Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Patterson, Bailey - Solid State Physics Introduction to theory

.pdf
Скачиваний:
1117
Добавлен:
08.01.2014
Размер:
7.07 Mб
Скачать

4.3 The Electron–Phonon Interaction

235

 

 

 

Thus

 

 

 

cb

(4.64)

P = [K (0) 1]ε0 E = d

E .

 

a

 

For the high-frequency or optic case r¨ → ∞, and r → 0 because the ions cannot follow the high-frequency fields so

P = dE =[K () 1]ε0 E .

(4.65)

From the above

 

 

d = [K () 1]ε0 ,

 

(4.66)

d bc = [K (0) 1]ε

0 .

(4.67)

a

 

 

We can use the above to get an expression for the polarization, which in turn can be used to determine the electron–phonon interaction. First we need to evaluate P.

We work out the polarization for the longitudinal optic mode, as that is all that is needed. Let

r = rT + rL ,

(4.68)

where T and L denote transverse and longitudinal. Since we assume

 

rT = v exp[i(q r +ωt)] , v a constant ,

(4.69a)

then

 

rT = iq rT = 0 ,

(4.69b)

by definition since q is the direction of motion of the vibrational wave and is perpendicular to rT. There is no free charge to consider, so

D = (ε0E + P) = (ε0 E + dE + cr) = 0

or

[(ε0 + d )E + crL ] = 0 , using (4.69b). This gives as a solution for E

 

E =

c

 

r .

 

 

ε0 + d

 

 

 

 

 

 

L

 

Therefore

 

 

 

 

 

 

P

= cr

+ dE =

 

cε0

r .

ε0 + d

L

L

 

 

L

If

rL = rL (0) exp(iωLt) ,

(4.70)

(4.71)

(4.72)

(4.73a)

236 4 The Interaction of Electrons and Lattice Vibrations

and

rT = rT (0) exp(iωTt) ,

then

rL = −ωL2rL ,

and

rT = −ωT2rT .

Thus by Eqs. (4.58a) and (4.71)

r

= ar

 

cb

r .

 

ε0 + d

 

L

 

 

L

 

 

L

 

Also, using (4.71) and (4.58a)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

rT = arT ,

 

 

 

so

 

 

 

 

 

 

 

 

 

 

 

 

 

 

a = −ωT2 .

 

 

 

Using Eqs. (4.66) and (4.67)

 

 

 

 

 

 

 

 

 

 

 

a

 

 

bc

 

 

= a

K (0)

,

 

 

ε0 + d

 

 

 

 

 

 

 

K ()

 

and so by (4.74a), (4.75) and (4.77)

 

 

 

 

 

 

 

 

 

2

 

 

K (0)

 

2

K (0)

 

ωL = −a

 

 

 

 

=ωT

 

 

,

K ()

K ()

 

 

 

 

 

 

(4.73b)

(4.74a)

(4.74b)

(4.75)

(4.76)

(4.77)

(4.78)

(4.79)

which is known as the LST (for Lyddane–Sachs–Teller) equation. See also Born and Huang [46 p. 87]. This will be further discussed in Chap. 9. Continuing, by (4.66),

ε0 + d = K ()ε0 ,

 

(4.80)

and by (4.67)

 

 

 

d [K (0) 1]ε0

= bc

,

(4.81)

 

a

 

 

from which we determine by (4.60), (4.77), (4.78), (4.80), and (4.81)

c =ωT

Nμ

ε0 K (0) K () .

