Sie sind auf Seite 1von 15

CCMS Summer 2007 Lecture Series

Fermi- and non-Fermi Liquids


Lecture 3: Fermi-liquid Theory
Dmitrii L. Maslov
maslov@phys.ufl.edu
(Dated: July 22, 2007)

Notation 1 Here and thereafter, L1 stands for Lecture 1: Basic Notions of the Many-body
Physics and L2 stands for Lecture 2: Examples of Interacting Fermi systems.

I.

GENERAL CONCEPTS

All Fermi systems (metals, degenerate semiconductors, normal He3 , neutron stars, etc.)
belong to the categories of either moderately or strongly interacting systems. For example,
in metals rs in the range from 2 to 5. (There are only few exceptions of this rule; for
example, bismuth, in which the large value of the background dielectric constant brings the
value of rs to 0.3 and GaAs heterostructures in which the small value of the effective mass
0.07 of the bare massleads to the higher value of the Fermi energy and thus to rs < 1).
On the other hand, as we learned from the section on Wigner crystallization, the critical rs
for Wigner crystallization is very high about 100 in 3D and about 40 in 2D. Thus almost
all Fermi systems occurring in Nature are too strongly interacting to be described by the
weak-coupling theory (Coulomb gas) but too weakly interacting to solidify. Landau put
forward a hypothesis that an interacting Fermi system is qualitatively similar to the Fermi
gas [6] . Although original Landaus formulation refers to a translationally invariant system
of particles interacting via short-range forces, e.g., normal He3 , later on his arguments were
extended to metals (which have only discrete symmetries) and to charged particles.
Experiment gives a strong justification to this hypothesis. The specific heat of almost
all fermionic systems (in solids, one need to subtract off the lattice contribution to get the
one from electrons) scales linearly with temperature: C (T ) = T. In a free Fermi gas,
= = ( 2 /3) F = (1/3) mkF . In a band model, when non-interacting electrons move in
the presence of a periodic potential, one should use the appropriate value of the density of
states at the Fermi level for a given lattice structure. In reality, the coefficient can differ
significantly from the band value but the linearity of C(T ) in T is well-preserved. In those
cases when one can change number density continuously (for example, by applying pressure
to normal He3 ), is found to vary. One is then tempted to assume that the interacting
Fermi liquid is composed of some effective particles (quasi-particles) that behave as free
fermions albeit their masses and other characteristics are different from the non-interacting
values.

A.

Quasi-particles

The concept of quasi-particles is central to the Landaus theory of Fermi liquids. The
ground state of a Fermi gas is a completely filled Fermi sphere. The spectrum of excited
states can be classified in terms of how many fermions were promoted from states below
the Fermi surface to the ones above. For example, the first excited state is the one with
an electron above the Fermi sphere and the hole below. The energy of this state, measured
from the ground state, is p = p2 /2m F . The net momentum of the system is p~. The next
state correspond to two fermions above the Fermi sphere, etc. If the net momentum of the
system is p~, then p = p~1 + p~2 , where p~1 and p~2 are the momenta of individual electrons. We
see that in a free system, any excited macroscopic state is a superposition of single-particle
states.
This is not so in an interacting system. Even if we promote only one particle to a state
above the Fermi surface, the energy of this state would not be equal to p2 /2m p because
the interaction will change the energy of all other fermions. However, Landau assumed
that excited states with energies near the Fermi level, that is, weakly excited states, can
be described as a superposition of elementary excitations which behave as free particles,
although the original system may as well be a strongly interacting one. An example of
such a behavior are familiar phonons in a solid. Suppose that we have a gas of sodium
atoms (which are fermions) which essentially do not interact because of low density. The
elementary excitations in an ideal gas simply coincide with real atoms. Now we condense the
gas into a metal. Individual atoms are not free to move on their own. Instead, they can only
participate in a collective oscillatory motion, which is a sound wave. For small frequencies,
the sound wave can be thought of consisting of elementary quanta of free particlesphonons.
The spectrum of each phonon branch is i (q) = si q , where si is the speed of sound and the
oscillatory energy is
E=

XZ
i

d3 q
i ni ,
(2)3

where ni is the number of excited phonons at given temperature. If the number of phonons
is varied, so is the total energy
E =

XZ
i

d3 q
i ni .
(2)3

Quite similarly, Landau assumed that the variation of the total energy of a single-component
3

