635
635
SOLID
-.
STATE
PHYSICS
,
~;;f'
By
ADRIANUS J. DEKKER
DEPARTMENT OF EI,ECTRICAL
ENGINEERING, UNIVERSITY OF GRONINGEN
LONDON
Published hy
MACMILLAN & CO LTD
--/---
f':
PREFACE
THE purpose of this book is, to introduce the reader to the study of the
physical properties of crystalline solids. It is based on notes which I used
for lectures in the Physics Department of the University of British Columbia, Canada, and in the Electrical Engineering Department of the University of Minnesota.
My aim has been to write an introductory text suitable for senior undergraduate and beginning graduate courses on the solid state in physics,
engineering, chemistry, and metallurgy. Also, I have attempted to make
it suitable for self study by scientists in industrial laboratories interested in
the physical properties of solids. The widely varying background of the
anticipated groups of readers has affected the organization and presentation of the subject matter. The general level of presentation has been
kept elementary, with emphasis on the physical reasoning underlying the
interpretation of the physical properties of solids. I have made an effort,
however, to remain as rigorous and up-to-date as possible within the limits
imposed by the level of presentation. The first eight chapters deal with
subjects which, at least in an introductory text, can..'be discussed without
reference to the details of the electronic structur' of solids. Prerequisite
for understanding this part of the book is an elementary knowledge of
statistical thermodynamics and of the quantized harmonic oscillator.
Chapters 9 through 20 deal with the electronic properties of solids and
require familiarity with the elements of wave mechanics, although in a
number of chapters no explicit use of wave mechanics is made. As a
consequence of the organization of,the material outlined above, the degree
of difficulty tends to increase as one progresses through the book. This in
itself does not compel the reader to follow the order in which the various
subjects are discussed. In fact, the chapters are organized in groups which
could be taken up in any order suitable to serve the particular needs of the
instructor or reader.
To some extent, my own interest and taste have determined the choice of
,
vi
PREFACE
,.
material; however, with the possible exception of Chapter 17, the material
is basic to a great variety of subjects in the field of solid state.
I am indebted to W. Opechowski for constructive criticism during the
preparation of Chapters 10 and I I, and to A. H. Morrish for his comments
on other parts of the manuscript. I also wish to acknowledge the cooperation of numerous publishers who kindly permitted me to reproduce illustrations. I am grateful to F. L. Vogel, W. G. Pfann, H. E. Corey, and
E. E. Thomas for a micrograph of a lineage boundary in germanium.
Finally, I wish to thank my wife for typing the manuscript and for her
encouragement.
A. J. Dekker
CONTENTS
1. The Crystalline State
The crystalline state of solids ....................... .
Unit cells and Bravais lattices ...................... .
Miller indices .................................... .
The diffraction of X-rays by a simple space-lattice according to von Laue ............................ .
X-ray diffraction according to Bragg ................ .
1-5.
The atomic scattering factor ........................ .
1-6.
X-ray intensity and atomic configuration of the unit cell ..
1-7.
Experimental methods of X-ray diffraction ........... .
1-8.
Diffraction of electrons by crystals .................. .
1-9.
1-10. Diffraction of neutrons by crystals .................. .
1-11. Interatomic forces and the classification of solids ...... .
1-12. Anisotropy of the physical properties of single crystals .. .
I-I.
1-2.
1-3.
1-4.
1
4
8
lO
13
14
16
19
20
21
23
27
32
2-1.
32
34
35
36
39
41
45
46
49
51
53
54
57
viii
CONTENTS
60
60
62
65
67
70
74
76
78
81
83
86
88
91
93
96
99
104
104
105
107
109
III
114
117
117
117
120
121
121
124
128
CONTENTS
IX
133
133
134
138
140
141
144
8. Ferroelectrics
8- I.
8-2.
8-3.
8-4.
8-5.
8-6.
8-7.
8-8.
148
150
154
160
160
164
166
168
171
175
178
184
184
186
192
195
196
198
20 I
207
CONTENTS
211
CONTENTS
11-8.
11-9.
11-10.
II-II.
The
The
The
The
xi
.r..
'1,,'
. ..1
,./w".:d
.
292
295
299
301
305
305
306
308
310
314
316
319
319
320
326
329
331
334
341
344
348
14-1.
348
349
351
354
356
357
361
366
366
369
CONTENTS
XlI
15-3.
15-4.
15-5.
15-6.
15-7.
15-8.
15-9.
15-10.
IS-II.
15-12.
15-13.
16. Luminescence
16-1.
16-2.
16-3.
16-4.
16-5.
16-6.
General remarks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Excitation and emission.... . . . . . . . . . . . . . . . . . . . . . . . ..
Decay mechanisms ........ " ... '" ....... ,. .. . .. . ..
Thallium-activated alkali halides. . . . . . . . . . . . . . . . . . . ..
The sulfide phosphors. . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Electroluminescence................................
398
398
399
402
406
410
413
418
446
446
448
CONTENTS
18-3.
18-4.
18-5.
xiii
451
454
457
458
459
460
464
Ferromagnetism
464
466
468
472
475
478
480
481
Antiferromagnetism
483
484
488
Ferrimagnetism
491
491
493
498
Paramagnetic Relaxation
20-1.
20-2.
20-3.
20-4.
Phenomenological description. . . . . . . . . . . . . . . . . . . . . ..
Relaxation mechanisms. . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Spin-lattice relaxation. : . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Spin-spin relaxation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
498
499
501
504
20-5.
20-6.
20-7.
505
506
508
xiv
CONTENTS
20-8.
51 I
513
516
517
518
APPENDIX
525
A.
B.
C.
D.
E.
INDEX
519
525
526
527
529
531
533
j"
I
.'
..
Chapter 1
lui
rb)
I
Fig. 1-1. Schematic illustration df the difference between a
crystal (a) and a glass (b). [After W. H. Zachariasen, J. Alii.
Chem. Soc., 54, 3841 (1932)]
[Chap. t
Sec. I-I]
16
_. -,
_riA)
10
12
14
10
12
14
(a)
.03
fir)
.02
.01
Ib)
; ,;
[Chap. I
specific distances from the origin. For example, there are eight atoms at a
distance ~aV3, six atoms at a distance a, etc. In the liquid state the
situation is rather different. Suppose the origin of the coordinate system
is attached to a given K atom and moves with this atom. At a given
instant there will be a certain configuration of the other atoms, but the
configuration changes continually with time. Taking the time average of
these different configurations, one could then plot the average number of
nuclei as function of the distance from the central atom. Such information
may actually be obtained from X-ray diffraction experiments. Thus, in
Fig. 1-2b the fully drawn curve fer) represents the density of K atoms per
Aa at 70C as function of the radial distance from an arbitrary K atom in
the liquid;3 the dashed line corresponds to 395C. Note that the set of
discrete lines of Fig. 1-2a has been transformed into a continuous curve.
Also, only the first few "shells" of other a4>ms are distinguishable in the
70C curve, whereas in the 39SoC curve only the first two are somewhat
pronounced. For distances larger than ,-,10 A the curves show little or no
structure and the density becomes independent of r; for the crystal,
however, the discrete lines extend over the whole piece of material, at
least when it is a single crystal. It is of interest to remark that the integral
of 47Tr2J'{r) over the first peak determines the average number of nearest
neighbors of the central atom; for the alkali metals we find that this is
approximately equivalent to 8 nearest neighbors, as it is in the solid (only
there, it is exactly 8).
1-2. Unit cells and Bravais lattices
We shall now discuss somewhat further the periodicity of structure,
which is the fundamental feature of a crystal. Consider part of a twodimensional crystal, the atoms of
M
which are arranged in a pattern
j----j---- ----'j
H '
as illustrated in Fig. 1-3. Each
G
/
cluster of atoms (in this case a dot
___ ___ !'V _ --o.k
/
and two open circles) will be
E
F
/
;'
referred to as a pattern-unit. It is
D '
C(
/
observed that when a parallelo-of- -- ~
gram such as ABC D is repeatedly
b
/
translated by the vectors a and b,
----~---A
K
corresponding respectively to AB
-----a-B
and A D, the whole pattern may be
obtained; thus ABCD is called a Fig. 1-3. Two-dimensional crystal and
ill'(- -
Sec. 1-2]
1Il1,
Fig. 1-4.
(bl
[Chap. 1
Sec. 1-2]
/
c
{3
I/a
2
!.. .
11
10
12
13
14
0"'0-
[Chap. I
Table 1-1. The Seven Crystal Systems and Their Essential Symmetry
System
Essential symmetry
f-----
Triclinic
Monoclinic
Orthorhombic
(rhombic)
Tetragonal
Cubic
Hexagonal
No planes, no axes
One 2-fold axis or one plane
Three mutually perpendicular
2-fold axes, or two planes
intersecting in a 2-fold axis
One 4-fold axis or a 4-foJd
inversion axis
Four 3-fold axes
One 6-fold axis
a = b
c; oc
{J
i' = 90
a = b ,~ c; oc = fJ = )J = 90
Three equal coplanar axes a at
120; fourth axis c 1. to
these; c i' a.
u~d~". ~ fl~y~, 90'
0
I
I
Rhombohedral
(trigonal)
'?"
.'
-, '
Sec. 1-3]
order to specify the orientation, one employs the so-called Miller indices;
these are defined as follows: Suppose a particular plane of a given set
has intercepts pa, qb, and re with the crystal axes (Fig. 1-8). The Miller
indices of the set of planes are then given by three numbers h, k, I such that
I. : k : 1= I/p : I/q : I/r
(1-1)
with the condition that h, k, and I are the smallest integers satisfying (I-I),
i.e., h, k, and I have no common factor> I. We shall adhere to the rather
general practice of using the notation (hkl) for a particular set of planes.
We emphasize once again tha.t these indices refer not to a particular plane
Fig. 1-8. Illustrating a plane with intercepts pa, qb, and rc; ON is the normal
to the plane.
but to a set of parallel planes. One or more of the indices may be negative
when the corresponding intercepts are negative; they are represented in a
form such as this: (likl), (hkl) etc. Miller indices for some planes are
given in Fig. 1-7.
When the indices are shown enclosed by braces, such as {hkl}, they
refer to planes which in the crystal are equivalent even though their Miller
indices may differ. For example, in a cubic lattice all cube faces are
equivalent; in order to specify this group of planes, one writes {100},
which includes the planes (100), (010), (001), (100), (010), (001).
In order to specify a certain direction in a crystal, one employs three
indices u, r, w enclosed in square brackets [u, v, w]; the indices are
integers and have no common factor larger than unity. The direction
specified by this symbol is obtain~d as follows: Move from the origin
over a distance ua along the a-axis; vb along the b-axis and we along the
c-axis. The vector connecting the origin with the point so obtained is
then the direction specified by the symbol [uvw]. Thus, in a cubic crystal
the direction of the x-axis is indicated by [100], the y-axis by [010], etc.
A full set of equivalent directions in a crystal is represented by a symbol
of the kind (uvw).
We may note here that the Miller indices of a set of planes are related
IO
[Chap. I
: cos y'
= (h/a)
: (k/b) : (l/e)
(l-2)
.; Sec for example C. S. Barrett. Structure o{ Metals. 2d ed., McGraw-Hill. New York,
952, p. 633.
r
I
Sec. J-4]
II
electrons in the atom are nearly all in phase. However, X-rays used in
diffraction work have a wavelength of the same order of magnitude as
the atomic diameter (this is necessary to obtain a diffraction pattern).
Thus the wavelets emitted by the electrons in an atom are in general out
of phase. Consequently, these wavelets will partially cancel each other by
interference and the amplitude of the radiation scattered by an atom
containing Z electrons IS less than that scattered by a free electron times
the number of electrons in the atom. We can, however, consider the
atom as a scattering center with an effective atomic scattering factor Is
which is given by the ratio of the amplitude of the wave scattered by the
t Zero order
Ix
Fig. 1-9.
Incident waves
atom and that of the wave scattered by a free electron (for the same
incident beam). This problem will be discussed further in Sec. 1-6.
In crystals we are concerned with the scattering by a large number of
atoms arranged according to a particular pattern. For simplicity, let us
consider a one-dimensional row of atoms with interatomic distance a.
Assuming the incident wave crests to be parallel to the row of atoms, we
obtain a picture such as Fig. 1-9. The envelope of the wavelets emitted
by the individual atoms forms new wave crests and we see that besides a
beam propagated in the same direction as the incident beam (zero-order)
there are a few diffracted beams of other direction (first-order, secondorder, etc.). Thus, even though th~ individual atoms scatter radiation in
all directions, there are only a few directions in which these wavelets
reinforce each other. The condition for such a diffracted beam to exist
may easily be found as follows: In Fig. 1-10, suppose that AB is a \\,ave
crest of the incident beam, and CD is a wave crest of the diffracted beam.
Then, because a wave crest is an assembly of points of the same phase,
we must require that the path difference (AC - BD) shall equal an integer
12
(Chap. I
cos ClO)
= eA,
with
e = 0, 1,2, 3, ...
' \ (1-4)
For given values of Cl O' a, A, and e there is only one possible value for Cl.
We note that such a value exists only if at the same time cos ex ~ I.
Suppose then that to a certain value of e there corresponds a value IX.
The direction of the diffracted beam then forms a cone of directions with
the row of atoms as axis, as indicated in Fig. 1-10. Thus a monochromatic
Sec. 1-4]
13
a (s -
so)
= b (s = c . (s -
so)
= I}.
= g).
so)
e},
(1-6)
where e,f, and g are integers; c(o, fJo, Yo and C(, fJ, y represent, respectively,
the angles between the incident and scattered beams and the axes a, b, c.
These are the von Laue equati.ons. It must be noted that for a given value
of ;, and an arbitrary direction of incidence so, it is in general not possible
to find a direction s which satisfies (1-6). In other words, for a monochromatic X-ray beam falling on a crystal with an arbitrary direction of
incidence, in general no diffraction is observed. This may readily be
understood by remembering that for a two-dimensional lattice there
exist only specific directions for a diffracted beam and these directions
are in general not part of the direction cones determined by the third
condition required for the three-dimensional case. Thus only for particular
angles of incidence will diffraction be observed; it is exact Iy this limitation
that makes X-ray diffraction a useful tool for investigating crystal
structures. This point will become more clear when we discuss the Bragg
formula below. Before doing this, it may be useful to rewrite (1-6) for
the case of a simple cubic lattice. Assuming a = b = c, we obtain from
(1-6) by squaring and adding
2a 2(1 - cos C( cos C(o - cos fJ cos fJo - cos y cos Yo) = ).2(e2 12
g2)
+ +
Now, if cp represents the angle between the incident and scattered beams,
we may write
2(1 - cos cp) = 4 sin 2 (cpj2) = ().2ja 2)(e2
+12 + g2)
(1-7)
In this form the von Laue equation is closely related to the Bragg formula.
1-5. X-ray diffraction according to Bragg
Bragg considered the problem of X-ray diffraction from a somewhat
different point of view. Although in itself it is not completely satisfactory
because it involves certain assumptions that are not immediately obvious,
it gives results identical with the Laue treatment and is therefore justified.
Bragg considers X-ray diffraction from a crystal as a problem of reflection
from atomic planes. In Fig. I-II consider a set of parallel atomic planes
of Miller indices (hk/), the distance between successive planes being d hk1 .
If we assume with Bragg that an X-ray beam is reflected by an atomic
plane according to Snell's law (i.e., incident beam, reflected beam, and
normal in one plane, and angle of incidence equals angle of reflection)
14
[Chap. I
we see that rays I and 2 can reinforce each other in the reflected direction
only if their path difference is an integer times A. This is necessary because
wave crests are points of equal phase. Thus from the figure we find as
condition for reflection from the set of planes under consideration,
with
n = 0, 1,2, 3, '"
(1-8)
2a sin () = An(h2
+ k 2 + /2)1/2
(1-9)
Thus, identifying the integers e,J, and g, respectively, with nh, nk, and nl,
a diffracted beam detIned in the Von Laue treatment by the integers e, f, g
may be interpreted as the nth order reflection from a set of planes (hkl)
in the Bragg theory. The order of the reflection n is simply equal to the
largest common factor of the numbers e,J, g.
Sec. 1-6]
15
(1-10)
On the other hand, the absolute value of the amplitude of the scattered
wave is of course independent of the location of the charge and simply
proportional to the amount of charge. The ratio of the complex amplitude
of the wave scattered by the element under consideration and the amplitude
of the wave that would be scattered by the same charge in A is thus simply
ei'l. The atomic scattering factor is 'therefore
j, =
/, =
(oo
I,
.0
sin kr
47Tr 2p(r) - - dr
kr
with k
.. '
= (417/,1)
sin
f)
(I-II)
...
16
[Chap. I
2[\
'"
To illustrate the problem to be discussed here, let us consider a particularly simple example. In Fig. 1-14 let us suppose, to begin with, that
only the corners of the cubic unit cell
are occupied by atoms. For such a
simple cubic lattice, (which, by the
3
way, does not occur in nature) the
first-order reflection from the set of
2
planes (001) would be observed for a
particular Bragg angle 0, determined
in accordance with (1-8) by
2doOl sin ()
2 a sin () = A
Fig. 1-14. Illustrating that the firstorder reflection from the planes {tOO}
is absent in a body-centered cubic
lattice. The path difference between
1 and 2 is ),; between I and 3 it is ).12.
For values of the scattering factors of atoms and ions see R. W. James and G. W.
Brindley, Z. Krist. 78,470, (1931); il1ternationale Tabellen zur Bestimmlll(f{ vall Kristallstrukturen, Vol. 2, Borntrager, Berlin, 1935.
Sec. 1-7]
17
In analogy with the von Laue equations (1-6) it then follows that the
phase difference between the beam scattered by atom k and the one
scattered by the atom at the origin is given by
where So and s are unit vectors, respectively, in the direction of the incident
and scattered beam. Substituting (1-12) and making use of (1-6), we
obtain
( 1-14)
18
[Chap. I
(1- 15)
where the summation extends over all atoms in the unit cell. In connection with this summation we must emphasize that an atom at a corner
belongs to eight unit cells, so that
such an atom in the summation
counts for only i. In other words,
all atoms at the corners together
produce only one term. We may
also look at this problem in this
fashion: if we add vectorially the
amplitudes of the waves scattered
Fig. 1-15. Showing the vectorial addition
by the atoms in a unit cell, we
of the amplitudes of the waves scatt<:red
obtain a picture such as in Fig.
by the different atoms in the unit cell.
1-15, where F is the resultant
F is the resultant of the individual j,k'S.
amplitude. Each amplitude has
two components, j~k cos tpk and f,le sin tple and the intensity, which is
proportional to the square of the amplitude, then becomes proportional to
(1-16)
IFI2
(1-17)
A simple example may illustrate the conclusions one may draw from the
above treatment. For a body-centered cubic lattice of similar atoms, the
summations extend over the values u, l" IV = 0, 0, 0 and u, V, w :_ !, t, t
(all corner atoms together represent one atom). According to (1-14)
ffl = 0 and tp2 = 7T(e f
g). Hence, for a body-centered cubic lattice
of similar atoms we have, according to (1-16),
+ +
IFI2 =
f~ ([I
...
19
Sec. 1-7]
etc. The results of such considerations for cubic lattices of similar atoms
are represented in Fig. 1-16 in the form in which they appear as lines in
1,1 1.3
3 4 5 6
17
l~ ~1
(e2+f2+g2)
8 9 10 ,: 12 : 14 16: 18 : 20:22 24
I I
Simple cubic
Body-centered cubic
Facecentered cubic
Diamond
20
[Chap. I
A =h/p
(1-18)
or
(1-19)
me 2
..1 2
2h2 (Z - /.) sin2 0
(1-20)
Here Is is the scattering factor for X-rays, Z is the nuclear charge, and
() is the Bragg angle. As for X-rays, the scattering factor for electrons
, C. Davisson and L. H. Germer, Phys. Rev., 30,707 (1927).
M. Born, Z. PhYSik, 38, 803 (1926).
9 N. F. Mott, Proc. Roy. Soc., 127A, 685 (1930).
'.
Sec. 1-9]
21
--
22
[Chap. J
(111)
(331
(511)
100
80
\
60
80 K
ao=B.85A.
0
(100)
(311)
60
40
293'K
MnO
Tc=120'K
ao=4,43A.
20
Scattering angle
Fig. 1-17. Neutron diffraction patterns for MnO at room temperature and at 80 o K. The magnetic unit cell is twice as large as the
chemical one. [After Shull and Smart, ref. 121
Sec. 1-10]
23
E(r)
= -rxlr"
+ Plr'"
(1-21 )
where r is the distance between the nuclei of the two atoms; rx, p, m, and n
are constants characteristic for the AB molecule. The zero of energy is
chosen such that for infinite separation E = O. The first term, which is
negative, corresponds to the energy associated with the forces of attraction,
the second (positive) term corresponds to the forces of repulsion. In
fact, the force between the two atoms as function of r is given by
nrx
F(r) = -dEldr = - - n 1
r+
+ -rmp
m +1
( 1-22)
The energy and the force between two atoms A and B which form a
13
(1946).
24
[Chap. I
t
Repulsive
(bl
(al
The dissociation energies are of the order of one or a few electron volts.
Assuming that the energy curve exhibits a minimum, one may express
the equilibrium distance ro and the corresponding binding energy E(ro)
in terms of the constants t:I., fJ, m, and n by making use of the condition
(d/dr),.~"rO =
0,
i.e.,
r~t-n =
(1-23)
(m/n)(fJ/r:x)
-r:x/r~
+ fJ/r3' =
(1-24)
-(a/r3)(1 - n/m)
Note that although the attractive and repulsive forces are equal in equi__~_. librium, the attractive and repulsive energies are not equal since n m. In
fact, if m ~ n, the total binding energy is essentially determined by the
energy of attraction -alr~.
As one may expect already by looking at Fig. 1-18, a minimum in the
energy curve is possible only if m > n; thus the formation of a chemical
bond requires that the repulsive forces be of shorter range than the
25
Sec. I-Ill
1)1X/r~+2
+ m(m +
l)fl/r~'+2
>
m>n
1.
2.
3.
4.
,,'
]t should be said from the outset that many intermediate cases occur
and in general one must be somewhat careful in employing very specific
labels.
26
[Chap. 1
bonds. Some further remarks may be found in Sec. 13-1. Valence crystals
are very hard (diamond, carborundum), are difficult to cleave, and have a
poor electrical and thermal conductivity.
27
Sec. I-II]
containing more than one type of atom; the two groups meet in the van
der Waals crystals (argon as an element and CH 4 as a compound would
be examples). Between the true alloys and ionic crystals there is a group
of intermetallic compounds for which the composing metallic components
have different tendencies for giving up electrons (Mg 3 Sb 2).
,
Monoatomic
metals
(Ag,Cu)
Valence
crystals
(diamond)
Ge, Si
Bi
-_c
~ .... l~
p
Se
van der Waals
crystals
(A,CH4 )
Si02
SiC
"
Alloys
(NiCu)
Fig. 1-19.
Mg 3 Sb 2
Ionic crystals
(NaCl)
FeS
Ti02
_\
28
[Chap. I
+ axzE.
I" = a"xEx + oyyE" + ayzE.
E y + azzE.
I z = azxEx +
Ix
= axxEx + axlI E y
(1-26)
0ZlI
IxEx
+ IyEy + I.E. =
a=E;
yz
(1-27)
where aI' a 2, and aa are the principal conductivities. Thus the electrical
properties of any crystal, whatever Jow symmetry it may possess, may be
characterized by three conductivities al , 2 , a3 or by three specific resistivities PI' P2, P3' Note that 1 and E have the same direction only when the
applied field falls along anyone of the three principal axes of the crystal.
In cubic crystals the three quantities are equal and the specific resistivity does not vary with direction. In hexagonal, rhombohedral (trigonal),
and tetragonal crystah the resistivity depends only on the angle 1> between
the direction in which P is measured and the hexagonal, trigonal, or
tetragonal axis, since in those crystals two of the three quantities PI' P2, P3
are equal. One finds
(1-29)
p(1)) = Pl. sin21> + PI! cos 2 1>
Sec. 1-12]
29
REFERENCES
:'c'
i"
'
30
[Chap. J
1
i
PROBLEMS
simple cubic:
= 2R; f = 7T/6
b.c.c. :
f.c.c. :
= 4R/Y2; f = (7TY2)!6
4R/Y3;
f=
(7TY3)/8
Chap. I]
31
I I
1-11. Discuss methods for growing single crystals (see, for example,
H. E. Buckley, Crystal Growth, Wiley, New York, 1951).
1-12. Discuss some physical effects which are due to anisotropy (see,
for example, C. Zwikker, Phys;cal Propert;eso.( SaUd Mate,;als, Interscience,
New York, 1954, Chapter 4).
,"
,
'<[1
;'1;.
, I
"'"
,,'1
:',j
,j
I"
.'
I,
<'."
'."
Chapter 2
con~tant
(2-1)
dE+ pdV
G;)v
dT+
G:)
dV
G;)v
dT+
[(~!) + p] dV
T
(2-2)
The specific heat in general is defined by dQ/dT, and unless stated otherwise, will be assumed to refer to 1 gram molecule of the solid. However,
unless one specifies in which way the increase in temperature takes place,
the specific heat is undetermined; in particular one must specify the
corresponding change in volume, as is evident from (2-2). Thus there
exist an infinite number of specific heats, but in general one is interested
in only two: the specific heat at constant volume Cv and the specific heat
at constant pressure Cpo According to (2-2), the former is given by
(2-3)
Sec. 2-1J
33
a solid at constant pressure than at constant volume. As shown in textbooks on thermodynamics, the second law leads to the following relation__
ship between Cll and CV :1
Cll
Cv = -T (oV)2 (op)
,
oT p oV T
(2-4)
(1/V)(oV/oT)ll
K = -(I/V)(eV/eph
and
(2-5)
(2-6)
7
~
Thus C v may be calculated from
~ 6
Cp measurements if at the same time
'"C
!Xv and K are known at the
~ 5
E
temperature of interest. Since both
~
4
!Xv and K are positive quantities,
.S
3
ep - Cv ;>- O.
Q
By way of illustration, we have
2
given in Fig. 2-1 Cp and Cv as
1
functions of temperature for copper; note that at low temperatures
800
400
1200
o
their difference becomes very small
-+- T (absolute)
and that both go to zero at T = O.
It is essentially the temperature
Fig. 2-1. The temperature variation of
variation of the specific heat at C and C for copper. [By permission
constant volume which will be from M. W.v Zemansky, Heat and Thermodiscussed in the present chapter. dynamics, 2d ed., McGraw-Hill, New York,
It may be noted that ifno direct
1943, p. 237]
compressibility data are available for the temperature range of interest,
one frequently employs the relationship
-t
(2-7)
34
[Chap. 2
From the atomic point of view one may distinguish between various
contributions to the specific heat of solids. In the first place, there is the
contribution resulting from the atomic vibrations in the crystal; an
increase in temperature is associated with a more vigorous motion of the
atoms, which requires an input of energy. Second, in metals and in
semiconductors there is an additional contribution to the specific heat
from the electronic system. Usually this contribution is small relative to
that of the lattice vibrations, as explained in Chapter 9. As the temperature
is raised from absolute zero, the
specific heat increases rather
10
rapidly from zero and finally levels
8
off to a nearly constant value. For
elements, the value at high temper~
atures is about 6 cal mole- I
4
degree-I. This is known as the law
of Dulong and Petit. Anomalies in
2
the specific heat curves are observed
in the ferromagnetic metals; for
o 200 400 600 800 1000 1200
example, in nickel, iron, and cobalt,
-+T
a peak is observed in the vicinity
Fig. 2-2. CD in cal mole- 1 degree- 1 for of the ferromagnetic Curie tem~
nickel as function of the absolute perature (see Fig. 2-2). The height
temperature.
of the peak is of the same
order of magnitude as the normal specific heat. The peak is associated
with the transition from the ferromagnetic (ordered) to the paramagnetic
(disordered) state. Similar peaks occur in the specific heat curves of alloys
which exhibit order-disorder transitions, and in ferroelectric materials.
These anomalies are discussed in the relevant chapters; in the present
chapter the discussion is confined to the specific heat associated with
atomic vibrations.
~p
,...
~
Sec. 2-2]
35
degree~l mole~l
(2-11)
36
[Chap. 2
(2-12)
nhve-nhv/kTj
n~O
e-nhv/kT
n~O
(2-13)
~ e-nhv/kT
(1 _
e-hv/kT)-l
n~O
oS
o(ljkT) - -
00
2:
hye- hv/ kT
(1 - e- hv/ kT )2
A. Einstein, Ann. PhYSik, 22, 180, 800 (1906); 34, 170 (1911).
See any Introduction to Modern Physics.
Sec. 2-4]
37
3N(E) = 3N
hv
. 1
-
hvlkT
(2-15)
~ E = 3R
oT
(_h_v ) 2 -,--;:--"e""hV,/k_T-----:
kT (e hv/ kT - 1)2
(2-16)
(n
+ t)hv
n = 0, 1,2, '"
(2-17)
rather than by (2-12). 5 This has the effect of shifting all energy levels by
the constant amount of hv/2, and instead of (2-14), one obtains
hv
(E) =
hv
2' + e hv/kT _
(2-18)
The first term is called the zero-point energy of the oscillator because
(E) = hv/2 for T = 0. Thus, according to quantum mechanics, the atoms
have vibrational energy even at absolute zero. The expression for the
specific heat is not altered by this result, because C v is determined by the
derivative of (E) with respect to T.
With regard to (2-16) it is observed that for kT,:?> hv, this expression
reduces in first approximation to the classical result (2-11). At low
temperatures, however, the specific heat decreases. To discuss this
5
38
[Chap. 2
\
(2-20)
De~ye
1.0
h'
.._
/ .~
.8
/V
k;
.6
.4
.2
~
..
EinstelU
VI
1/
/
1;
.2
.4
.6
.8
1.0
1.2
1.4
1.6
1.8
_Till
Fig. 2-4.
2.0
<
Sec. 2-5]
39
(2-21)
where Cs is the velocity of propagation of the waves. If it is assumed that
the end points of the string are fixed, the solutions of (2-21) are those
corresponding to standing waves:
u(x,t) = A sin (mrx/L) cos 27Tl1nt
(2-22)
An = 2L/n
and
Vn = cs/An = c_n/2L
(2-23)
3u
3u
1 3u
3u
-+-+-=_.ex2 of OZ2 c; ot 2
P. Debye, Ann. Physik, 39, 789 (1912).
(2-24)
(2-25)
40
[Chap. 2
(2-26)
where now nx , ny, and nz are positive integers 31. Substituting this
solution into the differential equation (2-25), one obtains the following
expression for the possible modes of vibration:
(2-27)
Z(I')
Zip)
_v
_v
(b)
,A.
+ dR
(2-29)
Each point occupies on the average a unit volume in the integer space.
Sec. 2-5]
41
Now, each point corresponds to a set of three integers nx , ny, n z , and each
set of integers determines, according to (2-26), a possible mode of
vibration; hence (2-29) immediately gives the number of possible modes
of vibration in a given range. Expressing R in terms of v in (2-29) one
thus finds
(2-30)
where V is the volume of the solid. For a perfect continuum, the possible
frequencies vary between 0 and IX, the number of such possible vibrations
increasing with the square of the frequency (see Fig. 2-5b). This situation
holds, for example, in the case of electromagnetic waves in a box of
volume V. Expression (2-30) is therefore basic in the theory of black-body
radiation.
In the case of elastic waves, we may distinguish between transverse and
longitudinal waves. In general, the velocities of propagation, say C t and
c" respectively, will not be equal. To set up an expression for Z(v) dv in
this case one should keep in mind that for each frequency or wavelength
there are two transverse modes and one longitudinal mode. 8 Thus,
instead of (2-30) one obtains
Z(I') dv
47TV
(.~ + ::l) v
I
dv
(2-31 )
How this expression has been used in the theory of the specific heat of
solids will be discussed in the following two sections.
