule m φlm = 0 (translation invariance). In words,
equ. 112 simply means that the force producing the devi-
ation on lattice site l only depends on the deviations from
The quantization of the energy levels in the presence
the other lattice sites. No deviation=no force (equilib-
of a magnetic field will lead to oscillations of almost any
rium). This equation is very general and does not assume
experimental quantity (resistance, thermal conductivity,
that we have a periodic lattice, but in order to calculate
magnetization, ...) as a function of B. These oscillations
things we will use a periodic lattice and start with 1D.
can be understood by looking at the density of states
In fig. 16 we illustrate two typical displacement waves in
due to the quantization (110). Indeed, (110) leads to a
2D.
peak in the density of states whenever E = (n + 1/2)~ωc .
Therefore, whenever the component of the Fermi energy
u s −2 u s −1 us u s +1 u
perpendicular to the magnetic field is equal to one of s+2

these quantized levels there is an extremum in the quan-


tity measured. The distance between two of these ex-
trema is given by K

qB1 qB2
~ (n1 + 1/2) = EF⊥ = ~ (n2 + 1/2)
m⊥ m⊥
1 1 q
⇒ − =~ ( EF⊥ )−1 (n2 − n1 ) us+4
B1 B2 m⊥ |{z} u s −1 u s u s +1 u s + 2 us+3
~2 K
2πm⊥ ·S
µ ¶
1 2πq −1
⇒ ∆ = S , (111) s −1 s s +1 s+2 s+3 s+4
B ~
H
where the maximum S is S = kk dk⊥ = π(kxF )2 = FIG. 16: Two types of lattice displacement waves in 2D, trans-
~2 (kx
F 2
)
~ S 2
π(kyF )2 and EF⊥ = = 2πm
2m⊥ ⊥
. Hence, the Magneto- verse and longitudinal modes
oscillations are periodic in 1/B and the period depends
on S. An example of these oscillations in resistance is
shown in figure 15 as a function of B and 1/B.
18

Mono-atomic phonon dispersion in 1D 1. Acoustic mode:


s
C
PIn 1D equ. 112 for nearest neighbors and using ω' ka (117)
2(M1 + M2 )
m φlm = 0 simply reduces to

M ül = K(ul+1 − 2ul + ul−1 ) (113) 2. Optical mode:


in the simplest approximation, where only the nearest s
neighbors are important and where the elastic constants 2C(M1 + M2 )
ω' (118)
are the same. Further we assume that the equilibrium M1 M2
case has a lattice constant a. In this case the solution
can be written as

ul = eikla−iωt
2
=⇒ M ω 2 =q2K(1 − cos(ka)) = 4K sin (ka/2) (114)
K
=⇒ ω = 2 M | sin(ka/2)|

which is illustrated below In this case there is only a

1.2
ω Mono-atomic dispersion
1.0
2 (K M)
0.8

0.6

0.4

0.2

0 k
-π/a 0 π/a 2π/a

FIG. 17: Phonon dispersion for a mono-atomic lattice

single dispersion mode, which is called the acoustic mode


or branch.

Optical branch

The situation is quite different when there is an addi-


FIG. 18: Phonon dispersion for a di-atomic lattice and the
tion symmetry breaking in the problem. In the 1D case
more general 3D case
this occurs when there is a second species with a different
mass, for example, or when there are two different elastic
constants. Let us discuss the case where there are two
different masses but only one elastic constant C:
Experimental determination of the phonon
½ dispersion
M1 ül = C(ul+1 − 2ul + ul−1 )
(115)
M2 v̈l = C(vl+1 − 2vl + vl−1 )
In order to determine the dispersion curve of the
To solve this set of equation we use a solution of the phonons there is essentially one trick. The idea is to
form ul = ueikla−iωt and vl = veikla−iωt , which inserted look for inelastic scattering of other particles, which are
into eq. 115leads to typically photons or neutrons. The incoming particles
scatter with phonons, thereby transferring some of their
s p energy and momentum to phonons. From the energy and
√ M1 + M2 ± M12 − 2M1 M2 cos(ka) + M22 momentum of the out-coming particles it is then possible
ω= C
M1 M2 to infer the phonon absorbtion. This process can also be
(116) inverted in that the energy of the out-coming particle in-
The following two limiting cases can easily be obtained creases by absorbtion of a phonon. In general, the energy
in the limit of small k. conservation can be expressed as
19

FIG. 19: Inelastic neutron scattering experiment in Chalk


River used to determine the phonon dispersion
~ωin (~k) − ~ωout
0
(k~0 ) = ±~Ω(K)
~ (119)
| {z } | {z }
in/out particle phonon

and the momentum conservation as

~~k − ~k~0 = ±~K


~ + ~G,
~ (120)

where G~ is a reciprocal lattice vector. A neutron scat-


tering experiment is illustrated below.

Origin of the elastic constant

From the TOE we have the potential term due to the


interactions between ions.

NX NX
ions
~2 ∇2i ions
qi qj
Hions = − + (121)
i
2mi i<j
|ri − rj |

This leads to a potential of the form

! displacment 
X ∂φ z}|{ 1 X ∂2φ 
φ(r1 , . . . , rn ) = φ(r10 , . . . , rn0 ) + ui + ui uj + · · · (122)
| {z } ∂ri 2 i,j ∂ri ∂rj
i ri0 0
ri
Cohesive energy | {z } | {z }
Equilibrium pos. 0 φi,j

The linear term has to be zero for stability reasons (no trated in fig. 18. From 123 we have
minimum in energy otherwise). The classical equation X ~ ~ ~
of motion is then simply given by equ. 112, because mω 2~² = φl,m eik·(Rl −Rm ) ·~² (124)
F~ = −∇φ.
~ To solve this equation in general, one can |
m
{z }
write a solution of the form φ̂(k)

,
where φ̂(k) is a 3 × 3 matrix, since φlm too. This leads
i~ ~l −iωt
k·R
to an eigenvalue equation for ω 2 with three eigenvalues:
u~l = ~² · e , (123) ωL2
(for ~k k ~²) and ωT2 (1,2) (for ~k ⊥ ~²). Hence, we have
on longitudinal mode and two transverse modes and all
phonon modes can be described by a superposition of
where R ~ l are the lattice sites. One then has to solve these.
for the dispersion relation ω(~k) in all directions, as illus- If we have two different masses, we have two addi-
20

tional equations of the form (124). This leads to the


general case, where one has two branches the optical one Z
and the acoustic one and each is divided in longitudinal 1 dSω
D(ω) = (127)
and transverse modes. The acoustic branch has always (2π)3 ω=const |∇k ω|
a dispersion going to zero at k = 0, whereas for the op-
tical mode at k = 0 the energy is non-zero. The acous- Let us now suppose that we have 3 low energy modes
tic phonons have a gapless excitation spectrum and can with linear dispersion, one longitudinal mode ωL (k) =
therefore be created at very low energy as opposed to cL k and two transverse modes ωT (k) R = cT k, hence
optical phonons which require a miniumum energy to be ∇ωL,T (k) = cL,T , which implies that dS = 4πk 2 . This
excited. leads to a density of state

Quantum case k2 ω2
DL,T (ω) = = (128)
2π 2 cL,T 2π 2 c3L,T
While many of the properties, like the dispersion rela-
or
tion, can be explained in classical terms, which are simply
vibrations of the crystal ions, others, such as statistical
properties need a quantum mechanical treatment. µ ¶
ω2 1 2
Phonons can simply be seen as harmonic oscillators Dtot (ω) = + 3 (129)
2π 2 c3 c
which carry no spin (or spin 0). They can, therefore, be | L {z T }
accurately described by bosons with energy 1
c3

in the isotropic case.


