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

Dressel.Gruner.Electrodynamics of Solids.2003

.pdf
Скачиваний:
66
Добавлен:
27.08.2013
Размер:
3.91 Mб
Скачать

128

5 Metals

Fig. 5.15. (a) Real part and (b) imaginary part of the dielectric constant ˆ(q, ω) of a freeelectron gas as a function of frequency and wavevector at T = 0 in three dimensions calculated using Eq. (5.4.21). The frequency axis is normalized to the Fermi frequency ωF = vFkF/2, and the wavevector axis is normalized to the Fermi wavevector kF.

5.4 Longitudinal response

129

Fig. 5.16. Frequency and wavevector dependence of (a) the real and (b) the imaginary part of the conductivity σˆ (q, ω) of a free-electron gas at T = 0 in three dimensions after Eq. (5.4.25). The frequency axis is normalized to the Fermi frequency ωF = kFvF/2 = EF/h¯ , and the wavevector axis is normalized to the Fermi wavevector kF.

130

5 Metals

and on substituting this into Eq. (5.4.15) we obtain the following expression for the real part of the dielectric constant:

 

 

 

 

 

 

 

 

e2

 

 

 

 

 

 

2

 

 

 

 

 

 

1(q, ω) = 1

 

 

 

d f 0

(Ek)

Ek

k .

 

 

 

 

π 2h2

ω2

k2

 

 

 

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

 

 

 

 

If we also expand the Fermi function in the numerator of Eq. (5.4.15),

 

 

f 0

 

 

 

 

 

 

 

 

 

f 0

h2

 

 

f 0

f 0

 

k

q

k

E

(k)

+ O

(q2)

=

f 0

+

 

k

q

¯

k

+ O

(q2) ,

 

 

 

 

k+q =

k

+ ∂E

·

 

 

 

k

∂E

· m

 

we find that for ω 0 the term linear in q gives the Thomas–Fermi result (5.4.19). In this limit the real part of the dielectric constant has the simple form

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

π e2

 

 

 

 

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1(q, 0) = 1

4

 

 

 

 

 

fk

D(E) dE .

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

q2

 

 

∂E

 

 

 

 

 

 

 

 

 

 

 

 

 

The integral yields at T = 0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

lim

(

 

,

 

)

 

 

 

 

 

4π e2

 

 

(

 

 

)

 

 

 

4π e2 3N0m2

 

 

4π e2 3N0

 

 

λ2

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

= 1+ q2

 

EF

= 1+ q2

 

 

 

h2kF2

 

= 1+ q2 2

 

 

 

= 1+ q2 ;

q0 1

 

q

0

 

 

D

 

 

 

 

 

 

E

F

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(5.4.24)

the result we arrived at in Eq. (5.4.10).

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

We can also use the relationship between σˆ (q, ω) and χˆ (q, ω) to obtain

 

 

 

 

 

ˆ

 

 

=

 

 

 

 

 

 

2

 

2

 

+ 4q

2kF

 

 

 

 

 

 

 

 

2

 

 

 

q

ω+i

 

 

1

 

 

 

4π i q2vF2

 

 

 

qvF

 

 

 

 

q

ω+i

 

 

σ (q, ω)

 

 

 

3ω

ωp

 

1

 

 

 

kF

1

 

q

 

 

 

 

 

ω + i

 

 

Ln

 

 

2kF

qvF

+ 1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

qvF

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ω+i

 

 

2kF

 

 

 

 

 

 

 

 

+

 

 

 

 

 

 

+

 

 

 

 

 

 

2

 

 

q

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

4q

2kF

 

qvF

 

 

 

 

q

 

+

ω+i

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2kF

 

 

 

 

 

 

 

 

 

 

 

 

 

kF

1

 

q

 

 

 

 

 

 

 

 

 

 

 

Ln

qvF

1

 

 

 

. (5.4.25)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

ω + i

 

 

 

 

2kF

+

qvF

+

 

 

 

The real and imaginary parts of the complex, wavevector dependent conductivity are displayed in Fig. 5.16 as a function of frequency ω and wavevector q. As discussed in Section 3.1.2, for small q values this expression is the same as σˆ (q, ω) derived for the transverse conductivity, and when q 0 the Drude model is recovered. For large q values – or short wavelength fluctuations – the results are significantly different: there is an appreciable conductivity in the case of the low frequency transverse response. However, in the longitudinal case σ1(q, ω) rapidly drops to zero with increasing q; this is the consequence of the screening for longitudinal excitations.

