Download as pdf or txt
Download as pdf or txt
You are on page 1of 145

OPTICAL ABSORPTION

AND DISPERSION
IN SOLIDS
OPTICAL ABSORPTION
AND DISPERSION
IN SOLIDS

J. N. HODGSON
Senior Lecturer in Physics,
University of Keele

CHAPMAN AN D HALL LTD


11 NEW FETTER LANE LONDON EC4
First published in 1970
© 1970 J. N. Hodgson

by Butler & Tanner Ltd


Frome and London
ISBN-13: 978-1-4613-3323-4 e-ISBN-13: 978-1-4613-3321-0
001: 10.1007/978-1-4613-3321-0
Softcover reprint of the hardcover 1st edition 1970

Distributed in the U.S.A.


by Barnes & Noble, Inc.
Preface
The electromagnetic theory of Maxwell and the electron theory of
Lorentz and Drude stimulated a great deal of experimental work on
the optical properties of solids in the late nineteenth and early
twentieth centuries. The time was not then ripe, however, for general
progress in this field. The experimental techniques were not available
to produce suitable specimens for optical measurements with well-
defined structure and purity. On the theoretical side, the classical
electron theory provided only a very incomplete account of the
interaction of light waves with matter. The centre of interest in
optical research moved to atomic and molecular spectroscopy where
quantitative results were easier to obtain. The quantum theory,
starting with Bohr's theory of 1913, provided a highly successful
basis for the interpretation of the optical spectra of atoms and
molecules.
The present-day theory of the optical properties of solids is based
on the quantum theory of electrons in solids, developed from the
early researches of Sommerfeld and Bloch, and the theory of lattice
vibrations originating in the research by Born. The formal con-
nection between optical absorption and electron wave functions in
solids has been well known since the 1930s but it is only recently that
electron energy band calculations have achieved sufficient accuracy
to make profitable a comparison of experimental and theoretical
results. Without some guidance from a theoretical band structure
calculation, it would be difficult to make any progress with the in-
terpretation of an optical absorption spectrum. When such guidance
enables specific transitions to be identified, the optical data can be
used to provide more accurate values for energy levels. This fruitful
interaction of theory and experiment is now being exploited in many
researches on the optical properties of solids. Although much re-
mains to be done in the measurement and interpretation of linear
optical properties, the advent of the laser has opened up the new
field of non-linear optical properties to experimental investigation.
The possibilities in this field are extremely varied and will no doubt
provide material for many years of research.
v
vi PREFACE

The aim of this book is to provide a bridge between the usual


textbook on solid state physics with one or two chapters on optical
properties, and the specialized review article or original paper. It
will be assumed that the reader has some knowledge of electro-
magnetic theory and elementary quantum theory. The scope is
limited to a treatment of optical absorption and dispersion in pure
crystals. An attempt has been made to relate the theory at each stage
with relevant experimental results. A selection of recent experi-
mental data is used to illustrate the various points in the theoretical
development. This selection represents, of course, only a small
fraction of the results published in recent years. References to
experimental data on other substances can be found in the review
articles and books listed as general references.
I would like to thank Miss K. B. Davies for her expert assistance
in typing the manuscript.
Contents
PREFACE page v

1. MACROSCOPIC THEORY 1
1.1 Electromagnetic field in a solid 1
1.2 Dielectric constant and optical conductivity 4
1.3 Crystal symmetry 7
1.4 Propagation of waves 8
1.5 Kramers-Kronig relations 13
1.6 The sum rule 19
1.7 Dispersion theory of classical oscillators 22

2. CRYSTAL LATTICE ABSORPTION 25


2.1 Vibrational modes of a crystal lattice 25
2.2 Photon-phonon interaction 28
2.3 Microscopic theory of infra-red dispersion 33
2.4 Two-phonon absorption 37

3. INTERBAND TRANSITIONS 41
3.1 Electron energy bands 41
3.2 Direct transitions 44
3.3 Critical points 48
3.4 Absorption band edges 53
3.5 Indirect transitions 56
3.6 Infra-red absorption in superconductors 59

4. FREE CARRIER ABSORPTION 62


4.1 Classical theory 62
4.2 Intraband transitions 65
4.3 Electron transport 69
4.4 Surface admittance 73
4.5 Infra-red absorption in metals 76
4.6 Free carrier absorption in semiconductors 80
viii CONTENTS

5. PLASMA EFFECTS 84
5.1 Free electron model 84
5.2 Volume plasmons 88
5.3 Surface plasmons 92
6. EXCITON EFFECTS 97
6.1 Electron-hole interaction 97
6.2 Optical absorption 100
6.3 Inert-atom solids and alkali halides 101
6.4 Semiconductors 104
6.5 Spatial dispersion 107

7. NON-LINEAR OPTICS III


7.1 Classification of non-linear effects 113
7.2 Non-linear susceptibilities 115
7.3 Second harmonic generation 119
7.4 Parametric amplification and oscillation 126
7.5 Third order effects 130

REFERENCES 135

INDEX 137
1
Afacroscopic Jrheory
1.1 Electromagnetic field in a solid
The classical theory of electric and magnetic fields in vacuum pre-
dicts the existence of electromagnetic waves travelling with a
characteristic velocity c. These waves are emitted by electric charges
which are moving with non-zero acceleration. A charge oscillating
in simple harmonic motion with an angular frequency w emits a wave
of the same frequency. The classical theory of an electron must be
replaced by the quantum theory when the electrons in a solid are
being considered. The electrons can make transitions between energy
levels with the emission or absorption of photons, the quanta of the
electromagnetic field. There remains a close analogy between the
quantum picture of an electron making a transition between energy
levels E1 and E 2 , with a photon of frequency w given by
Iiw = E1 - E 2 ,
and the classical picture of an electron oscillating with frequency w.
For the discussion of optical phenomena, the classical theory is an
adequate approximation for the electromagnetic field, while the
motions of electrons in a solid must be treated quantum-mechanically.
A solid is an assembly of electrons and nuclei each with a certain
mass and electric charge. In the presence of an electromagnetic wave
of frequency w, a particle of charge e is subject to an electric force
eE where' E is the electric field, which oscillates with frequency w.
The forces due to the oscillating magnetic field have a negligible effect
on the linear optical properties. Under the influence of the electric
force, the particles act as sources of secondary spherical waves which
combine with the original wave and influence other particles. A
microscopic theory of the propagation of electromagnetic waves in a
solid must lead to a self-consistent combination of the incident wave
and the scattered wave from each particle. As a simple example, let
us consider a plane light wave in vacuum incident on the plane sur-
face of a transparent crystal. We know from experiment that the
1
2 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

result is a reflected plane wave with velocity c and a transmitted


plane wave with phase velocity c/n, where n is the refractive index.
From the microscopic point of view, this means that the secondary
waves must cancel the incident wave inside the crystal and replace it
by a wave of phase velocity c/n. Outside the crystal the secondary
waves must combine to produce the reflected wave. A quantitative
treatment of these ideas, leading to the 'extinction theorem', was
first given by Ewald and Oseen.
The microscopic approach using only the vacuum field equations,
is unnecessarily complicated for most problems in the theory of
optical properties. It is more convenient to use the form of Maxwell's
equations appropriate for macroscopic fields in material media.
These equations can be justified as averages of the vacuum field
equations, as first shown by Lorentz. The phenomena discussed in
this book relate to the 'optical' region of the electromagnetic spec-
trum. This region will be taken to include vacuum wavelengths {iI.o}
from 50 nm to 500 pm or photon energies (fun) from 25 to 0·002 eV.
The lower limit on Ao is chosen so that Ao ~ a where a is a typical
atomic radius. The upper limit on Ao corresponds roughly to the
transition between optical and microwave experimental methods. It
also excludes low-frequency phenomena such as the various magnetic
resonance effects. These effects are related to optical properties in so
far as they are both due to the interaction of electric and magnetic
fields with matter.
The microscopic fields in a solid fluctuate widely between different
points depending on tht; proximity of electrons and nuclei. These
fluctuations can be smoothed out by taking averages of the fields;
equations relating the average fields can then be derived. In a crystal-
line solid the fields can be averaged over each unit cell of the lattice.
Suppose that the position vector of the centre of a typical cell is rio
The average field for the discrete values of ri can be smoothed into a
continuous function ofthe position vector r. This procedure assumes
that the variation of average field between adjacent cells is small.
An oscillating electric field of frequency m produces an oscillating
dipole moment in each lattice cell given by:
fpSrdXdydz
where p is the charge density, Sr is the displacement of a charge
element due to the applied electric field and the integration is taken
MACROSCOPIC THEORY 3
over one lattice cell. The oscillating displacement Sr can be expressed
as the real part of:
Sr = ro exp (-imt). (1.1)
The electric polarization P, the dipole moment per unit volume, is
given by:

P= (l/1,')fpSrdxdydz (1.2)

where v is the volume of a unit cell. The oscillations of charged


particles will also produce an electric current density:

J = (llv) f p Su dx dy dz (1.3)

where Su is the velocity of a charge element due to the applied field.


Since Sr and Su are relat~d by:
Su = a(Sr)lat
therefore:
J = aplat = -imP. (1.4)
We will use P to represent the effect of a medium on the electro-
magnetic field, including the case of a conducting medium. Assuming
that the medium contains no additional charges, Gauss's theorem
for the average electric field E becomes:
V.E = -4nV.P. (1.5)
Gaussian units will be used because they simplify the relations
between the electromagnetic equations and optical phenomena.
Introducing the electric displacement D = E + 4nP, equation (1.5)
becomes:
V.D=O (1.6)
The average magnetic field in a medium will be denoted by Band
the magnetization by M. The magnetic field vector H is defined by
H = B - 4nM. However, for oscillating fields of optical frequencies,
it is an unnecessary refinement to distinguish between Hand B. Any
paramagnetic or ferromagnetic moments will be unable to follow the
rapid oscillations of magnetic field because of their long relaxation
times. The remaining diamagnetic moments are so small as to have
an inappreciable effect on optical behaviour. In magneto-optical
phenomena, the correct use of B or H for the constant magnetic
field is important.
4 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The problem of the behaviour of a light wave at the surface of a


crystal can be dealt with very simply by means of the boundary
equations for the field vectors. These equations can be derived by
integrating the electromagnetic equations across a transition region
between two media and then letting the width of the region tend to
zero. The field vectors just on either side of the boundary are related
by:
EIt = E 2t ; Hit = H 2t ; DIn = D 2n ; BIn = B 2n (1.7)
where t and n refer to tangential and normal components. Given two
of these equations, the other two follow from the field equations for
the two media. The fields of incident and reflected waves at a crystal
surface are related by the boundary equations (1.7) to the field of the
refracted wave at the surface.

1.2 Dielectric constant and optical conductivity


Let us consider the effect of a monochromatic plane wave of angular
frequency wand wave vector k travelling through a crystalline
medium. The electric field E of the wave will cause forced oscillations
of the atomic particles. The electric fields in the outer parts of an
atom are of the order of 10 8 V cm- 1 so the perturbing effect of an
electromagnetic wave with E,...., 1 V cm- 1 is very small. As a con-
sequence, the relation between E and the electric polarization P is
accurately linear. Even with a laser beam where E,...., 10 5 V cm-1 is
possible, the non-linearities are usually small. The linear relation
between D (= E + 4nP) and E is:
3
Di = ~ eijE j • (1.8)
j=1
The labels ij on eii are written as superscripts because subscripts 1
and 2 will be used to distinguish real and imaginary parts. eii is a
symmetric tensor with eii = e ji so, in general, it has six independent
components. When referred to its principal axes, the non-diagonal
elements of eii are zero:
Dl = e11E 1 ; D2 = e22 E 2 ; Da = e 33E 3• (1.9)
The dielectric tensor eii represents the response of a crystal to a
perturbing field offrequency wand wave vector k; it should therefore
be considered as a function of wand k: eii(k,w). However, the pre-
vious assumption that fields could be averaged over a unit cell
MACROSCOPIC THEORY 5
amounts to neglecting any dependence of e ti on wavelength. The
same results are obtained for all wavelengths large enough so that
the variation of field over a unit cell is negligible. Expressed in
symbols, the assumption is:
eii(k,ro) = ei1(0,ro)= eH(ro). (1.10)
The neglect of possible 'spatial dispersion' of eii, that is dependence
on k, is not valid in some cases. Spatial dispersion arises whenever
the relation between D and E is not exactly a local relation with D
at a particular point determined solely by E at that point. The small
non-local effects in a transparent crystal can be represented by terms
proportional to VEi in the expression for D. These are very small
terms but they cause the new phenomenon of natural optical rota-
tion in some crystals, for example quartz. An electromagnetic wave
in a metal may be so strongly damped that the mean free path of
conduction electrons is greater than the penetration depth. This is
the condition for the occurrence of the anomalous skin effect and
the relation between D and E is then extremely non-local. Spatial
dispersion is also important in the interaction of light waves with
excitons.
The phase difference between D and E can be included in the theory
by allowing e to have complex values:
e = e1 + ie2. (1.11)
We will assume here that e refers to one of the principal values of
eii . Let the real electric field be represented by:
E(ro) = Eo exp (-irot) + Eo* exp (irot); (1.12)
the asterisk denotes the complex conjugate. The electric displace-
ment:
D(ro) = e(ro)Eo exp (-irot) + e( -ro)Eo* exp (irot).
Since D(ro) must be real for real E(ro), it follows that:
e( -ro) = e*(ro). (1.13)

< ataD).
The average rate of dissipation of electromagnetic energy density is:
1
W= <E. I) = 4n E·

Substituting for E and D, and averaging over a complete oscillation


we have:
W = (iro/4n){e( -ro) - e(ro)}Eo.Eo*
= {roe2(ro)/4n}(E2). (1.14)
6 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

(E2) is the mean square electric field (= 21 Eo 12).


Relations (1.13)
and (1.14) are shown in graphical form in fig. 1.1.
The optical properties can be expressed alternatively by a complex
optical conductivity:
(1.15)
a(w) expresses the relation between J and E, as sew) expresses the

Figure 1.1 Relations between electric field E, electric pOlarization P, electric


current density J, electric displacement D and power dissipation w.
relation between D and E; they both have a similar tensor character.
Using the relation (1.4) between J and P, we find:
<TI(W} - ia 2(w) = (w/4n}s2(w) - i(w/4n){sI(w) - I},
where <TI(W) and a2(w) are expressed in electrostatic units. Equating
real and imaginary parts:
<TI(W) = (w/4n)s2(w); <T2(W) = (w/4n){sI(w) - I}. (1.16)
<TI(W) and S2(W) each account for the current in phase with the
electric field and hence the absorption of an electromagnetic wave.
It is often convenient to use aI(w), the optical conductivity, to
represent the absorption because the power absorbed for a given
amplitude offield is simply a I (w)(£2) with no additional frequency-
dependent factor. As the frequency tends to zero, the optical con-
ductivity becomes the same as the electrical conductivity aI(O).
MACROSCOPIC THEORY 7

1.3 Crystal symmetry


The main subject of this book is the interpretation of optical absorp-
tion and dispersion in solids in terms of their electronic structures.
For this purpose it is sufficient to consider only linearly polarized
waves with their electric vectors parallel to one of the principal axes
of eii . The optical phenomena which arise with waves travelling in an
arbitrary direction in a crystal, form the subject of crystal optics
which has been treated in many well-known texts.
From the point of view of optical behaviour, crystals can be
classified into three groups. It will be assumed that the Cartesian
(x,y,z) axes coincide with the principal axes of eii . In the first group
of crystals the principal values of the dielectric tensor are all equal,
tP = e22 = e33 , and the crystal is optically equivalent to an iso-
tropic medium. All crystals of the cubic system belong to this group.
In the second group, ell = e 22 ~ e 33 and the crystal is said to be
uniaxial. The z-axis with unequal e33 is called the optic axis or
c-axis. A plane wave travelling parallel to the z-axis has its electric
vector E in a plane perpendicular to the z-axis. Since ell = e 22 , it
can be shown that the phase velocity and absorption of this wave are
independent of the direction of E in the plane. Crystals of the
hexagonal, tetragonal and trigonal systems are uniaxial and the
optic axis coincides with a crystal axis of six-, four- or three-fold
symmetry. In the third group of crystals, ell ~ e 22 ~ e 33 and the
crystal is said to be biaxial. There are two optic axes along which
propagation is independent of polarization, that is the direction of
E, but these axes are inclined to the principal axes. For a wave
travelling along the x-axis, D and E will be parallel only if E is
parallel to the y-axis or the z-axis. This is the condition for a wave
to be propagated with no change in the direction of E. Crystals of
the orthorhombic, monoclinic and triclinic systems are optically
biaxial. In crystals of the orthorhombic system the axes of eii co-
incide with crystal symmetry axes. In the monoclinic and triclinic
systems some of the axes of eii are not determined by crystal sym-
metry and their directions vary with frequency. In practice most
elements and simple compounds crystallize in lattices of high
symmetry and most of the published research on optical properties
is confined to these solids.
Absorbing crystals are included in the classification by allowing
B
8 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

the components of the dielectric tensor to have complex values,


eii= eli; + ie2ii. Two tensors of real dielectric constant eli! and
optical conductivity alii (= roe2iij4n) can be separated. In the
crystal systems where the principal axes of eii follow crystal sym-
metry axes, the principal axes of eli; and ali; coincide.

1.4 Propagation of waves


Maxwell's equations for the average fields in a solid can be written:
V.D=O· V.B=O
, (1.17)
V x E = -(ljc)(oBjot); V x H = (ljc)(oDjot).
Let us consider the solution corresponding to a plane wave in a
crystalline medium with E parallel to a principal axis of eii ; the
corresponding principal value of eii is denoted bye.
D = eE; E = Eo exp (ik. r - irot). (1.18)
The wave vector k will have complex components if a wave of fre-
quency ro is absorbed by the crystal. If we take k = kl + ik2' then:
exp(ik.r) = exp (ikl.r)exp (-k 2 .r). The planes of constant phase
are perpendicular to kl while the planes of constant amplitude are
perpendicular to k 2 • The operator V reduces to the factor ik for a
plane wave. Therefore, the field equations (1.17) become:
k.D=ek.E=O; k.B=k.H=O (19)
k x E = (rojc)B = (rojc)H; k x H = -(rojc)eE. 1.
Using the formula k x (k x E) = k(k.E) - E(k.k), the field E
can be eliminated to give a relation between k and ro:
c 2(k. k) = ro 2 e. (1.20)
If e has a real value for frequency ro, then one solution of (1.20) is a
real vector k of magnitude k = (rojc)e t . This corresponds to an un-
damped wave of phase velocity (rojk) = cje!. It follows from the
definition of refractive index n as the ratio, velocity in vacuum to
phase velocity in medium, that n = et. Equations (1.19) also show
that E, Hand k are mutually perpendicular and the magnitudes of
E and H are related by:
(1.21)
The ratio (EjB) or (4njc)(EjB) is known as the wave impedance;
for our purposes it will be more convenient to use its reciprocal, the
MACROSCOPIC THEORY 9
wave admittance Yo where:
Yo = (HIE) = 8 t . (1.22)
If the general complex form is taken for k, equation (1.20) can be
written:
(1.23)
There is a solution of this equation with k2 ~ 0 even if 8 is real.
This solution requires that k 1. k2 = 0; the planes of constant phase
and constant amplitude are perpendicular. This is called an evan-
escent wave and it can be generated near the boundary of a trans-
parent medium. The evanescent wave is damped but there is no
dissipation of energy since 8 2 = O.
When 8 has a complex value, equation (1.20) can be satisfied only
by complex k; all solutions correspond to damped waves. The
simplest case is when k1 and k2 are parallel. This type of wave is
produced when a plane wave is incident normally on the plane sur-
face of an absorbing medium. Equation (1.23) becomes:
c2(k12 - k a2 + 2ik1k 2) = W2(8 1 + i8 2)·
Separating real and imaginary parts:
c2(k12 :- k22) = W 28 1; c2(2k1k2) = W 28 2. (1.24)
The refractive index n and the extinction coefficient K of an absorb-
ing medium, often called the optical constants, are defined by:
n = (clw)k 1 ; K = (clw)k 2• (1.25)
Combining (1.24) and (1.25) we have:
8 1 = n 2 - K2; 8 2 = 2nK; 8 = (n + iK)2. (1.26)
Using (1.16) the relation between the optical conductivity and the
optical constants can be written:
(f1/c = nKjAo (1.27)
where 1.0 (= 21tcI w) is the vacuum wavelength. The field amplitudes
in the wave are damped according to the factor
exp (-k2Z) = exp (-21tKzIAo),
where the z-axis is normal to the surface. The penetration or skin
depth ~o is defined by:
~o = (llk 2) = Ao/21tK. (1.28)
When an external wave is incident obliquely on the surface of a
medium, the planes of constant amplitude remain parallel to the
10 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

z/Xo

Figure 1.2 Electric and magnetic fields in an absorbing crystal due to a normally
incident wave. Optical constants of crystal: n = 5·0, K = 3·5. (Silicon at
Ao = 0·31 ,urn.) Skin depth 15. Lattice constant a. Crystal surface in plane z = o.
Internal fields Ex and H y calculated for I = 0 when electric field of incident wave
is Ex = Eo cos wI at z = o.

surface. However, the magnitude of k2 depends on the angle of


incidence (j. If 1 e 1 ?> 1 then:
1/k2 = 15 ~ l5 o{1 - 1- sin 2 (j/(n 2 + K2)}. (1.29)
Since (n 2 + K2) = 1e I, we see that the variation of 15 with (j is small
when I e I ?> 1.
The reflected wave from the surface of a medium has a certain
amplitude ratio r and phase difference ~ with respect to the incident
wave. r and ~ depend on the angle of incidence (j and the optical
constants nand K of the medium. The Fresnel equations relating
these quantities can be derived from the boundary conditions for
the fields. Experimental measurements of r and ~ can be used to
deduce the values of nand K. For the case of normal incidence
((j = 0):
(1.30)
MACROSCOPIC THEORY 11
where eo is the dielectric constant of the transparent medium carry-
ing the incident and reflected waves. Replacing eT by (n + iK) and
eot by no, we have:
r exp (ill) = (no - n - iK)/(no + n + iK).
Hence: Ro = r2 = {(no - n)2 + K2}/{n o + n)2 + K2}
(1.31)
tan Ll = 2n oK/{n 2 + K2 - n02}.
Ro is the reflectance, defined as the fraction of the incident energy
which is reflected. In the generalization to oblique incidence «() ~ 0)
it is necessary to specify the polarization of the incident wave. For a
p-wave E is parallel to the plane of incidence, while for an s-wave E
is perpendicular to this plane. A measurement of two independent
reflection parameters at a given frequency is sufficient in principle to
determine nand K for this frequency. A choice of suitable parameters
must be made from considerations of accuracy and experimental
convenience.
The reflection of electromagnetic waves can also be treated by
using the concept of surface admittance defined by:
Y = H t/ E t • (1.32)
E t and H t denote the tangential components of E and H at the sur-
face; they are continuous on crossing the surface. The concept of
surface admittance remains valid under the conditions of the
anomalous skin effect when the spatial dispersion of e(k,w) cannot
be ignored. For normal incidence «() = 0) and if the approximation
e(k,w) ~ e(O,W) is valid, the surface admittance Y is equal to the
wave impedance (1.22). For oblique incidence Y is different for
p- and s-polarizations:
Y v = et{l - (n0 2 sin 2 ()/e}-t
Ys = et{l - (n02 sin2 ()/e}+t. (1.33)
From the boundary conditions for the continuity of E t and H t , the
following expressions for the reflection coefficients are derived:
rs exp (ills) = (no cos () - Ys)/(n o cos () + Ys);
(1.34)
rp exp (iLlp) = (no - Yp cos ()/(no + Yp cos ()
The difference between Yp and Y s becomes very small for metals
in the infra-red when I e I ~ I, and Yp ~ Y s ~ ef independent of ().
The wave in the metal is so strongly damped that the normal deriva-
tives of the field are much larger than the tangential derivatives.
The variation of the field within the metal becomes almost in-
12 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

dependent of the angle of incidence. The approximation Yp ~ Ys ~ Y


remains valid in the anomalous skin effect region. Theoretical
calculations of the anomalous skin effect usually assume () = 0 and
hence give Y for normal incidence. Equations (1.34) predict the
reflection coefficients for oblique incidence, assuming Yp = Ys = Y.
The optical constants of absorbing crystals can also be deduced
from measurements of the transmission and reflection by thin layers.
For measurement oftransmission the thickness of the layer must not
exceed a few times the skin depth 15 because the transmitted intensity
decreases as exp (-4:ns/l5) where s is the film thickness. The general
formulae for the transmission and reflection coefficients can be
derived by using the boundary conditions at each surface of the
layer. We will consider only one example, the transmission of a p-
wave through a very thin layer in vacuum. This example has the
advantage of introducing the concept of a loss function for longi-
tudinal fields. Let us assume that s ~ (A.o/2:n) and let ~ = (2:nS/A o).
If Eo is the amplitude of the incident electric field, then the electro-
magnetic power incident on the layer is:
(CE02/4:n) cos () erg cm- 2 sec-I. (1.35)
E 1- and Ell will be used to denote the components of E inside the
layer, perpendicular and parallel to the normal. The power absorbed
in the layer is:
(1.36)
If the layer is sufficiently thin, the variation of E 1- and Eil across the
layer can be neglected. The boundary equations for the tangential
and normal components of E are:
E 1- = Eo cos (); Ell = (1/8)E o sin (). (1.37)
Hence formula (1.36) can be written:
0'1E02s{cos 2 () + (1/1812) sin 2 ()} erg cm- 2 sec-I. (1.38)
The absorptance A is defined as the ratio of power absorbed to
incident power. From (1.35) and (1.38):
A = (4:n0'IS/C cos (){cos 2 () + (1/1 81 2) sin 2 ()}
= (Ucos (){8 2 cos 2 () + 82 sin2 ()/(8 I 2 + 82 2)}. (1.39)
A calculation of the reflectance shows that it is proportional to ~2;
it can therefore be neglected in this approximation and the trans-
mittance is equal to (1 - A). The absorptance due to the field
component Ell is seen to be proportional to 82/(8 1 2 + 82 2), or
MACROSCOPIC THEORY 13
+ B22) if all frequency-dependent factors are included. The
0'1/(B1 2
component Bit along the normal to the layer behaves similarly to the
field of a longitudinal electromagnetic wave, one in which E and k
are parallel. This type of wave will be discussed further in the
chapters on lattice absorption and plasma effects. Longitudinal
waves can be excited by high-energy electrons and their energy
losses are determined by the function O'd(B12 + B22).

1.5 Kramers-Kronig relations


Kramers-Kronig relations are relevant to physical phenomena
which can be described by the absorption and phase shift of waves.
An early practical application was to the design of electrical wave
filters for which the performance is specified by attenuation and
phase shift as a function of frequency. The application of these
relations to optical absorption and dispersion was first discussed
independently by Kramers and Kronig. Some similar relations in
nuclear physics are often called 'dispersion relations'; since this
name is commonly used in optics to denote relations between wave
number and frequency, it is probably better not to use it to signify
the Kramers-Kronig relations. The physical basis of the Kramers-
Kronig relations is the principle of causality. The cause in optical
phenomena is the incident electromagnetic wave, and the effect is
the response of the crystal, associated with the motion of electrons
and nuclei. The response at any point, at a time t = 0, can only
depend on the fields at that point for t <; O. It will be assumed that
the response is linear so that a Fourier component of field of fre-
quency w causes a response only at the same frequency w.
The Kramers-Kronig relations will be derived first for the electric
polarization P and then the generalization to other functions will
be considered. pet) will be used to denote the component of P
parallel to one of the principal axes as a function of time t. E(t)
represents the component of E in the same direction. The general
linear relation between pet) and E(t) is:

4nP(t) = J~F(t')E(t - t') dt'. (1.40)


The 4n has been introduced for convenience in later calculations.
pet) depends on all previous values of E, thus the relation between P
and E is 'non-local' with respect to time and we have the pheno-
14 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

menon of dispersion. The physical significance of the response


function F(t) can be made clearer by considering a delta function
pulse of electric field: E(t) = <5(t). Equation (lAO) gives:
4nP(t) = f: F(t')<5(t - t') dt' = F(t). (1.41)

Therefore F(t) represents the electric polarization due to a delta


function pulse of electric field. F(t) is zero for t < 0, and falls to
zero as t -+ 00 but may undergo oscillations on the way. If we take
E(t) = Eo exp (-iwt) then (lAO) becomes:

4nP(t) = Eo exp (-iwt) f: F(t') exp (iwt') dt'.

lt follows that the dielectric constant e(w) is related to the response


function F(t) by:
sew) - 1 = f:F(t') exp (iwt') dt'. (1042)

Hence 8l(W) - 1 = f:F(t') cos (wt') dt';


(1043)
82(W) = f:F(t') sin (wt') dt'.

f
Inversion of the Fourier transform in (1.42) gives:
F(t) = -1 00
{e(w) - I} exp (-iwt) dw
2n -00

foo {(8 l(W) - 1) cos wt + elw) sin wt} dw. (1.44)


= -1
n 0
The causality condition, F(t) = 0 for t < 0, is satisfied if for t > 0:

f:{8 l (W) - l} cos wt dw = f: 82(W) sin wt dw. (lAS)


Hence for t > 0:
F(t) = -2foo{81(W) - I} cos wt dw = -2foo 82(W) sin wt dw. (1.46)
n o n 0
By using these expressions for F(t) in (1.43) one can derive two
integral relations between 8l(W) and 82(W):
8l(W) - 1 = (2In)f:dw' e 2(w'){w'/(w'2 - w 2)};
(1.47)
e2(w) = (2In) f:dw' 8l (w'){wl(w 2 - W'2)}.
MACROSCOPIC THEORY 15
This derivation of the Kramers-Kronig relations (1.47) has the
virtue of showing clearly the connection with the causality condition.
A more elegant and rigorous proof makes use of integration in the
complex frequency plane. *
The first relation of (1.47) is particularly useful in the interpreta-
tion of experimental data. If a localized peak in 82(W), associated
with a particular set of transitions, can be isolated, then this relation
can be used to find 81(W) due to this peak. It is often convenient to
use the optical conductivity O'l(W) instead of 82(W) and the relation
then becomes:

f
81(W) - 1 = 8 dw' 0'1(w')/(w'2 - w 2) (1.48)

where the integration is taken over the frequency range of the peak.
The limiting value of 81(W) for zero frequency is:

f
81(0) - 1 = 8 dw' 0'1(w')lw'2

= (4jnc) f
0'1(.1') d.1'. (1.49)

0'1(.1) represents 0'1 as a function of wavelength .1. The high-frequency


value of 81(W) for w ~ w' is:

f
81(W) - 1 = -(8/w 2) O'l(W') dw'. (1.50)

Formula (1.48) can be put into a convenient form for computation


by transforming to a logarithmic scale offrequency, s = In w. With
this scale, equation (1.48) becomes:

81(W) - 1 = (-w4) fro smhs'


0
-.ds'
-{O'l(S + s') - O'l(S - s')}. (1.51)

It is clear from (1.51) that the contribution to the integral from


points near w' = w (s' = 0) is proportional to the gradient of O'l(S).
The application of (1.51) to some experimental data on tellurium
is illustrated in fig. 1.3. The optical constants were measured by
reflection from an evaporated film of tellurium. The tellurium crystal
is uniaxial so a film of randomly oriented crystallites should have
8 = t(8 11 + 822 + 833) = t(28 11 + 833). The low-fniquency experi-
mental values of 81 for the film do agree with t(28 111 + 8133),
* Landau and Lifshitz, pp. 256-62.
16 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

20

Ie
:t.
10
~
0

40 0

o 5
1rw(eV)
Figure 1.3 Optical properties of tellurium. The solid curve for the optical con-
ductivity «(/de) represents experimental values for polycrystalline films, with an
extrapolation at high frequencies. The solid curves for the dielectric constant 81
represent experimental values; the dotted curve was calculated from the measured
(/1 using the Kramers-Kronig relation (1.51).