En (k) = ~ω(k)(n + 1/2). (125) From statistics we know that the probability to find a
state at E = En is given by the Boltzmann
P distribution
We can now calculate the density of states of phonons,
Pn ∼ e−En /kB T , with normalization Pn = 1. Hence,
or the number of states between ω and ω + dω, i.e.,

Z ω+dω Pn = e−n~ω/kB T (1 − e−~ω/kB T ) (130)


1
D(ω)dω = dk (126)
(2π)3 ω
simply from the normalization condition. We can now
but dω = ∇k ωdk, hence calculate the average energy at ω,

X ∞
X
E(ω) = En Pn = (1 − e−~ω/kB T )~ω (n + 1/2)(e−~ω/kB T )n
n n=0
 
1 1 
 
= ~ω  + ~ω/k T  (131)
2 e B − 1
| {z }
hni

here hni is the expectation value of quantum number n zero energy state is totally degenerate (this does sup-
at T , which is nothing else but the Bose Einstein distri- pose that there are no interactions between bosons, with
bution. As is well known, Bosons obey the Bose-Einstein interactions the delta function will be a little broader).
statistics, where he Bose-Einstein distribution function is The total average energy can then be calculated from the
given by density of states as:

1 Z
fBE = (132)
eE/kB T − 1 hEi = dωD(ω)E(ω) (133)
Important: at T = 0, fBE is simply a delta function
δ(E). This is the Bose-Einstein condensation, where the This also allows us to calculate the specific heat CV =
21

d
dThEi in the isotropic case and in the linear dispersion If the electrons flow without collisions,
approximation D(ω) ∼ ω 2 /c3 , hence (137)=g(r, k, t) = fF D . The situation changes if
collisions (f. ex. between electrons, impurities, or
Z ∞ phonons) are included. In this case
d
CV = dωD(ω)E(ω)
dT 0
Z ∞
d ~ω g(r, k, t) − g(r − ṙdt, k − k̇dt, t − dt) = δgcoll. (138)
= dωD(ω) ~ω/k T
dT 0 e B −1
Z ∞ where gcoll. describes the collisions. gcoll. = 0 without
d 3 1 ~ω 3
= 2 3
dω ~ω/k T collisions. Expanding (137) to first order and using (138)
dT 2π c
|{z} 0 e B −1 we obtain
µ ¶| {z }
1 1 2 R∞ 3 (kB T )4
3 c3
+
c3 0
dω exx−1 ~3
L
Z
T

d 3(kB T )4 ∞
x3 ∂g ∂g ∂g dg
= dω x ṙ + k̇ + = (139)
dT 2π 2 (c~)3 e −1 | ∂r {z∂k ∂t
} dt coll.
0
| {z } dg
π 4 /15 dt

µ ¶3
2π 2 kB kB T This is Boltzmann’s equation.
= (134)
5 ~c

where we had defined the variable x = ~ω/kB T and Relaxation time approximation
where 1/c3 is the average over the 3 acoustic modes.
The Debye model assumes a linear dispersion (ω = ck) Solving Boltzmann’s equation is not easy in general
and uses the analogy to electrons where (n = kF3 /3π 2 ) to and therefore approximations are used. The simplest one
define kD . Since phonons have no spins one obtains: is the relaxation time (τ ) approximation. In this approx-
imation:
3
kD
n = and ¶
6π 2 dg 1
= − (g − fF D )
k B ΘD = ~ωD = ~ckD (ΘD : Debye temperature) dt coll. τ
µ ¶3
12π 4 kB T dg 1
=⇒ CV = ·n (135) =⇒ = − (g − fF D ) (140)
5 ΘD dt τ

The typical Debye temperature is of the order of 100K. It is quite intuitive to see from where this approxima-
Finally, combining this result with the contributions from tion comes from, since a kick at t = 0 would lead to
electrons (64), we obtain the expression valid for low tem- solution of the form
peratures:
e−t/τ
g(t) = + fF D → fF D (when t = ∞) (141)
CV = γT + βT 3 (136) τ
|{z} |{z}
electrons phonons This is the basic framework, which allows us to eval-
uate the effect of external fields on the system and to
estimate the linear response to them.
TRANSPORT (BOLTZMANN THEORY)

~ = −eE
Case 1: F ~
Transport allows us to calculate transport coefficients
such as resistances and thermal conductivities. While
several transport theories exist, Boltzmann’s approach In this case we look for a response in current density
is the most powerful and applies to most situations in and define the conductivity tensor in linear response as
condensed matter. The main idea in Boltzmann theory
is to describe the electrons by a distribution function g.
In equilibrium g is simply the Fermi-Dirac distribution ~
~j = σ E
Z
function fF D . In general, g(r, k, t) and at t − dt it can be d3 k
⇒ jα = σαβ Eβ = −2e vα g(r, k, t). (142)
written as: (2π)3

Since we are interested in the case of a homogenous


g(r − ṙdt, k − k̇dt, t − dt) (137) uniform electric field, ∂g/∂t = 0 and ∂g/∂r = 0 and
22

~k̇ = −eE, Boltzmann’s equation (139) in the relaxation Diffusion model of transport (Drude)
time approximation becomes
In the case where the scattering of electrons is domi-
∂g 1 nated by inelastic diffusion, we can write a very simple
k̇ = − (g − fF D )
∂k τ form for the conductivity or resistivity tensor. Indeed, in
|{z}
−eE~ ∂g ∂² a diffusive regime the average velocity is directly propor-
∂² · ∂~
~ k tional to the external force and to the inverse effective
~
eEτ ∂g ∂² mass. The proportionality coefficient is then simply the
⇒g = · +fF D (143) scattering probability 1/τ . Hence,
~ ∂² ∂~k
|{z} |{z}
∂fF D
'− ∂µ
~~
v
m∗
Here we used the first order approximation ∂g/∂² ' ~v = F~ = −eE
~ − e~v × B
~ (148)
τ
−∂fF D /∂µ, which is justified if the correction g − fF D is
smoother than ∂fF D /∂µ, which is close to a delta func- If we suppose that the magnetic field is small and in
tion when kT << EF . Hence, using (143) to evaluate ~ along x̂, then eq. (148) becomes
the ẑ direction and E
the current (142), we obtain

Z Z ~ = − m ~v −e~v × B
eE ~
d3 k ~ ∂fF D |{z} τ } | {z }
jα = 2e vα eτ E · ~v + ∼ ( vα fF D ), (144) eEx | {z
(2π)3 ∂µ 'jy B/n∗
| {z } ' mτ en
jx