(4.82)

 

V

 

 

 

 

 

 

 

 

 

 

 

4.3 The Electron–Phonon Interaction

237

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Using (4.72) and the LST equation we find

 

 

 

 

 

 

 

 

 

 

 

P = ωL

ε0

Nμ

 

 

1

 

 

 

 

 

 

K (0) K ()rL ,

(4.83)

 

V

 

 

 

K (0)K ()

 

 

 

 

 

 

 

 

 

 

 

or if we define

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

αc

=

 

 

e2

 

1

 

 

1

 

,

(4.84)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

8πε0 ωL r0 K

 

 

 

 

 

 

 

 

 

 

with

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1

 

=

 

1

 

1

 

 

 

,

(4.85)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

K

K ()

K (0)

 

 

 

 

 

 

 

 

 

 

and

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

r0

=

 

 

 

,

 

 

 

 

(4.86)

 

 

 

 

 

 

 

 

 

2mωL

 

 

 

 

 

we can write a more convenient expression for P. Note we can think of K¯ as the effective dielectric constant for the ion displacements. The quantity r0 is called the radius of the polaron. A simple argument can be given to see why this is a good interpretation. The uncertainty in the energy of the electron due to emission or absorption of virtual phonons is

 

E =

ωL ,

(4.87)

and if

 

 

 

 

 

E

2

( k)2 ,

(4.88)

 

 

 

2m

 

 

 

 

then

 

 

 

 

1 r

=

.

(4.89)

k

0

 

2mωL

 

 

 

 

The quantity αc is called the coupling constant and it can have values considerably less than 1 for for direct band gap semiconductors or greater than 1 for insulators. Using the above definitions:

P = ε ω

Nμαc 8π ωL r r

 

0 L

V

e2

0

L

(4.90)

 

 

 

ArL.

238 4 The Interaction of Electrons and Lattice Vibrations

The Electron–Phonon Interaction due to the Polarization (A)

In the continuum approximation appropriate for large polarons, we can write the electron–phonon interaction as coming from dipole moments interacting with the gradient of the potential due to the electron (i.e. a dipole moment dotted with an electric field, e > 0) so

H ep =

e

 

P(r)

 

1

 

dr =

e

P(

r) (r re )

dr .

(4.91)

4πε

 

 

 

r re

 

4πε0

 

 

0

 

 

 

r r

3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

e

 

 

 

Since P = ArL and we have determined A, we need to write an expression for rL. In the usual way we can express rL at lattice position Rn in terms of an expan-

sion in the normal modes for LO phonons (see Sect. 2.3.2):

r

= r

r

=

1

 

 

 

 

 

 

 

e

+

(q)

e

(q)

 

N

 

 

Q(q)

 

 

 

 

 

 

exp(iq R ).

Ln

n+

n

 

 

 

 

q

 

 

 

 

 

 

 

 

 

m

 

 

m

 

n

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+

 

 

 

 

 

The polarization vectors are normalized so

 

 

 

 

 

 

 

 

 

 

 

For long-wavelength LO modes

 

e+

 

2 +

 

e

 

 

2 =1.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

e

+

= −e

 

 

 

m.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+

 

 

 

 

 

 

Then we find a solution for the LO modes as

 

 

 

 

 

 

 

 

 

 

 

e+(q) = i

 

μ eˆ(q) ,

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

m+

 

 

 

 

 

 

 

 

 

 

 

e(q) = −i

 

 

 

μ eˆ(q) ,

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

m

 

 

 

 

 

 

 

where

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

eˆ(q) =

q

 

as q → ∞ .

 

 

 

 

 

 

 

 

q

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Note the i allows us to satisfy

e(q) = e (q) ,

as required. Thus

rLn = 1

q iQ(q)eˆ(q) exp(iq Rn ) ,

Nμ

 

(4.92)

(4.93)

(4.94)

(4.95a)

(4.95b)

(4.96)

(4.97)

4.3 The Electron–Phonon Interaction 239

or in the continuum approximation

rLn = 1

q iQ(q)eˆ(q) exp(iq r) .

(4.98)

Nμ

 

 

Following the usual procedure:

 

 

 

Q(q) =

1

(a+q aq )

(4.99)

 

i

2ωL

 

(compare with Eqs. (2.140), (2.141)). Substituting and making a change in dummy summation variable:

 

 

rL = −

 

 

 

+

iq r

+ aqe

iq r

)

q

.