Fermi liquid (or single band metal) can be written as


Z

E =

d3 p
p np .
(2)3

(1.1)

(For the sake of simplicity, I also assume that the system is isotropic, i.e., the energy and
n depend only on the magnitude but not the direction of the momentum but the argument
can be extended to anisotropic systems as well). In this formula, np is the distribution
function of quasi-particles which are elementary excitations of the interacting system. The
variation of np occurs when, e.g., a new particle is added to the system. However, these
quasi-particles are not completely free and, therefore, unlike in the phonon analogy, the
total energy is not equal to the sum of individual energies. However, the variation of the
total energy E is related to np via (1.1). In fact, this relation serves as a definition of the
quasiparticle energy via the variational derivative of E :
p =

E
.
np

One more and crucial assumptionis that the quasi-particles of an interacting


Fermi system are fermions, which is not at all obvious. For example, regardless of the
statistics of individual atoms which can be either fermions or bosons, phonons are always
bosons. Landaus argument was that if quasi-particles were bosons they could accumulate
without a restriction in every quantum state. That means that an excited state of a quantum
system has a classical analog. Indeed, an excited state of many coupled oscillators is a
classical sound wave. Fermions dont have macroscopic states so quasi-particles of a Fermi
systems must be fermions (to be precise they should not be bosons; proposals for particles
of a statistics intermediate between bosons and fermionsanyonshave been made). On
quite general grounds, one can show that quasi-particles must have spin 1/2 regardless of
(half-integer) spin of original particles [3],[4]. (Thus, quasi-particles of a system composed
of fermions with spin S = 3/2 still have spin 1/2.) Therefore, the quasiparticle energy and
the occupation number are the operators in the spin space represented by 2 2 matrices
and n
. As long as one has a system of fermions with doubly (Sz = 1/2) degenerate states,
the entropy on purely combinatorial grounds is
Z

S/vol = Tr

d3 p
[
np ln n
p (1 n
p ) ln (1 n
p )] ,
(2)3

which is the same expression in for a Fermi gas. The equlibrium occupation numbers are
obtained by equating the variation of S at fixed volume to zero. As for a Fermi gas, this
4

gives
n
=

exp

+1

(1.2)

In contrast to the Fermi-gas, however, is a functional of n


itself.
If the system is not in the presence of the magnetic field and not ferromagnetic,
= , n
= n .
In a general case, instead of (1.1) we have
Z

E =

d3 p
Tr
(p)
n (p) ,
(2)3

which, for a spin-isotropic liquid, reduces to


Z

E = 2

d3 p
Tr (p) n (p) .
(2)3

The occupation number is normalized by the condition


Z

N =

Z
d3 p
d3 p
Trn
(p)
=
2
n (p) = 0,
(2)3
(2)3

where N is the total number of real particles.


For T=0, the chemical potential coincides with energy of the topmost state
= (pF ) EF .
Another important property (known as Luttinger theorem) is that the volume of the Fermi
surface is not affected by the interaction. For an isotropic system, this means the Fermi
momenta of free and interacting systems are the same. A simple argument is that the
counting of states is not affected by the interaction, i.e., the relation
N =2

4p3F /3
(2)3

holds in both cases. A general proof of this statement is given in Ref. [ [3]].

B.

Interaction of quasi-particles

Phonons in a solid do not interact only in the first (harmonic) approximation. Anharmonism results in the phonon-phonon interaction. However, the interaction is weak at low
5

energies not really because the coupling constant is weak but rather because the scattering
1
rate of phonons on each other is proportional to a high power of their frequency: ph-ph
5.

As a result, at small phonons are almost free quasi-particles. Something similar happens
with the fermions. The nominal interaction may as well be strong. However, because of the
Pauli principle, the scattering rate is proportional to ( F )2 and weakly excited states
interact only weakly. In the Landau theory, the interaction between quasi-particles is introduced via a phenomenological interaction function defined by the proportionality coefficient
(more precisely, a kernel) between the variation of the occupation number and the corresponding variation in the quasi-particle spectrum
Z

d3 p0
f, (~p, p~0 ) n (~p0 ) .
(2)3

Function f, (~p, p~0 ) describes the interaction between the quasi-particles of momenta p~ and
p~0 (notice that these are both initial states of the of the scattering process). Spin indices
and correspond to the state of momentum p~ whereas indices and correspond to
momentum p~0 . In a matrix form,
Z
0