=0
m∗ B
which leads to Ex = 2
jx + jy (149)
ne τ ne
Z
∂jα d3 k ∂fF D
σαβ = = 2e2 3
vα vβ τ (²) (145) Since in general E ~ = ρ~j and because the resistivity
∂Eβ (2π) ∂µ ~
along B is not affected by the magnetic field we can write
This is one of the most important expressions for the ~ in the ẑ direction,
for B
conductivity. Since ∂fF D /∂µ is almost a delta function
for kT << EF this expression shows that only electrons  
m∗ B
close to the chemical potential µ will significantly con- ne2 τ ne 0
m∗
tribute to transport. It is possible to evaluate (145) in ρ= B
− ne ne2 τ 0  = σ −1 (150)

m
simple cases. 0 0 ne2 τ
Let’s assume that τ does not depend on the energy,
then we can rewrite (145) as This is the famous Drude formula in a magnetic field.
This formula is in fact equivalent to the relaxation time
Z µ ¶ approximation in the Boltzmann theory (147).
d3 k ∂² ∂fF D
σαβ = 2e2 τ v α −
(2π)3 ~∂kβ ∂²
Z 3
µ ¶
d k ∂fF D Case 2: Thermal inequilibrium
= e2 τ vα −
(2π)3 ~∂kβ
Z 3 Z We now consider the case, where we also have a spacial
2 d k ∂vα
= e τ fF D /~ − f F D vα gradient. Hence (139) and (140) become
(2π)3 ∂kβ k⊥β
|{z} | {z }
~δαβ /m∗ =0 ∂g ∂g 1
Z ṙ + k̇ = − (g − fF D ), (151)
e τ 2
d k 3 ∂r
|{z} ∂k
|{z} τ
= 2 fF D δαβ (146) ~ v ∂f
m∗ (2π)3 v · ∂f
'~ ∂~
r '−eE·~ ∂²
| {z }
n
here we used again that g − f is smooth so that ∂(g −
• If τ (²) = τ we therefore obtain the most important f )/∂² << ∂f /∂². Moreover, since
formula of conductivity (the Drude formula)
e2 τ n 1
σαβ = δαβ (147) fF D =
m∗ e(²−µ(r))/kT (r)
+1
µ ¶
Using the same approach but in the presence of a mag- ∂f ∂f ∇r T
⇒ = −∇r µ − (² − µ)
netic field, F~ = −eE−e~
~ v × B, ~ we can derive an equivalent ∂r ∂² T
µ ¶
expression for σαβ , but in this case the off-diagonal com- ∂f ~ ∇r T
⇒ g = ~v · τ eG + (² − µ) + fF D ,(152)
ponent is not zero anymore. (See assignment). ∂² T
23

~ = eE+
where we defined the generalized field eG ~ ∇µ(r)
~ Is obtained from eqs. (153,154,157), by setting j´e =
³
and the external forces are now −eE,~ ∇µ(r)
~ ~ (r),
and ∇T 11 ~
0, hence at low temperatures, L G+L 12 ~
∇T
− T =
corresponding to an external electrical field, a gradient 0 and
in the chemical potential (f.ex. a density gradient), and à !
a temperature gradient. The electrical current density is ~
∇T
the same as before but is now expressed (in the linear jQ = (L12 (L11 )−1 L12 −L22 ) −
| {z } T
response) in terms of the additional external fields: ∼O(T 4 )'0

L22
Z Ã ! = − ∇T
d3 k ~
−∇T T
j~e = −2e ~ + L12
~v g ' L11 G . (153)
(2π)3 T π2 k2 T
⇒ καβ = σαβ (158)
3 e2
An expression for the thermal current can be de- This is the well known Wiedemann-Franz law
duced from the following thermodynamical relation dQ =
T dS = dU − µdN , hence ~ = Q∇T with je = 0. In
• Thermopower (Q): E
this case
Z Ã !
d3 k ~
−∇T L12 π2 k2 T σ0
j~Q = 2e ~ + L22
(² − µ)~v g ' L21 G . Q= = (159)
(2π)3 T L11 T 3 e σ
(154)
Here Lαβ are the linear transport coefficients and σ = • Peltier effect (∇T = 0): jQ = Πje ⇒ Π =
L11 . We now want to evaluate these expressions in the L21 /L11 = T Q
low temperature limit, where we can use ∂fF D /∂µ '
δ(² − ²F ), hence the expression for σ in (145) can be Historical note:
The Seebeck effect: The discovery of thermoelectricity
written as
dates back to Seebeck [1] (1770-1831). Thomas Johann See-
beck was born in Revel (now Tallinn), the capital of Estonia
Z which at that time was part of East Prussia. Seebeck was a
∂jα d3 k
σαβ (x) = = 2e2 vα vβ τ (x)δ(² − x), (155) member of a prominent merchant family with ancestral roots
∂Eβ (2π)3 in Sweden. He studied medicine in Germany and qualified
as a doctor in 1802. Seebeck spent most of his life involved
where σαβ = σαβ (²F ). We will use this new function σ(x) in scientific research. In 1821 he discovered that a compass
to express the other transport coefficients. For instance needle deflected when placed in the vicinity of a closed loop
in linear response, formed from two dissimilar metal conductors if the junctions
were maintained at different temperatures. He also observed
that the magnitude of the deflection was proportional to the
∂jQ temperature difference and depended on the type of conduct-
e2 L22
αβ = ³ ´

~
−∇T ing material, and does not depend on the temperature distri-
T bution along the conductors. Seebeck tested a wide range of
β
Z 3 materials, including the naturally found semiconductors ZnSb
d k ∂f
= 2e2 vα vβ (² − µ)2 τ (²) and PbS. It is interesting to note that if these materials had
(2π)3 ∂µ been used at that time to construct a thermoelectric genera-
Z
∂f tor, it could have had an efficiency of around 3% - similar to
= dx (² − µ)2 σαβ (²). (156) that of contemporary steam engines.
∂µ
The Seebeck coefficient is defined as the open circuit volt-
2
Using that (²F − µ) ' π6 (kT )2 D0 (²F )/D(²F ), we ob- age produced between two points on a conductor, where a
uniform temperature difference of 1K exists between those
tain (derivation in assignment).
points.
The Peltier effect: It was later in 1834 that Peltier[2] de-
π2 scribed thermal effects at the junctions of dissimilar conduc-
L22
αβ = (kT )2 σαβ (²F ) and similarly tors when an electrical current flows between the materials.
3e2 Peltier failed however to understand the full implications of
π2 his findings and it wasn’t until four years later that Lenz[3]
L12 21
αβ = Lαβ = − (kT )2 σαβ 0
(²F )
3e concluded that there is heat adsorption or generation at the
L11 = σαβ (²F ) (157) junctions depending on the direction of current flow.
αβ
The Thomson effect: In 1851, Thomson[4] (later Lord
Kelvin) predicted and subsequently observed experimentally
the cooling or heating of a homogeneous conductor resulting
Physical quantities
from the flow of an electrical current in the presence of a
temperature gradient. This is know as the Thomson effect
• Thermal conductivity (κ): jQ = κ(−∇T ) and is defined as the rate of heat generated or absorbed in a
24

metal semiconductor insulator

FIG. 21: 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.

