Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

Condensed Matter Physics

Lecture notes — Baruch Horovitz and class of 2016

Contents
1 Transport in channels and Landauer-Büttiker formulas 2
1.1 Ideal wire . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Realistic wire . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 4 terminal conductance . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Incoherent scattering – the Drude case . . . . . . . . . . . . . . . . . 6
1.5 Multi terminal conductance . . . . . . . . . . . . . . . . . . . . . . . 6

2 Quantum Hall Effect 8


2.1 The Classical Picture . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Landau Levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Integer quantum Hall . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Edge states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

3 Fractional Quantum Hall Effect 17


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Laughlin’s wave function . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 Excitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4 Two holes and fractional statistics . . . . . . . . . . . . . . . . . . . . 23
3.5 Composite Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.6 FQHE edge states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

4 Quantum Dots 28
4.1 Coulomb blockade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 The capacitance model . . . . . . . . . . . . . . . . . . . . . . . . . . 31

5 Graphene 33
5.1 Band structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.2 Symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.3 Klein paradox . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.4 The Quantum Hall Effect in Graphene . . . . . . . . . . . . . . . . . 38
5.5 Edge states in graphene . . . . . . . . . . . . . . . . . . . . . . . . . 40

References and figure sources:

Yoseph Imry, Introduction to Mesoscopic Physics, Oxford University Press, 1997.


Thomas Ihn, Semiconductor Nanostructures Quantum States and Electronic Trans-
port Oxford University Press, 2010.
Jainendra K. Jain, Composite Fermions , Cambridge University Press, 2007.
A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov and A. K. Geim,
The electronic properties of graphene, Rev. Mod. Phys. 81, 109 (2009).

1
1 Transport in channels and Landauer-Büttiker
formulas
1.1 Ideal wire
Transport in channels is basic formalism which is used extensively in these notes.
We introduce first the notion of transport in an ideal wire with several transverse
modes, meaning that the wavefunctions in the wire are of the form
1
ψn± (r) = χn (y, z) · √ e±ikn x (1)
kn

where x is along the wire and χn (y, z) are eigenfunctions in the transverse directions
y, z, the normalization √1kn is chosen so that these states carry unit current (which
will be useful below). The eigenvalues have the form En (kn ) = En + (kn ), where
2 2
usually we take (k) = ~2mk∗n . The wires are connected to two reservoirs with different
chemical potentials µL , µR . The chemical potentials determines how many channels
are open, i.e. there is a propagating solution with kx real for energies in an open
2 2
channel, e.g. on the left µL = En + ~2mk∗n so that µL > En , see Fig. 1.

Figure 1: (a) One-dimensional channel connected to left and right electron reservoirs (gray) showing
ransverse modes with their propagation direction. (b) For each mode n, the energy En (kn ) has its
minimum at energy En . The gray-shaded energy interval is given by the applied voltage between
left and right reservoirs.

To evaluate the current we need to sum the current of modes kn > 0 that
propagate from the left with µL and subtract those of kn < 0 that propagate Pfrom
right with µR . We convert summation on kn to integration on energy by 2 kn =
2 dk
R n R 1

= π~v n
dE, i.e. a density of states factor with 2 for the number of spin
states, vn is the velocity vn = ~1 dEn
dkx
. The current is the evn times the probability of
occupation of energy E, i.e. the Fermi functions at temperature T
1
fL,R (E) = = f (E − µL,R ) (2)
e(E−µL,R )/kB T +1

2
The current is then
XZ 1 e X
Z
I=e dE vn [fL (E) − fR (E)] = dE[fL (E) − fR (E)] (3)
n En π~vn π~ n En

where the sum is on open channels and integration starts at the minimal energy En .
Note the remarkable cancellation of velocities vn . We focus on linear response, i.e.
the current for a small voltage µL −µR = |e|V  kB T , then expand fL (E)−fR (E) =
∂fL
∂µL
(µL − µR ) = − ∂fL
∂E
|e|V and the conductance becomes

I 2e2 X 2e2
G= = fL (En ) → N (4)
V h n h

where the limit is at low temperatures kB T  En+1 − En and N is the number of


open channels. See Fig. 2 for experimental confirmation.
Remarkably, the conductance of the ideal chan-
nel is finite even though there is no scattering in- Figure 2: Quantization of con-
side. From Ohms law, however, we are used to the ductance in a quantum point
fact that any finite resistance comes along with en- contact. Changing the gate volt-
age changes the numebr of open
ergy dissipation. We therefore face the question of channels in the contact, see inset
where energy is dissipated in our system. An elec- (van Wees et al., 1988).
tron with energy µL > E > µR travelling from left
to right leaves a nonequilibrium hole behind and be-
comes a nonequilibrium charge on the right, both hole
and charge eventually relax to the respective electro-
chemical potential. The energy dissipated by the hole
is µL − E and by the charge E − µR , together they
dissipate µL − µR = |e|V . This happens at a rate of
2
I/|e| hence the power dissipated is eV 2eh
N V = 2ehN V 2
2
identifying a contact resistance 1/Rc = 2ehN , just as
G above. This rough argument neglects details of the
nonequilibrium charge distribution near the contacts,
yet it clarifies the source of dissipation. The contact
resistance will be identified below more precisely.

1.2 Realistic wire


We proceed to consider a non-ideal wire that has impurities, leading to reflections
and scattering between channels. Assymptotically, we have free waves as given in
Eq. (1), hence the general form of a solution at x → ±∞ for an incoming wave from
the left has the form

(
√1 χn (y, z)eikn x + √1 −ikm x
P
kn m km rmn χm (y, z)e x → −∞
ψn (r) = (5)
√1 ikm x
P
m km tmn χm (y, z)e x→∞
0
For incident wave from the right there is a similar definition with reflections rmn
and transmissions t0mn . Define the total reflection and transmission into a channel

3
n on the left as Rn , Tn0 and into channel n on the right as Rn0 , Tn ,
X X
Rn = |rnm |2 , Tn0 = |t0nm |2
m m
X X
Rn0 = 0
|rnm |2 Tn = |tnm |2 (6)
m m

The current, e.g. on the right of the scattering region, is obtained by the transmitted
current with distribution fL (E), by the current from the right electrode with fR (E)
and by its reflected part,
XZ 1
I=e dE vn (fL (E)Tn (E) + fR (E)Rn0 (E) − fR (E)) (7)
i
π~vn

To proceed we need identities from current conservation. Consider a situation that


all incoming channels from both left and right are fully occupied, then all outgoing
channels n are also fully occupied, so on the right it is Tn + Rn0 = 1 and on the left
it is Tn0P
+ Rn = 1. Furthermore, the total P current originating from channel n on the
2 2
mn | + |rmn | ] = 1, hence
left is m [|tP n (Tn + Rn ) = N , while current from n on
the right is m [|tmn | + |rmn | ] = 1, hence n (Tn0 + Rn0 ) = N .
0 2 0 2
P
One can proceed now to Eq. 12, however, more insight into these identities is
gained by considering a scattering formulation. The asymptotic states for incoming
waves from either left or right are, in terms of the states Eq. (1),
X
ψ(r) = [an ψn+ (r) + bn ψn− (r)] x→∞
n
X
ψ(r) = [cn ψn− (r) + dn ψn+ (r)] x → −∞ (8)
n

We define an S matrix that relates the incoming waves to the outgoing ones

r t0
     
a b
S = S= (9)
c d r0 t

where r, t are reflection and transmission for incident waves from the left, as in Eq.
(5), while r0 , t0 are the same for incident waves from the right, all these elements are
N × N matrices, so S is a 2N × 2N matrix. The incoming current is |a|2 + |c|2 while
2 2
the outgoing√ one is |b| + |d| which is why we have chosen to normalize the states
(1) with 1/ kn , carrying unit current. Current conservation means therefore that
the S matrix is unitary, i.e. S · S † = 1. Hence the diagonal element is
X

1= (rnm rmn + t0nm t0∗ 0
mn ) = Rn + Tn (10)
m

and similarly Rn0 + Tn = 1. Furthermore, S † · S = 1 has diagonal elements


X X

1= (rnm rmn + t∗nm tmn ) = [|rmn |2 + |tmn |2 ] = 1 (11)
m m
P
hence n (Tn + Rn ) = N , recovering more formally the identities above.

4
Proceeding now from Eq. (7)
Z  
e X e X ∂fL
I= dE[fL (E) − fR (E)]Tn (E) = − (µL − µR )Tn (E)(12)
π~ n π~ n ∂E

At low temperatures − ∂f
∂E
L
→ δ(E − µL ) and the conductivity becomes

I 2e2 X 2e2 X 2e2


G2 = = Tn = |tnm |2 = T r(t† t) (13)
(µL − µR )/e h n h nm h

where t, as above, is an N × N matrix, Tn is evaluated at the chemical potential.


This is a simple and powerful expression for the 2-terminal conductivity.

1.3 4 terminal conductance


Additional information is obtained by measuring the voltage with separate electrodes
at the ideal part of the wire, outside the range of scatterers. The voltage probes
then local chemical potentials µA , µB . We consider, for simplicity, a one-channel
case. Define local densities at locations A, B
Z Z
2 1
nA = dE fA = dE · [(1 + R)fL + T 0 fR ]
π~v π~v
Z Z
2 1
nB = dE fB = dE · [T fL + (1 + R0 )fR ] (14)
π~v π~v
Note the factor 2 in the density of states, since here we sum both ±k. Expanding
in small differences of chamical potentials

(1 + R)fL + T 0 fR = 2fA T fL + (1 + R0 )fR = 2fB


(1 + R)µL + T 0 µR = 2µA T µL + (1 + R0 )µR = 2µB
⇒ µA − µB = 12 (1 + R − T )(µL − µR ) = R(µL − µR ) (15)
2e2
and R0 = R is used (1 channel). Since the current is I = h
T, the 4 terminal
conductance is
I G2 2e2 T
G4 = = = (16)
(µA − µB )/e R h R

Hence, for an ideal wire with no reflection G4 → ∞. We can now identify a universal
contact resistance Rc that is independent of the scattering,
1 1 h
= + Rc , ⇒ Rc = (17)
G2 G4 2e2
in support of identifying the source of dissipation at the contact. In section 1.5
we develop a multi-terminal formulation allowing transmissions and reflections at
all terminals. The result for G4 can be obtained as a delicate limit where the
transmissions from the voltage leads vanish [See M. Büttiker, IBM J. Res. Develop.
32, 317 (1988), Eq. 32].