2-6. The Debye approximation
One may wonder what the discussion of the preceding section could
have to do with the specific heat of crystals, which are by no means
continuous but are built up of atoms, i.e., of discrete "mass points."
The reason is the following: Consider an elastic wave propagated in a
crystal of volume V. As long as the wavelength of the wave is large
compared with the interatomic distances, the crystal "looks like" a
continuum from the point 'of view of the wave. The essential assumption
of Debye is now that this continuum model may be employed for all
possible vibrational modes of the crystal. Furthermore, the fact that the
crystal actually consists of atoms is taken into account by limiting the
total number of vibrational modes to 3N (see Sec. 2-2), N being the total
number of atoms. In other words, the frequency spectrum corresponding
to a perfect continuum is cut off so as to comply with a total of 3N modes
(see Fig. 2-6a). The Debye cut-off procedure leads to a maximum
8 In the longitudinal modes, the deflection is along the direction of propagation;
in the transverse modes the deflection is perpendicular to the direction of propagation,
which gives two independent components.
42
[Chap. 2
(2 I)
'I'lJ
+:;
II["lJ v2 dv
l Z(v) dv = 417V :;
C
C
0
t
1
or
..
9N
vJj'=
417 V
(2-32)
= 3N
(I
(2 + 1)-1 \
c:;
(2-33) ,~~
c:;
where Z(v) as given by (2-31) has been used. It should be noted that this
procedure assumes that the velocities c t and c/ are independent of the
Z(v)
Z(v)
t
\
_p
(bl
Fig. 2-6. The Debye cut-off takes place at the Debye frequency
lJ' common to the transverse and longitudinal modes (a), In
Born's procedure, the cut-off takes place at a common minimum
wavelength, corresponding to the maximum frequencies v, and v,
for the transverse and longitudinal modes respectively .(b). Note
that c, <: Ct.
wavelength, as in the continuum, It will be seen in Sec. 2-9 that this is not
correct for actual crystals. The order of magnitude of V]) may be obtained
by taking Nj V ~ 1022 per cm 3 and using for the velocity of sound
.--.!(}' em sec~I. This gives Vj) ~ 1013 per second. This corresponds to a
minimum wavelength of the order of one Angstrom, indicating that the
continuum theory may be at fault, especially in the high-frequency region.
Associating with each vibrational mode a harmonic oscillator of the
same frequency, one finds from (2-31) and Planck's formula (2-14) for
the vibrational energy of the crystal,
E=
r"}) Z(v)
.11
(kT)3
hI'
l'I m x dx
dl' = 9N h-. kT
--1
v D o e"-l
e''''lk1'
(2-34)
Sec. 2-6]
43
'l)
(I
x 3 dx
e~ - I = 6
t
CL
I
n4 =
15
for
1T4
so that
E
NOW,9
<
(j]) ( __
(2-36)
(0 /1'
I,
.0
(e -
(8..!.!.)
e X x4
I)
dx
= 3RFD
(2-38)
where FJ) is the Debye function. It has been represented in Fig. 2-4
together with the Einstein function. The reason that the Debye curve lies
above the Einstein curve is a result of the fact that in the Debye model,
the low-frequency modes are taken into account; at low temperatures
these have a higher average energy and temperature derivative than the
relatively high-frequency Einstein oscillators, as is evident from the
Planck formula (2-14).
To illustrate the agreement between the Debye theory and experimentally observed specific heat curves, we reproduce in Fig. 2-7 measurements on silver fitted to a Debye curve. From such curves it is possible
E. T. Whittaker and G. N. Watson, Modern
1935, p. 265.
Ana~ysis,
44
[Chap. 2
\1
Cv
t
1
'\' ~
40
80
120
160
200 240
--T
280
320
360
OIJ
Na
K
CII
Ag
Au
Be
150
100
315
215
170
1000
290
250
172
Mg
Zn
Cd
Solid
OIJ
Solid
On
Fe
Co
Ni
AI
Ge
Sn
Pb
420
385
375
390
290
260
C (diam.)
NaCl
KCI
KBr
AgCl
AgBr
CaF.
1860
281
230
177
PI
225
88
183
144
474
45
Sec. 2-6]
On
20
288
297
308
IS
10
KCI
10'C,IT'
0.388
0.356
0.334
14
8
4
3
213
222
236
227
Li
./
10'CvlT'
0.960
0.832
0.708
0.798
30
20
15
O/J
356
340
328
10'C .. IT'
0.101
0.118
0.131
'"
3.0
xlO- 3
"
:;..-,~
bI)
OJ
"0
OJ
-0
....
ta
2.0
--
.<:
-a'"
OJ
.g
h
--
1.0
cS
t
15
Fig. 2-8.
20
46
[Chap. 2
noted that in the Debye theory, the maximum frequency VJ) was common
to both the longitudinal and the transverse modes. Born proposed to
cut off the spectrum in such a manner that the longitudinal and transverse
modes have a common minimum ware/ength. This, as will becpme
evident from the discussion in the following sections, is actually more
sound theoretically speaking and in line with the theory of lattice vibrations
developed by Born and von Karman. 13 Thus if one takes the common
minimum wavelength equal to
Alliin
= (41TV/3N)l(3
(2-39)
one obtains two Debye frequencies, one for the longitudinal modes and
one for the transverse modes, viz.,
(2-40)
That this procedure leaves the total number of vibrational modes equal
to 3N follows immediately from (2-31) and (2-40), because
2 v2 dv
41TV ( ',. t"3
.0
c/
1 2 dv ) =
+ .0I,r" '--::lv
[/
3N
-I ""
,Sec. 2-8J
47
-/(x" --
f(x" - x,,-tl)
X,,_l) -
= /(X 71 - 1
+ X"+l
2x,,) (2-42)
e-i,o(l-"a1cs)
e-i(wt-qna)
(2-43)
= _/(e- iqa
+e
iqa
2)
= ,!(sin 2 (qaI2)
(2-44)
or
(,) =
wmnx
sin (q aI2)
with
w~laX
= 4flm
<
362 l
~'"
r-rc:
1"\1:' C A c l r
c::rfJ:'NrJ:'C::
48
[Chap. 2
'\
/\
\ b
bj
\
\
\
\
~
-2tr/a
I--- 2nd
-+----
tria
1st
- - - + - 2nd
2tr fa
..\
--l
qm
q + 2Trm!a with
111 =
I, 2, ...
(2-45)
-;..
..
r
Sec. 2-8]
49
the first zone as indicated in Fig. 2-10. Higher-order zones are defined in
a similar manner.
It is interesting to note that according to (2-44) there exists a maximum
frequency 1'lllilX which can be propagated through the chain, viz.,
V"",x
~
77
(L) 1/~
(2-47)
171
The chain may thus be considered a low-pass filter which transmits only
in the frequency range between zero and V lllax ' Tn contrast with this, the
continuous string has no frequency limit. The maximum frequency of the
chain of atoms occurs when the wave vector is equal to 7T/a, i.e., for a
wavelength Alliin = 2a. Now a'_' 10- 8 cm and the velocity of sound In
solids is of the order of 105-106 cm sec-I; this gives VllIilX -::::= 1013 sec-I.
2-9. Vibrational modes of a finite one-dimensional lattice of identical atoms
In the preceding section the discussion referred to an infinite lattice;
in the present section we shall see how the boundary conditions required
for a finite lattice lead in a natural manner to an enumeration of the
possible modes of vibration. The boundary conditions may be introduced
in either of two ways, which will now be discussed:
(2-48)
Here AI and A 2 are amplitudes, and (31' (32 are phase angles. The boundary
conditions are
xo(t) = 0 and xs(t) = 0 for all t
The first of these, when substituted in (2-48), requires Al = -A2 and
(32' Since the phase angles are equal, we shall choose (31 = (32 = O.
Taking the real part of the remaining solution, one obtains
~I =
(2-49)
= 0 or q = (7T/Na)j
(2-50)
50
[Chap. 2
where j is an integer. Note that j = 0 must be excluded, since this corresponds to q = 0, i.e., all particles arc at rest. The maximum value of q,
viz.,' 1Tla gives jllm = N; however, this value must be excluded for t'le
same reason as j = 0. We thus conclude that
j = 1,2,3, ... ,(N -
(2-51)
I)
(2~52)
= Xn+N(t)
eiqna =
eiq(n+N)a
or q
(21TINa)g
(21TjL)g
(2-53)
Sec. 2-9]
51
E = iM(dyJdt)2
+ Mw 2y 2J2
(2-57)
We shall now show that the energy associated with a vibrational mode
can indeed be written in the form (2-57). Let us consider a mode corresponding to a standing wave sin qna cos wt. The kinetic energy of the
particles in the lattice resulting from this vibrational mode is equal to
E kin
=-= 2m
Library
ANG RAU Central
d
Hili. Ne
Hyderaba
635
11111111111111111 \\III I111 1111
(2-58)
see also F.
52
[Chap. 2
where m is the mass per atom and the summation extends over aU particles
in the chain. The potential energy of the system due to the vibrational
mode q is a function of all coordinates Xn; let it be denoted by
V(xo, Xl' ... , X n , . ). The force exerted on patticle n is then, in accordance
with (2-42),
_ - \ .
oV
d 2x n
- ,
- d- = m d 2 = !(xn - 1 x n +1 - 2x,,)
(2-59)
xn
from this one may arrive at the following expression for the potential
energy
(Note that each of the mixed terms appears twice in the summation,
providing agreement between the last two equations; this may readily
be verified by writing out the sum explicitly.) Substituting the standing
wave solution int@ (2-60), one obtains after some manipulation,
(2-61)
Making use of the relation between wand q as given by (2-44), one may
write
(2-62)
v = imw2 cos2 wt 2: sin2 qna
n
E = tmw2S with S
2: sin 2 qna
(2-63)
(2-64)
tmw2S sin 2 wI
This expression is identical with (2-58), which proves the sought equivalence. The average energy associated with a particular mode of
vibration of angular frequency Wq is thus given by Planck's formula
(2-14), i.e.,
(2-65)
Sec. 2-101
53
_. 7T
In-fa
0
liwq
d
exp (liwqlkT) - I q
(2-67)
where the summation over the possible wave vectors defined by (2-50)
has been approximated by an integral. Employing the relation between
Wq and q as given by (2-44), one may replace dq by
dq dw =
2du)
dw
aw llH1X cos (qaI2)
2dw
a(w~lax -
(1)2)1/2
(2-68)
Hence
liw dw
E _ 2L (w max
[exp (liu)lkT) - l][w;nax - 7Ta.lo
(02]1/2
(2-69)
The specific heat as function of temperature may be obtained by differentiating with respect to T. The result is represented by the lower
curve in Fig. 2-11 for a critical temperature e = liwmaxlk = 200oK. It is
of interest to compare this result with the continuum theory corresponding
to the Debye approximation in one dimension. According to (2-24) the
54
[Chap. 2
E=
I'
w~.x
liw dw
1TC . 0
exp (liw/kT) - 1
Cv
(2-70)
f 1.0
.5
(L/1TC.)
20
40
60
fw~ax
Jo
dw
N
(2-71)
or
w:Uax = N1TC./L
Note that this limit is different from
that appearing in (2-69). The specific
heat calculated on the basis of (2-70)
is given by the upper curve in Fig.
2-11 again for a critical temperature
() = Iiw:Uax/ k = 200oK.
MX 2n
mX 2n +1 =
!(X2n
+ X 2n +2 -
2x 2n +1)
(2-72)
Ae- i (wt-2nQ a l
and
2n
+1 =
Be- i [w t -(2n+!lqal
(2-73)
55
Sec. 2-12]
are the amplitudes corresponding to particles of mass M and IJI, respectively. Substitution of the solution into (2-72) yields the following two
equations:
(Mw 2 - 2f)A
2Bfcos qa =- 0
~()B
(mu)2 -
+ 2Af cos qa =
(2-74)
This system flas nonvanishing solutions for A and B only if the determinant
of the coefficients of A and B vanishes, i.e.,
(Mw
2f)
~lcos qa
2f cos qa
(/17('j2 -- 2f)
I
=
(2-75)
This gives for the square of the frequency the following two possihilities:
(,,2
2qa]1/2
(I + MI) f [(I;;; + M1')24Sin
Mm
(2-76)
=f ;;;
[2f(~+ ~)]1/2
and
'"
(2f/mjl/2
(2f/M j l/2
-11'
/2a
11'
/2a
(')_=O
for q=O
(2-77)
From the form of (2-76) it is observed that here, as in the monatomic case,
the frequency is a periodic function of the wave vector. The first zone
thus limits the values of q to the range between -1TJ2a and +1TJ2a as
shown in Fig. 2-13. For q = 1Tj2a, the two angular frequencies are
evidently
w+
= (2fJm)1/2 and
w_
The complete curves for 0+ and (O_ versus q are illustrated in Fig. 2-13.
The larger the mass ratio MJm, the wider the frequency gap between the
two branches. The existence of a "f.k)rbidden" f~~~l'-JWl:""IWWJ-.__
, " " l ) eEl>! d{AL LIBRAR Y
56
(Chap. 2 ,
A=B
-MA =mB
for q = 0
(2-79)
Wopticai
~ [ 2/ ( ;;;
+ M1 )] 1/2
in accordance with (2-77). It is for this reason that the upper branch is
called the optical branch. The infrared absorption thus corresponds to a
vibration of the positive ion lattice relative to the negative ion lattice such
that the center of gravity in each cell remains at rest.
)ec. 2-13]
57
r-.-----,_..-_ _
8D
120
10
20
30
40
50
60
70
80
-T('K)
Fig. 2-15.
simple cubic latticeY From these results he was able to calculate the
specific heat in the manner outlined for the one-dimensional case in
Sec. 2-11. In the low temperature region the specific heat thus obtained
may be equated to the Debye formula (2-37) and the Debye temperature
On can be computed for different temperatures. The results obtained by
Z(wl
t
260
1.2
2.4
_
3.6
4.8
wXlO-13
(a)
6.0
20
40
60
-
80
100
T('K)
(b)
Blackman for the simple cubic lattice are represented in Fig. 2-15. It is
observed that ()n is by no means constant, indicating the possibility of
appreciable deviations from the Debye theory in actual crystals.
16 See M. Born and M. Goppert-Mayer, op. cit.; F. Seitz, op. cit.; L. Brillouin,
op. cit.
17 M. Blackman, Proc. Roy. Soc. (London), A148, 384 (1935); A159, 416 (1937);
PrOf. Cambridge Phil. Soc., 33, 94 (1937).
58
[Chap.2-'
REFERENCES
M. Born and M. Goppert-Mayer, Handbuch der PhYSik, 24 (2) (1933).
M. Born and K. Huang, Dynamical Theory of Crystal Lattices, Oxford,
New York, 1954.
M. Blackman, Reports on Progress in Physics, 8, 11 (1941).
L. Brillouin, Wave Propagation in Periodic Structures, Dover, New
York, 1953.
PROBLEMS
2-1. (a) Give a derivation of expression (2-4) for the difference
between C p and C v . (b) Calculate C p - C v per mole of sodium at room
temperature if at this temperature the compressibility of sodium is
12.3 X 10- 12 cm 2 dyne- I and the linear coefficient of expansion is
6.22 X 10 -5; compare the result with Cp - C v for a monatomic gas.
Also calculate the Griineisen constant for Na.
2-2. The possible energy levels of a rigid rotator according to quantum
mechanics are given by En = (fj2/2J)n(n
I) where J is the moment of
inertia and n = 0, 1,2, .... For the molecules H2 and Cl 2 calculate the
energy difference between the ground state and the first excited state for
rotation about an axis perpendicular to the line joining the nuclei.
(Answers. Resp., 14.7 X 10-3 and 0.06 X 10- 3 ev.) Also estimate the
value of E1 - Eo for rotation about the line joining the nuclei and show
18 E. W. Kellermann, Phil. Trans., A238, 513 (1940); Proc. Roy. Soc. (London),A178,
17 (1941).
1('
Chap. 2]
59
that this rotation does not in general contribute to the rotational specific
heat. At which temperatures for H2 and Cl 2 do quantum effects enter in
the rotational specific heat? If it is given that the number of possible
states corresponding to an energy level En for a rotator is equal to
2n(n
1), show on the basis of statistical mechanics that the rotational
specific heat for a molecule such as Cl z at room temperature is R cal per
mole. (Hint: According to statistical mechanics the average energy at T
is given by
(E)
It
where Zn is the number of possible states associated with En- For the
problem under consideration one can replace the summations by
integrals. )
Chapter 3
SOME PROPERTIES OF METALLIC LATTICES
3-1. The structure of metals _.
Most metals crystallize in one of the following three structures:
the body-centered cubic lattice (b.c.c.) in which each atom is surrounded
by eight nearest neighbors, the face-centered cubic lattice (f.c.c.) in which
a given atom has twelve nearest neighbors, and the hexagonal close
packed lattice (h.c.p.), also with a coordination number of twelve.
From the dimensions of the elementary cell, as obtained from X-ray
diffraction or otherwise, one may define a radius for the atoms on the
assumption that they are spherical in shape; the radius so defined is then
given by half the distance between nearest neighbors. That this procedure
has a physical meaning follows from the fact that for those metals which
crystallize in more than one structure, each structure being stable over
a certain range of temperatures, the radii so obtained are very nearly the
same. Table 3-1 gives the distances of closest approach (twice the atomic
Table 3-1. Structure and Distance of Closest Approach (at 20'C) (or Metals
which Crystallize in Any o( the Three Simple Metallic Structures. The
asterisks indicate the normal form.
I
Body-centered
cubic
Metal
d(A)
Li
3.039
3.715
4.627
2.632
2.860
2.498
2.725
Na
K
V
Ta
Cr
Mo
ex W*
at Fe"
r5 Fe (1425 C)
2.739
2.481
2.54
Face-centered
cubic
Metal
Cu
Ag
Au
Al
Th
Pb
y Fe (extrapolated)
fJ Co
Ni
P Rh*
Pd
Ir
Pt
60
d(A)
2.556
2.888
2.884
2.862
3.60
3.499
2.525
2.511
2.491
2.689
2.750
2.714
2.775
Hexagonal
close packed
d(A)
Metal
Be"
Mg
Zn
Cd
at TI*
C( Ti*
2.225
3.196
2.6E4
2.979
3.407
2.89
Zr*
Hf
oc Co"
Ru
Os
3.17
3.1 )
2.506
2.649
2.675
C(
Ot
Sec. 3-1)
61
radii) for metals which crystallize in one of the three structures mentioned
above. 1
The b.c.c. and f.c.c. lattices have been represented in Fig. 1-4. The
h.c.p. structure represented in Fig. 3-1 is closely related to the f.c.c.
structure, as may be illustrated with reference to Fig. 3-2. Let the dots in
Fig. 3-2 represent a layer of spheres in close packing. On top of this we
place another layer, represented by the crosses. The atoms of a third
layer may now be placed on top of the second one in either of two ways:
(1) they can be placed in positions corresponding to the open circles in
Fig. 3-2, or (2) they can be placed in positions identical in projection with
Fig.
close
,';
'h'
".~" "
First layer
)( Second layer
or 0 Third layer
62
[Chap. 3
may be made with reference to those metals which have different structures
in different temperature regions (allotropy). This phenomenon is exhibited
especially by the three- and four-valent metals and by the transition
metals. 3 For example, IX Fe (b.c.c.) is stable up to 910 o e; between 910 0 e
and l400 e the stable structure is y Fe (f.c.c.); between l400 e and the
melting point (l530C) the structure is again b.c.c. (a Fe). Here again, the
transformation from one structure to another is dictated by the requirement
of minimum free energy. This does not mean that such transformations
take place as soon as the existing structure becomes unstable. ]n fact, a
transformation of structure involves a rearrangement of atoms, and such a
process may take a long time. The reason is that even though the free
energy after the transformation is lower than in the initial state, the two
states are usually separated by an energy barrier or activation energy
(see Sec. 3-5). Thermodynamics specifies only the equilibrium condition
but does not give any information about the velocity of the reaction or
processes involved in establishing equilibrium. From the atomic point of
view, the stability of crystal structures is a problem of cohesive energy,
involving the interaction between the atoms. A brief discussion of the
cohesive energy of metals is presented in Sec. 10-13 based on the electron
theory of metals.
0
d,.r
Sec. 3-2)
63
atoms occupying positions in the lattice which in the perfect lattice would
be unoccupied. In discussions of this kind it is necessary to point out the
distinction between what we shall refer to
= 3Nkfl
+ log (kT/hv)]
(3-2)
= (N"
+ Nt')!
N,,!Nb!
':-'
(3-3)
64
[Chap. 3
Fo~ a perfect crystal containing identical atoms and in the absence of any
lattice defects, w,.[ = 1 and Se[ = 0 because there is only one possible
arrangement of the atoms. The total entropy occurring in the usual
thermodynamic formulas is equal to the sum of the thermal and
configurational entropies, i.e.,
\
\
(3-5)
The results obtained above may be used to explain qualitatively the
reason for the existence of lattice defects at any temperature T > o.
Suppose, for example, that in a perfect metallic crystal we produce a
certain number of vacant lattice sites by transferring atoms from the
interior of the crystal to the surface. This will require a certain amount of
energy, i.e., E increases. Consequently F
increases and this by itself is thus unfa vorable in the thermodynamic sense. On the
other hand, the creation of the vacancies
increases the disorder in the crystal and
thus increases the configurational entropy
from zero to a certain value determined
by the number of vacancies n produced.
In fact, according to (3-4) the configurational entropy associated with the
-TScf
possible arrangements of N atoms and n
Fig. 3-4. Schematic representavacancies over a total of (N
11) latrice
tion of the energy and the consites is
figurational entropy term as
(N
function of the fraction of vacant
(3-6)
Sef = k log [
"
N.n.
+ n)!]
Sec. 3-3]
65
+ ncpv -
nT!J.Sth - kTlog
(N
+ n)!
N! n!
(3-7)
In order to find the equilibrium value of n, we make use of the fact that in
equilibrium (o}/dn},p = O. Employing Stirling's formula m the form
log x! ,..._, x log x for x ~ 1, we find from (3-7),
(3-8)
l'
66
[Chap. 3
The thermal entropy of the imperfect crystal is then, in analogy with (3-2)
Sth
3nxk (I
(3-9)
Subtracting (3-2) from (3-9) and dividing the result by n, one finds for the
increase in thermal entropy per produced vacancy,
~Stll = 3xk
log (vly')
(3-10)
<Pv
= <Pvo (1
- ocT)
(3-12)
and one argues that if it were possible to measure nlN, a plot of log (nIN)
versus 1jkT would give <Pvo rather than CPv; furthermore, it is argued that
the pre-exponential factor is multiplied by exp (cxCPvo/k]. These arguments
are, however, incorrect since they neglect the temperature variation of ~S
accompanying the change in <Pv. 8 In fact, for zero pressure we have
d(~S)/dT
(lIT) dcpJdT
Sec. 3-3)
67
n <{ N
(3-13)
.. . I
+ Interstitia
(3-14)
From this, readers familiar with the law of mass action will readily
recognize that n should be proportional to (NNi )1/2 and that the
exponentials in (3-13) are correct.
3-4. The formation of lattice defects in metals
68
[Chap. 3
of the lattice site "occupied" by the vacancy (at least as long as it does not
become a nearest ncighbor of another vacancy or lattice defect). This is
represented schematically in Fig. 3-5b. It should be kept in mind, however, .
that although the energy before and after a jump may be the same, a
certain "activation cnergy" is always required to make the jump. In other
words, two possible neighboring lattice sites for a vacancy are separated by
a barrier, as indicated by E j in Fig. 3-5b. It is for this reason that the
A~
2(.c
4(Y
D.
_}
--+ Position
(al
(bl
Sec. 3-4]
69
atoms in the crystal be --O i . The total dissociation energy of the crystal E
is then equal to NOJ2, the factor 2 arising from the fact that the interaction
. energy between any two atoms should be counted only once; thus,
( i = 2O", which means that an atom at the surface is, on the average,
bound half as strongly as an atom in the interior. The physical meaning
of Oi may also be expressed in this way: it represents the energy required
to remove an interior atom to infinity if the position and the charge distribution of the other at0111s remain unchanged. The energy required to form a
vacancy, i.e., the energy required to transfer an atom from the interior to
the surface, may then be written in the form
cPv ==
Ei -
Er -
==
Es -
Er
(3-15)
70
[Chap. 3
""!\
\
( 3-16)
1= -D grad n
where I is the flux of atoms in cm- 2 sec-I, n is the number of atoms per cm3 ,
and D is the diffusion coefficient. In general, D is itself a function of the
concentration n. 12 Expression (3-16) is known as Fick's first law; the
minus sign indicates that the current flows from regions of high concentration to regions oflow concentration. Applying the continuity equation,
one obtains Fick's second law:
"[njct
DV 2n
(3-17)
p,867,
13 L. Onsager, Ann. N. Y. A cad. Sci., 46, 241 (1945),
HA, D, Ie Claire, Progress ill Metal PhYSiCS, Interscience. New York, Vol. I (1949),
Vol. 5 (1954).
J" G. 1. Dienes, 1. Appl. Phys., 21,1189 (1950).
Jij C. Zener in W. Shockley (ed.), Imperfections ill Nearly Perfect Crystals, Wiley,
New York, p. 299.
Sec. 3-5]
71
From the atomic point of view, the simplest type of diffusion in solids is
the diffusion of interstitial atoms. The reason is that in this case there
exists no doubt as to the actual atomic mechanism involved; the interstitial atoms presumably jump from one interstitial position to a neighboring
one. The diffusion of hydrogen, oxygen, nitrogen, and carbon in iron and
other metals are examples of this mechanism. In order to discuss this
type of diffusion from an atomic point of view, consider a set of parallel
atomic planes of interplanar distance A. We shall assume that there exists
a concentration gradient of the diffusing particles along the x-axis which is
perpendicular to the atomic planes. An atom in an interstitial position may
jump in the positive x-direction (forward) in the negative x-direction
(backward) or it may jump in a direction perpendicular to the x-axis.
We shall denote the probability for a given interstitial atom to make any
jump per second by p. Actually, the probability for a jump depends on the
probability that the neighboring interstitial site will be empty. We shall
assume, however, that the fraction of interstitial positions which is
occupied is <{I, so that p may be considered independent of the concentration of interstitials. 17 The probability for a jump per second in the
forward direction will be denoted by fp; furthermore, we shall assume that
the probabilities for a forward and backward jump are equal. The diffusion
problem is then reduced to a simple random-walk problem.
Denoting the number of diffusing particles per cm2 on the plane
located at x at the instant t by n(x) we have
n(x + ?.)
n(x - ?.)
+ (en/ex)?. + t,(2;2n/ax2)?.2 + .. .
n(x) - (an/ax). + ! (2Pn/ex 2)?.2 + .. .
~)
n(x)
(3-19)
Thus, when we consider the situation at the instant t + bt where bt <{ lip,
the increase bn of the number of particles on the plane located at x is
given by the number of particles jumping from (x - ?.) into x, plus the
number of particles jumping from (x + ?.) into x, minus the number of
particles jumping away from plane x. Since we have assumed bt <{ lip, it is
not necessary to consider other planes besides the three employed. Hence
e2n
bn(x) = fp bt eX2?.2
or
en
at =
e2n
fp?.2 ex2
(3-20)
72
coefficient must enter via the jump probability p. The simplest model that
can be set up to determine the temperature dependence of p is to consider a
particle moving in a fixed potential energy curve of the type illustrated in
Fig. 3-6. Let the potential minimum A correspond to the interstitial
position in which the particle finds
itself, and let B correspond to a
neighboring interstitial position. The
barrier of height Ei is a result of the
fact that as the particle moves from one
interstitial position to another it is
A
B
squeezed between the atoms conFig. 3-6. The energy barrier bestituting the host lattice. Assuming
tween two interstitial positions.
the potential to be parabolic, the atom
will vibrate as a harmonic oscillator.
The frequency of vibration Vi may be considered as the number of
attempts per second made by the particle to cross the barrier. However,
any attempt can succeed only if the energy of the particle is ?Ei' As
shown in Problem 3-6, the fraction of time spent by the particle in energy
states ;?:Ei is simply given by exp (-EJkT). Hence, for the probability of a
jump from A to B we find per second,
(3.22)
When the jumping problem is considered more rigorously than has been
done above, one obtains a formula of the same form ls as (3-22), but Ei is
then replaced by a free energy I1Fi = ( i - TI1Si , i.e.,
(3-23)
where I1Si is the entropy difference between the state in which the particle
is halfway between A and B, and the state in which the particle is in A.
Since an interstitial atom may jump into more than one neighboring
position, p is obtained by summation of (3-23) over all Pi' From (3-21) and
(3-23) we thus obtain
(3-24)
Let us now apply the results obtained above to a specific case. In Fig. 3-7
we have represented the diffusion coefficient of carbon in oc iron (b.c.c.)
as function of temperature according to WertJ8 Note that equation (3-18)
is satisfied for D-values covering 14 cycles of 10, with
.
1
z
(3-25)
E = 0.874 ev
and Do = 0.020 cm secThe interstitial positions in a b.c.c. lattice are indicated in Fig. 3-8;
they correspond to the centers of the faces and edges of the elementary cube.
18
For details, see C. Wert and C. Zener, Phys. Rev., 76, 1169 (1949); C. Wert, Phys.
..
.,.
-_
[Chap. 3 'lp..
Sec. 3-5]
73
~
I
8.....
-2
-6
;::;-
,/
U
0>
/
v
'"
""S
E. -10
Q
bIl
..s
.....
-14
-18
/
3.8
3.0
2.2
1.4
.6
10 3 fT
Fig. 3-7.
f =i
X t = -!. Since the four possible jumps from any interstitial position
are equivalent, we obtain from (3-24) with A = a12, where ais the cube edge,
(3-26)
Comparison of (3-18) and (3-26) shows that for interstitial diffusion of
the type under consideration, the activation energy for diffusion is identical
with the activation energy for the atomic jumps. The value of exp (tl.Sjk)
may be estimated as follows: for (X iron a = 2.86 A; putting y ~ k(}jh
where () = 420 K is the Debye temperature of iron, one obtains y ~ 1013
sec-I. From the known value Do = 0.020 cm 2 sec-lone then obtains
exp (tl.Sjk) ~ 7.
We may mention here that Zener has derived the following approximate
relationship :19
(3-27)
0
19
74
[Chap. 3
/'
I
}:(
I /
\
1//
~D----~-
.-
,/
,/
0.20 cm 2 sec-I,
2.05 ev
(Cu)
coefficient of Na.
Sec. 3-6J
75
e!!.S,./ke -.pjkl'
where ~SV refers to the thermal entropy change associated with the
creation of a vacancy. The probability for a jump of a vacancy to a
nearest neighbor site is given by a formula of the form (3-23). Hence the
self-diffusion coefficient for the vacancy mechanism may be written as
(compare 3-26):
(3-28)
where the subscripts j refer to jumps; y is a numerical factor determined
by the geometry of the lattice. Note that the
activation energy for diffusion is in this case
given by the sum of the energy of formation
of a vacancy and the jump activation energy,
i.e.,
(3-29)
We may recall that Overhauser lO found for
copper CPv = 1.39 ev and E; = 0.68 ev, the
sum of which is 2.07 ev. This is in good
agreement with the experimental value of E
quoted above. Huntington and Seitz 9 calcu- Fig. 3-9. Illustrating the twolated the activation energy for the direct inter- ring (direct interchange bechange (see Fig. 3-9) between two neighboring tween two atoms) and the fourring diffusion mechanisms.
atoms in copper and found a value four times
larger than the observed value. These authors
also found that the energy required to transfer a surface atom to an interstitial position requires an energy of nearly 13 ev. It thus seems that
the interstitial and direct exchange mechanism are very unlikely in eu;
the vacancy mechanism is evidently operating in this case. Nachtrieb 21
et al. believe that the vacancy mechanism is also responsible for the selfdiffusion in sodium.