Band Structure
FIG. 20: Thermoelectric cooling (left): If an electric current
is applied to the thermocouple as shown, heat is pumped from
the cold junction to the hot junction. The cold junction will Clearly the band structure of the semiconductors is
rapidly drop below ambient temperature provided heat is re- crucial for the understanding of their properties and their
moved from the hot side. The temperature gradient will vary device applications. Semiconductors fall into several cat-
according to the magnitude of current applied. Thermoelec- egories, depending upon their composition, the simplest,
tric generation (right): The simplest thermoelectric generator
consists of a thermocouple, comprising a p-type and n-type
type IV include silicon and germanium. The type refers
thermoelement connected electrically in series and thermally to their valence.
in parallel. Heat is pumped into one side of the couple and The band structure is quite rich and shown in figure
rejected from the opposite side. An electrical current is pro- (22) for silicon. Germanium is very similar to Si and does
duced, proportional to the temperature gradient between the
hot and cold junctions.

single current carrying conductor subjected to a temperature


gradient.
[1] Seebeck, T.J., 1822, Magnetische Polarisation der Met-
alle und Erzedurch Temperatur-Differenz. Abhand Deut.
Akad. Wiss. Berlin, 265-373.
[2] Peltier, J.C., 1834, Nouvelles experiences sur la calo-
riecete des courans electriques. Ann. Chem., LVI, 371-387.
[3] See Ioffe, A.F., 1957, Semiconductor Thermoelements
and Thermoelectric Cooling, Infosearch, London.
[4] Thomson, W., 1851, On a mechanical theory of thermo-
electric currents, Proc.Roy.Soc.Edinburgh, 91-98.

FIG. 22: Band structure of Si. (Figures from


ioffe.rssi.ru/SVA/NSM/Semicond)
SEMICONDUCTORS

not have a direct gap either. In general the dispersion


Semiconductors are distinguished from metals in that relation can be approximated with the use of the effective
they have a gap at the Fermi surface, and are distin- masses, noting that
guished from insulators in that the gap is smaller, which
is ambiguous. But generally, if the gap is close to 1eV 1 1 ∂ 2 E(~k)
it’s a standard semiconductor. If the gap is close to 3eV ∗ = 2 , (160)
mij ~ ∂ki ∂kj
it’s a wide band gap semiconductor and above 6eV it’s
usually called an insulator. Typically, the distinction is hence
simply made from the temperature dependence of the
conductivity. If σ → 0 for T → 0 it’s a semiconductor or à !
an insulator but the material is a metal if σ → σ0 > 0. kx2 + ky2 k2
E(~k) = Ec + ~ 2
+ x∗ (161)
However, even at room temperatures the conductivities 2mt∗ 2ml
of pure materials differ by many orders of magnitude
(σmetal >> σsemi >> σinsu ). for the conduction band and
25

FIG. 23: Band structure of Ge.

FIG. 24: Band structure of GaAs. Note the direct, Γ → Γ,


minimum gap energy. The nature of the gap can be tuned
~2~kh2 ~2~klp
2
~2~k 2 with Al doping.
E(~k) = Ev − ∗ − ∗ ' Ev − (162)
2mh 2mlp 2m∗p
conduction
hω ∼ Eg
for the valence band, where there are typically two bands, phonon vs k ≅ ω

the heavy holes and the light holes. These dispersion re- valence
photon ck ≅ ω ≅ E g/ h

lations are generally good approximations to the real sys-


tem. In addition, the valence band has another band due
FIG. 25: A particle-hole excitation across the gap can readily
to spin-orbit interaction. However, at zero wavenumber recombine, emit a photon (which has essentially no momen-
(~k = 0) this band is not degenerate with the other two tum) and conserve momentum in a direct gap semiconductor
light and heavy hole bands. (left) such as GaAs. Whereas, in an indirect gap semiconduc-
tor (right), this recombination requires the additional creation
m∗n m∗p m∗t m∗l m∗h m∗lp gap [eV] or absorption of a phonon or some other lattice excitation to
Si 0.36 0.81 0.19 0.98 0.49 0.16 1.12 conserve momentum.
Ge 0.22 0.34 0.0815 1.59 0.33 0.043 0.661
GaAs 0.063 0.53 0.063 0.063 0.51 0.082 1.424 no electrons in the conduction band (n = 0). At non-
zero temperatures, the situation is very different and the
TABLE I: The effective masses in units of the free electron carrier concentrations are highly T -dependent since all of
mass for the conduction band, the valence band, the trans-
the carriers in an intrinsic (undoped) semiconductor are
verse and longitudinal part in the conduction band and the
heavy and light hole mass in the valence band. thermally induced.
In this case, the Fermi-Dirac distribution defines the
temperature dependent density
The situation in III-V semiconductors such as GaAs
is similar but the gap is direct. For this reason GaAs Z Etop Z ∞
makes more efficient optical devices than does either Si n= DC (E)fF D (E)dE ' DC (E)fF D (E)dE
Ec Ec
or Ge. A particle-hole excitation across the gap can read- (163)
ily recombine, emit a photon (which has essentially no
Z Ev
momentum) and conserve momentum in GaAs; whereas,
in an indirect gap semiconductor, this recombination re- p= DV (E)(1 − fF D (E))dE (164)
−∞
quires the addition creation or absorption of a phonon
or some other lattice excitation to conserve momentum. To proceed further we need forms for DC and DV . Recall
2~ 2
For the same reason, excitons live much longer in Si and that in the parabolic approximation Ek ' ~2mk∗ we found
especially Ge than they do in GaAs.

material τexciton
Electron and hole densities in intrinsic (undoped) GaAs 1ns(10−9 s)
semiconductors Si 19µs(10−5 s)
Ge 1ms(10−3 s)
At zero temperature, the Fermi energy lies in the gap,
hence there are no holes (p = 0) in the valence band and TABLE II:
26

Eg
wk = v 2
k T=0 clearly kinetic
Thus, assuming that E − µ & 2 À kB T
energy increases
1
1 1
' = e−(E−µ)/kB T (168)
e(E−µ)/kB T +1 e(E−µ)/kB T
ie., Boltzmann statistics. A similar relationship holds for
E
0 ξ k = -E +
h k
2 2
holes where −(E − µ) & 2g À kB T
F 2m

1 1
FIG. 26: Partially filled conduction band and hole band at 1− = ' e(E−µ)/kB T
e(E−µ)/kB T +1 e−(E−µ)/kB T +1
non-zero temperature (169)
since e(E−µ)/kB T is small. Thus, the concentration of
3 √ electrons n
(2m∗ ) 2
that D(E) = 2π 2 ~3 E. Thus, 3 Z
(2m∗n ) 2 µ/kB T ∞ p
3 n ' e E − EC e−E/kB T dE
(2m∗n ) 2 p 2π 2 ~3 EC
DC (E) = E − EC (165) Z ∞
2π 2 ~3 3
(2m∗n ) 2 3 1
−(EC −µ)/kB T
= 2 3
(kB T ) 2
e x 2 e−x dx
2π ~
¡ ∗ ¢ 23 |0 {z }

