You are on page 1of 55

Chapter 6: The Fermi Liquid

L.D. Landau

December 22, 2000

Contents
1 introduction: The Electronic Fermi Liquid 3

2 The Non-Interacting Fermi Gas 5


2.1 Infinite-Square-Well Potential . . . . . . . . . . . . . . . . . . . . . . 5
2.2 The Fermi Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 T = 0, The Pauli Principle . . . . . . . . . . . . . . . . . . . . 10
2.2.2 T 6= 0, Fermi Statistics . . . . . . . . . . . . . . . . . . . . . . 13

3 The Weakly Correlated Electronic Liquid 23


3.1 Thomas-Fermi Screening . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Fermi liquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 Quasi-particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.1 Particles and Holes . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.2 Quasiparticles and Quasiholes at T = 0 . . . . . . . . . . . . . 33
3.4 Energy of Quasiparticles. . . . . . . . . . . . . . . . . . . . . . . . . . 39

4 Interactions between Particles: Landau Fermi Liquid 42


4.1 The free energy, and interparticle interactions . . . . . . . . . . . . . 42
4.2 Local Energy of a Quasiparticle . . . . . . . . . . . . . . . . . . . . . 46

1
4.2.1 Equilibrium Distribution of Quasiparticles at Finite T . . . . . 48
4.3 Effective Mass m∗ of Quasiparticles . . . . . . . . . . . . . . . . . . . 50

2
1 introduction: The Electronic Fermi Liquid

As we have seen, the electronic and lattice degrees of freedom


decouple, to a good approximation, in solids. This is due to the
different time scales involved in these systems.

τion ∼ 1/ωD À τelectron ∼ (1)
EF
where EF is the electronic Fermi energy. The electrons may be
thought of as instantly reacting to the (slow) motion of the lat-
tice, while remaining essentially in the electronic ground state.
Thus, to a good approximation the electronic and lattice degrees
of freedom separate, and the small electron-lattice (phonon) in-
teraction (responsible for resistivity, superconductivity etc) may
be treated as a perturbation (with ωD /EF as an expansion pa-
rameter); that is if we are capable of solving the problem of the
remaining purely electronic system.
At first glance the remaining electronic problem would also
appear to be hopeless since the (non-perturbative) electron-
electron interactions are as large as the combined electronic
kinetic energy and the potential energy due to interactions with
3
the static ions (the latter energy, or rather the corresponding
part of the Hamiltonian, composes the solvable portion of the
problem). However, the Pauli principle keeps low-lying orbitals
from being multiply occupied, so is often justified to ignore the
electron-electron interactions, or treat them as a renormaliza-
tion of the non-interacting problem (effective mass) etc. This
will be the initial assumption of this chapter, in which we will
cover

• the non-interacting Fermi liquid, and

• the renormalized Landau Fermi liquid (Pines Nozieres).

These relatively simple theories resolved some of the most


important puzzles involving metals at the turn of the century.
Perhaps the most intriguing of these is the metallic specific heat.
Except in certain “heavy fermion” metals, the electronic contri-
bution to the specific heat is always orders of magnitude smaller
than the phonon contribution. However, from the classical theo-
rem of equipartition, if each lattice site contributes just one elec-
tron to the conduction band, one would expect the contributions

4
from these sources to be similar (Celectron ≈ Cphonon ≈ 3N rkB ).
This puzzle is resolved, at the simplest level: that of the non-
interacting Fermi gas.

2 The Non-Interacting Fermi Gas

2.1 Infinite-Square-Well Potential

We will proceed to treat the electronic degrees of freedom, ig-


noring the electron-electron interaction, and even the electron-
lattice interaction. In general, the electronic degrees of freedom
are split into electrons which are bound to their atomic cores
with wavefunctions which are essentially atomic, unaffected by
the lattice, and those valence (or near valence) electrons which
react and adapt to their environment. For the most part, we are
only interested in the valence electrons. Their environment de-
scribed by the potential due to the ions and the core electrons–
the core potential. Thus, ignoring the electron-electron interac-
tions, the electronic Hamiltonian is
P2
H= + V (r) . (2)
2m
5
As shown in Fig. 1, the core potential V (r), like the lattice, is
periodic
V(r)

a V(r+a) = V(r)
Figure 1: Schematic core potential (solid line) for a one-dimensional lattice with
lattice constant a.