= Tr

d3 p0
p, p~0 )
n (~p0 ) ,
3 f (~
(2)

(1.3)

where Tr0 denotes trace over spin indices and . For a spin-isotropic FL, when =
and
n = n external spin indices ( and ) can also be traced out and Eq.(1.3) reduces
to

d3 p0
f (~p, p~0 ) n (~p0 ) ,
(2)3

where
1
f (~p, p~0 ) TrTr0 f (~p, p~0 ) .
2
For small deviations from the equilibrium, n (~p0 ) is peaked near the Fermi surface. Function
f (~p, p~0 ) can be then estimated directly on the Fermi surface, i.e., for |~p| = |~p0 | = pF . Then f
depends only on the angle between p~ and p~0 . The spin dependence of f can be established on
quite general grounds. In a spin-isotropic FL, f can depend only on the scalar product of
spin operators but on the products of the individual spin operators with some other vectors.
Thus the most general form of f for a spin-isotropic system is

0 ,
f (~p, p~0 ) = F s () I + F a ()

where I is the unity matrix,


is the vector of three Pauli matrices, is the angle between
p~ and p~0 , and the density of states was introduced just to make functions F () and G ()
dimensionless. (Star in means that we have used a renormalized value of the effective
mass so that = m kF / 2 , but this is again just a matter of convenience.) Explicitly,
f, (~p, p~0 ) = F s () + F a ()

.

(1.4)

In general, the interaction function is not known. However, if the interaction is weak,
one can relate f to the pair interaction potential. The microscopic theory of a Fermi
liquid [3] shows that the Landau interaction function is related to the interaction vertex
, (~p, p~0 0 |~p + ~q, + , p~0 ~q, ), which describes scattering of two fermions in initials states with momenta, energies, and spin projections p~, , and p~0 , 0 , , respectively
into the final states p~ + ~q, + , and p~0 , 0 + , via
f, (~p, p~0 ) = Z 2 lim , (~p, p~0 0 |~p + ~q, + , p~0 ~q, ) ,
q/0

where Z is the renormalization factor of the Greens function.


Example 2 At first order in the interaction, is represented by two diagrams, one of which
is obtained from the other by a permutation of outgoing lines. If the interaction conserves
spin,
, (~p, p~0 0 |~p + ~q, + , p~0 ~q, ) = U (~q) U (~p p~0 + q) .
Using an SU(2) identity
=

1
( +

) ,
2

this expression for reduces to

, (~p, p~0 |~p + ~q, + , p~ ~q, ) =

1
U (~q) U (~p p~0 + q)
2

1


U (~p p~0 + q) .
2

Exchanging indices and and taking the limit q 0 (at this order the vertex does not
depend on the energy), we obtain for f

f, (~p, p~ ) =

1
1
U (0) U (|~p p~0 |)

U (|~p p~0 |) .
2
2
7

[Because this relation holds only at first order of the perturbation theory, Z still equals to 1.]
On the Fermi surface, |~p p~0 | = 2pF sin /2. Comparing this expression with Eq.(1.4), we
find that

1
F s () = 1 U (0) U (2pF sin /2)
2
1
F a () = U (2pF sin /2) .
2

(1.5)

Notice that a repulsive interaction (U > 0) corresponds to the attraction in the spin-exchange
channel (F a < 0)this is the origin of the ferromagnetic instability.

C.

General strategy of the Fermi-liquid theory

One may wonder what one can achieve introducing unknown phenomenological quantities,
such as the interaction function or its charge and spin components. FL theory allows one
to express general thermodynamic characteristic of a liquid (effective mass, compressibility,
spin susceptibility, etc.) via angular averages of Landau functionsF s () and F a () . Some
quantities depend on the same averages and thus one express such quantities via each other.
The relationship between such quantities can be checked by comparison with the experiment.

D.