2mp p π/2
DV (E) = EV − E (166) µ ¶ 3
2π 2 ~3 2πm∗n kB T 2
= 2 e−(EC −µ)/kB T
for E > EC and E < EV respectively, and zero otherwise h2
EV < E < EC . C −(EC −µ)/kB T
= Nef fe (170)
In an intrinsic (undoped) semiconductor n = p, and so
EF must lie in the band gap. Physically, this also means Similarly
that we have two types of carriers at non-zero temper- µ ¶ 23
atures. Both contribute actively to physical properties 2πm∗p kB T
p=2 e(EV −µ)/kB T = Nef
V
fe
(EV −µ)/kB T
such as transport. h2
(171)
C V
DC
where Nef f and N ef f are the partition functions for a
classical gas in 3-d and can be regarded as ”effective
f(E)D (E)
densities of states” which are temperature-dependent.
EC C
EF Within this interpretation, we can regard the holes and
EV
f(E)
(1 - f(E))D (E)
V electrons statistics as classical. This holds so long as n
and p are small, so that the Pauli principle may be ig-
DV
nored - the so called nondegenerate limit.
In general, in the nondegenerate limit,
FIG. 27: The density of states of the electron and hole bands µ ¶3
kB T ¡ ∗ ∗ ¢ 23 −Eg /kB T
np = 4 2
mn mp e (172)
If m∗n 6= m∗p (ie. DC 6= DV ), then the chemical 2π~
potential, EF , must be adjusted up or down from the this, the law of mass action, holds for both doped and in-
center of the gap so that n = p. trinsic semiconductor so long as we remain in the nonde-
Furthermore, the carriers which are induced across the generate limit. However, for an intrinsic semiconductor,
gap are relatively (to kB T ) high in energy since typically where n = p, it gives us further information.
Eg = EC − EV À kB T .
µ ¶ 23
kB T ¡ ¢ 43
Eg (eV ) ni (cm−3 )(300◦ K) ni = pi = 2 m∗n m∗p e−Eg /2kB T (173)
2π~2
Ge 0.67 2.4 × 1013
Si 1.1 1.5 × 1010 (See table (II)). However, we already have relationships
GaAs 1.43 5 × 107 for n and p involving EC and EV
C −(EC −µ)/kB T V (EV −µ)/kB T
TABLE III: Intrinsic carrier densities at room temperature n = p = Nef fe = Nef fe (174)

V
1eV Nef f
' 10000◦ K À 300◦ K (167) e2µ /kB T = C
e(EV +EC )/kB T (175)
kB Nef f
27

or Carrier Densities in Doped semiconductor


à !
V
1 1 Nef f
µ= (EV + EC ) + kB T ln (176) The law of mass action is valid so long as the use
2 2 C
Nef f of Boltzmann statistics is valid i.e., if the degeneracy is
small. Thus, even for doped semiconductor
µ ¶
1 3 m∗p
µ= (EV + EC ) + kB T ln (177) # ionized
2 4 m∗n EC
+ 0
ED ND = N D + ND

Thus if m∗p
6= m∗n ,
the chemical potential µ in a semicon- EF
# un-ionized
ductor is temperature dependent. EA
+ 0
NA = NA + NA
EV

Doped Semiconductors
FIG. 30: Ionization of the dopants
Since, σ ∼ nτ , so the conductivity depends linearly
upon the doping (it may also effect µ in some materials, C
np = Nef V
f Nef f e
−βEg
= n2i = p2i , (180)
leading to a non-linear doping dependence). A typical
metal has where β = 1/kB T . Using (177) we can define
µ ∗¶
1 3 mp
nmetal ' 1023 /(cm)3 (178) µi = (EV + EC ) + kB T ln , (181)
2 4 m∗n
whereas we have seen that a typical semiconductor has hence,
n = ni e(−µi −µ)/kB T and p = ni e(µi −µ)/kB T (182)
1010 ◦
ni ' at T ' 300 K (179)
cm3 To a good approximation we can assume that all
donors and all acceptors are ionized. Therefore,
Thus the conductivity of an intrinsic semiconductor is
quite small! n − p = ND − NA and np = n2i
To increase n (or p) to ∼ 1018 or more, dopants are n2
used. For example, in Si the elements used as dopants ⇒ n = ND − NA + i
n
p
are typically in the third or fifth column. Thus P or B ND − NA (ND − NA )2 + 4n2i
= +
2
Si 3s 3p
2 2 2
⇒ n ' ND and p ' n2i /ND for ND À NA
Si Si Si Si Si Si Si Si
e+ 2 3
P 3s 3p

p ' NA and n ' n2i /NA for NA À ND (183)


Si Si Si Si Si B Si Si 1
e- B 3s2 3p
Si P Si Si Si Si Si Si
r
2
r = h ε 2 big! By convention, if n > p we have an n-type semicon-
Si Si Si Si Si Si Si Si m* e
ductor, n+ -Si, or n-doped Si. The same is true for p.
Quite generally, at large doping the semiconductor will
FIG. 28: The dopant, P, (left) donates an electron and the
behave like a metal, since the dopant will spill over into
acceptor, B, donates a hole (or equivalently absorbs an elec-
tron). the conduction band (or valence band for holes). In this
case there is no gap anymore for the carriers and the con-
ductivity remains constant to the lowest possible temper-
will either donate or absorb an additional electron (with
atures. The main difference to a metal is only that the
the latter called the creation of a hole).
density of carriers is still very much lower (several orders
In terms of energy levels
of magnitude) than in a metal. At low doping the num-
n - SeC p - SeC ber of carriers in the conduction band (or valence band)
vanishes at zero temperature and the semiconductor be-
EC EF
EC haves like an insulator.
unoccupied at T= 0
ED
(occupied by holes at T = 0) An effective way to describe a semiconductor at high
ef f
occupied at T= 0 EA temperature or high doping is to consider that n = ND
EV EV EF ef f
and p = NA , where n and p are mobile negative and
ef f
positive carriers, whereas ND and NAef f are effective
FIG. 29: Left, the density of states of a n-doped semiconduc- positive and negative fixed charges, respectively. In this
tor with the Fermi level close to the conduction band and, approximation, charge neutrality is automatically veri-
right, the equivalent for a p-doped semiconductor. fied and we can use it to discuss heterogenous systems,
ef f
where ND depends on position (but is fixed).
28