We should mention that besides the two-ring direct interchange referred
to above, there are other possibilities of direct interchange involving more
than two atoms. In Fig. 3-9, for example, we have indicated a four-ring
mechanism investigated by Zener. 22 Zener has shown that the activation
22
76
[Chap. 3
\.
Very low solute concentration. This is the simplest case since the
interaction between the impurity atoms may be neglected; furthermore,
complications arising from lattice defects associated with high-impurity
densities are avoided (see below). In order to illustrate the type of problems
encountered in this case, consider a certain metal A which is known to have
a self-diffusion governed by the vacancy mechanism. Suppose a very small
fraction of A atoms is replaced by B atoms and let us inquire about the
diffusion coefficient DBA of the B atoms in the A lattice. When PBv
represents the probability for a B atom to have a vacant lattice site as a
nearest neighbor and PBi represents the probability per second for a B
atom neighboring a vacancy to jump into the vacancy, then DBA will be
proportional to the product P BvP Bj' In a similar notation, the coefficient of
self-diffusion D.1A of the A lattice is proportional to PAvP Ai' Hence
DEA/ D AA
(PBv/P.1v)(PBi/PAj)
(3-30)
Sec. 3-7]
77
Ag
Cd
In
Sn
41.70
0.454
40.63
0.416
39.30
0.255
Sb
Do
45.50
0.724
II
38.32 keal/mole
0.179 em' /see
78
[Chap. 3
system. 28 They found that as the diffusion progressed, the markers moved
towards each other. The fact that the displacement of the m2.rkers was
proportional to the square root of the time indicated strongly that the
marker movements were related to the diffusion process itself.29 The
Kirkendall effect has since been found in many other systems; for example,
da Silva and Mehl have observed the effect in eu-Zn, Cu-Sn, Cu-Ni, Cu-Au,
Ag-Au. 30
Assuming that the markers are fixed relative to the system of lattice
sites, and assuming that the diffusion is governed by a vacancy mechanism,
a mass flow of atoms in a given direction must be compensated by a flow of
vacancies in the opposite direction. Thus in the copper-brass system, the
net flow of atoms out of the brass is balanced by a flow of vacancies from
the copper into the brass. For an excellent treatment of the theory of the
Kirkendall effect, the reader is referred to a papcr by Bardeen and Herring
in W. Shockley (ed.), Imperfections in Nearly Perfect Crystals, Wiley,
New York, 1952. Dislocations play an essential role in the atomic theory
of the effect as sources and sinks for vacancies (see Sec. 3-12).
3-8. The elastic constants of metals
For further reference and as an introduction to the following sections
of this chapter we may review very briefly some of the fundamental
principles of elastic stress-strain relations in crystals. 31 Let us first consider
an isotropic elastic medium under uniform stress along an arbitrarily
chosen x-direction. Let x' represent the distance of a given atom in the
material under stress relative to a fixed plane perpendicular to the x-axis.
When x represents the distance of the same atom in the unstressed material,
the strain Ex is defined by
EX
= (x' -
x)/x
(3-31)
29
Sec. 3-81
79
The proportionality factor E is called Young's modulus. When ax represents a tensile stress, there will be a contraction of the material perpendicular to the x-axis such that
(3-33)
IX
(3-34)
=T/G
(3-35)
where G is called the elastic shear modulus. It can be shown that, for
isotropic bodies, the three quantities E, l', and G satisfy the relation
G = E/2(y
1)
(3-36)
80
[Chap. 3
~----
axx
TXY
Tx.
.;.~
(1"ff
1',..
'1'e
'1'.11
However, the reaqer will readily convince himself that if rotation is absent,
y
U xx
--i---
Fig. 3-11. Illustrating the three couples of forces acting along the
x-direction; an is a tensile stress, Tx. and Txz are shear stresses;
'Tx. represents a force acting along the x-axis in a plane perpendicular
to the y-axis, etc. Similar forces act along the y- and z-directions.
the tensor must be symmetrical, i.e., TyX = TXY etc. The stress condition
may thus be specified by six independent stresses,
(3-37)
As a result of the stresses, the crystal is strained, i.e., an atom which in the
unstrained crystal occupied the position x, y, z will in the strained crystal
occupy the position x', y', z'. When the distortion is homogeneous, the
displacements are proportional to x, y, z and we have in analogy with
(3-31) the more general expressions
x' - x
y' - y
z' -
+ YXyy + yx. Z
= YyXx +
+ YyZz
= YzxX + YZyy + EzzZ
=
ExxX
EY?lY
(3-38)
where the E'S and Y's refer to normal strains and shearing strains, respectively. The strain tensor is again symmetrical if rotation is absent and the
~
Sec. 3-81
81
strain condition of the cube may be specified by the six strain components
(3-39)
When Hooke's law is satisfied the strain and stress components (3-39) and
(3-37) are linearly related. Thus in analogy with (3-32) we have, for
example,
axx = Cll xx + C12 YY + C13 zz
C14Yvz
C15Yzx
c1SYxv (3-40)
There are six such equations, and hence 36 moduli of elasticity or elastic
stiffness constants Ci ;.32 The relations, which are the inverse of type (3-40),
express the strains in terms of the stresses; for example,
(3-41)
The six equations of this type define 36 constants Si) which are called the
elastic constants. It can be shown that the matrices cij and Sij are symmetrical; hence a material without symmetry elements has 21 independent
elastic constants or moduli. Due to the symmetry of crystals, several of
these may vanish. In cubic crystals, as mentioned above already, there are
three independent elastic moduli which are usually chosen as Cn , C12 ' and
Cw Some representative values for cubic metals are given in Table 3-2.
The atomic theory of elasticity is based on the forces acting between the
atoms; we refer the reader to the books quoted at the end of this chapter
for a discussion of this subject.
Table 3-2. Elastic Moduli for Some Cubic Metals
in 10 1 dyne/cm'
Metal
Structure
C l1
C 12
";1.1
1.08
1.70
0.48
0.046
2.37
0.62
1.23
0.41
0.037
1.41
0.28
0.75
0.14
0.026
1.16
---"
AI
Cu
Pb
K
Fe
f.c.c.
f.c.c.
f.c.c.
b.c.c.
b.c.c.
82
[Chap. 3
P"'I~LLZ:Jl-Pull
Sec. 3-9)
83
= (F/A)
Slip
direction
(F/ A) cos 2 rx
(3-43)
')
84
[Chap. 3
\
x
T
---~T
,al
(b)
Fig. 3-14. Under influence of the shear stress T the upper plane of
atoms in (a) is displaced over a distance x (dashed circles). The
periodic behaviour of T, according to Frenkel, is indicated in (b).
Tc sin (21Tx/a)
(3-44)
The "amplitude" Tc is evidently the critical shear stress in this model and
it is this quantity that we wish to estimate. This may be done by realizing
that for x ~ a the usual theory of elasticity should apply; under these
circumstances
T(X) c::: (21Tx/a)Tc
for x~a
On the other hand, the elastic strain, in accordance with (3-34) and (3-35)
is given by
y = x/d = TIG
where G is the shear modulus. From the last two equations it follows that
in this model
(3-46)
Tc c::: (G/21T)(a/d) c::: G/21T
where the last approximation is justified because a c::: d. Since G c::: 1011
dynes per cm 2 (see C44 in Table 3-2), one obtains in this model a theoretical
shear stress Tc c::: 10 10 dynes per cm 2, which is several orders of magnitude
larger than the observed ones. Although it must be admitted that Frenkel's
model is open to objections, more refined calculations confirm the conclusion that it is impossible to obtain agreement between theory and
Sec. 3-10)
85
experiment on the basis of a model where atomic planes glide past each
other in the manner assumed above (see also Problem 3-12). In Fig. 3-14
it was assumed that the atoms of the upper atomic plane move simultaneously relative to the lower plane; this assumption is tied up with the
assumption of a perfect lattice and here we are at the root of the difficulty.
In an attempt to remove this difficulty, let us assume that the crystal
contains an imperfection of such a nature that the slip process is governed,
not by the simultaneous motion of the atoms of one plane relative to
another, but by the consecutive motion of these atoms. Before specifying
this model further, it may be useful to remind the reader of the fact that a
worm moves forward by displacing its segments one after the other rather
than by a simultaneous displacement of all the segments. The atomic
model for the progression of slip, based on the dislocation model, is
analogous to a wormlike motion.
The dislocation model for slip may be introduced with reference to the
crystal of Fig. 3-15a; let the plane PQR be a slip plane. This plane has
been redrawn in Fig. 3- I 5b. In the slip plane consider an arbitrary closed
curve ABC; the n:gion inside this curve is hatched in Fig. 3-15b. Suppose
now that in some way or other the material located over the hatched area
in the upper half of the crystal is displaced by an amoum b relative to
the lower half of the crystal; at the same time, the material in the upper
half lying over the area outside ABC is left undisplaced. In this manner we
have obtained a situation in which only a fraction of the upper half of the
crystal has slipped relative to the lower half. The ratio / of the area ABC
and the total area of the slip plane will be referred to as the fraction of
slip that has occurred in this plane. Thus, if in some way or other the
area ABC could be made to grow, / would increase and for / = I the
whole upper half of the crystal would be displaced by an amount b
relative to the lower half. For / < ], the average displacement of the upper
half relative to the lower half is/b.
The line ABC introduced above marks the boundary in the slip plane
between slipped and unslipped material; this line is called a dislocation
line. The vector b which defines the magnitude and direction of the slip
is called the Burgers vector. 34 Since the atoms always seek positions of
minimum energy, it will be evident that b must connect two atomic
equilibrium positions, i.e., the possible vectors b are determined by the
crystal structure. When the displacement equals one lattice spacing, the
dislocation is said to have unit strength. From calculations of the strain
energy associated with dislocations, Frank has shown that dislocations of
strength larger than unity are in general unstable; they dissociate into
dislocations of unit strength. 35
So far, we hav~ only given a definition of a dislocation. In order to see
3. J. M. Burgers, Proc. Koninkl. Ned. Akad. Wetenschap., 42, 293, 378 (1939).
35
'
, .'
86
[Chap. 3
how this model may account for a number of observations on plastic flow,
various questions must be raised; for example:
I. Assuming that a slip plane
such as PQR in Fig. 3-15 contains
a dislocation as indicated, one should
be able to show that under influence
of a shear stress, applied in the
proper direction to the crystal, the
dislocation tends to grow; under
la)
Ib)
these circumstances the slipped region
would
increase in size and slip would
Fig. 3-15. Schematic representation of
proceed.
Moreover, the calculated
a ring dislocation ABC in a slip plane
critical shear stress should agree
PQR. Slip has occurred only across the
hatched area.
quantitatively with the observed
values.
\
2. We have seen above that slip in a single slip plane may correspond
to displacements of the order of 1000 A; on the other hand, once a dislocation such as ABC in Fig. 3-15 has swept through the whole slip plane,
the slip produced is only b c:::: 2 A, and moreover, the dislocation has then
disappeared. It will thus be necessary to account for large numbers of
dislocations taking part in the slip process and for sources which supply
such dislocations.
3. Are other physical properties, besides plastic flow, determined to at
least a measurable degree by the presence of dislocations so that independent information regarding the properties of dislocations can be obtained?
Some of these questions will be discussed below; of course, many
more can be asked.
3-11. Motion of dislocations under influence of a uniform shear stress;
dislocation density
With reference to Fig. 3-15, suppose a uniform shear stress T is applied
to the crystal along the direction of the Burgers vector. Mott and Nabarro
have shown that this leads to a force on the dislocation line such that the
slipped area tends to groW. 36 Consider an element ds of the dislocation
line; suppose this element is displaced outwardly (Fig. 3-15b) by an
amount dt along a direction perpendicular to ds. The area swept out by
the line element is then ds dl. According to what has been said in the preceding section, this corresponds to an average displacement of the upper
part of the crystal relative to the lower part by an amount ds dt bfA, where
A is the area of the slip plane. The work done by the shear stress is equal
36
(LOlldol1). 1948,
~
...: I
't')
,,.
Sec. 3-11]
87
to the total shear force TA times the average shear displacement, i.e., equal
to Tb ds dl = dW. This corresponds to a force -dWjdl acting on the
element ds in the direction of the normal. Hence the force per unit length
is equal to
(3-47)
F=Tb
Thus the applied shear stress produces a force per unit length everywhere
along the dislocation line equal to Tb and perpendicular to the line element.
If the force is large enough to make the dislocation line move in the direction of F, the slipped area in Fig. 3-15 will grow and slip will occur under
influence of the shear stress.
On very general grounds one can show that the critical shear stress for
slip should be very small for the dislocation model. In order to see this,
let us consider the regions near the dislocation line somewhat more
closely. Because of the nature of interatomic forces, the boundary
between the slipped and unslipped regions is not sharp, but rather vague,
extending over several atomic distances. The atoms near the dislocation
line of Fig. 3-15, at the inside, have nearly completed the slip process;
those near the dislocation line on the outside are just beginning to slip. As
a result of the periodic nature of the potential for the atoms, those at the
outside of the dislocation line and close to it tend to push the dislocation
line inward, since this would allow them to occuPY their initial equilibrium
positions. On the other hand, the atoms inside the dislocation line and
situated close to it tend to push the line outward, since this would make it
possible for them to occupy their new equilibrium positions associated
with completed slip. Far away on either side of the dislocation line, the
atoms occupy normal lattice positions and are not affected by the dislocation. Thus to a first approximation, the forces on the dislocation line
balance and it should start moving under the smallest of shear forces.
It thus looks as if this model is too successful in explaining the relatively
low observed critical shear stress; however, when one goes to a second
approximation, one finds that the critical shear stress calculated for this
model is not zero, but in fact of the same order of magnitude as observed
values. 37
Density o/dislocations. It was mentioned above that a single dislocation
line sweeping across a slip plane gives rise to a displacement of the order of
a few Angstroms; thus any appreciable plastic deformation must be the
result of a large number of dislocations sweeping across many slip planes.
It will be evident that the rate of pla~tic flow will be determined by the rate
at which dislocation lines sweep through the slip planes, i.e., the rate of
flow may be expected to be proportional to the total length of all active
dislocation lines and the average velocity with which the elements of these
lines move. One has therefore introduced the concept of "dislocation
37
88
[Chap. 3
density," p = S/V, where Sis the total length of the dislocation lines and V
is the volume of the crystal. Note that p has the dimension length-2
More specifically, one arrives at this concept by the following reasoning:
Consider an element ds of a dislocation line such as ABC in Fig. 3-15.
Let v be the velocity of the element along the direction of the normal to ds
in the slip plane. When H is the height of the crystal and A is the area of
the slip plane, the increase in strain per second due to the motion of the
element ds is equal to
(3-48)
dyldt = v ds blAH
Considering the rate of flow resulting from all dislocations in planes
parallel to the plane PQR in Fig. 3-15, we have to sum expression (3-48) in
a suitable fashion, i.e., we must replace ds by the total length S of all these
dislocations and v by some average velocity (v). Hence
>
'.
dyjdt
(3-49)
Sec. 3-12]
89
D
(al
Ibl
The edge dislocation may then be obtained by cutting the block along the
plane EFGH, and putting the half plane of atoms initially above AB,
inside the cut. This gives rise to the "extra" half plane of atoms corresponding to HE in Fig. 3-16b, which is typical of an edge dislocation.
Note that if the extra half plane HE were displaced to the right, slip would
progress, and when HE has finally reached the right-hand side of the
block, the upper half of the block has completed slip by the amount b.
The slip process resulting from a
moving edge dislocation has been
J...
illustrated in Fig. 3-17. Edge dislocations for which the extra half
plane lies above the slip plane are
called positive. If the extra half Fig. 3-17. Motion of a positive edge
plane lies below the slip plane, one dislocation to the right, leading to slip.
speaks of a negative edge dislocation.
[After Taylor]
We leave it to the reader to show
for himself that the slip process of Fig. 3-17 resulting from a positive
edge dislocation moving to the right can also be achieved by motion of a
negative edge dislocation of the same strength to the left.
The definition of an edge dislocation does not necessarily imply that
the dislocation line is straight. In fact, any curved line will do as long as it
ClJ Gi
-. ... .
. ....
.... .. .. .. .. ...
... ..
90
[Chap. 3
Sec. 3-12]
91
92
[Chap. 3
for edge and screw dislocations. 41 In Fig 3-21 consider the cross section of
a cylindrical piece of material; the axis of the cylinder will be taken as the
z-axis of a Cartesian coordinate system. Suppose we produce a cut in the
plane y = 0, which extends between the axis and the outer surface as
indicated. We now let the material above the cut slip to the left by an
amount b, leading to the configuration
Y
indicated by the dotted line. We have
then produced a positive edge dislocation
along the z-axis with a Burgers vector
along the x-axis; the plane y = 0 is the
slip plane. In terms of the coordinates r
and e, the stress field of the dislocation
x
line may then be shown to be given by the
following tensile and shear stresses:
(Jrr
(Joo
T r6
TOr
= -
Gb
. ()
217r(1 _ v) SIll
Gb
217r( 1 -
~)
COS
(3-50)
(3-51)
For further details and references, see A. H. Cottrell, Di,'locations and Plastic Flow
ill Crystals, Oxford, New York, 1953; W. T. Read, Dislocations in Crystals, McGraw-
Sec. 3-131
93
To
Gb 2
Gb 2
dr =
log (R/ro)
27Tr(1 - '1')
47T(l - '1')
(3-52)
TO.?
Gb/27Tr
(3-53)
94
SOME PROPERTIES
OF
METALLIC LATTICES
[Chap. 3
energy determines the force between them. This can most easily be
demonstrated for two parallel screw dislocations; in this case the stress
fields have cylindrical symmetry and one expects the force between them
to depend only on their distance apart, i e., the force should be a central
force. To illustrate this, suppose a piece of material contains a screw
dislocation along the z-axis (Fig. 3-22) and let us produce a second screw
dislocation parallel to the first one, at a
distance r. As before, we produce a cut
extending from A to B in Fig. 3-22 and
displace the material on one side of the cut
relative to that on the other side over a distance b along the z-direction. Since at the
moment we are interested only in the inx teraction energy of the two dislocations, we
shall calculate only the work required to produce the second dislocation in so far as this
Fig. 3-22. Referring to the
calculation of the interaction work is determined by the presence of the
energy of two screw dis- first dislocation. Thus, if E; represents the inlocations running along the teraction energy per unit length of dislocation,
z-axis; one is located at the We may write
origin, the other in A. [After
Cottrell, op. cit., p. 50]
(3-54)
where Gb/21Tr is the shear stress produced along the z-direction in the cut by
the dislocation at the origin. The force between the two dislocations is
then
(3-55),
F(r) = -dEJdr = Gb 2/21Tr
Note that the force varies as l/r. For dislocations of opposite sign the
force is attractive; for equal signs the force is repulsive.
Similar considerations may be held for the interaction between edge
dislocations. In this case the force has a radial as well as a tangential
component. Thus for two-edge dislocations of equal sign, along the z-axis
and with a Burgers vector along the x-axis, one obtains 42
Gb 2
Eo = 21T(1 _ v)r sin 20
(3-56)
where 0 is the angle between r and the x-axis (see Fig. 3-23). Here again,
the radial force is repulsive or attractive depending on whether the
dislocations have equal or unequal signs. In the latter case the signs of
both Fr and Fo must be reversed in (3-56). We have mentioned earlier that
the motion of an edge dislocation is mainly confined to the slip plane
(conservative motion). For this reason, the force component along the
.. See Cottrell, op. cit., p. 47.
'l,'
Sec. 3-14)
95
(j -
Fo sin
(3-57)
(j
(3-58)
27T(l - v)r 4
0 and for. x
y.
y
parallel edge dislocations of the same sign
repel each other in the direction of the slip
plane; for x < y, or (j > 45, they attract
each other along the x-axis. The stable
configuration for the two dislocations
occurs when they lie vertically above each
other. This conclusion is also true when
a large number of edge dislocations of the
same sign are involved. In fact, such an
array of dislocations has been suggested
by Burgers as a model for a grain bound- Fig. 3-23. Radial and tangential
ary between two crystallites of different components of the force exerted
by an edge dislocation at the
orientation. 43
origin on an edge dislocation at
Dislocations also interact with a free A. Both dislocations lie along
surface. In fact, any dislocation will be the z-axis and have a Burgers
attracted by a free surface, since a motion vector along the x-axis. The
towards the surface would reduce the strain calculation in this case involves a
cut along AB.
energy. According to Koehler the force of
attraction is approximately given by an
image force, i.e., approximately equal to the force of attraction produced
by a dislocation of opposite sign located at the image position of the first
one relative to the surface. 44
While on the subject of stress fields around dislocations, we may
mention that impurities are in general attracted by edge dislocations. If
the impurity atoms are "larger" than those of the host lattice, they will
tend to move toward the region of tension, since in this way the tension will
be somewhat released in this region. On the other hand, if the impurity
atoms are "smaller" than the host atoms, they tend to be deposited in the
region of compression.
In the preceding section we mentioned that dislocations are not in
thermal equilibrium with the lattice and the question was raised as to why
.3 J. M. Burgers, Proc. Koninkl. Ned. Akad. Wefenschap., 42, 293 (1939); Proc. Phys.
Soc. (London), 52, 23 (1940). For experimental evidence, see next section.
H J. S. Koehler, Phys. Rev., 60, 397 (1941).
96
[Chap. 3
la)
Ib)
assume that the crystal in this state contains a number of edge dislocations
in a pattern such as the one illustrated in Fig. 3-24b. In order to calculate
the density of dislocations required to bend a certain specimen to a certain
radius of curvature, consider a single glide packet. When L is the length of
the outer arc and t is the thickness of the packet, the length of the inner arc
is evidenly L(l - tfR) where R is the radius of curvature. Suppose now
that the packet contains n positive edge dislocations; we must then have
'1b = tLfR, where b is the absolute value of the Burgers vector. Since the
iensity of dislocations in this case is simply given by the number of
F. C. Frank, Report of Pi((sbu~r:h Conference on Plastic Deformation of Crystals,
950, p. 100.
..
;
.: t
~,
Sec. 3-15]
97
dislocation lines piercing through a unit area of the plane of the paper,
we..obtain
(3-59)
p = niL! = llRb
For example, to bend a crystal to a radius of 3 cm one requires, with
b c::::: 3 X 10- 8 cm, a dislocation density pc::::: 107 cm- 2
2. Estimates from X-ray diffraction measurements. In Chapter 1 we
discussed. the conditions for X-ray diffraction from crystals and found that
reflection occurs only if the Bragg condition is fulfilled. Now, if a crystal
had perfect periodicity, the angular spread about the Bragg angle should be
not more than about 5 seconds. However, most crystals show an angular
spread of the order of several minutes. In order to explain discrepancies of
this kind, Darwin 46 and Ewald 47 introduced, many years ago, the notion of
a mosaic crystal, i.e., they assumed that an actual crystal is made up of a
number of small blocks which themselves are perfect but which are slightly
misoriented relative to each other. It is presently believed that this mosaic
structu. e may be the result of the three-dimensional network of dislocations mentioned in the preceding section. Assuming that this is the
case,48 an estimate of the dislocation density in terms of the observed. total
angular spread () of the X-ray pattern may be made in the following manner:
Suppose the surface area of the crystal involved in the X-ray measurements
has a side L. Let the density of dislocations in the crystal be p, so that p
dislocation lines pierce through a unit area of the surface of the crystal.
Let us now define the edge A of each block in the mosaic structure in such a
way that A2 = lip, i.e., we associate one block with each dislocation line
coming out of the surface. The average angular misfit between blocks is
then ex = blA radians; these misfits may be positive or negative. Thus as
we pass across the crystal over a distance L, we pass LI A blocks, and the
probable angle of misfit between the first and the last block is ex (LIA)1/2.
Identifying this angle with the total observed spread of 0 radians, we find
() = ex(LIA)1/2
or
{j =
bV/2p3/4
(3-60)
For a typic~l case of a pure crystal let us take L = 0.1 cm; () = 10-2
radian, and b = 3 X 10-8 cm. We then obtain p c::::: 108 cm- 2 . Note that in
this case A c::::: 10-4 em, in agreement with other estimates.
3. In heavily cold-worked metals the density of dislocations is
sufficiently high to produce an increase of a few per cent in the electrical
resistivity. According to calculations by Dexter, densities of dislocations
of the order of 1012 cm- 2 are required to explain measurements made on
cold-worked copper. 49 Similar estimates have been obtained from
.6
48
98
[Chap. 3
-.'.....
iii:
. . ...
.....
/O,M
99
(seconds)
Deale.
(cm)
17.52.5
65.0:': 2.5
85.0::": 2.5
(1.3
DubS.
10~'
10~'
10~'
I
I
(cm)
(5.3 0.3)
0.1)
(0.99 ::I- 0.2)
(1.3
X IO~'
X 10~'
X IO~'
conclude that the etch pit method is presently the most direct method for
determining the dislocation density.
3-16. The
~rank-Read
100
[Chap. 3
REFERENCES
c. S.
1952.
N. F. Mott and H. Jones, Theory of the Properties of Metals and Alloys,
Oxford, New York, 1936.
W. T. Read, Dislocations in Crystals, McGraw-Hill, New York, 1953.
Chap. 3)
101
PROBLEMS
3-1. When N is an integer ~ 1, one may approximate the expression
log (N!) = log 1 + log 2
+ ... + log N
N log N.
3-2. Suppose one has N boxes and n balls, where both Nand n are
I; the balls are all of the same color and indistinguishable.
Nk [1
+ log (kT/hv)]
for
hv <{ kT
This proves expression (3-2). Also find an expression for Stll for
temperatures which are not high compared with hv/k. Furthermore,
derive an expression for the free energy F of the system for both temperature
ranges.
3-4. Consider a system of N one-dimensional harmonic oscillators in
contact with a temperature bath T; all oscillators have the same frequency
v. Assume thai in equilibrium the oscillators have a total of n vibrational
quanta hv; consider n for the moment as a variable and find an expression
for the free energy F. From the equilibrium condition (3F/onh = 0,
derive Planck's formula for the average energy of an oscillator at a
temperature T. (Note; In the preceding problem one makes explicit use of
Planck's formula, in contrast with the present problem.)
3-5. Estimate the number of vacancies per atom in thermal equilibrium
for a crystal at T = 300 and T = 600, assuming that the energy required
to form a vacancy is lev.
102
(Chap. 3
3-8. Consider a low-temperature modification A and a high-temperature modification B of a certain element. I n order to discuss the equilibrium
between these modifications, assume for simplicity that the lattice vibrations
may be represented by Einstein models of frequencies v" and VO' Suppose
that the binding energies per atom in the two modifications are given by
- E " and -E/). Explain why E" > Eli' Set up an expression for the free
energy per atom in each of the modifications. Then write an expression for
the change in the free energy t::.F if one atom is transferred at constant T
from A to B. In equilibrium, t::.F = O. Show that there is only one temperature for which the two modifications are in equilibrium, viz., Tee( =
(Eb - Ea)/[k log (vu/vJ. What can one conclude about the ratio vb/Va, and
explain why the answer is reasonable.
3-9. Consider a b.c.c. lattice built up of atoms which may be assumed
to be hard spheres of radius R. Calculate the maximum radius of a hard
spherical atom that would fit in an interstitial position. Do the same for a
f.c.c. lattice (in this case there are two types of interstitial positions !).
How many interstitial positions are there in both types oflattices per normal
lattice site?
3-10. Consider a particle restricted to motion in one dimension.
Suppose the particle undergoes consecutive displacements with equal
probability to the left or to the right, the absolute magnitude of the
displacements being A. Show that the mean square displacement for N
steps is equal to NA2. Also show that the probability for the particle to be
found at a distance nA relative to the origin, after N steps, is given by
1)""
l(N
pen)~ = ( :2
N!
From this, show that for N'}:> 1 the probability to find the particle after a
time interval t at a distance x relative to the origin is given by
P(x,t) c::::
where
2T)1/2
exp (-X 2T/2A 2t)
( TTt
From this and the results obtained in the previous problem, show that the
diffusion coefficient in terms of the random walk problem is given by
D = }.2j2T. Also show that the mean square displacement for a time t
is given by (2Dt)1/ 2
Chap. 3]
103
..
"
J
'
_ ,
',t
,~:);:
'~
Chapter 4
\
\
104
Sec. 4-1]
105
4.068
4.064 '---'--'---L--'-_'_-'---L.--1~--'
o 20 40 60 80 100
- - Atomic per cent Ag
Here epAA' epnn, and epAB represent, respectively, the dissociation energy of
an AA, BB, and AB pair of nearest neighbors; z is the coordination
number, i.e., the number of nearest neighbors of a given atom. Whether
or not the two metals will have a wide or narrow solubility range depends
on the quantity ep defined by (4-1). When ep > 0, like atoms attract each
other more strongly than unlike atoms, and hence we expect a limited
solubility. On the other hand, when ep < 0, there is a preferential attraction
between unlike atoms, and the solubility may cover the whole range of
compositions, as in the Ag-Au system.
We shall consider the case ep > (limited solubility), inquiring about
the variation of the solubility with temperature. Suppose an alloy of the
substitutional type contains Na A atoms and Nb B atoms. For convenience
we shall introduce the atomic concentrations c = Nal Nand 1 - c = Nbl N,
N b We shall now express the free energy F = E - TS
where N = Na
of the alloy in terms of c; in order to find the equilibrium concentration
106
[Chap. 4
Ebinding
+E
vibr -
TSth
TS"f
(4-2)
The terms here are, respectively, the binding energy of the alloy relative
to the system of infinitely separated atoms, the vibrational energy of the
lattice, the thermal entropy term, and the configurational entropy term.
The binding energy of the alloy may be found from the following
arguments, assuming nearest neighbor interaction only: The total number
of pairs of nearest neighbors is Nz/2 (the factor i arises since otherwise
each pair is counted twice). If we assume a completely random distribution
of A and B atoms over the lattice (which is probably never exact in
practice), the probability for a pair chosen at random to,be
of the AA type is e2
of the BB type is (l - c)~
of the AB type is 2e(l - e)
The factor 2 in the last case enters because we count AB as well as BA
pairs; the sum of the probabiliti/ _equals unity, as it ought to. Employing
the dissociation energies introd'uced above, we may write
Eb
-tNZ[C24>AA
+ (l -
e)24>BB
+ 2c(~ -
C)4>AB]
-tNz[C4>AA
+ (1
(4-3)
The minus sign arises from the fact that the mutual potential energy of
two atoms equals minus the dissociation energy. In order to simplify
matters we shall assume that the thermal entropy is independent of c, i.e.,
we assume that when an A atom is substituted for a B atom, the vibrational
spectrum of the lattice does not change; this assumption does not impair
the general conclusions.
The configurational entropy is determined by the number of different
ways in which Na atoms of kind A and N b atoms of type B may be distributed
over Na + Nb lattice sites (vacancies are neglected). Hence, according to
the Boltzmann relation,
N!