5
1.4 Incoherent scattering – the Drude case
So far we studied coherent scattering, i.e. elastic scattering from impurities where
the phase coherence is maintained. Consider now an incoherent situation where
inelastic scattering leads to the loss of phase information. Considering one channel
with many scatterers M, scattering probabilities Ri , Ti are added. For 2 scatterers,
their combined transmission is


(2) 2
X T1 T2
T = T1 T2 + T1 R2 R1 T2 + T1 (R2 R1 ) T2 + ... = T1 T2 (R1 R2 )n =
n=0
1 − R1 R2
(2)
1−T 1 − R1 R2 − T1 T2 T1 + T2 − 2T1 T2 1 − T2 1 − T1
⇒ = = = + (18)
T (2) T1 T2 T1 T2 T2 T1

using R1 = 1 − T1 and R2 = 1 − T2 . We find that (1 − T )/T is additive. Since


Ti are random, we can use the central limit theorem and replace the sum on many
scatterers by an average

1 − T (M ) 1 − Ti N
(M )
= Mh i≡ (19)
T Ti ν`e
D E
where ν = M/L is the density of scatterers and `e = [ν 1−TTi
i
]−1 is an elastic mean
free path. For weak scatterers Ri = 1 − Ti  1 and `e ≈ 1/(ν hRi i). The product
ν hRi i can be interpreted as an average backscattering probability per unit length.
The probability Pbs that an electron has not been backscattered after travelling a
distance L therefore obeys the equation
dPbs
= −ν hRi i Pbs ⇒ Pbs = e−νhRi iL = e−L/`e (20)
dL
showing exponential localization with a decay length `e . The conductance is G =
2e2 (M )
h
T N is additive for N incoherent channels; for width W and Fermi wavevector
kF the number of transverse modes is N = kF W/π and since T (M ) = `e /L we have
2
G = 2eh kFπL
`e W
, similar to the Drude result. (The contact resistance is negligible in
this case of large M).

1.5 Multi terminal conductance


We consider next a general arrangement of terminals
(Fig. 3), each has current Iα (going away from the Figure 3: 4 terminal arrange-
reservoir) and voltage Vα . We seek the linear conduc- ment, possibly with a flux Φ.
tance that relates the currents to the voltages,
X
Iα = Gαβ Vβ (21)
β
P
Current conservation gives α Iα = 0. ThePmatrix
Gαβ is constrained: if only one Vβ 6= 0 then α Iα =

6
P P
αGαβ Vβ = 0, hence α Gαβ = 0. Also,P if all Vβ = V
are equal then no response, Iα = 0, hence β Gαβ = 0.
Reflection and transmission coefficients are defined
by the asymptotic form, where x → −∞ is far into a given terminal and x → ∞ is
far into other terminals,
 α
α
 √1kα χαi (y, z)eiki x + j rji √ α χj (y, z)e−ikj x
P α 1

kj x → −∞
ψiα (r) =
i
P βα 1 β ik β
x (22)
 β,j tji
q χ (y, z)e j
β j x→∞
k

j

The current is then given by


" #
XZ 1 α X X X ij
α 2
Iα = −e dE v fα (E) −
α i
|rij | fα (E) − |tαβ |2 fβ (E) (23)
i∈α
π~v i j∈α α6=β j∈β

This yields the Landauer-Büttiker formalism,


Z " #
2|e| X
Iα = − dE Nα (E)fα (E) − Rα (E)fα (E) − Tαβ (E)fβ (E) (24)
h β6=α

α 2
P
where Nα is the number of open channels, Rα = ij∈α |rij | is the reflection back
P αβ 2
into terminal α, Tαβ = j∈β,i∈α |tji | is the transmission from β to α, and fα,β are
the Fermi functions at the chemical potentials of the terminals.
In linear response the potential difference between the various leads is small,
hence we expand to first order in µβ − µα :
∂f ∂f
fβ (E) ≈ fα (E) + (E − µβ − (E − µα )) = fα (E) − (µβ − µα ) (25)
∂E ∂E
In equilibrium µα = µβ the current must vanish (Iα = 0), hence the relation
X
Nα = Rα (E) + Tαβ (E) (26)
β6=α

and the current becomes


Z
2|e| X ∂f
Iα ≈ − dETαβ (E) (µβ − µα ) (27)
h β6=α ∂E

The resulting conductance is, with µα = |e|Vα ,


2e2 2
Z  
∂f T →0 2e
Gαβ = dETαβ (E) − −−−→ Tαβ (EF ) (28)
h ∂E h
∂f
where at low temperatures − ∂E → δ(E − EF ) and the Fermi energy EF is the
common limit of equal µα .
In conclusion within linear response we have
X X
Iα = Gαβ (Vβ − Vα ) = Gαβ Vβ (29)
β6=α β

where formally only source voltages appear, a convenience in calculations. As shown


above the matrix G is constrained so as all sums on either rows or columns vanish.

7
2 Quantum Hall Effect
2.1 The Classical Picture
The discovery of the Hall effect in 1879 has shown that a magnetic field normal to
a current flow results in an electric field that is normal to both magnetic field and
the current direction, see Fig. (4). The discovery of the electron in 1897 has led to
the basic understanding of this phenomena, i.e. the moving electrons in the current
(x) direction feel a Lorenz force in the y direction, leading to a charge accumulation
on the surface and a field Ey that cancels the Lorenz force, so that no current flows
in the y direction. Hence

Figure 4: Hall Bar with source (S) drain (D) electrodes, magnetic field B, current I and the Hall
voltage VH .

 
1 cEy
E + v × B = 0 ⇒ vx = (30)
c y B

Since the current density is j = −ne ev for an electron density ne , the Hall conduc-
tivity σH becomes
cEy
jx = −ne evx = −ne e
B
jx ne ec (31)
⇒ σH ≡ =−
Ey B
Note that in 2-dimensions (2D) the Hall conductance equals the Hall conductivity
and is independent of size Ly , i.e.
jx jx Ly Ix
σH = = =
Ey Ey Ly VH

We next consider Newton’s equation for the time dependent momentum p(t) of
an electron,
dp  p 
= −e E + ×B (32)
dt mc
This can be solved by the transformation p = p0 + mc E×B B2
, reducing this equation
dp0 p0 0
to dt = −e mc × B. Hence p performs circular motion with the cyclotron frequency
eB
ωc = mc while p describes a trochoid, Fig. (5), i.e. a circular motion superimposed

8
on a drift velocity in the y direction vD = c E×B
B2
so that the conductance in the x
direction vanishes, σxx = 0.
We extend now the classical model to include random impurities via Drude’s
model. The impurities cause scattering between shifted trochoids, as in Fig. (5),
leading to an average velocity in the x direction so that σxx 6= 0.
The average effect of impurity scattering is
included via a relaxation time τ Figure 5: Effect of collisions on the elec-
tron’s trajectory.
dp  p  p
= −e E + ×B − (33)
dt mc τ
dp
Steady state solution dt
= 0 yields
eB px
0 = −eEx − py −
mc τ (34)
eB py
0 = −eEy + px −
mc τ
ne e2 τ
or in terms of the current j = −ne ep/m and σ0 = m
(the Drude conductivity for
B = 0),
1 B
Ex = jx + jy
σ0 ne ec
B 1
Ey = − jx + jy (35)
ne ec σ0
A resistance tensor is defined by E = ρ̂j where
 1 B   B

σ0 ne ec ωc τ →∞ 0 ne ec
ρ̂ = −−−−→ (36)
− nB
e ec
1
σ0
− nB
e ec
0
so that at large B the effect of impurities is weaker. A conductivity tensor is defined
by inverting to j = σ̂E, or σ̂ = ρ̂−1 ,
!
σ0 σ0 ωc τ  ne ec

1+ωc2 τ 2 1+ωc2 τ 2 ωc τ →∞ 0
σ̂ = σ0 ωc τ σ0 −−−−→ B (37)
− 1+ω 2τ 2 1+ω 2 τ 2
− nB
e ec
0
c c

Note in particular that σxx 6= ρ−1


xx . By requiring jy = 0 in Eq. (35) we find Hall
conductance for any B
jx ne ec
σH = =− (38)
Ey B
which is independent of τ ! Note also that all electrons contribute to σH since
σ H ∼ ne .
We conclude this classical treatment by addressing the question of dissipation.
In Eq. 36, we noticed that for high magnetic fields ρxx → 0, a result that may be
surprising as it implies that there is no dissipation in the sample. It is useful to
consider equipotential lines. For B = 0 the field Ex produces vertical equipotential
lines in y direction, Fig. 6 (left). For ωc τ → ∞ Eq. (35) gives for the Hall geometry
jy = 0, E x
Ey
= ω1c τ → 0. Hence the equipontial lines are horizontal in the bulk, as

9
in Fig. 6 (right). The difficulty arises at the contacts, where the magnetic field is
excluded, Ex dominates and the field lines become vertical. Since the equipotential
lines are continuous there is an accumulation of lines at the corners, as in Fig. 6
(right). The voltage drop at the contacts is then the same as the voltage difference
of the upper and lower lines in the bulk, i.e. VH , and leads to dissipation at the
contact.

Figure 6: Equipotential lines at B = 0 (left), and in a strong magnetic field (ωc τ  1) on the
right. The electrodes are source (S) and drain (D) for driving a current Ix .