Metal-Insulator transition where the constants (ρ0 , A and B) depend on the mate-
rial. In most metals and heavily doped semiconductors
In n-type semiconductors, when EF < EC , carriers ex- the temperature dependence of the resistivity is domi-
perience a gap ∆ = EC −EF . Hence at low temperatures nated by these three mechanisms, which means that the
the system is insulating. Indeed, importance of impurities, electron-phonon and phonon-
µ−EC
phonon interactions can be extracted from the tempera-
n = NC e kB T
(184) ture dependence of the resistivity.
|{z}
∼T 3/2
A special case is the magnetic impurity case (Kondo),
which gives rise to an additional term in 1/τK ∼
Since from Drude −(T /TK )2 .
ne2 τ
σ= → 0 when T → 0. (185)
m∗
In practice
When EF > EC , the semiconductor behaves like a
metal. In this case we obtain again from Drude that
2 To determine the density (n-p) the Hall resistance can
σ = ne τ
m∗ > 0 even for T → 0, since the density does be used. The ration τ /m∗ can then be obtained from
no vanish. In general, τ also depends on temperature.
the Drude conductivity and m∗ can be obtained from
Indeed, in the metallic phase the most important tem-
magneto-oscillations due to the Landau levels. This can
perature dependence comes from τ . To evaluate the con-
in principle be done for all temperatures, hence it is possi-
tributions from different scattering mechanism we can
ble to extract m∗ , n(T ), and τ (T ) simply by using trans-
consider 1/τ as the scattering probability. This follows
port and to deduce the dominant scattering mechanisms
directly from Boltzmann’s equation, where
¶ in the systems under study.
dg X
= (Γ(k 0 k) − Γ(kk 0 )), (186)
dt coll 0 k
p-n junction
where
(See also pn junction supplement)
0 0 0
Γ(kk ) = W (kk ) · g(k) · (1 − g(k ))
| {z } |{z} | {z }
prob. k→k0 # of states in k # of empty states in k 0
(187)
is the transition rate from state k to k 0 . Local equilibrium
implies Γkk0 = Γk0 k , hence g = fF D .
The most important scattering cases are the following:
• Impurity scattering (for a density of impurities nI ):
1
∼ nI (188)
τe−imp

• Electron-electron scattering:
1
∼ (T /TF )2 (189)
τe−e

• Electron-phonon scattering:
FIG. 31: Formation of a pn junction, with the transfer of
1 5 charges from the n region to the p region and the alignment
∼ (T /TD ) (190) ~ r) = 0. The depletion
of the chemical potential so that ∇µ(~
τe−ph
and accumulation regions are delimited by xn and −xp , re-
spectively.
In general, the total scattering probability is the sum
of all possible scattering probabilities, hence
When two differently doped semiconductors are
brought together they form a pn junction. In general,
1 1 1 1
= + + electrons form the more n-type doped region will trans-
τtot τe−imp τe−e τe−ph fer to the less doped or p-type region. This leaves a

µ ¶2 µ ¶5
m T T positively charged region on the n side and and accumu-
⇒ρ= ' ρ0 + A +B (191)
e2 nτ TF TD lation of negative charges on the p side. The potential
29

distribution can then be obtained by solving Poisson’s


equation v
Z R
z}|{
kF
dk ~k
∂ 2 V (x) e I = −env = −e 2·
= − ρ(x) (192) L
kF 2π m∗
∂x 2 ²0
~(kFL )2 − (kFR )2
= −e
where ρ(x) = −n(x)+p(x)+ND (x)−NA (x) is the charge 2πm∗
distribution of the mobile carriers (electrons and holes) 2e
= − ( µR − µL )
plus the fixed charges (donors and acceptors) and ²0 is h |{z} |{z}
−eVR −eVL
the dielectric constant.
2
2e
= (VR − VL ) (193)
h

- +
- +
- +
p type - + n type
- +
- +

no holes no electrons
charge density in depletion region

electric field
potential

Vo=(n-p)/e

FIG. 33: Schematic cross-sectional view of a quantum point


contact, defined in a high-mobility 2D electron gas at the in-
terface of a GaAs-AlGaAs heterojunction. The point contact
FIG. 32: Sketch of the charge density, the electric field and is formed when a negative voltage is applied to the gate elec-
the internal potential distribution due the transfer of charges trodes on top of the AlGaAs layer. Transport measurements
in a pn junction are made by employing contacts to the 2D electron gas at ei-
ther side of the constriction. (Beenakker, PHYSICS TODAY,
July 1996).
The most important consequence of a pn-junction is
the diode behavior. Indeed, when applying a negative Hence I = (2e2 /h)∆V ⇒ R = h/2e2 and G = 2e2 /h,
bias on the n region, the conduction bands tend to align where R and G are the resistances and conductances,
more and current can flow (forward bias). If the bias is respectively. This result can seem surprising at first since
positive the internal potential is increased and almost no it implies that the resistance does not depend on the
current can flow (reverse bias). This leads to the well length of the system. A very short quantum wire has the
known asymmetry in the current-voltage characteristics same resistance as an infinitely long wire. This result
of a diode. This is one of the most important elements of however, only applies for a perfect conductor. As soon
electronics. These ideas can be extended to three termi- as impurities lie in the wire this result has to be modified
nal devices such as a pnp junction of npn junction (which to take into account scattering by the impurities and then
is equivalent to two pn junction put together). In this R would typically depend on the length of the system.
case we have a transistor, i.e., by varying the potential
on the center element we can control the current through
the device. MORE THAN ONE CHANNEL, THE QUANTUM
POINT CONTACT

The previous result is specific to the purely one-


ONE DIMENSIONAL CONDUCTANCE dimensional case. In general the system can be extended
to a system of finite width, W . In this case the electron
Suppose that we have a perfect one-dimensional con- energy is given by
ductor (quantum wire) connected by two large electron
~2 kx2 ~2 (πn/W )2
reservoirs. In real life they would be electrical contacts. ²= + (194)
2m ∗ 2m∗
The left reservoir is fixed at chemical potential µL and
the right one at µR . The current is then given as usual if the boundary of our narrow wire is assumed to be
by sharp, since in this case the wave function has to vanish
30

FIG. 34: Conductance quantization of a quantum point con- FIG. 35: Imaging of the channels using an AFM (1 to 3 chan-
tact in units of 2e2/h. As the gate voltage defining the con- nels from left to right). The images were obtained by applying
striction is made less negative, the width of the point contact a small negative potential on the AFM tip and then measur-
increases continuously, but the number of propagating modes ing the conductance as a function of the tip scan and then
at the Fermi level increases stepwise. The resulting conduc- reconstruct the 2D image from the observed change in con-
tance steps are smeared out when the thermal energy becomes ductance. (From R.M. Westervelt).
comparable to the energy separation of the modes.

-M (magnetization)
at the boundary (like for an electron in a box of width
W ). In general, if

~2 (π/W )2 ~2 (2π/W )2 Type Type B>O (Vortex phase)


< E F < (195) H
2m∗ 2m∗ HC1 HC2
B (internal field)
then we recover the ideal case of a one-dimensional quan- B=O (Meissner)
tum wire, or single channel. If, however,
H
HC1 HC2
~2 (N π/W )2 ~2 ((N + 1)π/W )2
< E F < (196) T
(Abrikosov
2m∗ 2m∗ TC vortex crystal)
we can have n channels, where each channel contributes H
equally to the total conductance. Hence in this case G = HC1 HC2

N · 2e2 /h, where N is the number of channels. This


implies that a system, where we reduce the width of the FIG. 36: Magnetic field dependence in a superconductor.
conductor will exhibit jumps in the conductance of step (Vortex picture from AT&T ’95)
2e2 /n. Indeed, this is what is seen experimentally.