For the moment, ignore the core potential, then the electronic
wave functions are plane waves ψ ∼ eik·r . Now consider the
core potential as a perturbation. The electrons will be strongly
effected by the periodicity of the potential when λ = 2π/k ∼ a
1
. However, when k is small so that λ À a (or when k is large,
so λ ¿ a) the structure of the potential may be neglected, or
we can assume V (r) = V0 anywhere within the material. The
1
Interestingly, when λ ∼ a, the Bragg condition 2d sin θ ≈ a ≈ λ may easily be satisfied, so the
electrons, which may be though of as DeBroglie waves, scatter off of the lattice. Consequently states
for which λ = 2π/k ∼ a are often forbidden. This is the source of gaps in the band structure, to be
discussed in the next chapter.

6
potential still acts to confine the electrons (and so maintain
charge neutrality), so V (r) = ∞ anywhere outside the material.

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

Thus we will approximate the potential of a cubic solid with


linear dimension L as an infinite square-well potential.





 V0 0 < r i < L
V (r) = 

(3)


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

7
The normalize plane wave solution to this model is

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

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


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

ky

3
(π/L)
kx

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

eigenenergies of the wavefunctions are


h̄2∇2 h̄ X 2 h̄2π 2 µ 2 2 2

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

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

9
2.2 The Fermi Gas

2.2.1 T = 0, The Pauli Principle

Electrons, as are all half-integer spin particles, are Fermions.


Thus, by the Pauli Principle, no two of them may occupy the
same state. For example, if we calculate the density of electrons
per unit volume
Z ∞
n= 0
D(E)f (E, T )dE , (11)

where f (E, T ) is the probability that a state of energy E is oc-


cupied, the factor f (E, T ) must enforce this restriction. How-
ever, f is just the statistical factor; c.f. for classical particles

f (E, T ) = e−E/kB T for classical particles , (12)

which for T = 0 would require all the electrons to go into


the ground state f (0, 0) = 1. Clearly, this violates the Pauli
principle.
At T = 0 we need to put just one particle in each state, start-
ing from the lowest energy state, until we are out of particles.
Since E ∝ k 2 in our simple square-well model, will fill up all

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

ky 2 2
kf h kf
= Ef
2m

occupied
states

kx
kf D(E)

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

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

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

12.3 A ◦
λe = 1/2
≈4A (18)
(E(eV))
thus our original approximation of a square well potential, ig-
noring the lattice structure, is questionable for electrons near
the Fermi surface, and should be regarded as yielding only qual-
itative results.

12
2.2.2 T 6= 0, Fermi Statistics

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


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

1 2

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

13
this system, in equilibrium,
X ∂F
0 = δF = δni (19)
i ∂ni
P
electrons are conserved so i δni = 0. Thus, for our two orbital
system
∂F ∂F
δn1 + δn2 = 0 and δn1 + δn2 = 0 (20)
∂n1 ∂n2
or
∂F ∂F
= (21)
∂n1 ∂n2
A similar relation holds for an arbitrary number of particles.
Apparently this quantity, the increased free energy needed to
add a particle to the system, is a constant
∂F
=µ (22)
∂ni
for all i. µ is called the chemical potential.
Now consider an ensemble of orbitals. We will treat the ther-
modynamics of this system within the canonical ensemble (i.e.
the system is in contact with a thermal bath, and the particle
number is conserved) for which F = E − T S is the appropriate
potential. The system energy E and Entropy S may be written
14
as functions of the orbital energies Ei and occupancies ni and
the degeneracy gi of the state of energy Ei. For example,

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

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

Figure 6: states from an ensemble of orbitals.

X
E= ni Ei . (23)
i

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


of ways of distributing the electrons among the states, then

S = kB ln P . (24)

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


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

1.0
+1)
β(ω−µ)

0.5
1/(e

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

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


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

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

the phonons contribute essentially a classical result to the spe-


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

18
Specific Heat Calculation Of course, since we know the free energy
of the non-interacting Fermi gas, we can calculate the form of
the specific heat. Here we will follow Ibach and Lüth and Kittel;
however, since the chemical potential does depend upon the
temperature, I would like to make the approximations we make
a bit more explicit.
Upon heating from T = 0 to finite T , the Fermi gas will gain
energy
Z ∞ Z E
F
U (T ) = 0
dE ED(E)f (E, T ) − 0
dE ED(E) (35)