-TScf = -kTlog N
IN I
a'
b'
+ (1 -
c) log (1 - c)]
(4-4)
0, one
Sec. 4-2]
107
CJ
e
(4-5)
For given values of <p, <P AA' and <PBB' this equation can be solved numerically
for c. If we consider the special case <PAA = <PilE we obtain
e/(l - c)
(4-6)
exp (-z<p/2kT),
.. - f - - :
(4-7)
In Fig. 4-2 we have plotted 2kT/z<p as function of c. The region above the
curNe corresponds to a homogeneous solid solution; the region below the
curve corresponds to temperatures that are
Homogeneous solution
too low to give a true solid solution. The
.5
symmetry of the curve is, of course, due to
our assumption <PAA = <PHIl' In practical
cases, a great part of the solubility curve
may lie above the melting point so that
Phase mixture
in the phase diagram only those parts will
enter that are close to either of the pure
metals. When the reader takes a look
at phase diagrams of binary alloys he
.5
o
1.0
will readily recognize the occurrence
_c
of the domelike shapes similar to that Fig. 4-2. The solubility curve
of Fig. 4-2. To give a numerical example, for a binary alloy, according to
suppose that the maximum solubility of
equation (4-6).
a certain metal in another is I per cent
at 300C. In that case, kT,-...J 0.05 ev and from (4-7) one finds
z<p ~ 0.46 ev. Treatment similar to that given here for substitl,ltional alloys
I
may be given for interstitial ones.
,
108
[Chap. 4
Similar sequences of phases with the same structures are found in many
alloys, but the compositions at which they occur may be quite different
from those given for the brass system above. It is obvious that the alloy
compositions corresponding to the various phases cannot be explained in
terms of the usual chemical valence rules. It was pointed out by HumeRothery, however, that the electron to atom ratio. is approximately the
same for a given phase of different alloys.3 A few examples are given in
Table 4-1 to illustrate this.
Table 4-1. Compositions and 'Electron-to-Atom Ratio for Structurally Analogous Phases
Electron-atom ratio 3:2
fJ structure (b.c.c.)
If
CuZn
CuBe
Cu.AI
Cu.Sn
AgCd
AgMg
Cu.Zns
Cu.AI.
Cu.Cd.
Au.Cd.
Ag.Cd.
CU31Si.
CuZn 3
CuCd.
Cu 3 Ge
AgZn 3
Ag.Sn
AuZn.
Since the phases have a certain range of compositions over which they
are stable, the chemical formulas and the electron-to-atom ratios given in
the table are approximate. However, there is a striking regularity when
these "compounds" are considered from this point of view. For the alloys
given in the table the electron-to-atom ratios are calculated on the basis of
the normal number of \ ..1ence electrons associated with the atoms involved.
In order to fit alloys containing transition metals such as Fe, Co, Ni into
this scheme, one must assume that these atoms contribute zero valence
electrons. For example, FeAI has the {J structure corresponding to an
electron-to-atom ratio 3 : 2.
An interpretation of the change in structure associated with an increase
in the electron-to-atom ratio has been given by Jones in terms of the band
theory of metals. 4 Essentially, the picture is the following: In the
expression for the total energy of an alldy there occurs a term associated
3 W. Hume-Rothery, J. Inst. Metals, 35, 295, 307 (1926); see also by the same author
Atomic Theory for Students of Metallurgy, Institute of Metals, London, 1946.
H. Jones, Proc. Roy. Soc. (London), AI44, 225 (1934); A147, 396 (1934). See also
N. F. Mott and H. Jones, Theory of the Properties of Metals and Alloys, Oxford, New
York, 1936; and C. Zener, Phys. Rev., 71, 846 (1~47); also references in footnote 3.
Sec. 4-3]
109
with the kinetic energy of the conduction electrons; since this is a positive
term, it is unfavorable for the cohesive energy. As one increases the
electron-to-atom ratio it may be advantageous for the lattice to change
its structure if this permits a reduction of the total energy of the system.
We may give here the results obtained by Jones from calculations of the
band structure for the electron-to-atom ratios for which a new phase
should appear in the alloys.
Phase ............................. {3
1.615
1.75
1.538
1.7
I \
I
I
I
I __.
.-f.--;;:""--- -.
1"- I
'-...J, '
f ','. ,
I
"
0---'<
,./
,.
OZn atoms
Cu atoms
.
I
0-+-
,./
OAu atoms
Cu atoms
./
Fig. 4-3.
random over the available lattice sites, there is a great deal of experimental
evidence which shows that this is frequently not the case. For example,
the structure of {3 brass (CuZn) at low temperatures approaches an ordered
structure in which corner points of a cubic unit cell are occupied by Zn
atoms and the center by Cu atoms (see Fig. 4-3a). Thus, in the completely
ordered state, brass may be vizualized as two interpenetrating simple
cubic lattices of Cu and Zn. As the temperature is raised the degree of
order decreases, as will be further discussed below; at a critical temperature Tc the degree of order drops rapidly. Another example of an ordered
110
[Chap. 4
.250
co
...
~ .225
.s'..,"
co
.a'"
'-'
.200
ci. .175
rf.l
.150
.125
440
f'ig. 4-4.
480
500'C
Sec. 4-4)
111
Y = (R -
(4-8)
= k(Nlog N -
R log R -
Wlog W)
(4-9)
W. L. Bragg and E. J. Williams. Proc. Roy. Soc. (London), A145, 699 (1934).
112
[Chap. 4
-k(L\R log R
+ L\Wlog W) =
\
2k log (R/W)
(4-10)
or
R/ W
. \.
(4-11)
Let us now inquire about the nature of the function 4>(R, W). Consider
a given right A atom; the probability for an arbitrarily chosen nearest
neighbor of this atom to be a B atom (right) is R/ N; the probability that
a nearest neighbor atom is an A atom is WIN. The potential energy of the
A atom in the field of its nearest neighbors is then
where we use the same symbols as in Sec. 4-2. Similarly, the potential
energy of a right B atom in the field of its nearest neighbors is
We leave it as Problem 4-6 for the reader to show that this model of
nearest neighbor interactions leads, for the energy required to interchange
the positions of the right A and B atoms, to the expression
4>(R,W)
4>o(R -
W)/N
4>0[1'
(4-12)
where 4>0 = z(24)AB - 4>AA - 4>BB) is a positive quantity since the dissociation energy 4>AB for an unlike pair is larger than that for a pair of similar
atoms. The physical meaning of 4>0 is that it represents the energy required
to produce 2 wrong atoms in the completely ordered lattice ([I' = 1). It
is observed that according to (4-12) the energy required to produce 2
wrong atoms decreases as the amount of order decreases. Qualitatively,
this can readily be understood; for example, if two atoms in the completely
disordered state ([I' = 0) are interchanged, the energy is, on the average,
zero because in the long-distance theory the distribution of A and B atoms
around a given A or B atom is then random. Actually the simple linear
relationship (4-12) between 4>(R, W) and [I' was introduced by Bragg and
Williams as an assumption; we see that this assumption is equivalent to
the model employed above in which the interaction between the atoms is
simplified to nearest neighbor interactions with constant 4> AA, 4> AB, and
CPBn values.
When (4-11) and (4-12) are substituted into (4-8) we obtain the
following implicit equation for the long-distance order parameter.
[I' =
tanh (4)o[l'/4kT)
(4-13)
Sec. 4-5J
113
(4-14)
_x
Fig. 4-5.
114
[Chap. 4
For fJ brass, Tc
740oK, so that according to (4-15) the quantity CPo
is approximately 0.25 ev in this case.
A few words may be said here about the "extra" specific heat associated
with the order-disorder transition. The energy required to increase R by
dR is, according to the definition of cp, equal to -CPo.9"d R/2. Making use
of (4-8) one may thus write
,-...J
dE = -(Ncpo/4).9" d.9" ;
(4-16)
The specific heat per atom associated with the order-disorder transformation if thus given by
!
llcv
= (l/N)(dE/dT) = -(CPo.9"/4)(d.9"/dT)
&:vlk
1.5
1.0
.6
o
-TITe
(4-17)
roo C v dT =
.10
kTe/2
(4-18)
Sec. 4-6]
115
(I -
(I
+ a)/(I
- a)
exp (cp/kT)
(4-19)
.00
c5
- __
<7
A;
OB;
~)
A or B
-T
vicinity of the critical point than for constant cp. As an example we give
in Fig. 4-9 the long-range and short-range order parameters Y' and a for
a superlattice of the AB3 structure. For the details of short-distance order
theory we refer the reader to the literature. We may mention here an
approximate theory developed by Cowley in which the order parameter is
expressed in terms of the coefficients of the Fourier series, which determines
the intensity of X-ray scattering. 10 This makes a direct comparison
between theory and experiment possible; the agreement is very good.
REFERENCES
Besides the references given at the end of Chapter 3, see also:
J. M. Cowley, Phys. Rev., 77, 669 (1950); J. Appl. Pltys., 21, 25 (1950).
116
[Chap. 4
',j
..
~'''
r "
Chapter 5 {
.i
117
118
[Chap. 5
= -
:2
(:1- ~~ :3 -:4
+
+ ~; - ... )
(5-1)
where e is the charge per ion. Note that because Coulomb forces decrease
relatively slowly with distance, it is not sufficient to consider only a few
shells of ions around the central ion.
Evidently, the coefficient of e2 /r is a pure number, determined only
by the crystal structure. Series of this type have been calculated
by Madelung,3 Ewald,4 and Evjen. 5 For the
_C./'F--+-'_-+/-;.-'--a/
NaCl structure the result is
f'/~I--1I~/'-i-I-,-1~/
I,
I
I,
,I
lo-_1-L-
!~:J_
/'
',,/1
..."/9-1-
77
~-r---r-~;--+-<1
NaCI
Fig. 5-1. The sodium
chloride structure.
Ec
-Ae2 /r with A
1.747558 ...
(5-2)
A
A
A
=
=
=
1.762670
1.6381
1.641
Sec. 5-2]
119
Focusing our attention again on one particular ion, we may thus write
for the repulsive energy of this ion due to the presence of all other ions,
(5-3)
where B is related to B' by a numerical factor. In view of the fact that
repulsive forces depend so strongly on the distance between the particles,
the repulsive energy (5-3) is mainly determined by the nearest neighbors
of the central ion. The total energy of one ion due to the presence of all
others is then obtained by adding (5-2) and (5-3):
(5-4)
Assuming that the two types of forces just discussed are the only ones
we have to take into account and neglecting surface effects, we thus find
for the total binding energy of a crystal containing N positive and N
negative ions,
E
E(r)=N(-A~+~)=N(r)
,
r
rn
(5-5)
(dE/dr)
r~ao =
'"
la
I
I
(5-6)
From the last two expressions one thus obtains the following relation
.,
"'".
between the two unknown parameters Band n:
B
(Ae 2 /n)ag- 1
(5-7)
120
[Chap. 5
V= cNr3
(ddV2E) a = 9cWag
1
[-4Ae
~+
2
KocNa~ =
n(n
+ 3)B]
a~H
(5-12)
+ 9cao 1Koe A
4
(5-13)
5.9
8.0
- 8.7
11 =
11 ~
11
NaCl
NaBr
II =
11 =
9.1
9.5
Sec. 5-3]
121
It will be obvious that the repulsive forces acting between two ions
will depend on the distribution of the electronic charges in the ions and
especially on the number of electrcns in the outer shells. For example,
we would expect n to be larger for NaCl than for LiCl, because the Na+
ion has eight outer electrons, the Li I- ion has only two. From an approximate treatment of the interaction between closed-shell electronic configurations, Pauling arrived at the following values of n as a function of
) , .,
the occupation of electronic shells. 9
Table 5-1. Repulsive Exponent as Function of Electron Configuration
Electron configuration'
Ion type
He
Ne
Ar (eu)
Kr (Ag)
Xe (Au)
..... .
2
2
8
8
8
8
. .....
..... .
..... .
..... .
8(18)
18
18
......
8(18)
18
......
. .....
. , . , ..
8(18)
5
7
9
10
12
This table should be used by taking the average value of n for the two
ion types occurring in the crystal. For NaCl, for example, one takes the
average of 7 and 9; for NaF the average of 7 and 7, etc. Note that this
table is in qualitative agreement with the experimental values of Slater
referred to above.
5-5. Calculated and experimental lattice energies
the
the
for
the
L. Pauling, Proc. Roy. Soc. (London), 114, 181 (1927); J. Am. Chern. Soc., 49, 765
(1927); Z. Krist., 67, 377 (1928).
122
[Chap. 5
Compound
a o in
Angstroms
"
in ev
calc.
LiF
NaF
KF
RbF
CsF
2.07
2.31
2.66
2.82
3.00
6.0
7.0
8.0
8.5
9.5
10.5
9.3
8.3
7.9
7.5
......
......
......
......
......
LiCI
NaCl
KCl
RbCI
CsCI
2.57
2.81
3.14
3.27
3.56
7.0
8.0
9.0
9.5
10.5
8.4
8.0
7.1
6.9
6.5
8.6
7.9
7.1
7.0
6.7
LiBr
NaBr
KBr
RbBr
CsBr
2.74
2.97
3.29
3.42
3.71
7.5
8.5
9.5
10.0
11.0
7.9
7.5
6.8
6.6
6.2
8.2
7.5
6.8
6.6
6.4
LiI
NaI
KI
Rbi
CsI
3.03
3.23
3.53
3.66
3.95
8.5
9.5
10.5
11.0
12.0
7.4
7.0
6.5
6.2
5.9
7.8
7.2
6.6
6.S
6.3
MgO
CaO
SrO
BaO
2.10
2.40
2.57
2.75
7.0
8.0
8.5
9.5
41.0
36.5
34.5
32.5
('L
('L
......
......
......
......
Sec. 5-5]
123
--+
NaCl solid
+ ECI + L
The quantities introduced all refer to the formation of one ion pair of
solid NaCl. Here SNa represents the sublimation energy of sodium per
atom. Sublimation energies in general can be determined experimentally
IJ'
CsCl
Fig. 5-3. The CsCI and the ZnS (sphalerite or zincblende) structures. The open circles in the ZnS structure are located at points
obtained by displacements of 1/4 along three cube edges of the
corresponding corner point. For one of the open circles we have
indicated how it is surrounded by four black dots occupying the
corner points of a regular tetrahedron, with the open circle at the
center.
124
(Chap. 5
where q refers again to the heat of formation per "molecule" NaCI formed.
Subtracting this equation from the one obtained above, we find for the
lattice energy per ion pair,
ELexp. = SNa + INa + iDcI, - ECI + q
(5-14)
For NaCl, all quantities on the right-hand side are known from experiments
and thus we are able to give an experimental value for E]. which may be
compared with the one calculated with the Born theory. For NaCl we
find, for example, from (5-14),
ELexp.
1.1
+ 5.1 + 1.2 -
3.8
+ 4.3 = 7.9 ev
whereas Born's theory yields 8.0 ev. The experimental values obtained in
this way are listed in Table 5-2, and we see that theory and experiment
agree within a few per cent, indicating that the relatively simple approach
is essentially correct.
For the fluorides and oxides, the electron affinities are not known
from experiment, and they are usually calculated by replacing EL expo by
ELcalc. in (5-14). We note that for oxygen the electron affinity is negative,
i.e., it requires energy to add 2 electrons to the atom. This is not surprising,
because after the first electron has been added, we have a negative 0- ion
and we would expect addition of a second electron to require appreciable
energy. An experimental determination of the affinity of a neutral oxygen
atom for the first electron added gave 2.2 ev according to Lozier. l l Now
the total electron affinity for the addition of 2 electrons is -7.3 ev when
calculated from the lattice energy of oxides in the manner indicated above.
Thus addition of the second electron requires about 9.5 ev. The usually
accepted values of the electron affinities are given in Table 5-3 together
with the dissociation energies of the diatomic molecules (in electron volts).
Table 5-3. Electron Affinities and Dissociation Energies
Electron affinity
F
CI
Br
1
0
S
Se
Dissociation energy
4.25 ev
4.0
3.8
3.45
-7.3
-3.5
-4.2
2.75 ev
2.50
2.01
1.58
1.52
2.75
2.50
Sec. 5-6]
125
structure (ZnS). The latter two are represented in Fig. 5-3. In the CsCI
structure each ion is surrounded by 8 nearest neighbors of opposite sign;
in NaCl by 6, and in zincblende by 4. One may thus ask why a certain
compound crystallizes in a particular structure.
The answer must obviously be sought in the fact that the energy should
be a minimum, and the problem is thus reduced to explaining why for a
given compound its natural structure has a lower energy than any other
structure. We shall see that some insight into this problem may be
obtained from considerations of the size of the ions.
For metals one defines the atomic radius as half the distance between
nearest neighbors, although it is recognized that the meaning of the size
of an atom is necessarily vague. For ionic crystals one could try a similar
approach, but one is immediately faced with the difficulty that these
compounds consist of at least 2 types of ions, so that the lattice constant
provides information only about the sum of two radii. A little consideration of the interionic distances as given in the preceding section shows
that to a fair approximation ionic radii are additive quantities. For
example, if one calculates the difference (r K + - r Na +) from the values
given in Table 5-2 for the halides of these metals, one finds from the
fluorides,
f
q
rK+ -
r x ,,+
aKF -
ax"",
0.35 Angstrom
'!
.' 4." 1
and from the chlorides, bromides, and iodides in the same manner 0.33 A,
0.32 A, and 0.30 A, respectively. We see that the difference is roughly
constant and that it has meaning to associate a rather definite radius
with each ion. It is also obvious that a table of ionic radii can be obtained
only if the radius of one ion is known. Goldschmidt in 1927 has tabulated
ionic radii based on a radius of the F- ion of 1.33 A, a value which he
decided upon on the basis of work by Wasastjerna on the relation between
polarizability and ion size. 12 Pauling, in the same year, independently
published ionic radii based on theoretical calculations of the radii of some
ions_I2 The two sets are not equal, which is not surprising because of the
inaccuracies involved. One commonly refers to the Goldschmidt and the
Pauling radius of a given ion. In Table 5-4 the Goldschmidt radii (G) do
not refer to the original set but include many recent X-ray diffraction
data, especially those of Zachariasen. 13 Contrary to the tables by Goldschmidt, the radius for 0 2- is 1.45 A rather than 1.35 A. The radii
according to Pauling are also given in Table 5-4.
Returning now to the question of stability, we would expect at first
sight that the CsCI structure should always be more stable than the other
'" V. M. Goldschmidt, Chefll. Berichte, 60, 1263 (1927); L. Pauling, J. Am. Chell/.
Soc., 49, 765 (1927).
13 W. H. Zachariasen, Acta Crystal/., 1,265 (1948); Ph),s. Rev., 73,1104 (1948);
Chefll. Ph),s., 16,254 (1948).
126
[Chap. 5
'I
'I
Ion
Ion
.'_
H
F
Cl
Br
I
1.54
1.33
1.81
1.96
2.19
2.08
1.36
1.81
1.95
2.16
0'
1.45
1.90
2.02
2.22
1.40
1.84
1.98
2.21
S'Se 2
Te'
Li'
Na"
K+
Rb'
Cs"
Cu+
AgT
Au+
TP
0.68
0:98
1.33
1.48
1.67
0.95
1.13
......
1.51
0.60
0.95
1.33
1.48
1.69
0.96
1.26
1.37
1.44
Mg2+
CaH
Sr 2 +
BaH
Zn2+
Cd 2+
Hg2+
Pb 2+
0.30
0.65
0.94
1.10
1.29
0.69
0.92
0.93
1.17
0.31
0.65
0.99
1.13
1.35
0.74
0.97
1.10
1.21
Mn 2+
Fe"
Co2+
Ni'+
Cu2+
0.80
0.76
0.70
0.68
0.92
0.80
0.75
0.72
0.69
BeH
1:1
......
B3+
AP'
Sea.
y3+
La 3+
Ga 3+
In3+
TI3+
0.2
0.45
0.68
0.90
1.04
0.60
0.81
0.91
0.20
0.50
0.81
0.93
1.15
0.62
0.81
0.95
Fe3+
Cr 3+
0.53
0.55
......
......
CH
Si H
TiH
ZrH
Ce H
Ge H
Sn4+
Pb H
0.15
0.38
0.60
0.77
0.87
0.54
0.71
0.81
0.15
0.41
0.68
0.80
1.01
0.53
0.71
0.84
Suppose the ion of radius 'I is the ion in the center of the cube. If we
now increase the radius '2 gradually, leaving 'I constant, we reach a value
of /"2 such that further increase makes it impossible for the central ion to
touch the ones at the corners. This critical value is clearly reached when
127
Sec. 5-6]
2'2
= ('1
+ '2)V2
, '-/'
or
'2
= 2.44'1
Again, if the ratio becomes larger than 2.44, positive and negative ions
cannot touch each other, leading to an increase of the energy and consequently to the formation of the more stable zincblende structure (Fig.
5-3). For this structure, positive and negative ions cannot touch each
other if '2> 4.55r1 . The stability limits as derived from the above
simplified billiard ball model for the ions are therefore
,.11'1 < 1.37
r2/'1 < 2.44
zincblende .. """ ....... 2.44 < r2/r, < 4.55
rough rule. In general, however, one may say that the esCI structure is
found in those compounds for which the ionic radii are nearly equal,
whereas the zincblende structure occurs only when the ratio of the radii is
about two or more. This may be illustrated by a few examples in Table 5-5.
Table 5-5. Ratio of Negative and Positive Ion Radii for Salts with the
Cesium Chloride and Zincblende Structure
, _Ir,
Zincblende
structure
r_lr
CsC!
1.1
CsBr
1.2
1.3
ZnS
ZnSe
BeS
BeSe
CuCl
CuBr
CuI
2.1
2,3
5.1
5.6
1.9
2.0
2,3
Cesium chloride
structure
Csl
'.
n~1
1.2
I";2':J,; ~(
1.3
L5
1)
"
128
[Chap. 5
(5-16)
Sec. 5-7]
129
ex.
= -
(:r qEdx
Jo
= _
(E qE': dE = -iex.E2
Jo
(5-18)
For a field strength varying with time, one would have for the average
energy,
E
(5-19)
-(rx/2)(E2)
-constant/Ii
')
(5-20)
The mutual energy of two atoms would then be given by the sum of two
terms of the type (5-20). From the classical point of view, therefore, these
forces are a consequence of the dipole-dipole interaction between the
atoms.
Actually, the energy corresponding to (5-16) is only part of the van
der Waals energy and there is an infinite series of rapidly converging
terms. The next one corresponds to dipole-quadrupole interaction and
varies as ,8.
For the alkali halides, the attractive energy corresponding to (5-16) is
of the order of a few per cent of the total lattice energy. For the silver
halides it is appreciably more; e.g., for AgBr it is about 14 per cent.
This is a consequence of the relatively high polarizability of the silver ion.
We should note that the van der Waals energy sometimes plays an
important role in the discussion of the stability of different lattice
structures. 19
The zero-point energy of the crystal is also a consequence of quantum
mechanics. The possible energy levels of a harmonic oscillator are given
by
E
= (n
+ l)hJl
(5-21)
130
[Chap. 5
(2
47TV 3
Ct
+ 3c,1 )
112
dll
(5-22)
where V is the volume of the crystal and C t and c! are, respectively, the
velocities of propagation of transverse and of longitudinal elastic waves
. Making use of the definition of the Debye frequency Y D , one may write
(5-23)
where N stands for the total number of atoms or ions in the crystal.
Hence, at absolute zero, the contribution of the zero-point energy is
.\
t Jo(Vn F(Y)hv dy =
(5-24)
sNhllD
Per ion pair this corresponds to 9hv D /4. With a Debye frequency of the
order of 1012_1013 sec-1 this gives about 0.1 ev. As a correction to the
lattice energy the zero point energy thus contributes about 1 per cent.
Note that this correction reduces the values given in Table 5-2, wherea~
the van der Waals correction raises them. In general, the van der Waals
correction is more important for heavy elements (large polarizabilities),
and the zero-point energy for light elements (high Debye frequency). As
an example, we give here the various contributions to the lattice energy
for the two extreme cases LiF and CsI (all energies in ev).
Coulomb ...............
Repulsive ...............
Dipole-dipole .........
Dipole-quadrupole ...
Zero-point ..............
LiF
-12.4
+ 1.9
- 0.17
- 0.03
+ 0.17
CsI
-6.4
+0.63
-0.48
-0.04
+0.3
"
j'
}{:t
f~',;t:.r{t$i
REFERENCES
M. Born, Atomtheorie des festen Zustandes, Teubner, Leipzig, 1923.
M. Born and M. Gappert-Mayer, Handbuch der Physik, Vol. 24/2,
Springer, Berlin, 1933, pp. 723-794.
1. A. A. Ketelaar, Chemical Constitution, Elsevier, New York, 1953.
N. F. Mott and R. W. Gurney, Electronic Processes in Ionic Crystals,
2d ed., Oxford, New York, 1948, Chap. 1.
F. Seitz, The Modern Theory of Solids, McGraw-Hill, New York, 1940,
Chap. 2.
Chap. 5]
131
PROBLEMS
5-1. Show that the Madelung constant for a one-dimensional array of
ions of alternating sign with a distance a between successive ions is equal
to 2 log 2.
5-2. Calculate the compressibilities at absolute zero from (5-13) for
LiF and BaO, assuming the values of n given in Table 5-1.
5-3. A molecule of the vapor of an alkali halide is presumably built
up of a positive and a negative ion. Assuming the quantities B' and n in
the repulsive energy to be the same as in the solid state, show that
g/s = 's/Arg
and
(rs/rg)n-l = 6/A
= (rte<X. l
+ 2re<x' <x'2)/(,s 1
4oc l 0(2)
with a similar expression for #2; r denotes the separation between the
132
[Chap. 5
nuclei. (See, for example, P. Debye, Polar Molecules, Dover, New York,
1945, p. 60.)
5-11. Discuss the binding energy and dipole moment of alkali halide
molecules on the basis of an ionic picture and compare the results with
experiment (see E. S. Rittner, J. Chern. Phys., 19, 1030, 1951).
5-12. Give a simplified discussion of van der Waals forces (see, for
example, S. Glasstone, Theoretical Chemistry, Van Nostrand, New York,
1944, p. 423).
"
~.,
. 1,
l',(
,,<
Chapter 6
DIELECTRIC AND OPTICAL PROPERTIES
OF INSULATORS
In the present chapter a brief survey will be given of the atomic
interpretation of the dielectric and optical properties of insulators. The
theory given here is essentially classical; for the quantum theory of
dielectrics we refer to the literature (see, for example, J. H. van Vleck,
Theory of Electric and Magnetic Susceptibilities, Oxford, New York, 1932).
This chapter is divided into two parts: in part A we shall essentially be
concerned with the static dielectric constant, in part B, the frequency
dependence of the dielectric properties, including optical absorption and
dielectric losses, will be discussed. It may be emphasized that only isotropic
substances, for which E, D, and P are parallel vectors, will be considered.
,. f
~.
47Tq
(6-1)
Evac . d
flh
evac =
Aq/4>vac = Q/4>vac
(6-3)
Suppose now that the space between the plates is filled with an insulating
substance, the charge on the plates being kept constant. It is then observed
that the new potential difference 4> is lower than 4>vac' and similarly, the
133
134
[Chap. 6
= D = EsE
(6-5)
In other words, the effective surface charge density on the plates is now
q' = E/47T rather than q = Evac/47T, and one may say that introducing the
dielectric is equivalent to reducing the surface charge
density by an amount
P = q -q' = (Evae/47T)(1 - liEs) = (E., - 1)E/47T
(6-6)
+
Fig. 6-1. Schematic
illustration of
charges induced at
the surface of a
dielectric.
+ 47TP = EsE
(6-7)
The link between the macroscopic quantity Es and the atomic theory of
the dielectric constant is provided by the relation (6-6). In fact, it will be
shown below that P may be expressed in terms of the properties of the
atoms and molecules composing the dielectric.
i
6-2. The static electronic and ionic polarizabilities of molecules
\"
Sec. 6-2]
135
For systems which, as a whole, are neutral, i.e., when Liei = 0, one can
show readily that M is independent of the origin chosen (see Problem
6-1); it is with such neutral systems that we shall concern ourselves. As
an example, the dipole moment corresponding to two charges +e and -e,
separated by a distance d, is ed.
In a free atom, the charge distribution is such that the dipole moment
in the absence of an external field vanishes; the center of gravity of the
electron distribution coincides with the nucleus. Consider now an atom
in a static homogeneous external field E. The force exerted on the positive
nucleus will then be oppositely directed to the forces exerted on the
electrons. As a result, the external field tends to draw the center of gravity
of the electrons away from the nucleus. On the other hand, the attractive
forces between the electrons and the nucleus tend to preserve a vanishing
---+E
/'
/
/
'.
"\ \
+e
\
J
Ze-
/
/
Fig. 6-3. Simplified model for estimating the magnitude of the electronic polariz.ability of an atom, as
described in the text.
136
[Chap. 6
ZeE = (Ze)2djr3
(6-10)
,,\z
(6-11)
For this simple model, therefore, the polarizability IXe is equal to r3.l Note
that IX. has the dimensions of a volume. For r::: 10-8 cm, we see that IXe
is of the order of 10- 24 cm 3 Hence, for an external field of 300 volts per
cm one finds d::: 10-15 cm, which shows that for most practical field
strengths the condition d
r is satisfied. It is for this reason that in (6-9)
one usually retains only the first term. For atoms with more than one
electron, similar considerations are valid, and with each atom or ion one
may associate a certain electronic polarizability IX It will be evident that
in general atoms with many electrons tend to have a larger polarizability
than those with few electrons. Electrons in the outer electronic shells will
contribute more to !Xc than do electrons in the inner shells, because the
former are not so strongly bound to the nucleus as the latter. Positive
ions therefore will have relatively small polarizabilities compared with
the corresponding neutral atoms; for negative ions the reverse is true.
We give a few examples in Table 6-1; more complete tables are available
elsewhere. 2
I
<
Table 6-1. Some Electronic Polarizabilities ill 10- 24 cm 3 The values for
the alkali and halide ions are those given by Bottcher; the others are due
to Pauling
IX,
He
Ne
Ar
Kr
Xe
0.20
0.39
1.62
2.46
3.99
01.
1)(,
Li+
Na+
K+
Rb+
Cs+
0.02
0.22
0.97
1.50
2.42
FCIBr1-
0.85
3.00
4.13
6.16
lIFn
~- -~--------
Sec. 6-2]
137
So far, we have considered only simple atoms and ions. For molecules
one is faced with two more possible influences of an external field:
I. Molecules may have permanent dipole moments which may be
aligned in an external field.
2. The distances between ions or atoms may be influenced by an
external field.
For exaJ'!1ple, a molecule such as HCI may in first approximation be
considered to consist of two ions; the permanent dipole moment is thus
equal to the effective charge per ion times the separation of the ions.
Symmetric molecules like H 2' CO 2 , CCI 4 , etc. evidently have no permanent
dipole moment. An external electric field will tend to orient permanent
dipoles along the field direction, and one speaks of orientational polarization. This contribution to the total polarizability of a molecule will be
discussed in Sec. 6-3.
In molecules as well as in atoms an external field will displace the
electrons with respect to the corresponding nuclei. Over and above this,
however, a displacement of atoms or ions within the molecule may be
caused by an external field. For example, in an HCI molecule an external
field will change the interionic distance to some extent, leading to a change
in the dipole moment. Similarly, in a molecule like CCl 4 (which has no
permanent dipole moment) a change in the bond angles between the CCI
groups will produce a dipole moment because each of these groups by
itself does have a dipole moment. This kind of induced polarization is
called atomic or ionic polarization because it is a consequence of the
displacement of atoms within the molecule. The induced electric dipole
moment resulting from elastic displacements of ions within the molecule
may again be represented by an expression of the type (6-9), by replacing
7.,. by the atomic polarizability OCa' It should be noted that OC n refers to
an average over all possible orientations of the molecule with respect to
the field. In Sec. 6-9 it will be shown that OC a may be considered a constant
up to frequencies in the infrared spectrum. For most molecules, oc" is of
the order of I 0 per cent of (Xc'
Summarizing, one may conclude that the electric properties of a
molecule may be characterized by the following three quantities:
(a)
(b)
(X"
(Xa' representing the polarizability due to atomic or ionic displacements within the molecule (changes in bond angles and interatomic
distances).