 

 

 

 

q (aq e

 

 

q

 

 

 

 

2NμωL

 

 

 

 

 

 

 

 

 

 

Thus

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

H ep = −

ωL

4παcr0

dr r re q q (aq+eiq r + aqeiq r ) .

 

 

4π

 

V

r re

3 q

 

 

 

 

 

 

 

Using the identity from Madelung [4.26],

 

 

 

 

 

 

 

 

exp[±exp(iq r)]

 

(r re )

dr = 4πi

q

exp(±iq r ) ,

 

 

 

 

 

 

 

 

r re

 

3

 

 

 

q2

 

 

 

e

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

we find

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

H ep = i

ωL r0

 

4παc

q 1

[aq exp(iq re ) aq+ exp(iq re )] .

 

 

 

 

V

q

 

 

 

 

 

 

 

 

 

 

 

(4.100)

(4.101)

(4.102)

(4.103)

Energy and Effective Mass (A)

We consider only processes in which the polarizable medium is at absolute zero, and for which the electron does not have enough energy to create real optical phonons. We consider only the process described in Fig. 4.6. That is we consider the modification of self-energy of the electron due to virtual phonons. In perturbation theory we have as ground state |k, 0q with energy

Ek =

2k 2

(4.104)

2m

 

 

and no phonons. For the excited (virtual) state we have one phonon, |k – q, 1q . By ordinary Rayleigh-Schrödinger perturbation theory, the perturbed energy of the ground state to second order is:

Ek,0 = Ek(0),0 +

k,0 H ep k,0 + q

k q, 1 H ep k,0

2

(4.105)

E(0)

E(0)

.

 

 

 

 

 

 

k,0

k q, 1

 

 

240 4 The Interaction of Electrons and Lattice Vibrations

But

 

 

 

 

 

E

(0)

=

 

2k 2

 

 

 

 

 

 

 

 

k

,0

 

 

 

 

,

 

 

 

 

 

 

 

 

 

2m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

k,0 H ep k,0

= 0 ,

 

 

 

 

 

 

Ek(0)q, 1 =

 

 

2

 

 

(k q)2 + ωL ,

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

so

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

E(0)

E(0)

=

 

2

 

 

(2k q q2 ) ω

 

,

(4.106)

 

 

 

 

 

 

 

L

 

 

 

 

 

 

 

 

 

k,0

 

k q, 1

 

 

2m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

and

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

k q, 1 H

ep

k,0 = −i

ω

L

r

 

4παc

 

1 k q, 1 e(iqre )a+k,0

. (4.107)

 

 

 

0

 

 

V

 

 

 

 

 

qq

 

q

 

 

 

 

 

 

k q

 

 

 

k

 

q

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

k

Fig. 4.6. Self-energy Feynman diagram (for interaction of electron and virtual phonon)

Since

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1 aq+ 0

=1 ,

 

 

 

 

 

 

(4.108a)

k q exp(iqre ) k

 

= δq,q,

 

 

(4.108b)

we have

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

k q, 1 H

ep

k,0

2 = (

 

ω

L

)

2 r

 

4παc

1

CH2

,

(4.109)

 

 

 

 

 

 

0

V

q2

q2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

where

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

C2

= ( ω

L

)2 r

 

4παc

.

 

 

 

(4.110)

 

 

 

 

 

 

 

 

 

H

 

 

 

0

 

 

V

 

 

 

 

4.3 The Electron–Phonon Interaction 241

Replacing

 

 

 

q

by

 

V

 

dq ,

 

 

 

 

 

 

 

(2π)3

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

we have

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Ek,0 =

2k 2

+

VCH2

1

 

 

 

 

 

dq

 

 

 

.

2m

(2π)3

q2 2k 2

 

(2k q q2 )

 

 

 

 

 

 

 

 

 

 

 

 

 

ω

L

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2m

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

For small k we can show (see Problem 4.5)

Ek,0 αc ωL +

2k 2

,

2m

 

 

where

m =

m

 

.

1(αc / 6)

(4.111)

(4.112)

(4.113)

Thus the self-energy is increased by the interaction of the cloud of virtual phonons surrounding the electrons.