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

0.3
x

0.2
2 x

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

µ ¶
x
T ¿ TF , D β + EF ≈ D (EF ), since the density of states
usually does not have features which are sharp on the energy
scale of 10kB T . Thus
Z ∞
2 ex
CV ≈ kB T D (EF ) dx x x
−βEF (e + 1)2
π2 2
≈ k T D(EF ) . (42)
3 B
Note that no assumption about the form of D(E) was made
other than the assumption that it is smooth within kB T of the
Fermi surface. Thus, experimental measurements of the specific
heat at constant volume of the electrons, gives us information
21
about the density of electronic states at the Fermi surface.
Now let’s reconsider the DOS for the 3-D box potential.
  1/2  
1  2m 3/2 1/2 E
D(E) = 2 E = D(EF )   (43)
2π h̄ EF
R EF
For which n = 0 D(E)dE = D(EF ) 23 EF , so
π2 T 3
CV = nkB ¿ nkB (44)
2 TF 2
where the last term on the right is the classical result. For
room temperatures T ∼ 300K, which is also of the same order
of magnitude as the Debye temperatures θD ,
3
CV phonon ∼ nkB À CV electron (45)
2
So, the only way to measure the electronic specific heat in most
materials is to go to very low temperatures T ¿ θD , for which
CV phonon ∼ T 3. Here the total specific heat

CV ≈ γT + βT 3 (46)

We will see that gives us some measurement of the electronic ef-


fective mass for our Fermi liquid theory. I.e. it tell us something
about electron- electron interactions.
22
3 The Weakly Correlated Electronic Liquid

3.1 Thomas-Fermi Screening

As an introduction to the effect of electronic correlations, con-


sider the effect of a charged oxygen defect in one of the copper-
oxygen planes of a cuprate superconductor shown in Fig. 9.
Assume that the oxygen defect captures two electrons from the
metallic band, going from a 2s22p4 to a 2s22p6 configuration.
The defect will then become a cation, and have a net charge
O

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

O O O O O O

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

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

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

O O O O O O

Figure 9: A charged oxygen defect is introduced into one of the copper-oxygen planes of
a cuprate superconductor. The oxygen defect captures two electrons from the metallic
band, going from a 2s2 2p4 to a 2s2 2p6 configuration.

of two electrons. In the vicinity of this oxygen defect, the elec-

23
trostatic potential and the electronic charge density will be re-
duced.
If we model the electronic density of states in this material
with our box-potential DOS, we can think of this reduction in
the local charge density in terms of raising the DOS parabola
near the defect (cf. Fig. 10). This will cause the free electronic
near charged Away from
defect charged defect

e
EF

-eδU

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

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

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

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

If |eδU | ¿ EF , then

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

We can solve for the change in the electrostatic potential by


solving Poisson equation.

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

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


qe−λr
δU (r) = (52)
r
The length 1/λ = rT F is known as the Thomas-Fermi screening
length.
µ ¶−1/2
2
rT F = 4πe D(EF ) (53)
3
The solution is actually Ce−λr /r, where C is a constant. C may be deterined by letting D(EF ) =
0, so the medium in which the charge is embedded becomes vacuum. Then the potential of the charge
is q/r, so C = q.

25
Lets estimate this distance for our square-well model,
a0 π a0
rT2 F = ≈
3(3π 2n)1/3 4n1/3
 
1  n −1/6
rT F ≈ (54)
2 a30

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

r r
/r

/r
-r/rTF

rTF=1/4
-r/rTF

rTF=1
-1/6
-e

rTF= n
-e

bound states
free states

Figure 11: Screened defect potentials. As the screening length increases, states that
were free, become bound.

Now consider an electron bound to an ion in Cu or some other


metal. As shown in Fig. 11 the screening length decreases, and
bound states rise up in energy. In a weak metal (i.e. something
26
like YBCO), in which the valence state is barely free, a reduction
in the number of carriers (electrons) will increase the screening
length, since
rT F ∼ n−1/6 . (56)

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


or bind more states–making the one free valance state bound.
Now imagine that instead of a single defect, we have a con-
centrated system of such ions, and suppose that we decrease
the density of carriers (i.e. in Si-based semiconductors, this is
done by doping certain compensating dopants, or even by mod-
ulating the pressure). This will in turn, increase the screening
length, causing some states that were free to become bound,
causing an abrupt transition from a metal to an insulator, and
is believed to explain the MI transition in some transition-metal
oxides, glasses, amorphous semiconductors, etc.