2.2 Landau Levels


We start the quantum description with a Hamiltonian of 2D free electrons in the
presence of a perpendicular magnetic field in the z direction (B = B ẑ):

1  e 2
H= p+ A (39)
2m c
Using Landau gauge we take A = (0, Bx), and we get:
 2 
1 2
e
H= p + Bx + py (40)
2m x c

Hence the eigenfunctions have the form ψ(x, y) = √1 eiky χ(x) where χ(x) satisfies
Ly

~2 ∂ 2
 
1 2 2
2
− + mωc x − ` k χ(x) = Eχ(x) (41)
2m ∂x2 2
where the fundametal energy and length scales are the cyclotron frequency ωc and
the cyclotron radius, or the magnetic length `,
r
eB ~c
ωc = `= (42)
mc eB
The equation for χ(x) is a harmonic oscillator one whose center is at `2 k. The well
known eigenstates are
√ −1/2 − (x−`2 k)2 x − `2 k
 
N
χN (x) = 2 N ! π` e 2` HN
2
(43)
`

10
where HN (x) is the N -th Hermite polynomial. and the energy eigenvalues have
integers N ≥ 0

EN = ~ωc N + 12 (44)
These states are known as Landau levels.
Since EN do not depend on k each level is degenerate. Periodic boundary con-
dition in the y direction yields k = 2πn
Ly
and the integer n is such that the center
Lx Ly
position of the oscillator states is 0 < k`2 = 2πn
Ly
`2 < Lx , hence 0 < n < 2π`2
and
the degeneracy of each Landau level in area Lx Ly = A
A AB
Ns = 2
= (45)
2π` hc/e
Note that Ns is the total flux AB divided by the flux quanta hc/e, i.e. the number
of flux quanta in the whole area. The area 2π`2 accommodates one flux quanta, i.e.
2π`2 B = φ0 = hc/e.
1
A full Landau level corresponds to an electron density of 2π`2 . We define the

filling factor that measures how many Landau levels are occupied for an electron
density ne
ν = 2π`2 ne (46)
so that ν is controlled by either the density ne or by the field B.

2.3 Integer quantum Hall


We proceed now to the major discovery of the integer quantum Hall effect (von
2
Klitzing, 1980). Hall conductance in Drude’s model Eq. (38) is σH = ν eh , however
the data on 2D samples has shown remarkably flat plateaus for ν centered around
2
an integer n with ρxy = eh n, while ρxx = σxx = 0 in the range of these plateau, see
Fig. 7.
To understand this remarkable phenomena we
need to consider first the effect of disorder and lo-
calization. For B = 0 we know that all states in 2D
are localized, while in 3D one usually has a mobil-
ity edge, separating extended states at high density of
states (DOS) and localized states at low DOS. In a
strong magnetic field we expect that disorder broad-
ens the Landau levels, leading to maxima of the DOS
at the Landau levels, Fig. 8. At such maxima there
are many states that facilitate an extended state, sim-
Figure 8: Density of States for
ilar to the 3D case at high DOS. The dissipative ρxx Landau levels with disorder. Ex-
is determined by states near the Fermi level, hence if tended states are in white while
these states are extended near ν = N + 21 then ρxx is localized states are in gray.
finite, otherwise the states are localized with ρxx = 0.
This gives a qualitative explanation of the experimental ρxx (ν).
To understand the quantization of ρxy we note first that a change in density near
an integer ν adds (or subtracts) localized states that are not expected to change

11
Figure 7: Longitudinal, and Hall resistance in Integer Hall effect.

the conductance, hence a plateau. This argument, however, does not determine the
value of this plateau. We consider next a situation with ρxx = 0 and evaluate σxy
following an argument by Laughlin (1981).
Recall first the meaning of gauge invariance. Consider j = 1, 2, ..., N electrons
with interaction V, so the Hamiltonian is
X e
H{Ag } = [−i~∇j + A(rj )]2 + V (r1 , ..., rN ) (47)
j
c

The vector potential A(r) can be separated as A(r) = Ap (r) + Ag (r) where Ap is
physical, i.e. ∇ × Ap is the magnetic field in the region where the electrons are
present while Ag is pure gauge, i.e. ∇ × Ag is a magnetic field confined to holes
whereH electrons are excluded. Defining Ag (r) = ∇χ(r) the flux through the hole is
Φ = Ag · d` = ∆χ.
Consider an eigenstate ψ = ψ(r1 , ..., rN ) so that H{Ag }ψ = Eψ, then we can
find an eigenstate for the Ag = 0 system
e P e P e P
H{Ag = 0}ei ~c j χ(ri )
ψ = ei ~c j χ(ri )
H{Ag }ψ = Eei ~c j χ(ri )
ψ (48)

so that Ag is absorbed in a phase. Boundary conditions, e.g. periodic, of ψ do


e
not change when rotating rj for one j around the hole if ei c~ ∆χ = 1, hence Φ =
φ0 ×integer. Thus the gauged ψ is an acceptable solution and the spectrum of
Ag 6= 0 is identical to that of Ag = 0. The ground state E0 (Φ) is therefore periodic,
E0 (Φ + φ0 ) = E0 (Φ).

12
Figure 9: Laughlin’s gedanken-experiment for explaining the integer quantum Hall effect. A con-
stant magnetic field B is perpendicular to the cylinder’s surface, the current I flows around the
cylinder and the voltage V is along the length of the cylinder.

To show quantization of ρxy Laughlin considered a gedanken-experiment, Fig.


(9). A time dependent flux generates an electromotive force that in turn changes
the ground state energy in time dt (recall no dissipation σxx = 0)

1 dΦ dE0
dE0 = I dt ⇒I=c (49)
c dt dΦ
If there are extended states E0 (Φ) is not strictly periodic since an integer number
n of charges e can be transferred between the edges, leading to an energy change of
neV , hence

E0 (Φ + φ0 ) = E0 (Φ) + neV (50)

The average slope gives then a current I = c neV


φ0
, hence

I nec ne2
σxy = = = (51)
V φ0 h
Hence integer plateaus, though the argument does not determine the value of the
integer n. A final comment concerns the possibility of a fractional charge e∗ so that
∗ ∗
I = c neφ0V and σxy = nee
h
, as realized in the fractional quantum Hall, next chapter.

2.4 Edge states


Considerably more insight into the quantum Hall effect is gained by studying effects
of edges and of the contacts. Consider a potential V (y) that increases smoothly at
the y boundaries, confining the electrons to a strip of finite, though large, width Ly .
Choosing the Landau gauge for the vector potential A = (−By, 0) the Hamiltonian
is
1 h e 2 2
i
H= (px − By) + py + V (y) (52)
2m c
We can write the eigenfunctions in the form ψj,k (x, y) = eikx fj (y) so that the equa-
tion for fj (y) is
~2 ∂ 2
 
1 2 2
− + mωc (y − y0 ) + V (y) fj (y) = Efj (y) (53)
2m ∂y 2 2

13
where y0 = k`2 . Far from the edges where V (y) = 0 the Landau levels are Ejk =
~ωc (j + 21 ) and are degenerate in y0 , which is the center of the harmonic oscillator
wavefunction. If V (y) changes in a range ` less than the harmonic energy, i.e.
` dVdy(y)  ~ωc , then one can still define locally harmonic oscillator wavefunctions
with energies Ejk = ~ωc (j + 21 ) + V (y0 ). The energy levels become nondegenerate
near the edges since y0 = k`2 . These levels are plotted schematically in Fig. (10).

Figure 10: Energy spectrum of a 2D conductor in high magnetic field in presence of edge potentials.
The Landau levels are at Ej = ~ωc (j + 12 ) in the bulk. The center position y0 of the eigenstates is
y0 = k`2 .

From the energy dispersion we can calculate the velocity of electrons in the x
direction (a relation well known in band theory)
1 dEj,k 1 dEj,k dy0 `2 dEj,k
vj,k = = = (54)
~ dk ~ dy0 dk ~ dy0
dE dE
As seen from Fig 10 dyj,k 0
> 0 in one edge (large y) while dyj,k 0
< 0 in the other
edge, hence edge state currents are in opposite directions, as illustrated in Fig. (11).
dE
In the bulk dyj,k0
= 0 hence no current. The velocity
Eq. (54) is also consistent with the semiclassical de-
scription where the potential gradient is balanced by
dE
the Lorenz force, dyj,k 0
= dVdy(y) = v eB
c
. The solutions
are known as skipping orbits.
In equilibrium maximal energies occupied in each
Landau level in the two edges are equal, Fig. (10).
However in presence of a Hall voltage these maximal Figure 11: 2D conductor in
energies differ by eV , hence the total current is a perpendicular magnetic field
connected to reservoirs. The
Z 2 X 2 chemical potential of the reser-
e X dk dEjk e Ne
I= = V = V (55) voirs are µ1 and µ2
~ j 2π dk h j h

showing the quantized Hall conductance. These uni-


directional edge states, also known as chiral states, account for the remarkable accu-
racy of the quantized Hall plateaus. To produce backscattering one needs to scatter
electrons between two edges, this scattering is strongly suppressed since the states

14
are localized on a scale `  Ly . The only possible scattering is between states
propagating in the same direction, having the same sign of velocity and therefore
the same current. When the Fermi energy is close to a bulk Landau level (ν is near
half filling) then degenerate levels in the bulk allow backscattering and deviation
from quantization is possible.
The edge state picture is a natural framework for using the Landauer-Büttiker
formalism. We consider first ideal contacts (no reflection). The edge channels are
ideal conductors, hence thePtransmission amplitudes are Tj = 1. In equilibrium
e
R
the edges carry current ± h j Tj (E)fµ̄ (E) where fµ̄ (E) refers to an equilibrium
chemical potential µ̄; the current is to the right in the top edge and to the left in
the lower one (see Fig. 11), the total current vanishes. When applying a voltage the
current in the top (bottom) edge is distributed with its source µ1 (µ2 ), so with the
expansion f1 (E) = f¯(E) − ∂E∂f
(µ1 − µ̄) the excess current I1 (I2 ) in the top (bottom)
edge is
Ne Ne
I1 = (µ1 − µ̄) I2 = (µ2 − µ̄)
h h
Ne
⇒ I = I1 − I2 = (µ1 − µ2 ) (56)
h
2
Hence the 2-terminal conductance is Nhe , exactly as for an ideal wire Eq. (4), except
that here I1 , I2 are separated in space and except for the factor 2 for spin states (here
the spin is polarized by the magnetic field).
Ideal contacts lead to the corner problem, Fig. (6) (right), caused by the con-
tinuity of the equipotential lines across the contacts. The situation of horizontal
equipotential lines in this figure does not necessitate now the condition ωc τ  1, it
is rather the exact situation whenever the quantum Hall effect is realized. To resolve
the corner problem we consider the geometry of Fig. (12) that allows reflections from
the contacts.