5.4.4 Single-particle and collective excitations

In the previous section we have addressed the issue of screening, the rearrangement of the electronic charge in response to transverse and longitudinal electric fields.

5.4 Longitudinal response

131

The formalism leads to the dielectric constant or conductivity in terms of the excitations of electron–hole pairs. Besides those excitations, a collective mode also occurs, and it involves coherent motion of the system as a whole.

Collective excitations of the electron gas are longitudinal plasma oscillations. As we saw in the derivation leading to Eq. (4.3.2) for vanishing 1(ω), an infinitesimally small external perturbation ext produces a strong internal field ; and in the absence of damping the electron gas oscillates collectively. For the uniform q = 0 mode simple considerations lead to the frequency of these oscillations. Let us consider a situation in which a region of the electron cloud is displaced by a distance x without affecting the rest of the system. The result is a layer of net positive charge on one side of this region and an identical negative layer on the opposite side, both of thickness x. The polarization is simply given by P = −N er, and this leads to an electric field E = −4π P. In the absence of damping (10) the equation of motion (3.2.22) reduces to

d2r

= −eE(t) = 4π N e2r .

(5.4.26)

m dt2

The solution is an oscillation of the entire charged electron gas, with an oscillation frequency

ωp =

4

π N e2

 

1/2

 

 

m

 

called the plasma frequency. Of course such collective oscillations of the electron plasma could be sustained only in the absence of a dissipative mechanism; in reality the oscillations are damped, albeit weakly in typical materials. Note that ωp here is the frequency of longitudinal oscillations of the electron gas. There is – as we have discussed in Section 5.1 – also a significant change in the transverse response at this frequency. The dielectric constant 1(q = 0, ω) switches from negative to positive sign at ωp, and this leads to a sudden drop of the reflectivity.

In order to explore the finite frequency and long wavelength (small q) limit, we expand 1 in terms of q; we obtain

 

 

ωp2

1 +

3 q2v2

+ · · ·

 

1(q, ω) = 1

 

 

 

F

(5.4.27)

ω2

5

ωp2

at frequencies ω ωp. Setting 1(q, ω) = 0 yields for the dispersion of the plasma frequency

 

 

 

 

 

3 q2v2

ωp2(q) ωp2

1 +

 

 

F

 

5

ω2

 

 

 

 

p

132 5 Metals

or

 

 

 

+ 10 ωp2

 

+

 

8m2

ω2

280 ωp4

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

3 v2

 

 

 

h

 

 

3 v4

 

 

 

ωp(q)

 

ωp

1

 

 

 

 

 

F

q2

 

 

¯

 

 

 

 

F

q4

,

(5.4.28)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

where the term

3 vF2

 

2

2

2

, with λ the Thomas–Fermi screening parameter

 

 

 

q

 

 

q

10

ωp2

 

 

as defined earlier in Eq. (5.4.6). The collective oscillations can occur at various wavelengths, and we can refer to these quanta of oscillations as plasmons. In Eq. (5.4.28) we have added the last term in O(q4) without having it explicitly carried through the calculation [Gei68, Mah90]. Plasma oscillations are well defined for small q since there is no damping due to the lack of single-particle excitations in the region of plasma oscillations. For larger q, the plasmon dispersion curve shown in Fig. 5.11 merges into the continuum of single-electron excitations at a wavevector qc and for q > qc the oscillations will be damped and decay into the single-particle continuum. Using the dispersion relation ωp(q) and the spectrum of electron–hole excitations, the onset of the damping of the plasmon called the Landau damping can be derived for particular values of the parameters involved. For small dispersion, we obtain by equating Eq. (5.3.2) to ωp(q = 0) the simple approximate relation

ωp

(5.4.29)

qc vF

for the onset of damping of plasma oscillations by the creation of electron–hole pairs.