Effective mass

As a first application of the Landau theory, lets consider the effective mass. In a Galileaninvariant system, the momentum per unit volume coincides with the flux of mass:
Z

Tr

Z
d3 p
d3 p
p
~
n

=
Tr
m n
,
(2)3
(2)3 ~p

where m is the bare mass. Taking the variation of both sides of this equation, we obtain
Z

Tr
Now,

"

Z

d3 p
d3 p

p
~

n
=
Tr
n
+
n .
3
3m
~p
~p
(2)
(2)
Z
d3 p0 f (~p, p~0 )

= Tr0

n (~p0 ) .
3
~p
~
p
(2)

For a spin-isotropic liquid, Eq.(1.4) reduces to


Z

"Z
#
Z
Z
d3 p
d3 p
d3 p
d3 p0 f (~p, p~0 )
0
p~n =
m n +
m
n (~p ) n (~p) .
~p
(2)3
(2)3 ~p
(2)3
(2)3

(1.6)

In the second term re-label the variables p~ p~0 , p~0 p , use the fact that f (~p, p~0 ) = f (~p0 , p~)
and integrate by parts
Z

Z
Z
d3 p
d3 p
d3 p Z d3 p0
p0 )
0 n (~
n

p
~
n
=
m
f
(~
p
,
p
~
)
n (~p) .
~p0
(2)3
(2)3 ~p
(2)3 (2)3

Because this equation should be satisfied for an arbitrary variation n (~p) ,


p~
Z d3 p0
p0 )
0 n (~
=

f
(~
p
,
p
~
)
.
m
~p
~p0
(2)3
Near the Fermi surface, the quasi-particle energy can be always written as
(p) = vF (p pF )
so that
pF

= vF p = p,
~p
m
where p is the unit vector in the direction of p~, and thus
p~
pF p Z d3 p0
p0 )
0 n (~
=
f
(~
p
,
p
~
)
m
m
~p0
(2)3
Now

n (~p0 )
(~p0 ) n
=
.
~p0
~p0 0

Near the Fermi surface,


n
= (0 EF )

and

p~
pF p 1 Z d0
pF p0
= + F
f (~p, p~0 )
m
m
2
4
m

or

p~
pF p 1 Z d0
0
pF p 1 pF Z d0
0 pF p
= + F
f (~p, p~ ) = +
f (~p, p~0 ) pF p0 .
2
m
m
2
4
m
m
2
4
Setting |~p| = pF and multiplying both sides of the equation by p, we obtain
1
1
1 pF Z d
= +
f () cos
m
m
2 2
4
or

1
1
pF Z d
=

f () cos .
(1.7)
m
m 2 2 4
This is the Landaus formula for the effective mass. Noticing that f () = (1/2)TrTr0 f () =
(2/F ) F () = (2 2 /m pF ) , the last equation can be reduced to
Z
m
d
=1+
F () cos 1 + F1 .
m
4

Although we do not know the explicit form of f () , some useful conclusions can be made
already from the most general form [Eq.(1.7)]. If hf () cos i is negative, 1/m > 1/m
m < m; conversely, if hf () cos i is positive, m > m. For the weak short-range interaction,
we know that m > m, whereas for a weak long-range (Coulomb) interaction m < m. Now
we see that both of these cases are described by the Landaus formula. Recalling the weakcoupling form of F () [Eq.(1.5)], we find

"

!#

Z
d
d
1

F1 =
F () cos =
U (0) U 2pF sin
cos
4
4
2
2

!
Z
d 1

=
U 2pF sin
cos .
4 2
2
If U is repulsive and picked at small where cos is positive, F1 < 0 and m < m. This

case corresponds to a screened Coulomb potential. If U is repulsive and depends on only


weakly (short-range interaction), the angular integral is dominated by values of close to
, where cos < 0. Then m > m. In general, forward scattering reduces the effective mass,
whereas large-angle scattering enhances it.

E.
1.

Spin susceptibility
Free electrons

We start with free electrons. An electron with spin s has a magnetic moment = 2B s,
where B = eh/2mc is the Bohr magneton. Zeeman splitting of energy levels is E =
E E = B H (B H) = 2B H. The number of spin-up and -down electrons is found
as an integral over the density of states
1 Z EF B H
d () .
n, =
2 0
The Fermi energy now also depends on the magnetic field but the dependence is only
quadratic (why?), whereas the Zeeman terms are linear in H. For weak fields, one can neglect the field-dependence of EF . Total magnetic moment per unit volumemagnetizationis
found by expanding in H

"
#
Z EF B H
B Z EF +B H
M = B (n n ) =
d ()
d ()
2
0
0
#
"
Z EF
B Z E F
=
d () + B HF
d () + B HF
2
0
0
= 2B HF .