Figure 12: Hall bar with realistic contacts. The middle region has N = 2 edge states, the contacts
are connected to reservoirs with a large number of incoming and outgoing states (external R1 , R2
may differ to R10 , R20 ).

The contact resistances lead to chemical potentials at the upper and lower edges
µA and µB that are somewhere between µ1 and µ2 . We aim to evaluate the currents
I1 , I2 in the upper and lower edges, respectively. The method is similar but not
exactly the multi-terminal one since these are not terminal currents, but rather

15
internal ones. The currents are evaluated relative to the equilibrium ones, hence the
chemical potentials appear relative to µ̄.
I1 has two contributions: carriers incident from left of the contact with transmis-
sion T1 contribute a current he T1 (µ1 − µ̄) while carriers coming from the lower chiral
states with reflection R1 contribute he R1 (µB − µ̄). I1 represents also the current of
N chiral states at µA , hence
e e e
I1 = T1 (µ1 − µ̄) + R1 (µB − µ̄) = N (µA − µ̄)
h h h
e e e
I2 = T2 (µ2 − µ̄) + R2 (µA − µ̄) = N (µB − µ̄) (57)
h h h
where I2 has a T2 from the right reservoir and an R2 contribution from reflection
of the upper chiral states. The conservation laws R1 + T1 = N, R2 + T2 = N and
T2 = T20 (transmission from the right reservoir) show that µ̄ cancels, hence we could
choose any reference chemical potential. Yet the choice µ̄ gives a meaningful value
to I1 , I2 themselves, e.g. for an ideal contact R1 , R2 → 0 and µA → µ1 , µB → µ2 ,
Eq. (57) reproduces Eq. (56) with finite I1 , I2 . The chemical potentials of the edges
can be solved,
N T1
µA = µ2 + (µ1 − µ2 )
N2
− R1 R2
R2 T1
µB = µ2 + 2 (µ1 − µ2 )
N − R1 R2
T1 T2
⇒ µA − µB = 2
(µ1 − µ2 ) (58)
N − R1 R2
showing the reduction of the Hall voltage relative to the 2-terminal voltage. The
Hall resistance is seen directly from Eq. (57), with I = I1 − I2 ,

µA − µB h
RH = = (59)
eI N e2
and is quantized even in present of resistive contacts. The 2-terminal resistance is
increased by (µ1 − µ2 )/(µA − µB ),

µ1 − µ2 h N 2 − R1 R2
R2 = = (60)
eI N e2 T1 T2

We can identify the left contact resistance by a 3-terminal measurement µ1 −µ eI


A
=
h R1 µB −µ2 h R2
N e2 T1
and for the left contact eI = N e2 T2 . The total contact resistance is also
found by

h N 2 − R1 R2
   
h R1 R2
Rc = R2 − RH = −1 = + (61)
N e2 T1 T2 N e2 T1 T2

Hence the contact reflections R1 , R2 lead to a finite resistance and dissipation in a 2-


terminal measurement. Yet, the 4-terminal measurement shows the ideal quantized
result.

16
3 Fractional Quantum Hall Effect
3.1 Introduction
In addition to quantized plateaus of the Hall conductance at integer filling factors
in high purity samples additional plateaus where found at fractional filling factors,
such as ν = 13 , 32 , 25 , 27 (Tsui, Stormer and Gossard, 1982). At such plateaus where
2
σxy = ν eh the resistance ρxx vanishes, producing sharp minima in the data, Fig.
(13).

Figure 13: Integer and fractional quantum Hall transport data showing the plateau regions in the
Hall resistance RH and associated dips in the dissipative resistance ρxx . The numbers indicate the
Landu level filling factors, Stomer H.L., Physica B177 (1992) 401.

The fractional quantum Hall effect (FQHE) is an effect of electron-electron


Coulomb interaction. We expect that at low densities, ν . 0.1, a (quantum) Wigner
solid is the ground state, since zero point motion is small and classical Coulomb re-
pulsion yields a periodic solid. It is expected that this electronic solid will be pinned
by impurities so that both σxx , σxy vanish. In the FQHE range we look for a cor-
related liquid that has a gap in the excitation spectrum at fractional ν, in analogy
with the Landau gap in the integer Hall effect. The Hamiltonian that we consider
is
e2 e2
Z X
1X
H = H0 + − ρ0 0
d2 r0 (62)
2 j6=k |~rj − ~rk | j
|~rj − ~r |

where ρ0 is background charge needed for overall neutrality.


We reconsider first the free part H0 in the symmetric gauge A = 12 H(yx̂ − xŷ)
so that
" 2  2 #
1 ∂ y ∂ x
H0 = ~ωc −i` − + −i` + (63)
2 ∂x 2` ∂y 2`

17
using the cyclotron frequency ωc and length `. It is useful to define complex dimen-
sionless variables z and z̄ (note the non-standard signs):
1 1 ` i`
z= (x − iy) z̄ = (x + iy) ; x= (z + z̄) y= (z − z̄) (64)
` ` 2 2
   
∂ ∂x ∂ ∂y ∂ ` ∂ ∂ ∂ ` ∂ ∂
= + = +i = −i (65)
∂z ∂z ∂x ∂z ∂y 2 ∂x ∂y ∂ z̄ 2 ∂x ∂y
∂ ∂
Hence ∂z z̄ = ∂ z̄
z = 0, so that z and z̄ are independent variables. The Hamiltonian
becomes,
 
1 ∂ ∂ 1 ∂ ∂
H0 = ~ωc −4 + z z̄ − z + z̄ (66)
2 ∂z ∂ z̄ 4 ∂z ∂ z̄
Define the following operators
   
† 1 1 ∂ 1 1 ∂
a = √ z̄ − 2 a= √ z+2 ; [a, a† ] = 1
2 2 ∂z 2 2 ∂ z̄
   
† 1 1 ∂ 1 1 ∂
b = √ z−2 b= √ z̄ + 2 ; [b, b† ] = 1 (67)
2 2 ∂ z̄ 2 2 ∂z
where a, b commute [a, b] = [a, b† ] = 0. The Hamiltonian can be finally written as,
 
† 1
H0 = ~ωc a a + (68)
2

The eigenfunctions ψnm (z, z̄) satisfy


 
1
H0 ψmn = ~ωc n + ψmn (69)
2
b† bψmn = mψmn (70)
where m, n ≥ 0 are integers, as usual for harmonic oscillator operators. n represents
the Landau level index while m labels degenerate states within each Landau level.
1
The lowest state ψ00 satisfies aΨ00 = bψ00 = 0 whose solution is ψ00 = Ce− 4 zz̄
where C is a normalization constant. All other eigenstates ψnm can be created by
applying the ladder operators a and b, The state with m, n is then
m † n m  n
b†

a C 1 ∂ 1 ∂ 1
ψmn = 1/2
ψ =
1/2 00 1/2 1/2
z−2 z̄ − 2 e− 4 zz̄
(m!) (n!) (2m m!) (2n n!) 2 ∂ z̄ 2 ∂z
(71)

consistent with n, m ≥ 0. In particular the first Landau level, n = 0 is


1 1 2
ψm,0 = √ z m e− 4 |z| (72)
2π`2 2m m!
The advantage of the symmetric gauge is that it preserves rotation symmetry around
the z axis (while the Landau gauge preserves translation in one direction). In terms

18
of the angle θ in the complex plane z we have ψm0 ∼ eimθ , hence the angular
momentum is
 
∂ ∂ ∂
Lz ψm,0 = −i~ x −y ψm,0 = −i~ ψm,0 = ~mψm,0 (73)
∂y ∂x ∂θ

We consider next the degeneracy of the Landau level, that relates to the number
of allowed angular momenta ~m. Note first that the average area of a state m is
Z

π r m = π d2 r ψm,0

2
r2 ψm,0 = 2π (m + 1) l2 ≤ A (74)

where this area is limited by the sample area A. The allowed values 0 ≤ m ≤ Ns − 1
determine the degeneracy,
A
Ns = (75)
2π`2
Hence the standard result of one state per Landau level per flux quantum. We
conclude that the states at n = 0, up to normalization factors, are
2
1, z, z 2 , ..., z Ns −1 × e−|z| /4

(76)

3.2 Laughlin’s wave function


We study now a trial wavefunction for the interacting system (Laughlin 1983). The
interaction strength is estimated by e2 /` since ` is the range of wavefunctions. Con-
sider a strong magnetic field such that e2 /`  ~ωc , hence the n > 0 Landau levels
can be neglected when constructing the interacting wavefucntion. Also the Zeeman
energy is large, so that we use n = 0 spin polarized states. Consider the following
requirements:

1. ψ (z1 , . . . , zN ) is comprised solely of one-particle wavefuntionsPin lowest Lan-


mN − 1 i |zi |2
dau level, i.e. it is a linear combination of z1m1 z2m2 · · · zN e 4 , where all
mi are non-negative integers.

2. ψ is totally antisymmetric.
1 P 2
e− 4 i |zi | is eigenstate of Lz , therefore f (z1 , . . . , zN )
3. ψ (z1 , . . . , zN ) = f (z1 , . . . , zN )P
is a polynomial of degree M : N i=1 mi = M .