References: R. s. CALDWELL and H. Y. FAN, Phys. Rev., 114, 664--75 (1959);


J. N. HODGSON, J. Phys. Chern. Solids, 23,1737-42 (1962).

measured for a single crystal, showing that the assumption of


random orientation is valid. The experimental values of B1(W) are
eonsistently greater than {B1(W) - I} calculated from (1.51) by a
nearly constant difference of 6·0 ± 0·5. This shows that the optical
conductivity at frequencies above the measured range is contributing
about 5·0 to 81(W) at low frequencies. The approximate constancy of
81(W) in the infra-red shows that there are no strong absorption
bands in this region. In the far infra-red and radio regions the free
carriers in tellurium will have an appreciable effect on dispersion.
A value for f <11(w') dw' can be derived from the high-frequency
MACROSCOPIC THEORY 17
variation of 81(W) or by integration of O'l(W). The interpretation of
this integral is discussed in Section 1.6.
Insulating crystals have a range of frequencies extending from
W = 0, in which O'l(W) = O. If the crystal has lattice absorption
bands there is a region of non-zero O'i(w) in the infra-red. At higher
frequencies strong absorption associated with interband transitions
commences but there is usually a region between the absorbing
regions, in which O'l(W) = O. When O'l(W) = 0 the crystal is trans-
parent and the optical properties are given by the real dielectric
constant 81(W) or the refractive index n(w) (= 8 11). The frequency
variation of 81(W) in a region of transparency can be expressed by
means of dispersion formulae which have been in use for many years.
It is the intention here, to show the connection of these formulae
with the Kramers-Kronig relations, which have quite general
validity. In terms of wavelength A (= 21'&cjw), (1.48) becomes:

81(A) - 1 = (4j1'&c) fdA' O'l(A')j{1 - A'2jA 2}. (1.52)

If O'l(A) can be represented as a sum of peaks at mean wavelengths


A1, A2, etc. then the weighting factor {I - A'2jA 2} can be taken as
{I - A12jA2}, etc. for wavelengths far from the peaks. Thus in
regions of transparency 8 1(A) is given approximately by:
8 1(A) - 1 = ~ {AiA. 2j(A2 - A.i 2)}. (1.53)
i

This formula was first suggested by Sellmeier on the basis of a classi-


cal oscillator model. Ai is the value of (4j1'&c) f O'l(A') dA.' for the ith
peak. For infinite wavelength, zero frequency, (1.53) becomes:
8 1(0) - 1 = ~ Ai. (1.54)
i

The dispersion of crystalline sodium chloride in the visible region can


be accurately expressed by a four-parameter equation of the form
(1.53). The values of Ai and Ai are:
Ai 0·052 1·005 0·271 3·537
Ai 0·0347 0·1085 0·1584 61·67 pm
Some recent experimental data on 0'1(.1.) for sodium chloride in the
ultra-violet is shown in fig. 1.4. There are two main peaks at 0·111
and 0·159 pm and the area under the curve can be divided as shown
to give corresponding Ai values of 1·008 and 0·274. The two main
18 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

20

15

Ie
3- 10
u
......
6"

,,
.
,

0
"" 0·20

Figure 1.4 Optical conductivity u 1 of crystalline sodium chloride as a function


of vacuum wavelength A in the ultra-violet.
Reference: D. M. ROESSLER and w. c. WALKER, Phys. Rev., 166, 599-606 (1968).

peaks in the experimental 0"1(1,.) curve correspond closely in position


and area to the two main ultra-violet peaks predicted by the dis-
persion formula. The dispersion depends mainly on the broad
distribution pattern of 0"1(1,.) and is not sensitive to fine detail in the
0"1(1,.) curve. From the dispersion analysis, el(O) = 1 + ~ Ai = 5'86,
i
while the experimentally measured el(O) = 5·90.
The Kramers-Kronig relations (1.47) were derived for the function
e(W) but they can also be applied to other functions. For example,
(n + iK) = e! is the ratio between the magnitudes of Hand E, and
the arguments leading to (1.47) can be applied with H in place of
4nP. This leads to relations of the form (1.47) between n(w) and
K(W). The general significance of the 1 in the first equation of (1.47)
is that it represents the value of el(W), n(w), etc. for w ~ 00. The
Kramers-Kronig relations apply in fact to any function of e(W) for
which the integrals converge. This includes the various reflection
MACROSCOPIC THEORY 19
coefficients which can be expressed as functions of sew). The causality
argument can also be applied to the surface admittance Y and the
Kramers-Kronig relations can be derived for Yl and Y2 , including
frequency ranges where the relation Yew) = stew) is not valid.
One relation which has been widely used in the analysis of experi-
mental data gives the phase change on reflection, Ll, in terms of the
measured reflectance R(w). The experimental data usually refer to
nearly normal incidence where R is independent of polarization for
cubic crystals. The logarithm of the reflection coefficient is:
In (Rt exp ill) = 1- In R + ill. (1.55)
The Kramers-Kronig relation which gives Ll in terms of R is:

Ll(w) = (2/n) J~ dw' a In R(w')}{w/(w 2 - W'2)}.

A convenient form of this equation for calculation is obtained by

J:
integrating by parts:

Ll(w) = (1/2n) dw' (d/ dw'){ln R(w')}ln {Iw - w'I/(w + w')}. (1.56)

The contribution to Ll(w) for frequencies near w' is proportional to


the gradient of In R(w'), and the weighting factor is
In {I w - w' !I(w + w')},
which has a logarithmic infinity at w' = wand tends to zero for
w' --+ 0 and w' --+ 00. The optical constants nand K can be calculated
from Ro and Ll using the formulae (1.31). This method of measuring
optical constants has the advantage of experimental simplicity since
it requires only the measurement of normal reflectance Ro at each
wavelength. It has been widely used to determine optical constants
in the ultra-violet and infra-red regions. The main difficulty of the
method lies in the extrapolation of R(w) beyond the measured
region to perform the integration of (1.56).

1.6 The sum rule


An expression for Sl(W) at high frequencies has been given in (1.50):

Sl(W) - I = J
-(8/w 2) O"l(W') dw'.

This formula is valid when the photon energy liw is much greater
than the energies liw' of strong interband transitions. In classical
20 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

terms, the frequency W is much greater than the resonant frequencies


of the electrons. If this condition is satisfied for all electrons in a
solid, including core electrons, then the behaviour will correspond
to a free electron gas. A simple calculation given in Section 4.1
shows that for a free electron gas:
8 1(W) - 1= _Wp2/W2 (1.57)
where wp = (4nNe 2/m)! is called the plasma frequency; N is the
number of electrons per cm 3 • A comparison of (1.50) and (1.57)
shows that:

(1.58)

This is one of the sum rules for electrons. The limits of integration
have been taken as 0 to 00 since (1.58) refers to all electrons in the
solid including the core electrons for which w' extends well into the
X-ray region.
The sum rule can often be applied approximately to certain groups
of electrons in a solid, such as the valence electrons. If O'l(W') due to
the valence electrons is appreciable only in the range 0 < w' < WI
and WI ~ W2 the threshold frequency for core electron transitions,
then approximately:

(4nN"e 2/m) =8 f"'l


0 O'l(W') dw' (1.59)

where N'f) is the number of valence electrons percm3 • A more


sophisticated form of the sum rule, which includes the effect of
interaction between different groups of electrons, is considered in
Section 3.2. The effective number of electrons or the effective oscilla-
tor strength per atom for transitions in the frequency range
o <w' <w,
is defined by:

(1.60)

No is the number of atoms per cm3 • Experimental curves ofn*(w) for


aluminium, silicon and germanium are shown in fig. 1.5. The curve
for aluminium rises from the origin since 0'1(0) is non-zero for a
metal. The small number of free carriers in pure silicon or ger-
manium makes no appreciable contribution to n*(w) on the scale of
MACROSCOPIC THEORY 21

~w.(eV)

Figure 1.5 Effective number of electrons per atom n* defined by formula (1.60).
Reference: H. R. PHILIPP and H. EHRENREICH, J. Appl. Phys., 35, 1416-19 (1964).

this graph. n*(w) does not rise appreciably above zero until well
above the minimum energy gaps. The curves· for aluminium and
silicon are rising towards 3·0 and 4'0, respectively, at the highest
energies. These are the values predicted by the approximate sum rule
(1.59). For germanium the curve of n*(w) rises well above 4·0
indicating an appreciable effect of core electrons.
The considerations of this section have been limited so far to
absorption associated with electronic transitions. The infra-red
lattice absorption bands of ionic solids are due to the vibrations of
ions with masses of the order of 10 4 to 10 5 electron masses. Since
ionic charges are of the order of the electron charge, the plasma
frequency for ions is approximately:
wp = (41tNe 2 jM)1 (1.61)
where M is the mean ionic mass. It follows from (1.58) that the
effective oscillator strength of lattice absorption bands is of the order
(mjM) times smaller than that of electronic absorption bands.

1.7 Dispersion theory of classical oscillators


Lorentz formulated a classical theory of optical dispersion in terms
of electrons bound by elastic forces to their equilibrium positions.
Under the influence of an electromagnetic wave these electrons per-
form forced oscillations with amplitude and phase depending on
22 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

their resonant frequencies and damping coefficients. Let us consider


the equation of motion for the x-displacement of a vibrating electron:
(d 2x/dt 2) + y(dx/dt) + W02x = (-eEo/m) exp (-iwt).
Eo exp (-iwt) is the x-component of the electric field; -my(dx/dt)
represents the damping force; -mw02x represents the restoring
force. The solution for forced oscillations of frequency w is:
x = (-eEo/m) exp (-iwt)/(w 02 - w 2 - iwy).
The electric polarization P z due to this displacement is ( - Nex). The
dielectric constant 8(W) is given by:
8(W) - I = 4nPz/Ez •
Hence:
(81(W) - I} + i8 2(W)
= wp 2{(W 02 - w 2) + iwy}/{(w 02 - W2)2 + w 2y2} (1.62)
where wp 2 = (4nNe /m). The optical conductivity is:
2
O'I(W) = W8 2(w)/4n = (w 2N4ny)/{1 + Q2(w o/W - w/w o)2} (1.63)
where Q = wolY. The maximum of O'I(W) occurs at w = Wo and the
peak is symmetrical about Wo on a logarithmic scale of frequency;
the half-width is (ro o/ Q) if Q ~ 1.
The applications of the Lorentz formulae (1.62) and (1.63) are
much wider than the simple model used for their derivation might
imply. We saw in Section 1.5 that the electric polarization P(w)
could be considered as the Fourier transform of a response function
F(t). This function represents the displacement of electrons or ions
due to a delta function pulse of electric field. For the Lorentz
oscillator:
F(t) = A exp (-tyt) sin wot. (1.64)
Assuming that the damping is sufficiently small so that wolY ~ 1.
The only assumption needed to obtain (1.62) and (1.63) is that the
response function has the form {1.64). This means that the free
oscillations of the system decay exponentially with a constant relaxa-
tion time T (= 2/y). The mechanism of relaxation has been in-
vestigated in most detail for conduction electrons; note that in this
case Wo = 0 and T = IIY. The conditions under which a relaxation
time can be defined for conduction electrons are fairly well known.
In other cases the assumption that F(t) has the form (1.64) can only
be justified on the grounds of simplicity.
MACROSCOPIC THEORY 23
The Lorentz formulae have proved very useful in analysing the
infra-red dispersion of ionic crystals. If the variation of optical con-
ductivity with frequency can be represented by a sum of Lorentz
peaks, then the total CI(W) is equal to the sum of the corresponding
expressions for CI(W) given by the real part of (1.62). Spitzer and
104r---------------------------------------------------------~

04

~ 0·2

Figure 1.6 An infra-red absorption band of quartz. Theoretical curve for (uI/c)
according to (1.63) with Ao = (2nclw o) = 12'85 ,urn, (wo/Y) = 100, (Wp2/w0 2)
= 0·10 (contribution to Experimental curve for normal reflectance Ro
8 1 (0)).
with E parallel to c-axis.
Reference: w. O. SPITZER and D. A. KLEINMAN, Phys. Rev., 121, 1324-35 (1961).

Kleinman have shown that the complicated infra-red absorption


and dispersion spectrum of quartz can be represented by a sum of
seven Lorentz terms for each polarization, Ell
c and E -'-= c. The
experimental data consisted of normal reflectance values for wave-
lengths between 5 and 37 p,m. A set of Lorentz terms was chosen by
trial and error to give the experimental variation of Ro with A. In
regions of low absorption, between the absorption peaks, Ro is in-
sensitive to the actual value of absorption. As a practical limit, when
K <: 0·1 the effect of non-zero K on Ro is less than 1%. However,

experimental measurements of the transmission of quartz in the


c
24 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

regions of low absorption show generally good agreement with


values calculated from the Lorentz formulae, even though K " " 0·001.
The measured absorption peaks have the Lorentz form out to very
low absorption levels. Thus the analysis into Lorentz terms has more
physical significance for quartz than just a mathematical representa-
tion of the reflection spectrum.
2
Crystal Lattice Absorption
2.1 Vibrational modes of a crystal lattice
A crystal lattice consists of a regular array of unit cells each con-
taining a similar group of nuclei and electrons. At 0° K, each nucleus
is at rest in its equilibrium position, ignoring for the moment the
quantum zero-point vibration. If a nucleus is displaced from its
equilibrium position, a restoring force will act on the nucleus. The
restoring force F can be expressed as a polynomial in the displace-
ment ~:
(2.1)
This is a simplified version of the correct vector equation with tensor
coefficients. For sufficiently small ~, the first term predominates and
the oscillation of the nucleus about its equilibrium position is simple
harmonic. The approximation F = a~ is called the harmonic
approximation. The higher order, anharmonic terms in (2.1) can be
treated as small corrections in the theory of lattice vibrations.
The nuclei in a crystal lattice are coupled by interatomic forces so
they cannot vibrate independently of one another. We are dealing
with a system of coupled oscillators whose vibrations can be analysed
into normal modes. Each normal mode has an angular frequency ro,
a wave vector q and a polarization vector e, such that the displace-
ment of a nucleus due to the mode is given by:
~ = ~oe exp i(q.r - rot). (2.2)
If Q is a vector of the reciprocal lattice, then wave vectors q and
+
q Q are equivalent in (2.2) because of the periodicity of the lattice.
It is therefore only necessary to consider q-values in one Brillouin
zone of the reciprocal lattice. The zone centred on the origin q=(O,O,O)
is usually chosen. The quantum-mechanical theory of lattice vibra-
tions requires that each normal mode (2.2) be quantized into phonons
of energy nro and momentum nq.
2S
26 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

So far we have been considering the displacement of nuclei in the


lattice vibrations. Because of their much smaller mass, the electrons
will arrange themselves in nearly equilibrium configurations for each
nuclear displacement. In a lattice of inert gas atoms, the electrons
form closed shells around each nucleus and the interatomic forces
are very weak. The electrons and nuclei therefore vibrate together as
atomic units. In lattices of molecules with closed electron shells,
such as saturated organic molecules, the internal vibrations of each
molecule will not be greatly affected by intermolecular forces. The
ions in a simple ionic crystal, such as sodium chloride, also have
closed electron shells. To a first approximation, the electrons and
nuclei can be considered as moving together in the lattice vibrations.
For example, a sodium chloride crystal can be considered as equiva-
lent to a lattice of point charges +e and -e, with masses Ml and
M 2 , at the centres of the Na + and CI- ions. This is called the rigid
ion model because it neglects the distortion of the electron cloud
round an ion caused by a lattice vibration. In a covalent crystal the
valence electrons outside the closed shells are shared among neigh-
bouring atoms to form covalent bonds. The closed shell electrons
and the nuclei, which form the ion cores, are displaced together in
the lattice vibrations. The situation in metals is similar but the
valence electrons here form a continuous background instead of the
localized bonds of a covalent crystal.
The (w,q) relation for normal modes has several branches which
can be classified as either acoustic or optic branches. In an acoustic
mode of vibration all the nuclei in a unit cell move together in phase.
The acoustic branches of the (w,q) relation reduce to w = uq when
q -? 0, where u is the velocity of low-frequency acoustic waves. In
an optic mode of vibration there are phase differences between
the displacements of different nuclei in a unit cell, :n; radians for the
simplest case of two nuclei per unit cell. The optic branches of the
(w,q) relation go through non-zero values of w when q = 0. The
(w,q) relation for a cubic ionic crystal (KBr) is illustrated in fig. 2.1.
The curves show the relation between liw and I/A (= q/2:n;) for q in
the <111> directions; by crystal symmetry w(-q) = w(+q). The
normal modes in a cubic crystal are linearly polarized with e either
parallel or perpendicular to q; that is, longitudinal or transverse.
The large separation between the LO and TO branches is typical of
an ionic crystal.
CRYSTAL LATTICE ABSORPTION 27

0-021="""-------

TO

>
Q) 0-01
2

1/>" (10 6 em- 1 )


Figure 2.1 Phonon dispersion curves for potassium bromide at 90° K measured
by neutron diffraction. Wave vectors in the (111) directions; longitudinal optic
(LO), transverse optic (TO), longitudinal acoustic (LA) and transverse acoustic
(TA) branches. Wave number q = 2n/)..
Reference: W. COCHRAN and R. A. COWLEY, 'Handbuch der Physik', Vol.
XXV/2a, pp. 59-156 (1967).

The number of normal modes can be found by the application of


boundary conditions for the displacement. The normal modes cor-
responding to a particular branch can be represented by points on
q-space with position vectors equal to the wave vectors of each mode.
The number of representative points in a volume Vq of q-space is
(VqJ8:n: 3) for a non-degenerate branch. The total number of rep-
resentative points in the Brillouin zone is N, the number of unit cells
in the crystal. A degenerate branch occurs when normal vibrations
with two independent polarization vectors e have the same (w,q)
relation. For example, the transverse modes in a cubic crystal are
degenerate and the density of representative points is 2 X (1 J8:n: 3). The
total number of modes in the acoustic branches is 3N, while the
optic branches contain the remaining 3(s - l)N modes where s is
the number of nuclei in a unit cell. In the calculation of optical
absorption it is sometimes necessary to know the number of modes
in a certain branch as a function of energy liw. This can be deduced
from the number of representative points in q-space between neigh-
bouring constant energy surfaces. The degree of excitation of a
28 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

given mode in thermal equilibrium, that is the number of phonons


in that model, will depend on the temperature T. The average
number of phonons in a mode of energy noo is given by the Bose-
Einstein formula:
ii = {exp (noo/kT) - 1}-1. (2.3)

2.2 Photon-phonon interaction


The frequencies of lattice vibrations lie in the infra-red region of the
spectrum. An electromagnetic wave of higher frequency is trans-
mitted by an ionic crystal with very little absorption if the frequency
is below the threshold for interband electronic transitions. The phase
velocity of this wave is determined by the electronic polarizability of
the ions since the large masses of the nuclei prevent them from
responding to a high-frequency field. The dielectric constant associ-
ated with the electronic polarizability is usually denoted by e( 00).
The 'infinite' frequency is really a frequency much larger than the
lattice vibration frequencies, but less than the interband threshold.
The phase velocity of high-frequency waves is C/et(oo) where
etc 00) is usually between 1·3 and 2·5 for ionic crystals. The sloping
dotted line in fig. 2.2 represents the dispersion relation for high-
frequency waves in crystalline potassium bromide. A comparison
with the (oo,q) relation in fig. 2.1 shows that the phase velocity of
acoustic lattice waves is much smaller than c/e!( 00), which is to be
expected as we are comparing sound waves with light waves. The
sloping dotted line in fig. 2.2 intersects the optic branches of the
(oo,q) relation close to q = O. For the range of q shown in fig. 2.2,
the optic branches can be approximated by horizontal lines,
oo(q) = 00(0).
At the intersections of the photon dispersion line with the TO and
LO phonon branches, the frequency and wavelength conditions for
a resonant interaction are satisfied. It is also necessary for a reson-
ance that there should be coupling between electromagnetic and
lattice waves. We will show below that electromagnetic waves are
coupled to the TO vibrations but not to the LO vibrations.
The displacement of ions from their equilibrium positions pro-
duces an electric dipole moment in each unit cell and an average
polarization P. We will limit our considerations to ionic crystals
CRYSTAL LATTICE ABSORPTION 29
0·04.------------------,

0·03

">
~ LO
0·02 ,
3
.c: ,,

----------------~:~~~----------------~~
0·01 ////-

~/,//
400
1/)., (em-f)

Figure 2.2 Photon-phonon interaction in potassium bromide. W and TO are


the phonon branches of fig. 2.1 for small wave numbers. The sloping dotted line
represents photons without phonon interaction; OJ = ck/st(OJ). The curves refer
to mixed photon-phonon excitations. Wave numbers k or q = 2n/A.

with just one positive and one negative ion per unit cell. The dipole
moment will depend on the relative displacements of the positive
and negative ions, say (u+ - u-). It is convenient to use another
variable HI to specify the relative displacement, where HI is defined by:
HI = {M+M-/(M+ + M-)}iv-i(u+ - u-). (2.4)
M + and M - are the ionic masses and v is the volume of a unit cell.
In the harmonic approximation, the equation of motion of the ions
and the equation for the polarization are:
+
(d 2H1/dt 2) = buHl buE; (2. Sa)
(2.5b)
Microscopic models for the calculation of the coefficients b u , bn
and b 22 will be considered later. The equality of the two coefficients
b12 follows from the conservation of energy. For solutions with a
time-dependence exp (-iwt), we can put (d 2H1/dt 2) = -w 2 w.
Eliminating HI between (2.5a) and (2.5b) we obtain a relation
30 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

between P and E and hence an expression for the dielectric constant


s(w):
s(w) = (1 +
4nP/E)
= 1 4nb 22 + +
4nb122/( -b ll - w 2)
= s(oo) +
{s(O) - s(oo)}/{l - W 2/W 0 2} (2.6)
where
s(oo) = (1 + 4nb 22 ); s(O) - s(oo) = 4nb 12 2/(-b ll); W 02 = (-b n ).
The transverse wave solutions of Maxwell's equations with dielectric
constant s(w), given by (2.6), represent the coupled photon-phonon
modes of the crystal. The dispersion relation between the frequency
and wave number k of the coupled wave, w = {c/st(w)}k, has two
branches as shown by the curves in fig. 2.2. At low frequencies
w = {c/s~(O)}k, but as the frequency increases towards Wo, the lower
branch approaches the TO branch of lattice vibrations. s(w) is
negative for frequencies between Wo and WL' where WL is given by:
WL = wost(O)/s~(oo). (2.7)
For frequencies above WL the (w,k) curve tends asymptotically to
w = {c / st( (0) }k. When s(w) has a negative value, the optical con-
stants are n = 0 and K = (-s)t, corresponding to an exponentially
damped field without energy loss: 8 2 = O. An external wave in this
frequency range is completely reflected by the crystal. This is the
Reststrahlen band of the crystal, a region of high reflectivity in the
far infra-red. An experimentally observed Reststrahlen band for
indium antimonide is shown in fig. 2.3. Although the binding in this
crystal is mainly covalent, there is an appreciable imbalance between
the charge distributions around indium and antimony atoms, causing
a dipolar coupling with an electromagnetic wave. The rounded curve
represents the experimentally measured reflectance while the curve
with sharp corners was calculated using equation (2.6). Outside the
central part of the Reststrahlen band the experimental and calcu-
lated curves are seen to agree closely. However, the experimental
reflectance does not rise to 100% in the central part of the band as
required by theory. The reason for this discrepancy is the neglect of
damping in the theory; the inclusion of damping will be considered
in the next section.
The displacement of ions in a transverse electric field will also be
transverse, as shown by the form of (2.5a). Similarly, longitudinal
displacements, w II k, will be associated with a longitudinal field,
CRYSTAL LATTICE ABSORPTION 31

40

20

o~~~--------~~--~--------~
160 200 240
11 hO (em-I)
Figure 2.3 Reststrahlen band of indium antimonide. Curves of normal re-
flectance Ro; Ao = vacuum wavelength. Full curve: experimental reflectance.
Dashed curve: theoretical reflectance without damping.
Reference: R. B. SANDERSON, J. Phys. Chern. Solids, 26, 803-10 (1965).

Ell k. It was assumed in Section 1.4 that Maxwell's equation


V.D = 0, for a crystal not containing any extra charges, implies
that Ei, = O. For a plane wave of wave vector k, the condition
V.D = 0 becomes:
e(w)k.E = O. (2.8)
This equation (2.8) can be satisfied by a non-zero Ei" that is, k .E :;C. 0,
if there is a frequency or frequencies for which:
e(w) = O. (2.9)
From equation (2.6) for the dielectric constant of an ionic crystal,
we find that condition (2.9) is satisfied for w = WL defined in (2.7).
Longitudinal optic waves can exist only at frequencies which satisfy
(2.9), independently of the wavelength A.. This conclusion is only
accurate within the limitations of macroscopic theory and correc-
tions will be needed when A. is of the order of the lattice constant.
32 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The dispersion curve for a typical LO mode over the whole range of
A is shown in fig. 2.1. When damping is included the expression for
e(W) gives a complex value for all real frequencies. The condition
(2.9) can however be satisfied for a complex value of w:
WL = WL' - iWL". (2.10)
The time-dependence of a longitudinal oscillation at frequency WL is
given by exp {( -iWL' - WL")t}. In quantum language, the longi-
tudinal phonon has a finite lifetime (l/WL") owing to scattering. If

~
I- 50

500 250
1/).o(cm- 1 )

Figure 2.4 Transmission spectrum of lithium fluoride film. Transmittance T;


vacuum wavelength Ao. Wave of p-polarization; angle of incidence"'" 30°; film
thickness 0·20 pm.
Reference: D. w. BERREMAN, Phys. Rev., 130,2193-8 (1963).

the damping is small, W L' is approximately equal to the real fre-


quency at which I e(W) I is a minimum or I e(W) 1-1 is a maximum.
An excitation similar to a longitudinal wave can be produced near
the surface of a crystal by an incident electromagnetic wave. A
p-wave incident obliquely on a plane surface has a component of
E normal to the surface, say .til' The surface charge produced by
.til is similar to the distribution of charge associated with a longi-
tudinal wave. A calculation of the absorptance of a thin layer in
Section 1.4 showed that the part of the absorptance corresponding
to .til is proportional to e 2/1 e 12. This will be a maximum for W ~ W L.
An experimental transmission spectrum of a p-wave through a
lithium fluoride film is shown in fig. 2.4. The main dip represents the
CRYSTAL LATTICE ABSORPTION 33
absorption of the transverse component near 0) = 0)0 while the
smaller dip is centred near the longitudinal optic frequency O)L.