5.5 Summary of the ω dependent and q dependent response

Let us finally recall the results we have obtained for the complex q and ω dependent conductivity (or, equivalently, dielectric constant). Electron–hole excitations – and the absorption of electromagnetic radiation – are possible only for certain q and ω values, indicated by the shaded area on Fig. 5.11. This area is defined by the

boundaries hω

h2

 

2qkF q

2

and hω

 

h2

 

2qkF q

2

.

¯

 

 

 

¯

 

 

¯

= 2m

1

 

= 0 and no

+

 

 

 

 

 

¯

=

2m

 

 

 

Outside this region σ (q, ω)

 

 

 

absorption occurs. In the so-called

homogeneous limit, for qvF ω, we find that the transverse and longitudinal responses are equivalent (we cannot distinguish between transverse and longitudinal fields in the q 0 limit), and to leading order we find

 

iN e2

1 +

q2v2

 

σˆ L(q, ω) = σˆ T(q, ω) =

 

 

F

+ · · ·

(5.5.1)

mω

5ω2

with, of course, the associated δ{ω = 0} response for the real part. The dielectric

References

133

constant has the form

 

ωp2

2

 

1L(q, ω) = 1T(q, ω) = 1

+ · · · ;

(5.5.2)

ω

the zeros of which lead to the ultraviolet transparency of metals and the longitudinal plasma oscillations with the dispersion relation ωp(q) given above.

In the opposite, quasi-static, limit for qvF ω screening becomes important, and indeed in this limit the transverse and longitudinal responses are fundamentally different. Here we find

ˆ

=

3π N e2

 

 

ω2

+ · · ·

 

4qvFm

q2vF2

σ T(q, ω)

 

 

 

1

 

 

(5.5.3)

and the response is primarily real; we recover dissipation in the small q (but still qvF ω) limit. In contrast,

ˆ

=

3N e2ω

 

 

kFω2

+ · · ·

q2vF2

 

2q3vF2

σ L(q, ω)

 

 

i

1

 

 

(5.5.4)

is primarily imaginary; there is no dissipation, and we also recover the Thomas– Fermi screening in this limit.

All this is valid for a three-dimensional degenerate electron gas. Both the single-particle and collective excitations are somewhat different in twoand onedimensional electron gases for various reasons. First, phase space arguments, which determine the spectrum of single-particle excitations, are somewhat different for a Fermi sphere in three dimensions, a Fermi circle in two dimensions and two points (at +kF and kF) in one dimension. Second, polarization effects are different in reduced dimensions, with the consequence that the frequency of the plasmon excitations approach zero as q = 0. These features are discussed in Appendix F.

A final word on temperature dependent effects is in order. Until now, we have considered T = 0, and this leads to a sharp, well defined sudden change in the occupation numbers at q = kF. At finite temperatures, the thermal screening of the Fermi distribution function leads to similar smearing for the single-particle excitations. With increasing temperature, we progressively cross over to a classical gas of particles, and in this limit classical statistics prevails; the appropriate expressions can be obtained by replacing mvF2 by 3kBT .

References

[Abr72] A.A. Abrikosov, Introduction to the Theory of Normal Metals (Academic Press, New York and London, 1972)

134

5 Metals

[Bla57]

F.J. Blatt, Theory of Mobility of Electrons in Solids, in: Solid State Physics, 4,

 

edited by F. Seitz and D. Turnbull (Academic Press, New York, 1957), p.

