Electron-Phonon Interaction: 2.1 Phonons and Lattice Dynamics
Electron-Phonon Interaction: 2.1 Phonons and Lattice Dynamics
Electron-Phonon Interaction: 2.1 Phonons and Lattice Dynamics
ElectronPhonon Interaction
The electronphonon interaction is important not only in creating the phonon scattering of the electrons but also in the formation of Cooper pairs. This interaction is
indeed the cause of the superconductivity. The Frohlich Hamiltonian is derived. A
phonon exchange can generate an attraction between a pair of electrons.
(2.1)
j
=x,y,z
u j
V
u j
+
0
2 V
1
u j u k
+ ,
2 j k
u j u k 0
(2.2)
V
u j
= 0.
(2.3)
15
16
ElectronPhonon Interaction
For small oscillations we may keep terms of the second order in u j only. We
then have
1
A jk u j u k ,
(2.4)
V V
2
j
k
A jk
2 V
u j u k
.
(2.5)
The prime ( ) on V indicating the harmonic approximation, will be dropped hereafter. The kinetic energy of the system is
T
m
j
r 2j =
m
j
u 2j
m
j
u 2j .
(2.6)
A jk u j u k .
(2.7)
m
j
u 2j
1
j
3N
1
=1
(P2 + 2 Q 2 ),
(2.8)
where {Q , P } are the normal coordinates and momenta, and { } are normalmode frequencies. Note that there are exactly 3N normal modes.
Let us first calculate the heat capacity by means of classical statistical mechanics. This Hamiltonian H is quadratic in canonical variables (Q , P ). Hence the
equipartition theorem holds. We multiply the average thermal energy for each mode,
k B T , by the number of modes, 3N , and obtain 3N k B T for the average energy H .
Differentiating this 3N k B T with respect to T , we obtain 3N k B for the heat capacity,
which is in agreement with DulongPetits law: C = 3R. It is interesting to observe
that we obtained this result without knowing the actual distribution of normal-mode
frequencies. The fact that there are 3N normal modes played an important role.
Let us now use quantum theory and calculate the heat capacity based on Equation
(2.8). The energy eigenvalues of the Hamiltonian H are given by
E[{n k }] =
1
+ n ,
n = 0, 1, 2, .
(2.9)
17
We can interpret Equation (2.9) in terms of phonons as follows: the energy of the
lattice vibrations is characterized by the set of the numbers of phonons {n } in the
normal modes {}. Taking the canonical-ensemble average of Equation (2.9), we
obtain
E[{n}] =
1
1
+ n =
+ f 0 ( ) E(T ),
2
2
(2.10)
where
f 0 ()
1
exp(/k B T ) 1
(2.11)
E(T ) = E 0 +
E0
1
2
d f 0 ()D(),
(2.12)
d D(),
(2.13)
0
0
where D() is the density of states (modes) in angular frequency defined such that
Number of modes in the interval (, + d) D()d.
(2.14)
E
CV =
T
=
V
d
f 0 ()
D().
T
(2.15)
This expression was obtained under the harmonic approximation only, which is
expected to be valid at very low temperatures.
To proceed further, we have to know the density of normal modes, D(). To
find the set of characteristic frequencies { } requires solving an algebraic equation
of 3N -th order, and we need the frequency distribution for large N . This is not
a simple matter. In fact, a branch of mathematical physics whose principal aim is
to find the frequency distribution is called lattice dynamics. Figure 2.1 represents
a result obtained by Walker [2] after the analysis of the X-ray scattering data for
aluminum, based on lattice dynamics. Some remarkable features of the curve are
arbitrary units
18
ElectronPhonon Interaction
0.6
0.4
0.2
angular frequency
(2 x 1013 sec1)
0.8
1.0
max
(2.16)
for
max .
(2.17)
(C) A few sharp peaks, called van Hove singularities [3], exist below max .
The feature (A) is common to all crystals. The low frequency modes can be
described adequately in terms of longitudinal and transverse elastic waves. This
region can be represented very well by the Debyes continuum model [4], indicated
by the broken line and discussed in Book 1, Section 2.2. The feature (B) is connected
with the lattice structure. Briefly, no normal modes of extreme short wavelengths
(extreme high frequencies) exist. Hence there is a limit frequency max . The sharp
peaks, feature (C), were first predicted by van Hove [3] on topological grounds, and
these peaks are often refered to as van Hove singularities.