2.3 Microscopic theory of infra-red dispersion


The macroscopic electric field E in a crystal is an average over one
unit cell. To obtain the effective field E' acting on an ion the field of
the ion itself must be subtracted from E. The usual method of calcu-
lating E' is to find the field at the site of an ion due to all the other
ions. Lorentz showed that for a cubic lattice the effective field is
given by:
E' = E + (4n/3)P. (2.11)
The same result is valid for the random arrangement of molecules
found in a liquid or gas. The coefficient of Pin (2.11) has other
values for non-cubic crystals and may be anisotropic for lattices of
lower symmetry. The earlier theories of ionic crystals considered the
ions as well-separated, rigid units. However, to explain the stability
of the crystals it is necessary to assume some overlap between the
wave functions of neighbouring ions. The agreement of the experi-
mental cohesive energies with the values calculated from the Coulomb
forces between rigid ions indicates that the overlap is small. A refine-
ment of the rigid ion model allows for deformation of the ions by
electric fields and the overlap repulsion forces.
Let us consider the polarization P due to displacements u+ and
u- of the ions from their equilibrium positions. The deformations of
the ions by electric fields can be represented by the electronic
polarizabilities oc+ and oc-. The deformations by the overlap forces
can be represented in a simple approximation by dipole moments
proportional to the displacements. This correction is included in the
formulae by using the Szigeti effective charges ±(Z*e) in place of
the ionic charges ±(Ze):
p = v- 1{Z*e(u+ - u-) + (oc+ + oc-)E'}. (2.12)
After substituting for E' from (2.11) and introducing w defined by
(2.4), equation (2.12) is comparable with the macroscopic equation
(2.5b). By equating coefficients one obtains:
b12 = Z*e(Mv)-t{l - (4n/3)(oc+ + oc-)V-1}-1 (2.13)
(2.14)
34 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

where M = {M+M-/(M+ + M-)}. The coefficient b 22 represents


the electronic polarization and is related to the high-frequency di-
electric constant s( (0):
1 + 4nb 22 = s( (0).
Therefore:
{s(oo) - 1}/{s(oo) + 2} = (4n/3)(~+ + ~-)V-l. (2.15)
A comparison of experimental values of s( (0) for the alkali halides
with (2.15) has shown that it is possible to choose a self-consistent
set of values for ~+ and ~-. This lends support to the original assump-
tion that the overlap of ions is quite small.
Equation (2.13) can be rewritten in terms of 10(0), s( (0) and w o, as
follows:
10(0) - 10(00) = 4nb 12 2w o- 2.
Therefore:
10(0) - 10(00) = 4n(Z*e)2(Mv)-lw o-2{1 - (4n/3)(~+ + ~-)V-l}-2.
Hence, using (2.15):
10(0) - 10(00) = 4n(Z*e)2(Mv)-lwo-2{(j-)(s(00) + 2)P (2.16)

TABLE 2.1: SZIGETI EFFECTIVE CHARGE NUMBERS

Crystal NaCI KBr CsBr MgO ZnS SiC GaAs InSb

Z* 0·74 0·76 0·79 1·76 0'96 0-94 0-51 0-42

The deviations of Z* from 1·00 for the alkali halides are in the
direction to be expected when the negative ion is more easily deform-
able than the positive ion. The values of Z* for the covalent semi-
conductor crystals are seen to be comparable with the values for the
alkali halides.
The equation of motion of the ions, corresponding to the macro-
scopic equation (2.5a), is:
Md 2(u+ - u-)/dt 2 = -G(u+ - u-) + Z*eE'. (2.17)
-G(u+ - u-) is the overlap force between neighbouring ions. The
coefficient G can be related to the compressibility of an ionic crystal.
A comparison of equations (2.5a) and (2.17) yields a relation be-
tween b n (= -W 0 2) and G, from which a relation between Wo and
CRYSTAL LATTICE ABSORPTION 35
compressibility follows. This relation agrees quite well with the
experimental data for the alkali halides.
The disagreement between the theoretical and experimental re-
flectance curves for the central part of the Reststrahlen band, as
illustrated in fig. 2.3, shows that the theory leading to equation (2.6)
is incomplete. A comparison of (2.6) with the first of the dispersion
relations (1.47), shows that the real part of s(w), as given by (2.6),
must be associated with an imaginary part, S2(W), which is propor-
tional to a delta function at w = wo. In terms of the optical con-
ductivity:
(2.18)
The theory predicts an infinitely sharp absorption line because of the
neglect of damping terms in the equation of motion of the ions. In
the harmonic approximation, the lattice vibrations are undamped
and the phonons are independent non-interacting excitations of the
lattice. The harmonic approximation includes only second order
terms in the expression for the lattice potential energy in terms of the
ion displacements. If the third order terms are included, the normal
modes no longer represent independent excitations and phonon-
phonon scattering occurs. This scattering can be calculated by
perturbation theory assuming that the anharmonic forces are much
smaller than the harmonic forces.
A simple classical theory of the effect of damping on the optical
conductivity is obtained by assuming that the TO mode has a relaxa-
tion time (1/r). The discussion of Section 1.7 shows that the
corresponding O'l(W) is given by a Lorentz peak:
!!02 2rw 2
O'l(W) = 4n {s(O) - s(OO)}(!!02 _ W2)2 + 4r 2w 2' (2.19)

The delta function of (2.18) has become a peak centred at


w = !!o = (W 02 + r2)1,
with half-width 2r. Typical experimental values of O'l(W) for an
alkali halide in the region of the Reststrahlen band are shown by the
points in fig. 2.5. The Lorentz curve, which was fitted to the experi-
mental maximum and half-width, represents the experimental values
quite well near the peak. The difference between Wo and !!o is only
0·1 % in this case. The explanation of the side-band at frequencies
above Wo is taken up in the next section. To obtain better agreement
36 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

o
o
o
o

10ofL-------------~~-__110

1/Ao(cm-1)
Figure 2.5 Optical conductivity (1, and normal reflectance Ro of caesium bromide
in the region of the Reststrahlen band. Optical conductivity: experimental
points and theoretical Lorentz oscillator curve. Reflectance: experimental curve.
Reference: R. GEICK, Z.Jiir Physik, 163,499-522 (1961).

with the shape of the main band, a quantum-mechanical theory of


anharmonic broadening must be used.
The quantum-mechanical theory shows that the interactions of
phonons satisfy the laws of conservation of energy and momentum.
The sums ~ nWi and ~ nq are conserved in a scattering process. A
i i
phonon has a finite lifetime because it can be scattered by other
phonons or it can change into two or more phonons. The lifetime r
in the quantum-mechanical theory is a function of frequency and
can be conveniently calculated by the methods of quantum field
theory. An imaginary part is added to r to represent the effect of
damping on the self-energy of the phonon; this means that Q 0 is
also a function of the frequency. The maximum of O"l(W) occurs at a
frequency wo' which is the solution of wo' = Qo(wo') and the half-
width of the peak is approximately 2r(wo'). The real part of the
CRYSTAL LATTICE ABSORPTION 37
dielectric constant, £l(W), can be derived from O'l(W) by using the
dispersion relation:
£l(W) - £1(00) = 8J:dw' 0'1(W')/(W'2 - w 2). (2.20)

When rand Q 0 are functions of w (2.20) will not give the classical
oscillator formula for £l(W).
The temperature-dependence of the width of the main peak can be
deduced from the phonon occupation numbers of the relevant modes.
At high temperatures, the width due to third order anharmonic
terms increases linearly with temperature, T. Experimental values
of the width for lithium fluoride and sodium chloride show a
temperature-dependence between T and P at high temperatures.
This indicates that the influence of fourth order anharmonic terms
on the width is not negligible at high temperatures. These terms
give a contribution to the width which varies as P.

2.4 Two-phonon absorption


Besides the main Reststrahlen band, ionic crystals have other
weaker, absorption bands in the infra-red such as the side-band
shown in fig. 2.5. These bands can be explained by processes in
which a photon is absorbed by the lattice with the creation of two or
more phonons. The anharmonic terms described in the previous
section can cause this type of absorption or it may be caused by the
second order electric moment discussed below. Two-phonon ab-
sorption of radiation occurs at frequencies which are the sum or
difference of phonon frequencies. They correspond to the classical
combination frequencies of normal modes mixed by non-linearities.
The infra-red absorption of the four-valent crystals, diamond,
silicon and germanium, consists of a series of peaks, as shown in fig.
2.6 for silicon, with a total strength only about 0·001 of the Rest-
strahlen bands of ionic crystals. The diamond lattice structure,
shared by silicon and germanium, consists of two interpenetrating
face-centred cubic lattices. The unit cell contains two atoms which
vibrate relative to each other in the optic modes. It follows from
crystal symmetry that the dipole moment associated with the optic
modes is zero, even when anharmonic terms are included. The one-
phonon absorption and the two-phonon anharmonic absorption are
therefore zero. There remains the absorption associated with the
38 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

10

lIAo (em-I)
Figure 2.6 Two-phonon lattice absorption bands in silicon. Absorption co-
efficient K (= 4:rtK/J. o); vacuum wavelength J. o•
Reference: F. A. JOHNSON, Progress in Semiconductors, 10, 180-235 (1966).

second order electric moment to explain the experimental results. A


normal mode vibration causes local distortions of the crystal lattice
and hence local unbalance of the charge distributions. A second
mode can interact with these charge oscillations to produce a second
order electric dipole moment, M:
M=U.. C·Us' (2.21)
where C is a tensor and Us and Us' are displacements associated with
the modes sand s'. C· Us' represents the effective charge distribution
produced by the s' mode. The absorption associated with M can be
calculated from its matrix elements between initial and final states.
Details of a similar calculation for interband absorption are given in
Chapter 3. Frequencies WI and Ws appearing in Us and Us' will be
mixed by (2.21) to give frequencies (WI ± ws) in M. The frequency
(WI + ws) corresponds to the creation of two phonons while
(WI - ws) corresponds to the creation of one phonon and the
annihilation of another phonon. The conservation of energy and
momentum is expressed by:
liwi ± liws = liw; (2.22)
Iiq1 ± liqa = lik (2.23)
CRYSTAL LATTICE ABSORPTION 39
liw and lik are the energy and momentum of the absorbed photon.
k is much less than q for most of the phonons so (2.23) can be re-
written:
(2.24)
The absorption in summation bands is proportional to:
~ I M(q,s,s') 12 WI-1Wa-1F(Wt>wJ ~(w - WI - wJ; (2.25)
M(q,s,s') is the matrix element of M between the two phonon states
and the sum is taken over all pairs of states with an energy difference
Ii(WI + wJ. F(Wl,WJ is a function of the phonon occupation num-
bers of the two modes. So far, very little progress has been made with
the detailed calculation of the matrix elements M(q,s,s'). It seems
reasonable as a first approximation to assume that M(q,s,s') is It.
slowly varying function of q for a given pair of branches (s,s'). The
dominant factor determining the structure of the absorption spec-
trum will then be the density of pairs of phonon states with an
energy difference Ii(WI + wa)' If the density of states has strong
maxima at WI and Wa we expect to find a two-phonon absorption
maximum at (WI + wa). The density of states for a given branch of
the vibration spectrum usually has one main maximum at a charac-
teristic energy. By assigning suitable values to these characteristic
energies, the peaks in the two-phonon spectrum can be assigned to
pairs of phonon branches. For example, the two-phonon spectrum
of silicon, shown in fig. 2.6, has been. analysed in this way using the
following characteristic energies: TO (0·0598 eV), LO (0·0513 eV),
LA (0·0414 eV), TA (0·0158 eV). These values agree quite well with
estimates from the calculated density of phonon states for silicon.
The agreement between measured and predicted peak energies is
shown in Table 2.2.
A more detailed comparison of theory and experiment has been
made using the method of critical point analysis. The concept of
critical points of a combined density of states will be discussed in
Chapter 3 in connection with interband transitions.
Similar calculations can be made for the difference bands,
W = (Wl- wJ
and three-phonon bands such as
W = (WI + Wa + w 3).
The maximum energy for two-phonon absorption in silicon is about
0·12 eV (1000 cm- 1) and the absorption at higher energies can be
D
40 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

TABLE 2.2: TWO-PHONON ABSORPTION BANDS OF


SILICON

Phonon branches Peak energies

Measured Calculated

TO+TO 0·1195 eV 0·1196 eV


TO+LO 0·1111 0·1111
TO+LA 0·1015 0·1012
LO+LA 0·0917 0·0927
TO+TA 0·0756 0·0756
LO+TA 0·0702 0·0671

interpreted as three-phonon absorption. The temperature variation


of the absorption spectrum is very useful in assigning peaks to
appropriate phonon processes. Assuming that M(q,s,s') is nearly
independent of temperature, the temperature-dependence of absorp-
tion is given by the function F(WloW2). In each type of absorption,
F(WloWJ is the difference between terms representing absorption
and stimulated emission of photons. Expressions for F(WI,W2) in
terms of the occupation numbers fi., fi." defined by (2.3), are given
below:
co F(C01oC02)
Summation band: COl + CO2 (1 + ih)(1 + ii 2) - ii l ii 2 = 1 + iii + ii2
Difference band: COl - COs ii l (1 + ii2) - ii2(1 + iii) = iii - ii2•
As the absolute temperature tends to zero, fil and fi2 tend to zero, so
that the summation bands approach a limiting intensity while the
difference bands disappear.
3
Interband Transitions
3.1 Electron energy bands
An isolated atom has a characteristic set of wave functions and
energy eigenvalues merging at high energies into an ionization
continuum. All the excited states have a finite lifetime T for spon-
taneous transitions to a state of lower energy. It follows from the
uncertainty principle that the energy levels have a width of order
Ii/T. This natural width is usually less than 10-6 eV and is often
masked in practice by broadening due to other causes. The absorp-
tion spectrum of an atom consists of a series of lines and a con-
tinuum above the ionization limit. The frequencies of the absorption
lines are given by the combination rule:
liw = E; - Ei (E; > E i) (3.1)
where Ei and E; are energy eigenvalues. The strength of an absorp-
tion line can be expressed in terms of matrix elements of the inter-
action between the atom and the radiation field. A typical matrix
element is defined by:
Mi; = f 1p;* V1pi dv (3.2)

where 1pi and 1p; are the wave functions of the initial and final states.
When the wavelength of the radiation is much greater than the
atomic radius, Mi; is proportional to the matrix element of the
dipole moment:
(3.3)

where e and m are the charge and mass of the electron.


If we imagine N similar atoms brought together to form a crystal,
each non-degenerate energy level of an atom will be spread into a
band of N levels. If N is large these levels can be treated as a con-
tinuum of energy states. The wave function of a system of N isolated
atoms is equal to the product of N atomic wave functions. When
41
42 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

the atoms are brought together in a crystal the wave functions will
be altered by the interactions between atoms. Both the width of
energy bands and the perturbation of the wave functions will be
least for electrons in the inner closed shells. The inert element atoms,
with all electrons in closed shells, have small interaction energies.
The absorption spectrum of an inert element solid may therefore be
expected to resemble that of the corresponding gas. On the other
hand, atoms with unsaturated outer electron shells have larger inter-
action energies, of the order of several electron volts. This is the
same order as the excitation energies of the outer electrons in iso-
lated atoms. The wave functions and energy eigenvalues of the outer
or valence electrons are quite different in the atomic and crystalline
states.
The wave functions and energies of electrons in a crystal can be
calculated by several approximate methods. Each electron is usually
assumed to move in a potential field of the nuclei and the average
charge distribution of the other electrons. This is the one-particle
approximation, which leads to a SchrOdinger equation for a single
particle wave function 'ljJ. The potential function V(r) in this equation
has the same periodicity as the crystal lattice. Bloch showed that the
wave functions must therefore have the form:
'ljJ(r) = exp (ik· r) u(k,r). (3.4)
u(k,r) has the periodicity of the lattice and modulates the plane wave
function exp (ik·r). The eigenvalues of energy E(k) have discon-
tinuities on certain planes in k-space which define the boundaries of
the Brillouin zones. It is often convenient to consider only the
first zone, enclosing the origin in k-space. The plane wave factor
exp Uk· r) can be reduced to the first zone by subtracting a suitable
vector of the reciprocal lattice, say K, from k. Exp (iK·r) is a
periodic function, so (3.4) can be rewritten:
'ljJ(r) = exp (ik·r) usCk,r) (3.5)
where k lies within the first zone and s is a band index labelling the
periodic function us(k,r). In this reduced zone scheme the energy
function Es(k) has a separate branch for each band.
The simplest approximation to 'ljJ(r) is obtained by taking VCr )=0,
which gives a constant u(k,r). 'ljJ(r) is the same as the wave function
of a free electron and the energy is (112m )1i 2k 2 where m is the electron
mass. This is in fact a reasonable approximation for many metals as
INTERBAND TRANSITIONS 43
shown, for instance, by its success in interpreting the topology of the
Fermi surfaces of polyvalent metals. Some calculated energy bands
for crystalline germanium are shown in fig. 3.1. To calculate the
optical absorption in the visible and near infra-red regions, we need
to know the form of the energy bands to an accuracy of at least
±O'l eV. Most theoretical calculations of energy bands have a

4
r 15
3

2 .AI
:;
~
l<J

-1 - _ _ _ _oolL3

-2

r L
1c<1II>
Figure 3.1 Energy bands of germanium in <111 >directions. r is tht: centre of
the zone and L is a point at the intersection of a <111 >axis with the zone boundary.
Reference: D. BRUST, Phys. Rev., 134, A1337-53 (1964).

maximum accuracy of about ± 1 eV. It is very difficult to improve on


this accuracy because many small terms of uncertain size, such as
relativistic corrections, need to be included in the calculation. A
semi-empirical approach has proved successful in the interpretation
of the energy bands of some crystals. Optical data can be used to
establish accurate values of energy band gaps at one or more points
in the zone. The theoretical calculations then provide an accurate
scheme of extrapolation to other points of the zone. This method
has been used very successfully for germanium and silicon.
The band theory approximation for the electron-lattice inter-
action, replacing it by an average potential V(r), is invalid in some
cases. An excited electron in an insulating crystal causes a local dis-
tortion of the charge distribution which can trap the electron in a
44 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

series of discrete energy levels. These are the exciton levels which can
lie in the forbidden gaps between energy bands. Also in a super-
conductor one finds electron wave functions and energies not in-
cluded in the band approximation.

3.2 Direct transitions


We will consider in this section the transitions caused by the inter-
action of an electromagnetic wave with electrons in Bloch states.
The interaction term in the Hamiltonian is:
Her = (e/mc)A·p = (eli/imc)A. V (3.6)
where A is the vector potential of the field and p is the momentum
operator of an electron. The scalar potential can be assumed to be
zero for the field of an electromagnetic wave, so the electric field E
is given by:
E = -(1/c)oA/ot). (3.7)
The real vector potential of a travelling wave of frequency wand
wave vector ko will be represented by:
A = Aon {exp i(k o. r - wt) + exp -i(ko. r - wt)}, (3.8)
where n is a unit vector in the direction of A. The corresponding
electric field is:
E = Eon{exp i(ko.r - wt) - exp -i(ko.r - wt)}, (3.9)
where Eo = (iw/c)A o. The mean square amplitude of Eo is 2E02.
The question arises, whether a local field correction to E is neces-
sary. Lorentz showed that charges localized on a cubic lattice
experience an effective field given by:
E' = E + (4n/3)P. (2.11)
On the other hand, for perfectly free electrons with a constant
probability density, E' = E. The Bloch wave functions are inter-
mediate between localized atomic wave functions and free electron
wave functions. This suggests that the effective field for Bloch
electrons can be written:
E'=E+(l.P (3.10)
where 0 < (I. < 4n/3. The value of (I. depends on the degree of
localization of the electrons. Highly non-localized electrons, which
are associated with a large dielectric constant, have a small value of (I..
INTERBAND TRANSITIONS 45
The subsequent calculations on interband transitions assume that
IX = 0 in view of the lack of exact values for IX. These calculations
can be corrected for a non-zero IX by a suitable redefinition of the
oscillator strengths and transition frequencies.
The probability Wet) of an electron making a transition from
state i to state j in a time interval t can be calculated by perturbation
theory. The result is:

W(t) J: J
= n- 2 1 dt' "P/ exp (iwjt')Her"Pi exp (-iWi t ') dv 12. (3.11)

"Pi and "Pj are eigenfunctions of energies Ei = nWi and E j = nWj;


dv is an element of volume. For sufficiently large t (3.11) increases
linearly with t and a constant transition rate is obtained:
dW
dt = (2ne 2n/m 2w 2)Eo2 ~(Ej - Ei - nw)

J
X 1 "P/ exp (iko·r)n. V"Pi dv 12
(3.12)

The delta function has been written for the case when E j > Ei and
the transition is accompanied by the absorption of a photon. The
transition rate is non-zero only for photons which satisfy the con-
servation of energy:
E j - Ei = nw. (3.13)
When E j < E i , an expression similar to (3.12) gives the rate for
transitions accompanied by stimulated emission. The rate of absorp-
tion of energy by an electron making a transition from state i to
statej (E j > E i ) is:
nw(dW/dt) = 0"1(w)(E2) (3.14)
where 0"1(w) is the optical conductivity. Hence:
O"l(W) = (ne 2n2/m 2w) ~(Ej - Ei - nw)
x 1 J
"Pj* exp (iko·r)n. V"Pi dv 12.
(3.15)

By substituting Bloch functions for "Pi and "Pj in (3.15) we can


calculate the optical conductivity associated with one-electron
transitions in the band model. The reduced zone scheme will be used:
"Pi = exp (ik . r) us(k,r); (3.16)
"Pj = exp (ik' .r) us,(k',r);
46 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

where sand s' are band indices. The matrix element in (3.15):
ftp/ exp (ik o·r)\1tpi dv

= f exp {i(k o - k').r} us.*(k',r) \1{exp (ik.r) u.(k,r)} dv.


(3.17)

The integral in (3.17) is zero because of the periodicity of the in-


tegrand, unless:
ko - k' + k = O. (3.18)
For photons in the optical region ko .;;;: 10 6 cm-I, while the maxi-
mum k in a Brillouin zone is about 10 8 cm- I • The selection rule (3.18)
corresponds to approximately vertical transitions between bands s
and s' in the reduced zone. The selection rule for exactly vertical
transitions is:
k=k'. (3.19)
The symbol Mij will be used to represent the matrix elements (3.17)
when ko = 0 and k = k':
Mij = f
Us *(k,r) \1u..(k,r) dv. (3.20)

The formula (3.15) for the optical conductivity can be written:


a 1 (w) = (ne 2n 2/m 2w) ~(Ej - Ei - nw) 1n.MiJ 12. (3.21)
This expression for aI(w) associated with direct transitions of an
electron between states i and j will now be compared with the
corresponding expression for a classical electron oscillator. Formula
(1.63) gives the optical conductivity associated with a classical
oscillator of resonant frequency Wo and damping constant y. For a
single electron, (1.63) can be written:
aI(w) = (e2Q/mwo)j{1 + Q2(WO/W - W/WO)2} (3.22)
where Q = wo/y. Let us write (3.22) in the form:
a1(w) = (ne 2n/2m)L(w,wo,Q) (3.23)
where L(w,wo,Q) represents the Lorentzian peak. In the limit, as the
damping tends to zero, y --+ 0 and Q --+ 00, the peak becomes a delta
function:
(3.24)
where Eo = nwo. The classical formula (3.23) has a similar form to
the quantum expression (3.21) if we equate Eo and (Ej - Ei)'
Formula (3.21) can be written:
a 1(w) = j;,j(ne 2n/2m) ~(Ej - Ei - nw), (3.25)
INTERBAND TRANSITIONS 47
where /;.i is called the oscillator strength of the transition; note that
/;.i= 1 for a single classical electron oscillator. From (3.21) and
(3.25) we obtain:

(3.26)

(Ei - E i ) has been written instead of liw in the denominator so that


the definition (3.26) can also be used when Ei < E i ; in that case Iii
is negative, corresponding to stimulated emission.
A sum rule can be derived for the /;i by using the boundary con-
ditions for Bloch functions:
m 02
'1/;i(k,s,s') +h2 ok;' E.(k) = I (3.27)

assuming that /;i refers to radiation with E parallel to the x-axis.


Equation (3.27) can be summed over all values of k in band s. The
sum of (0 210kllJ2) E.(k) is zero, so we obtain:
~ "£/;i(k,s,s') = N. (3.28)
k .'
where N. is the number of states in band s. The summations in (3.27)
and (3.28) refer to all bands both above s (s' > s) and below s
(s' < s). Formula (3.28) can be used to discuss the effective electron
number per atom n* associated with transitions from a filled band.
n* is defined by an expression similar to (1.60):
f
n* = (2mlne 2N o) O'l(W') dw' (3.29)

where No is the number of atoms per cm3 • The integral is extended


over the frequency range in which transitions from the filled band
have appreciable oscillator strength. It may be necessary to subtract
from the observed O'l(W) contributions due to transitions from other
bands. Let us consider the case when band s is the highest filled band
in the crystal, for example the valence band in germanium. Substi-
tution of (3.25) in (3.29) gives:
n* = (1INo) ~ "£/;.i (s' > s); (3.30)
k .'
the sum is taken over all upward transitions from band s. Since the
bands below s give a negative contribution to the sum in (3.28), it
follows from (3.30) that:
(3.31)
48 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The difference between n* and (Ns/No) will be large when there is


another energy band just below the valence band with appreciable
matrix elements Mi;. The graph in fig. 1.5 shows that n*, for the
valence band of germanium, considerably exceeds (Ns/No) = 4·00.
This can be explained by the influence of the d-band electrons. In
silicon, which has no d-electrons, the difference between n* and
(Ns/No) is much smaller. The application of the sum rule (3.27) to
metals is considered in Section 4.2.

3.3 Critical points


The optical conductivity given by (3.21) is that associated with up-
ward transitions of a single electron. The probability that a state of
energy E contains an electron at temperature T is given by the
Fermi-Dirac function F(E,T). If F(Ez,T) < 1 and F(Ej,T) > 0,
there will be a finite rate of stimulated emission and (3.21) must be
corrected by a factor {F(E.,T) - F(E;,T)}. In most cases of interest,
F(Ei,T) ~ 1 and F(E;,T) ~ 0, and these values will be assumed in
the further developments. To obtain the total optical conductivity,
(3.21) must be summed over all electrons which satisfy the energy
conservation law, E; - E. = liw. The density of Bloch states in
k-space is (1 /8x 3) and each state can contain two electrons of opposite
spins. The formula for the total optical conductivity associated with
direct interband transitions is:

O'l(W) = ( xe
21i2)(4x1 )f dkz dkll dkz 1n.Mij 12
m 2w 3
(3.32)
X c5{E,.(k) - E.(k) - liw},
where the integration is taken over the Brillouin zone. The volume
integral in (3.32) can be transformed into a surface integral in k-space
by using the properties of the delta function. The points in k-space at
which the delta function of (3.32) is non-zero lie on the surface:
Es:(k) - Es(k) - liw = O. (3.33)
The integral in (3.32) is of the form:

fdVk U c5(V). (3.34)

Let dS be an element of area of the surface V = 0 in k-space and dt


an element of length along the normal to dS; then (3.34) can be put
INTERBAND TRANSITIONS 49
in the form:

f dS dt U c5(V) = f dS dVI V\VI-l U c5(V)


(3.35)
= fdSUI V'kVI v':o.
Similarly (3.32) can be put in the form:

<il(W) = (::~2)(4:a) f dSI V'k{1s~~7~ ~.(k)} I


where the integration is taken over the surface defined by (3.33). The
joint density of states function for the bands sand s' is defined by:

Nss{w) = (4: )f V'k{Es{k~S_ Elk)}


a l 1 (3.36)
The integrand in (3.36) has singularities at certain critical points
where the denominator is zero; these are points where:
V' k{E.(k)} = V' k{Es,(k)}. (3.37)
Near to a critical point, the energy difference E(k) = {Es{k) - Elk)},
can be expanded in a Taylor series:
3
E(k) = Eo + ~ alki - k oi)2 + ... (3.38)
i=l
where the suffix 0 denotes a value at the critical point. The signs of
the coefficients ai are important in determining the behaviour of
Ns •.(w) near a critical point. If all the ai are positive, so that the
critical point is a minimum of E(k), then N.8.(w) increases like
(liw - EoY' for liw > Eo. Critical points are classified as M o, M 1 ,
M 2 , Ma where the suffix is equal to the number of ai which are
negative. MI and M2 correspond to saddle points of E(k) and Ma to
a maximum. The form of N S8 {w) in the region of critical points is
illustrated in fig. 3.2. The optical conductivity <i1(W) will show
critical point structure if the variation of 1n. Mij 12 with k is not too
rapid. The particular case when 1n. Mij 12 = 0 at a critical point
will be considered in Section 3.5.
Critical point analysis has been applied in detail to the ultra-violet
absorption spectra of several metals and semiconductors. The
optical absorption as a function of frequency can be represented by
curves of C2(W), <i1(W) or {W<il(w)}. This last quantity appears to be
the best choice if the matrix elements Mii are independent of w;
50 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

Figure 3.2 Joint density of states near critical points. Curves show the shapes of
discontinuities in N ..,(w) at critical points of types Mo. Mh M2 and Ma.

{WO't(W)} is then proportional to the joint density of states Ns.{w).


Calculations of Mil for germanium at various points in the zone
show that Mij varies typically by a factor of the order of two for a
given pair of bands, even when transitions are allowed for all values
of k. This variation in Mij makes it doubtful whether there is any
advantage in using curves of {wO't(w)} rather than C2(W) or O't(w),
for critical point analysis.
A graph of the optical conductivity of germanium at room
temperature is given in fig. 3.3. Ideally, for critical point analysis,
the optical properties should be measured at a temperature well
below the Debye temperature to reduce thermal broadening. Some
possible critical point identifications are indicated on the graph and
more details are given in Table 3.1. Most of the identifications would
be impossible without a fairly accurate band structure of germanium
as a guide. Measurements of piezo- and electro-reflectance provide
additional methods of identifying critical points. Some fine details of
the absorption spectrum associated with spin-orbit coupling, are
not shown in fig. 3.3. A splitting of 0·3 eV in the r~5 state causes a
INTERBAND TRANSITlONS 51

40

30
.,
E
.:!-
t>
'-
6" 20

10

1
+
o 6
nw (eV)
Figure 3.3 Optical conductivity of germanium at room temperature. Experi-
mental curve of optical conductivity IJ 1 against photon energy hw. Identification
of critical points: 1 - Mo, 2 - Mo, 3 - M" 4 - Mo, 5 - M" 6 - M 2 (see Table 3.1).
Reference: H. R. PHILIPP and E. A. TAFT, Phys. Rev., 113, 1002-5 (1959).