 

199

[Cal91]

J. Callaway, Quantum Theory of Solids, 2nd edition (Academic Press, New

 

York, 1991)

[Cha90]

R.G. Chambers, Electrons in Metals and Semiconductors (Chapman and Hall,

 

London, 1990)

[Gei68]

J. Geiger, Elektronen und Festkorper¨ (Vieweg Verlag, Braunschweig, 1968)

[Gru77]

G. Gruner¨ and M. Minier, Adv. Phys. 26, 231 (1977)

[Hau94]

H. Haug and S.W. Koch, Quantum Theory of the Optical and Electronic

 

Properties of Semiconductors, 3rd edition (World Scientific, Singapore,

 

1994)

[Kit63]

C. Kittel, Quantum Theory of Solids (John Wiley & Sons, New York, 1963)

[Koh59]

W. Kohn, Phys. Rev. Lett. 2, 395 (1959)

[Lin54]

J. Lindhard, Dan. Mat. Fys. Medd. 28, no. 8 (1954)

[Mad78]

O. Madelung, Introduction to Solid State Theory (Springer-Verlag, Berlin,

 

1978)

[Mah90]

G.D. Mahan, Many-Particle Physics, 2nd edition (Plenum Press, New York,

 

1990)

[Mat58]

D.C. Mattis and G. Dresselhaus, Phys. Rev. 111, 403 (1958)

[Pin66]

D. Pines and P. Nozieres,` The Theory of Quantum Liquids, Vol. 1

 

(Addison-Wesley, Reading, MA, 1966)

[Pip47]

A.B. Pippard, Proc. Roy. Soc. A 191, 385 (1947)

[Pip54a]

A.B. Pippard, Proc. Roy. Soc. A 224, 273 (1954)

[Pip54b]

A.B. Pippard, Metallic Conduction at High Frequencies and Low

 

Temperatures, in: Advances in Electronics and Electron Physics 6, edited by

 

L. Marton (Academic Press, New York, 1954), p. 1

[Pip62]

A.B. Pippard, The Dynamics of Conduction Electrons, in: Low Temperature

 

Physics, edited by C. DeWitt, B. Dreyfus, and P.G. deGennes (Gordon and

 

Breach, New York, 1962)

[Reu48]

G.E.H. Reuter and E.H. Sondheimer, Proc. Roy. Soc. A 195, 336 (1948)

[Zim72]

J.M. Ziman, Principles of the Theory of Solids, 2nd edition (Cambridge

 

University Press, London, 1972)

 

Further reading

[Car68]

M. Cardona, Electronic Optical Properties of Solids, in: Solid State Physics,

 

Nuclear Physics, and Particle Physics, edited by I. Saavedra (Benjamin,

 

New York, 1968), p. 737

[Dru00a]

P. Drude, Phys. Z. 1, 161 (1900)

[Dru00b]

P. Drude, Ann. Physik 1, 566 (1900); 3, 369 (1900)

[Gol89]

A.I. Golovashkin, ed., Metal Optics and Superconductivity (Nova Science

 

Publishers, New York, 1989)

[Gro79]

P. Grosse, Freie Elektronen in Festkorpern¨ (Springer-Verlag, Berlin, 1979)

[Huf96]

S. Hufner,¨ Photoelectron Spectroscopy, 2nd edition, Springer Series in

 

Solid-State Sciences 82 (Springer-Verlag, Berlin, 1996)

 

Further reading

135

[Hum71]

R.E. Hummel, Optische Eigenschaften von Metallen und Legierungen

 

 

(Springer-Verlag, Berlin, 1971)

 

[Jon73]

W. Jones and N.H. March, Theoretical Solid State Physics (John Wiley &

 

 

Sons, New York, 1973)

 

[Lif73]

I.M. Lifshitz, M.Ya. Azbel, and M.I. Kaganov, Electron Theory of Metals

 

 

(Consultants Bureau, New York, 1973)

 

[Pin63]

D. Pines, Elementary Excitations in Solids (Addison-Wesley, Reading, MA,

 

1963)

 

[Pla73]

P.M. Platzman and P.A. Wolff, Waves and Interactions in Solid State Plasmas

 

(Academic Press, New York, 1973)

 

[Sin77]