138
[Chap. 6
(6-12)
-lL' E
where 27T sin () dO is the solid angle between 0 and 0 d(). Hence the
average component of the dipole moment along the field direction is equal
to
"
f-l cos () sin 0 dO exp [(f-lE cos ()/kT]
f-l(cos (j)
D=O
:',-,r:"}i
"
(6-13)
and
(f-lE/kT)
a 'cf'"'W'
(6-14)
Sec. 6-3]
139
We then obtain
+a
f xex dx = e
(cos 0) = - f-+
1
ex dx
+ e-
e-
- -a = L(a)
(6-15)
-a
The function L(a) is called the Langevin function, since this formula was
first derived by Langevin in 1905 in connection with the theory of paramagnetism. 3 In Fig. 6-5,L(a) has been
L(a)
plotted as a function of a = p,E/kT.
)1/3)a
/
Note that for very large values of
a, i.e., for high field strengths, the 1.0
.8
function approaches the saturation .6
value unity. This situation would .4
correspond to complete alignment .2
of the dipoles in the field direction,
5
4
6
2
o
3
because then p(cos 0) = p.
_. a=/JoE/kT
As long as the field strength is
not too high and the temperature is Fig. 6-5. The Langevin function L(a).
not too low, the situation may be
For a
1, the slope is 1/3.
strongly simplified by making the
approximation a ~ 1 or pE ~ kT. Under these circumstances the
Langevin function L(a) = a/3, so that then
,"
<
pE ~ kT ,
(6-16)
See, for example, P. Debye, Polar Molecules, Dover, New York, 1945.
P. Debye, Phys. Z., 13,97 (1912).
,o'
ni
;1;".
140
[Chap. 6
E. Suppose the gas contains N molecules per unit volume; the properties
of the molecules will be characterized by an electronic polarizability OCe>
an atomic polarizability ()(a' and a permanent dipole moment fl. From the
discussions in the preceding two sections it follows. that, as a result of the
external field E, there will exist a resulting dipole moment per unit volume:
(6-17)
Note that only the permanent dipole moment gives a temperaturedependent contribution, because ()(e and ()(a are essentially independent of T.
If the gas fills the space between two capacitor plates of area A and
separation d, the total dipole moment between the plates will be equal to
M=PAd
This simple relation shows immediately that the same total dipole moment
would be obtained by assuming that the dielectric acquires an induced
surface charge density P at the boundaries facing the capacitor plates, as
discussed in Sec. 6-1. Hence the quantity P introduced here as the dipole
moment per unit volume is identical with the quantity P introduced in
Sec. 6-1, where it represented the induced surface charge density at the
dielectric-plate interface. Therefore, combination of (6-17) and (6-6) leads
immediately to the Debye formula for the static dielectric constant of a gas. 6
s -
1 = 47TP/E
47TN(iJ. e
+ iJ. + f-l2/3kT)
a
(6-18)
Sec. 6-4)
141
~CH2CI2
4
2
_ _ _ CHCI 3
CC14
142
[Chap. 6
e) =
(47T/3)P
(6-20)
D - 47TP
Es
or
+ 47TP/3 =
+2
E., = -3- E
for
+ 47TP/3
E4 = 0
(6-21)
E4 -
Ilk'"
3x~ - ,~
5
'k
(6-22)
Sec. 6-5]
143
symmetry of the problem, Ph = Pkz = 0, and Pka; will be the same for
all atoms. Furthermore, for the atoms inside a spherical region,
(6-23)
Obviously, (6-22) vanishes for a simple cubic lattice and (6-21) should
hold if the assumption of point like dipoles is accepted. We leave it up
to the reader to show that (6-22) also vanishes for b.c.c. and f.c.c. lattices
and for crystals such as NaCI. It must be emphasized that (6-21) does
not hold for all cubic crystals. For example, in barium titanate, which
has cubic symmetry, the oxygen ions are surrounded by TiH ions in such
a way that their contribution to (6-22) does not vanish.lO One must
therefore be careful in applying (6-21); one should start. from (6-22) in
order to evaluate 4 for the particular problem encountered. ll It will also
be evident from the above discussion that each type of atom in a given
crystal has, in general, its own internal field because the environment of
the different atoms is generally different. Thus the internal field at the
location of atoms of type 1,2, etc. may be written in the form
(6-24)
where the y's are the internal field constants. Only if 4 = 0 do we have
y = 41T/3. The internal field for tetragonal and simple hexagonal lattices
has been calculated by Mueller. 12
Even if the crystal symmetry is such that (6-21) applies, it does not
mean that the Lorentz field gives results in agreement with experiments.
This may be due to an overlapping of atoms as well as to the fact that
the dipolar fields produced by atoms which are only a few Angstroms
away are far from homogeneousP The latter makes it doubtful whether
one may employ the relation
(6-25)
}'
"
Pinuucctl = r:xi
to calculate the dipole moment induced in the central atom, as is done
in the theory outlined in the next section.
As a side line, it may be of interest to remark that the application of
(6-21) to polar liquid dielectrics has led to a great deal of confusion in
the literature. It was not until 1936 that Onsager realized that the internal
field cannot be used as the field which tends to orient the dipoles. 14 The
'I)
144
[Chap. 6
I)E/47T
(6-27)
(6-28)
where N represents the number of atoms per unit volume. In the particular
case for which the Lorentz expression for the internal field (6-21) is valid,
y = 47T/3. The resulting expression is then usually written in the form
of the Clausius-Mosotti formula, which may be obtained by substitution
of (6-24) into (6-27):
(6-29)
The main experimental test of the correctness of either (6-28) or (6-29) is
provided by measurements of the dielectric constant as function of the
number of atoms per unit volume. It has therefore been applied mainly
to gases. For solid elements one would have to vary the temperature in
order to vary N and the possible range of N values is of course very
limited. We do not know of any such measurements on, say, diamond or
other possible solids which may fall in this class of dielectrics.
Sec. 6-6]
145
Pe
+P
(Es -
Pe
(6-30)
1)E/47T
EO,
defined by
(6-31)
1)E/47T
(EO -
LiF
LiCI
LiBr
Lir
NaF
NaCI
NaBr
Nal
9.27
11.05
12.1
11.03
6.0
5.62
5.99
6.60
0 =
n2
1.92
2.75
3.16
3.80
1.74
2.25
2.62
2.91
KF
KCI
KBr
KI
RbF
RbCI
RbBr
Rbi
'. ".
6.05
4.68
4.78
4.94
5.91
5.0
5.0
5.0
Eo =
/I?
1.85
2.13
2.33
2.69
1.93
2.19
2.33
2.63
I
Hence P a is about two or three times P e in these compounds. In nonionic compounds, on the other hand, Pais usually a relatively small
fraction of P e
S. Whitehead and W. Hackett, Proc. Phys. Soc. (London), 51, 173 (1939).
For a number of other ionic solids, see for example N. F. Mott and R. W.
Gurney, op. cit., p. 12.
15
16
146
[Chap. 6
Let us now investigate if a simple theory can account for the observed
difference between the static and high-frequency dielectric constants.
Suppose the positive and negative ions acquire induced dipole moments
of, respectively, p+ and p_, under influence of a static field E. Furthermore,
suppose the positive ion lattice is displaced over a distance x relative to
the negative ion lattice. The atomic polarization may then be represented
by a point dipole ex at the location of each positive ion, because x is very
small compared with the lattice constant. The total electric dipole moment
per unit volume is then
P = N(p+
+ p_ + ex) =
(6-32)
(Es - I)Ej41T
where N represents the number of ion pairs per unit volume. For the
moment let us assume that the internal field at the location of a positive
ion is the same as that at a negative ion site and let it be represented by
E;. We may then write
: J\
x ; , ;J
Xu-33)
eEi
(6-34)
= Ix or x = eEilf
}'
(6-35)
If it were assumed that E; is given by the Lorentz expression (6-21), the
last expression could be rewritten as
(6-36)
This expression is analogous to the Clausius-Mosotti equation (6-29),
with the additional term e 2lfassociated with the elastic ionic displacements.
To investigate whether or not (6-36) describes the alkali halides satisfactorily, it is convenient to set up an equation relating Es and the highfrequency dielectric constant EO' determined by (6-31). Thus, if there were
only electronic polarization, which is the case if one measures the index
of refraction for visible light (EO = n 2 ), and if again the validity of the
Lorentz expression were assumed, one would have
Pe
EO -
~ E
N(cx e+
EO
+2 E
+ cx e_) - 3 -
(6-37)
Sec. 6-6]
147
Substitution of the factor N(r:t.e+ + O(e-) from this equation into (6-36)
yields the following relation between lOs and EO:
(6-38)
EO
47TNe 21f
1 -4:--7T--:N-'e=-:2--:j3=-1j
-C-
(6-39)
gives much better agreement with the experiments. This formula was first
derived by Hojendahl,I7 who introduced rather special assumptions about
the internal field at the location of positive and negative ion sites. From
the above discussion it is evident that the theory of the static dielectric
constant of simple crystals such as the alkali halides is not in a completely
satisfactory state, mainly because of the difficulties involved in calculating
quantitatively the internal field.
It may be noted here that the force constant f and the masses of the
positive and negative ions determine the infrared frequency associated
with the lattice vibrations. It is therefore possible to express the difference
(lOs __'_ EO) in terms of the infrared absorption frequency of the lattice. IS
A discussion of recent work on this topic may be found in H. Frohlich,
Theory of Dielectrics, Oxford, New York, 1949, Sec. 18.
(iii) In substances composed of molecules which bear permanent
electric dipole moments, the total polarization is made up of three
contributions,
(6-40)
where P d corresponds to the dipolar contribution. There exists no general
quantitative theory for dipolar solids because first of all the same
difficulties arise in evaluating the internal fields as in class (ii), and furthermore, the dipoles in such solids may not be able to rotate at all or only
to some extent. The discussion must therefore be limited to some
qualitative remarks. As an example of a dipolar solid which behaves in a
relatively simple manner, we show in Fig. 6-8 the dielectric constant
measured as function of temperature for CsH5N02 (nitrobenzene).19 It is
observed that at the melting point there is a large increase in the dielectric
constant. This is interpreted as an indication that in the solid the dipoles
K. HojendahI, Kgl. Danske Videnskab. Selskab, 16, No.2 (1938).
,. B. Szigeti, Trans. Faraday Soc., 45, 155 (1949); Proc. Roy. Soc. (London), A204,
51 (1950).
19 C. P. Smyth and C. S. Hitchcock, J. Am. Chern. Soc., 55, 1830 (1933).
17
148
[Chap. 6
16
fs
12
10
<-,'
0
80
290
300
-T('K}
120
160
;(r~'ft
200
_T('K)
has been plotted for H 2S.20 The melting point of H 2S is 187.7K. In this
case, the dipoles are apparently "frozen in" at temperatures below !03.5 D K;
at this temperature the structure changes in such a manner that the dipolar
groups become mobile; as the temperature is further increased, the
dielectric constant decreases as a result of increased thermal motion.
The other changes evidently affect essentially the density of the material,
i.e., N is reduced at these transition points.
for example, if
E = Eo cos wI
2.
(6-41)
C. P. Smyth and C. S. Hitchcock, J. Am. Chel11. Soc., 55, 1296 (1933); 56, 1084
(1934).
Sec. 6-7]
149
we have
D = Do cos (0)/
b)
DI cos
('It
D2 sin
0)/
(6-42)
,,"(I)
(6-44)
,J
(6-48)
where use has been made of (6-1) and (6-42); q is the surface charge
density on the capacitor plates. The energy dissipated per second in the
dielectric per unit volume is
I
W 2fT!,"
W=JEd!
277 .
(6-49)
By substitution of (6-48) and (6-41) into (6-49) one readily finds that the
integral containing D1 vanishes and one is left with
(6-50)
The energy losses are thus proportional to sin 0; for this reason sin 0 is
called the loss factor and 0 is the loss angle. 21
21 One frequently calls tan (j the loss factor; this is correct only for small values of <5,
because then tan D ~ sin /)
D.
150
[Chap. 6
(6-54)
Now, our final goal is to express the real and imaginary parts of the
dielectric constant in terms of the frequency wand the relaxation time T.
For this purpose we shall define the "instantaneous" dielectric constant
E,,, by
P
e
+ Pa =
_e_a_ _
417
(6-55)
"" For a proof that this procedure is correct, see, for example, M. Gevers, Philips
Research Rep/s., 1, 279, 298 (1946); M. Gevers and F. K. Du Pre, Trans. Faraday Soc.,
42.-\, 47 (l946).
Sec. 6-8)
151
Es
dPd=~(E8-E'IlE
.
\:
dt
41T
iwt_
oe
p )
d
(6-57)
+ +
D(t)
E*E(t)
E(t)
+ 41TP(t)
(6-59)
where E* is the complex dielectric constant. From the last two equations
and from the definition E* = E' - iE" the following expressions result:
(6-60)
(6-61)
These equations are frequently referred to as the Debye equations. In
Fig. 6-10 the quantities E' and E" are represented as functions of WT. It is
observed that the dielectric loss, which is proportional to E" according to
(6-50), exhibits a maximum for (r)T = I, i.e., for an angular frequency
equal to liT. Also, for frequencies appreciably less than l/T, the real part
of the dielectric constant E' becomes equal to the static dielectric constant.
In this frequency range, therefore, the losses vanish and the dipoles
contribute their full share to the polarization. On the other hand, for
frequencies larger than liT, the dipoles are no longer able to follow the
field variations and the- dielectric constant E' approaches E.".
The question may now be raised as to which physical models actually
satisfy the above phenomenological theory. We shall discuss here a
particular one as an example, viz., the case in which certain positive ions
in a solid may have two equilibrium positions separated by a distance 2a.
For simplicity it will be assumed that the line joining the two positions
A and B is parallel to the external field direction, as indicated in Fig. 6-11.
[Chap. 6
152
L--- E
....
__ 2a--
-logw
energy barrier separating the two types of sites is cp, the probability per
second for an ion in an A site to jump into a B site is, according to
statistical mechanics, of the form
Po =
'P
(6-62)
exp (-cp/kT)
where 'P is a frequency factor of the order of 1012 per second. Thus, without
external field, the ions will continuously change over from A and B sites,
but on the average there are per unit volume NI2 in A sites and Nf2 in
B sites if N is the total number of such ions per unit volume.
Suppose now that suddenly a static field E is applied in the direction
as indicated in Fig. 6-11. Particles in A sites will then see a potential
barrier (cp - eaE) and particles in B sites see a barrier (cp
eaE); hence
the ions will prefer B sites over A sites. In equilibrium we must evidently
have just as many particles making transitions A -+ Bas B -+ A, so that
(6-63)
or
where Naao and Nboo represent the equilibrium values for the number of
particles in A and B sites with an external field E. Let us now consider
the transient phenomenon as we go from the initial state NaO = Nbo = NI2
to the final state Naoo and N boc ' Particularly, we are interested in the time
dependence of (Nb - N a ) because the dipole moment per unit volume
resulting from this effect is
(6-64)
Sec. 6-8]
153
dNa/dt = -NaPab
dNb/dt
+ NbPba
(6-65)
NaPab - NbPba
+ Nb =
N, one may
Now, one may assume for not too high field strengths that eaE <{ kT,
so that
+ eaE/kT)
and similarly,
Pba
c:::'. Po(1
_---t'
}rt!
.,
- eaE/kT)
(dfdt)(Nb - N a)
-T
-2Po(Nb -- N a)
2poNeaE/kT (6-67)
NeaE
2kT
N - N = - - (1 - e
b
t
2Po)
(6-68)
and the polarization due to this mechanism is, according to (6-64) and
(6-68), given by
'i.n~:h
Ne 2a2 E
1
Ptl(t) = - k - (1 - e- t /
with T = (6-69)
2 T
2po
T
'i
,,' \.
Note that this equation has the same form as (6-52), which was the basis
on which the Debye equations were derived. The relaxation time is thus
equal to the reciprocal of the jumping probability per unit time in the
absence of an external field. Note also, that for this type of mechanism
the relaxatioq time decreases with increasing temperature and so does the
saturation polarization. It is of interest to observe that if the quantities
t:' and t:" are measured at a constant frequency but at different temperatures,
the curves as indicated in Fig. 6-12 may be expected to result.
For other possible models which lead to the Debye equations, see
Frohlich, op. cit., Sec. 11. It should be pointed out that the interpretation
of experimental results on dielectric losses frequently requires a distribution of relaxation times rather than a single one as assumed in the above
154
[Chap. 6
\
E"
(E
ea
foo F(T)wTdr
0
+ w2r
(6-70)
where F(T) is the distribution function of the relaxation times, such that
fooo F(T) dT = 1
(6-71)
We first of all note that in a static field, i.e., for W = 0, this reduces simply
to
(6-74)
x = eEo/mw~ or as = ex/Eo = e2jmw5 for w = 0
23 For a review, see M. Gevers, Philips Research Repts., 1,279,298 (1946); see also
M. Gevers and F. K. Du Pre, Trans. Faraday Soc., 42A, 47 (1946).
2' For dielectric losses in alkali halides resulting from Schottky defects and divalent
impurities, see R. G. Breckenridge, J. Chern. Phys., 16, 959 (1948); 18, 913 (1950);
also his article in W. Shockley (ed.), Imperfections in Nearly Perleet Crystals, Wiley,
New York, 1952 .
.. A proof that this leads to a term proportional to dx/dt may be found in R. Becker,
Theorie der Elektrizitiit, 6th ed., Teubner, Leipzig, 1933.
Sec. 6-9]
155
E*-I.
. .
47T
E~
Ii
47TN - . )
m (w6 - (
EO
)2 _. 0)2
e2
E~
47T N
e2
- . -2C;---O-::---~'
m (1)0 - (1)2)2 + y 2w 2
(6-77)
Y(l)
2
)2
+y
(6-78)
2w 2
E;;
156
[Chap. 6
<)
p*
6 -
417
1E
eiwt =
NIJ,.*
e
(6 +
3
which leads to
eiwt
0
(6-79)
,.
or
Substituting
2) E
47TN
1/1J,.: - 417N/3
= 1 + -----
(6-80)
IJ,.:
+ 417N -me . w~
---:;-----:------,,-,-,_ w 2 + iyw - 417Ne 2 /3m
(6-81)
Comparing this with (6-76) for a gas, we see that by defining a new
frequency
(6-82)
the same behavior is obtained as above; in the formulas obtained for a
gas, one only has to replace w~ by
i.e., the absorption frequency is
displaced.
In optical work it is usual to introduce instead of the quantity 6 the
complex index of refraction. A few remarks in this connection may therefore be in order. It is well known that Maxwell's equations for a nonmagnetic insulator give for the velocity of propagation of light the expression v = ely'; On the other hand, the index of refraction is defined as
n = e/v. This leads to the Maxwell relation = n 2 Now, when there is
absorption, the electric component of a light wave polarized in the ydirection and propagated in the x-direction may be represented by
wi,
(6-83)
Aeiw(t-n'x/c)
(6-84)
(6-85)
157
Sec. 6-9]
(n 2 -
k 2)
and the formulas (6-77) and (6-78) are thus also valid for
and
2nk, respectively. Note that the absorption per unit volume is proportional
to nk.
-.
.
~. . i'i.:~~ ~ _;_.~,'
: ,"
P (real part)
I,
J -_
Pe
Micro-
--
Infrared
waves
~>;"
~>,/
"_
The above considerations may be applied equally well to ionic displacements. To summarize the frequency-dependence of the polarization
we have represented, fn Fig. 6-14, pew) for a dipOlar substance with a
single atomic and electronic absorption line.
,
;
REFERENCES
R. Becker, Theorie der Elektrizitiit, 6th ed., especially volume II, Teubner,
Leipzig, 1933.
C. J. F. Bottcher, Theory of Electric Polarization, Elsevier, Amsterdam,
1952.
W. Fuller Brown jr., Encyclopedia of PhYSiCS, 27, Springer, Berlin, 1956,
pp. 1-154.
P. Debye, Polar Molecules, Dover, New York, 1945.
Dielectrics discussion, Trans. Faraday Soc., 42A, (1946).
H. Frohlich, Theory of Dielectrics, Oxford, New York, 1949.
R. J. W. LeFevre, Dipole Moments, 2d ed., Methuen, London, 1948.
O. Fuchs and K. L. Wolf, "Dielektrische Polarisation," Hand- und
lahrbuch der chemischen Physik, Vol. 6, Leipzig, 1935.
E. J. Murphy and S. O. Morgan, "Dielectric Properties of Insulating
Materials," Bell System Tech. I., 16, 493 (1937); 17, 640 (1938);
18, 502 (1939).
J :'H)~~;;~ Lil"J~t. ')d; Lcd;
,1
t,
158
[Chap. 6
I'
PROBLEMS
6-1. Consider a system of positive and negative charges, the system
being neutral as a whole. Show that the dipole moment of the system as
defined by (6-8) is independent of the location of the origin of the coordinate system.
6-2. Show that the potential energy of a dipole fJ. in an external field
E may be written -fJ. E. Also show that if IX is the polarizability of an
atom, the energy of the atom in an electric field E is given by -(1X/2)2.
6-3. From the electronic polarizabilities for the alkali and halide ions
given in Table 6-1 and from the lattice constants for the alkali halides as
obtained from X-ray diffraction data, calculate the high-frequency
dielectric constant for some of these salts on the assumption that the
internal field is given by the Lorentz expression; compare the results with
the experimental values given in Table 6-2.
6-4. Calculate the field strength required to reach 0.1 per cent of the
saturation value of the orientational polarization of a dipolar gas at room
temperature if the dipoles have a strength of 1 Debye unit.
6-5. Consider a system of noninteracting dipoles which are confined
to two possible orientations relative to an applied field E: either parallel
or antiparallel. Show that at a temperature T the average dipole moment
along the field direction is equal to ft2/kT (which differs by a factor 3 from
formula (6-16).
6-6. (a) A sphere of dielectric constant Ei and radius R is brought in a
homogeneous field E; the sphere is surrounded by vacuum. Show that
the field inside the sphere is homogeneous and given by 3E/(E i + 2).
(b) A substance of dielectric constant EO contains a spherical cavity of
radius R. If the field at large distances from the sphere is homogeneous
and equal to E, show that the field inside the cavity is homogeneous and
Chap. 6]
159
equal to 3EoE/(2Eo + I). (This field is called the cavity field; note that it
is independent of R.)
(c) Consider a homogeneously polarized sphere of radius R in vacuum;
there is no applied field. If P is the polarization of the sphere show that the
field inside the sphere (the self-field) is given by Es = -(47T/3)P = -M/R3,
where M is the total dipole moment of the sphere .
(E-l) 2
f= ~
R3
V'2V
V = -(A/r2
+ Br) cos 0
'<\..,
-,.. Chapter 7
l "\
160
Sec. 7-1]
161
Thus even if the energy 1>+ required to produce a single positive ion
vacancy were appreciably different from the energy 1>- to produce a
single negative ion vacancy, they would occur in approximately equal
numbers in the interior of the crystal. It is obvious from this that their
number will be determined only by the sum of the formation energies
(7-1)
As in Chapter 3, it will again be assumed that the external pressure may be
taken as zero, so that the equilibrium condition requires the free energy
E - TS to be a minimum. The free energy of the fictitious perfect crystal
will be represented by
" _.J_
(7-2)
\
<
F"
Fp
+ n1> -
T(S" -
+ n)!/ N!n!]
(7-4)
where Sa is the thermal entropy of the actual crystal. Let us define the
increase in thermal entropy LlSth resulting from the production of a
positive plus a negative ion vacancy by
(7-5)
Note that the essential factor is the Boltzmann factor containing 1>/2, i.e.,
half the energy required to produce a positive plus a negative ion vacancy.
The exponential term containing the change in thermal entropy per vacancy
LlStl J2 may be calculated on the basis of a particular model, for example,
an Einstein model. In that case the calculation is similar to that given in
Sec. 3-3 for metals. Thus let us assume that the Einstein frequency
associated with the ions in the perfect lattice is 1'.4 In the actual lattice,
It would be more realistic to introduce two Einstein frequencies, one for the positive
and one for the negative ions; the essential conclusions, however, would remain the
same.
"',
1(";;'.
162
[Chap. 7
+ (6N -
+ 6Nk
+ 6Nk
hv. Hence
Sa
Sv
(7-7)
According to (7-7) and (7-5), we may then write for the increase in thermal
entropy per vacancy formed,
b..Sth /2
(7-8)
For this model, the expression for the density of vacancies may then be
be obtained by substituting (7-8) into (7-6), giving
n/N = C exp (-4>/2kT)
with
(v/v')3Z
(7-9)
Note that the frequency ratio is larger than unity, so that the thermal
entropy changes favor the formation of vacancies.
Here, as in the case of metals, one frequently finds the argument in the
literature that if 4> depends on temperature in accordance with a relation
of the type
(7-10)
4>(T) = 4>0 + T (d4>/dT) = 4>0 - yT
the actual expression for the density of vacancies should be
-
(7-11)
or
(7-13)
Sec. 7-1]
163
(7-14)
Hence, one actually measures 4> in this manner. Also, the pre-exponential
factor A as determined experimentally is always equal to exp (6.s/2k).
Approximate methods to calculate the energy 4> required to create a
positive plus a negative ion vacancy will be discussed in Sec. 7-3 and we
shall therefore postpone giving numerical values for the quantities
involved.
+
y+
It will be evident that a positive and a
+
+
negative ion vacancy will attract each other as +
a result of the Coulomb field between them.
+
+
For large distances, the energy of attraction is
+
equal to -e2 /a, where If is the dielectric con- +
stant of the medium. They may therefore com+
+'
bine to form pairs of vacancies (Fig. 7-1). At
+
+
a given temperature there will exist a certain +
tatio between the number of single vacancies _ I c I +
+
and the number of pairs, the ratio depending
on the dissociation energy required to separate +
+
+
a pair into two singlets. There are evidently Fig. 7-1. The sequence
certain degrees of dissociation depending on of jumps I, 2, 3 may lead
whether the distance between the single vacancies to the formation of a
is small or large; in a sense one may therefore positive ion vacancy A;
speak of a thermally excited state of a pair if tile B represents a negative ion
vacancy; C an associated
distance between the vacancies is only a few pair of vacancies formed
atomic diameters. Reference to the importance as a result of Coulomb
of pairs of vacancies for diffusion in ionic
attraction.
crystal will be made later.
Interstitial ions in combination with vacancies may also occur: for
example, a positive ion may jump into an interstitial position, leaving a
vacancy behind. If the vacancy and interstitial ion are far enough apart to
prevent an immediate recombination, one speaks of Frenkel defects.
In this case it is not necessary to have equal numbers of positive and
negative Frenkel defects, because their formation does not require the
setting up of space charges over macroscopic distances. In general,
depending essentially on the energy required to form them, either the
positive or negative Frenkel defects will predominate. Also, they may occur
in combination with Schottky defects. The calculation of their density as
function of the energy 4> required to produce a Frenkel defect is essentially
the same as that given above for Schottky defects, i.e., one finds an
expression for the free energy of a crystal containing n defects and
165
Sec. 72)
(7-18)
all vectors having radial direction. Hence the dipole moment induced in a
volume clement dT located at a distance r from the center is equal to
(7 -19)
This dipole moment produces a potential at the center of the sphere of
P dT/r 2 , and thus the reaction potential in the center is
I(
DO
. I'=R
P
- dT
r2
cf)
- e ( 1 - -1)' . -I '
R 47Tr2
r2
47Tr2 dr
= -e ( 1 - -1) (7-20)
R
Thus we conclude from (7-17) and (7-20) that the energy of the ion in the
reaction field is
~2 eV= ~
(1 -~)
2R
(7-21)
A e
ao
(7-22)
1. '
(7-23)
166
[Chap. 7
energy of the ions, plays a major role in the theory of solubility of ionic
solids. The simple electrostatic problem mentioned above will also enter
'
in the discussion of the next section.
7-3. The activation energy for the formation of defects in ionic crystals
In Sec. 7-1 we derived an expression for the number of vacancies in an
ionic crystal in thermal equilibrium at a temperature T. According to
(7-6) or (7-9) this number is essentially determined by the formation energy
1> = cp+ cp_. Let us first consider the energy 1>+ involved in the formation
-+
+
t
+
+-_
"0
;t
+
~+
--+
+
1)
EL =
e ( 1 -Aao
n
(7-24)
"-
------
if we use the simple Born lattice theory. Putting the ion from infinity on
the surface of the crystal leads to a gain in energy of
~2 EL = ~A
~.
2 a
o
(1 -~)n
(7-25)
Thus if nothing else would happen, cp+ would be equal to the difference of
(7-24) and (7-25). However, the removal of a positive ion will affect the
neighboring ions in such a way that an adjustment takes place by which
energy is gained, thereby lowering cp+. Referring to Fig. 7-2, we note that
from the point of view of the surroundings of a positive ion vacancy, it looks
167
egative charge has been added at this lattice site. In other words,
an excess of negative charge in the vicinity of the missing positive
Consequently the surrounding material will become polarized.
polarization consists first of the formation of dipoles induced in the
ions by the Coulomb field of the missing ion, second of a slight ionic
displacement as indicated in Fig. 7-2. Because of the long range of
Coulomb forces, it is not sufficient to take into account only nearest
neighbors; the effect will spread over distances many times the lattice
constant. The calculation of this polarization energy P+ is very complicated
although it may be understood in principle on the basis of a simplified
model, first introduced by Jost. 8 If we consider the vacancy as a spherical
hole inside a homogeneous dielectric constant the hole bearing a charge e
at its center, we obtain the situation given in Fig. 7-3. The charge e, due
to the missing ion, polarizes the dielectric and thus in turn will create a
reaction potential Vat the location of the charge. We see that this problem
is identical with the one treated in the preceding section. Thus the polarization energy is given by
IS
p+
= ~ eV = ~. ~(I ~. ~.)
2
2 R+
From (7-24), (7-25), and (7-26), one obtains
(7-26)
(7-27)
For negative ion vacancies the same reasoning applies, so that the energy
required to produce a positive and a negative ion vacancy is equal to
cp
=cp+ + cp_ =
ao
(I -~)n _~e2
(1 _ ~) (_1 +.!_ )=
2
R+
R_
L-
P+ _ P_
(7-28)
The first term, of course, can be calculated with good accuracy, but as far
as the remainder is concerned the problem arises as to what values one
should assign to R+ and R_ in this idealized model. These values can be
found with good approximation only by comparison with more accurate
calculations of P+ and P_ based on an actual ionic picture rather than on a
continuum. Calculations of this kind have been made by Mott and
Littleton 9 and more recently to a higher approximation by Rittner,
Hutner, and Du Pre.lO As an example of the results of the former authors,
we cite those for NaCl and KCI below (all energies in electron volts).
*' L
NaCL........
KCI..........
7.94
7.18
P.,
3.32
2.71
P2.76
2.39
cp
1.86
2.08
R ,.
0.58a o
0.61a o
R
0.95a.
0.8Sa o
8 W. lost, J. Chem. Phys., 1,466 (1933); Trans. Faraday Soc., 3", 860 (1938); W.
Jost and G. Nehlep, Z. Physik. Chem., 832, I (1936); 834, 348 (1938).
N. F. Mott and M. J. Littleton, Trans. Faraday Soc., 34, 485 (1938).
10 E. S. Rittner, R. A. Hutner, and F. K. Du Pre, J. Chem. Phys., 17, 198 (1949).
[Chap. 7
]68
')f ionic
The values of R+ and R_ are obtained by substituting the valut enter
and P_ given by Mott and Littleton into (7-26). We note that the nun.
vacancies depends on 4>0/2, according to (7-6). Also, we see how impor.
the polarization energy is for the formation of vacancies in ionic crystals.