TABLE 3.1: CRITICAL POINTS OF GERMANIUM

Energies
Type Transition
Experimental Theoretical

1. r~5 -+r~ 0·8 eV 0'6eV


2. L~ -+ L1 2·1 1·8
3. .,13 -+.,11 2·3 2·0
4. r~5 -+ r I5 3·2 3·6
5. X4 -+ Xl 4·2 3-6
6. E4 -+E1 4'4 3-8

Reference: D. BRUST, Phys. Rev., 134, A1337-53 (1964).


52 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

subsidiary absorption edge at 1·1 eV, and a splitting of 0·2 eV in the


L; state leads to a double peak near 2·1 eV.
The direct transitions described in previous paragraphs occur
between one-electron energy eigenstates in the bands s and Sf. In an
actual crystal the energies of these states are broadened by electron-
phonon and electron-electron scattering. Each excited state of the
crystal has a finite lifetime T and hence an energy width lilT. If the
decay from the excited state is exponential, the effect of relaxation
can be included in the theory by replacing the delta function of (3.32)
by a Lorentzian function:

O'l(W) = (::~2)(4:3) f 1 12
dkx dky dk. n.Mii L(w,wo,Q) (3.39)

where liwo = Es{k) - Elk) and Q = wolY = tWoT. Each transition


should be assigned an appropriate value of Q in (3.39). Owing to the
complication of the theory for calculating Q, only rough estimates
are available at present. The theory of direct transitions between
energy bands is valid so long as lilT is much less than the widths of
the energy bands. If lilT is comparable with the band width the
selection rule for k loses its significance.
The transitions caused by the electron-phonon interaction are
not limited to nearly vertical transitions in the reduced zone because
the phonon has a range of k comparable with the electron. An
electron-phonon transition can be associated with either the emission
or absorption of phonons. The absorption processes disappear at
low temperatures because of the absence of phonons while the
emission processes tend to a constant level. The effect of phonon
interaction on an interband transition in aluminium is shown in
fig. 3.4. A pair of bands in aluminium run approximately parallel
over a considerable part of the zone, resulting in a narrow peak of
large oscillator strength in the optical conductivity. An increase of
temperature causes a broadening of the peak by phonon inter-
action and a shift of the mean energy.
The theory of direct transitions also ignores several aspects of
electron--electron interaction, including the interaction between an
electron in band Sf and a hole in band s produced by absorption of a
photon. This electron-hole interaction is responsible for the exciton
effects which will be considered in Chapter 6. The excited states are
also broadened by electron-electron scattering. In a metal, an
INTERBAND TRANSITIONS 53

• 97°K

1·0 2·0
hw (eV)
Figure 3.4 Absorption band in aluminium. Optical conductivity 0'1; photon
energy nO). Experimental curves from previously unpublished data.

electron with energy E above the Fermi energy ~F is subject to


electron--electron scattering which increases initially as (E - EF)2.
At higher energies of excitation, an electron can cause an interband
transition of another electron, leading to further broadening of the
energy levels.

3.4 Absorption band edges


Each crystal has a characteristic minimum photon energy for direct
interband transitions. For an insulating crystal or a semiconducting
crystal with a small number of free carriers, the direct interband
absorption is associated with electron transitions from a filled band
to an empty band. The minimum energy for such transitions is
located at a critical point of type Mo. If the matrix element (n. M ii )
is non-zero at the critical point, the transition is allowed and the
optical conductivity rises initially as:
(Tl(W) = A(liw - Eo)!. (3.40)
If (n.Mij) = 0 at a critical point, the transition is forbidden at this
point in the zone. In the neighbourhood of a critical point the
54 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

matrix element can be expanded in a Taylor series:


n.Mij = (n.Mij)k=o + k. 'Yk(n.Mij) + ... (3.41)
k is the wave vector measured from the critical point. For a forbidden
transition and small k, (3.41) becomes approximately:
(n.Mij) = k. 'Yin. Mij). (3.42)
Substitution of (3.42) into the formula (3.32) for the optical con-
ductivity introduces a factor k 2 , which is proportional to (liw - Eo).
The initial frequency-dependence of O'l(W) near a forbidden absorp-
tion edge is therefore:
O'l(W) = B(liw - Eo)~. (3.43)
The slope of O'l(W) is zero at liw = Eo for a forbidden absorption
edge. This simple theory of allowed and forbidden absorption edges
has neglected the possible influence of exciton states. The modifica-
tions of the theory to allow for exciton effects will be considered in
Chapter 6.
The unit vector n represents the direction of the electric field E of
an electromagnetic wave. It has been assumed that n is parallel to a
principal axis of the crystal so that 0'1 is a principal value of the
tensor O'lii. The matrix elements Mij are functions of the electron
wave vector k and the initial and final band indices sand s', as given
by (3.20); n.Mii is the component of Mij in the direction of n. For
given k, sand s', one or more components of Mij may be zero
because of the symmetries of the initial and final wave functions.
The symmetries of the relevant wave functions can be determined by
group theory and hence the selection rules for (n.Mij) determined.
For cubic crystals (n.Mij) has the same value for any direction of n.
For a uniaxial crystal (n.Mij) can have different values according to
whether n is parallel or perpendicular to the c-axis. In particular,
(n.Mij) at an Mo critical point may be zero for n perpendicular to c
and non-zero for n parallel to c. The absorption edges for radiation
of the two polarizations will have the different forms (3.40) and
(3.43). This will cause an apparent shift of the absorption edge as
the polarization of the radiation is varied.
Another type of absorption edge, associated with direct transitions
from a filled band to states just above the Fermi level, has been
observed in metals and degenerate semiconductors. These absorption
edges reflect the discontinuity in the electron distribution at the
lNTERBAND TRANSlTIONS 55
Fermi level. The discontinuity in optical absorption will be most
marked for those metals with a narrow energy band containing a
large number of electrons, situated a few electron volts below the
Fermi level. The metals copper, silver and gold, with their narrow
d-electron bands, satisfy these criteria. The optical conductivity of
gold for photon energies of 0·5 eV to 20 eV is shown in fig. 3.5. The

10

.,
E
->
u
"-
tl 5

nw(eV)

Figure 3.5 Optical conductivity of gold. Experimental curves of optical con-


ductivity 0'1 against photon energy liw.
References: L. R. CANFIELD, G. HASS, and w. R. HUNTER, J. de Physique, 25,
124-9 (1964); J. N. HODGSON, J. Phys. Chern. Solids, 29, 2175-81 (1968).

curve below 1 eV represents absorption associated with intraband


transitions of conduction electrons. This absorption has dropped to
a low level before the threshold for interband transitions at about
2·0 eV. The rise in (jl(W) above the threshold depends on the inter-
section of the surfaces of constant interband energy, defined by:
Es,(k) - Es(k) = liw (3.44)
E
56 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

with the empty and full states in the conduction band. If the lower
band has negligible width, the surfaces (3.44) coincide with constant
energy surfaces in the conduction band. As E is increased, O'l(W)
rises abruptly when the surface (3.44) crosses the Fermi surface,
E•.(k) = EF • In an actual metal the absorption edge is spread out
because of the appreciable width of the lower band. The shapes of
the absorption edges in copper, silver and gold have been calculated
using the detailed knowledge of band structures and Fermi surfaces
which is now available for these metals.

3.5 Indirect transitions


Below the absorption edge of germanium at 0·80 eV, No. I in fig. 3.3,
there is a region of weak absorption extending to about 0·63 eV at
room temperature. This absorption can be explained by indirect
interband transitions in which the wave vector of an electron changes
by a considerable amount. Absorption curves for germanium in the
indirect transition region are shown in fig. 3.6. The shift with tempera-
ture is not peculiar to the indirect absorption but is also shown by
the direct absorption at higher energies. It is caused mainly by the
change in volume with temperature and the associated shifts in
energy bands. The change in shape of the absorption curves with
temperature, as shown in fig. 3.6, is a typical feature of the indirect
transition region. The general level of absorption coefficient in this
region is I to 10 cm-I, compared with 10 5 to 106 cm- 1 in the direct
transition region. This large difference means that indirect transitions
are generally observable only in regions of the spectrum where there
are no direct transitions. Indirect interband transitions in metals also
tend to be masked by intraband transitions of conduction electrons.
There has been a possible identification of indirect transitions in the
near infra-red absorption spectra of the alkali metals.
An indirect transition is produced by the interaction of a photon,
an electron and a lattice defect. The defect can be a lattice vibration,
an impurity centre, a structural imperfection or a bounding surface
of the crystal. In this section we will consider indirect interband
transitions associated with lattice vibrations. The absorption due to
these transitions can be calculated by perturbation theory. Terms
representing the electron-photon and electron-phonon interactions
are included in the Hamiltonian. Each transition involves two
INTERBAND TRANSITIONS 57

"j
E
~

o~--~--~------~~----~--~

hw{eV)
Figure 3.6 Indirect absorption in high purity germanium. Experimental curves
of absorption coefficient K (= 4nK/Ao) against photon energy Iiw.
Reference: T. P. MCLEAN, Progress in Semiconductors, 5, 54-102 (1960).

successive steps, corresponding to the two interactions, through an


intermediate state which does not have to satisfy the conservation of
energy. The theory of Section 3.2 shows that transitions caused by
the electron-photon interaction must be nearly vertical in the E-k
diagram. This is not the case for transitions caused by the electron-
phonon interaction because phonons have a range of k comparable
to that of electrons. The conservation laws for the complete transi-
tion are:
(3.45)
The plus sign refers to phonon emission and the minus to phonon
absorption; ko is the photon wave vector and can be neglected;
klJ and ElJ are the wave vector and energy of the phonon.
In germanium the conduction band state at the point L on the
zone boundary is about 0·15 eV below the state r~ at the centre of
the zone. The energy difference Eg between r~5 and Ll is therefore
less than the direct energy gap between r~5 and r~. Absorption
associated with indirect transitions starts at an energy of about
0·15 eV below the direct absorption edge. The theoretical formula
58 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

for the absorption coefficient associated with indirect transitions in


the region near liw = Eg is:
K = (A/liw) ~ Ev -1{(1 + ii) 1 XB(k.,o) 12 F(liw - Eg - Ev)
+ ii 1 XMk.,o) 12 F(liw - Eg + Ev)}. (3.46)
The sum is taken over all branches of the phonon spectrum; ii is the
occupation number of phonons as given by equation (2.3). The
matrix element Xi;(k.,O) is calculated for an initial state r~5 at
k = 0 in the valence band and a final state Ll at k = k. in the con-
duction band. For a forbidden transition Xi;(k.,O) is replaced by
k. VkXij as in the theory of direct transitions. If the electron-hole
interaction is neglected, the function F(E) = 0 for E < 0 and
F(E) = E' for E>- 0 where r = 2 for an allowed transition and
r = 3 for a forbidden transition. The two terms in (3.46) have
threshold energies given by:
liw = Eg ± Ev. (3.47)
The plus sign corresponds to the threshold with phonon emission
and the minus sign to the threshold with phonon absorption.
The inclusion of electron-hole interaction in the theory leads to
important changes in formula (3.46) when applied to a semiconduct-
ing or insulating crystal. The binding energy Eeg: of an exciton in
germanium, with the electron at Ll and the hole at r~5' is about
0·003 eV. The energy thresholds (3.47) are lowered by this amount
and the form of F(E) is altered. For an allowed transition F(E)
increases as El for 0 < E < Eez and as E~ for E ~ Eez. Referring
to the curve in fig. 3.5 for germanium at 4° K, two thresholds at
0·749 and 0·769 eV can be distinguished. Since at 4° K, n ~ 0, these
thresholds must be associated with phonon emission. Further
analysis shows that (Eg + Eeg:) = 0·741 eV and Ev = 0·008 or
0·028 eV; or in terms of temperature Ev/k = 90° K or 320° K. These
values of Ep agree well with the energies of TA and LA phonons at
k = k •. A careful analysis of the 195° K curve has shown the presence
of components associated with the absorption of TA and LA
phonons as well as components with two more phonon energies.
Some solids, for example the silver halides, anthracene and tri-
gonal selenium, perpendicular to the c-axis, have absorption edges of
exponential form. The absorption near the edge is given by Urbach's
rule:
K = Ko exp {IX(E - Eo}/kT}. (3.48)
INTERBAND TRANSITIONS 59
The parameter ~ determines the steepness of the edge and may be
weakly temperature-dependent; the energy gap parameter Eo is also
temperature-dependent. The dependence of absorption coefficient K
on temperature T suggests that phonon-assisted transitions are in-
volved in the absorption. Several theories have been suggested to
explain Urbach's rule but none has been generally accepted to date.

3.6 Infra-red absorption in superconductors


Although the phenomena to be described in this section are not due
to interband transitions, it was felt that because of the analogies with
interband absorption, this would be an appropriate place to discuss
the infra-red absorption of metals in the superconducting state.
Early measurements of the infra-red reflectance of metals at low
temperatures showed no change at the transition from the normal to
superconducting state. The change in the optical conductivity O"l(W)
at infra-red frequencies is therefore inappreciable at the super-
conducting transition while the zero frequency conductivity 0"1(0)
has a first order discontinuity. It is clear that a change between these
two extremes of behaviour must occur in some intermediate fre-
quency range. Later experiments have confirmed this and shown that
the change occurs in a frequency range which depends on the transi-
tion temperature Te. The mean frequency w of the change in be-
haviour at temperatures well below T e, is approximately (5kTe/Ii), or
in terms of photon energy and wave number:
liw ~ 0-43Te (meV); ko = w/2nc ~ 3·5Te (cm- I ) (3.49)
For metals with low transition temperature, for example aluminium
(Te = 1·18° K), w lies just within the upper end of the microwave
region. For superconductors with higher values of T e, W lies in the
far infra-red region.
Experiments on far infra-red and microwave absorption in super-
conductors have provided useful tests of the theory of supercon-
ductivity put forward by Bardeen, Cooper and Schrieffer in 1957.
The basic idea of the BCS theory is that the electrons form bound
pairs in the superconducting state owing to the electron-phonon
interaction. A minimum energy Ey(T) is needed to break a bound
pair and transfer the electrons to the normal state. The absorption
spectrum of a superconductor should therefore have a threshold at
60 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

liw = Eg(T). The BCS theory predicts that Eg(T) is 3'5kTc at T = 0


and decreases to zero as T increases to Te.
Recent experiments on superconductors have been designed to
measure the absorptance A or reflectance R (= 1 - A) of bulk
specimens or optically thick films. Other experiments have measured
the transmission of radiation through very thin films. The complex
surface admittance Y(w) can be calculated if A(w) is known for a
sufficiently wide range offrequencies. Alternatively, the experimental

10 -

z
~ 05-
U>
<:!

Or-.----'

25 3·0 35
nw(meV)

Figure 3.7 Absorptance of lead at 1'08° Kt Absorptance in superconducting


state As, absorptance in normal stage AN, photon energy liw in 0·001 eV (8'0 cm-l).
Reference: s. L. NORMAN, Phys. Rev., 167, 393-407 (1968).

curves for A(w) can be compared directly with theoretical predic-


tions. The results of a typical experiment on the absorptance of an
optically thick film of lead are shown in fig. 3.7. The film was
alternated between superconducting and normal states by applying
a strong magnetic field. The ratio of absorptances As(w)jAN(W) is
zero below the threshold energy of2·64 meV (4·28kTe). The tempera-
ture of the lead film was sufficiently low so that EiT) ~ Eg(O) and
the threshold should be at 3·5kTe according to the BCS theory. The
discrepancy in the coefficients of kTe has been explained by recent
refinements of the BCS theory. The experimental curve also shows a
second threshold at 2·86 meV. The existence of two thresholds can
INTERBAND TRANSITIONS 61
be explained by anisotropy of the energy gap and the orientations of
the crystallites in the lead film. The conditions for an isotropic
energy gap, as given by the original BCS theory, are satisfied in
superconductors with a small relaxation time in the normal state.
The energy gap is isotropic when:
Wg''C ~ 1 (3.50)
where liw y = Ey and 7: is the relaxation time of conduction electrons
in the normal state. The detailed calculation of A.(w) and AN(W) is
complicated because the absorption occurs in the region of the
anomalous skin effect. Methods of calculating the optical properties
in this region will be discussed in Chapter 4.
4
Free Carrier Absorption
4.1 Classical theory
The interband transitions discussed in the previous chapter require a
certain minimum energy corresponding to the separation between
energy bands. If all the electrons in a solid are in filled bands, the
solid cannot absorb lower energy photons by electron transitions.
On the other hand, if one or more bands are only partially occupied
by electrons then transitions within these bands, intraband transi-
tions, can cause absorption of radiation down to zero frequency.
These transitions are indirect because a change of energy within a
single band must be associated with a change in crystal momentum.
The quantum theory of intraband transitions will be considered in
the next section. In this section some general consequences of
classical macroscopic theory are considered.
The real optical conductivity a 1(w) approaches the electrical con-
ductivity a 1(0) as the frequency tends to zero. The electrons in
partially occupied bands are responsible for al(w) at frequencies
below the interband threshold and for a 1(0). It will be useful for the
development of this section to consider the complex conductivity
a(w):
(4.1)
In terms of the current density vector J(w), assumed parallel to a
principal crystal axis:
J(w) = a(w)Eoexp(-iwt). (4.2)
Introducing J(t) to represent the current density produced by a unit
delta function pulse of electric field at t = 0, we can write:

a(w) = J:J(t') exp (iwt') dt'. (4.3)

The derivation of (4.3) is analogous to the derivation of equation


(1.42) for {feW) - I}. A delta function pulse of electric field gives
62
FREE CARRIER ABSORPTION 63
each electron an initial velocity (-elm) and hence the initial current
density is:
(4.4)
where N is the number of electrons per cm 3 •
Note that any inter-
actions can be ignored in calculating J o since a delta function pulse
exerts an infinite force on each electron. The current subsequently
decays due to interactions between the conduction electrons and the
electrons and nuclei forming the crystal lattice. The response of the

o ',0
t (sec)
Figure 4.1 Current density J(t) in aluminium due to a delta function pulse of
electric field at t = O. Logarithmic scale for J(t); straight parts of curve represent
exponential decays. Electron mass m, optical effective mass m*, interband
energy ,...., li/tB, phonon energy""" li/tD.

lattice to the displacement of an electron takes place in two stages,


the first associated with the motion of other electrons and the second
with the motion of nuclei. Fig. 4.1 shows a schematic graph of J(t),
on a logarithmic scale, which corresponds roughly to the measured
a(w) of aluminium at room temperature. The initial fall in J(t) due
to electron-electron interaction is shown as an exponential decay
with a relaxation time of about 10-16 sec. The oscillatory part of
J(t) which follows, arises from the energy gap of about 1·6 eV in the
excitation spectrum of conduction electrons in aluminium. The
decay of J(t) between 0-4 and 2·0 X 10- 14 sec corresponds to
64 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

electrons of effective mass m* ~ 1·5m being slowed down by inter-


action with lattice vibrations. The limit 2 X 10-14 sec is the reciprocal
ora typical lattice vibration frequency and represents the time taken
for the lattice to reach a steady state response to the conduction
electrons. After 2 x 10-14 sec we can consider J(t) to consist of
quasi-particles formed by electrons and co-moving clouds of elec-
tronic polarization and lattice displacements. J(t) gradually falls
to zero as the quasi-particles are slowed down by collisions with
phonons and lattice imperfections.
The representation of the non-oscillatory parts of J(t) by ex-
ponential decays can only be justified by an analysis of the relaxation
processes in each case. It is a simplifying assumption, to be revised
in the light of further experimental data and theoretical calculations.
If J(t) has the form:
J(t) = J o exp (-yt) (4.5)
for t = 0 to 00, then:
a(w) = Jo/(Y - iw). (4.6)
If J(t) has the form shown in fig. 4.1, formula (4.6) is approximately
valid for the frequency ranges w ~ wB, wB ~ W > Wn and w < Wn.
The appropriate value of y in each case is given by the slope of the
J(t) curve for t ~ tB, tB ~ t < tn and t> tn. In the middle range
of frequency:
a(w) = (Ne 2/m*)(y - iW)-l. (4.7)
The quantity y is the reciprocal of the relaxation time 7: of the current
density; we will call y the 'collision frequency' of the conduction
electrons. Ignoring for the moment any change in y near t = tB,
formula (4.7) can be extended to w = 0:
alO) = (Ne 2/m*)y-1 = (Ne 2/m*)7:. (4.8)
Taking real and imaginary parts of (4.7), we have:
a 1(w) = (Ne 2/m*)y(y2 + ( 2)-1 = a 1 (0)(1 + W27: 2)-1; (4.9)
a 2(w) = -(Ne 2/m*)w(y2 + ( 2)-I. (4.10)
alw) is related to C1(W), the real part of the dielectric constant by
{CI(W) - I} = (4n/w)a 2(w) = -(4nNe 2/m*)(y2 + ( 2)-1. (4.11)
These formulae were first derived by Drude for a classical model of a
free electron gas. Two calculated curves of <11(W) for gold are shown
FREE CARRIER ABSORPTION 65

o 005 010
hw(eV)

Figure 4.2 Drude curves of optical conductivity for gold. Calculated from
formula (4.9) for a1(w) in e.s.u.

in fig. 4.2. The area under each curve is (nNe 2j2m*c) which varies
slightly with temperature owing to thermal expansion. The fre-
quency variations of (fl(W) and {I - Cl(W)} predicted by (4.9) and
(4.11) have the same form. At high frequencies, w}> y, (4.9) and
(4.11) reduce to:
(fl(W) ~ (Ne 2jm*)yw- 2 = (lj4n)wp2yw-2; (4.12)
(4.13)
where wp2 = (4nNe 2jm*). For most metals and semiconductors at
room temperature and below, (y j2nc) <; 10 3 cm -1, so the condition
W }> Y is satisfied in the near infra-red. Note that the first correction
terms to (4.12) and (4.13) are ofthe order of (y2jw 2). Equation (4.13)
for Bl(W) is the same as if there were no collisions. The optical
conductivity (fl(W) given by (4.12) is proportional to the collision
frequency y, whereas the electrical conductivity (fl(O) is inversely
proportional to y.

4.2 Intraband transitions


The optical effective mass m* introduced in the previous section is
an average over all the conduction electrons. The band theory of
66 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

crystals gives us an expression for the inverse effective mass tensor of


a single electron:
(4.14)

Let us suppose that the x-axis is a principal axis of the conductivity


tensor so that electric field and current are parallel in this direction.
For an electric field in the x-direction:

(;) = 4~3 JJJ (~*)xx dkxdky dkz


= 4n~/i2 JJJ (:;~) dk dky dk x z• (4.15)
The non-diagonal terms integrate to zero. When the occupied states
are separated from the empty states by a well-defined Fermi surface,
as in a metal or degenerate semiconductor, (4.15) can be trans-
formed into an integral over the Fermi surface:

JJJ(:;~) dkx dk dkz JJ (:~) dky dk


=

= ~ J (vaNv) dS. (4.16)


Therefore:
(m/m*) = (vll,2/v)S/voSo. (4.17)
v is the velocity of an electron on the Fermi surface; (v a,2/v) is the
average of VIJ)2/V over the Fermi surface; S is the area of the Fermi
surface. Vo and So refer to the same quantities calculated for a free
electron gas. For a cubic crystal (v,,2/v) = (v1l 2 /v) = (v z2 /v) and
m* is independent of the direction of the electric field.
The sum rule for electrons with Bloch wave functions is:
m 02E.(k) ,
/i 2 OklJ)2 = 1 - l'fi;(k,s,s). (4.18)

The fi; are interband oscillator strengths for the electric field parallel
to the x-axis. Summing over N conduction electrons per cms in
band s, we have:
(m/m*) = 1 - N-l ~ ~fi;(k,s,s'). (4.19)
Ie 8'

It follows from (4.19) that m* can be greater or less than m depending


on the distribution of energy bands and the interband matrix ele-
FREE CARRIER ABSORPTION 67
ments. A band s' just below s and connected to it with large matrix
elements gives a large negative contribution to the sum in (4.19).
This corresponds to a small m*, as commonly observed in semi-
conductors.
If there were no interaction between conduction electrons and
lattice vibrations, J(t) would maintain a constant value (Ne 2 /m*).
The optical conductivity corresponding to a constant J(t) is:
(11(£0) = (ne 2JiN/m*) ~(Ji£O). (4.20)
This formula is similar to the delta function approximation for inter-
band transitions given in Section 3.2. When the electron-phonon
interaction is included, the delta function becomes a broadened peak
with a maximum at £0 = O. The value at £0 = 0 is just the electrical
conductivity (11(0).
The calculation of electrical conductivity has received a great deal
of attention from solid state theorists. The basic problem is the
calculation of the scattering of conduction electrons by phonons,
lattice defects or impurities. The matrix element for electron-phonon
scattering is:
M(k,k') = (n I f
1p*(k,r)Hsp1p(k',r) dr I n - I). (4.21)

n is the initial number of phonons and (n - 1) the final number;


Hsp is the Hamiltonian of the electron-phonon interaction. An
electron with wave vector k absorbs a phonon and is scattered to a
state of wave vector k'. The conservation laws for this transition are:
k'-k=q+K; (4.22)
E(k') ~ E(k) = Ep; (4.23)
q is the wave vector of the phonon; K is zero or a vector of the
reciprocal lattice ; Ep is the energy ofthe phonon. The scattering must
also be compatible with the exclusion principle. A detailed evalua-
tion of (4.21) must take account of energy-momentum relations of
electrons and phonons, that is E(k) and Ep(q). We will assume for
simplicity that the transition probability can be represented by an
average value for conduction electrons. Let us define an electron-
phonon collision frequency yp by:
oJ/at = -ypJ (4.24)
where (-oJ/at) is the rate of decrease of current owing to electron-
phonon transitions. From formula (4.21) one finds a transition
68 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

probability proportional to the number of phonons. The consequent


variation of y p with temperature is shown by the lower curve of
fig. 4.3. A Debye model for the lattice vibrations with characteristic
temperature e was assumed in calculating this curve. At high

TIS
Figure 4.3 Temperature variation of electron-phonon collision frequency in
metals. r = (Yp/g) where yp is the collision frequency and yp -+ (gT/f9) for
T ~ f9; Debye. temperature f9. Curve A for low frequencies (liw ~ k(9) and
curve B for high frequencies (liw ~ k(9). Typical values of kf9: 0,01 to 0·04 eV.
Reference: A. I. GOLOVASHKIN and G. P. MOTULEVICH, Sov. Phys. JETP, 20,
44-9 (1965).

temperatures the number of phonons is proportional to temperature


and:
yp = (gTje). (4.25)
The ratio r = (Ypjg) is plotted in fig. 4.3. The collision frequency
includes a contribution from phonon emission processes. The
energy available from electron transitions within the thermal layer
around the Fermi surface is about kT which is enough for the crea-
tion of phonons with energies less than kT. When we consider the
effect of electron-phonon scattering on the optical conductivity it is
necessary to distinguish two frequency regions: liw ~ ke and
FREE CARRIER ABSORPTION 69
liw ?> k@; k@ is a typical phonon energy. In the low-frequency
region the calculations of yp for a steady electric field should be
applicable. In the high-frequency region, at low temperatures, the
photon gives an absorbing electron sufficient energy to create
phonons throughout the phonon spectrum. The collision frequency
at low temperatures therefore tends to a constant value associated
with phonon emission processes. This is illustrated by the upper
curve in fig. 4.3 which shows that the limiting value of yp at low
temperatures is 0·4g.
Equations similar to (4.21) can be written for scattering of elec-
trons by lattice defects and impurities. Using the same approximation
of representing the scattering by a collision frequency, and assuming
that the scattering processes are independent, we can write the total
collision frequency:
(4.26)
where Yd and Yi are the collision frequencies of electrons with defects
and impurities. Y d and Yi vary only slightly with temperature so the
main part of the temperature variation of Y is caused by yp.