10

Spin susceptibility
=

M
= 2B F
H

(1.8)

is finite and positive, which means that spins are oriented along the external magnetic field.
Notice that if, for some reason the bare g- factor of electrons is not equal to 2, then
Eq.(1.8) changes to
g
= 2B F
2
2.

(1.9)

Fermi liquid

In a Fermi liquid, the energy of a quasi-particle in a magnetic field is changed not only
due to the Zeeman splitting (as in the Fermi gas) but also because the occupation number
is changed. This effect brings in an additional term in the energy as the energy is related
to the occupation number. This is expressed by following equation
Z

= B H
+ Tr

d3 p0
p, p~0 ) n (~p0 ) .
3 f (~
(2)

The first term is just the Zeeman splitting, the second one comes from the interaction. Now,
n

n=
and we have an equation for
Z

d3 p0
n

p, p~0 ) 0 (~p0 ) .
3 f (~

(2)

= B H
+ Tr0
Replacing

by the delta-function and setting |~p| = pF


Z

(pF p) = B H

Tr0 F

d0
f (pF p, pF p0 ) (pF p0 ) .
4

(1.10)

Lets try a solution in the following form


=

B
gH
,
2

(1.11)

where g has the meaning of an effective g factor. For free electrons, g = 2. Recalling that

0 and substituting (1.11) into (1.11), we obtain


F f = F () I + G ()

Z
i
B
B
d0 h s
0
gH
= B H
Tr
F () I + F a ()

0 gH
0 .
2
4
2

11

The term containing F in the integral vanishes because Pauli matrices are traceless: Tr = 0.
The term containing G is re-arranged using the identity Tr(

0 )
0 = 2
upon which we
get
g
g
= 1 F0a ,
2
2
where

F0a

d a
F () .
4

g=

2
.
1 + F0a

Therefore,

Substituting this result into Eq.(1.9) and replacing F by its renormalized value F , we
obtain the spin susceptibility of a Fermi liquid
=

2B F
1 + F1s
=

,
1 + F0a
1 + F0a

where is the susceptibility of free fermions. Notice that differs from both because
the effective mass is renormalized and because the g-factor differs from 2. When F0a = 1,
the g-factor and susceptibility diverge signaling a ferromagnetic instability. However, even
if G0 = 0, is still renormalized in proportion to the effective mass. If m diverges (which
is a signature of a metal-insulator transition of Mott type), diverges as well. Recall
that [Eq.(1.5)] in the weak-coupling limit F a () = 12 U (2pF sin /2) . Thus, for repulsive
interactions g > 2 which signals a ferromagnetic tendency. This is another manifestation of
general principle that repulsively interacting fermions tend to have spin aligned to minimize
the energy of repulsion. For normal He3 , G0 2/3.

F.

Zero sound

All gases and liquids support sound waves. Even ideal gases have finite compressibilities
and therefor finite sound velocities

s=

P
.

For example, in an ideal Boltzmann gas P = nT = T /m and


s

s=

1
T
= vT ,
m
3

12

where vT =

3T /m is the rms thermal velocity. In an ideal Fermi gas, P = (2/3)E where

E = (2/5) nEF is the ground state energy and


1
s = vF .
3
It seems that interactions are not essential for sound propagation. This conclusion is not
true. In a sound wave, all thermodynamic characteristicsdensity n (r, t) , pressure P (r, t) ,
temperature T (r, t) (if the experiment is performed under adiabatic conditions), etc.vary
in space and time in sync with the sound wave. Thus we are dealing with a non-equilibrium
situation. To describe a non-equilibrium situation by a set of local and time-dependent
quantities, one needs to maintain local and temporal equilibrium. When the sound wave
arrives to an initially unperturbed region, this region is driven away from its equilibrium
state. Collisions between molecules (or fermions in a Fermi gas) has to be frequent enough
to establish a new equilibrium state before the sound wave leaves the region. For sound
propagation, the characteristic spatial scale is the wavelength and characteristic time is
the period 2/. If the mean free path and time are l and , respectively, the conditions of
establishing local and temporal equilibrium are
l
2/ .
Therefore, collisions (interactions) are essential to ensure that the sound propagation occurs
under quasi-equilibrium conditions. For classical gases and liquids these conditions are
satisfied for all not extremely high frequencies. For example, the mean free path of molecules
in air at P = 1 atm and T = 300 K is about l ' 105 cm and s ' 300 m/s=3104 cm/s.
Condition l takes the form