In fact, for NaCl it reduces the value of 4> for NaCI from 7.94 to only
1.86 ev. Because of the relatively small value of 4>, vacancies in alkali
halides close to the melting point may occur in concentrations of the order
of 10-4 per ion.
From the results quoted above it follows that for alkali halides
Rc :::::: 0.6a o and R_ :::::: 0.9a o. We shall see in Sec. 7-5 that the calculated
values of 1> are in fairly good agreement with experiments.
Similar estimates have been made of the activation energies required
for the formation of Frenkel defects in ionic crystals. It turns out that in
alkali halides, defects of the vacancy type are much more likely to occur
than interstitial ions. However, for silver halides there is theoretical and
experimental evidence for the occurrence of Frenkel defectsY
We have seen above that the "effective" charge of a positive ion
vacancy is negative, of a negative ion vacancy is positive. Thus there will
be attraction between vacancies of opposite sign and one may expect them
to form pairs. The binding energy of a pair of vacancies is about 0.9 ev in
the alkali halides_12 A pair of vacancies is neutral and thus will not lead to
ionic conductivity. On the other hand, they correspond to dipoles and
consequently may give rise to dielectric losses at relatively low frequencies;
for a review of recent work on this subject we refer to the literatureP
Also, pairs of vacancies of opposite sign appear to be very mobile in the
alkali halides and are therefore important for diffusion in these crystals. a
Besides single vacancies and pairs of vacancies, higher aggregates of
course are possible, as triplets, quadruplets, etc.
7-4. Example of self-diffusion in alkaJi haJides
As an example of diffusion in ionic crystals, some measurements by
Mapother, Crooks, and Maurer will be discussed briefly.1 5 These investigators measured the self-diffusion of radioactive sodium in NaCI and
NaBr in the following manner: A thin layer of about 5 X 10-4 cm of
radioactive saIt containing the isotope Na 24 was deposited on one face of a
cubic crystal, approximately 1 cm on edge. The crystal was then held at a
11 J. Tetlow, Ann. Physik, 5, 63, 71 (1949); Z. physik. Chem., 195, 197,213 (1950);
for further references see the article by F. Seitz in W. Shockley (ed.), Imperfections in
Nearly Perfect Crystals, Wiley, New York, 1952.
12 J. R. Reitz and J. L. Gammel, J. Chelll. Pllys., 19,894 (1951).
13 See, for example, the paper by R. G. Breckenridge in W. Shockley (ed.), Imperf<!ctiolls in Nearly Perfect Crystals, Wiley, New York, 1952.
It J. G. Dienes, J. Chem. Phys., 16, 620 (1948); F. Seitz. Phys. Rev., 79,529 (1950).
1> D. Mapother, H. N. Crooks and R. Maurer, J. Chem. Phys., 18, 1231 (1950).
Sec. 7-4]
169
constant temperature for a certain length 'of time. After this diffusion
anneal, the distribution of radioactive sodium was determined by means of
a sectioning technique, employing a microtome. In a similar fashion,
Schamp has investigated the diffusion of bromine in NaBr. 16 What one
measures in this way is the self-diffusion of the radioactive ions in the salt.
It must be emphasized that this type of experiment is altogether different
from one in which one heats the salt in the vapor of one of the constituents;
in such experiments one obtains information about the diffusion of color
centers in the lattice (see Sec. 15-6).
According to Fick's first law, the net flux of ions is proportional to the
concentration gradient (see Sec. 3-5), i.e.,
J
-D grad n*
(7-29)
on*
Tt =
(7-30)
Assuming the diffusion coefficient D to be independent of the concentration of radioactive ions, one obtains for the one-dimensional problem
under consideration,
on*
o2n*',
-=D-., '.
(7~31)
ox 2
at
The solution of this equation for the boundary conditions in the experiment
mentioned above iS17
n*
r
(7-32)
n*(x,t) = (7TD~)112 exp 4Dt
(_X2)
170
[Chap. 7
'"I
S
x
'i:
500
400
350C
-9
Q0 -10
]_
"'<=
.Q
;>
0
-11
-12
100
,,
" ".,
,,
,,
",
"
200
300
unit=2.34XlO- 7 cm 2)
Fig. 75. The fully drawn curve represents the directly measured self-diffusion
coefficient of Na in NaCl as function of
temperature. The dashed curve is calculated from the measured conductivity
by means of equation (7-45). [After
Mapother, Crooks, and Maurer, ref. 15]
-+- IDepth)2 II
_lOOO/T
(7-33)
'-----._---
where E is an activation energy. It must be noted, however, that the highand low-temperature regions have different activation energies of,
respectively, 1.80 ev and 0.77 ev. This point will be taken up below;
it is believed that the low activation energy results from the presence of
divalent positive impurities. For diffusion measurements on crystals
containing intentionally added divalent positive ions, see the work by
Witt and Aschner. 19 In the next section, the diffusion measurements will be
interpreted in terms of the migration of lattice defects.
19 H. Witt, Z. Physik, 134, 117 (1953);
J. F. Aschner, Thesis, University of
Illinois, 1954.
Sec. 7-5]
171
---.x
e.
N_ = _1_ . !!.. .
_!_ (n*
2a2 N 3 N
+ dn* a)
dx
Here 1/2a2 represents the total number of positive lattice sites per cm 2
172
[Chap. 7
*
*
np dn*
J=N...... -N..... =-6N2a dx
(7-34)
1 n
_a2 _p
3 N
(7-35)
The reader may compare this result with expression (3-21). As expected,
the self-diffusion coefficient is proportional to the number of vacancies
per unit volume n and to the jump probability of a vacancy per second p.
As in Sec. 3-5, p may be written in the form
(7-36)
The constant C arises from the thermal entropy change associated with the
production of vacancies, as discussed in Sec. 7-1. We note that in a plot of
log D versus II kT, the slope of the line according to the above interpretation
is determined by the sum (Ej
rf/2), i.e., by the energy required. for the
formation of vacancies plus the activation energy for jumping. Thus, from
the diffusion measurements of Na in NaCl, represented in Fig. 7-5, it
follows from the slope in the high-temperature region that E j
rf/2 =
1.80 ev.
The break in the log D versus liT curve leading to a smaller slope in
the low-temperature region may in principle be a result of either or both
of the following two causes: (1) the presence of divalent positive impurities,
(2) the freezing-in of positive ion vacancies. The explanation is as follows:
Suppose that a salt like NaCl contains in solid solution a small amount of
SrCI 2 or of the chloride of another divalent metal, the divalent positive
ions occupying sites which are normally occupied by the singly charged
Na+ ions. The condition of electric neutrality then requires that for each
divalent positive ion present, there must be a positive ion vacancy. Such
crystals then may contain at lower temperatures more positive ion
Sec. 7-5]
173
(7-38)
From (7-37) it follows that the pre-exponential factor in the expression for
the diffusion coefficient is equal to
when v'::: 1013, a'::: 3 . 10-8 cm and C,::: 100. The experimental value
of Do for the intrinsic region in NaCl, according to the work of Mapother,
Crooks, and Maurer, is 3.1 cm 2 sec-I, and 0.67 cm 2 sec-1 for NaBr.
Diffusion of positive ions does not necessarily take place as a result of
174
[Chap. 7
migration of single positive ion vacancies only. In fact, at least two other
possible diffusion mechanisms must be considered in the alkali halides:
(i) Diffusion resulting from migration of pairs.
- IA
+
r-,
,+
:c
I
I
:B
+
+
+
+
cf> -
11
kT
j1
(7-39)
Fig. 7-7. The pair AB may diffuse by the positive ion C jumping
into the vacancy A, or by a
negative ion jumping into B.
The associated complex divalent
positive ion-positive ion vacancy
may migrate as a result of interchange between the divalent ion
and the vacancy D, combined
with singly charged positive ions
jumping into the vacancy.
,f 'i
Sec. 7-5]
175
values of the binding energy between divalent positive iQns and positive ion
vacancies as calculated by Bassani and Fumi :21
NaCl ........ .
KCI ......... .
Cd2+
Ca2+
Sr2+
0.38 ev
0.38 ev
0.32
0.45 ev
0.32
0.39
Thus these binding energies are roughly half as large as those for pairs
of vacancies in the alkali halides. The study of the influence of divalent
impurities on the physical properties of alkali halides receives a good deal
of attention at present.
Although we have limited ourselves to the discussion of a rather
restricted area of the field of diffusion in ionic crystals, the same general
ideas apply to other cases. For further study we therefore refer the reader
to the literature. 22
+
7-6. Ionic conductivity in "pure" alkali halides
22
176
[Chap. 7
t_
t+
+ t_ =
l.
__
23
Sec. 7-6]
177
positive ion sites per cm3 by N, the number of positive ion vacancies per cm3
by n. If the electric field in Fig. 7-6 is directed to the right, a positive ion
vacancy will jump with a higher probability to the left than to the right,
because it is negatively charged. The potential energy along the line of
motion may therefore be represented by the full curve in Fig. 7-9 which is
the resultant of the dashed field-free curve and the linear potential due to the
external potential difference. Clearly then, the probabilities per second for
a jump to the left and to the right are, respectively,
p_
p_ =
iv exp [--(j -
t aeE)jkTJ
(7-41)
where the notation used is identical with that of Sec. 7-5; E represents
the field strength. The current density, i.e., the net flux of charge passing
per second through 1 cm2, is then equal to
(7-42)
because Ij2a 2 is the number of positive ion sites in a plane perpendicular
to the x-axis of an area of 1 cm 2 and njN is the probability for such a site
to be vacant. Now, for nearly all practical cases, aeE ~ kT, so that in first
approximation
n e2vEe-';lkT
(7-43)
1= - . - = aE
N 6akT
Now the number of vacancies n is given by (7-9), so that the conductivity
is equal to
Ce 2v
~,
(7-44)
a = 6akTexp [-(j
t~)jkT] ,
Ne 2 jkT
(7-45)
178
[Chap. 7
Sec. 7-7]
179
that all curves come together to a single straight line, the intrinsic region.
In that region, the conductivity is determined essentially by the density
of vacancies produced thermally. Thus the slope of the intrinsic curve
is determined by the sum of the activation energies j and cfo/2 in accordance
with (7-44). Now for each divalent ion there is one positive ion vacancy.
500
400
600
700C
s::
]' -6
t
-7
1.8
1.6
1.4
1.2
1.0
_lOOO/T
,"
H"
\"
'
180
[Chap. 7
a and the concentration of charge carriers n are known, the mobility !.t
(i.e., the velocity per unit field) may be calculated from the relation
a
ne!.t
(7-46)
\
Comparison of this expression with (7-43) shows that this in turn allows
one to calculate the jump probability. As mentioned before, the
probability for a jump of a positive ion vacancy at room temperature is
about I per second for the alkali halides. Once the mobilities are known,
the density of Schottky defects in the intrinsic range may be determined
from the measured conductivity. In this fashion Etzel and Maurer find
for the density of Schottky defects in the intrinsic range for NaCI,26
(7-47)
where 1> = 2.02 ev is the energy required for the formation of a positive
and a negative ion vacancy. Close to the melting point, this gives a
density of Schottky defects of about 1018 per cm3, i.e., about 1 vacancy
per 1()4 ions. At room temperature n c:::: 106 per cm3 It is of interest to
compare (7-47) with the theoretical expression (7-9). With N c:::: 1022
it follows that the constant Cc::::I O.
Information about the binding energy of a divalent positive ion and a
positive ion vacancy may be obtained from the fact that the "induc~d"
conductivity is not exactly proportional to the concentration of the added
divalent salt. In this way, Etzel and Maurer conclude that a fraction of
the vacancies is associated With the divalent impurities, the binding energy
being about 0.3 ev for NaCl containing CaC1 2 .26 However, this topic
is still in a state of flow and will not be discussed here any further. We
may refer to page 175, where calculated binding energies are given.
We mentioned in Secs. 7-5 and 7-6 that the break in the log D and log (]
versus liT curves is now generally interpreted as resulting from the
presence of divalent impurities rather than as a freezing-in of vacancies.
As experimental evidence we reproduce in Fig. 7-11 measurements by
Kerkhoff 27 of the conductivity and positive ion transport number for three
KCI crystals. It is important to compare the position of the knees in the
three cases; as the materials become purer, the knee shifts to lower
temperatures, in agreement with the above interpretation. It is also of
interest to note the influence of the divalent ions on the measured positive
ion transport numbers, mentioned in Sec. 7-6. Evidently most of the transport numbers quoted in the literature are unreliable as a consequence of the
presence of impurities. The recrystallization of the "analytically pure"
KCI carried out by Kerkhoff corresponds to a tenfold increase in purity.
26 H. W. Etzel and R. J. Maurer, J. Chern. Phys., 18, 1003 (1950).
2, F. Kerkhoff, Z. Physik, 130,449 (1951).
Sec. 7-7]
.~
C)
"Cl
<lI
'2
<Il
'"
VI
<lI
;::;
" /:
\":;
""
...;
"'"
...;
'"
i:-
t)
181
5'"
8....
:.::
....
""
I
8....
0;0
It:l
OC!
...;
',I,'j'
C)
8
0;0
'"
r-r---.------,;::;
,.,:;j'
..t
'"
oS
<lI
bI)
t:l
1
o
....
::s
""+
~
""
0;0
...;
....
gj
<lI
;:!l
.~
0;0
182
[Chap. 7
REFERENCES
Pe~rect
PROBLEMS
7-1. From equation (7-9) calculate the number of vacancies per unit
volume, assuming N = 1022 cm-3 , 1> = 2 eV, and Vo = vv'2.
7-2. Assuming a simple Coulomb interaction between positive and
negative ion vacancies, estimate the binding energy of a pair of vacancies
in LiF, NaCI, and KI.
7-3. Neglecting ionic displacements, set up a general expression for
the energy required to produce a Frenkel defect in a crystal of the sodium
chloride structure, employing the simple Born theory. Calculate the
energy required to form a Frenkel defect in NaCi and compare the result
with that required to form a positive and a negative ion vacancy. (To
check your results, see, for example, W. Jost, Diffusion in Solids, Liquids,
Gases, Academic Press, New York, 1952, p. 108.)
7-4. Assuming only a Coulomb interaction between a divalent
positive ion and a positive ion vacancy, employing the static dielectric
constant of the medium, calculate the association energy of the complex
for NaCi. Compare the result with the more detailed calculations of
Reitz and Gammel, J. Chem. Phys., 19, 894 (1951) and of Bassini and
Fumi (footnote 21).
. I'
7-5. Neglecting thermal entropy changes, set up an expression for
the free energy of a crystal with the NaCi structure containing n 1 single
positive ion vacancies, n 1 single negative ion vacancies, and n 2 pairs of
28 R. G. Breckenridge, J. Chern. Phys., 16,959 (1948); 18,913 (1950); ~ee also his
article in W. Shockley (ed.), Imperfections in Nearly Perfect Crystals, Wiley, New York,
1952, p. 219; Y. Haven, J. Chern. Phys., 21,171 (1953) .
Chap. 7]
=0
and of/2n 2
183
= 0,
where represents the binding energy of a pair and 4> is the energy required to produce a single positive and negative ion vacancy.
7-6. On the basis of the simple Born lattice theory calculate the energy
required to create a positive and negative ion vacancy in MgO. Assume
that in. the Jost model R+ = 0.6a and R_ = 0.9a and use for the dielectric
constant the value 9.8 (Answer: The lattice energy is 41 ev per ion pair;
the total polarization energy in 34 ev; 4> = 7 ev).
7-7. Consider a crystal of monovalent ions of the NaCl structure.
Let N represent the number of positive ion sites per cm3 , nd the number of
added divalent positive ions per cm3 . Furthermore, let nc be the number
of associated complexes per cm3, so that (nd - nc) equals the density of
free positive ion vacancies and free divalent ions. Show that in thermal
equilibrium
::'!"",
ncno ' = I2e/kT
(n d - n c)2
",
\
Chapter 8
FERROELECTRICS
8-1. General properties of ferroelectric materials
The dielectrics discussed in the preceding chapter show a linear relationship between polarization and applied electric field. In the present chapter
we shall deal with dielectrics for which this relationship exhibits hysteresis
effects. Since the dielectric behavior of these materials is in many respects
analogous to the magnetic behavior of ferromagnetic materials, they are
called ferroelectric solids, or simply ferroelectrics. A ferroelectric is
spontaneously polarized, i.e., it is
p
polarized in the absence of an external
field; the direction of the spontaneous
polarization may be altered under
""",.. )',.'" influence of an applied electric field.
In general, the direction of spon-----r..;---;,.t---.f----~ E
taneous polarization is not the same
throughout a macroscopic crystal.
Rather, the crystal consists of a
number of domains; within each
domain the polarization has a specific
direction, but this direction varies
Fig. 8-1. Schematic representation of from one domain to another. On the
hysteresis in the polarization versus basis of the domain concept, the
applied field relationship.
occurrence of hysteresis in the P
versus E relationship can be explained
as follows: With reference to Fig. 8-1, consider a crystal which initially has
an over-all pOlarization equal to zero, i.e., the sum of the vectors representing the dipole moments of the individual domains vanishes. When an
electric field is applied to the crystal, the domains with polarization components along the applied field direction grow at the expense of the
"antiparallel" domains; thus the polarization increases (0 A). When all
domains are aligned in the direction of the applied field (BC), the polarization saturates and the crystal has become a single domain. A further
increase in the polarization with increasing applied field results from "normal" polarization effects discussed in the preceding chapter; rotation of
domain vectors may also be involved if the external field does not coincide
184
Sec. 8-11
FERROELECTRICS
185
with one of the possible directions of spontaneous polarization. The extrapolation of the linear part BC to zero external field gives the spontaneous
polarization Ps The value of P" so obtained is evidently the same as the
polarization which existed already within each of the domains in the virgin
state corresponding to 0 in Fig. 8-1. Thus, when we speak of "spontaneous
polarization" we have in mind the polarization within a single domain and
not the over-all polarization of a crystal. We note here that the spontaneous polarization and its dependence on temperature, or on other
external conditions that might be imposed, can be measured by displaying
the hysteresis loop on an oscilloscope screen. When the applied field for a
crystal corresponding to point B in Fig. 8-1 is reduced, the polarization
of the crystal decreases, but for zero applied field there remains the
remanent polarization Pr where Pr refers to the crystal as a whole. In order
to remove the remanent polarization, the polarization of approximately
half the crystal must be reversed and this occurs only when a field in the
opposite direction is applied. The field required to make the polarization
zero again is called the coercive field Ec. It is evident that if the coercive
field is larger than the breakdown field of the crystal, no change in the
direction of spontaneous polarization can be achieved, i.e., under those
circumstances we cannot speak of the solid as a ferroelectric.
In connection with the last remark a few words may be said here about
the crystal structure of ferroelectrics. A necessary, but not sufficient,
condition for a solid to be ferroelectric is the absence of a center of
symmetry. ]n total there are 21 classes of crystals which lack a center of
symmetry; the classes are based on the rotational symmetry of crystals.
Of these 21 classes, 20 are piezoelectric, i.e., these crystals become polarized
under influence of external stresses. As soon as the crystal structure of a
particular solid falls within this group, it can be predicted to be piezoelectric; piezoelectricity is thus determined solely by the symmetry
properties of a crystal. Ten out of the 20 pieozelectric classes exhibit
pyroelectric effects. These pyroelectric crystals are spontaneously polarized.
However, the polarization is usually masked by .urface charges which
collect on the surface from the atmosphere; when the temperature of such
a crystal is altered, the polarization changes and this change can be
observed, hence the name pyroelectricity. As in the case of piezoelectricity,
pyroelectric properties can be predicted as soon as the crystal structure of
the solid has been determined. The ferroelectric materials discussed below
are part of the group of spontaneously polarized pyroelectrics. However,
they have the additional property that the polarization can be reversed by an
applied field. This additional feature cannot be predicted from the crystal
structure; it can be established only on the basis of a dielectric experiment.
The ferroelectric properties of a ferroelectric disappear above a
critical temperature Tc; this temperature is called the ferroelectric Curie
temperature. Associated with the transition from the ferroelectric to the
186
FERROELECTRlCS
"
[Chap. 8
C' j(T - 8)
+ EO
(8-1)
Sec. 8-2]
FERROELECTRICS
187
chemical formula NaKC1H 4 0S4H 2 0.1 The salt was tlrst prepared in
1672 by a pharmacist Seignette, living in Rochelle; it is therefore also
Jknown under the name Seignette salt. It is representative of the "tartrate
group." Other members of this group are those in which a fraction of the
potassium in Rochelle salt is replaced by NH 4 , Rb, or TI. Lithium
ammonium tartrate and lithium tantalum tartrate also belong to this
group.
Rochelle salt has the peculiar property of being ferroelectric only in the
temperature region between -18C and 23C, i.e., it has two transition
temperatures. In the region above 23C and below -18C it crystallizes
in the orthorhombic structure (three mutually perpendicular axes a, b, c).
In thc ferroelectric phase the crystal is monoclinic and the angle bctween
the a- and c- axes differs from 90. The spontaneous polarization
occurs along the direction of the original orthorhombic a-axis. Thus
Rochelle salt has only one polar axis and two possible polarization
directions (+ and -- along the a-axis). The domain pattern of this salt is
therefore rather simple.
The dielectric constants for Rochelle salt along the three axes are
given in Fig. 8-2, according to Halbliitzel. 2 Note that Ea reaches values as
high as 4000 near the transition temperatures. In the region above 23C the
susceptibility along the a-axis can be represented by the Curie-Weiss law,
"
FER ROELECTRICS
255'K
[Chap. 8
!l
\
296'K
I
I
I
:l
~
..s""
2
E"
100
150
200
--TI'K)
-T('K)
"
',.L
189
FERROELECTRICS
Sec. 8-2]
P.
"bD
..Q
--~-'->.-
1
OL_~~~~~~~~~
100
105
110 115
-TtOK)
120
125
Fig. 11-4.
3
2
1
0
50
100
150
200
-TtOK)
250
300
= 4.5
+ 3100/(T -
121)
G. Busch and P. Scherrer, Naturwiss., 23, 737 (1935); G. Busch, Helv. Phys. Acta,
11,269 (1938).
6 B. C. Frazer and R. Pepinsky, Acta Cryst., 6, 273 (1953); S. W. Peterson, H. A.
Levy, and S. H. Simonsen, J. Chem. Phys., 21, 2084 (1953); Pllys. Rev., 93, 1120 (1954);
G. E. Bacon and R. S. Pease, Proc. Roy. Soc. (London), A220, 397 (1953).
, See L. Pauling, Nature olthe Chemical Bond, Cornell University Press, Ithaca, 1945.
"
FERROELECTRICS
190
[Chap. 8
Replacement of hydrogen by deuterium in KH2PO~raises its Curie temperature from 123 0 to 213K, an increase of 90C. 8 It thus seems fairly certain
that the hydrogen bonds are essential in the polarization of this group of
ferroelectrics.
3. Wainer and Salomon in 1942 \ )served a number of anomalous
dielectric properties of barium titanate (BaTi0 3 ). It was recognized in this
country as a ferroelectric material by von Hippel and coworkers 9 and
independently, by investigators in England, Holland, and Switzerland.
This brings us to the third group of ferroelectrics, viz., the so-called
oxygen octahedron group. This group
i.
can be subdivided into others, one of
which is the subgroup of the perovskites
P Ba2+ with the general chemical formula A B0 3 ,
o ()2- where A is a di- or monovalent metal
Ti4+
and B is a tetra- or pentavalent metal.
BaTi03 is the most important and most
thoroughly studied representative of the
perovskites. In the nonpolarized phase
Fig. 8-6. The structure of BaTi0 3
it has cubic symmetry; the Ba2t ions
in the cubic phase.
occupy the corners of a cube, the
oxygen ions are located at the centers
of the faces, and the Ti4+ ion is at the center (see Fig. 8-6). Typical for the
BaTiO;j structure and for the other members of this group is the arrange. ment of the highi) j'olarizable oxygen ions in the form of an octahedron
with a small metallic ion at the center.
Barium titanate has an upper transition temperature of 120C; above
this temperature it is non ferroelectric and has the cubic structure of
Fig. 8-6. In this region the dielectric constaD t is well described by the
.. L.
Curie-Weiss law,
= 1. 7 X
105 j(T -
393)
Dir. ofpo/.
S/ruc/lire
278-393
193-278
<193
[001)
[011]
tetragonal
orthorhombic
rhombohedral
[III]
The transition points are evident from Fig. 8-7 and Fig. 8-8, representing,
respectively, the dielectric constant and spontaneous polarization as
8 B. T. Matthias, Science, 113,591 (1951); see also Phase Trans/ormations ill Solids,
National Research Council, Wiley, New YOlk, 1951.
" For a review, see A. von Hif'lpc1 R, us. Mod. Pllys., 22, 221 (1950).
Sec. 8-2]
FERROELECTRICS
191
...
10,000
'"
8000
i
4000
2000
o~==~~~=c~~~~~~
90
130
170
210
250
290
330
__~
370
410
-+T('K)
Fig. 8-7.
"u.
'"S
-...
u" 12
<0
0
.....
8
:j", :
I
o..~
\ ,. j
o ---'--
120
.'.
180
240
360
300
-T('K)
Fig. 8-8.
v2
vi
be multiplied, respectively, by
and
Thus the "Spontaneous polarization is nearly constant in the region below say 300 K.
It is interesting to note that the Curie temperature of barium-strontium
titanate mixtures varies linearly with the lattice constant of the mixed
0
FERROELECTRICS
[Chap. 8
;tals. n In this way Curie temperatures between 83K and 393K can
obtained.
Other compounds of the perovskite structure which are known to be
'oelectric are KTa0 3 , NaTa0 3 , KNb0 3 (Tc = 708K) and NaNb0 3
= 913K).
,
4. Recently a fourth group of ferroelectrics has been found which is
related to the groups mentioned above. Th's group is exemplified, by
lllidine aluminium sulfate hexahydrate, Nfl ':NH 2)zAIH(S04h6H 20.12
e structure of these compounds is present] unknown; they apparently
compose before a Curie temperature is reached.
3. The dipole theory of ferroelectricity
Ei= E+ yP
(8-2)
12
..t
193
FERROELECTRICS
Sec. 8-3]
the following manner. As long as one is far away from saturation of the
polarization, one may write in accordance with (6-16)
P = Nfl (cos e) = N ((l2j3kT)Ei
(8-3)
volume. l4
Nfl2j3kT
1 _ NYfl2j3kT
ejy
T_ e
(8-4)
1/
= Nfl (cos () =
= N flL
NflL
[:T (E
(:~)
+ yP) ]
,/~
o
(8-5)
10
Fig. 8-9. The fully drawn curve represents the Langevin function; the dashed
lines are those given by expression (8-7)
for various temperatures. The slope of
L(x) at the origin is 1/3.
L(x)
(8-6)
where
(8-7)
14 For simplicity, the contributions to P resulting from electronic and ionic displacements will be neglected in this section, because it does not impair the essential arguments.
15 It can be shown that the origin, which is also a common point of the straight line
and the Langevin function, corresponds to an unstable physical state; PI> however,
corresponds to a stable physical state.
'. ,
194
FERROELECTRICS
[Chap. 8
decreases, the slope of the straight line (8-7) decreases and the solution
P/Poat approaches unity. Also, when the temperature is higher than a
critical value determined by
\
kTc/Np2y
=!
or
Tc = N p 2y /3k =
(8-8)
it is observed that (8-6) and (8-7) intersect only at the origin. (Note that
in this model Tc = 0.) In other words, there is no spontaneous polarization
for T > e. By means of the method outlined above, it is thus possible,
to find PIP,at as function of Tie and the
result is represented in Fig. 8-10. It is
observed that just below the Curie temperature, the spontaneous polarization
increases rapidly, in agreement with experiment. (Compare, for example, Fig.
_. TIB
8-4.)
(8-9)
Thus, if Ce(T) and P sat are known from experiment, (8-9) allows one to
calculate the internal field constant y. However, y may also be obtained
"
195
FERROELECTRlCS
Sec. 8-3]
from (8-4) if the Curie constant and the Curie temperature are known.
According to Blattner and Merz one obtains the following results :16
y from (8-9)
Rochelle salt.........
BaTiO a .............. .
2.1
0.044
KH 2 PO ........ : .... .
0.37
y /rolll
(8-4)
2.2
0.049
0.48
Nfl
28120 esu
I.
buted by the timeaverage of the reaction field; this part is equal to Er(cos 0)
where 0 is the angle between !L and E and, as emphasized above, does
not produce a torque on the dipole. To find the actual field strength tending to orient the dipole, one must subtract the reaction field component in
the external field direction. This may be done simply by first taking away
the dipole and calculating the field inside the cavity. The field so obtained
_
._
\
is called the cavity field and is equal to
Ec
= [31:/(21:
1)]
_ (8-10)
Ec = E
+ ~P/(21: + 1) = E + Y(I:)P
(8-11)
.t
8-5. Ionic displacements and the behavior of BaTi03 above the Curie
temperature
In Sec. 8-3 we have seen that the dipole theory, with an expression of
the type (8-2) for the internal field, led to a Curie-Weiss law for the susceptibility above the Curie temperature. However, in Sec. 8-4 it was pointed
J. M. Luttinger and L. Tisza. Phys. Rev., 70, 954 (1946); 72,257 (1947).
C. Kittel, Phys. Rev., 82, 313, 729 (1951).
22 S. Roberts, Phys. Rev., 83, 1078 (1951); E. Sawaguchi, H. Mariwa, and S. Hoshino,
Phys. Rev., 83, 1078 (1951) .
20
21
.)oc. ~L.onaO~), 14,}, 1 (1<J35): !t is illustrative to compare the dipole theory of ferr~
electriCIty wIth the Bragg-Wllhams theory for order-disorder transitions in alloys (see
Chapter 4).
FERROELECTRlCS
Sec. 8-5]
197
out that the internal field (8-2) could not be considered the field producing a
torque on the dipoles. On the other hand, the objection of Onsager raised
there does not refer to electronic and ionic displacements, and for these an
internal field of type ~-2) may stilI be applied. In this section it will be
shown that in case the dielectric constant of a material is large compared
with unity, a Curie-Weiss law may be obtained which is solely due to
electronic and ionic displacements. 23 At first sight this may seem somewhat
surprising because one generally connects a strong temperature dependence of with the existence of permanent dipoles. .. .. ::;" .~ ..... J
For the sake of argument, let us
_ -_
assume that for a particular nondipolar
" t - - ~.--,-solid the Clausius-Mosotti expression
4
_.' . _.
'--r
holds (which is based on the Lorentz
'"
internal field formula):
;:: 3
(-I)J(+2)=(47TJ3)NIX=fJN (8-12)
_-
+ 2)(
d
1 dN
I) dT = N . dT =
- 3A.
(8-13)
Jd = - JA dT
2
or
= T__!_f!__
_ ()
(8-14)
The last expression has indeed the form of the Curie-Weiss law; the
Curie temperature () enters as a constant of integration. It is of interest to
note that the Curie constant is equal to the reciprocal of the linear coefficient of expansion. For BaTi03 , ;. ~ 10-5 per degree, which gives fair
agreement with the experimental value for the Curie constant quoted in
23
198
[Chap. 8
FERROELECTRICS
where (Xri and (X((i are the polarizabilities of the ions assr Jated with
electronic and ionic displacements, respectively.23 Now, if t:lt: right-hand
side of this expression becomes unity, E becomes infinite, and spontaneous
polarization will occur. Thus the TiH displacement would have to account
for less than 38 per cent of the total polarization.