4.3 Electron transport


When the electron current density is given by:
I(w) = a(w)E(w) (4.27)
Maxwell's equations have solutions representing damped plane
waves:
E(w) = Eo exp (-Kko.r) exp (inko.r - iwt) (4.28)
where ko = w/c. The complex refractive index (n + iK) is related to
a(w) by:
(n + iK)2 = e(w) = 1 + 4niw- 1 a(w). (4.29)
Equation (4.27) is a local relation referring to the current and field ~t
the same point in the medium. According to (4.28) the field ampli-
tude varies by a factor e in a distance:
15 = (KkO)-l = Ao/2nK. (4.30)
When the mean free path t of the conduction electrons is more than
a small fraction of the skin depth 15, a local relation between I and E
is no longer valid. I at a given point depends on E within a region of
70 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

approximate radius t, the mean free path. This is the origin of what
is known as 'the anomalous skin effect'. The natural generalization
of (4.27) to deal with spatially varying fields is to take a Fourier
spectrum with respect to the space co-ordinates:
J(q,w) = a(q,w) E(q,w) (4.31)
where E(q,w) = Eo(q,w) exp (iq.r - iwt). (4.32)
q has been used to denote the electromagnetic wave vector because
k will be used for the electron wave vector. The strongest part of the
Fourier spectrum is near wave number q = 15- 1 •
To obtain a(q,w) from a microscopic model, one requires the
simultaneous solution of the electromagnetic equations and the
electron transport equation, subject to appropriate boundary con-
ditions at the surface of the medium. Most theoretical research on
electron transport in metals has been based on the Boltzmann
equation, treating conduction electrons as classical particles but also
satisfying the exclusion principle. This can be justified by quantum
mechanics if electron wave packets of well-defined position and
velocity can be constructed from the electron wave functions. The
region of validity of classical transport theory can be expressed in
terms of the collision frequency 'Y and the Fermi energy EF :
(4.33)
This inequality is well satisfied by most solid pure metals and alloys
but only marginally satisfied by many liquid metals.
The Boltzmann distribution function F(k,r,t) expresses the frac-
tion of electron states occupied as a function of electron wave vector,
position and time. The equilibrium function will be denoted by Fo
and the change caused by an applied electric field will be calledf To
simplify calculations and prepare the way for the theory of surface
admittance in the next section, we will consider a conducting medium
filling the half-space above the plane z = 0; the field in the medium
is caused by a normally incident wave with:
Ex = Eox exp {+ikoZ - iwt}; Ey = E z = O. (4.34)
With this geometry the Boltzmann equation reduces to:
v.(ojliJz) + (of/at) - (oj/ot)c = (eEx/Ii)(oF%kx). (4.35)
Only the first order terms injand Ex have been retained since we are
considering the linear response. The term (oj/ot)c represents the
FREE CARRIER ABSORPTION 71
effects of collisions; the simplest assumption about the form of this
term is:
(4.36)
This means that the return of F to its equilibrium value after a dis-
turbance is an exponential decay with relaxation time T. We will not
assume at this stage that l' has the same value for all electrons; a
possible. dependence of l' on k will be retained in the transport
equation. If we are considering the response of electrons to a Fourier
component of field like (4.32), then (4.35) can be written
(iqvz - iw + y)f = (eE,,/n)(oF%kaJ (4.37)
The current density in the x-direction is given by:

Jf1) = 4n~ fffvJ dkf1) dkll dkz• (4.38)

Substituting for f from (4.37) and assuming that Fo is the Fermi

f
distribution for a degenerate electron gas, we have:
e2Ef1) (Vf1)2/V2)t dS
Jf1) = (]f1)iq,w)Ef1) = 4 31;; 1 + .( Ii 1 )' (4.39)
n " l q VzV- - WT

where t = VT and the integral is over the Fermi surface. Since the
significant values of q are of the order of d-1 , it follows thatqt '" tid.
At sufficiently low frequencies we have t ~ d and W ~ y; (4.39)
then reduces to an expression for the electrical conductivity:

(]f1)f1)(0) = 4::n f
(Vf1)2V-2)t dS. (4.40)

If l' is constant over the Fermi surface, (4.40) can be put in the
familiar form:
(4.41)
with m* defined by (4.15) and (4.16). As the frequency is increased,
the skin depth d decreases and in pure metals at low temperature one
can have t ~ d while WT < 1. This is the extreme anomalous limit
which will not be discussed further because the relevant frequencies
lie mainly below the optical region. The bracketed terms in the
denominator of (4.39) represent the spatial and temporal dispersion
of (]f1)f1). The ratio of temporal to spatial dispersion is approximately
(d/~) where ~ = v/w; the maximum value for vz, that is v, has been
assumed. ~ represents the distance travelled by an electron while the
F
72 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

phase of the field changes by one radian. When 15 ~ the relation e


between J and E will be a local relation, as shown by the negligible
size of the spatial dispersion term in (4.39). When 15 ~ (4.39) e,
reduces to:

(4.42)

'Yith the assumption of a constant T on the Fermi surface:


(J=(w) = (Ne 2/m*)(y - iW)-l (4.43)
in agreement with the classical Drude formula. The variation of
(l5/e) and (15/t) with frequency, for gold at 20° K and 293° K, is

100

10
AO(jLm)

Figure 4.4 Anomalous skin effect region in gold. Short wavelength limit (jj~ ~ I;
long wavelength limit (jjt ~ 1. Vacuum wavelength Ao, skin depth 13, mean free
path t, parameter ~ = vjw where v is the Fermi velocity.

shown in fig. 4.4. The region where spatial dispersion is significant,


that is· the anomalous skin effect region, lies between the limits
(15 Ie) ~ 1 and (15/t) ~ 1. Fig. 4.4 shows that for gold at room
temperature there is a limited region in the near infra-red where the
skin effect is weakly anomalous. At 20° K the anomalous region is
more extensive, reaching the short wavelength end of the microwave
spectrum.
FREE CARRIER ABSORPTION 73

4.4 Surface admittance


It was shown in Section 1.4 that reflection coefficients for electro-
magnetic waves can be calculated from the surface admittance:
Y = H t/ E t (4.44)
where H t and E t are the tangential components of electric and
magnetic fields at the surface. The simplest geometry for calculating
Y is that given in Section 4.3 with a normally incident wave of the
form (4.34). Y does not vary appreciably with angle of incidence
when I Y I ~ 1, which holds for most metals in the infra-red. With
the geometry of Section 4.3, Maxwell's equations for the fields in the
medium reduce to:
oE.';oz = (-iw/c)Hy ; oHy/oz = (4n/c)Jx - (iw/c)Ex • (4.45)
Integrating the second of these equations through the depth of the
medium, we have:

J00
o
(OH
a;y ) dz =
(4n)
--;
Jco ,
0 Ja; dz.

J~ includes the displacement current density. Hence:


Hy(O) = -(4n/c)I~ (4.46)
where I; is the total surface current. Equation (4.44) can therefore
be written:
Y = (4n/c)I~/ExCO). (4.47)
So far we have assumed that the current is parallel to the electric
field. To allow for cases when this is not so, Y can be redefined as a
two-dimensional tensor YiJ. If the x- and y-axes are the principal
axes of this tensor then (4.47) is the correct expression for Y IJJIJJ.
For frequency ranges in which spatial dispersion is negligible, we
can put:
(4.48)
except for points near the surface. In a layer below the surface of
depth (J at low frequencies or ~ at high frequencies, the current is
affected by scattering of electrons at the surface of the medium. The
surface scattering is important at low frequencies when the con-
ducting medium has one or more dimensions which do not contain
many mean free paths. For the semi-infinite medium assumed here,
74 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

surface scattering can be neglected at low frequencies and we can


take:
0'(00) ~ 0'1(0). (4.49)
The electric field in the medium is given by (4.28) with
(n + iK)2 ~ 4niw-10'1(0).
Th~refore the surface admittance:
Y = (n + iK) ~ (1 + i){2nW-10'1(0)}it. (4.50)
The low-frequency skin depth from (4.30):
~ = c{2nwO"I(0)}-it. (4.51)
>
Formula (4.51) confirms that ~ t at sufficiently low frequencies
since t is independent of w.
In the frequency range where spatial dispersion is important, the
current density J at a given point can be expressed as an integral of
the electric field E over a finite region surrounding the point. When
the integral for J is substituted in (4.45) and H'U is eliminated, one
obtains an integro-differential equation for Ea:. The observable
quantity Y can be expressed in terms of Ea: and (oEa:/oz) at the
surface:
Y = -(ic/w){(oEJ/oz)/Ea:}z=o. (4.52)
The integral for J must take account of electron scattering at the
surface of the medium. To simplify the calculations, two extreme
types of surface scattering have been assumed, completely diffuse
and completely specular. Diffuse scattering means that after scatter-
ing the electrons have an equilibrium distribution function Fo.
Specular scattering reverses the normal component of velocity and
momentum while conserving tangential components. The conditions
for specular scattering are easy to define for electrons on a spherical
Fermi surface but difficulties arise with non-spherical Fermi surfaces.
Momentum and velocity may not be parallel and some parts of the
surface may be cut off by zone boundaries. In order to allow for
scattering intermediate between diffuse and specular, one can intro-
duce parameters p and (1 - p) to represent the fractions of electrons
specularly and diffusely reflected. The equation for Ea: can be solved
by analysing Ea: and Ja: into spatial Fourier spectra and using the
non-local conductivity a(q,w) with appropriate allowance for surface
scattering effects.
FREE CARRIER ABSORPTION 75
At high frequencies when ~ ~ b, the spatial dispersion is again
negligible and the local conductivity a(w) is given by (4.42) except
within distances '"-'~ of the surface. In the range of frequencies
Y ~ w ~ W p , the skin depth b is approximately independent of fre-
quency. The ratio (wplY) lies in the range 10 2 to 103 for most pure
metals at room temperature and lower temperatures. It follows that
the range of frequencies defined above covers a considerable part of
the visible and near infra-red. The approximately constant value of b
is given by:
(4.53)
This follows from:
s ~ -W p2W- 2 ; K ~(-s)! ~(wplw). (4.54)
A typical value of b for a metal is 20 nm, about 100 atomic layers.
Under the condition ~ ~ b, the scattering of electrons at the sur-
face can be represented by a surface current i. A corresponding
surface admittance y( w) is defined by:
i = (cI4n)y(w)Et • (4.55)
In general, yew) should be a two-dimensional tensor yij(w). The total
surface admittance which determines the reflection of waves by the
surface is {Y( w) + y(w)}. Since spatial dispersion is negligible,
Yew) is given by:
Yew) = {sew)}! = (n + iK). (4.56)
The ratio I y III Y I can be estimated from the equivalent ratio
I i III I I, to be about Ub, which is much less than one. It must be
remarked, however, that the real part of Yew) is small in the fre-
quency range y ~ w ~ W p , so yew) can have a considerable effect
if it is mainly real. The real part of yew) corresponds to surface
scattering which dissipates energy. Thus diffuse scattering gives a
yew) with a non-zero real part, whereas specular scattering gives a
purely imaginary yew).
An expression for yew), when the surface scattering is completely
diffuse, can be derived by considering the dissipation of energy by
electron collisions with the surface. Let us consider the frequency
range y ~ w ~ W p , so that:
Yew) ~ !wpyw- 2 + i(wplw) (4.57)
from the Drude formulae (4.12) and (4.13). Adding a real yew) to
(4.57) is equivalent to increasing the collision frequency y. This is
76 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

reasonable since the surface scattering effectively reduces the mean


free path of electrons in the surface layer. The surface density of
electrons interacting with the field is of the order of N c5 and the
number of surface collisions per second is about Nv. The correction
to I' to allow for surface scattering is therefore of the order of
(v/c5) = wiv/c). A typical value of liwiv/c) is 0·02 eV, which is
comparable with iiI' for electron-phonon scattering at low tempera-
tures. A more exact calculation, taking account of the distribution
of velocity on the Fermi surface, confirms that the effect of yew) is
equivalent to the addition of a frequency-independent term to 1'.
The formula for Yx.,,(w) is:

Yxx(w) = (Wp2 /CW 2 )J (vx2 vz/v) dS/J(v",2/V) dS. (4.58)


Vz>O
y(w) may depend on the direction of the field if the Fermi surface is
non-spherical. For a spherical Fermi surface (4.58) reduces to:
yew) = Us)(v/c)(W p2 /W 2 ). (4.59)
The equivalent correction term to 1', the 'surface collision fre-
quency' 1'8' is:
1'8 = (t)(v/c)wp. (4.60)
The surface admittances Y and y have been defined for normally
incident waves. For an obliquely incident wave Y depends on the
angle of incidence (), as shown by equations (1.33). The values of Y
given by (1.33) yield the Fresnel formulae (1.34) for the reflection
coefficients. On the other hand, y does not change for oblique in-
cidence so the reflection coefficients corresponding to {Y + y} will
show deviations from the Fresnel formulae. These deviations may
be large enough to be detected experimentally when Yew) is small,
that is for w ""' wp. In the frequency range I' ~ w ~ Wp, we have
I Yew) I ~ I and the total surface admittance is approximately
independent of e. The Fresnel equations are approximately valid
with an effective complex refractive index:
(n + iK) = (Y + y). (4.61)

4.5 Infra-red absorption in metals


The complex dielectric constant of a metal can be separated into two
components, sf(w) associated with intraband electron transitions
FREE CARRIER ABSORPTION 77
and eb ( w) associated with inter band transitions. The total complex
dielectric constant is:
e(w) = 1 + eb(w) + ef(w). (4.62)
Since interband transitions have a minimum threshold energy, the
imaginary part of eb(w), e2b(w), is zero below the threshold fre-
quency. This frequency usually lies in the visible or near infra-red
but can be in the middle infra-red for the semi-metals; for example,
bismuth has a minimum threshold energy of 0·05 eV. The frequency
variation of e 2 f (w) below the threshold of e 2b (w) can usually be
extrapolated to frequencies above the interband threshold. This
means that elw) can be separated into e 2f(w) and e2b(W) over the
whole range of frequencies. The corresponding real parts el f(w) and
el b(w) can be calculated by means of the dispersion relations (1.47).
At frequencies below the interband threshold, e1b (w) tends to a
constant value elb(O).
The Drude formula (4.7) forms a convenient starting point for
the interpretation of experimental data on infra-red absorption and
dispersion in metals. Written in terms of ef(w), the formula becomes:
ef(w) = -wzNw(w + iy) (4.63)
where w p 2 = 4nNe 2/m*. For most metals in the infra-red, the
optical absorption and dispersion is dominated by the effects of
intraband transitions. We can put:
e(w) ~ ef(w).
The frequency variation represented by (4.63) can be conveniently
expressed in terms of {e(w)}-l.
{e(w)}-1 = -Wp-2(W 2 + iyw). (4.64)
Most experiments on the infra-red properties of metals take the form
of some type of reflection measurement at the surface of a metal.
The results of such experiments can be summarized in experimental
values of Yew), the surface admittance. Let us define another com-
plex quantity '(w) by:
'(w) = -{ Y(W)}-2. (4.65)
For frequencies where spatial dispersion is negligible, (4.65) reduces
to:
'(w) = -{e(w)}-I.
Using formula (4.64) for {e(w)}-l, we have:
'leW) = wp -2W 2; '2(W) = wp -2yW. (4.66)
78 OPTICA L ABSORPTION AND DISPERSION IN SOLIDS

We have seen that the effect of diffuse surface scattering on the sur-
face impedance is approximately equivalent to a constant correction
term 7'8 added to 7' in (4.63).
The use of formulae (4.66) will be illustrated by an analysis of
some infra-red reflection measurements on lead films. Experimental
values of {(IiW)2/C t } and {Cs/liw} are shown by the points in fig. 4.5.

a
56 f- a 0 0- 0
293°K
'">CD JlO a
0 0
0 0
54 I-
~Iv
52 I- 0

0
0
4 x10-3 f- 0 0
293°K
0 0
0 0 0 0 0 9

"i
>CD
2f-
~I.i 0
0 0
7soK
0 Q 0 Q 0 0 0 0 1:1 (5
0 0 Q Q 0 0 0 0 0 0 0
~

I 1 I 4°K
0 5-0 10-0
2-0
AO(j-Lm)

Figure 4.5 Infra-red properties of lead. Vacuum wavelength Au, photon energy
Roo, complex parameter (Cl + iC 2) = -(81 + i82)-1 for normal skin
effect.
Experimental points.
Reference: A. E. GOLOYA SHKIN, Sov. Phys. JEPT, 21, 548-53 (1965).

According to (4.66) these should be independent of frequency:


(IiW)2/C t = (IiWp)2; Cs/(liw) = (liy)(liwp)-2. (4.67)

It is seen from fig. 4.5 that {(IiW)2/Ct} at room temperature has an


approximately constant value; the scatter can be explained by
experimental errors. Theoretical calculations of the anomalous skin
effect corrections to Ct(w) for diffuse scattering show that deviations
from (4.67) increase with wavelength reaching about 3% at Ao = 12
pm. This is of the same order as the experimental errors in
the
present data. Similar conclusions can be drawn from the experi-
FREE CARRIER ABSORPTION 79
mental values of {(liw) 2/C I} at 78° K and 4° K. Average experimental
values of (IiWll)2 and n*, the effective number of electrons per atom,
are given in Table 4.1. n* is defined by:
n* = (N/Na)(m/m*) (4.68)
where Na is the number of atoms per cm3 and (N/m*) is calculated
from (liw ll ) 2. The values of n* are considerably less than 4'00, the
number of valence electrons per atom in lead. Other measurements
on lead have shown that the area of the Fermi surface is not greatly
different from that of the free electron sphere. It must be concluded
that the optical effective mass m* is several times the electron mass m.
The experimental values of {C2/liw}, as shown by the points in
fig. 4.5, are approximately constant in the middle range of wave-

TABLE 4.1: OPTICAL PARAMETERS FOR LEAD

T 293 78 4 OK

(IiWp) 2 55·2 53-1 51'5 eV2


n* 1·22 1'17 1·14
1i1' 16·4 6·3
3'2} 10- 2 eV
li1'p 15·0 4·4 1·8
1i(1'd + 1'8) 1-4 1·9 1·4

Experimental Idata !from: A. I. GOLOVASHKIN, Sov.


Phys. JETP, 21,548-53 (1965).

lengths. At longer wavelengths there are some small systematic


deviations which may be caused by the anomalous skin effect. The
rise in g~liw} at shorter wavelengths, particularly marked in the
values for 293 K, is approximately proportional to Ao -2, or w 2 • A
0

similar effect has been noticed in other metals, for example gold
where the magnitude of the w 2 term has been found to depend on the
microstructure of the skin layer. One possible explanation is that the
collision frequency y contains a term proportional to w 2 , associated
with electron-electron scattering. A simple consideration of the
initial and final states available for electron-electron scattering gives
a proportionality to w 2 • It is not clear, however, whether this ex-
planation is compatible with the temperature- and structure-
dependence of the w 2 term. The values of IiYll given in Table 4.1
were deduced from the temperature variation of liy using the
80 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

theoretical curve for the electron-phonon collision frequency in


fig. 4.3 with g = 86° K. The values of n* and liy'P for 293° K can be
combined to give a value of 20 ,un-cm for the electrical resistivity of
lead, which agrees well with the measured value. Note that, although
lead is superconducting at 4° K, the measurements were made in a
frequency range where W ~ Wg and consequently there is no 'dis-
continuity at the superconducting transition. The residual collision
frequency (Ya + Ys) is approximately independent of temperature.
The maximum value of Ys is obtained with completely diffuse scatter-
ing. Formula (4.60) in this case gives liys '" 1·0 X 10- 2 eV, assuming
v "'10 8 cm sec-i. Thus the data are consistent with completely
diffuse scattering of electrons at the surface but it is not possible to
separate Ya and Ys without further information.

4.6 Free carrier absorption in semiconductors


At frequencies below the threshold for interband transitions, semi-
conductors have a continuous absorption spectrum associated with
transitions of electrons within the conduction band or holes within
the valence band. This absorption due to intraband transitions is
much weaker than for metals because the concentration of carriers
in a semiconductor does not usually exceed 1019 per cm3 , whereas in
metals the concentrations are of the order ,of 1022 per cm3 • This
means that the skin depths are much larger in semiconductors and
the skin effect is normal at all frequencies. In semiconductors with a
fairly low concentration of carriers, it is possible to determine the
free carrier absorption by transmission measurements on samples
cut from a bulk crystal. This method can be used when the absorp-
tion coefficient does not exceed about 103 cm-i . The measured
infra-red absorption in a semiconductor may include some lattice
absorption and absorption due to low-energy interband transitions
between different branches of the conduction or valence band.
Accurate experimental data on free carrier absorption is available
for many semiconductors in the infra-red for wavelengths up to 20
,urn. The condition W'r ~ 1 is well satisfied by all semiconductors in
this wavelength range. If we compare metals and semiconductors,
the electron mean free paths are of the same order but the electron
velocities in semiconductors are much smaller. For a non-degenerate
semiconductor the electron kinetic energy is !kT, about 0·04 eV at
FREE CARRIER ABSORPTION 81
room temperature, compared with a Fermi energy of several electron
volts in a metal. The relaxation time 7:(= t Iv) is therefore larger for
electrons in a semiconductor and the condition 0)7;' ~ 1 is satisfied
for wavelengths up to about 50 ftm at least. Infra-red photon energies
are comparable with the range of conduction electron energies in a
semiconductor and it is not possible to calculate the infra-red
absorption from the low-frequency relaxation time. The absorption
must be calculated by means of the quantum theory of indirect
transitions.
The complex dielectric constant sew) of a semiconductor can be
expressed as the sum of terms due to interband and intraband
transitions:
(4.69)
A lattice contribution can also be added where appropriate. Intro-
ducing the complex refractive index (n + iK), (4.69) can be written:
(n + iK)2 = {I + Slb(w) + Slr(W)} + iS2f(w) (4.70)
assuming that there are no real interband transitions for the fre-
quencies under discussion. In practice the extinction coefficient K is
sufficiently small for K2 to be neglected in (4.70). Hence, approxi-
mately:
(4.71)
The free carrier absorption in a given semiconductor is a function of
the frequency w, the concentration of free carriers N and the tem-
perature T. In terms of S2f(W), we can write:
S2f(w) = F(w,N,T). (4.72)
We have seen in Section 3.5 that an indirect electron transition re-
quires a lattice defect to maintain the momentum balance in the
transition. The important types of defects for a calculation of
F(w,N,T) are phonons and impurity atoms. If the concentration of
defects is not too large, the total function F(w,N,T) is equal to the
sum of the functions calculated for each type of defect. As an ex-
ample we will consider the calculation of F(w,N,T) for free carrier
absorption in a III-V semiconductor.
The transition probability for electron transitions associated with
each type of defect, as given by second order perturbation theory, is:
Wii = 2nn- 3w- 2 1 ~ (j I HI m)<m I H I i) IY(k)
x {l - f(k')} (j(E j - Ei)' (4.73)
82 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

H is the interaction operator representing the electron-phonon or


electron-impurity interactions. The matrix elements of H are be-
tween initial, intermediate and final states, i, m and j. f(k) is the
Fermi distribution for the initial wave vector k and similarly f(k') for
the final wave vector k'. The sum of matrix elements is taken over all
intermediate states m and over the various combinations of photon
or phonon, emission and absorption. The delta function IJ(Ej - E i )
which represents overall conservation of energy, can be written:
IJ{E(k') - E(k) ± liw ± liw(q)}. (4.74)
w(q) is the frequency of a phonon of wave vector q; the + or -
signs refer to emission or absorption processes. The absorption co-
efficient is obtained by summing (4.73) over all initial and final
electron states.
For the III-V compounds, important contributions to (4.73) arise
from interactions of electrons with longitudinal acoustic (LA)
phonons, longitudinal optic (LO) phonons and ionized impurities.
The interaction with LA phonons is caused by the local changes in
density associated with the LA mode. The interaction energy is
proportional to a deformation potential El which is defined as the
shift of the conduction band edge per unit dilatation of the lattice.
The interaction with LO phonons is particularly strong in crystals
with some ionic character. The strength of the interaction is given by
a coupling constant 0(. When 0( > 1, as in strongly ionic crystals like
the alkali halides, the perturbation theory calculation of the transi-
tion probability is not valid. For the III-V compounds, which are
weakly ionic, 0( < 1 (for example 0( = 0·05 in InAs), and the per-
turbation theory is valid. The interaction with ionized impurities can
be represented by a screened Coulomb potential. It is assumed that
the impurities are of the donor or acceptor type so that they each
carry unit electronic charge. Their effective field is reduced by
lattice polarization and screening by free carriers.
The theoretical formulae for F(w,N,T) are complicated, but the
frequency-dependence can often be expressed approximately as a
simple power law over a certain range of frequencies:
B2f(W) f"OooI Cw-B. (4.75)
The Drude formula (4.63) for w ~ y (wt' ~ 1) becomes:
B2f(W) ~ wp 2yw- 3 = Cw- 3• (4.76)
FREE CARRIER ABSORPTION 83
The value s = 3·0 is obtained from the exact formula when liw <{ EF
for a degenerate semiconductor or liw <{ kT for a non-degenerate
semiconductor, and the scattering is due to acoustic phonons and
impurities. Other values of s between 2·5 and 4·5 are obtained
according to the degree of degeneracy and the proportions of various

4
Ao(fLm)

Figure 4.6 Free carrier absorption in n-type indium antimonide. Vacuum wave-
length Ao, imaginary part of free carrier dielectric constant e2F. Points represent
experimental values at 80 K for sample A (N = 4'7 x 1016 cm- 3) and sample B
0

(N = 4·5 x 10 17 cm- 3 ). Lines represent power laws, e2 F (w) = Cw-' with


s = 4'5 and 3·5.
Reference: R. M. CULPEPPER and J. R. DIXON, J. Opt. Soc. America, 58, 96-102
(1968).

scattering processes. The frequency variation of B2f(w) for indium


arsenide is shown by a log-log graph in fig. 4.6. The samples A and
B had different concentrations of electrons; the straight lines rep-
resent empirical relations of the form (4.75) with s = 3·5 and 4·5.
The theoretical values of F(w,N,T) for indium arsenide are in good
agreement with experimental values over the investigated range of
each parameter.
5
Plasma Effects
5.1 Free electron model
The simplest model of an electrical plasma is a gas of free electrons
moving in a uniform distribution of positive charge which maintains
average charge neutrality. For this model, the real part of the di-
electric constant is:
(5.1)
where wp 2 = (4nNe /m); N is the number of electrons per cm 3 •
2

Inelastic electron collisions have been neglected and in this model


the plasma does not absorb any energy from the electromagnetic
field at non-zero frequencies:
O'l(W) = C2(W) = O. (5.2)
This model can be made more realistic by replacing the continuous
positive charge with a lattice of positive ions. The motion of electrons
in an ionic lattice can be represented under suitable conditions by the
effective mass approximation. For our purpose this will lead to a
redefinition of Wp with m* instead of m :
w p 2 = (4nNe 2/m*). (5.3)
In addition, the ionic lattice will give a contribution to the dielectric
constant due to electronic and lattice polarizability. If real transi-
tions between energy levels of the lattice are sufficiently weak at the
frequencies under consideration, then (5.2) remains approximately
valid. Formula (5.1) must be modified to include the contribution to
Cl(W) from the lattice:
(5.4)
We will now consider the optical properties of a medium with
dielectric constant given by (5.2) and (5.4). The dielectric constant
is always real; it is negative at low frequencies and positive at high
frequencies. The frequency Wo at which Cl(W) changes sign is given by:
Cl(W O) = 0 :. Wo = wp{l + ClC(WO)}-t. (5.5)
84
PLASMA EFFECTS 85
Wo is called the plasma frequency. The condition for the existence of
transverse electromagnetic waves of the form:
E = Eo exp {i(k l + ik2).r - iwt} (5.6)
can be expressed as:
kl2 - k22 + 2ik l . k2 = s(w )(w 2Jc2). (5.7)
For w < W o, sew) is real and negative so k2 ~ 0 and kl.k2 = O. The
waves are evanescent with kl perpendicular to k2 and:
k22 - k l 2 = -SI(W)(W 2/C 2). (5.8)
External waves incident on the surface of the medium are totally
reflected for w <; woo At frequencies above the plasma frequency,
sew) is real and positive so both ordinary and evanescent waves can
exist in the medium. The refractive index new) is given by:
new) = {c(w))!'. (5.9)
For a range offrequencies immediately above the plasma frequency,
CI(W) and hence the refractive index new) < 1. External waves are
totally reflected for angles of incidence larger than the critical angle
eo where:
sin eo = new). (5.10)
Some experimental values of the normal reflectance Ro for a
crystal of indium antimonide in the region of the free carrier plasma
frequency are shown by the points in fig. 5.1. The theoretical curve
was calculated from CI(W) given by (5.4) with N = 4·0 X 1018 cm- a,
m* = 0'048m and CIC(w) = 15·7. Typical free carrier densities for
semiconductors correspond to plasma frequencies in the infra-red.
The large value of CIC(w) for indium antimonide is due to interband
transitions with a threshold energy near 0·2 eV. The general trend of
experimental values agrees with the theoretical curve but the mini-
mum Ro is not quite zero and Ro does not rise quite to unity below
the plasma frequency. These differences can be explained as con-
sequences of the finite relaxation time of the plasma electrons. The
relaxation time in indium antimonide is particularly long; reflectivity
curves of other semiconductors with· smaller relaxation times show
a less well-defined plasma edge.
The densities of valence electrons in solids are several orders of
magnitude greater than the free carrier densities in semiconductors.
The calculated values of nw'P for the valence electrons, assuming
86 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

1·001------,
o
o
o
o

rf 0·50 o

o
o 0·20
nw (eV)
Figure 5.1 Plasma edge of n-type indium antimonide. Normal reflectance Ro,
photon energy nw.Experimental points and theoretical curve (N = 4·0 X 1018
em-a, m* = 0'048m, 88w) = 15'7).
Reference: w. G. SPITZER, and H. Y. FAN, Phys. Rev., 106, 882-90 (1957).

m* = m, lie in the range 3 to 20 eV. The alkali metals with one


valence electron per atom have the smallest values. The valence
electrons behave like a free electron plasma with m* ~ m for
excitations with energies well above the transitions of appreciable
oscillator strength. Excitations of energy ;> liw p approximately
satisfy this criterion for free electron behaviour of the valence
electrons, in most solids: metals, semiconductors or insulators. If
the core electron contribution to feW) is just the real term f1C(W) then
the solid behaves like a free electron plasma for a range of frequencies
which is defined below. The upper limit of the range is determined by
the threshold for interband transitions of the core electrons, say Wc'
The lower limit of the range is determined by the maximum fre-
quency of appreciable interband transitions of the valence eleCtrons,
PLASMA EFFECTS 87
say WB' The free electron model is approximately valid in the
frequency range:
We > W > WB. (5.11)
For most insulators and semiconductors, and many metals,
we> Wp > WB;
the range 'of free electron behaviour includes Wp and the plasma
frequency Wo is defined by (5.5). For example, aluminium has
liwe ~ 80 eV, liwp = 15·8 eV and IiWB ~ 2 eV. The behaviour of
aluminium approximates to the free electron model over a wide
range of frequencies in the visible and ultra-violet. In some metals
the threshold for core electron transitions lies at a frequency below
Wp. For example, copper has liwe = 2·1 eV, liwp = 9·5 eV and
IiWB ~ 5 eV. The core electrons in copper include the 3d shell which
is separated by only 2·1 eV from the Fermi level. The overlapping of
valence electron and d-electron transitions prevents the existence of
a frequency range in which the free electron behaviour of the valence
electrons is dominant. This is also true of the transition elements
with their partially filled d-bands. From the sizes of its parameters,
liwe = 3·9 eV and liwp = 9·0 eV, silver would seem to resemble
copper but because of the considerable value of C1C(W), the plasma
frequency defined by (5.5) is liwo = 3·8 eV. This falls just below the
threshold for d-electron transitions and so condition (5.2) for free
electron behaviour is approximately satisfied at the plasma frequency.
In an actual solid the condition (5.2) for zero absorption cannot
be exactly satisfied. A residual absorption is always present due to
weak interband transitions or relaxation effects. The simple calcula-
tion of electron relaxation in Section 4.1 gave:
C2(W) = wp 2 yw- a (5.12)
for W ~ y, where y is the collision frequency. Experimental values of
C2(W) for aluminium in the ultra-violet fit (5.12) quite well with
y ~ 0·9 X 1015 sec-1 for w.;;;;; Wp2; at higher frequencies C2(W) is
somewhat larger. This value of y is several times that deduced from
the infra-red absorption. The difference can be explained by an in-
crease in electron-electron scattering at high frequencies. For the
alkali metals in the region of the pla~ma frequency, C2(W) varies more
slowly with W than (5.12) and is probably associated with interband
transitions. The free electron model predicts complete transparency
at frequencies above the plasma frequency. Experiments on metal
G
88 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

films have shown that they are sufficiently transparent for W > Wo
to allow measurement of refractive index from the critical angle or
interference effects in thin layers. Some experimental values of el(W)
for sodium and aluminium above their plasma frequencies are shown
in fig. 5.2. They confirm the frequency-dependence of el(W) predicted

£-2 (eV-2)

Figure 5.2 Dielectric constant of metals above the plasma frequency. Real part
of dielectric constant 81> photon energy E (= liw).
References: 1. C. SUTHERLAND, E. T. ARAKAWA and R. N. HAMM, J. Opt. Soc.
America, 57, 645-50 (1967); w. R. HUNTER, J. de Physique, 25,154-60 (1964).

by (5.4); the values of liwp from the slopes of the graphs are 5·8 eV
(Na) and 14·8 eV (AI). These values are slightly less than the calcu-
lated values for the valence electrons with m* = m, 5·9 eV (Na) and
15·8 eV (AI). It seems that m* is a few per cent larger than m even
for excitations above the plasma frequency.