The cause of superconductivity lies in the electronphonon interaction [5]. The
microscopic theory can be formulated in terms of the generalized BCS Hamiltonian
[5], where all phonon variables are eliminated. In this sense the details of lattice dynamics are secondary to our main concern (superconductivity). The following point,
however, is noteworthy. All lattice dynamical calculations start with a real crystal
lattice. For example, to treat aluminum, we start with an fcc lattice having empirically
known lattice constants. The equations of motion for a set of ions are solved under
the assumption of a periodic lattice-box boundary condition. Thus, the k-vectors used
in lattice dynamics and Bloch electron dynamics are the same. The domain of the
k-vectors can be restricted to the same first Brillouin zone. Colloquially speaking,
phonons (bosons) and electrons (fermions) live together in the same Brillouin zone,
which is equivalent to say that electrons and phonons share the same house (crystal
19
lattice). This affinity between electrons and phonons makes the conservation of momentum in the electronphonon interaction physically meaningful. Thus, the fact that
the electronphonon interaction is the cause of superconductivity is not accidental.
(2.18)
(2.19)
20
ElectronPhonon Interaction
Since electrons follow phonons immediately for all q , the coefficient Cq can be
regarded as independent of q . If we further assume that the deviation is linear in
the scalar product uq q = qu q and again in the electron density n(r), we obtain
Cq = Aq qu q n(r).
(2.20)
(2.21)
(2.22)
1 2
( p + 2 q 2 ),
2
q u,
p q,
q ,
(2.23)
where we dropped the mode index q. If we assume the same quantum Hamiltonian
H and the quantum condition:
[q, p] = i,
[q, q] = [ p, p] = 0,
(2.24)
the quantum description of a harmonic oscillator is complete. The equations of motion are
q =
1
[q, H ] = p,
i
p =
1
[ p, H ] = 2 q,
i
(2.25)
(Problem 2.2.2).
We introduce the dimensionless complex dynamical variables:
a (2)1/2 ( p + iq),
a (2)1/2 ( p iq).
(2.26)
a = ia.
(2.27)
21
p = (/2)1/2 (a + a).
(2.28)
Thus, we may work entirely in terms of (a , a). After straightforward calculations, we obtain (Problem 2.2.3)
a a = (2)1 ( p + iq)( p iq)
1
= (2)1 [ p 2 + 2 q 2 + i(qp pq)] = H ,
2
1
aa = H + ,
2
(2.29)
aa a a [a, a ] = 1,
(2.30)
1
1
1
H = (a a + aa ) = a a +
n +
.
2
2
2
(2.31)
r
r
r
r
In summary, the quantum Hamiltonian and the quantum states of a harmonic oscillator can be simply described in terms of the bosonic second quantized operators
(a, a ).
We now go back to the case of the lattice normal modes. Each normal mode
corresponds to a harmonic oscillator characterized by (q, q ). The displacements
u q can be expressed as
uq = i
2q
1/2
(aq aq ),
(2.32)
[aq , ap ] aq ap ap aq = pq ,
[aq , ap ] = [aq , ap ] = 0.
(2.33)
22
ElectronPhonon Interaction
(2.34)
where (r) and (r) are annihilation and creation electron field operators, respectively, satisfying the following Fermi anticommutation rules:
(r), (r ) (r) (r ) + (r )(r) = (3) (r r ),
(r), (r ) = (r), (r ) = 0.
(2.35)
1
exp(ik r)ck ,
(V )1/2 k
(r) =
1
(3)
{ck , ck } ck ck + ck ck = k,k
,
(2.37)
d 3r
1
q
Aq qu q exp(iq r) (r)(r) + h.c. ,
(2.38)
where h.c. denotes the Hermitian conjugate. Using Equations (2.20), (2.32) and
(2.36), we can re-express Equation (2.38) as (Problem 2.2.3):
HF =
1
k
(2.39)
Vq ck+q ck aq
(Vq ck ck+q aq )
(2.40)
23
k
k q
q
k q
(a)
(b)
Fig. 2.3 Feynman diagrams representing (a) absorption and (b) emission of a phonon by an
electron
k
k cks cks
1
k
q
1
+ aq aq
2
Hel + H ph + HF H0 + V,
(V HF )
(2.41)
where the three sums represent: the total electron kinetic energy (Hel ), the total
phonon energy (H ph ), and the Frohlich interaction Hamiltonian HF , [see Equation
(2.39)].
For comparison we consider an electron gas system characterized by the
Hamiltonian
Hc =
k
k cks cks +
1
3, 4 | vc | 1, 2 c4 c3 c1 c2 Hel + Vc , (2.42)
2ks
k s
1 1
4 4
24
ElectronPhonon Interaction
3, 4|vc |1, 2 k3 s3 , k4 s4 | vc | k1 s1 , k2 s2
=
4 e2 k0 1
k +k ,k +k k k ,q s3 s1 s4 s2 .
V q2 1 2 3 4 1 3
(2.43)
The elementary interaction process can be represented by the diagram in Fig. 2.4.