One theory based on the assumption that the Ti4+ ions are mainly
responsible for the ferroelectric properties of BaTiO a has been developed
by Mason and Matthias. 25 These authors assume that the stable position
for the Ti 4+ is not in the center of the unit cell (Fig. 8-6) but that there
exist six stable positions corresponding to slight displacements from the
center toward the six surrounding oxygen ions. In each of these positions,
the unit cell would thus bear a dipole moment. They furthermore assumed
an internal field of the type (8-2) and essentially their theory is similar
to the dipole theory discussed in Sec. 8-3. With this theory it is not
H. D. Megaw, TrailS. Faradav Soc., 42A, 224, 244 (1946).
l'iP. Mason and B. T. Matthias, Phys. Rev., 74, 1622 (1948); also W. P. Mason,
Piezoelecfric Crystals and Their Applications in Ultrasonics, Van Nostrand, New York,
1950.
..
I .
Sec. 8-6]
FERROELECTRICS
199
200
FERROELECTRICS
[Chap. 8
six ions belongs to two unit cells), which may be denoted by Ox, Oy and 0 .
As the crystal is cooled from above the Curie point, the cubic lattice
contracts and at the Curie point one of the three oxygen ions is squeezed
out of the plane of the barium ions, let us say the ion Oz. This produces a
dipole moment per unit cell along the z-axis, part of which is equal to
2edz , where d z is the displacement of the ion 0z relative to the plane of BaH
ions .. At the same time, this allows for a possible contraction of the
lattice in the plane of the barium ions. The direction of polarization
corresponds to the c-axis of the tetragonal structure and sets in along
one of the cube edges at the Curie temperature. As the temperature is
lowered further, the Oy and Ox ions are successively squeezed out of their
normal positions, leading to a polarization along a face diagonal [OIl] and
a body diagonal [Ill], respectively, (by combination of their own effect
with the polarization already existin~. This model is in agreement with
the changes of structure associated with the changes in polarization
direction mentioned in Sec. 8-2. Also, X-ray diffraction studies have
shown that the oxygen ions are indeed displaced by 0.08-0.1 A relative to
the BaH ions; the displacement of the TiH is about 0.06 A according to
these measurements (the cube edge of BaTi0 3 is 4.00 A).31 The essential
feature of this model is that it combines the mechanical forces with the
electric forces.
On the quantitative side, the following simple ar~ment may b(!
put forward: Experimentally it is found that in the tetragonal region
the contraction of the lattice is proportional to the square of the polarization
and satisfies the relation
dala = 1.2 X 1O-12P2
/
(8-16)
where a is the cube edge just above the Curie point and da is the (vntraction
in the tetragonal phase. Now, in the cubic phase, the sum of the radii of the
BaH and 0 2- ions is equal to alv2. Suppose now that the oxygen ion is
displaced out of the plane of Ba 2+ ions by an amount z and let it be assumed
that the radii of the ions remain constant and that the oxygen and barium
ions remain in contact. With reference to Fig. 8-13 it then follows that if
(a - fla) is the new edge of the square of Ba2+ ions, we must have
___~_
(a -
As long as fla/a
flan2 = a 2/2 -
Z2
'.
."
I, this yields
flala
(8-17)
(zla)2
The dipole moment per unit volume resulting only from the displaced
oxygen ions is equal to Po, = 2ezla3 and it thus follows from (8-17) that
fla/a
31
(8-18)
H. T. Evans, Acta Cryst., 4, 377 (1951); W. Kanzig, Helll. Phys. Acta, 24, 175
(1951).
. . . , ;
Sec. 8-6]
FERROELECTRICS
201
Comparison of (8-16), and (8-18) shows that both expressions are of the
same form, and that if po. represents two thirds of the total polarization,
the agreement is quantitative. Although the oxygen displacement theory
has attractive features, recent neutron diffraction studies suggest that the
oxygen octahedra suffer little distortion in passing through the transition,
in contradiction with the theory. One must therefore conclude that the
problem is still not solved satisfactorily. 32
(8-19)
The coefficients c are functions of temperature; the numerical factors
are introduced for later convienence. Note that since we want the
free energy to be the same for "positive" and "negative" polarization
along the polar axis, only even powers of P are included. In thermal
G. Shirane, F. Jona, and R. Pepinsky, op. cit.
For a review of this work see A. F. Devonshire, "Theory of Ferroelectrics,"
Advallces ill Physics (quarterly suppl. of Phil. Mag.), 3, April 1954, p. 85.
32
33
202
Chap. 8
FERROELECTRICS
equilibrium (oF/? Ph
the equation
_p
ta)
tb)
Sec. 8-7)
FERROELECTRICS
203
Hence the applied field may be written E = (oFloPh .. Above the transition temperature the polarization will be small for small applied fields,
and in this region we may neglect all terms on the right-hand side of
(8-19) except the first. We thus obtain for T> To,
.'
,.-.~-
Ilx
.. I(
= dE/dP
C1
(8-23)
where Xa is the susceptibility above the critical temperature; the coefficient ci is evidently equal to the reciprocal of the susceptibility Xn'
However, we know that in this temperature range the susceptibility is
given by the Curie-Weiss law X" = C;(T - 0), so that c i = (T - O)/C,
where C is the Curie constant. However, since the transition at To
corresponds to CI = 0, we have 0 = To and thus
(8-24)
In the ferroelectric region we obtain likewise from (8-19) and (8-23),
(8-25)
where Xb is the susceptibility below the transition temperature; the
terms with powers :;::'6 have been neglected in (8-19). For small applied
P, in this region, so that according to (8-25) and (8-21) we
fields, P
have
(8-26)
r--J
204
[Chap. 8
FERROELECTRICS
where So is the entropy of the un polarized crystal. To Q first approximation we may then write
\"
i
(8-28)
Since P is a continuous function of temperature for the case under consideration and since the slope of p2 has a discontinuity at T = Te, there
l/xa
I
(a)
()
""
'Tc
(b)
:}<
,"l
205
FERROELECTRICS
Sec. 8-7]
--T
(b)
(a)
0- 2CIPs(Tc)
-t.
+ 4,C 2 P.(Te) +
sc3 Ps(Tc) +
~{>~f,Uc :"t:}l
...
(8-30)
-i(C 2 /C 3 );
C1 = 13S(c~/C3);
P;(TJ = 3c 1/C 3
(8-31)
C/(T - 0)
and
c1 = (T - O)/C
(8-32)
206
[Chap. 8
FERROELECTRICS
making use of the relations (8-31) we find for the susceptibility below the
critical temperature,
(8-33)
At the critical temperature C1 is, according to (8-32), equal to (Tc - ()/C
and the susceptibilities just above and just below Tc are given by
l/X"
(Tc - ()/C
and
l/Xb
= 4(Tc -
()/C
for
Tc (8-34)
The reciprocal susceptibility as one passes through the transition temperature is illustrated in Fig. 8-ISb.
We should mention here that decisive evidence as to whether a particular ferroelectric transition is of the first order may be obtained from
a so-called "double loop" experip
ment in which the transition is induced slightly above the critical
temperature Tc by application of a
strong electric field. Such an induced
transition was first produced by
Roberts in ceramic material and has
more recently been demonstIated for
a good single crystal of BaTi03 by
Merz. 34 A strong a-c field is applied
to the crystal a few degrees above its
normal transition temperature. At
zero applied field the crystal if
Fig. 8-17. Schematic representation of
non ferroelectric but at a critical
a double hysteresis loop, of the type
of the apphed field the polarivalue
observed for BaTiO a, slightly above the
zation increases rapidly and upon
transition tt:lllperature.
reversal of the field hysteresis is
observed. The hysteresis loop is not complete, however, and for
low applied fields the behavior is normal again (see Fig. 8-17).
A double hysteresis loop obtained in this manner can only occur if the
transition is of the first order, as may be understood in the following
manner: In the absence of an applied field the transition occurs when in
Fig. 8-16a the minimum of the free energy for P s = 0 is equal to the
minimum associated with nonvanisning value of the spontaneous polarization. For a crystal subjected to a field E, however, the induced transition occurs when F--EP rather than Fhas the same value as the minimum
at the origin. Such induced transitions can evidently occur only if the free
energy curves are of the type illustrated in Fig. 8-16a and not if they are
of the type corresponding to Fig. 8-14a. Hence the double loop experiment distinguishes between first- and second-order transitions. Since a
at
S. Roberts, Phys. Rev., 85, 925 (1952); W. J. Merz, Phys. Rev., 91, 513 (1953).
Sec. 8-7]
FERROELECTRICS
207
double lo'op has been observed for BaTi0 3 , the upper transition of this
material is evidently of the first order. We should note that it is usually
not possible to obtain a clear-cut distinction between a first- or secondorder transition from measurement of the spontaneous polarization as
function of temperature, since P s rises rapidly just below Tc even for a
second-order transition. For further details on the thermodynamic theory
of ferroelectricity and for a treatment of antiferroelectric transitions we
refer the reader to A. F. Devonshire, op. cit.
8-8. Ferroelectric domains
It was mentioned in Sec. 8-2 that when a Rochelle salt crystal is cooled
to below the Curie temperature, spontaneous polarization along the
a-axis of the orthorhombic structure sets in. In general, however,
the direction of spontaneous polarization is not the same throughout the
crystal; certain regions are polarized in the +a direction, others in the
-a direction. These regions are referred to as domains. The boundaries
between domains are called domain walls. In a Rochelle salt crystal the
domains are polarized in opposite directions. For KH2PO~ there is also
only one axis along which spontaneous polarization takes place, viz., the
c-axis of the tetragonal structure. The domain structure is thus similar
to that of Rochelle salt. In the case of BaTiO:l , spontaneous polarization
may occur along anyone of the three edges, leading to six possible
directions for the spontaneous polarization. The domain structure for
BaTi0 3 is therefore more complicated than in the other two groups of
ferroelectrics.
The ferroelectric domains are the electrical analogues of the Weiss
domains in ferromagnetic materials, although there are certain interesting
differences in their formation and growth, as we shall see below. The
existence of domains, which has been confirmed by X-ray investigations and
optical studies 35 , explains the possibility for a crystal below the Curie
temperature to have a zero or very small total polarization. By applying
an dectric field to such a crystal, the number and size of domains polarized
in the external field direction may be increased. This process leads,
upon reversal of the field direction, to hysteresis in the P versus E curves,
and gives rise to dielectric losses. These losses are proportional to the
area of the hysteresis loop and to the frequency of the applied a-c field.
Optical observation of ferroelectric domains is possible since ferroelectrics are birefringent. In BaTi0 3 for example, the optical axis coincides
with the direction of spontaneous polarization. Thus a domain polarized
in a direction perpendicular to the surface of a crystal plate looks dark
'" B. T. Matthias and A. yon Hippel, Phys. Rev., 73,1378 (1948); P. W. Forsbergh,
Phys. Rev., 76, 1187 (1949); Blattner, Kanzig, Merz, and Sutter, Helv. Phys. Acta, 21,
207 (1948); W. J. Merz, Phys. Rev., 95, 690 (1954).
208
FER ROELECTRICS
. ""
[Chap. 8
--
-t::====-
--
3.
Sec. 8-8]
209
FERROELECTRICS
REFERENCES
W. G. Cady, Piezoelectricity, McGraw-Hili, New York, 1946.
PROBLEMS
8-1. Let P be the spontaneous polarization of a ferroelectric solid and
let yP be the internal field. Show that the "extra specific heat" of the
material is given by C = -(y/2)(dp2/dT). For the dipole theory, draw the
curve for specific heat versus temperature and show that at the critical
temperature the specific heat is equal to 3k/2 per dipole.
8-2. In the theory of Mason and Matthias 25 of ferro electricity of
barium titanate it is assumed that the TiH ion has six stable positions
corresponding to small displacements from the center of the unit cell
toward the six surrounding oxygen ions. If the absolute value of the dipole
moment due to this displacement is p, show that in a field appJied parallel
to a cube edge, the polarization due to these dipoles is given by
where N is the number of unit cells per unit volume and Ei is the internal
field. Introduce the approximation pEi ~ kT and compare the result with
31 T. Mitsui and J. Furuichi, Phys. Rev., 90, 193 (1953);
R. Sommerhalder, Heir,. Phys. Act.a, 26,603 (1953).
W. .Kiinzig and
210
FERROELECTRICS
[Chap. 8
that for freely rotating dipoles. Assume further that the internal field is
E
y(P,[
rx.E;) where
represents the polarizability
given by Ei
per unit volume with the exclusion of the Ti4+ displacements. Show that
spontaneous polarization can occur only below a critical temperature
Tc = (yN;-t2/3k)/O - yrx.). Show further that the dielectric constant of the
material is given by
. .
. \.
= +
rx.
E=l+~(rx.+!'~)
1- yrx.
y T- Tc
8-3. Discuss an experimental method for growing large platelike
single crystals of BaTi03 . (See l. P. Remeika, J. Am. Chem. Soc., 76, 940
(1954).)
8-4. Discuss the results of X-ray and neutron diffraction studies of
ferro- and antiferr.oelectric ~aterials. See, fOf example, G. Shirane,
F. lona, and R. Pepmsky, op. Cit.
'
8-5. Consider a system of dipoles and assume that the field acting on a
given dipole is equal to the cavity field (8-1 1). If there are N dipoles percm3 ,
show that the dielectric constant of the system is
E
= 1
+ ! [47TNIX -
+ ( 1 + 87TNIX/3 + 167T2N21X2)1/2]
where IX = ;-t2/3 kT. This shows that E remains finite for any finite temperature, i.e., the system is nonferroelectric.
..f ,
., .'1
-'
;)
Chapter 9
212
"
[Chap. 9
',\
Sec.9-2J
213
_----CEI,2
T=O
Fig. 9-2. The curve CE* represents Z(E) in accordance with (9-3):
the energy distribution N(E) is
obtained by lllultiplyingZ(E) by F(E).
(9-2)
It is frequently convenient t~have an expreSsion for the number of allowed
states in an energy range between E and E + dE. This may readily be
obtained from (9-2) by replacing p2/2m by E, yielding
Z(E)dE=CEl!2dE with
C=47TV(2m)3/2jh3
(9-3)
The function Z(E) is represented schematically il~ Fig. 9-2. To find the
states actually occupied by the free electrons at a temperature T, we must
3 C. Davissun and L. H. Germer, PhI'S. ReL"., 30, 705 (1927): H. Bethe, AIIII. Physik.
87, 60 (1928).
' . ,,~
'
'OJ,
[Chap. 9
214
N(E) dE = Z(E)F(E) dE
wIth
1
F(E) = -e-:X+~E~/k~T-+-1
(9-4)
where ex is a parameter and F(E) is called the Fermi function. Note that
F(E) simply represents the fraction of possible states which is occupied.
When there is about one free electron per atom in the metal, the electron
gas may be expected to be highly degenerate at room temperature. This
implies that e? < I and we shall therefore write (see Appendix D)
e? = e- EF / k1 '
1'1E)
t
l~----------~,,~~T=O
kT
.5
~EF./"
\
\, --+-E
OL---------------~~
F(E) =
EFo
1. T
O.
(9-5)
e (E-E F )/k1'
+1
(9-6)
for
E<EFo
E>
(9-7)
EF o
Thus at absolute zero, all possible states below EFo are occupied. all
those above EFo are empty. The physical meaning of EFo is, therefore,
that it represents the highest occupied energy level at T = 0 (see Figs.
9-2 and 9-3). It is of interest to calculate EF II in terms of the number of
free electrons per unit volume. In general, one must satisfy the condition
(9-8)
I,rEF
.0
E1!2 dE
or
h2 (3n)2/3
E., = __
"
0
2m 81T
(9-9)
'._
Sec. 9-3]
215
classical statistics and Fermi statistics. In the former case, all electrons
would have zero energy. For "classical electrons" to have an energy of
1 ev, a temperature of about 50000 K would be required.
Table 9-1. Fermi Energy Calculated from (9-9) and Work Function
cJl (Exp.) for Some Metals
Metal
Na
K
Cu
Ag
Ba
Al
Valence
I
(ev)
3.1
2.1
7.0
5.5
3.8
EFo
l' ~ t
T-~
)'
-;}
2.1
3
'" (ev)
2.28
_._2.~
;4:4f"-
. 4.46
_,--.l..S l __ ~__ .
4.20
11.7
(Eo> = !EF o
J j;;
e-(E-EF)/kT
for
E - Ep';P kT
(9-12)
roo
dE
=.10 Z(E) e(E-E,)/kT +
(E)
= N10
roo
EdE
Z(E)
e(E
E F )/k1'
(9-13)
+1
(9-14)
216
[Chap. 9
~ EFo
[1 -7; (;TfJ
~ (Eo)
[1 + ~~2 (;:rJ
Ep
(E)
Fo
\\
(9-15)
(9-16)
c"
= 7T2(kT/2EF.)k
= 7T2(T/2TF)k
(9-17)
Cv
AT+ BT3
(9-18)
Sec. 9-4]
217
When one calculates the coefficient A on the basis of (9-17) and uses
Ep o = 7.04 ev, calculated from (9-9), one obtains A = 1.24, which is
ftB
(9-18)
(9-19)
If we assume classical statistics M may be calculated in the same way
as the orientation polarization of electric dipoles in the Oebye-Langevin
theory. We leave it to the reader (Problem 9-4) to show that in that case,
.mi""
(9-20)
218
[Chap. 9
as long as flH ~ kT. The quantity X", is called the paramagnetic susceptibility. Note that for freely rotating dipoles the average component in
the field direction is (flJd3kT)H. The fact that the factor 3 is missing in
(9-20) is a consequence of the fact that the dipoles can accept only two
possible orientations relative to H. If (9-20) were correct, one would find for
the susceptibility of metals at room temperature with n ~ 1022 per cm3 ,
Xv ~ 10-4 per cm3 Also, Xv should vary as liT. Experimentally, however,
one finds Xv ~ 10- 6 per cm3 and practically no temperature dependence.
The disagreement with experiment disappears when one applies FermiDirac statistics, as was first shown by Pauli. 6 For simplicity let us first
consider the situation at T = O. Without external magnetic field all energy
ME)
levels below E F 0 are occupied and all those above E F 0 are empty. Leaving
for a moment all electrons in their original state and applying an external
field H, all electrons with a magnetic moment parallel to H would suffer
a shift in energy of -flEH, all antiparallel ones of +flnH. This is indicated
in Fig. 9-4a. It must be noted that fl BH ~ E Fo; in fact, even for a field
strength of 10 5 gauss, flBH ~ 10-3 ev as compared with EFo ~ 5 ev.
The situation as depicted in Fig. 9-4a is, of course, unstable and a number
of anti parallel spins will enter the group of parallel ones. In equilibrium
both halves are filled to the same level, as in Fig. 9-4b. Now, according
to (9-3), the number of allowed states in each of the halves is per cm3
equal to
z(E) dE = 27T(2m)3/2El/2 dElh 3
(9-21)
W, Pauli, Z Physik, 41, 81 (1927).
Sec. 9-5)
219
:;
'.
t
because fl IJH ~ E F'o The total excess of electrons with parallel orientation
is twice this, so that one finds by making use of expression (9-9) for EIi'.'
_ 47Tm
M - h2
(3n)1/:lflnH 2
7T
(9-22)
XI,H
1O-14n1!3
( :::'1':1.
(9-24)
~. nfl~
2 E Fo
[1 _ 12 (kT)2]
E,
2
7T
(9-25)
1'0
The factor in brackets is identical with that occurring in (9-15) for the
temperature-dependence of Ep. For T = 0, (9-25) reduces to (9-23).
With n ,..._, 1022 it thus follows from (9-23) that the paramagnetic susceptibility of the free electron gas is of the order of 10-6 per cm3 , in agreement
with experiment.
"I
A quantitative comparison between the results obtained above and
experiment is rather difficult. First of all, the magnetic susceptibility of a
metal consists of three contributions:
(i) The paramagnetic contribution of the free electrons
Oi) A diamagnetic contribution of the free electrons, first calculated
by Landaus
(iii) The diamagnetic contribution of the ionic cores
Thus, in order to obtain Xv' the last two contributions must be
subtracted from the total susceptibility measured. For completely free
electrons, contribution (ii) is equal to - Xv13. Contribution (iii) is usually
calculated from susceptibility data on ionic solutions; this involves the
7 E. C. Stoner, Proc. Roy. Soc. (London), A152, 672 (1935).
L. Landau, Z. Physik, 64, 629 (1930).
220
[Chap. 9
assumption that the susceptibility per ion is the same in the solution and
in the metal. Furthermore, the experimental data are sometimes impaired
by the presence of ferromagnetic impurities. Finally, the free electron
model can be expected to hold in good approximation only for the alkali
metals, as we shall see in Chapter 10.
",\_,
For further details, we refer to the literature. 9
\
P;')2171
Es
(9-26)
1 '
(9-27)
'P.l'n
Sec. 9-6)
221
per unit volume. The number of electrons occupying states with momenta
between Px, Px
dpx; PY' py
dpy; Pz, pz
dp, is therefore
where E
(p;
+ p; + p;)/2m.
2
n(px) dpx
2
h3
dp:rdpydpz
e(E -
+1
Ep)/O'
,t!l (l!
(9-28)
Hence
f+ocl'-1W
= Iii dpx.l_
'l]
dpudpz
oc e(t: -Hp)/"],
+1
(9-29)
Now we are interested only in those electrons for which Pr ;:::: P.r o' i.e., the
total energy of the electrons of interest is at least equal to E,. On the other
hand, E, - Ep = ?:> kTfor all metals at temperatures below the melting
point (see Fig. 9-1). Hence the term of unity in the distribution function
may be neglected; we are interested only in the Boltzmann tail of the
Fermi dis.tribution. The quantity is called the work function of the
metal; it represents the energy difference between an electron at the Fermi
level and the vaClJum level.
( __
One thus obtains from (9-29),
(9-30)
Substituting this expression into (9-27) one finds upon integration for the
emission current density,
1= A(I -
r)T2 e -</>lkT
.:L
),
(9-31)
where A
47Tcmk 2 /h 3 = 120 amp/cm 2/deg 2 . This is the DushmanRichardson equation.
From the form of (9-31) one may be inclined to conclude that by
simply plotting log U/T2) versus liT olle obtains from the slope of the
resulting straight line and A(l -- r) from the interc,pt at I IT = 0 (see
Fig. 9-5). A number of complicating factors in the thermionic emission
of an actual metal must, however, be considered. 12
(i) The apparent work function increases if a negative space charge
exists in the vicinity of the emitter; the anode potential should
therefore be sufficiently positive to prevent space charge build-up,
i.e., one should work in the region of saturation-current density.
(ii) The apparent work function decreases with increasing external
field strength, as explained in Sec. 9-8. Thus J(T) should be
I; ,
measured for different external fields and then extrapolated by
means of a so-called Schottky line to zero field strength (see
Fig. 9-7).
12 For an evaluation ot thermionic emission data and a thorough discussion of the
theory, see C. Herring and M. H. Nichols, "Evaluation of Thermionic Data," Revs.
Mod. Phys., 2], 185 (1949).
'---...
222
[Chap. 9
t10lOg (l/T2)
\,
\,
\,
\,
\,
-2
\,
\,
\,
-4
'I'
\,
\,
'J
,I'
\,
\,
-6
\,
1'11'
-&
~'"
; ~: f '
-10
-12
.2
I
Fig. 9-5.
,/
where
~o
~o
+ (d~/dT)T
(9-32)
J3 For an extensive study of this and other aspects of the thermionic emission of
tungsten, see G. F. Smith, Phys. Rev., 94, 295 (1954) .
.. Illustrative in this respect is a table of values for 4> and A(l - r) for platinum in
chronological order in G. Herrmann and S. Wagener, The Oxide Coated Cathode,
Chapman and Hall, London, 1951, Vol. 2, p. 78.
"I
.I .:
Sec. 9-6]
223
+ log (I
- r) - (d</l/dT)/k - </lo/kT
(9-33)
On the basis of this expression one thus determines from the slope of a
Richardson plot, such as represented for tungsten in Fig. 9-5, a value for
</lo rather than for cP. Also, it is evident that the constant obtained from
the intercept at liT = 0 may differ appreciably from A = 120 amp/
cm 2/deg 2 A number of experimental results obtained by various methods 15
indicate that for metals d</l/dT ~ 10-4 ev per degree. Work functions for
a number of metals are given in Table 9-2.
'-ty"
AI
Ag
Au
Ba
Cd
Co
Cr
4.20
4.46
4.89
2.51
4.10
4.41
4.60
Cs
Cu
Fe
K
Li
Mg
Mo
1.93
4.45
4.44
2.22
2.48
3.67
4.24
Na
Ni
Pd
Pt
Ta
W
Zn
2.28
4.96
4.98
5.36
4.13
4.54
4.29
veo: dveo: =
Vo:
dvo:
(9-36)
For a review of methods to determine r/>, see Herrmann and Wagener, I.c. Chap. 2.
224
[Chap. 9
(9-37)
(9-38a)
dS = ~k dN + dE/T+ P dV/T
(9-39)
dN= -1,
dE = -(El!'
+ cP + 2kT)
(9-41) . - ---
From the last three equations one obtains for the heat lost by the metal
per emitted electron,17
dQ
T dS
cP
+ 2kT
(9-42)
Note the important physical meaning of the work function in this result
16 C. Davisson and L. H. Germer, Pllys. Rev., 20, 300 (1922); 30, 634 (1927);
G. M. Fleming and J. E. Henderson, Phys. Rev., 58, 887 (1940).
17 A detailed thermodynamical study shows that an additional term must be added
to the right-hand side of (9-42), containing the Thomson coefficient; this term is of the
order of 10- 2 ev. See C. Herring, Pllys. Rev., 59, 889 (1941).
Sec. 9-71
225
for the latent heat of evaporation per electron. The power consumed by
the emitter per cm 2 due to this process is thus
P = (I/e)(c/>
+ 2kT)
(9-43)
From the power input and correcting for losses due to thermal radiation
and heat conduction, it is possible to determine c/> at a given temperature.
This method has been used to determine dc/>/dT.lB
9-8. Field-enhanced electron emission from metal..
."":-';'
<
--~'::
,.
,,
See, for example, F. Krueger and G. Stabenow, Ann. Physik, 22, 713 (1935).
[Chap. 9
V(x)
~e2/4x
- eEx
(9-45)
vhere the last term corresponds to the external field. The maximum of
his curve occurs for x = xm and from (9-45) one finds Xm = He/E)l!'I..
5ubstituting, one finds for the change in work function
/lrp =
V(x m )
-e(eE)l!2
(9-46)
log A
+ log (1
- r) - rpjkT + e(eE)1/2jkT
(9-47)
Thus, if one plots the logarithm of the saturation current for a given
temperature as fu'nction of the square root of the anode voltage, one
expects a straight line (the Schottky line). A comparison of theory and
experiment for tungsten is given in Fig. 9-7; the agreement is good for
anode voltages above 100 volts; the deviations below 100 volts are
ascribed to variations of the work function over the surface. 21
It should be noted that the actual change in the work function is
relatively small. For example, for E = 103 volts per em, one obtains
X lIl ~ 10- 5 em and /lrp ~ 0.01 ev.
19
20
Zl
'I
"
Sec. 9-81
227
Field emission. When the external electri<: field becomes of the order
of J06 volts per em, cold emission or field emission sets in. This phenomenon is quite different from the Schottky effect: in the latter case the
electrons cross over the potential barrier, in field emission they tunnel
through the barrier. For simplicity, consider a metal at absolute zero and
let us assume the surface potential barrier to be abrupt. The potential
energy of an electron outside the metal is then equal to -eEx, represented
by the line AB in Fig. 9-8. If the distance d in Fig. 9-8 is of the order of
10 Angstroms or less, electrons in the vicinity of the Fermi level will be
I;"
20
"";
......___
----<-;:--e.
.....
x
'0
o
....
t
2~
.
,
____- L_ _ _ _ _ _k -_ _ _ _
10
20
__ IV
30
)1/2
high-field
d should
less for
to take
able to tunnel through the barrier.22 For <p '::::' 3 ev, this requires a field of
the order of tO i volts cm-I . As the field strength becomes larger, more
and more electrons below the Fermi level begin to contribute to the
emission current. According to Fowler and Nordheim, the emission
current as function of the field strength E for a triangular barrier may be
written in the form
(9-48)
where Band fJ are constants containing the wor~ function. 23 Note that
E plays the same role in this formula as T in the Dushman-Richardson
expression for the thermionic current. Thus if log (II E2) is plotted versus
lIE, a straight line should result. This has been confirmed by experiment. 2i
Usually, field emission sets in at fields of the order of 10 6 volts cm- I ,
probably as a consequence of high fields occurring at surface irregularities.
" See, for example, N. F. Mott and I. N. Sneddon, op. cit .
.. R. H. Fowler and L. Nordheim, Proc. Roy. Soc. (London), 119A, 173 (1928).
"See, for example, R. Haefer, Z. Physik, 116, 604 (1940).
228
[Chap. 9
It will be evident from the above discussion that field emission is not
strongly influenced by temperature. Of course, the temperature should
be kept low enough to assure the absence of thermionic emission.
:\
.2
.4
.6
.8
1.0
1.2
-Ii
i
9_9. 26
......
Sec. 9-9]
229
D; - I
+ cP >
D"
11q, =
47Taed = 47TNe 2d
(9-49)
where d is the distance between the positive and negative charges and N
" J. B. Taylor and I. Langmuir, Phys. Rev., 44, 423 (1933); J. A. Becker, Phys. Rev ..
28, 341 (1926).
230
[Chap. 9
') !,
'
Vacuum
):,
(a)
,1 .. '
- 4>1
(9-50)
Sec. 9-10]
231
by the two work functions and is independent of the depths of the potential
energy wells.
In connection with the great importance of the Fermi level in equilibria
between two or more sets of electronic energy levels, let us consider the
problem from the thermodynamic viewpoint. Suppose two sets of energy
levels, distinguished by the subscripts 1 and 2, are in thermal equilibrium
at constant pressure and temperature. This means, according to Appendix
A that the Gibbs thermodynamic potential of the combined systems should
be a minimum; i.e., when one electron is transferred from system 1 to
system 2, the resulting change dG = dG1
dG 2 should vanish. Now,
according to thermodynamics,
T dS
dE
+ p dV -
(9-51)
fl dN
+ p V) =
fl dN
(9-53)
.Equilibrium thus requires that the fl'S of the two systems be the same.
However, from (9-39), (9-40), and (9-52) it follows that fl = E p , showing
that for two (or more) sets of electronic levels, the Fermi levels must be
the same in equilibrium. This conclusion is of importance in the discussion
of contacts between metals, semiconductors, and insulators.
Fig. 9-13.
conta~t
potentials.
232
[Chap. 9
and B be two SL!ch surfaces with different w0rk functions. Thus, without
external voltage. the plates will be charged as explained above. A sudden
change in the distance between the plates (switch S open) will lead to a
voltage pulse resulting from the change in capacitance and can be measured.
]f an external voltage is applied by means of a potentiometer (S closed),
the levels of one metal are raised or lowered relative to those in the other.
For a particular value of the external voltage the charges on the plates
vanish and a change in distance (S open) will not yield a voltage pulse.
Clearly, the external voltage then just compensates thc contact potential
and thus (~.1 ~ ~JI) may be obtained. A method dcvised by Zisman
employs a vibrating plate so that a-c techniques can be used and work
functions can be measured in a matter of seconds. 28 ]n this way one has
measured. for example, the change of work functions with temperature,
with the result that for metals ~ increases with about 10-4 ev per degree.
p <;Po
-+ hv/c
or
(9-55)
+ hv/c
(9-56)
+ (2Eo/111C2)1!2 ;?: 1
(9-57)
29
Sec. 9-11]
233
One thus concludes that in the free electron approximation, the conservation laws cannot be satisfied and thus free electrons cannot absorb photons.
This argument would hold fot the electrons in the interior of the metal.