5.2 Volume plasmons


As well as the transverse electromagnetic waves considered in the
previous section, longitudinal electromagnetic waves can also occur
in a plasma. Consider a slab of plasma as shown in fig. 5.3, with the
PLASMA EFFECTS 89
l

: +
I

:
I

+,1
1
1
I
E 1
+',I
1
'I 1
>x< >x<

Figure 5.3 Volume plasma oscillation. E = 4nNex~

electrons displaced a distance x perpendicular, to the plane of the


slab. The average equation of motion of an electron is;
m(d 2x/dt 2) = -4nNe 2x (5.13)
ignoring inelastic electron collisions. This is an equation of un-
damped harmonic motion of angular frequency:
OJ1) = (4nNe 2 /m)t. (5.14)
The oscillation is equivalent to a longitudinal electromagnetic wave
in the limit as the wave number tends to zero. To investigate this
further let us see how a longitudinal wave can be compatible with
Maxwell's equations. Assuming a plane-wave form for E, the
equation V'. D = 0 can be written:
e(k,OJ)k.E = O. (5.15)
In a longitudinal wave k and E are parallel so k . E :#- 0 and therefore:
e(k,OJ) ~ O. ' (5.16)
Equation (5.16) is an implicit relation between k and OJ for plasma
waves. The quanta of these waves are called plasmons. Theoretical
expressions for e(k,OJ) have been calculated for the free electron
model. An approximate form of (5.16), valid for small k is:
OJ2 = OJ p 2 + tVB 2k 2• (5.17)
VB is the Fermi velocity. Equation (5.17) gives real values of k, cor-
responding to undamped waves, for frequencies above OJ1)' The
corresponding equation for transverse electromagnetic waves is:
(5.18)
90 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

where c is the velocity of light. The ratio of wave numbers for the
same frequency is:
(5.19)
hence kL ~ k T • Plasma waves can be excited by a p-polarized
electromagnetic wave incident obliquely on the surface of a plasma.
This possibility causes appreciable deviations from the Fresnel
formula for the reflectance rp when w >- Wp. Coupling between
external p-waves and plasma wave resonances in thin slabs has also
been predicted. These effects of plasma waves on optical properties
have not been verified experimentally to date.
The dielectric constant c(k,w) in (5.15) and (5.16) refers to longi-
tudinal electrical fields, whereas the previously defined c(k,w)
referred to transverse fields. It can be shown that the two quantities
are identical for k r-J 0, which is a good approximation for optical
waves. The equation (5.16) for plasma waves is not limited to any
specific model and in general c(k,w) must be assumed complex, so
(5.16) becomes:
(5.20)
Equation (5.20) can be satisfied if we allow k or w to have complex
values; in either case the solution corresponds to a damped wave.
We will assume that k is real and w complex, and consider plasma
waves with k ~ 0 so as to establish a connection with the optical
properties. Let us write the solution of (5.20) for k r-J 0:
w = Wo - iyl2. (5.21)
This corresponds to a time variation of the field of the plasma wave:
exp (-iwt) = exp (--!yt) exp (-iwot). (5.22)
The plasmon is a well-defined excitation if the damping is not too
large, that is (wo/Y) ~ 1. For zero damping, C2(W) = 0 and equation
(5.20) becomes:
(5.23)
For small damping we can approximately represent c(w) in the region
of the plasma frequency by:
c(w) r-J (w - wo)(dct/dw)w=wo + iC2(W O). (5.24)
Combining (5.20), (5.21) and (5.24) we have:
y ~ 2c2(wo)/(dclldwL=wo (5.25)
PLASMA EFFECTS 91
If sew) can be represented by the Drude formulae (5.4) and (5.12)
then y in (5.25) is the same as the collision frequency y in (5.12).
The most direct evidence for the existence of plasmons in solids
has come from electron energy loss experiments. When electrons of
energy 10 to 50 keY are passed through a thin layer of a solid
the energy loss spectrum has strong peaks at integral multiples of the
plasmon energy. The interaction between a moving electron and the
plasma is purely electromagnetic but it differs from the interaction
between an electromagnetic wave and the plasma because the field
of an electron is longitudinal. For a stationary distribution of charge
of density p, the Coulomb potential V is given by:
V 2 V = -4np. (5.26)
If p is analysed into spatial Fourier components, (5.26) becomes:
k 2 V(k) = 4np(k). (5.27)
When the charge distribution is moving with velocity v, each com-
ponent V(k) becomes a travelling wave with frequency:
w =k.v. (5.28)
The associated electric field, - VV, is parallel to k, and its interaction
with the medium can be described by s(k,w) for longitudinal fields.
The energy loss in a longitudinal wave is proportional to:
O"ik,w)/1 s(k,w) 12; (5.29)
0"1(k,w) is the optical conductivity; notice the analogy with the ex-
pression in Section 1.4 for optical absorption in a thin film associated
with the normal component of field. In (5.29) 0"1 represents the
energy dissipation and 1 sl-2 is a screening factor. The energy loss
spectrum of an electron is obtained by an integration of (5.29) over
the appropriate range of k. This integral is weighted towards small
values of k and the energy loss spectrum is approximately propor-
tional to (5.29) for k = o. It is convenient to express (5.29) in terms
of ~(w), introduced in Section 4.5 and defined by:
~(w) = ~1(W) + i~2(W) = -{S(W)}-l. (5.30)
Function (5.29) is proportional to {W~2(W)} but {C2(W)} is usually
taken to represent the electron energy loss spectrum. However, the
positions of less well-defined peaks in the energy loss spectrum are
92 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

given more accurately by {WC2(W)}. From (5.24) and (5.25), the


energy loss function near the plasma frequency is:

C2(W) = ty . (53 )
(del/dw)",=",.{(w - WO)2 + ty2} . I
The maximum value of {C2(W)} is {e 2(wo)}-1 and the half-width of the
peak, in energy units, is iiI'. Fig. 5.4 shows the height and width of

20

10

~2 5

o 10 20
. flw(eV)

Figure 5.4 Plasmon peaks in the electron energy loss function C2(W). C2 = 8 2/
(81' + 822). Positions and half-widths of peaks calculated from optical data.
plasma loss peaks in some solids. The half-width iiI' was obtained
from (5.25) using measured values of the optical constants. The
positions and widths of the plasma loss peaks calculated from optical
data agree well with the directly measured electron energy loss
spectrum.

5.3 Surface plasmons


Let us consider the possible electtolnagnetic waves in the region of
a plane boundary. between a plasma, .and a transparent medium,
z = 0 in fig. 5.5. We will consider frequencies below the plasma
frequency and rteglect damping, so e(w) is real and negative and
n = 0, K = (-e)l. An ordinary wave inthe transparent medium is
totally reflected at the interface, whatever the angle of incidence or
PLASMA EFFECTS 93
z

s~--------~-------------

or-----------------------~
PLASMA (£<0)

Figure 5.5 Optical excitation of surface plasmons. Transparent media (co and
co'). Plane interfaces z = 0 and z = s.

polarization. Three waves, the incident and reflected waves in the


transparent medium and an evanescent wave in the plasma, together
satisfy the boundary conditions at the interface. Whereas the wave
in the plasma must be evanescent for w < wo, both ordinary and
evanescent waves are possible in the transparent medium. We will
consider the possibility of matching a pair of evanescent waves, one
in each medium, propagating along the interface. Assuming p-
polarization, the electric vector of an evanescent wave in the trans-
parent medium can be written:

(E",Ey,E.) = (Eo,0'(k2 ~::::jC2)t)


X exp {-z(k 2 - Eow2jc 2)t + i(kx - wt)}. (5.32)
The form of (5.32) has been chosen to satisfy equation (1.23) for the
wave vector and the condition k. E == O. The planes of constant
amplitude are parallel to the interface, z = 0, and the wave is propa-
gated with zero attenuation in the x-direction. The matching wave
in the plasma must have the same wave number k in the x-direction;
its electric field can be written:

(E",Ey,E.) = ( Eo,0'(k 2 _ ikEo)


Ero 2jc 2)t

X exp {+z(k 2 - Ew2jc 2 l + i(kx - wt)}. (5.33)


94 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The boundary conditions at the interface are the continuity of the


tangential field, Ex, and the normal displacement, Dz. The first con-
dition is satisfied by putting the same constant Eo in (5.32) and (5.33).
The second condition gives:
e(k 2 - ew2jc 2)-i = -eo(k 2 - eow2jc 2)-i. (5.34)
Equation (5.34) can only be satisfied if e is negative; this is a neces-
sary condition for the existence of this type of surface wave. It is also
necessary for the wave to have p-polarization; the roles of E and H
are interchanged for an s-wave and a medium of negative perme-
ability would be needed to satisfy the boundary equation for Bz •
Rearranging (5.34), we have:
k 2 = (w 2jc 2){e o- 1(w) + e-1(W)}-1. (5.35)
For low frequencies I e(w) I is large and (5.35) is approximately the
same as the dispersion relation for ordinary waves in the transparent
medium. As the frequency is increased, k -+ 00 for a limiting fre-
quency w. given by:
(5.36)
w. is called the surface plasmon frequency. Assuming for the plasma
dielectric constant:
e(w) = {1 + e1 (w)}{1
C - W 0 2W- 2}, (5.37)
where Wo is the volume plasmon frequency, then:
w. = wo{l + e1 (w.)}! j{1 + e1 (w.) + eo(w.)}t.
C C (5.38)
If e1C(w S) ~ 0 and eo = 1, then w. = wo/v'2; for a metal like silver
with a large value of e1C (w.), the ratio (wsjw o) is nearer unity.
Formulae (5.35) and (5.37) have been used to calculate the graph in
fig. 5.6. The straight line through the origin corresponds to an
ordinary wave in the transparent medium with phase velocity
(wjk) = (cje ot ), here assumed constant. The phase velocity of a sur-
face plasma wave decreases from (cje ot ) at k = 0 to zero, as k -+ 00.
We have assumed a plasma without damping but similar surface
waves are possible with damping so long as the real part of e is
negative, e1 < O. A non-zero imaginary part, e 2 , means that k has
a complex value and there is some attenuation in the x-direction.
Also, equation (5.36) for the surface plasmon frequency has a com-
plex root corresponding to a finite plasmon lifetime.
We will now consider the excitation of surface plasmon waves by
PLASMA EFFECTS 95
0-8~------------~/------------~------~
I
~ I
Wo ~----------,/~---------------------;
I
I
I

,I
I

I
I
I
I
I
I
0-4 I
I
l

o 2
klko
Figure 5.6 Dispersion curve of surface plasmons. Volume plasmon frequency
000;dashed line (00/000) = (k/k o), where ko = (ooo/C)Eot, represents the dispersion
relation for ordinary waves in the transparent medium. Curve calculated from
(5.35) and (5.37) with El c = 0 and Eo = 1.

electrons and electromagnetic waves. A surface plasma wave has a


component of field, EIIJ , in the direction of propagation, and can
therefore be excited by a moving electron. The energy loss of an
electron to the surface plasma wave is quantized in units of the
surface plasmon energy. The electron couples with surface plasma
waves of phase velocity equal to the parallel component of the
electron velocity. For an electron velocity much less than the velocity
of light, the electron couples with plasmons of large k and energy
nearly lim.. Experiments on electron energy losses have confirmed
this mode of excitation of surface plasmons. An energy loss function
for surface plasmons can be defined, analogous to (5.30) for volume
plasmons:
(5.39)
A surface plasma wave can be excited by an ordinary electro-
magnetic wave by means of the arrangement shown in fig. 5.5. A
second transparent medium of higher index is placed a small distance
s from the interface z = O. An ordinary wave in this medium falls
at an angle of incidence () on the interface z = s, where () is greater
96 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

than the critical angle. An evanescent wave of phase velocity


(C/eo'l sin 0) is generated and if the separation s is small enough, it
has an appreciable amplitude at the interface with the plasma. The
phase velocities of the evanescent wave from z = s and the surface
plasma wave can be matched since:
C/(eo')t sin 0 < C/(eo)l. (5.40)
Another possible method of coupling an ordinary wave to a surface
plasma wave is by means of an interface with periodic structure in
the x-direction, for example a grating surface with rulings perpen-
dicular to the x-axis. Let p(x) be the charge density associated with
a surface plasma wave on a smooth interface:
p(x) = Po exp (ikx). (5.41)
If the interface has structure of fundamental period ~ in the x-
direction, the charge density is modulated by a function which can
be represented by a Fourier series with fundamental wave number:
qo = 21'(,/~. (5.42)
The modulated charge density contains components with wave
numbers:
k±nqo (5.43)
where n is an integer. An ordinary wave incident on the interface has
an x-component of wave vector given by:
ko sin 0 =eol(ro/C) sin O. (5.44)
Resonant coupling with the surface plasma wave can occur when
(5.44) and (5.43) are equal:
ko sin (J = k ± nqo. (5.45)
These resonances cause anomalies in the p-wave reflectance of
gratings at frequencies for which (5.45) is satisfied. Measurements of
these anomalies can be used to derive an experimental dispersion
curve for surface plasmons. .
6
Exciton Effects
6.1 Electron-hole interaction
The theory of interband transitions given in Chapter 3 neglects the
interaction between an electron which has made a transition to the
conduction band and the hole which it leaves in the valence band. It
is clear that there will be an electrostatic interaction between the
electron and hole, screened by other electrons and ions. When this
interaction is included in the theory it is found that excited states
occur with energies inside the forbidden energy gap of band theory.
Transitions to these states can therefore cause absorption at energies
below the minimum energy for interband transitions. The excited
states are associated with bound electron-hole pairs, called excitons,
which can propagate through the crystal.
A simple type of exciton, called a Frenkel exciton, occurs in
crystals of weakly bound atoms or molecules, for example the
crystals formed by condensation of noble gases. Each atom has
complete electron shells which interact weakly with neighbouring
atoms and there is little overlap between the wave functions of
neighbouring atoms. Let us suppose that one of the atoms is raised
to its first excited state. If the wave function of the excited state does
not extend appreciably beyond one lattice cell, the excited state will
be the same as an atomic state except for a small perturbation by the
neighbouring atoms. AlthOugh the excitation is localized on one
atom there is a finite probability of transfer to a neighbouring atom.
In a perfect lattice the excitation can propagate without scattering
and we can define a corresponding excitation wave. Using Bloch's
theorem on wave functions in a periodic lattice, one can show that
the wave function of the excited state can be expressed in the form:
1J'(K) = (llyN) ~ exp (iK.R)1J'(R). (6.1)
R
K is the wave vector of an excitation wave and IiK is the crystal
momentum of the associated exciton; 1J'(R) is the wave function of
97
98 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

the crystal with an excited atom at the lattice site specified by R. The
energies associated with lJ'(K) can be calculated by perturbation
theory. The exciton energy band, E(K), has a width which depends
on the interaction between an excited atom and its neighbours.
So far we have been considering the first excited state of an atom
in a crystal but similar considerations apply to higher excited states
as long as the wave functions are localized in a single lattice cell.
This condition will obviously break down as the energy approaches
the ionization limit and the wave functions merge into the non-
localized functions of the continuum states. In many crystals even
the wave function of the lowest excited state extends over several
lattice cells. In that case the excitation is quite different from the
excitation of an isolated atom. The electrostatic electron-hole
interaction is· screened by the polarization of the crystal. If the
exciton wave function extends over a large number of lattice cells,
the screening effect of the crystal polarization can be represented by
a dielectric constant e. The electron-hole interaction energy with
dielectric screening is:
VCr) = -e 2/er. (6.2)
The appropriate value of e is obtained from a consideration of e(W)
for the particular crystal and the eigenfrequency w of the exciton
wave function; w = Eb/fi, where Eb is the binding energy of the
exciton. For a crystal like germanium, withe solely due to electronic
polarization, it is appropriate to take the constant infra-red value,
e ~ 16, for most exciton states. For an ionic crystal, with a lattice
contribution to e for frequencies below the Reststrahlen band, the
appropriate value of e may be e(O) or e( (0) depending on the eigen-
frequency of the exciton. In addition, the lowest energy excitons of
an ionic crystal may be nearly localized in a single lattice cell with
negligible screening.
A theory of excitons which extend over several lattice cells has
been formulated by Wannier and Mott. We will consider its applica-
tionto excitons in a semiconducting solid where the electron-hole
interaction is substantially reduced by dielectric screening. Let us
consider excitons formed from electrons and holes with momenta
near extrema of the band energies Ee(k) and Ev(k). If the extrema
have the same value of k, the excitons are called direct excitons and
they can be produced by direct absorption of photons. Indirect
EXCITON EFFECTS 99
excitonsare formed from electrons and holes near energy extrema
of different k. As a simple example of a direct exciton let us consider
electrons and holes near k = 0 in a semiconductor with energy
bands:
Ec(k) = Eo + /i 2k 2 /2m. *;
(6.3)
Ev(k) = -/i 2k2j2mh*'
The electrons and holes behave like particles with charges -e and
+e, and masses m.* and mh*' The Hamiltonian operator for an
electron-hole pair is:
H = -(/i 2 /2m.*)V.2 - (/i 2/2mh*)Vh2 - (e 2/sr). (6.4)
The first two terms represent the kinetic energies of the electron and
hole, and the third term represents their interaction. Formula (6.4)
also represents the Hamiltonian of a hydrogen atom if we assume
me * = free electron mass, mh * = proton mass and s = 1. The wave
functions and energy eigenvalues deduced from Schrodinger's
equation for a hydrogen atom can easily be modified to apply to an
exciton. The energy eigenvalues are:
En(K) = Eo - n-2{pe 4 j2/i 2s 2) + (/i 2K2/2M*), (6.5)
where n is the principal quantum number (n = 1, 2, 3 etc.); p, is the
reduced mass (m.*mh*)/(m.* + mh*); M* is the total effective mass
of an exciton (m.* + mh *). The final term in (6.5) represents the
kinetic energy of translation of an exciton. Eo is the minimum energy
needed to produce an unbound electron-hole pair from the ground
state. At energies above Eo there is a continuum of states which
correspond to the states of the band model but the wave functions
are modified by the electron-hole interaction. The binding energy of
an exciton in the lowest state is:
(6.6)
The corresponding binding or ionization energy of a hydrogen atom
is 13-6 eV. It follows from (6.6) that for an exciton:
Eb = 13'6(p,/ms 2) (eV). (6.7)
It is unusual for p, to exceed the free electron mass m and it may be
considerably less than m; for a typical semiconductor s "" 10. The
binding energies of excitons in semiconductors are therefore typically
of the order of several hundredths of an electron-volt. The Bohr
radius of an exciton state is given by:
an = O'52(sm/p,)n 2 (A). (6.8)
100 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The effective mass approximation, as used in (6.4), is valid when an


is much larger than the lattice constant.

6.2 Optical absorption


The optical absorption associated with the formation of excitons
can be calculated by perturbation theory. The method is similar to
that described for direct interband transitions in Section 3.2. The
optical conductivity associated with transitions between a ground
state 1J'0 and an exciton state 1J' is:
O"l(W) = (ne 2Ji2/m2w) ~(E - Eo - Jiw)
f
x 1 1J'* exp (iko.r)n. V1J'o dv 12, (6.9)

where.. Eo and E are the initial and final energies, ko is the wave
vector of the electromagnetic field and n is a unit vector in the direc-
tion of the electric field. The exciton state 1J' can be written in the
form 1J'(K) given by (6.1). It follows from the oscillatory character
of 1J'*(K) that the integral in (6.9) is zero unless:
ko - K = O. (6.10)
Since ko is muchless than typic!!.l vectors in the Brillouin zone, (6.10)
can .be written:
, .
K '"" o. (6.11)
Thus a photon can only cause transitions from the ground state to
K = 0 states of exciton bands. This means that the exciton absorp-
tionspectrum (,:onsists of discrete lines in the ranges of energy where
the exciton bands are discrete. Using the selection rule (6.11), the
matrix element for optical absorption can be written:
f 1J'*(O)V1J'o dv. (6.12)
The crystal wave functions 1J' and 1J'0 can be expressed as determin-
ants of one-electron wave functions 1jJ and 1jJo. In the Frenkel model
of an exciton, 1jJ and 1jJo are atomic wave functions, the same as for
an isolated atom. In the Wannier-Mott model, 1jJ and 1jJo are localized
wave functions formed from the Bloch wave functions of band
theory, such as the solutions to Schrodinger's equation with the
Hamiltonian (6.4). Substitution of the determinantal form for 1J' and
Po reduces (6.12) to:
(6.13)
EXCITON EFFECTS 101
where N is the number of atoms per unit volume. It follows that the
absorption spectrum of Frenkel excitons in a crystal is the same as
the absorption spectrum of similar atoms in a low-pressure gas.

6.3 Inert-atom solids and alkali halides


The noble gases argon, krypton and xenon condense at low tempera-
tures into crystalline solids with face-centred cubic structures. The
minimum energy for electronic transitions in these solids is of the
order of 10 eV, corresponding to an ultra-violet photon. The ultra-
violet absorption spectrum of a thin film of solid krypton is shown in
fig. 6.1. The peak absorption coefficient in the strongest lines is about

Figure 6.1 Ultra-violet absorption spectrum of a film of solid krypton at 20° K


(annealed at 44° K). Absorption coefficient K (maximum value --..10 6 cm- 1),
photon energy liw. Atomic absorption lines shown by arrows.
Reference: G. BALDINI, Phys. Rev., 128, 1562-7 (1962).

10 6 cm -1. The arrows above the absorption curve represent the


energies of atomic resonance lines of krypton. The two lines corres-
pond to transitions from the ground state of atomic krypton to a
doublet excited state, split by spin-orbit coupling. The energies of the
atomic lines and the exciton peaks at 10·2 and 10·8 eV are nearly
the same, as required by the Frenkel model. A calculation of the
exciton radius in the lowest state, n = 1, gives a 1 ~ 2 A which is
102 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

comparable with the interatomic spacing. The Frenkel model is


justified for the lowest exciton state but higher states overlap several
lattice cells and the Wannier-Mott model must be used. The exciton
peaks at 10·2 and 10·8 eV both correspond to n = 1 in the hydro-
genic model. Two series of peaks for n = 2, 3 follow the n = 1
peaks. Fitting the n = 2 and n = 3 peaks of the first series to the
hydrogenic formula (6.5) gives Eo = 11·7 eV and Eb = 1·7 eV. Eo is
the ionization limit of the first series of lines and is equal to the direct
energy gap in the band model. The parameter Eb is not exactly equal
to the exciton binding energy 1·4 eV, because of the variation of di-
electric screening with n. From the hydrogenic formula (6.7) for E b,
we have:
(p,/me 2) = 0·13. (6.14)
The dielectric constant in the visible spectrum e ~ 1'80, so the
reduced exciton mass '" ~ 0·4m. It is significant that the n = 2 and
n = 3 peaks only appear with annealed films of krypton. This can be
explained by the effect of crystal defects on the exciton energy levels.
The wave functions increase in extent with increasing n and the
energy levels of higher n are therefore more susceptible to broadening
by crystal defects than the n = 1 level.
Alkali halide crystals resemble the inert-atom crystals in having
their electrons grouped in closed shells. The electrons of both the
alkali and halogen ions form clos-ed shells but the outer electrons of
the halogen ions are less strongly bound and fill the highest energy
bands. The band structure of a typical alkali halide crystal for two
symmetry directions is shown in fig. 6.2. Of the four bands shown,
the three lower are filled and the upper one is empty; the splitting d
is due to spin-orbit coupling. The minimum energy gap Eo is at the
centre r of the zone and the threshold for direct· interband transi-
tions is liw = Eo. The values for a crystal of potassium chloride are
Eo = 8·7 eV and d = 0·11 eV. There are two series of exciton levels
r
at associated with the two bands separated by d. The optical con-
ductivity for potassium chloride at 10° K, shown in fig. 6.3, has a
strong double peak associated with transitions to the n = 1 exciton
states. This is similar to the doublet in the exciton absorption
spectrum of an inert-atom solid. The n = 2 peaks, which are observ-
able only in single crystals at low temperatures, are broad and
poorly defined. It is believed that the strong fields between ions
EXCITON EFFECTS 103

L <111> <100> x
k
Figure 6.2 Schematic valence and conduction bands of face-centred cubic alkali
halide crystals in <Ill) and (100) directions. Parameters for potassium chloride:
Eo = 8·7 eV, Ll = 0·11 eV.

80~------------------------------~

lOOK

75 8·0 8·5
hw(eV)

Figure 6.3 Ultra-violet optical conductivity of a single crystal of potassium


chloride. Optical conductivity 0' h photon energy lim.
Reference: T. TOMIKI, J. Phys. Soc. Japan, 26, 738-58 (1969).
H
104 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

broaden the energy levels of the higher exciton states and smear out
the absorption peaks.
The temperature has a very marked effect on the exciton absorp-
tion spectra of alkali halide crystals. This is illustrated in fig. 6.3 by
the curves of optical conductivity for potassium chloride at 10 K 0

and 295 0 K. The curve for 295 0 K shows only a trace of doublet
structure and a considerable lowering and broadening of the main
peaks. The energy shift of the peaks with temperature is caused by
the change in lattice spacing. To explain the thermal broadening of
exciton absorption lines it is necessary to extend the theory to in-
clude phonon effects. The phonons may be considered as broadening
the exciton states between which the photons cause transitions.
Alternatively, the transitions can be considered as the result of two-
stage photon-phonon absorption processes. The latter approach was
used in the discussion of indirect interband transitions in Section 3.5.
The delta functions in the theory of Section 6.2 are replaced by
functions of finite width. The theory predicts line profiles which are
approximately Lorentzian but with some degree of asymmetry
because of the predominance of phonon emission processes at low
temperatures.