• Superconductivity can be destroyed by an exter-


QUANTUM HALL EFFECT nal magnetic field Hc which is also called the criti-
cal field (Kammerlingh-Onnes, 1914). Empirically,
The quantum hall effect is a beautiful example, where Hc (T ) = Hc (0)(1 − (T /Tc )2 )
the concept of a quantized conductance can be applied
to (see additional notes on the quantum Hall effect). • The Meissner-Ochsenfeld effect (1933). The mag-
netic field does not penetrate the sample, the mag-
netic induction is zero, B = 0. This effect distin-
SUPERCONDUCTIVITY guishes two types of superconductors, type I and
type II. In Type I, no field penetrates the sample,
The main aspects of superconductivity are whereas in type II the field penetrates in the form
of vortices.
• Zero resistance (Kammerlingh-Onnes, 1911) at T <
Tc : The temperature Tc is called the critical tem- • Superconductors have a gap in the excitation spec-
perature. trum.
31

Due to retardation, the electron-electron Coulomb repul-


sion may be neglected!
The net effect of the phonons is then to create an
attractive interaction which tends to pair time-reversed
quasiparticle states. They form an antisymmetric spin
k↑
e

Vanadium ξ ∼ 1000Α°

e

- k↓

FIG. 39: To take full advantage of the attractive potential


illustrated in Fig. 38, 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
FIG. 37: The critical magnetic field, resistance and specific
a singlet.
heat as a function of temperature. (Ref: superconductors.org)

singlet so that the spatial part of the wave function can


The main mechanism behind superconductivity is the be symmetric and nodeless and so take advantage of the
existence of an effective attractive force between elec- attractive interaction. Furthermore they tend to pair in
trons, which favors the pairing of two electrons of oppo- a zero center of mass (cm) state so that the two electrons
site momentum and spin. In conventional superconduc- can chase each other around the lattice.
tors this effective attractive force is due to the interaction Using perturbation theory it is in fact possible to show
with phonons. This pair of electrons has now effectively that to second order (electron-phonon-electron) the effect
zero total momentum and zero spin. In this sense this of the phonons effectively leads to a potential of the form
pair behaves like a boson and will Bose-Einsteein conden-
sate in a coherent quantum state with the lowest possible (~ωq )2
Ve−ph ∼ (197)
energy. This ground sate is separated by a superconduct- (²(k) − ²(k − q))2 − (~ω(q))2
ing gap. Electrons have to jump over this gap in order to
be excited. Hence when the thermal energy exceeds the This term can be negative, hence effectively produce and
gap energy the superconductor becomes normal. attraction between two electrons exceeding the Coulomb
repulsion. This effect is the strongest for k = kF and q =
The origin of the effective attraction between electrons
2kF since ²(kF ) = ²(−kF ) and ω(2kF ) ' ωD (the Debye
can be understood in the following way:
frequency). Hence electrons will want to form opposite
momentum pairs (kF , −kF ). This will be our starting
+
e- point for the microscopic theory of superconductivity à
ions
e- la BCS (Bardeen, Shockley and Schrieffer).
+ + + + + +

8
region of + +
+ + + +
vF ∼ 10 cm/s positive charge
attracts a second
+
electron BCS theory

FIG. 38: Origin of the retarded attractive potential. Elec-


trons at the Fermi surface travel with a high velocity vF . As To describe our pair of electrons (the Cooper pair) we
they pass through the lattice (left), the positive ions respond will use the formalism of second quantization, which is
slowly. By the time they have reached their maximum excur- a convenient way to describe a system of more than one
sion, the first electron is far away, leaving behind a region of particle.
positive charge which attracts a second electron.

p2
When an electron flies through the lattice, the lat- H1particle = ⇒ H1p ψ(x) = Eψ(x) (198)
2m
tice deforms slowly with respect to the time scale of
the electron. It reaches its maximum deformation at Let’s define
a time τ ' ω2πD ' 10−13 s after the electron has passed. c+ |0i = ψ(x) and h0|c1 (x) = ψ ∗ (x)
1 (x) (199)
In this time the first electron has travelled ' vF τ ' |{z}
108 cm −13
s · 10 s ' 1000Å. The positive charge of the lat-
vacuum

tice deformation can then attract another electron with- With these definitions, |0i is the vacuum (or ground
out feeling the Coulomb repulsion of the first electron. state), i.e., state without electrons. c+
1 (x)|0i corresponds
32

to one electron in state ψ(x) which we call 1. c+ is also


called the creation operator, since it creates one elec- (
tron from vacuum. c is then the anhilation operator, i.e., c1 = A1 cos(θ) + A+ 2 sin(θ)

c1 c+ c+ = −A sin(θ) + A+2 cos(θ)


1 |0i = |0i, which corresponds to creating one electron 2 1
(
from vacuum then anhilating it again. Other properties
A1 = c1 cos(θ) − c+
2 sin(θ)
include ⇒ (204)
A2 = c1 sin(θ) + c+
+
2 cos(θ)
c1 |0i = 0 and c+ +
1 c1 |0i = 0 (200)
It is quite straightforward to see that A+ +
i Aj +Aj Ai = δij ,
+ + + +
The first relation means that we cannot anhilate an Ai Aj + Aj Ai = 0, and Ai Aj + Aj Ai = 0 using the
electron from vacuum and the second relation is a con- properties of ci . We can now rewrite HM F in terms of
sequence of the Pauli principle. We cannot have two our new operators Ai :
electrons in the same state 1. HM F = t(c+ + + + 2
1 c1 + c2 c2 ) − ga(c1 c2 − c1 c2 ) + ga
Hence,
= t(A1 cos(θ) + A2 sin(θ))(cos(θ)A1 + sin(θ)A+
+
2 ) + ···
c+ + + + +
1 c1 = 0 ⇒ (c1 c1 ) = 0 ⇒ c1 c1 = 0 = + +
(A1 A1 + A2 A2 )(t cos(2θ) − ga sin(2θ))
⇒ (c+ + +
1 c1 )c1 |0i = c1 |0i (201) + t(1 − cos(2θ)) + ga sin(2θ) + ga2
+ (A+ +
1 A2 − A1 A2 ) (t sin(2θ) + ga cos(2θ)) (205)
This shows that c+ 1 c1 acts like a number operator. It | {z }
counts the number of electrons in state 1. (Either 1 or =0 to diagonalize HM F
0).
Hence the diagonalization condition for HM F fixes the
We can now extend this algebra for two electrons in
angle θ of our Boguliubov transformation:
different states c+ 1 |0i corresponds to particle 1 in state 1
and c+ |0i to particle 2 in state 2. The rule here is that ga −ga
2 tan(2θ) = − ⇒ sin(2θ) = p (206)
c+
i cj + c j c+
i = δi,j . t t + (ga)2
2