Consider first 2 electrons. The general form satisfying the requirements above is
2
1
+|z2 |2 )
ψmm0 (z1 , z2 ) = (z1 − z2 )m (z1 + z2 )m e− 4 (|z1 |
0
(77)

where m > 0 is odd, to satisfy antisymmetry and m0 ≥ 0 is integer to satisfy having


states in the lowest Landau level. It is remarkable that we do not need to solve a
radial equation – this solution is enforced by the 3 conditions and is exact for any
central potential. It is also remarkable that the states are bound even in presence
of repulsive interactions – there is no kinetic energy to be gained by separating the

19
0
particles. The center of mass dependence (z1 + z2 )m does not affect the interaction
and can be ignored.
Laughlin solved the N = 3 problem which led him to suggest a wavefunction for
N electrons, known as a Jastrow form,
N
Y 1 P 2
ψ(z1 , ..., zN ) = f (zj − zk ) e− 4 i |zi | (78)
j<k

We require that f (z) → 0 as |z| → 0 reducing the amplitude of pairs to approach


(this neglects the approach of 3 electrons or more to a point, which is less probable).
The requirements 1-3 determine f (z) = z m with m odd, leading to the celebrated
result,
N
Y 1 P 2
ψm (z1 , ..., zN ) = (zj − zk )m e− 4 i |zi | (79)
j<k

Adding all powers gives the total angular momentum M = 12 (N − 1)N m. This
wavefunction turns out to be an excellent approximation for some fractional fillings
and for a large variety of repulsive interactions.
Increase of m leads to a lower energy since the repulsion energy becomes weaker.
There is, however, a limit on how large can m be. To find this limit consider one
coordinate, e.g. z1 , and restrict its maximal power by Ns − 1, so that πhr12 i ≤ A.
The highest power of z1 in the product

(z1 − z2 )m (z1 − z3 )m · · · (z1 − zN )m (80)

is m (N − 1). Equating this to Ns − 1 gives, for N, Ns  1


N 1
ν= = (81)
Ns m
hence m is fixed by the filling factor. This remarkable result is independent of the
details of the interaction. For m odd, i.e. m = 3, 5, 7, ... this wavefunction describes
the 1/m filling fraction cases.
For m = 1 the filling factor is ν = 1 and the wavefunction (79) recovers the
expected non-interacting Slater determinant, known also as a Vandermonde poly-
nomial:
N
Y N (N −1) X
0 1 2 N −1
(zj − zk ) = (−1) 2 sgn(σ) × zσ(1) zσ(2) zσ(3) · · · zσ(N ) (82)
j<k σ

where σ (i) is a permutation of the N electrons. For example, the case of N = 3,


(z1 )0 (z2 )0 (z3 )0

Y3
(z1 )1 (z2 )1 (z3 )1 = z2 z32 − z3 z22 − z1 z32 + z3 z12 + z1 z22 − z2 z12 = −

(zi − zj )
(z )2 (z )2 (z )2 i<j
1 2 3

20
We proceed to study another method for deriving ν = 1/m. Consider the norm
of the wave function Eq. (79) and write it in the form |ψ (z1 , ...zN )|2 = e−βφ(z1 ,...,zN ) ,
so that
X 1X 2
βφ (z1 , ...zN ) = −2m ln |zj − zk | + |zi | (83)
j<k
2 i

The potential φ(z1 , ..., zN ) is the classical potential energy of N point charges in
2D with a uniform background, the so called one-component plasma. To show this
it is convenient to choose β = m and identify the charge ρ(r) that generates the
potential φi (r) = −2 ln |r − ri |. Poisson’s equation and its area integral yield
 
2 1 ∂ ∂
∇ (ln r) = r ln r = 0 r 6= 0
r ∂r ∂r
Z I I
2 ∂ ln r
∇ (∇ ln r) d r = ∇θ (ln r) · rdθ = rdθ = 2π
∂r
⇒ −∇2 φi (r) = 4πδ(r − ri ) = 4πρi (r) (84)
i.e φi (r) is generated by a unit charge localized at ri . The second term is a back-
1 2 2
ground charge density: φbg (r) = 2m r /l ,
 
2 1 1 ∂ ∂ 2 2 1
−∇ φbg (r) = − 2
r r = − 2 = 4πρbg =⇒ ρbg = − (85)
2ml r ∂r ∂r ml 2πml2
The plasma has to be charge neutral, otherwise the long range forces produce a
divergence and the probability e−βφ = |ψ|2 vanishes. Hence the total background
charge has to balance the number of point charges
A Ns 1
N= 2
= ⇒ν= (86)
2πm` m m
1
Hence our wavefunction Eq. (79) must correspond to the filling fraction ν = m
.

Hard core model:


To see the effect
R of interactions consider the matrix elements between 2-particle
states Eq. (77) |ψmm0 (z)|2 V (z)d2 z = Vm where z = z1 − z2 is the relative position.
Since V (z) is repulsive Vm > 0 and decreases with m. In the state (79) each 2
particles have relative angular momentum equal to m or larger (e.g. (z1 − z2 )m
multiplies many other factors with z1 , z2 ). Consider first a model with Vm0 = 0 for
all m0 ≥ m, but Vm−2 6= 0, therefore hψm |V |ψm i = 0. Imagine now increasing ν so
that it crosses m1 , then at least one pair must have a smaller angular momentum
m − 2 to be compatible with the given area, hence the interaction average is at least
Vm−2 . More generally, for Vm with all components m the interaction average for
ν < m1 is ∼ Vm while it is at least Vm−2 for ν > m1 . Hence at ν = m1 there is gap
∼ Vm−2 − Vm for adding one particle. E.g. to cross 13 at least one pair has (zi − zj )1
with interaction energy V1 . The chemical potential and the gaps are illustrated in
Fig. 14.
The gap is essential for having dissipationless current, ρxx = 0. It also account
e2
for the flatness of σxy = mh near ν = m1 since a few added or subtracted particles
are localized by disorder and do not change σxy .

21
1/7 1/5 1/3
Figure 14: The gap in the chemical potential due to the gap in the potential when ν crosses 1/m.

3.3 Excitations
Pierce the plane with flux ±φ0 = ±hc/e, the ground state evolves then to an excited
state. Aθ = ± hc/e
H
2πr
corresponds to this flux since A · dl = ±hc/e The effect on a 1
particle wave function is
 
 e  1 ∂ ~  
~ e±iθ ψ (r, θ) (87)
−i~∇ + A ψ (r, θ) = −i~ ± ψ (r, θ) = −i~∇
c θ r ∂θ r θ

Since eiθ ∼ z we propose that a quasi hole with charge |e|/m is described by:
Y
Sz0 |mi = (zi − z0 ) |mi (88)
i

|mi is the ground state when ν = 1/m.


Proof I : Count the maximal power of z1 to determine the radius of the system:
|mi : m(N − 1) =⇒ πR2 = 2π [m (N − 1) + 1] `2
Sz0 |mi : m (N − 1) + 1 =⇒ π (R + δR)2 = πR2 + 2π`2
1  2 2

change in charge= π (R + δR) − πR = 1/m (89)
2πm`2
using the density 1/2πm`2 of the |mi state. Hence a charge e/m (e < 0) is spilled
out of the boundary at R leading to a charge deficit e/m near z0 , i.e. a hole with
positive charge.
Proof I’ : This is also power counting, but instead of area changes we assume
now that the area is fixed but formally the total charge changes, N → N + δN . The
maximal power of z1 in Sz0 |mi is m (N − 1 + δN ) + 1, to keep the same area this
power has to equal to that of the power in |mi, i.e. m(N − 1). Hence δN = − m1 .
Proof II : Adding a time dependent flux, such that the total flux change is φ0
produces an electric field. The flux change is done slowly on the time scale of ~/gap
so that |mi is maintained without dissipation.
−1 dφ
I
dφ 1
=− E · dl =⇒ Eθ = (90)
dt c 2πcr dt

22
The Hall conductivity produces current perpendicular to the field:
e2 −1 dφ
jr = σxy Eθ = ν (91)
h 2πcr dt
The transported charge across a circle of radius r due to this current is
e2 e2
Z Z

Qout = jr · 2πrdt = −ν dt = −ν · φ0 = −ν|e| (92)
hc dt hc
hence a deficit of charge νe, i.e. positive charge inside that radius, near z0 .
0
Proof III : Consider the 2D plasma analog with |Sz0 |mi|2 = e−βφ where
X X X
βφ0 = −2m ln |zj − zk | + 21 |zi |2 − 2 ln |zi − z0 | (93)
i<j i i

We note an extra ”phantom” charge at z0 , with charge 1/m times smaller P than the
2
plasma charges. The background charge due to the physical particles ∼ i |zi | does
not change, hence the neutrality condition is that charge of particles at zi (which
is N ) plus the charge at z0 (which is m1 ) = Aρbg = 2πm` A
2 . Hence the number of
1
physical charges at zi is reduced by m , i.e. we have a quasi hole with charge |e|/m.
We finally note that for a quasi - particle, rather that quasi-holes, we cannot use
1
z
∼ z̄, because this produces states in the second landau level. Here we just note
that one uses
Y ∂ Y
− |zi |2 /4
(zj − zk )m ] .
P

Sz0 |mi = e 2 − z̄0 (94)
i
∂z i
j<k

3.4 Two holes and fractional statistics


Considering two holes, we propose that their wavefunction has the form
1 1 2 +|z 2)
|zA , zB i = (zA − zB ) m e− 4m (|zA | B|
SzA SzB |mi (95)
The reason for the unusual prefactor comes from the 2D plasma analogy. Defining
00
hzA zB |zA zB i = e−βφ where
X X X
βφ00 = −2m ln |zj − zk | − 2 ln |zi − zA | − 2 ln |zi − zB |
j<k i i
2 1 X
− ln |zA − zB | + [|zA |2 + |zB |2 ] + 1
2 |zi |2 (96)
m 2m i
1
Choosing as above β = m the phantom charges are and therefore need additional
m
1
background charge from the terms 2m [|zA |2 + |zB | ]. Furthermore, the phantom
2

charges should interact with each other as − m22 ln |zA − zB |, producing the prefactor
of Eq. (95). Since the plasma is fully neutral the charges at zA , zB are allowed to
separate, creating two independent quasi-holes.
The quasi-holes exhibit unusual quantum statistics. We recall that exchanging
two particles is equivalent to rotating one around the other by π, plus a translation
that is irrelevant. The exchange produces a new wavefunction
ψ(zA , zB ) = eiθs ψ(zB , zA ) (97)

23
where θs is called the statistical phase. Performing the exchange twice is equivalent
to a 2π rotation, producing

ψ̃(zA , zB ) = e2iθs ψ(zA , zB ) (98)

In 3D a circle on a sphere can be deformed continuously to a point, hence ψ̃ = ψ


and this operation is equivalent to a unity. The two solutions, θs = 0, π, correspond
to the well known bosons and fermions. In 2D, however, one cannot shrink the circle
since one of the particles is a singularity inside the circle and other statistics are
possible. In particular, our two holes have a factor
1 1 2π rotation 1
(zA − zB ) m = |zA − zB | m eiθ/m −−−−−−−−→ e2πi/m |zA − zB | m eiθ/m (99)

where θ is the relative angle of the two particles. Hence the statistical phase is
π
θs = m , i.e. fractional statistics.