27
3.2 Fermi liquids

The purpose of these next several lectures is to introduce you


to the theory of the Fermi liquid, which is, in its simplest form,
a collection of Fermions in a box plus interactions.
In reality , the only physical analog is a gas of 3He, which
due its nuclear spin (the nucleus has two protons, one neutron),
obeys Fermi statistics for sufficiently low energies or temper-
atures. In addition, simple metals, from the first or second
column of the periodic table, for which we may approximate
the ionic potential
V (R) = V0 (57)

are a close approximant to Fermi liquids.


Moreover, Fermi Liquid theory only describes the ”gaseous”
phase of these quantum fermion systems. For example, 3He also
has a superfluid (triplet), and at least in 4He-3 He mixtures, a
solid phase exists which is not described by Fermi Liquid The-
ory. One should note; however, that the Fermi liquid theory
state does serve as the starting point for the theories of super-

28
conductivity and super fluidity.
One may construct Fermi liquid theory either starting from a
many-body diagrammatic or phenomenological viewpoint. We,
as Landau, will choose the latter. Fermi liquid theory has 3
basic tenants:

1. momentum and spin remain good quantum numbers to de-


scribe the (quasi) particles.

2. the interacting system may be obtained by adiabatically


turning on a particle-particle interaction over some time t.

3. the resulting excitations may be described as quasi-particles


with lifetimes À t.

3.3 Quasi-particles

The last assumption involves a new concept, that of the quasi-


particles which requires some explanation.

3.3.1 Particles and Holes

Particles and Holes are excitations of the non-interacting system


at zero temperature. Consider a system of N free Fermions
29
each of mass m in a volume V . The eigenstates are the anti-
symmetrized combinations (Slater determinants) of N different
single particle states.
1
ψp(r) = √ eip·r/h̄ (58)
V
The occupation of each of these states is given by np = θ(p−pF )
where pF is the radius of the Fermi sphere. The energy of the
system is
X p2
E= np (59)
p 2m
and pF is given by
N 1 Ã p F !3
= 2 (60)
V 3π h̄
Now lets add a particle to the lowest available state p = pF
then, for T = 0,
∂E0 p2F
µ = E0(N + 1) − E0(N ) = = . (61)
∂N 2m
If we now excite the system, we will promote a certain number
of particles across the Fermi surface SF yielding particles above
and an equal number of vacancies or holes below the Fermi
surface. These are our elementary excitations, and they are

30
quantified by δnp = np − n0p





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

. (62)
 0

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

D(E)

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

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


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

Now lets take our system and place it in contact with a par-
ticle bath. Then the appropriate potential is the free energy,

31
E
2
δF = p’ /2m - µ
2
EF = µ = p /2m
2 F
δF = µ - p /2m

2
δF = | µ - p /2m |

D(E)

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

which for T = 0, is F = E − µN , and


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

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


32
3.3.2 Quasiparticles and Quasiholes at T = 0

2 -a/r TF
U ≈ ea e
a

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

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


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

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

p p p

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

Figure 15: We add a particle with momentum p to our noninteracting (U = 0) Fermi


liquid at time t = 0, and slowly increase the interaction to its full value U at time t.
As the particle and system evolve, the particle becomes dressed by interactions with
the system (shown as a shaded ellipse) which changes the effective mass but not the
momentum of this single-particle excitation (now called a quasi-particle).

switched on, we slowly begin to perturb the particles close to


the additional particle, so the particle becomes dressed by these
interactions. However since momentum is conserved, we have
created an excitation (particle and its cloud) of momentum p.
We call this particle and cloud a quasiparticle. In the same way,

34
if we had introduced a hole of momentum p below the Fermi
surface, and slowly turned on the interaction, we would have
produced a quasihole.
Note that this adiabatic switching on procedure will have
difficulties if the lifetime of the quasi-particle τ < t. If so,
then the process is not reversible. If we shorten t so that again
τ À t, then the switching on of U may not be adiabatic (ie., we
will evolve to a system which is not in its ground state). Such
difficulties do not arise so long as the energy of the particle is
close to the Fermi energy. Here there are few states accessible
for creating particle-hole excitations. In fact, one can formulate
a perturbative argument that the lifetime of a quasi-particle is
proportional to the square of its excitation energy above the
p2 p2F
Fermi energy ε = 2m − 2m ≈ v(p − pF ).
To estimate this lifetime consider the following argument
from AGD: A particle with momentum p1 above the Fermi
surface (p1 > pF ) interacts with one of the particles below the
Fermi surface with momentum p2. As a result, two new par-
ticles appear above the Fermi surface (all other states are full)