Finally we can write down the two particle hamiltonian


Hence HM F now becomes
as
p p
HM F = (A+ + 2 2 2
1 A1 +A2 A2 ) t + (ga) +(t− t + (ga) )+ga
2 2
H2p = t1 c+
1 c1 + t2 c+
2 c2 − gc+ +
1 c1 c2 c2 (202) (207)
It is now immediate to obtain the solutions of the
where ti is the kinetic energy of particle i and g is
Hamiltonian, since we have the ground state |ai and we
the attraction between particle 1 and 2. We will also
have a diagonal Hamiltonian in terms of Ai hence the
suppose that t1 = t2 for the Cooper pair. The job now
Ground state energy E0 is simply given by HM F |ai =
is to find the ground state of this Hamiltonian. Without
E0 |ai and Ai |ai = 0. The first degenerate excited states
interactions (g = 0) we would simply have E = t1 +
are A+ + +
i |ai with energy E1 , where HM F Ai |ai = E1 Ai |ai
t2 . The interaction term is what complicates the system + +
and the next energy level and state is A1 A2 |ai, with
since it leads to a quadratic term in the Hamiltonian.
energy E2 given by HM F A+ + + +
1 A2 |ai = E2 A1 A2 |ai, hence
The idea is to simplify it by getting rid of the quadratic
p
term. This is done in the following way. From eq. (202) E2 = t + t2 + (ga)2 + ga2
we have
E1 = t + ga2
p
E0 = t − t2 + (ga)2 + ga2 (208)
H = t(c+ + + +
1 c1 + c2 c2 ) − gc1 c1 c2 c2 If we take t → 0 and define ∆ = ga we have
= t(c+ + + +
1 c1 + c2 c2 ) + gc1 c2 c1 c2 

 E2 = ∆ + ga
2
= t(c+ + + +
1 c1 + c2 c2 ) − ga(c1 c2 − c1 c2 ) + ga
2

+ g(c+ c+ + a)(c1 c2 − a) (203) E1 = ga2 (209)




| 1 2 {z } E0 = −∆ + ga2
'0 (Mean field approx.)
We can now turn to what a is since,
⇒ HM F = t(c+ + + +
1 c1 + c2 c2 ) − ga(c1 c2 − c1 c2 ) + ga
2

We used the mean field approximation, which re- a = ha|c1 c2 |ai


places c1 c2 by its expectation value ha|c1 c2 |ai = a ⇒ = − cos(θ) sin(θ)ha|A1 A+1 |ai
ha|c+ +
1 c2 |ai = −a, where |ai is the ground state of the
Hamiltonian. The idea now is to = − cos(θ) sin(θ) ha|ai
Pdiagonalize HM F , i.e. | {z }
a Hamiltonian in the form H = i c+ i ci . The trick here =1
is to use the Boguliubov transformation: ⇒ a = − sin(2θ)/2 (210)
33

Combining (206) and (210) we obtain another one with momentum −k − q, hence momentum
is conserved in this scattering process and a phonon with
ga momentum q is exchanged.
2a = p (211) How can we relate BCS theory to the observed Meiss-
t2 + (ga)2 ~ into the
ner effect? By including the vector potential A
which is the famous BCS gap (∆) equation. Indeed, it BCS Hamiltonian it is possible to show that the current
has two solutions, density is then given by
2
a = 0 ⇒ ha|c1 c2 |ai = 0 ⇒ Normal ~j = ne A~ (216)
a 6= 0 ⇒ t2 + (ga)2 = g 2 /4 ⇒ Superconductor(212) mc
| {z }
√ 2
∆=ga= 2
g /4−t The magnetic induction is B ~ =∇~ ×A ~ as usual. This is
in fact London’s equation for superconductivity. We can
'=ga now take the rotational on both sides of (216), hence
Normal

~ × ~j ne2 ~ ~
∇ = ∇×A
Superconductor |{z} mc
~ B=
∇× ~ 4π ~j
c
t 2
g/2
FIG. 40: The gap of a BCS superconductor as function of the
~ ×∇
⇒∇ ~ = − 4πne B
~ ×B ~
mc2
kinetic energy.
⇒ Bx ∼ e−x/λL (217)
This gives us the condition for superconductivity g ≥ q
mc2
2t. Hence the attraction between our two electrons has with λL = 4πne2 which is the London penetration
to be strong enough in order to form the superconducting length.
gap ∆. Typically, t is directly related to the temperature,
hence there is a superconducting transition as a function
of temperature.
We now want to find the expression for our supercon- B
ducting wavefunction |ai. The most general possible form
is e -x/ L

Vacuum
|ai = α|0i + β1 c+ + + +
1 |0i + β2 c2 |0i + γc1 c2 |0i (213) Inside the
superconductor

In addition the condition Ai |ai = 0 has to be verified, x


which leads after some algebra to
FIG. 41: The decay of the magnetic field inside the supercon-
ductor. The decay is characterized by the London penetration
|ai = α(1 + tan(θ)c+ +
1 c2 )|ai (214) length λL .

This state clearly describes an electron pair, the


Cooper pair and represents the superconducting ground If we had a perfect conductor, this would imply that
state of the Hamiltonian. In our derivation we only con- the current would simply keep on increasing with an ex-
~ hence
ternal field E,
sidered two electrons, but this framework can be gener-
alized to N electrons, where the generalized BCS Hamil-
∂~j ne2 ~
tonian can be written as = E
∂t mc
X X ∂ ~ ~ ne2 ~ ~
⇒ ∇ ×j = ∇×E
HBCS = tk c+
k, σ ck,σ − Vq c+ +
k+q,↑ c−k+q,↓ ck,↑ c−k,↓ ∂t mc
|{z} 2
k,σ spin
q ∂ ~ ~ ×B~ = ∂ 4πne B ~
⇒ ∇ ×∇ (218)
(215) ∂t ∂t mc2
Here Vq is positive and represents the effective phonon
induced attraction between electrons at the Fermi level. This equation is automatically verified from (217).
It’s maximum for q = 0. The second term describes the Hence, (217) implies both the Meissner effect and zero
process of one electron with momentum k and another resistance, which are the main ingredients of supercon-
ductivity. (Reminder: Maxwell gives ∇~ ×B ~ = 4π ~j and
electron with momentum −k which are anhilated in or- c
der to create one electron with momentum k + q and ∇ ~ = − 1 ∂ B~ ).
~ ×E
c ∂t
34

A remarkable aspect of superconductivity is that one where U (1) ⇔ c → ceiα (Gauge invariance), U (2) ⇔ c →
of the most fundamental symmetries is broken. Indeed, Ac (A = eiφ/2 a) and a+ a = 1, SU (2) ⇔ φ = 0.
~ What hap-
Gauge invariance is broken because ~j ∼ A.
pens is that below Tc we have a symmetry breaking,
which leads to new particles, the Cooper pairs. Math-
ematically, we have

U (2) = SU (2) ⊗ U (1) , (219)


| {z } | {z } | {z }
N ormal state SC state Gauge invariance

You might also like