Experiments and Numerical Results (A)

A discussion of experimental results for large polarons can be found in the paper by Appel [4.2, pp. 261-276]. Appel (pp. 366-391) also gives experimental results for small polarons. Polarons are real. However, there is not the kind of comprehensive comparisons of theory and experiment that one might desire. Cyclotron resonance and polaron mobility experiments are common experiments cited. Difficulties abound, however. For example, to determine m** accurately, m* is needed. Of course m* depends on the band structure that then must be accurately known. Crystal purity is an important but limiting consideration in many experiments. The chapter by F. C. Brown in the book edited by Kuper and Whitfield [4.23] also reviews rather thoroughly the experimental situation. Some typical values for the coupling constant αc (from Appel), are given below. Experimental estimates of αc are also given by Mahan [4.27] on p. 508.

Table 4.4. Polaron coupling constant

Material αc

KBr 3.70

GaAs 0.031

InSb 0.015

CdS 0.65

CdTe 0.39

242 4 The Interaction of Electrons and Lattice Vibrations

4.4Brief Comments on Electron–Electron Interactions

(B)

A few comments on electron–electron interactions have already been made in Chap. 3 (Sects. 3.1.4 and 3.2.2) and in the introduction to this chapter. Chapter 3 discussed in some detail the density functional technique (DFT), in which the density function plays a central role for accounting for effects of electron–electron interactions. Kohn [4.20] has given a nice summary of the limitation of this model. The DFT has become the traditional way nowadays for calculating the electronic structure of crystalline (and to some extent other types of) condensed matter. For actual electronic densities of interest in metals it has always been difficult to treat electron–electron interactions. We give below earlier results that have been obtained for high and low densities.

Results, which include correlations or the effect of electron–electron interactions, are available for a uniform electron gas with a uniform positive background (jellium). The results given below are in units of Rydberg (R), see Appendix A. If ρ is the average electron density,

 

 

3

1/ 3

r

=

 

 

 

s

 

 

 

 

 

4πρ

is the average distance between electrons. For high density (rs << 1), the theory of Gell-mann and Bruckner gives for the energy per electron

E

=

2.21

 

0.916 + 0.062 ln r 0.096 + (higher order terms)(R

 

) .

N

r2

 

 

 

rs

s

 

 

 

s

 

 

 

 

 

For low densities (rs >> 1) the ideas of Wigner can be extended to give

E

= −

1.792 +

2.66

+ higher order terms in r1/ 2 .

N

r3 / 2

 

rs

s

 

 

 

s

 

In the intermediate regime of metallic densities, the following expression is approximately true:

E = 2.21 0.916 + −

0.031ln rs 0.115 (R) ,

N rs2 rs

for 1.8 ≤ rs ≤ 5.5. See Katsnelson et al [4.16]. This book is also excellent for DFT. The best techniques for treating electrons in interaction that has been discussed in this book are the Hartree and Hartree–Fock approximation and especially the density functional method. As already mentioned, the Hartree–Fock method can give wrong results because it neglects the correlations between electrons with antiparallel spins. In fact, the correlation energy of a system is often defined as the

4.4 Brief Comments on Electron–Electron Interactions (B) 243

difference between the exact energy (less the relativistic corrections if necessary) and the Hartree–Fock energy.

Even if we limit ourselves to techniques derivable from the variational principle, we can calculate the correlation energy at least in principle. All we have to do is to use a better trial wave function than a single Slater determinant. One way to do this is to use a linear combination of several Slater determinants (the method of superposition of configurations). The other method is to include interelectronic coordinates r12 = |r1 r2| in our trial wave function. In both methods there would be several independent functions weighted with coefficients to be determined by the variational principle. Both of these techniques are practical for atoms and molecules with a limited number of electrons. Both become much too complex when applied to solids. In solids, cleverer techniques have to be employed. Mattuck [4.28] will introduce you to some of these clever ideas and do it in a simple, understandable way, and density functional techniques (see Chap. 3) have become very useful, at least for ground-state properties.