35
p4

p3

p1 p2

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

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


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

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


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

k1 − k 3 = k 4 − k 2 . (68)

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


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

N(E)

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


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

µ ¶
ε1 2
Thus, the scattering rate 1/τ is proportional to EF so that
excitations of sufficiently small energy will always be sufficiently
long lived to satisfy the constraints of reversibility. Finally, the
fact that the quasiparticle only interacts with a small number of
other particles due to Thomas-Fermi screening (i.e. those within
a distance ≈ RT F ), also significantly reduces the scattering rate.

38
3.4 Energy of Quasiparticles.

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


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

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


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

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

²p ≈ µ + (p − pF ) · ∇p ²p|pF (72)

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


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

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


²p,σ = ²−p,σ under BZ reflection (74)

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


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

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

vpF = pF /m∗ (75)

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


fermi surface. Recall, that in the non-interacting system,
 
1  2m 3/2 1/2 mpF
D(EF ) = 2 EF = (76)
2π h̄2 πh̄3
where p = h̄k, and E = p2/2m. Thus, for the interacting
system at the Fermi surface
m∗ p F
Dinteracting (EF ) = , (77)
πh̄3
where the m∗ (generally > m, but not always) accounts for the
fact that the quasiparticle may be viewed as a dressed particle,
and must “drag” this dressing along with it. I.e., the effective
mass to some extent accounts for the interaction between the
particles.

41
4 Interactions between Particles: Landau Fermi
Liquid

4.1 The free energy, and interparticle interactions

The thermodynamics of the system depends upon the free en-


ergy F , which at zero temperature is

F − F0 = E − E0 − µ(N − N0) . (78)

Since our quasiparticles are formed by adiabatically switching


on the interaction in the N + 1 particle ideal system, adding
one quasiparticle to the system adds one real particle. Thus,
X
N − N0 = δnp , (79)
p

and since
X
E − E0 ≈ ²pδnp , (80)
p

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

As shown in Fig. 18, we will be interested in excitations of the


system which distort the Fermi surface by an amount propor-
tional to δ. For our theory/expansion to remain valid, we must
42
δ

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


1 P
that N p |δnp | ¿ 1.

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

so, to be consistent we must add the next term in the Taylor


series expansion of the energy to the expression for the free
energy.
X 1 X
F − F0 = (²p − µ) δnp + fp,p0 δnpδnp0 + O(δ 3) (84)
p 2 p,p0
43
where
δE
fp,p0 = (85)
δnpδnp0
The term, proportional to fp,p0 , was added (to the Sommerfeld
theory) by L.D. Landau. Since each sum over p is proportional
to the volume V , as is F , it must be that fp,p0 ∼ 1/V . However,
it is also clear that fp,p0 is an interaction between quasiparticles,
each of which is spread out over the whole volume V , so the
probability that they will interact is ∼ rT3 F /V , thus

fp,p0 ∼ rT3 F /V (86)

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


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

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


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

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

and, in a system with reflection symmetry

fpσ,p0 σ0 = f−pσ,−p0 σ0 . (89)

Then
fpσ,p0 σ0 = fp−σ,p0 −σ0 . (90)

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


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

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

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

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


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

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

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


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

Conventionally these f parameters are expressed in terms of


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

4.2 Local Energy of a Quasiparticle

Figure 19: The addition of another particle to a homogeneous system will yeilds in
forces on the quasiparticle which tend to restore equilibrium.

Now consider an interacting system with a certain distribu-


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

(recall that fp,p0 = fp0 ,p ). Both terms here are O(δ). The
second term describes the free energy of a quasiparticle due to
the other quasiparticles in the system (some sort of Hartree-like
term).
The term ²̃p plays the part of the local energy of a quasi-
particle. For example, the gradient of ²̃p is the force the system
exerts on the additional quasiparticle. When the quasiparticle
is added to the system, the system is inhomogeneous so that
δnp0 = δnp0 (r). The system will react to this inhomogeneity
by minimizing its free energy so that ∇r F = 0. However, only
the additional free energy due the added particle (Eq. 95) is
inhomogeneous, and has a non-zero gradient. Thus, the system
will exert a force
X
−∇r ²̃ = −∇r fp0,p δnp0 (r) (96)
p0

on the added quasiparticle resulting from interactions with other


quasiparticles.
47
4.2.1 Equilibrium Distribution of Quasiparticles at Finite T

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


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

δnp = hδnpi + (δnp − hδnpi) (98)