3.5 Composite Fermions


We study here a method for generating additional fractions showing FQHE. Recall
A BA
first that The number of states at each Landau level NS = 2πl 2 = φ , i.e. the total
0
flux in units of φ0 . The filling fraction ν satisfies
Ns AB No. of flux quanta
ν −1 = = = (100)
N N φ0 No. of electrons
We construct now a ”composite electrons” by attaching two flux quanta to each
electron. The composite electron is also a fermion since by their exchange (equivalent
to π rotation) the additional flux yields phase factor ei(2φ0 )π/φ0 = e2πi = 1. Fig. 15a
shows the original electron system, where 2 flux lines symbolize a uniform magnetic
field. The transformation to composite electrons is then shown in Fig. 15b. Now
comes an approximation, i.e. the magnetic flux of the added flux lines are averaged
into a uniform magnetic field, a mean field process, as shown in Fig. 15c. It is
assumed that the averaging keeps a finite gap in the excitation spectrum so that the
qualitative FQHE properties are not changed. The new filling fraction is, just by
counting flux lines per electron, is ν ∗ , where
1 1 ν

= +2 ⇒ ν∗ = (101)
ν ν 1 + 2ν
In particular consider the quantum Hall state with ν = n integer. Add 2m flux
quanta to each electron (m is integer), then spreading the flux into a uniform field
yields
1 1 n

= + 2m ⇒ ν∗ = (102)
ν n 1 + 2mn
Hence the gap ~ωc is expected to become an interaction induced gap during the
mapping. This remarkably simple process can be appreciated in Fig. 15. Fig. 15a
shown 4 electrons with 2 flux lines (representing a uniform field), hence an integer
quantum Hall with n = 2. Adding 2 flux lines (m = 1) to each electron is adding a

24
Figure 15: The composite electron transformation. (a) Electrons in a uniform magnetic field,
symbolized by 2 flux lines. (b) Composite electrons with 2 flux lines attached to each electron. (c)
Spreading the magnetic flux to a uniform field, symbolized by the 10 flux lines.

total of 8 flux lines, so after spreading to a uniform field the total is 10 flux lines.
4
The filling fraction is then 10 = 52 .
A phase factor of 2π under a π rotation is also generated by the operation of
Y
D= (zj − zk )2 (103)
j<k

Starting from a full n Landau levels with a Slater determinant Φn leads to a trial
wave function χν ∗ = Dm Φn which can be tested numerically; this is in fact a test of
the mean field assumption. In particular Dm Φ0 is Laughlin’s state Ψ2m+1 , Eq. (79).
The use of higher n > 1 Landau levels implies higher energy, so we need to
calculate the relative fraction of such states. Consider n = 1 states η1,s defined here
as generated from the n = 0 states η0,s+1 states by the operator a† = √12 ( 12 z̄ − 2 ∂z

)
zs − zz ∗
η0,s = 1 e
4

(2π2s s!) 2
† z s (2s + 2 − z̄z) z z̄
η1,s = a η0,s+1 = 1 e− 4 (104)
(2π2s+2 (s + 1)!) 2

Applying Dm , involves a factor of z q , where q ∼ 2m(N − 1) is large. Consider its


effect on n = 1 states which can be written as
s
q 2q (s + q)! h 1
i
z · η1,s = (s + q + 1) η1,s+q − qη0,s+q
2 (105)
(s + 1)!

hence for large q η0 dominates. In numerical studies one can use alternatively the
state χν ∗ projected on η0,s .

A hierarchy of composite electron states can be obtained with the following


operations:
• D: attaching 2 flux lines to each electron. Dm shifts the filling fraction by
ν → ν ∗ = 2mν+1
ν
.

• C: particle-hole conjugation with respect to a completely occupied lowest


Landau level. This causes ν → 1 − ν.

• L: addition of a filled Landau level. This causes ν → ν + 1

25
A table showing the first levels of the hierarchy is given by Jain and Goldman,
PRB 45, 1255 (1992) and is reproduced in Fig. 16. The authors note that these
operators generate all odd denominator fractions and that each fraction is obtained
by applying a unique sequence of the above operators on an integer ν. The key
question is which of these states are indeed realized in experiment, and what is their
stability related to their excitation gap, the larger the gap the more stable it is. It
4
was suggested that the L operation can be ignored, yet 11 obtained by DLDΦ1 and
7
11
obtained by CDLDΦ1 are actually observed.

Figure 16: FQHE hierarchy.

3.6 FQHE edge states


In the integer quantum Hall we saw in Fig. 10 that the

Fermi level crosses various Landau levels at distinct


points and therefore the density decreases roughly as
shown in Fig. 17. However, since the Hall current
changes with ν, the current difference must occupy a
finite range, i.e. not steps as in this figure. In the
integer case these are edge channels of Fig. 11.
We note that the Coulomb interaction favours Figure 17: The electron density
near the edge. The filling ν3 = 3
small changes in density, however the actual filling
disappears first, then ν2 = 2 and
fractions near the edge is a complex issue. Here so on.
we just assume a given change ∆ν and explore its
consequences, allowing also fractional ν values (see
Beenakker and van Houten, arXiv:cond-mat/0412664, section IVC).
We evaluate the edge current (in the x direction)
due to a Hall voltage ∆µ/e (in the y direction) at
the interface between the adjacent quantized ν states.
The density depends on the difference µ − φ becoming
smaller near the edge, where φ(y) is the edge potential,
see Fig. 18. Therefore
    Figure 18: The edge potential.
δn δn
∆n = · ∆µ = − · ∆µ (106)
δµ φ δφ µ

26
This density change moves in the x direction with a drift velocity, as determined by
balancing the Lorentz force with the force −∂φ/∂y,

c ∂φ
vd = − · (107)
|e|B ∂y

This corresponds to a current density


 
δn c ∂φ |e| ∂ν
j(y) = −|e|∆n · vd = −|e| · ∆µ · · = − ∆µ · (108)
δφ µ |e|B ∂y h ∂y
hc
since ν = 2πl2 n = eB · n. Hence the total current density and the conductance are

|e|
Z
I∆ν = j(y)dy = − ∆µ · ∆ν
h
2
eI e
G = = · ∆ν (109)
∆µ h

Figure 19: Edge currents between quantized ν states.

We see that the current is confined to the range where the filling factor changes,
as illustrated in Fig. 19. The whole edge current is, summing up all differences ∆ν
is
X |e|
I= I∆ν = − ∆µ · ν (110)
∆ν
h

27
4 Quantum Dots
In previous chapters we studied transport properties where the electron-electron
Coulomb interaction was neglected. We consider now the other extreme situation
where the Coulomb interactions dominate. A typical system in Fig. 20 shows a
quantum dot (QD) coupled by various capacitances to the environment. It has a
current source and drain with weak tunneling that probes the properties of the QD.

Figure 20: A typical setup for a quantum dot. The gate has only capacitive coupling Cg while the
current source and drain have capacitances CS , CD as well as large resistances RS , RD , respectively.

e 2
The charging energy is defined as Ec = 2C with a capacitance C, Ec is typically
∼meV. Focusing now on the role of one gate voltage the Hamiltonian describing the
coulomb interaction in the quantum dot is
 2
2
H = Ec N̂ − αN̂ |e| Vg = Ec N̂ − N0 + const. (111)

where N̂ is the electron number operator, N0 = α|e|V 2Ec


g
and α is a geometric factor
known as lever arm. Tunneling to the electrodes is neglected so that in the ground
state N̂ is well defined and has eigenvalues at the integer nearest to N0 . Therefore,
as Vg increases the charge on the QD increases by steps of one unit at the half
way positions, e.g. at N0 = 21 , 32 , 52 , .... At these values of Vg weak tunneling from
the electrodes allows for charge fluctuations and for a finite conduction. Otherwise
charge transport has to overcome a large Coulomb barrier and the conductance
vanishes at low temperatures. This phenomena of Coulomb blockade is shown in
Fig. 21. To estimate the resistance needed to see sharp peaks in the conduction,
consider each electrode as an RC circuit whose charging time is ∆t = RC. To
2 2
resolve the charging energy of e2 /C we need eC ∆t > h, hence R > eh . The sharpness
of the conductance peak is therefore due to a large resistance, i.e. weak tunneling.
An additional condition is low temperatures kB T  Ec . The sharpness of the
Coulomb blockade peaks allowed the development of a device known as ”single
electron transistor” – by tuning the system to be near a conductance peak the
conductance becomes an extremely sensitive measure of Vg .

28
Figure 21: Conductance of a quantum dot versus the gate voltage ; the conductance peak locations
are periodic in the gate voltage (Lindemann et al., 2002).

The QD energy involves also single particle energies, e.g. kinetic energy. For N
~2 k2
electrons in a 2D QD of radius r and Fermi energy EF = 2mF∗

ZkF ∗
2 d2 k 2 m
N = 2πr = πr EF (112)
(2π)2 π~2
0

dN m∗
so that the density of states dE
= πr2 π~ 2 is independent of the energy. The kinetic

confinement energy is then

ZEF
dN m∗ 2 ~2
Ekin = E dE = πr2 2
E F = ∗ 2
N2 (113)
dE 2π~ 2m r
0

2 2 2
The Coulomb energy dominates therefore if m~∗ r2 < 2Ec ≈ er so that r > m~∗ e2 = a∗B ,
i.e. the QD is not too small. E.g., in a QD made of GaAs, r ≈ 100nm and
a∗B ≈ 10nm the coulomb energy is dominant, while in a InAs QD r ≈ 20nm and
a∗B ≈ 30nm the kinetic energy is dominant.

4.1 Coulomb blockade


We proceed to study the Coulomb dominated case, as defined by Eq. (111). The
electrochemical potential is defined as the energy required for adding one electron
to the QD

µN (Vg ) = EN (Vg ) − EN −1 (Vg ) = 2Ec (N + 1) − α|e|Vg (114)

In linear response when µS → µD conduction is possible only when µN is very close


to these external chemical potentials, as illustrated in Fig. 22; define VcN as the gate
voltage at which µS = µD = µN (VcN ). For other Vg values

µN (Vg ) = µN (VcN ) − |e|α(Vg − VcN ) (115)

29
Increase now Vg to reach the next conduction peak at VcN +1

µS = µD = µN +1 (VcN +1 ) = µN +1 (VcN ) − |e|α(VcN +1 − VcN )


µN +1 (VcN ) − µN (VcN ) 2Ec
⇒ ∆Vc = VcN +1 − VcN = = (116)
|e| α |e|α
giving the spacing between conductance peaks.