It is well to keep in mind that most calculations of electronic properties in real solids have been done in some sort of one-electron approximation and they treat electron–electron interactions only approximately. There is no reason to suppose that electron correlations do not cause many types of new phenomena. For example, Mott has proposed that if we could bring metallic atoms slowly together to form a solid there would still be a sudden (so-called Mott) transition to the conducting or metallic state at a given distance between the atoms.6 This sudden transition would be caused by electron–electron interactions and is to be contrasted with the older idea of conduction at all interatomic separations. The Mott view differs from the Bloch view that states that any material with well separated energy bands that are either filled or empty should be an insulator while any material with only partly filled bands (say about half-filled) should be a metal. Consider, for example, a hypothetical sodium lattice with N atoms in which the Na atoms are 1 meter apart. Let us consider the electrons that are in the outer unfilled shells. The Bloch theory says to put these electrons into the N lowest states in the conduction band. This leaves N higher states in the conduction band for conduction, and the lattice (even with the sodium atoms well separated) is a metal. This description allows two electrons with opposite spin to be on the same atom without taking into account the resulting increase in energy due to Coulomb repulsion. A better description would be to place just one electron on each atom. Now, the Coulomb potential energy is lower, but since we are using localized states, the kinetic energy is higher. For separations of 1 meter, the lowering of potential energy must dominate. In the better description as provided by the localized model, conduction takes place only by electrons hopping onto atoms that already have an outer electron. This requires considerable energy and so we expect the material to behave as an insulator at large atomic separations. Since the Bloch model so often works, we expect (usually) that the kinetic energy term dominates at actual interatomic spacing. Mott predicted that the transition to a metal from an insulator as the interatomic spacing is varied (in a situation such as we have

6 See Mott [4.31].

244 4 The Interaction of Electrons and Lattice Vibrations

described) should be a sudden transition. By now, many examples are known, NiO was one of the first examples of “Mott–Hubbard” insulators – following current usage. Anderson has predicted another kind of metal–insulator transition due to disorder.6 Anderson’s ideas are also discussed in Sect. 12.9.

Kohn has suggested another effect that may be due to electron–electron interactions. These interactions cause singularities in the dielectric constant (see, e.g., (9.167)) as a function of wave vector that can be picked up in the dispersion relation of lattice vibrations. This Kohn effect appears to offer a means of mapping out the Fermi surface.7 Electron–electron interactions may also alter our views of impurity states.8 We should continue to be hopeful about the possibility of finding new effects due to electron–electron interactions.9

4.5 The Boltzmann Equation and Electrical Conductivity

4.5.1 Derivation of the Boltzmann Differential Equation (B)

In this section, the Boltzmann equation for an electron gas will be derived. The principle lack of rigor will be our assumption that the electrons are described by wave packets made of one-electron Bloch wave packets (Bloch wave packets incorporate the effect of the fields due to the lattice ions which by definition change rapidly over inter ionic distances). We also assume these wave packets do not spread appreciably over times of interest. The external fields and temperatures will also be assumed to vary slowly over distances of the order of the lattice spacing.

Later, we will note that the Boltzmann equation is only relatively simple to solve in an iterated first order form when a relaxation time can be defined. The use of a relaxation time will further require that the collisions of the electrons with phonons (for example) do not appreciably alter their energies, that is that the relevant phonon energies are negligible compared to the electrons energies so that the scattering of the electrons may be regarded as elastic.

We start with the distribution function fkσ(r,t), where the normalization is such that

dkdr fkσ (r,t) (2π)3

is the number of electrons in dk (= dkxdkydkz) and dr (= dxdydz) at time t with spin σ. In equilibrium, with a uniform distribution, fkσ fk0σ becomes the Fermi–Dirac distribution.

7See [4.19]. See also Sect. 9.5.3.

8See Langer and Vosko [4.24].

9See also Sect. 12.8.3 where the half-integral quantum Hall effect is discussed.

Соседние файлы в предмете Химия