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

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

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


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

X 1 X
≈ h²̃piδnp − fp0,p hδnpihδnp0 i + O(δ 4) (100)
p 2 p,p0
At this point, we may repeat the arguments made earlier to de-
termine the fermion occupation probability for non-interacting
48
Fermions (the constant factor on the right hand-side has no
effect). We will obtain
1
np(T, µ) = , (101)
1 + exp β(h²̃pi − µ)
or
1
δnp(T, µ) = − θ(pf − p) . (102)
1 + exp β(h²̃p i − µ)
However, at least for an isotropic system, this expression bears
closer investigation. Here, the molecular field (evaluated within
kB T of the Fermi surface)
X
h²̃p − ²pi = fp0 ,phδnp0 i (103)
p0

must be independent of the location of p on the Fermi sur-


face (and of course, spin), and is thus constant. To see this,
reconsider the Legendre polynomial expansion discussed earlier
X
h²̃p − ²pi = fp0,p hδnp0 i
p0
XZ
∝ d3pfl Pl (cos θ)hδnp0 i
l Z
∝ f0 d3phδnp0 i = 0
(104)

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

4.3 Effective Mass m∗ of Quasiparticles

This argument most closely follows that of AGD, and we will


follow their notation as closely as possible (without introducing
any new symbols). In particular, since an integration by parts
is necessary, we will use a momentum integral (as opposed to a
momentum sum) notation
X Z d3 p
→V . (106)
p (2πh̄)3
The net momentum of the volume V of quasiparticles is
Z d3 p
Pqp = 2V pnp net quasiparticle momentum (107)
(2πh̄)3
which is also the momentum of the Fermi liquid. On the other
hand since the number of particles equals the number of quasi-
particles, the quasiparticle and particle currents must also be
50
equal
Z d3 p
Jqp = Jp = 2V vpnp net quasiparticle and particle current
(2πh̄)3
(108)
or, since the momentum is just the particle mass times this
current
d3 p
Z
Pp = 2V m vpnp net quasiparticle and particle current
(2πh̄)3
(109)
where vp = ∇p²̃p, is the velocity of the quasiparticle. So
Z d3 p Z d3 p
pnp = m ∇p²̃pnp (110)
(2πh̄)3 (2πh̄)3
Now make an arbitrary change of np and recall that ²̃p depends
upon np, so that
XZ d3 p
δ²̃p = V fp,p0 δnp0 . (111)
σ0 (2πh̄)3
For Eq. 110, this means that
Z d3 p Z d3 p
pδnp = m ∇p²̃pδnp (112)
(2πh̄)3 (2πh̄)3
Z d 3 p X Z d 3 p0 ³ ´
+mV ∇ f
p p,p 0 δn p np ,
0
(2πh̄)3 σ0 (2πh̄)3

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

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

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

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


Z 1 2
−1
dxPn(x)Pm(x)dx = δnm (119)
2n + 1
and finally that
V m ∗ pF α
D(0)flα = 2
α
3 fl = F l , (120)
π h̄
we find that
1 1 F1s
= + , (121)
m m∗ 3m∗
53
Quantity Fermi Liquid Fermi Liquid/Fermi Gas
m∗ pF 2 CV m∗
Specific Heat Cv = k T
3h̄3 B CV 0 = m =1+ F1s/3
κ 1+F0s
Compressibility κ0 = 1+F0s /3
µ ¶2
p2F c 1+F0s
Sound Velocity c2 = 3mm∗ (1 + F0s) c0 = 1+F1s /3
m∗ pF β 2 χ 1+F1s /3
Spin Susceptibility χ = π 2 h̄3 1+F0a χ0 = 1+F0a
Table 1: Fermi Liquid relations between the Landau parameters Fnα and some exper-
imentally measurable quantities. For the latter, a zero subscript indicates the value
for the non-interacting Fermi gas.

or m∗/m = 1 + F1s/3.
The effective mass cannot be experimentally measured di-
rectly; however, it appears in many physically relevant measur-
able quantities, including the specific heat
 
∂E/V  1 ∂ X
CV = 
  = ²̃pnp. (122)
∂T V N V ∂T p

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


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

55

You might also like