6.4 Semiconductors
As we have seen in the previous section, the effects of excitons and
the electron-hole interaction on the absorption spectrum of a solid
are found principally in the neighbourhood of interband absorption
edges. These edges occur in the visible and infra-red regions of the
spectrum for semiconducting crystals so we can expect to see exciton
effects in these regions also. The binding energies of excitons in
semiconductors are generally small, in the range 0·001 to 0·1 eV,
because of the large dielectric constants. The excitons extend over
many lattice cells and the effective mass Hamiltonian, with modifica-
tions to allow for anisotropic and degenerate energy bands, is a good
approximation. Well-developed series of exciton absorption lines
below the interband threshold have been observed in large gap semi-
conductors, particularly in cuprous oxide crystals. Smaller gap
semiconductors, such as germanium, have such weakly bound
excitons that it is impossible to resolve the individual lines of the
exciton spectrum. However, the effect of the electron-hole inter-
EXCITON EFFECTS 105
action on the continuous absorption above the interband threshold
has been clearly established.
Let us consider the effect of the electron-hole interaction on the
absorption spectrum of a semiconductor near a threshold for direct,
allowed interband transitions. A series of absorption lines below the
threshold of continuous absorption are produced by transitions to
the K = 0 states of the exciton bands. These lines occur at energies
Eb(0)/n2 below the absorption edge, where n = 1, 2 etc. The con-
tinuous absorption above the threshold is associated with the
formation of unbound electron-hole pairs. The attractive force
between the electron and hole increases the probability that they will
be close together although in free particle states. This enhances the
continuous absorption over that calculated from the band model,
particularly near the threshold. The continuous absorption is in-
creased by a factor:
(IX exp IX/sinh IX) (6.15)
where IX = n{Eb/(nw - Eo)}!. Near the threshold, where IX ~ 1, the
band theory formula (3.40) predicts an optical conductivity O'l(W)
proportional to (nw - Eo)~. The correction factor (6.15) changes
this and 0' 1(w) tends to a constant value, rather than zero, at nw = Eo.
For energies well above the threshold, IX ~ 1 and
(IX exp IX/sinh IX) -+ 1.
Fig. 6.4 shows the absorption spectrum of germanium near the
threshold for direct interband transitions; Eo ~ 0·89 eV at low tem-
peratures. The binding energy Eb of the K = 0 exciton, deduced
from an analysis of the absorption spectrum, is 0·0011 eV. Because
of the small value of E b , the exciton lines overlap and can be seen
only as a single peak just below nw = Eo. The considerable enhance-
ment of the continuous absorption due to the electron-hole inter-
action is apparent.
When the interband transitions at the edge are forbidden, the
electron-hole interaction again causes an enhancement of the con-
tinuous absorption. The exciton line spectrum is changed by the
absence of the n = 1 line. This type of exciton spectrum is found in
the yellow series of cuprous oxide where the n = 1 line, although not
absent, is extremely weak. The residual intensity of the n = 1 line
can be explained by electric quadrupole transitions which are much
106 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

.(hw-Eo) (eV)

Figure 6.4 Exciton effects near direct interband threshold in germanium. Ab-
sorption coefficient K (= 4nK/A o), photon energy liwo, direct interband threshold
energy Eo. Dashed curve: theoretical interband absorption neglecting electron-
hole interaction. Full curve: experimental curve at 20° K, which agrees with
theory including electron-hole interaction and exciton formation.
Reference: T. P. MCLEAN, Progress in Semiconductors, 5, 54-102 (1960).

weaker than allowed dipole transitions but have different selection


rules.
The calculation of absorption associated with indirect transitions
in the band model, has been outlined in Section 3.5. These results
are also modified when the electron-hole interaction is included. As
in Section 3.5, we will assume that the valence band maximum is at
k = 0 and the conduction band minimum at k = k c ; the indirect
energy gap is Eg. The indirect transitions are phonon-assisted from
valence band states around k = 0 to conduction band states near
k = k c• When the exciton bands are considered it is seen that the
bands near K = kc can also act as final states for indirect transitions.
The selection rule that K does not change in a transition is relaxed
because phonons can provide the balance of momentum. There are
no absorption lines associated with indirect transitions to exciton
states because transitions can occur to any part of the exciton bands.
The threshold photon energy for indirect transitions to exciton
states is:
(6.16)
EXCITON EFFECTS 107
where E1) is a phonon energy and the ± signs correspond to phonon
emission or absorption. The function F(E) in (3.46), which gives the
shape of the absorption edge, is changed by the electron-hole inter-
action. When E ~ Eb, F(E) increases as Ei for allowed transitions
and as E-t for forbidden transitions.

6.5 Spatial dispersion


Let us consider the optical dispersion in a crystal with a strong
exciton transition at a frequency wo. The optical behaviour of the
crystal can be represented by a classical model of a distribution of
electric dipoles with resonant frequency wo. It was assumed in
Section 6.2 that the frequencies of transitions to exciton states are
given with sufficient accuracy by the approximate selection rule
K = O. In fact, some interesting phenomena arise from the K-
dependence of the exciton energy which has been neglected. When
the small but non-zero photon wave vector ko is not neglected, the
selection rule becomes:
K= ko = k. (6.17)
Assuming a simple exciton band as given by (6.5), the energy of a
transition is:
(6.18)
The corresponding resonant frequency (Eln) contains a term
dependent on k. The classical model can still be used to discuss the
optical behaviour of the crystal if the k-dependence of the resonant
frequency is included.
The classical formula for the dielectric constant e(k,w), neglecting
damping, is:
e(k,w) = e(oo) + (4nNe2jlm)/(wo2 - w2 + Bk2). (6.19)
B = nWolM* and a term in k4 in the denominator has been neglected;
f is the oscillator strength of the transition; e( (0) is the dielectric
constant of the crystal for frequencies well above wo. The appearance
of kin (6.19) shows that spatial dispersion has been included in the
theory. The transverse wave solutions of Maxwell's equations
satisfy the dispersion relation:
(c 2k2/w2) = e(k,w). (6.20)
Substitution of (6.19) into (6.20) leads to a quadratic equation for k 2
in terms of w 2 and the other parameters. For each value of w, there
108 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

are two values of k representing waves of different phase velocities


(w/k). The curves in fig. 6.5 represent (c 2k2/W2) against (w/w o),
calculated from (6.19) and (6.20) with hypothetical values for the
parameters f, Wo and B. At frequencies below wo, the upper branch
represents a propagating wave with phase velocity approximately
C/ei(O), while the lower branch has (c 2k2/W2) < 0 and represents a
damped wave. At frequencies above wo, the lower branch represents

101=--------

"'I'"
~ 3
u
O~--------------_+~~----------__1

-10

(w1wol
Figure 6.5 Coupled exciton-photon dispersion curves. Exciton energy at zero
wave number /iwo. Calculation from (6.19) with hypothetical parameters:
8(00) = 5·00, (4nNe2jlm) = 0.01010', Blc = 1·0 x to- 5•

a propagating wave with phase velocity approximately c/ et ( 00),


while the upper branch represents a second propagating wave with a
large exciton component. The two propagating waves for w > W o
have the same polarization and can therefore interfere with each
other. The difference in phase velocity will make the transmission of
a thin layer of crystal have an oscillatory dependence on thickness.
A phenomenon of this sort has been observed in thin crystals of
anthracene, thickness 500 A to 3000 A, in the region of an exciton
absorption line. When exciton relaxation processes are included in
EXCITON EFFECTS 109
the theory, it is found that the conditions for observing the inter-
ference effect in transmission -are not easily satisfied in practice.
To calculate the transmission of a thin layer of crystal or reflection
by a crystal surface, it is necessary to define the boundary conditions.
Since there are two propagating waves in the crystal, the usual
electromagnetic boundary conditions are not sufficient. An addi-
tional boundary condition is provided by the behaviour of an
exciton at a surface. There is a resemblance here to the anomalous
skin effect, in which spatial dispersion is important also and the
boundary conditions include a specification of electron reflection at
06.-----------,

o
a:

0-2

(hw-2-550) (eV)

Figure 6.6 Normal reflectance of a crystal of cadmium sulphide at 4° K in the


region of an exciton line.
Reference: J. J. HOPFIELD and D. G. THOMAS, Phys. Rev., 132, 563-72 (1963).

a surface. Some experimental reflection spectra of crystals in the


region of an exciton line can be explained by postulating a repulsive
potential for excitons near a surface. This potential effectively ex-
cludes excitons from a barrier layer of thickness t at the surface. The
reflectance is calculated by assuming a layer of dielectric constant
e( CX) and thickness t between the external medium and the inner
region of the crystal having spatial dispersion. The theory is able to
explain experimental reflectance curves in the region of an exciton
transition, like the curve in fig. 6.6 for a cadmium sulphide crystal.
The normal reflectance Ro calculated from classical theory with no
damping and no spatial dispersion, rises to 100% in the main peak.
110 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

Damping and spatial dispersion both reduce the maximum Ro but


the main effect in cadmium sulphide seems to be due to spatial
dispersion. Also, the subsidiary peak in Ro for the frequency at
which [; = 0 in the lower branch, is predicted by the barrier layer
theory but not by simple classi~al theory. Recent developments in
the theory of optical effects associated with excitons use the concept
of polaritons: excitations composed of coupled photons and excitons.
Thermal broadening effects can then be treated in terms of polariton-
phonon interactions.
7
Non-linear Optics
7.1 Classification of non-linear effects
We have assumed in previous chapters that the electric current
density and polarization in a medium are linear functions of the
electric field of an electromagnetic wave. The linear optical properties
are defined in terms of these functions. The justification for the
linear theory is that the applied field is much less than the atomic
fields which act on an electron in a solid. An electromagnetic wave
can therefore be treated as a small perturbation for which linear
theory is an accurate approximation. We can estimate the order of
magnitude of the atomic field acting on a valence electron in a solid
as (e/a 2) e.s.u. where a is a typical interatomic distance. Assuming
a = 3 A, this gives a field of about 10 8 volt cm- l • The field pro-
.duced by an incoherent light source is such a small fraction of this
atomic field that any non-linear response by the medium is un-
observably small. With the advent of the laser, however, the experi-
mental possibilities have been transformed. It is possible to have
fields of about 106 volt cm-l in the focussed beam of a high-power
laser. This is still only a small fraction of an atomic field, so the
non-linearities are unlikely to be large. However, a small non-
linear response by a single atom or lattice cell can lead to large
effects when many atoms or cells act coherently. The heat produced
by a laser beam sets a practical upper limit to the field which can be
applied without damaging the medium.
The non-linear response of a medium to an electromagnetic wave
can be represented by a distribution of electric current density and
polarization, as in the linear case. It is usual to consider principally
the electric dipole polarization P of the medium. The current density
J is related to P by:
J = oP/ot. (7.1)
It is sometimes necessary to include also a term representing the
quadrupole polarization of the medium, when the symmetry of the
111
112 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

medium requires that the non-linear dipole polarization be zero.


Let us consider first a highly simplified theory to see qualitatively the
sort of effects caused by non-linearities. We will assume that the
polarization pet) and electric field E(t) can be treated as real scalar
functions of time t, and dispersion will be neglected. If the non-
linearities are not too large, P(t) can be expressed as a power series
in E(t):
Pet) = aIE(t) + a2£2(t) + aaE3(t) + . . . (7.2)
Let us suppose that two laser beams of frequencies WI and W 2 are
passing simultaneously through the medium and the total field is:
E(t} = Al COS (WIt - ~l)
i .. ,,,,
+ Aa cos (W2t - ~2).
The second order term a 2£2(t) can be expressed as the sum of
sinusoidal terms at frequencies 0, (WI - wJ, (WI + wil), 2WI and
2w 2• The non-linearity produces combination frequencies and har-
monics of the original frequencies. The higher order terms produce
more combination frequencies and higher harmonics.
In the exact formulation of linear optics, the field is analysed into
a Fourier spectrum, and a linear electric susceptibility tensor Xii(W)
is defined by: '
(7.3)
\, Suffixes i and j refer to the different space components of the vectors
pew) and E(w); the convention of summation over repeated suffixes
is assumed. The field component E;(w) = Eo; exp (-iwt); X;;(w)
can have complex elements to represent the phase difference between
P;(w) and E;(w). Formula (7.3) allows for dispersion and crystal
structure but neglects spatial dispersion. A similar theory has been
developed for the non-linear effects by defining the Fourier com-
ponents of P for each non-linear mixing process. Each Fourier
component of P is related by a susceptibility tensor to the relevant
combination of field components. When considering non-linear
effects it is important to represent each Fourier component of the
field by a real function. For the component of frequency w we will
assume:
(7.4)
This can be expressed in terms of exponential functions:
Eiw) = !A; exp (i~;) exp (-iwt) + !A; exp (-i~;) exp (iwt)
_Ej(w) +
E;*(-w). (7.5)
NON-LINEAR OPTICS 113
Similarly for the polarization, linear or non-linear:
P;(w) = P;(w) + P;*( -w). (7.6)
An example of a second order non-linear effect is the generation of
a difference frequency from two waves of frequencies WI and W2. The
non-linear polarization at the difference frequency W (= WI - (2)
can be written:
Pi(W) = Xiik(W = WI - wJE;(w 1 )Ek*(-W 2)
+ XUk( -w = -WI + ( 2)E;*( -(1)Ek(wJ. (7.7)
The susceptibility for this process is represented by the third order
tensor Xiik(W = WI - wJ. Since Pi is a real number:
Xi;k( -w = -WI + ( 2) = X:k(W = WI - ( 2). (7.8)
Writing XUk as a complex number (X' + iX") and omitting subscripts,
we have:
X'( -w) + iX"( -in) = X'(w) - iX"(w). (7.9)
(7.9) holds for all susceptibilities, linear or non-linear. In linear
optics a non-zero X"(w) is associated with absorption at frequency w.
In general, the average power absorbed by a medium can be calcu-
lated from:
(7.10)
W is zero unless W = WS' In the example of difference frequency
generation given above, there is no average absorption of energy due
to the interaction ofthe driving fields E(W1) and E(w 2) with the non-
linear polarization pew).
Some care is needed in defining the second order non-linear
susceptibility in the limiting case when WI and W2 are equal. The
equation for the non-linear polarization at the difference frequency:
Pi(W) = Xi;k(W = WI - ( 2)E;(W1)Ek*(-W2) (7.11)
becomes formally identical with the formula for the zero frequency
non-linear polarization:
Pi(O) = Xiik(O = WI - ( 1)E;(w 1)Ek*( -WI)' (7.12)
An analysis of the transition from (7.11) to (7.12) shows that the
two susceptibilities are not equal but:
Xi;k(W = WI - ( 2) -+ 2Xiik(O = WI - wJ (7.13)
as WI -+ W2 • Similarly, the sum frequency and second harmonic
generation become identical as WI -+ W2 but:
XUk(W = WI + wJ -+ 2Xiik(W = WI + WI)' (7.14)
114 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

Third order non-linearities can produce combination frequencies


of up to three original frequencies. From a single laser wave of fre-
quency Wl> a third order non-linearity can generate the third har-
monic 3WI' The non-linear polarization at the third harmonic
frequency is:
Pi(3wI) = Xiik,(3wI = WI + WI + wl)Ej(WI)Ek(WI)E,(w l ). (7.15)
A fourth order tensor Xiik' must be used to represent the suscepti-
bility for a third order effect. Among the many combination fre-
quencies which can be produced by third order non-linearities, one
interesting possibility is the generation of non-linear polarization at
the frequency of one of the driving fields. For example:
Pi(WI) = Xiik,(WI = WI + W2 - wJE;(w l )Ek(w 2)E,*( -wJ. (7.16)
The real parts of the tensor elements, X:jkb represent a change in the
refractive indices at frequency WI' proportional to the power at
frequency W2' The imaginary parts, X;;kl> represent absorption or
generation of power at Wl> proportional to the power at W 2• From
the formula (7.10) for the average power absorption:
W = iwl{Ei(wl)Pl( -WI) - El( -WI)Pi(WI)}
= 2WIX;;kl(WI)Ei(WI)Ej*( -wI)Ek(wJE,*( -(2)' (7.17)
A positive X;;kl(WJ represents absorption of power at WI while a
negative X;;kl(WI) represents generation of power at WI' When
several waves are interacting, power absorbed from one wave need
not be changed into heat but can increase the p.9wer in another wave.
The general third order tensor Xi;k has (3)3 elements and the fourth
order tensor Xi;kl has (3)4 elements, but symmetry considerations can
be used to reduce the number of independent elements. The suscepti-
bility tensors must. conform to the symmetry of the medium as
expressed by its point group. It is convenient to choose reference
axes for the tensors, which coincide with symmetry axes of the
medium. In. some cases it can be shown by symmetry that some or
all elements of a particular susceptibility tensor must be zero. For
example, if the point group symmetry includes a centre of inversion,
all second order non-linear effects are zero. A reversal of sign of all
components of E in a formula like (7.7) for second order polarization,
leaves Pi(W) unchanged, which is not compatible with inversion
symmetry unless Xiik = O. The fourth order tensors which represent
NON-LINEAR OPTICS 115
the third order non-linear effects, are not identically zero for any
point group symmetry. Third order effects can occur in an isotropic
medium. For example, let us consider the non-linear polarization:
Pi(wJ = XiikZ(WI = WI + WI - wJEi(wJEk(wI)Ez*( -WI). (7.18)
This represents changes in refractive index and absorption induced
by a single laser wave. From considerations of symmetry, P(w) and
E(w) must be parallel or anti-parallel in an isotropic medium. Only
two combinations of E(w l ), E(wI) and E*( -WI) satisfy this con-
dition and P(WI) can be expressed in the form:
P(WI) = IXE(wl){E(wI).E*(-wI)}
+
IX'E*( -wI){E(wI).E(wI)}. (7.19)
All the non-zero elements of XiikZ(W I = WI + WI - WI) can be ex-
pressed in terms of IX and IX'.

7.2 Non-linear susceptibilities


In the previous section we considered the classification of non-linear
effects and the definitions of the associated susceptibility tensors.
We now turn to a consideration of the factors which determine the
magnitude of a susceptibility tensor for a given non-linear effect. To
calculate a non-linear susceptibility, one must treat the interaction
between an electromagnetic wave and a solid in a higher approxima-
tion than the first order linear theory. Formulae for the suscepti-
bilities can be derived by quantum-mechanical perturbation theory
and the reader is referred to the general references for details of these
calculations. The formulae are complicated and require a knowledge
of the wave functions of the solid for their evaluation.
A simple classical model can be used to give estimates of the
magnitudes of various non-linear susceptibilities. We saw in Section
1.7 that a classical oscillator model is a useful guide to the linear
optical properties of a medium. If the oscillators are assumed to have
a small degree of anharmonicity, then this model is also useful for
the discussion of non-linear effects. The equation of motion of a
classical electron oscillator with an anharmonic restoring force, can
be written:
(d 2x/dt 2) + y(dx/dt) + wo 2x + f3x 2 = -(e/m)E.,(t). (7.20)
116 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

- {JX2 represents an anharmonic term in the restoring force; only the


motion in the x-direction will be considered. The real electric field
Ea;(t) consists of one or more sinusoidal fields of the form:
Ea;(w) = Ea;(w) + Ea;*(-w) (7.21)
with the same definitions as (7.5). In the linear approximation
(fJ = 0) the electric polarization Pa;(w) and the displacement x(w)
due to Ea;(w) are given by:
Pa;(w) = -Ne x(w) = (Ne 2/m)Ea;(w)/ D(w) (7.22)
where D(w) = {wo - £02 - iyw}. This is the Lorentz oscillator
model of Section 1.7. When {J is non-zero but small, the displacement
will not differ greatly from x(w) given by (7.22). The anharmonic
term in (7.20) is approximately {Jx 2 with x given by the linear theory.
(_{JX2) can then be considered as an extra driving force on the right
hand side of (7.20). Let us suppose that Ea;(t) is the sum of two
sinusoidal terms at frequencies WI and £0 2• The total linear displace-
ment is:
x = X(Wl) + x*( -WI) + x(W 2) + x*( -(02) (7.23)
with x(w) as defined by (7.22). The anharmonic term {JX2 contains
terms at the sum, difference and second harmonic frequencies. The
non-linear polarization Pa;(WI + (0 2) and thedisplacementx(wl + (02)
at the sum frequency (WI + (0 2) are given by:
Pa;(WI + (02) = -Ne x(w 1 + (0 2)
= 2(fJNe 3 /m 2)E,,(Wl)Ea;(w 2)/ D(Wl)D(wJD(w l + (0 2 ). (7.24)
The corresponding non-linear susceptibility is:
Xzzz(w = WI + (0 2) = 2({JNe 3 /m 2 )/ D(Wl)D(W2)D(WI + (02). (7.25)
The form of (7.25) indicates that the susceptibility will be enhanced
if one of the frequencies Wb £0 2 or (WI + (0 2) lies near the resonant
frequency £0 0 • The non-linear susceptibility (7.25) can be related to
linear susceptibilities derived from (7.22). For example, when
WI = £02' (7.25) gives the susceptibility for second harmonic genera-
tion and it can be related to the linear susceptibilities at WI and 2£01:
Xza;a;(2Wl = WI + WI) = ({JNe 3/m2)/{D(Wl)}2D(2wl)
= ((Jm/N2e3){Xa:z(Wl)}2Xa:z(2wl). (7.26)
From (7.26) the magnitude-ratio of non-linear polarization at 2£01
to linear polarization at WI is:
(fJm/N2e3) I Xa;z(Wl) II Xa;z(2Wl) II E Z (w 1) I. (7.27)
NON-LINEAR OPTICS 117
The linear susceptibility X....((0) is of the order of 0·1 for transparent
crystals in the visible and ultra-violet but it can be of the order of
1·0 for semiconductor crystals in the infra-red. The order of magni-
tude of f3 can be estimated by assuming that the linear force W02X and
the non-linear force f3x2 are equal when x is equal to the interatomic
distance. This estimated value of f3 gives:
({Jm/N 2e3)-1 106 e.s.u. = 3 X 10 8 V cm-l •
f"OooI (7.28)
Some values of ({Jm/N2e3)-1, calculated from experimental suscepti-
bilities, are given in Table 7.1. The order of magnitude is in good

TABLE 7.1: PARAMETER RELATING FUNDAMENTAL AND SECOND


HARMONIC POLARIZATION

Crystal Quartz KDP GaAs Te

0'7 1·4 0'5 X 106 e.s.u.

agreement with (7.28). The field in a laser beam does not usually
exceed about 0·3% of (7.28) and the ratio of non-linear to linear
polarization is therefore small:
The third order non-linear effects can be considered as the result
of mixing between the fundamental frequencies and the frequencies
produced by second order effects. For example, non-linear third
order polarization at frequency WI can be produced by mixing Ws
and (WI - 0( 2), or (-00 2) and (WI + 0( 2), Let us suppose that the
difference frequency (WI - 0( 2) is near the resonant frequency 00 0 so
that X...,..(w = WI - w s) is much larger than other second order
susceptibilities. The non-linear third order polarization at WI is:
P(WI) = -Ne X(WI = Ws + WI - 0( 2)
= 2Nef3 X(W2)X(W = WI - (2)! D(w l ).
The corresponding susceptibility is:
X...,....(WI = WI + Ws - w s)
=2 ({J2Nel /m 3)/{D(Wl)}2 D(wJD( -WS)D(WI - ws). (7.29)
Formula (7.29) can alternatively be expressed in terms of linear
susceptibilities:
X........ (WI = WI + Ws - w s)
= 2({Jm/N2e3)2{X....(Wl)}2Xz.-(Ws)X ....(-ws)X ....(WI - ws). (7.30)
118 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The third order susceptibility contains an extra factor (fJm/N2e3)


compared with the second order susceptibility (7.26). When (Wl - w 2)
is near to the resonant frequency wo, X.,,,(w l - w 2) will have a
positive imaginary part corresponding to the absorption of electro-
magnetic energy, assuming Wl"> W2. If W l and W 2 are sufficiently
far from wo, X.,a;(w) at these frequencies will be real, corresponding

Figure 7.1 Imaginary parts of linear and third order non-linear susceptibilities
according to modelof Section 7.2. Variable frequency w, fixed frequency WL,
resonant frequency of electron oscillator wo, WB = WL - Woo Vertical scales are
arbitrary; XU > 0 represents absorption, XU < 0 represents generation.

to a solid which is transparent at Wl and W2. Under these conditions,


interchanging Wl and W2 in (7.29) and using
XxiW2 - Wl) = Xxx *(Wl - W2),
we have:
Xxxxx(w 2 = W2 + Wl - Wl) = X!xxx(W l = W l + W2 - W2). (7.31)
Hence, X.xxx(W2 = W 2 + W l - Wl) has a negative imaginary part
corresponding to the generation of electromagnetic energy at fre-
quency W2. This third order effect causes a transfer of power from a
wave at the higher frequency W l to a wave at the lower frequency W2.
The relation between the imaginary parts of the linear and non-
linear susceptibilities is shown in fig. 7.1.
NON-LINEAR OPTICS 119
7.3 Second harmonic generation
The Fourier component of non-linear polarization at a frequency ru
radiates an electromagnetic wave of the same frequency. The propa-
gation of this wave is governed by the linear optical properties of
the medium at frequency ru. For example, the second order non-
linear polarization PNL(rul + ruJ, produced by mixing two plane
waves of frequencies rul and ru2, and wave vectors kl and k2' has a
space-time dependence given by:
exp {i(k 1 + k 2).r - i(ru 1 + ruJt}, (7.32)
where kl = (rudc)et (rul); k2 = (ru 2/c)et (ruJ. The dielectric constant
eCru) represents the linear optical properties of the medium and is
related to the linear susceptibility Z(ru) by:
eCru) = 1 + 4nz(ru). (7.33)
The magnitude of the wave vector of a propagating wave at fre-
quency (rul + ruJ is:
k = {{rul + ruJ/c}et(rul + ruJ. (7.34)
In general k and I (kl + kJ I are not equal; this means that the
linear polarization pL(rul + ruJ and the non-linear polarization
pNL(rul + ruJ oscillate in and out of phase as the waves propagate.
The propagation of waves created by non-linear polarization can
be treated quantitatively by including a term for pNL in the definition
of D(ru), the Fourier component of the electric displacement at
frequency ru:
D(ru) = E(ru) + 4:n:{pL(ru) + pNL(ru)}. (7.35)
From Maxwell's equations (1.17), assuming B = H, we have:
V x V x E = -(1/c2)(o2D/ot 2).
Hence:
V x V x E - (ru 2/c 2)e(ru)E(ru) = (4:n:ru 2jc2)pNL(ru) (7.36)
An equation of the form (7.36) can be written for each Fourier
component of the field. The non-linear polarization pNL(ru) depends
on fields at other frequencies and therefore couples these equations
together. In many experiments the conversion of power from the
original frequencies to the mixed frequencies is very small. It is then
a good approximation to neglect the depletion of the original waves
and to consider their amplitudes as fixed parameters. Each equation
like (7.36) can then be solved separately.
I
120 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

The general solution of (7.36) consists of a solution of the linear


wave equation with PNL(W) = 0 and a particular integral corres-
ponding to the specified pNL(W). As an example, let us consider
second harmonic generation in a uniaxial crystal, such as KDP,
with the experimental geometry shown in fig. 7.2. The non-zero

Figure 7.2 Fundamental (00.) and second harmonic (200.) waves near the surface
of a KDP crystal. Wave vectors and polarizations of waves are shown; 0
represents a vector parallel to Oy. The optic axis of the crystal is assumed to lie
in the zOx plane.

elements of the susceptibility tensor for second harmonic generation,


referred to the optic axis as the z-axis, are:
X~z = XZZ1J = X"zal = X"zz; Xza:u = X",:/:· (7.37)
With the geometry of fig. 7.2, the crystal y-axis can be chosen to
coincide with the y-axis defined as the normal to the plane of
incidence. The electric field of the fundamental wave has no
y-component, E,,(Wl) = 0, and therefore by (7.37), the non-linear
polarization pNL(2wl)' is in the y-direction. From (7.32):
P"NL(2w 1 ) = poNL exp (2ik 1T .r - 2iwlt). (7.38)
For a plane wave with wave vector k, (7.36) becomes:
k 2E(w) - (w 2/c 2)n 2(w)E(w) = (4:nW2/C 2)pNL(W). (7.39)
The medium is assumed transparent and the refractive index n( w)
has a real value. The general solution of (7.39) for the field in the
NON-LINEAR OPTICS 121

crystal at the second harmonic frequency, W = 2w l , is:


Ey(2wl) = E2T exp (ik2T.r - 2iwlt)
(l6nw 2 /c 2)
+ (2klT)2 _1 (k2T)2 exp (2ikl .r - 21wlt)

(7.40)
where kl = (wdc)n(w l ), k2T = (2wdc)n(2w l ). The directions and
amplitudes of transmitted and reflected waves can be determined
from the boundary conditions for the fields at the surface of the
crystal. From the phase relations between the various waves, one
finds:
fJ l R = fJ 2R = fJ; sin fJ = n(wl) sin fJ l T = n(2w l ) sin fJ 2T • (7.41)
From the continuity of Ey(2wl) at the surface:
E2R = E2T + (16nwl2/c 2)PoNL/{(2kl)2 - (k2T)2}. (7.42)
The continuity of Hx(2wl) gives another equation which can be
combined with (7.42) to give exact expressions for E2R and E2T. In
practice E2R is often much smaller than E2T and it is a reasonable
approximation to put E2R = 0 in (7.42). With this approximation
and assuming normal incidence (fJ = 0), (7.40) and (7.42) can be
combined to give an expression for 1Ey (2w l ) 12:
2 _ (8nP ONL ) 2 (W l 2/C 2) sin 2 {(~k)z}
1 Ey(2wl) 1 - {n(2wl) + n(wl)}2 (~k)2' (7.43)
where ~k = (wdc){n(2w l ) - n(wl)}. {(~k)z} represents the phase
mis-match between the linear and non-linear polarization at the
second harmonic frequency. When ~k = 0:
sin 2 {(~k)z}/(~k)2 = Z2. (7.44)
The intensity of the second harmonic grows in proportion to Z2;
under this condition the assumption of constant fundamental in-
tensity does not remain valid. When ~k ~ 0, the second harmonic
intensity oscillates as a function of z between zero and a maximum
value. The maximum value is equal to the intensity produced by a
thickness tc when ~k = 0, where:
tc = (1/~k) = (A o/2n)/{n(2w l ) - n(wl)}, (7.45)
tc is the coherence length; Ao is the vacuum wavelength of the funda-
mental wave. In a typical crystal {n(2w l ) - neWt)} is of the order of
0·01 so tc '?> Ao. The reflected second harmonic wave is effectively
produced by the non-linear polarization in a surface layer about
122 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

half a wavelength in thickness. The neglect of E2R in calculating E2T


from (7.42) is therefore usually justified. In the case of an absorbing
crystal, such as gallium arsenide in the visible spectrum, it is simpler
experimentally to measure second harmonic generation in the re-
flected wave. A similar calculation to that outlined above gives the
dependence of reflected second harmonic intensity on the incident
wave polarization and the crystal orientation.
The formulae derived so far in this section have not included any
explicit reference to the double refraction of the crystal medium.
The second harmonic wave in the medium has its electric vector in
the y-direction, perpendicular to the optic axis. This means that the
second harmonic wave is an ordinary wave with an index no in-
dependent of the direction of propagation. On the other hand, the
transmitted fundamental wave has EuCw) = 0 and is an extraordinary
wave with an index ntb which depends on the angle 4> between the
wave vector and the optic axis. In some uniaxial crystals, for example
KDP, it is possible to find an angle 4> for which:
(7.46)
where WI is a frequency in the red or near infra-red. When the index
matching condition (7.46) is satisfied, a large fraction of a pulsed
laser beam can be converted to second harmonic. A measurable
second harmonic conversion factor can even be obtained under
index matched conditions, with continuous laser beams of a few
milliwatts power. The variation of second harmonic intensity with
phase and index mis-match is shown in fig. 7.3.
According to the symmetry conditions outlined in Section 7.1, the
second order non-linear susceptibility for second harmonic genera-
tion is identically zero in crystals with a centre of inversion. How-
ever, experiments on calcite crystals, which have a centre of inversion,
show some second harmonic generation from focussed laser beams
along an index matched direction. The intensity is about 10- 5 of that
from KDP under similar conditions. This effect in calcite can be
explained by the electric quadrupole polarization of the crystal. An
electric dipole moment is the first term in a series which represents
the electrical properties of a distribution of balanced positive and
negative charges. Although the second harmonic dipole polarization
must be zero in a crystal with a centre of inversion, the same sym-
metry conditions do not apply to the electric quadrupole polariza-
NON-LINEAR OPTICS 123

I
z2

Figure 7.3 Phase matching in second harmonic generation. Phase mis-match


tlk, index mis-match !1nq" second harmonic intensity I, path length z. Experi-
mental data for KDP: index matching angle </>0 = 41°, angular deviation 0·06°
for !1nq, = tAo/Z (Ao = 1·15 {tm, Z = 1·23 em).
Reference: A. ASHKIN, G. D. BOYD, and J. M. DZIEDZIC, Phys. Rev. Lett., 11,
14-17 (1963).

tion. The non-linear quadrupole polarization tensor QNL(W) can be


defined in terms of a susceptibility tensor 'YJiikl and the applied field
E(w 1 ) :
(7.47)
The quadrupole polarization contributes an extra term to the wave
equation (7.36) for the second harmonic wave. The right hand side
of (7.36) becomes:
(4nw 2 / C2){pNL(W) _ V. QNL(W)}. (7.48)
The quadrupole term {V. QNL(W)} is generally of the order of (a/A)
times the dipole term {pNL(W)}, where a is a typical interatomic
distance. Since (a/A) "" 10- 3, this means that quadrupole effects are
unobservable except when the dipole effect is forbidden.
When a steady electric field is applied to a crystal with a centre of
inversion, the symmetry conditions are modified and dipole genera-
tion of second harmonic is possible. A steady field can be included
124 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

in the theory by representing it as a field of zero frequency, E(O).