The wavy horizontal line represents the instantaneous Coulomb interaction vc . The
net momentum of a pair of electrons is conserved:
k1 + k2 = k3 + k4 ,
(2.44)
k3
k4
k1
k2
time
k q
k +q
k +q
k q
(a )
( b)
Fig. 2.5 A one-phononexchange process generates the change in the momenta of two electrons
similar to that caused by the Coulomb interaction
25
used in quantum field theory [8, 9].) In the diagrams in Fig. 2.5, we follow the
motion of two electrons. We may therefore consider a system of two electrons and
obtain the effective Hamiltonian ve through a study of the evolution of two-body
density operator 2 . Hereafter, we shall drop the subscript 2 on indicating twobody system.
The system-density operator (t) changes in time, following the quantum Liouville equation:
i
(t)
= [H, ] H.
t
(2.45)
We assume the Hamiltonian H in Equation (2.41) and study the time evolution
of (t), using quantum many-body perturbation theory. Here, we sketch only the
important steps; more detailed calculations were given in FujitaGodoys book,
Quantum Statistical Theory of Superconductivity [10, 11].
Let us introduce a quantum Liouville operator
H H0 + V,
(2.46)
(2.47)
which is reasonable at 0 K, where there are no real phonons and only virtual phonons
are involved in the dynamical processes. We can then choose
phonon = |0 0| ,
(2.48)
for any q.
(2.49)
(2.50)
t
0
d V exp(i 1 H0 )V(t ) av .
(2.51)
26
ElectronPhonon Interaction
(2.52)
a > 0.
(2.53)
Let us now take momentum-state matrix elements of Equation (2.53). The lhs is
k1 s1 , k2 s2 | (t) | k3 s3 , k4 s4
(1, 2 ; 3, 4, t),
t
t
(2.54)
where we dropped the upper bar indicating the phonon vacuum average. The rhs
requires more sophisticated computations due to the Liouville operators (V, H0 ).
After lengthy but straightforward calculations, we obtain from Equation (2.53)
(1, 2 ; 3, 4, t) =
i1 [1, 2 | ve | 5, 6 2 (5, 6 ; 3, 4, t)
t
k s k s
5 5
6 6
3, 4 | ve | 1, 2 |Vq |2
q
k +k ,k +k k k ,q s3 s1 s4 s2 .
(3 1 )2 2 q2 1 2 3 4 3 1
(2.56)
Kroneckers delta k1 +k2 ,k3 +k4 in Equation (2.56) means that the net momentum
is conserved, since the phonon exchange is an internal interaction.
For comparison, consider the electron-gas system. The two-electron density matrix c for this system changes, following
c (1, 2 ; 3, 4, t) =
i1 [1, 2 | vc | 5, 6 c (5, 6 ; 3, 4, t)
t
k s k s
5 5
6 6
5, 6 | vc | 3, 4 c (1, 2 ; 5, 6, t)] ,
(2.57)
which is of the same form as Equation (2.55). The only differences are in the interaction matrix elements. Comparison between Equations (2.43) and (2.56) yields
References
27
4 e2 k0 1
V q2
q
|Vq |2
(k1 +q k1 )2 2 q2
(Coulomb interaction),
(phononexchange interaction).
(2.58)
(2.59)
(2.60)
then the effective interaction is attractive. Fourth, the atraction is greatest when
k1 +q k1 = 0, that is, when the phonon momentum q is parallel to the constantenergy (Fermi) surface. A bound electron-pair, called a Cooper pair, may be formed
by the phononexchange attraction, which was shown in 1956 by Cooper [12].
References
1. T. W. B. Kibble, Classical Mechanics, (McGraw-Hill, Maidenhead, England, 1966), pp.
166171.
2. C. B. Walker, Phys. Rev. 103, 547 (1956).
3. L. Van Hove, Phys. Rev. 89, 1189 (1953).
4. P. Debye, Ann. Physik 39, 789 (1912).
5. J. Bardeen, L. N. Cooper and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).
6. H. Frohlich, Phys. Rev. 79, 845 (1950).
7. H. Frohlich, Proc. R. Soc. London A 215, 291 (1950).
8. R. P. Feynman, Statistical Mechanics (Addison-Wesley, Reading, MA, 1972).
9. R. P. Feynman, Quantum Electrodynamics (Addison-Wesley, Reading, MA, 1961).
10. S. Fujita and S. Godoy, Quantum Statistical Theory of Superconductivity, (Plenum, New York,
1996), pp. 150153.
11. S. Fujita and S. Godoy, Theory of High Temperature Superconductivity, (Kluwer, Dordrecht,
2001), pp. 5458.
12. L. N. Cooper, Phys. Rev. 104, 1189 (1956).
http://www.springer.com/978-0-387-88205-5