K.P. Sinha, Electromagnetic Properties of Metals and Superconductors, in:

 

 

Interaction of Radiation with Condensed Matter, Vol. 2 IAEA-SMR-20/20

 

(International Atomic Energy Agency, Vienna, 1977), p. 3

 

[Sok67]

A.V. Sokolov, Optical Properties of Metals (American Elsevier, New York,

 

1967)

 

[Som33]

A. Sommerfeld and H. Bethe, Handbuch der Physik, 24/2, edited by H. Geiger

 

and K. Scheel (Springer-Verlag, Berlin, 1933)

 

6

Semiconductors

The focus of this chapter is on the optical properties of band semiconductors and insulators. The central feature of these materials is the appearance of a single-particle gap, separating the valence band from the conduction band. The former is full and the latter is empty at zero temperature. The Fermi energy lies between these bands, leading to zero dc conduction at T = 0, and to a finite static dielectric constant. In contrast to metals, interband transitions from the valence band to the conduction band are of superior importance, and these excitations are responsible for the main features of the electrodynamic properties. Many of the phenomena discussed in this chapter also become relevant for higher energy excitations in metals when the transition between bands becomes appreciable for the optical absorption.

Following the outline of Chapter 5, we first introduce the Lorentz model, a phenomenological description which, while obviously not the appropriate description of the state of affairs, reproduces many of the optical characteristics of semiconductors. The transverse conductivity of a semiconductor is then described, utilizing the formalisms which we have developed in Chapter 4, and the absorption near the bandgap is discussed in detail, followed by a summary of band structure effects. After discussing longitudinal excitations and the q dependent optical response, we briefly mention indirect transitions and finite temperature and impurity effects; some of the discussion of these phenomena, however, is relegated to Chapter 13.

Optical properties of semiconductors are among the best studied phenomena in solid state physics, mainly because of their technological relevance. Both experimental data and theoretical description are quite advanced and are the subject of many excellent textbooks and monographs, e.g. [Coh88, Hau94, Yu96].

136

6.1 The Lorentz model

137

6.1The Lorentz model

6.1.1Electronic transitions

Before discussing optical transitions induced by the electromagnetic fields between different bands in a solid, let us first examine a simple situation: the transitions between a ground state and excited states of N identical atoms, the wavefunctions of which are 0(r) and l (r). The electrodynamic field is assumed to be of the form

E(t) = E0 (exp{−iωt} + exp{iωt})

(6.1.1)

and leads to the admixture of the excited states to the ground state. This can be treated by using the time dependent Schrodinger¨ equation with the perturbation given for an electric field by

H = eE · r .

Because of this admixture, the resulting time dependent wavefunction is

(r, t) = 0 exp iEh0t +

l

al (t) l exp iEhl t

(6.1.2)

 

¯

 

 

 

¯

 

 

with El denoting the various energy levels. The coefficients al (t) are obtained from the solution of the Schrodinger¨ equation

H0

+ H (r, t) ,

 

ih¯ t (r, t) = −

(6.1.3)

where H0 l = El l describes the energy levels in the absence of the applied electromagnetic field. Inserting the wavefunction into Eq. (6.1.3) we find that the coefficients are

al (t)

 

E

1 exp

i(h¯ ω+Eh¯l E0)t

 

 

1 exp

i(h¯ ω+h¯El E0)t

 

l ( er) 0 dr.

 

 

 

 

= 2h

 

h

ω

(

l

 

0

)

 

h

ω

 

(

l

 

0

)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

¯

¯

+

 

 

E

E

 

 

 

 

¯

 

 

E

E

 

 

 

 

(6.1.4)

Here

l er 0 dr = erl0

is the matrix element of the dipole moment P = −er. The induced dipole moment

P(t) = −er(t) =

 

E

 

 

2ω

 

(r, t)(e r) (r, t) dr =

 

 

e2 |rl0|2

 

l

= αE

2h

l

ωl2

ω2

 

 

¯

 

 

 

 

(6.1.5)