Second harmonic generation by a crystal subject to a steady field,
can then be considered as a third order effect with a susceptibility XiikZ
defined by:
P;(2W l) = XiikZ(2wl = W l + W l + O)Ej(wl)Ek(Wl)Ez(O). (7.49)
Some experimental values for quadrupole and field-induced second
harmonic generation in calcite are shown in fig. 7.4. In the field-
induced effect the second harmonic power is proportional to 1 E(O) 12.

OL---~----~----~----L-----L- __~

£(0) (kVlcm)
Figure 7.4 Second harmonic generation in calcite. Steady polarizing field E(O),
second harmonic power W. Focussed laser beam of 48·5 mW power along index
matched direction. QNL and pNL indicate contributions of second order quad-
rupole and third order dipole polarization to W.
Reference: J. E. BJORKHOLM and A. E. SIEGMAN, Phys. Rev., 154, 851-60 (1967).

A non-zero quadrupole term in (7.48) for a crystal with a centre of


inversion, means that a reflected wave will have a second harmonic
component. Gallium arsenide and germanium are two crystals with
similar linear optical properties, but germanium has a centre of
inversion while gallium arsenide does not. Under similar conditions
the reflected second harmonic intensity from a germanium surface is
about 10-3 times that from a gallium arsenide surface. This agrees
with the ratio of quadrupole to dipole effects which is of the order
NON-LINEAR OPTICS 125
of (a/b)2 for an absorbing medium, where b is the second harmonic
skin depth. Some experimental second harmonic conversion factors
are given in Table 7.2. They indicate the fraction of the incident
intensity converted to second harmonic reflected intensity with a
pulsed beam (J,...., 10 7 W cm- 2) from a ruby laser. The value of
the conversion factor depends on the polarization of the funda-
mental and second harmonic waves and the angle of incidence,
so the values in Table 7.2 should be taken only as a guide to the
order of magnitude.
In an isotropic medium or a cubic crystal with a centre of in-
version, the quadrupole term {V. QNL(W)} has two transverse
components which are zero within the medium and a longitudinal

TABLE 7.2: SECOND HARMONIC REFLECTION FROM MEDIA WITH INVERSION


SYMMETRY

Material Si0 2 (glass) NaCl Ge Ag

Power conversion factor 0·8 0·4 100 3900 X to- 16

Reference: c. C. WANG and A. N. DUMINSKI, Phys. Rev. Lett., 20, 668-71


(1968).

component which does not couple with a propagating wave. How-


ever, at a surface there is a sudden change in QNL(W) and the
source term {V. QNL(w)} is non-zero. In the limit as the change in
QNL(w) is assumed to be a mathematical discontinuity its normal
derivatives become delta functions. {V. QNL(W)} is then equivalent
to a surface dipole polarization pS(w). The contribution of pS(w)
to the reflected second harmonic wave can be calculated by modifying
the boundary conditions for the second harmonic fields to allow for
pS(w). At a surface the external waves can also couple to the longi-
tudinal component of {V. QNL(W)} and this will contribute to the
reflected second harmonic wave. Calculations of the quadrupole
susceptibilities for insulating crystals indicate that they are approxi-
mately proportional to the square of the low-frequency linear
susceptibility. This explains the large difference between Ge and
NaCI in Table 7.2. For metals both the valence and core electrons
contribute to the quadrupole polarization; in the case of silver and
126 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

gold these contributions are of comparable magnitude for funda-


mental frequencies in the visible spectrum. Since the effective surface
polarization pS(w) is determined by the properties of the first few
atomic layers at the surface, the investigation of reflected second
harmonic waves may give information about the electronic structure
of the surface layers. The limited experimental data available at the
moment agrees with the assumption of a discontinuous change
between bulk properties at the surface.

7.4 Parametric amplification and oscillation


The theory given in the previous section suggests that a large fraction
of a coherent wave can be converted to second harmonic when the
phase matching condition (11k = 0) is satisfied. Similarly with other
non-linear processes, a large effect is expected when a phase matching
condition is satisfied. In practice, phase matching can be used to
select the dominant non-linear process in a given experimental
situation. When non-linear processes cause considerable transfer of
power between different waves, it is no longer valid to treat the
amplitudes of the driving waves as constants. An equation like (7.36)
can still be written for each Fourier component of the field but the
equations must be solved simultaneously. Let us consider the equa-
tion for the Fourier component of field labelled with subscript s:
V X V x E(w s) - (ws2/C2)e(ws)E(ws) = (4nw s2 /c 2 )PNL(w s). (7.50)
To simplify the mathematics, let us assume that the k-vectors of all
the waves are parallel and in the direction of Oz; E(w s ) is assumed
parallel to Ox or Oy which are principal axes of the dielectric tensor.
Under these conditions:
(7.51)
The field and the non-linear polarization at Ws can be written:
E(w s) = E(w.,z) exp {iksz - iWst}; (7.52)
pNL(w s) = pNL(W.,Z) exp {i(ks - I1k)z - iWst}, (7.53)
where ks = (Ws/C)et(Ws). Using (7.52) and (7.53), equation (7.50)
can be reduced to:
dE(w.,z)/ dz = {2nik s/ e(Ws)}pNL(W.,z) exp (-il1kz). (7.54)
NON-LINEAR OPTICS 127
A term {d SE(w.,z)ldz 2} has been neglected on the assumption that
the relative change in E(w.,z) per wavelength is very small.
As a simple example, the coupled equations for second harmonic
generation will be solved for the phase matched condition, !:1k = O.
The simplifying assumptions of the previous paragraph correspond
to propagation in a uniaxial crystal, in a direction perpendicular to
the optic axis. Index matching in this direction (n O(wl) = n.(2w 1)
for q, = n12) is possible in crystals of lithium niobate. The theory is
also a reasonable approximation for other crystals, such as KDP, in
which the index matching direction is not perpendicular to the optic
axis. The coupled equations of the form (7.54) are:
dE",(WbZ)ldz =
{2nik 1/e=(wJ}XX1/",(Wl = 2001 - wl)Ey{2wl,Z)E",*(-WbZ);
(7.55)
dE,i2wbZ)ldz =
{4nikt/eyy(2wl)}xyu(2wl = WI + wl)E",(w 1,z)E",(Wl,Z).
The index matching condition is: eu t (Wl) = eyy!(2w 1 ). If the crystal
is transparent for frequencies between WI and 20010 then the suscepti-
bilities satisfy a symmetry relation:
XX1/",(W 1 = 2001 - WI) = 2XYU(2wl = OJ 1 + WI)' (7.56)
If we put:
E",(w 1,z) = A 1(z); Ey(2wbZ) = iA 2(z) (7.57)
then equations (7.55) reduce to:
dA 11dz = -IXA 1A 2; dA 2/dz = iXA12, (7.58)
where IX = {4nkt/eyy(2w1)}XYU(2wl = WI + WI)' Assuming that the
initial second harmonic amplitude is zero, the first equation of (7.58)
represents the depletion of the fundamental wave and the second
equation represents the growth of the second harmonic wave. An
integral of (7.58) is:
A I2(Z) + As2(Z) = AI2(0), (7.59)
which expresses the conservation of electromagnetic energy. The
solutions of (7.58) are:
A 2(z) = A 1(0) tanh (zit); A 1(z) = A 1(0) sech (zit), (7.60)
where ( = {IXA 1(0)}-1; (is called the interaction length. In terms of
the vacuum wavelength Ao and the refractive index n (= ellyi (2w 1»:
( = (Aon)/{8n 2A 1(0)Xyu(2w 1 = WI + wJ}. (7.61)
JO
128 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

o
zll
Figure 7.5 Interaction of fundamental (WI) and second harmonic (2Wl) waves.
Intensity J, path length z, interaction length I.

For lithium niobate, assuming Xyxx = 1-4 X 10- 8 e.s.u. for Ao = 1·06
pm, and 2A 1(0) = 3 X 10 5 V cm- 1 , formula (7.61) gives t = 0·3 cm.
The variation of A12 and A22 with z is illustrated in fig. 7.5.
Besides the generation of harmonics of a laser wave, another
possible use of non-linear effects is in the generation of lower fre-
quencies from a coherent pump wave. Second order non-linearities
can be used to mix three waves of frequencies WI' w 2 and W3 such
that:
(7.62)
If WI is the pump frequency, waves are generated at the lower
frequencies W 2 and W3. The phase mis-match between the waves can
be represented by the wave vector difference:
(7.63)
The largest effects will occur for waves which make f)"k = O. Under
usual experimental conditions, the conversion of the pump power to
lower frequencies is sufficiently small for the depletion of the pump
wave to be neglected. The coupled equations for the fields at W 2 and
W3 are solved with the assumption of a constant amplitude at WI.
Making the assumptions of the previous paragraph about the wave
NON-LINEAR OPTICS 129
vector directions, field polarizations and crystal axes, the coupled
field equations become:
dEx(W2,Z)/dz =
{2nik2/8xx(W2)}Xxyiw2 = WI - w 3)Ey(w l )Ex*( -W3,Z);
(7.64)
dEx*( -W3,Z)/dz =
{ - 2nik3/8xx(W 3)}x"",x( -W3 = -WI + w2)Ey*( -wl)E",(W2,Z).
The pump wave at WI is polarized in the y-direction and the waves
are phase matched (!J..k = 0). The phase matching condition can be
expressed in terms of the dielectric constants:
WI8yyi(w1) = W28xxi (W2) + W38XXi(W3)' (7.65)
Eliminating Ex *( -W3,Z) between equations (7.64), we have:
(7.66)
where:
go2 = {4n2k2k3/8xxCW2)8xx(W3)}
x X"",iW2 = WI - W3)X':;'X(W3 = WI - w2) 1 EY(WI) 12.
ExCw 3,z) obeys the same differential equation (7.66) as Ex(w 2,z).
The general solution of (7.66) is:
ExCW2'Z) = BI exp (+goZ) + B2 exp (-goZ). (7.67)
Thus waves at frequency W2 or W3 grow exponentially when travelling
in either direction along the z-axis. The growth rate is determined by
the parameter go which is proportional to the pump power. When the
waves are not perfectly phase matched (11k ~ 0) then the gain
parameter becomes:
g = go{l - (!J..k/2go)2}i. (7.68)
The mis-match decreases the gain and g becomes zero when !J..k = 2go'
Equation (7.67) represents parametric amplification at a frequency
W 2 • If a crystal showing this effect is placed in a resonant cavity,
parametric oscillation at frequencies W2 and W3 is possible. Giord-
maine and Miller (Phys. Rev. Lett., 14, 973-6 (1965» demonstrated
this oscillation in the optical region with a crystal of lithium niobate
pumped by a coherent wave of 0·529 !lm wavelength. The pump
wave was produced by second harmonic generation from a laser
wave in another lithium niobate crystal. A plane Fabry-Perot con-
figuration was used as the resonant cavity. The plane ends of the
crystal were coated with multilayers which had high reflectance at
CO 2 and W3 but low reflectance and high transmittance at WI' The
130 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

condition for a spontaneous oscillation to build up from the noise at


£0 2 and £Os, is that the gain should be sufficient to overcome losses due
to incomplete reflection at the end faces and scattering in the crystal.
If reflection losses are predominant, the condition for oscillation can
be written:
gt> (1- R), (7.69)
where t is the length of the crystal and R is the reflectance of the
multilayers. Since g has its maximum value go when f).k = 0, the
oscillation tends to occur at the frequencies which satisfy the index
matching condition (7.65). The temperature-dependence of the re-
fractive indices was used to tune the frequencies of oscillation over a
small range; in terms of wavelength, the range of £0 2 was from 1·04
to 0·97 pm. The subharmonic oscillation, when w. = £Os = tWl
(equivalent wavelength 1·06 pm), is theoretically possible but was
not achieved in this experiment.

7.5 Third order effects


A wave at the third harmonic frequency can be generated by a laser
wave in a crystal from the non-linear polarization:
Pi(3wl) = XiJW(3Wl = £01 + £01 + wl)E;(wJEk(Wl)E,(Wl)' (7.70)
The third harmonic susceptibility tensor Xiik,(3wJ has non-zero
elements for isotropic and cubic materials, as well as crystals of
lower symmetry. The magnitude of XUkl(3w 1) is of the order of 10-15
e.s.u. The generation of the third harmonic in a crystal can be
analysed by a method similar to that of Section 7.3 for second
harmonic generation. A coherence length analogous to te given by
(7.45) can be defined for the third harmonic generation. Assuming
typical values for Xiikl(3wJ and t e , and a field amplitude
I 2E(wJ I = 3 X 10 5 V cm-t,
the third harmonic power conversion factor from the fundamental is
of the order of 10-15 • A conversion factor of 3 X 10-8 has been
obtained by using a focussed laser beam along an index matched
direction in calcite.
Stimulated Raman scattering is an example of a third order non-
linear process which can produce a large resultant effect. Raman
scattering occurs when an optical photon of frequency £0 L interacts
NON-LINEAR OPTICS 131
with a medium to produce an optical phonon of frequency Wv and a
scattered photon of frequency ws, less than WL' In gases and liquids,
molecular vibrations can play the same role as optical phonons in
solids. Raman scattering is associated with third order non-linear
polarization at frequencies WL and Ws:
Pi(WL) = XiikZ(WL = WL + Ws - wS)Ej(WL)Ek(ws)Ez*(-ws); (7.71)
Pi(ws) = XiikZ(WS = Ws + WL - wL)Ei(ws)Ek(WL)Ez*( -WL). (7.72)
It was shown in Section 7.2 for a classical model of oscillating
electron that X=(ws) = X!,.,.,(WL). When the difference frequency
(WL - ws) is equal to the electron resonant frequency 00o, the
susceptibilities have imaginary values which represent the absorption
of power at WL and the generation of power at Ws. This classical
model does not correspond exactly to the Raman effect because the
optical phonon frequency lOy is not an electron resonant frequency.
It is a resonant frequency of lattice or molecular vibrations. The
propagation of high-frequency waves is governed by the electronic
polarizability of the medium. It is the coupling between the elec-
tronic polarizability and lattice or molecular vibrations which causes
the Raman effect. However, there is some correspondence between
the Raman effect and the model of Section 7.2; in particular, the
correct formula for XiikZ(WL) includes a resonant denominator
D(WL - ws) where:
D(w) = {wy2 - w 2 - iyw}. (7.73)
The magnitudes of XiikZ(WL) and XiikZ(WS) at resonance (WL - wS=Wy)
are increased by the Q-factor (wy/y) over their magnitudes for
00 ~ Wy.
The non-linear polarizations (7.71) and (7.72) can produce large
effects if a phase matching condition is satisfied. This condition can
be written:
kL - ks = ky (7.74)
where kL = (wL/c)n(wL) and ks = (ws/c)n(ws). kL and ks will be
identified with the wave vectors of a laser wave and a Stokes wave,
that is a Raman scattered wave at frequency (WL - Wy). A typical
dispersion curve for an optical phonon with weak phonon-photon
coupling is shown in fig. 7.6. If the wave vectors are all in the same
direction and the difference between WL and Ws is sufficiently small
132 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

k
Figure 7.6 Dispersion curves of electromagnetic waves and transverse optic
vibrations (weakly coupled). (WL - ws) = wv. Dashed line illustrates approxi-
mate phase matching condition (7.75) for the stimulated Raman effect.

so that n(wL) ,...., news), then the phase matching condition (7.74)
becomes approximately:
n(wL)(wL - ws)/c = n(wL)(wV/C) = kv. (7.75)
Condition (7.75) can be satisfied if the index at WL exceeds the index
at frequencies just below Wv. The intersection of the dashed line in
fig. 7.6 with the phonon dispersion curve defines a kv which satisfies
(7.75). If kL and ks are not in the same direction, a larger kv is
needed to satisfy (7.74). Inspection of fig. 7.6 shows that a suitable
kv is available if the condition (7.75) for phase matching in the for-
ward direction is satisfied. In crystals with strong phonon-photon
coupling, such as strongly ionic crystals, the low-frequency index is
quite high and phase matching is not possible in the forward direc-
tion, at least in cubic crystals.
The gain at the Stokes frequency £Os can be great enough to pro-
duce oscillations if suitable optical feedback is provided. The inter-
action volume is usually greatest when the laser and Stokes waves
are parallel, and so the gain is greatest in that direction. The oscilla-
tion at £os, called the Raman laser effect, has been demonstrated in
a large number of molecular liquids and a few crystalline solids, for
example diamond, calcite and IX-sulphur. Raman laser action occurs
NON-LINEAR OPTICS 133
usually at the frequency of the strongest line in the Raman scattering
spectrum.
We will now consider some third order effects associated with a
single coherent wave of frequency WI' The third order non-linear
polarization produced by a single coherent wave is represented by:
PlWl) = Xijkl(W I = WI + WI - wl)EJCWl)Ek(w 1)E1*( -WI)' (7.76)
The tensor Xiikl is symmetrical for the subscript pairs jk and iI. It was
shown in Section 7.1 that Xiikl for an isotropic medium can be ex-
pressed in terms of two parameters oc and oc'. The tensor elements
which relate P,iw 1) to E z(Wl) and Ey(Wl) are:
+
Xzzxz = oc oc'; Xzzyy +
Xzyzy = oc; Xzyyz = oc'. (7.77)
The imaginary part of Xiikl is associated with the absorption or
generation of power at frequency WI' Assuming an isotropic medium
and a coherent wave polarized in the x-direction, the imaginary part
of Px (w 1) is:
iX~(WI = WI + WI -
w1)E,iWl)EzCw1)Ez*( -WI)' (7.78)
The corresponding absorption of power is proportional to:
X~zCWI = WI + WI - w 1){/(w 1)}2, (7.79)
where I(Wl) is the intensity of the coherent wave. If the linear
absorption, proportional to X~~(wl)/(Wl) is non-zero, then (7.79)
represents a change in the absorption coefficient proportional to the
intensity. If the medium has two energy levels separated by (2Iiw 1 ),
then (7.79) can be used to represent two-photon absorption processes.
The selection rule for two-photon electric dipole transitions differs
from the rule for single-photon transitions in requiring that the
initial and final states have the same parity.
The real part of Pi (W 1) is associated with a change in the effective
refractive index. For an x-polarized wave in an isotropic medium:
PzCWl) = X~ZZzCWI = WI + WI - w 1)EzCw 1)EzCw 1)Ez*( -WI)' (7.80)
The corresponding intensity-dependent refractive index can be
written:
(7.81)
nl(W) is positive for most materials. The increase in index with in-
tensity explains the self-focussing effects which are often observed
with laser beams. In a typical laser beam the intensity has a maximum
on the axis and decreases towards the edges. Hence the index near
134 OPTICAL ABSORPTION AND DISPERSION IN SOLIDS

the axis exceeds that near the edges, creating an effect like a con-
verging lens. If the power of the equivalent lens is sufficient to
overcome the spreading of the beam, the wave converges to a narrow
cylinder. The wave is trapped by total internal reflection in the
cylinder which acts as a wave-guide. The conditions for self-trapping
are given approximately by the following simple theory. The critical
angle Oe for total internal reflection at the surface of the cylinder, is
given by:
(7.82)
The fractional change of index is assumed to be small and hence
(n12 - ( 0 ) is a small angle. It follows that approximately:
(n12 - Oe)2 = (2ndno)J. (7.83)
The divergence of the trapped wave due to diffraction is about
(AolnoAI) where A is the cross-sectional area of the cylinder. The
condition for self-trapping requires that the divergence be less than
the glancing angle for total internal reflection:
(A02In 02A) < (2n dn 0)1. (7.84)
The power W in the laser beam is AI, so condition (7.84) can be
written:
(7.85)
This suggests that a laser beam of any cross-section can be self-
trapped ifits power exceeds a critical value We, estimated from (7.85)
at about 10 6 watt. This theory explains the propagation of high-
power laser pulses in glass. The pulsed beam is focussed to a narrow
filament which then continues for lengths up to several centimetres.
It can be shown that when (x' in (7.77) is non-zero, a circularly
polarized wave in an isotropic medium imparts optical rotatory
properties to the medium. This has been confirmed in experiments
on liquids but it is difficult to eliminate the effects of strain double
refraction in experiments on solids.
General References
Chapter 1
L. D. LANDAU and E. M. LIFSHITZ, 'Electrodynamics of Continuous Media'
(Pergamon Press, Oxford, 1960).
E. E. BELL, Optical constants and their measurement, 'Handbuch der Physik',
Vol. XXV/2a, pp. 1-57 (Springer Verlag, Berlin, 1967).

Chapter 2
M. BORN and K. HUANG, 'Dynamical Theory of Crystal Lattices' (Oxford Uni-
versity Press, 1954).
w. COCHRAN, Phonons in perfect crystals, 'Handbuch der Physik', Vol.
XXV/2a, pp. 59-156 (1967).

Chapter 3
J. TAUC (ed.), 'The Optical Properties of Solids' (Academic Press, New York &
London, 1966).
T. S. MOSS, 'Optical Properties of Semiconductors' (Butterworths, London, 1959).
N. F. MOTT and H. JONES, Motion of electrons in an applied field, 'The Theory
of the Properties of Metals and Alloys,' Chap. III (Oxford University Press,
1936).
H. Y. FAN, Photon-electron interaction, crystals without fields, 'Handbuch
der Physik', Vol. XXV /2a, pp. 157-228 (1967).
J. C. PHILLIPS, The fundamental optical spectra of solids, Solid State Physics,
18,56-164 (Academic Press, 1966).
J. TAUC, Optical properties of semiconductors in the visible and ultra-violet
ranges, Progress in Semiconductors, 9, 87-133 (Haywood, London, 1965).

Chapter 4
F. ABELES (ed.), 'Optical Properties and Electronic Structure of Metals and
Alloys' (North Holland, Amsterdam, 1966).
A. v. SOKOLOV, 'Optical properties of Metals' (Blackie, London, 1967).
A. B. PIPPARD, 'The Dynamics of Conduction Electrons' (Blackie, London,
1965).
M. P. GIVENS, Optical properties of metals, Solid State PhysiCS, 6, 312-52 (1958).
G. P. MOTULEVICH, Optical Properties of polyvalent non-transition metals
Soviet Physics Uspekhi, 12, 80-104 (1969).
135
136 GENERAL REFERENCES

Chapter 5
H. RAETHER, 'Solid State Excitation by Electrons', Springer Tracts in Modem
Physics, Vol. 38 (Springer, Berlin, 1965).

Chapter 6
D. L. DEXTER and R. S. KNOX, 'Excitons' (Interscience, New York, 1965).
s. NIKITINE,Exciton spectra in semiconductors and ionic compounds, Progress
in Semiconductors, 8, 233-322 (1962).

Chapter 7
N. BLOEMBERGEN, 'Nonlinear Optics' (Benjamin, New York, 1965).
R. W. MINCK, R. W. TERHUNE, and c. C. WANG, Nonlinear optics, Proc. IEEE,
54, 1357-74 (1966).
A. F. GIBSON, Nonlinear optics, Science Progress, 56, 479-97 (1968).
Index
Absorption, exciton, 100-1 Duminski, A. N., 125
Reststrahlen, 30-1 Dziedzic, J. M., 123
two-phonon, 37-40
two-photon, 133 Effective field, 33
Admittance, wave, 9 Ehrenreich, H., 21
surface, 11, 73-6 Electric susceptibilities,
Aluminium, 53, 63, 88, 92 linear, 112
Anomalous skin effect, 71-2 non-linear, 112-18
Arakawa, E. T., 88 Electromagnetic waves, 8-12
Ashkin, A., 123 Electrons, effective density, 20
energy bands, 41-4
Baldini, G., 101 mean free path, 71-2
Berreman, D. W., 32 optical mass, 63-7
Bjorkholm, J. E., 124 relaxation time, 71
Boundary equations, 4, 120-1 surface scattering, 75-6
Boyd, D. G., 123 transport theory, 69-72
Brust, D., 43, 51 Excitons, Frenkel, 97-8
Mott-Wannier, 98-100
Cadmium sulphide, 109 in semiconductors, 104-7
Caesium bromide, 34, 36 optical absorption, 100-1
Calcite, 124 Extinction coefficient, 9
Caldwell, R. S., 16
Canfield, L. R., 55 Fan, H. Y., 16, 86
Classical oscillator, harmonic, 22-4 Focussing, self, 133-4
anharmonic, 115-18
Cochran, W., 27 Geick, R., 36
Coherence length, 121 Germanium, 43, 51, 106, 125
Cowley, R. A., 27 Giordmaine, J. A., 129
Critical points, 48-53 Gold, 55, 65, 72
Culpepper, R. M., 83 Golovashkin, A. I., 68, 78, 79

Damping, lattice vibrations, 35 Hamm, R. N., 88


electrons, 23, 67-9 Hass, G., 55
Dielectric, constant, 4-6 Hodgson, J. N., 16, 55
tensor, 7-8 Hopfield, J. J., 109
Dispersion, formulae, 17 Hunter, W. R., 88
free electrons, 84-8
spatial, 107-10 Index matching, 122-3
Displacement, electric, 3 Indium antimonide, 31, 34, 83,
Dixon, J. R., 83 86
Drude formulae, 64-5 Interaction length, 127-8
137
138 INDEX

Ionic crystals, Quadrupole effect, 123-4


infra-red dispersion, 28-37
Raman scattering, stimulated, 130-2
Johnson, F. A., 38 Reflection coefficients, 10-11
Refractive index, 9
Kleinman, D. A., 21 Reststrahlen bands, 30-1
Kramers-Kronig relations, 13-19 Roessler, D. M., 18
Krypton, 101
Sanderson, R. B., 31
Lattice vibrations, 25-8 Scattering, electron-phonon, 68, 81-2
Lead, 60, 78-9 Second harmonic, 119-28
Lithium fluoride, 32 Sellmeier formula, 17
Longitudinal waves, 30-1 Siegman, A. E., 124
Lorentz formulae, 22--4 Silicon, 38, 92
Silver, 92, 125
Maxwell's equations, 8 Skin depth, 9, 69-70
McLean, T. P., 57,106 Sodium, 88, 92
Miller, R. C., 129 Sodium chloride, 18, 34, 125
Motulevich, G. P., 68 Spitzer, W. G., 21, 86
Sum rule, 19-22,66-7
Norman, S. L., 60 Superconductors,
infra-red absorption, 59-61
Optical conductivity, 6, 9 Sutherland, J. c., 88
Optical mass, 63-7 Szigeti effective charges, 34
Oscillations, parametric, 128-30
plasma, 88-92 Taft, E. A., 51
Oscillator strength, 20 Tellurium, 16
Third harmonic, 130
Phase matching, 122-3, 132 Thomas, D. G., 109
Philipp, H. R., 21, 51 Tomiki, T., 103
Phonons, 25-8 Transitions, intraband, 65-9
Plasmons, volume, 88-92 interband, 41-61
surface, 92-6 direct, 44-8
Polarization, electric dipole, 3 indirect, 56-61
non-linear dipole, 111-15
non-linear quadrupole, 123--4 Urbach's rule, 58-9
Potassium bromide, 27, 29, 34
Potassium chloride, 103 Walker, W. c., 18
Power dissipation, 5, 114 Wang, C. C., 125
Waves, damped, 9
Quartz, 5, 21 evanescent, 9, 93-4

You might also like