3 104 cm/s
s
=
= 3 109 s1 .
l
105 cm

In a liquid, l is of order of the inter-molecule separation 'few 107 cm. As long as 107
cm, equilibrium is maintained.
In a Fermi liquid, the situation is different because l and increase as temperature goes
down. Local equilibrium is maintained as long as (T ) 1. Now, (T ) = 1/AT 2 so
at fixed temperature sound of frequency 1 (T ) = AT 2 cannot proceed in a quasiequilibrium manner.
13

What happens when we start at small frequency and then increase it? As long as
(T ) 1, a Fermi liquid supports normal sound (in the FL theory it is called first
sound). The velocity of this sound is linked to the sound velocity in the ideal Fermi gas but
differs from because of renormalization
vF
s1 = (1 + F0s )1/2 (1 + F1s )1/2 ,
3
where F0s =

R d s
F () . The first sound corresponds to the oscillations of fermion number
4

density in sync with the wave. Because the number density fixes the radius of the Fermi
sphere, pF oscillates in time and space. As the product increases, sound absorption does
so too, at when becomes impossible. However, as increases further and becomes 1,
another type of sound wave emerges. These waves are called zero sound. These waves can
propagate even at T = 0, when = as they do not require local equilibrium. Another
difference between zero- and first sound is that in the first sound-wave the shape of the Fermi
surface remains spherical whereas its radius changes. In the zero-sound wave, the shape of
the Fermi surface may be changed without changing its volume. In a simplest zero-sound
wave [4], the occupation number is

~
n = ( EF ) k ei(k~rt)

= C

cos
s0 /vF cos

where s0 > vF is the zero-sound velocity. The Fermi surface is an egg-shaped spheroid, with
the narrow end pointing in the direction of the wave propagation. The condition s0 > vF ,
always satisfied for zero sound, means absence of Landau damping.
As well as s1 , the zero-sound velocity s0 depends on Landau parameters F0s and F1s ,
although the functional dependence is more complicated. At the same time, F1s can be
extracted from the effective mass (measured from the specific heat) and F0s is extracted from
the compressibility. Having two experimentally determined parameters F0s and F1s , one can
substitute them into theoretical expressions for s1 and s0 and compare them with measured
sound velocities. The agreement (see the plot) is quite satisfactory. This comparison provides
a quantitative check for the Fermi-liquid theory.
H. J. Schulz, Fermi liquids and non-Fermi liquids, in: Proceedings of Les Houches Summer
School LXI, edited by E. Akkermans et al., Elsevier (Amsterdam), 1995, p. 553. Available

14

at xxx.lanl.gov/abs/comd-mat/9503150.

[1] D. Yue, L. I. Glazman and K. A. Matveev, Phys. Rev. B 49, 1966 (1994).
[2] Mahan, Many-body physics
[3] A. A. Abrikosov, L. P. Gorkov, and I. E. Dzyaloshinski, Methods of quantum field theory in
statistical physics, (Dover Publications, New York, 1963).
[4] E.M. Lifshitz and L.P. Pitaevskii, Statistical Physics II, (Course of Theoretical Physics, v.
IX), Pergamon Press, Oxford, 1980.
[5] D. Pines and P. Nozieres, The Theory of Quantum Liquids, Benjamin, New York, 1969.
[6] L. D. Landau, Zh. Eksp. Teor. Fiz. 30, 1058 (1956)[Sov. Phys. JETP 3, 920 (1957)]. Reprinted
in: D. Pines, The Many-Body Problem, Benjamin, Reading, MA.
[7] B. L. Altshuler and A. G. Aronov, in Electron-electron interactions in disordered conductors,
edited by A. L. Efros and M. Pollak (Elsevier, 1985), p. 1.
[8] G. Zala, B. N. Narozhny, and I. L. Aleiner, Phys. Rev. B 64, 214204 (2001).
[9] E. Abrahams, S. V. Kravchenko, and M. P. Sarachik, Rev. Mod. Phys. 73, 251 (2001).
[10] B. L. Altshuler, D. L. Maslov, and V. M. Pudalov,
Physica E 9, 209 (2001).

15

Das könnte Ihnen auch gefallen