Figure 22: schematic drawing of the electrochemical potentials. (a) Coulomb blockade situation,
(b) conductance peak.

Consider next a finite bias voltage between the source and drain, µS − µD =
− |e| VSD 6= 0. Conduction is possible if electrons can lower their energy (which is
dissipated) when passing the QD, as in Fig. 23, i.e.

µS ≥ µN (Vg ) ≥ µD (117)

leading to a window ∆Vg where the system exhibits a finite conductance.

Figure 23: Finite conduction region due to bias voltage VSD

(1) (2)
This window is at Vg < Vg < VG , see Fig. 24, where

µN = 2Ec (N + 1) − α|e|Vg(1) = µS , µN = 2Ec (N + 1) − α|e|Vg(2) = µD


µS − µD VSD
⇒ ∆Vg = Vg(1) − Vg(2) = − = (118)
α|e| α
The crossing point labeled 1 in the left Fig. 24 is the solution of

µN = 2Ec (N + 1) − α|e|Vg = µD
2Ec
µN +1 = 2Ec (N + 2) − α|e|Vg = µS ⇒ VSD = − (119)
|e|

30
and similarly the crossing 2 has VSD = 2E
|e|
c
. The insulating diamonds have therefore
unequal digaonals, as shown in the right Fig. 24. To evaluate the slopes β, γ, where
in general β 6= γ, we need a more detailed model, as we consider next.

Figure 24: Coulomb blockade diamonds. Left: General form with conducting regions shaded.
Right: Parameters of the insulating diamond.

4.2 The capacitance model


Consider a quantum dot with n electrodes. The charge on each conducting element
is determined by relative capacitances Cij
n
X (0)
Qi = Cij Vj + Qi (120)
j=0

(0)
where i = 0 is the dot index and i = 1, 2, ..., n denote the gates, Qi are reference
charges at Vj = 0. If we shift all Vj → Vj +constant then the charges should not
Pn Pn
change, hence Cij = 0, and in particular C00 = − C0j > 0. By rearranging
j=0 j=1
Eq. (120)
(0) n
Q0 − Q0 X C0j
V0 (Q0 ) = − Vj . (121)
C00 j=1
C 00

The change in the charging energy after adding N electrons is


(0)
Q0 Z−|e|N
n
e2 N 2 X C0j
E(N ) = V0 (Q0 )dQ0 = + |e| N Vj (122)
2C00 j=1
C00
(0)
Q0

(0)
where Q0 is the initial charge on the dot. Therefore, the electrochemical potential
of the dot is
n
e2
 
1 X C0j
µN = E (N ) − E (N − 1) = N− + |e| Vj (123)
C00 2 j=1
C 00

31
Our previous treatment considered varying just one gate, e.g. j = 1. Comparison
with Eq. (114) we identify

e2 C01
Ec = , α=− (124)
2C00 C00

The general shape of the diamonds is then maintained, while the slopes β, γ in the
right Fig. 24 can be identified using the other capacitances in the system. E.g. the
model of Fig. 20 with C00 = C, C01 = −CG , C02 = −CS , C03 = −CD is left as an
exercise.

32
5 Graphene
5.1 Band structure
Graphene is an allotrope of carbon in the form of a two-dimensional honeycomb
lattice, see Fig. 25 (left); the unit cell includes two atoms. Each Carbon has four
electrons, one electron is bonded with with each of the three neighbors and one
electron is unpaired, forming the band of interest. We choose the unit vectors, with
a the nearest neighbor distance, as
a √ a √
a1 = (3, 3), a2 = (3, − 3) (125)
2 2
The reciprocal lattice is hexagonal, shown in the right Fig. 25. Its unit vectors,

Figure 25: The graphene lattice (left) and its reciprocal space (right).

satisfying bj · ai = 2πδij are


2π √ 2π √
b1 = (1, 3), b2 = (1, − 3) (126)
3a 3a
We choose two special vectors, K and K0 , as
2π 1 2 1
K = (1, √ ) = b1 + b2
3a 3 3 3
2π 1 1 2
K0 = (1, − √ ) = b1 + b2 (127)
3a 3 3 3
K and K0 are inequivalent, but related by time reversal, i.e. under time reversal
K → −K, while −K is equivalent to −K + b1 + b2 = K0 .
We find the electron dispersion by the method of tight binding. Consider local-
ized orbitals on the 2 atoms A, B in Fig. 25, φA (r) and φB (r), respectively. Form
Bloch states by summing on lattice vectors R
1 X
ψk,A (r) = √ φA (r − R)eik·R
N R
1 X
ψk,B (r) = √ φB (r − R)eik·R (128)
N R

33
where N is the number of unit cells; the normalization neglects the overlap of orbitals
on different unit cells. These Bloch states satisfy e.g. ψA (r + a1 ) = eik·a1 ψA (r).
In terms of the of the Hamiltonian overlap t between the neighbors A, B the
Hamiltonian matrix element involves overlap with 3 nearest neighbors,
Z
t = φ∗A (r)HφB (r)
Z r

ψk,A (r)Hψk,B (r) = t[1 + eik·a1 + eik·a2 ] (129)
r

In the A, B basis the Hamiltonian in reciprocal space is then a 2 × 2 matrix


 
0 1 + eik·a1 + eik·a2
h(k) = t (130)
1 + e−ik·a1 + e−ik·a2 0

with eigenvalues

Ek = ±|1 + eik·a1 + eik·a2 | (131)


2 1 4π 2π
In particular for k = K we have eiK·ai = ei 3 b1 ·ai +i 3 b2 ·ai which is ei 3 for a1 and ei 3
4π 2π
for a2 , hence EK = |1 + ei 3 + ei 3 | = 0 and similarly EK0 = 0.
Expanding near K for small q where k = K + q we obtain Eq = v(qy σx − qx σy )
with v = 23 ta and σi are the Pauli matrices. We relabel qx → −qy , qy → qx so that

h(K + q) = v(qx σx + qy σy )
h(K0 + q) = h(−K + q) = h∗ (K − q) = v(−qx σx + qy σy )
Eq = ±v|q| = ±vq (132)

Hence a 2D Dirac Hamiltonian with with linear dispersion, except that the light
velocity c in Dirac’s equation is replaced by a Fermi velocity which in Graphene is
v/c ≈ 1/300. For neutral Graphene the Fermi level is at E = 0, the Dirac point, and
the system is a semi-metal, i.e. the Fermi surface has just 2 points. The full spectra
of Eq. (131) is shown in Fig. 26 with the region near the Dirac point expanded.

Figure 26: Band structure of graphene with the region near a Dirac point enlarged.

34
5.2 Symmetries
We consider next symmetries of Graphene and their significance. Time reversal T
involves complex conjugations while space indices are not affected T hij T −1 = hij .
To identify the effect in reciprocal space consider
X X X
T eik·(Ri −Rj ) h(k)T −1 = e−ik·(Ri −Rj ) h∗ (k) = eik·(Ri −Rj ) h(k)
k k k
(−1) ∗
⇒ T h(k)T = h (−k) = h(k) (133)

Space inversion P with respect to the center of the hexagon in the left Fig. 25
involves reflecting the unit cell coordinate Ri → −Ri as well as interchanging the
sites A, B. The latter operation is represented by σx , hence
X X X
P eik·(Ri −Rj ) h(k)P −1 = e−ik·(Ri −Rj ) σx h(k)σx = eik·(Ri −Rj ) h(k)
k k k
(−1)
⇒ P h(k)P = σx h(−k)σx = h(k) (134)

Combining both T, P symmetries lead to h(k) = σx h(−k)σx = σx h∗ (k)σx which is


a local k symmetry. This local symmetry is a sensitive test for the presence of a
Dirac point, i.e. a gapless linear spectrum. Considering e.g. the K vicinity, the
only way to producepa gap is by adding to h(K + q) of Eq. (132) a mass term mσz
leading to Eq = ± v 2 q 2 + m2 . This term, however, does not commute with the
combined P, T operation, i.e. σx (mσz )∗ σx = −mσz . We conclude then that both
T, P symmetries are needed for having a Dirac point.
A more compact notation is to use a 4-spinor state
 K 
ψA,q
 ψB,qK 
ψq = 
 ψA,qK0  ,
 h(q) = v(qx σx τz + qy σy ) (135)
K0
ψB,q

where the Pauli matrices τi act in the K, K0 space and the τz factor acounts for the
sign change of qx in the Hamiltonian near K0 , Eq. (132). To identify the symmetry
operations in this notation we note that for time reversal T h(K+q)T −1 = h∗ (K0 −q)
since −K is equivalent to K0 . Hence time reversal mixes K, K0 and therefore for the
4 × 4 Hamiltonian

T h(q)T −1 = τx h∗ (−q)τx (136)

For parity we note P h(K + q)P −1 = σx h(K0 − q)σx , hence P also mixes K, K0

P h(q)P −1 = τx σx h(−q)σx τx (137)

It is easy to check that the 4 × 4 Graphene Hamiltonian, Eq. (135), is indeed


invariant under both T, P operations.
To appreciate the role of symmetry, consider a system that has different atoms
on the sites A, B, e.g. as realized by Boron Nitride (BN). The different site energies

35
Table 1: Effect of mass terms on symmetries. + is maintained symmetry, − is broken symmetry.

type of mass T P realization


mσz + − BN
mσz τz − + Haldane’s model
mσz τz Sz + + spin-orbit

lead to the term mσz in the Hamiltonian which is the same for both K, K0 , hence
the Dirac spectrum acquires a mass term
h(q) = v(qx σx τz + qy σy ) + mσz BN lattice
p
⇒ E q = ± v 2 q 2 + m2 (138)
In this case T symmetry is still valid, but not P symmetry, as is also obvious from
the lattice figure, Fig. 25, with inequivalent A, B atoms. Hence, as also noted
above, both T, P symmetries are essential for having a Dirac point. Therefore,
small deformations of the Graphene lattice that maintain these symmetries keep the
Dirac point. E.g. shifting qx , qy → qx + a1 , qy + a2 maintains the symmetries and
the Dirac point, which is merely shifted. This corresponds to change in the overlap
integrals between the 3 nearest neighbors, i.e. t1 6= t2 6= t3 so that the 3-fold rotation
symmetry is broken, or one can add 2nd or 3rd nearest neighbor overlaps. Yet, large
perturbations may cause 2 Dirac points to annihilate each other, e.g. if t1 = t2 = 0
but t3 6= 0 these are isolated dimers and Eq = ±t3 are flat bands.
Other options of producing gaps are collected table 1. Haldane’s model (1988)
has opposite sign mass terms at the K, K0 points, hence in a full band model it is
a k dependent term. The term mσz τz Sz involves the actual spin operator Sz and
describes the effect of spin-orbit coupling in Graphene (Kane and Mele, 2005). Time
reversal involves now also spin rotation iSy and T symmetry is restored (the actual
spin was implicit so far). This case shows that while T, P symmetry is a necessary
condition for a Dirac point, it is not a sufficient one.

5.3 Klein paradox


An important implication of a relativistic-like dispersion relation for fermions, is in
the question of scattering. Shortly after the publication of Dirac’s equation, Klein
studied its scattering and found unusual results. For our 2D Dirac electrons we
consider the Hamiltonian,
H = −ivσ · ∇ + V (x)

V0 , 0≤x≤D
V (x) = (139)
0, else
In a potential free regions we need momentum eigenfunctions of vσ · k which we
choose as
 
1 1
ψ(k) = √ iθ (140)
2 ±e k

36
 
where θk ≡ tan−1 kkxy and the ± sign refers to the energy eigenvalue ±v|k|. [This
choice has a Berry phase π upon rotation instead of the sign change embedded in
the other common choice multiplying Eq. (140) by a global phase factor of eiθk /2 ].
We further assume that the scattering does not mix momenta near K and K0 . We
show In Fig. 27 the three regions which define the scattering solution,

Figure 27: Scattering process in graphene. Top: The potential barrier parameters and the particle’s
energy E. Bottom: scattering angles used for identifying the wavefunctions in the regions I, II, III.


   
√1
1 i(kx x+ky y) 1 i(−kx x+ky y)
ψI (r) = iφ e +r i(π−φ) e
 se  se
2
 
1 1 i(q x+k y) 1 i(−q x+k y)
ψII (r) = √2 a 0 iθ e x y
+ b 0 i(π−θ) e x y
(141)
se   se
1
ψIII (r) = √t2 ei(kx x+ky y)
seiφ
Here θ = tan−1 (ky /qx ) , φ = tan−1 (ky /kx ), the energy is E = v kx2 + ky2 =
p
p
v qx2 + ky2 + V0 , qx > 0 is chosen and assumed real (i.e. incidence angle φ not
too close to π/2) and s =sgn(E), s0 =sgn(E − V0 ). Backscattering conserves ky but
reverses the sign of kx or qx , accounting for the π factor in the phases.
The coefficients r, a, b, t are determined by continuity of the wave function ψI (x =
0, y) = ψII (x = 0, y), ψII (x = D, y) = ψIII (x = D, y). Since the Dirac equation
is first-order in space, there is no need to match derivatives as in Schrödinger’s
equation. The solution for the transmission coefficient is [Castro Neto et al. (2009)]
cos2 θ cos2 φ
T = |t|2 = (142)
[cos(Dqx ) cos φ cos θ]2 + sin2 (Dqx )(1 − ss0 sin φ sin θ)2
We see that for normal incidence, φ = θ = 0, there is complete transmission T = 1;
this never happens for non-relativistic electrons. Furthermore, for Dqx = nπ with

37
n integer again T = 1, i.e. cases with a finite scattering angle. Thus, even for a
barrier of height much greater than the electron energy, the transmittance for normal
incidence is unity, i.e. the barrier becomes completely transparent. This unexpected
result, contradicting both classical and non-relativistic quantum mechanics, is known
as the Klein paradox.
Another peculiar phenomena of Dirac’s equation is known as Zitterbewegung, a
jittery motion that occurs when trying to confine a Dirac electron. Localization is
space implies uncertainty in momentum and therefore uncertainty in energy that
may cause a particle-like state to evolve in time into hole-like state and back. These
fluctuations have well known manifestations in quantum field theories, and possibly
also in Graphene.

5.4 The Quantum Hall Effect in Graphene


We proceed now to study Graphene in a magnetic field, in particular we aim at
understanding the spectacular data on the Hall quantization, shown in Fig. 28.
The cyclotron energy turns out to be ∼ 1000K, so that quantum Hall effect can be
seen even at room temperature.

Figure 28: Hall plateaus in σxy and longitudinal resistivity ρxx as function of electron density
relative to the Dirac point.

~ = B(−y, 0).
Consider a magnetic field in the z direction, and Landau’s gauge A
~ hence
The canonical momentum undergoes the usual transformation: pµ 7→ pµ + ec A,
the Hamiltonian for Graphene near one Dirac point becomes
e
vσ · (−i~∇ + A)ψ(x, y) = Eψ(x, y) (143)
c
The wavefunctions can be chosen as ψ(x, y) = eikx φ(y)
∂y − k + ~ce By
 
0
−~v φ(y) = Eφ(y) (144)
−∂y − k + ~ce By

38
In terms of the usual cyclotron length ` with `2 = ~c
eB
we define the dimensionless
variable ξ = y` − `k and the operators
1
a = √ (∂ξ + ξ)
2
1
a† = √ (−∂ξ + ξ) (145)
2
with the harmonic oscilator commutation [a, a† ] = 1. The Hamiltonian becomes
 
0 a
−~ωc † φ(ξ) = Eφ(ξ) (146)
a 0

and ωc = 2v/`. This ωc is very different from that of the usual quantum Hall case
that involves massive particles, allowing much higher values. We note an E = 0
solution with a state
 
0
φ0 (ξ) = (147)
ψ0 (ξ)
1 2
where φ0 (ξ) ∼ e− 2 ξ so that aψ0 (ξ) = 0. The other states correspond to Harmonic
oscillator states
1 1 1 2
ψN = √ e− 2 ξ HN (ξ) (148)
2N/2 N!
where HN (ξ) are Hermite polynomials, since
  √

     
0 a ψN −1 (ξ) ± N ψ N −1 ψ N −1
· = √ =± N (149)
a† 0 ±ψN (ξ) N ψN (ξ) ±ψN (ξ)

so that the eigenstates and eigenvalues are



 
ψN −1 (ξ)
φN (ξ) = , EN = ± N ~ωc (150)
∓ψN (ξ)

The degeneracy of each Landau level is found as in the usual quantum Hall,
except that here we have spin degeneracy (neglecting the relatively small Zeeman
efect) and another factor 2 due the presence of 2 Dirac points, i.e. the K, K0 de-
generacy. Hence k = L2πx m with an integer m and the oscillator center position `2 k
yields the degeneracy Ns and the filling factor ν
2π Lx Ly N φ0
0 < `k = `2 m < Ly ⇒ Ns = 4 ⇒ ν= = ne (151)
Lx 2π`2 Ns 4B
where ne is the electron density. Peaks of the resistance ρxx correspond to half
filled Landau levels and have a periodicity in density of ∆n = 4B φ0
, as indeed in
Fig. 28. However, the values of the quantized plateaus of σxy do not follow simply
by extending Laughlin’s argument of Fig. 9. The latter assumes that in the non-
dissipative phase ρxx = 0 passing a flux φ0 in Fig. 9 causes N 0 charges e to move

39
from one edge to the other, where N 0 is the number of extended states. A naive
2
guess N 0 = 4N would give σxy = 4N eh , which is inconsistent with the data, in
particular there is no plateau at N = 0.
In a semiclassical picture a full band does not contribute to σxy since the equation
of motion dk
dt
∼ v × B just rearranges states in the bands without changing overall
occupation. In graphene the states E < 0 form a band as in Fig. 26. A question
then arises how to count the number of extended states that belong to the N = 0
Landau level – how many of them contribute to N 0 in Laughlin’s argument? To
identify the correct N 0 we turn to study edge states.

5.5 Edge states in graphene


We wish to add a potential term to graphene’s Hamiltonian that represents edges,
i.e. confines the electrons in a large but finite width. A term V (y) · 1 with a unit
matrix is not appropriate, since in view of Klein’s paradox it does not confine the
electrons. We therefore choose a term m(y)σz where m(y) = 0 in the bulk and
diverges at the edges. A mass term, as in Eq. (138), produces a gap in the spectrum
and for energies |E| < |m| there are no propagating states, i.e. they are evanescent
as needed beyond the edge. The Hamiltonian is then
   
Be
H = ~v iσy ∂y − k − y σx τz + m(y)σz (152)
~c

where the τz is in the K, K0 space, as in Eq. (135).

Figure 29: Energy levels of Graphene in presence of edges. The Fermi energy here crosses the 1st
Landau level with degeneracy 4 and 2 degenerate levels of the 0th level.

We note a particle-hole transformation

Ĉ = σx τy ⇒ ĈHĈ = −H (153)

40
so that eigenvalues come in ± pairs, i.e.

HψE = EψE ⇒ HĈψE = −E ĈψE (154)

The significant consequence is that the 4 fold degenerate zero energy Landau level
must split into two 2-fold degenerate states near the edges, as illustrated in Fig. 29.
Consider now a positive Fermi energy EF that is between the levels N, N + 1 in
the bulk. It crosses N + 1 edge states with degeneracy 2 for the lowest one and 4
for each of the next levels. The Hall conductance is therefore
e2
σxy = ±2(2N + 1) ± = sign of EF (155)
h
accounting for the data in Fig. 28. If the Fermi energy is negative and is in between
the levels −EN , −EN +1 in the bulk, then the Fermi energy again crosses N + 1 levels
with total degeneracy 4N + 2. These levels have now opposite slope than those with
positive energy, hence the − sign in Eq. (155). Therefore, the correct number of
transported charges N 0 in Laughlin’s argument is N 0 = 4N + 2.
The remarkable phenomena found in graphene concludes our study of quantized
transport. It also provides a starting point for the very active present research of
topological insulators, to be presented in a separate course.

41

You might also like