The reason that there actually exists a volume effect is a consequence of
the fact that the free electron approximation is not valid; even in the case
of the alkali metals, for which this approximation is better than for any
'i
f.... f,'i
-r-l~
Vacuum
I
hv
'-
hv
_J
E
- E'"./e
n(EI
(al
other metals, the volume effect may contribute to the surface effect (b).30
The discussion given below is confined to process (b); because of lack of
space, only some qualitative remarks will be made.
Notwithstanding the arguments given above, the free electron approximation applied to electrons near the surface leads to the possibility of
photon absorption for these electrons; this follows from the wavemechanical treatment of the problem. 31 One might say that the presence
of the potential barrier at the surface makes it possible to satisfy the
conservation laws in the sense that the surface itself acts as a possible
source or sink for momentum. In other words, the system under consideration is no longer the electron plus a photon, but electron plus photon plus
surface. With reference to Fig. 9-14a the following conclusions may then
be drawn for the emission characteristics of a metal at absolute zero.
3<1 H. J. Fan, Phys. Rev., 68, 43 (1945). In the volume effect, the excitation of electrons
is governed by the selection rule that the transition should be "vertical" in the reduced
zone scheme (see Chapter 10).
31 K. Mitchell, Proc. Roy. Soc. (London), A 146,442 (1934); 153,513 (1936); I. Tamm
and S. Schubin, Z. Physik, 68, 97 (1931); A. G. Hill, Phys. Rev., 53,184 (1938); R. E. B.
Makinson, Pllys. Rev., 75, 1908 (1949).
/
234
[Chap. 9
(9-58)
For V < V t , no emission occurs. Evidently the work function of a
metal may be obtained by measuring the threshold frequency V t At
Na
3600
4400
5200
6000
6800
_AinA
I
j
,YO
E,;, =
hv -
cp
(9-59)
Sec. 9-11]
235
Thus for a potential -E~Je applied to the collector, all emitted electrons
are stopped and the collected current I = O. As the collector potential is
made less negative, the collector current increases until for zero collector
potential the current reaches its saturation value Is for the particular
incident frequency. The I versus V('()lkdor curve thus has the. form as
indicated in Fig. 9-14b; this is in agreement with the observations. 33 By
differentiation of such curves, the energy distribution of the emitted
electrons may be obtained.
Quantitatively, the theory may be set up in the following way: let
neE) dE be the number of electrons in the metal occupying states in the
energy range between E and E + dE. Also, let P(v,E) be the probability
that an incident photon of frequency v excites an electron from a state E
into the state E + hv. The number of electrons emitted by the metal
originating from the range E, E
dE is then per incident photon,
n(E)P(v,E)Q(E + hv) dE
(9-60)
F(v)(E - Ep - 4 dE = F(v)E'dE'
(9-61)
Here F(v) is a function of the frequency only, and E' is the energy of the
electron in the excited state relative to the vacuum level. From (9-61) an
expression for the collector current as function of the retarding potential
can be obtained. 33
'.
REFERENCES
L. Brillouin, Die Quantenstatistik, Springer, Berlin, 1933.
236
[Chap. 9
PROBLEMS
9-1. Discuss how one can determine E, in Fig. 9-1 from electron
diffraction experiments.
9-2. Calculate the average velocity of a conduction electron in sodium
at T = O. Compare the corresponding "classical temperature" with the
0
melting point of Na. What is the electronic specific heat at T = 300 K?
9-3. Show that the derivative of the Fermi function is symmetrical
about E]<' and that
J('(F/aE)dE
-1.
-oc
)
Chap. 9]
.
FREE ELECTRON THEORY OF METALS
237
Chapter 10
Introductor~'
remarks
,l "
238
,
Sec. 10-1]
239
references we refer the reader to the article by Reitz cited at the end of this
chapter. The problem as outlined above involves essentially that of the
behavior of an electron in a potential which has the periodicity of crystal
lattice. We shall see that this leads, among other things, to a natural
distinction between metals, insu)ators, and semiconductors.
Before discussing the actual problem it may be useful to point out the
analogy which exists between (i) electronic motion in a constant and a
periodic potential, and (ii) the propagation of elastic waves in a continuum
and in a periodic structure.
For elastic waves in a continuous medium the frequency is inversely
proportional to the wavelength, i.e., there exists a linear relationship
between frequency and wave number (or wave vector). This implies a
velocity of propagation which is independent of the wavelength. Furthermore, there exists no upper limit for the frequency of the vibrational
modes in a continuous medium. However, when one considers the modes
of vibration in a lattice of discrete mass points which form a periodic
structure, two characteristic features appear (see Chapter 2):
I. There exist allowed frequency bands, separated by forbidden
regions.
2. The frequency is no longer proportional to the wave number but a
periodic function of the latter.
Returning now to the motion of electrons, the reader is reminded of the
fact that in a constant potential (free electron theory) the energy of the
;;,
electron as function of the wave vector k is given by
E = 1i2k 2 /2m
where
k = 2771). = pili
Here i. is the wavelength associated with the electron and p is the momentum
of the electron; the potential energy has been assumed zero. In this case,
there is no upper limit to the energy, i.e., the energy spectrum is quasicontinuous (quasi, because the limited dimensions of the potential box
produce closely spaced but discrete energy levels). However, if we
consider the motion of an electron in a periodic potential we arrive at the
following resulTs:
I. There exist allowed energy bands separated by forbidden regions.
2. The functions E(k) are periodic in k.
These results will be derived below. The analogy pointed out above is
not too surprising if one recognizes that in both problems one deals with
waves in periodic structures; in one case they are elastic waves, in the
other they are waves associated with the electrons. For further details
with regard to the general problem of wave motion in periodic structures
we refer the reader to Brillouin. l
1 L. Brillouin, Wave Propl{l{afion in Periodic Structures, Dover, New York, 1953.
The existence of energy bands for electrons in crystals was first pointed out by M. S. O.
Strlltt, Ann. Phvsik, 84, 485 (1927); 85, 12<J (1928).
,i~ ,
240
[Chap. 10
d 21p/dx2
+ (2m/1i2)(E -
Vo)1p = 0
(10-1 )
=eik.t
Upon substitution one obtains for the kinetic energy of the electron
E kin
= E --
Vo
= /i 2k2/2m =
p2/2m
It
vex
-+-
( 10-2)
a)
d 21p/dx2
+ (2m/1i2)[E -
V(x)]1p = 0
( 10-3)
In other words, the solutions are plane waves modulated by the function
uk(x), which has the same periodicity as the lattice. This theorem is known
as the Bloch theorem;2 in the theory of differential equations it is known as
Floquet's theorem. Functions of the type (10-4) are called Bloch functions.
Before giving a proof of this theorem we note that the Bloch function
1p(x) = exp (ikx)uk(x) has the property
1p(x
+ a) =
1p(x
2
exp [ik(x
+ a)] uix + a) =
+ a) =
Q1p(x)
where
= exp (ika)
(10-5)
Sec. 10-2]
241
It will be evident that if we can show that the Schrodinger equation (10-3)
has solutions with the property (10-5), the solutions can be written as
.Bloch functions and the theorem is proved. This will now be done. 3
Suppose g(x) and f(x) are two real independent solutions of the
Schrodinger equation. Now a differential equation of the second order has
only two independent solutiorfs, and all other solutions are expressible as a
lin~ar combination of the independent ones. Then, since j(x
a) and
g(x
a) are also solutions of the Schrodinger equation, we must have the
relations
f(x
+ a) =
ocd(x)
+ oc~(x)
g(x
+ a) =
PI/(x)
+ P~(x)
( 10-6)
where the oc's and P's are real functions of E. The solution of the Schrl"idinger equation may be written in the form
) .
1p(x)
Af(x)
+ Bg(x)
+."
In view of what has been said above about the property (10-5) of the
Bloch functions, let us choose A and B such that
, . , '.. ,,"" ":.IIi'
+ B{JI =
Aoc 2 + BP2 =
AocI
QA
( 10-7)
QB
+ a) =
Q1p(x)
(10-8)
(10-9)
or
Now, one can show that OC I {J2 - OC 2PI = 1 in the following manner:
from equations (10-6) one can derive that
f(X
F(x
I
...
+ a)
-1-
a)
g(x
g'(x
+ a)
+ a)
If(X)
g(x)
f'(x)
g'(x)
lOCI
oc 2 \
If1I f12
( 10-10)
3 See H. A. Kramers, Physica, 2,483 (1935); F. Seitz, The Modem Theory of Solids,
McGraw-Hili, New York, 1940, p. 279; N. F. Mott and H, Jones, Theory of the Properties of Metals and Alloys, Oxford, New York, 1936, p. 57; A. H. Wilson, TheOl~v ol
Metals, 2d ed., Cambridge, London, 1953, p. 21.
242
[Chap. 10
o = Ig" -
gf"
= (d/dx)(fg' - gf')
If'(x)
g(x)
.-~
= constant
g'(x)
This result, together with equation (10-10), leads to the conclusion that
(10-12)
The corresponding functions 1J'l(X) and 1J'2(X) then have the property
and thus are Bloch functions (see 10-5). In other regions of the energy E,
viz., those corresponding to (oc l + 132)2 > 4, the two roots Ql and Q2 are
real and the reciprocals of each other. These roots correspond to
solutions of the Schrodinger equation of the type
"~heX) =
e"l'u(x)
and
1J'2(X)
= e~!'l'u(x)
--------------~
Sec. 10-3]
243
(10-14)
li21p/dx2 + (2m/1i2)(E -
Vo)1p
for
-b
<
<
(10-15)
:x 2
(10-16)
and making use of the fact that the solutions must be Bloch functions of
the form eik.T uk(x), one obtains upon substitution into (10-14) and (10-15)
the following equations for uk(x):
(10-17)
d 2u/dx2 2ik(du/dx)
(oc 2 - P) U = 0
O<x<a
d u/dx2
2
+
+
+ 2ik(du/dx) - (f32 + k 2) U =
-b
<
<
( 10-18)
u1 =
Ae i (oc-k)2'
-t- Be-i(x-l-k)I
Ce(P-ik)x
De-(fJtil')I
O<x<a
(10-19)
-b<x<O
(dUl/dx)x~o
= u 2(O),
(dU2/dx)x~o'
u1(a)
u2(-b)
( 10-20)
, R. de L. Kronig and W. G. Penney, Proc. Roy. Soc. (London), ABO, 499 (1930);
see for an extension of this work D. S. Saxon and R. A. Hutner, Philips Research Repts.,
4,81 (1949); J. M. Luttinger, Philips Research Repts., 6, 303 (1951); G. Allen, Phys.
Rev., 91, 531 (1953). The case Vex) ~- A sin x has been discussed by Morse, Phys. Rev.,
35, 1310 (1930). For another calculable one-dimensional case see J. C. Slater, Phys.
Rev., 87, 807 (1952).
244
[Chap. 10
The two conditions on the left are imposed because of the requirement of
continuity of the wave functions and of their derivatives; the two on the
right are required because of the periodicity of uk(x). It is evident that
application of (10-20) on (10-19) leads to four linear homogeneous
equations in the constants A, B, C, D; thus the wave functions may be
calculated. However, for our purpose we are more interested in determining
the values of the energy for which satisfactory solutions are obtained. The
four equations just mentioned have a solution only if the determinant of
the coefficients of A, B, e, D vanishes. It can be shown that this leads to
the following condition:
p-~
+- cosh ph cos xa =
cos k(a
+ b)
.
(10-21)
Fig. 10-2. The left hand side of (10-24) for P = 31T/2, plotted as
. \
function of IXa. The allowed regions are heavily drawn.
to infinity and b approaches zero, but the product Vob remains finite. Under
these circumstances (10-2 J) reduces to
(m Vob/h 2 x) sin xa
+- cos xa =
cos ka
iL
(10-22)
( 10-23)
which is evidently a measure for the "area" Vob of the potential barrier.
In other words, increasing P has the physical meaning of binding a
given electron more strongly to a particular potential well. From the last
two equations we find that solutions for the wave functions exist only if
sin oca
P -rxa
(10-24)
Sec. 10-3] -
245
equation as function of oca for the value P = 37Tj2. The reader is reminded
that ex2 is proportional to the energy E, i.e., the abcissa is a measure for the
energy. Furthermore, it is important to realize that the right-hand side
can accept only values between - I and I, as indicated by the horizontal
lines in Fig. 10-2. Therefore the condition (10-24) can be satisfied only for
values of exa for which the left-hand side lies between I.
From the figure, the following interesting conclusions may be drawn:
o
P/47r +-
o
-+ 47r/P
n = 1,2,3, ...
;~ola f; hi,'
(10-26)
(Chap. 10
246
These k-values define the boundaries of the first, second, etc. Brillouin
zones. It must be noted that Fig. 1O-4a- gives only half of the complete E(k)
curve; thus the first zone extends from --7T/a to +1T/a. Similarly, the
second zone consists of two parts; one extending from 7T/a to 21T/a, as
shown, and another part extending between --1T/a and -21T/a.
A further important conclusion may be drawn from (10-24):
(e) Within a given energy band, the energy is a periodic function of k.
For example, if one replaces k by k + 27Tn/a, where n is an integer, the
right-hand side of (10-24) remains the same. In other words, k is not
15
10
vi
00
'(
-7r /a
1st
2nd
7r/a - k
3rd
(bl
(al
vector.
-7T/a
~ k ~
7T/a
--
(to-27)
The energy versus reduced wave vector is represented in Fig. 10-4b. It may
be noted here that the fact that k is not uniquely determined also follows
quite generally from the form of the Bloch function (10-8). Consider the
function eik,ruk(x) and introduce a new wave vector V - k
21Tn/a,
where n is an integer. One may then write
(10-28)
It will be noted that uk.(x) is also periodic with the lattice so that (10-28)
is just as good a Bloch function as the initial function eik.ruk(x).
Sec. 10-3]
247
The number ofpossible wave functions per band. So far, we have assumed
the crystal to be infinite, but it will now be necessary to investigate the
consequences of imposing boundary conditions. Since we have employed
the running wave picture, it will be convenient to use cyclic or periodic
boundary conditions (see Sec. 2-9 for the same problem in the theory of
elastic waves in a chain of atoms). For a linear crystal of length L the
boundary condition may be taken as
tp(X
+ L) =
tp(x)
(10-29)
k = 2-rrn1L
with
+ L) =
n = l, 2, ...
(10-30)
See Appendix C.
248
[Chap. 10
v = dw/dk
\
Here w is the angular freq uency of the de Broglie waves; it is related to the
energy of the particle by the relation E = nu). Thus instead of (10-32),
one may write in general for the velocity of the particle,
D
= n-l(dE/dk)
(10-33)
(a)
(h)
(e)
_k
Fig. 10-5. Energy, velocity, effective mass and
.I~ as function of k. The
dashed lines correspond to
the inflection points in the
E(k) curve.
(10-34)
See, for example, M. Born, Atolllic PhYSics, 5th ed., Hafner, New York, 1951.
, To avoid confusion with the energy, the electric field will be represented by F.
Sec. 10-4]
249
(10-35)
eF/1l
,) ..,..'
F-
_,_
= (eF/1l2)(d 2E/dk2)
(10-37)
\.
250
[Chap. 10
where grad" grad" E(k) is a tensor with nine components of the general
form eJ2E/ok; ok; with i, j = x, y, Z.9
10-5. The distinction between metals, insulators, and intrinsic semiconductors
up
(10-40)
I
where the summation extends over all occupied states in the band. Now,
according to (10-31) the number of states in an interval dk (excluding the
spin) for a one-dimensional lattice of length L is equal to L dlt/27T.
Because two electrons occupy each of these states in the shaded region of
Fig. 10-6, one may write instead of (10-40),
where we used (10-39). Thus the effective number of electrons in the band is
(10-41)
9
,"
Sec. 10-5]
251
?Z4V2ZZZZT/2l2Z/
Insulator
(al
Semiconductor
(bl
i:
Metal
'\
(el
Fig. 10-7.
252
[Chap. 10
1= -e ~
Vi
= -e [Vj
+ '*)
.2:. Vi] =
(10-42)
,.
= -":_e '"
.~.
I rJ
vt
= ev
)
(10-43)
Applying an external field F, the rate of change of the current l' due to the
field is
(10-44)
dl'ldt = e(dvjldt} = -e2 Flmj
Now, since holes tend to reside in the upper part of a nearly filled band,
mj is negative and the right-hand side of (10-44) becomes positive. In
other words, a band in which an electron is plissing behaves as a "positive
hole" with an effective mass
This concept is of great importance in
the theory of conductivity and Hall effect as 'we shall see later. It explains,
for example, why certain materials show a positive rather than a negative
Hall coefficient (free electrons give a negative Hall coefficient}V
Iml.
Sec. 10-7]
253
Vectors such as d are called direct lattice vectors; the adjective "direct" is
included to distinguish such vectors from the "reciprocal" lattice vectors
to be introduced below. In order to di,scuss the behavior of an electron in
a periodic potential it will be convenient to consider first how one represents
periodic functions such. as (10-46) in terms of three-dimensional Fourier
series. For a one-dimensional periodic potential which satisfies the
condition
Vex) = Vex + d1a) where d1 = integer
one may of course always write
Vex)
= L
"
Vg exp (27Tigx/a)
g = integer
(10-47)
where the summation extends over all integers from --00 to +00; the
coefficients Vq are the Fourier coefficients. That this series indeed satisfies
the periodicity requirement may readily be shown as follows. Replacing x
in (10-47) by x + d1a, where d1 is an integer, we obtain
i
[Chap. 10
254
represents a potential with the periodicity (l0-46) in terms of a threedimensional Fourier series. It can be done rather easily, however, if we
introduce the so-called reciprocal lattice. The reciprocal lattice is defined
by three primitive translations b I , b 2 , b 3 which satisfy the conditions
bj
ij
I
0
{
if
if
i = j
i =1= j
(l0-50)
Thus the vector bi is perpendicular to the plane through the direct lattice
vectors 02 and 03' The explicit expressions for the b's are evidently of the
form
(l0-51)
from which the absolute magnitudes of the b's may be obtained in terms
of the primitive translations of the direct lattice. Any vector
(l0-52)
is called a reciprocal lattice vector. The end points of these vectors
define the reciprocal lattice points. The reader may show himself that the
reciprocal lattice of an f.c.c. lattice is b.c.c. and vice versa. IO
We shall now show that the three-dimensional Fourier series,
V(r)
(l0-53)
Vn exp (27Tin r)
0
+ d) =
~
n
Vn exp [27Ti(rno r
+ nod)]
+
10 For other properties of the reciprocal lattice, see, for example, Brillouin's book
quoted at the end of this chapter.
11 Some authors define the reciprocal lattice by means of the relations a i
bi = 27T15ij;
with this definition the factor 27T in the exponential of (10-53) is absent.
0
Sec. 10-7]
255
10-2, may be extended to three dimensions. The result is that the wave
:t!,! ,,1
functions are, in analogy with (10-4), of the type )Ci
.\
(10-54)
where uk(r) has the periodicity of the lattice. Hence, in general, we may
write
(10-55)
uk(r) = ~ en exp (27Tin r)
n
where n is a vector in the reciprocal lattice. In analogy with what has been
said in connection with equation (10-28), one can show that any two Bloch
functions for which the wave vectors differ by 27T times a reciprocal lattice
vector are physically equivalent. For example, let n be a reciprocal lattice
vector and let us introduce instead of k another wave vector k' = k
27Tn
in (10.54). We may then write
'I/'(r)
e:1:ik"'e T2 "in"Uk(r)
eik"rUk (r)
-7Tb 2
~ k2 ~
7Tb 2
-7Tb 3
7Tb 3
ka
(10-56)
In this case we refer to k as the reduced wave vector; the region of k-space
defined by (10-56) is referred to as the first Brillouin zone or reduced zone.
As in the Kronig-Penney model, a given reduced wave vector k
corresponds to a set of energy values E1(k), E 2(k), ... , where the subscripts
refer to a particular energy band. Within each energy band the k-values
are restricted in accordance with (10-56). We shall now show that for a
finite crystal the number of possible reduced k-values within a single energy
band is equal to the number of unit cells contained in the crystal. This
statement is the analogue of conclusion (f) in Sec. 10-3 for the KronigPenney model. Consider a crystal in the form of a parallelepiped with
edges N10 1 , N 20 2 , N 30 3 , where N 1 , N 2 , N3 are large integers. Employing
cyclic boundary conditions (compare 10-29), the wave functions should
satisfy the condition
( 10-57)
256
[Chap. 10
Since lp(r) is a Bloch function of the type (10-54), for which uir) is periodic
with the lattice, this condition is equivalent with the requirement
k . (NIa l
+N
2a 2
+N a
3 3)
2rr[(n1/NI )b1
+ (n
2 /N 2 )b 2
(10-58)
+ (n3/N3)b31
(10-59)
1 ."
dns
(2/8rr3 ) dO,.
(10-60)
The factor 2 arises from the spin. The quantity dns is referred to as the
density of states corresponding to the Hement dO" in k-space. In
subsequent discussions it will frequently be desirable to introduce the
number of states per unit volume of the crystal per unit energy interval.
Thus, consider in the k-space two surfaces of constant energy, one of E,
the other of E + dE. The volume element dD,. in k-space corresponding
to a differential area dS and bounded by the constant energy surfaces, is
then given by
dO,. = dS[lgrad,. E(k)IJ-I dE
:rH
so that the density of states per unit energy interval is given by
dns/dE
(2/81T
3
)
dS
JIgrad"E(k) I
(10-61)
where the integral extends over the whole area of the constant energy planes.
Sec. 10-8]
257
..
~.
-(/i 2 j2m)V2cp
+ Va(r)rfo =
Eo
We 'shall assume that the level is nondegenerate, i.e., there is only one
wave function corresponding to Eo. Furthermore, we shall assume that
the wave functions are normalized. Suppose then that similar atoms are
brought together in the form of a crystal. The potential energy of the
electron in the crystal then looks like the dashed curve in Fig. 10-8; the
potential energy in this case will be represented by VCr), where V(r) has the
periodicity of the lattice. Taking a particular atom as the origin of our
coordinate system, the position of any atom may then be represented by a
vector R; where R; is a lattice vector. In the tightly bound electron
approximation it is assumed that the electron in the vicinity of a particu!.ar
nucleus j is only slightly influenced by the presence of other atoms, i.e.,
when the end point of the vector r lies in the vicinity of R j , the wave
function for the electron is approximately given by cp(r -- R j ) and the
258
[Chap. 10
energy of the electron is still very close to the value Eo in the free atom.
Consequently, one calculates the energy of an electron with a wave vector k
in the crystal on the basis of a linear combination of the form
(10-63)
since this expression satisfies the approximation just mentioned: if
r lies close to R j all contributions in the sum will be small except that from
cjJ(r - R;). However, since we are dealing with an electron in a periodic
field, the wave function must be a Bloch function, and this restricts the
choice of the coefficients C j lfin expression (10-63) we take the coefficient
clk) equal to exp (ik . R j ) we obtain
(10-64)
which indeed has the properties of a Bloch function. This can be seen by
applying a transformation corresponding to a lattice vector, say R "'.
This gives
"Pk(r
Rm) =},: eik'RjcjJ [r - (R j - - Rm)]
.i
= i k ' Rm },: eik'(Rj-Rm)cjJ[r -- (R - R",)]
.J
The sum in the last expression, however, is equal to "Pk(r), so that (10-64)
satisfies the characteristic property of a Bloch wave. We shall now calculate
the energy of an electron with wave vector k in the crystal, based on the
wave function (10-64). This can be done by starting from the expression
(10-65)
where :Yt' is the Hamilton operator for an electron in the crystal; the
denominator takes care of the proper normalization of the Bloch functions.
The denominator becomes
I
Now cjJ(r - Rm} has appreciable y.alue only when the end point of the
vector r lies in the vicinity of atom m; similarly, cjJ(r - R;) has appreciable
value only in the vicinity of atom j. In other words, there is very little
overlap between the wave functions, even for nearest neighbors. To a
first approximation, therefore, we shall neglect all overlapping, so that of
the summation over j only the term j = m will be retained. Since we have
assumed that the atomic wave functions were normalized, we may then
write
(10-66)
J"Pk"Pk d-r = .2 J cjJ*(r - Rm)cjJ(r -. Rm) d-r = N
In
where N is the total number of atoms in the crystal. Let us now consider
Sec. 10-8]
259
+ VCr) =
+ Vir -
+ VCr) - Va(r -- R j )
-(/i2j2m)V2 + V'(r - R j ) + V" (r -
-Wj2m)V2
R j) =
Rj)
(10-67)
V'(r -
R;)
Va (r -
VCr) -
( 10-68)
R;)
The reason for this will become obvious below. The physical meaning
of V'(r - R j ) is that it represents the potential energy of the electron in the
crystal at the point r, minus the potential energy of the electron in the same
point if there were only a single atom, viz., the one located at R j In other
words, V'(r - R j ) represents the potential energy of the electron in point r
resulting from the presence of all atoms excl!pt the one located at R j . It is,
in a sense, a perturbation potential. According to Fig. 10-8, V'(r - R j )
is a negative quantity. Substituting the Hamilton operator (10-67) into
(10-65), making use of (10-66), and realizing that
-Wj2m)V2cp(r -
R j)
Va(r -
Rj)cp(r -
R j)
Eo q,(r - Rj )
L L eik'(Rj-Rm)S cp*(r j
Rm)[EO
+ V'(r -
Rj)]cp(r -
njl,!
Rj)
dT
'In
First consider the term containing Eo. Since the overlapping is small
anyway, we may neglect in the summation over m all terms except m = j.
Thus the term containing Eo' becomes
" ,
(ljN)
L Scp*(r -
Rj)Eocp(r -
R j)
dT
Eo
Ul:ju,r.J.
IX
-Jcp*(r -
-S cp* (r -
Rj)V'(r -
R;) cp(r -
Rm)V'(r -
R j)
Rj)cp(r -
dT
R j)
dT
(10-69)
(10-70)
Eo -
IX --
Y~
eik'(Rj-R m )
(10-71)
260
[Chap. 10
observed that the energy of the electron in the crystal differs from the
energy of the electron in the free atom by a constant factor ex: plus a term
which depends on the wave vector k. It is this last part which transforms
the discrete atomic level into an energy band in the solid. In order to see
this more clearly, we shall apply this result to the case of cubic crystals in
the next section.
Another important approximation is the'so-called nearly free electron
approximation. In this case it is assumed that the Fourier coefficients of
the periodic potential are small relative to the constant potential. This
approximation may therefore be expected to be applicable to the conduction
electrons in monovalent metals. The energy versus wave vector curves
obtained for a one-dimensional lattice on the basis of this approximation
resemble closely those given in Fig. 10-4 for the Kronig-Penney model.
_For a discussion of the nearly free electron approximation as well as
other approximations we refer the reader to the literature.
10-9. Application to a simple cubic lattice
In order to appreciate the consequences of the results obtained in the
preceding section we shall first apply expression (10-71) to a simple cubic
lattice. In this lattice a given atom has six nearest neighbors, located such
that
R j - Rm = (a, 0, 0); (0, a, 0); (O,O,a)
Evaluation of the sum in (10-71) then yields for the energy of an s electron
in the crystal
E(k)
= Eo -
ex: -
2)'(cos krra
(10-72)
Eo -
ex: -
61'
From the definition (10-70) of I' it follows that the width of the band
Sec. 10-9]
261
'""---~"
,~i-,_I;
-.2
f
-.4
-.6
_r
12
10
t::::::'.
Eo -
0( -
6y
+ ya 2k 2
for small k
(10-73)
262
[Chap. 10
one atom to another; they have a high effective mass, and the acceleration
produced by an electric field will be relatively small.
The top of the band corresponds to cos kxa = cos k lla = cos kza = -1,
i.e., to kx' k ll , k z = TT/a. Thus the corner points of the reduced zone
correspond to states at the top of the band. In the vicinity of such a
corner point we may expand the cosines again; for example, if we expand
about the point kx = ky = k z = TT/a we may write cos kxa = COS(TT - k~a)
where the new component k: = TTla - k" is measured relative to the
ky
(a)
(h)
Thus, relative to the top of the band, the electron energy is proportional
to k' 2, where the new wave vector k' is measured from the corner point of
the Brillouin zone. Constant-energy surfaces in this region are therefore
again spherical, but with the corner point as center . .ijy way ofiIlustration
we give in Fig. 10-JOa a schematic representation of constant-energy
surfaces in k-space for a two-dimensional square lattice, based on the
tightly bound electron approximation. In the nearly free electron approximation, the proportionality with k2 of the energy relative to the bottom of
the band extends to much larger values of the wave vector than in the
tightly bound approximation. Here again, however, constant energy
surfaces near the top of the band are spherical relative to corner points of
the reduced zone. This is illustrated for comparison in Fig. 10-lOb.
I t is left as a problem for the reader to discuss in a similar manner the
application of the tightly bound electron approximation to body-centered
and face-centered cubic lattices.
Sec. 10-10]
263
(10-76)
where n,r and n. are integers. The first Brillouin zone is enclosed by the
264
[Chap. 10
+ nyky + nzkz =
7T(n;
+ n; + n;)/a
(10-77)
(10-78)
At the zone boundaries, the energy exhibits a discontinuity as in the
one-dimensional case (see Fig. 10-4). It is of interest to note that the values
of the wave vector satisfying (10-78) are those for which the electron
suffers a Bragg reflection.13 We leave it as a problem to show that this is
the case. An electron which satisfies the Bragg condition cannot penetrate
the lattice, since it suffers reflections. Such an electron therefore does not
correspond to a wave propagating through the crystal, but to a standing
wave. The energy discontinuities or energy gaps occurring at the Brillouin
zone boundaries represent the energy ranges for which it is impossible
for an electron to move through the crystal. This is clearly borne out by the
fact that if such electrons are incident on the crystal from the outside,
they are totally reflected and unable to penetrate into the crystal. For
the structure of Brillouin zones for various crystal structures we refer the
reader to the literature. 14
The denSity of states as function of energy. The number of electronic
states per unit volume associated with a volume element dn k in the
k-space is, according to (10-60), equal to (2/87T3) dn k The density of
states per unit energy interval is given by the general expression (10-61).
Let us now consider the consequences of this for a si91ple cubic lattice.
In order to simplify the problem, let us assume that tne constant-energy
surfaces are spheres or parts of spheres around the center of the first
Brillouin zone. This situation is approached in the nearly free electron
approximation (see Fig. 1O-lOb), although even there it does not hold
in the vicinity of the corner points of the zone. With this assumption we
have, as for free electrons, E(k) = /i 2k2/2m*. Since the density of states
correspoflding to wave vectors for which the absolute magnitude lies
between k and k + dk is given by (2/87T 3 )47Tk2 dk, we obtain for the
density of states Z(E) dE, corresponding to an energy interval dE
Z(E)dE=CE 1 / 2 dE
with
C=47T(2m*/h2)3/2
13
(10-79)
1...../(./-1
Sec. 10-10]
265
Hence Z(E) increases as El/2; also note that as the effective mass increases,
Z(E) increases. For narrow energy bands, therefore, Z(E) rises more
rapidly than for broad bands. For the example under consideration,
expression (10-79) will hold up to values of the wave vector equal to
k = Tria because for this k-value the spherical constant energy surface
just touches the Brillouin zone boundary. For larger values of k and E,
only the corners of the cube are available for electronic states, and (10-79)
can no longer be used. In fact, for k = (Trla)v3 the density of states
becomes zero. One thus obtains a Z(E) curve as represented schematically
in Fig. 10-12; the energy E1 corresponds to k = Tria.
Z(El
Z(E)
266
[Chap. 10
E B < Ec
the first Brillouin zone are lower
This is likely to be the case when
In the second case, however, the
(ER ) lies below the top of the llrst
_
'
.;.
~ mi. t,~
~t
.,
'.'
Z{E)
band (Ec). Thus the two bands overlap to some extent, and this may
possibly happen when t::..E is relatively small. It is instructive to consider
the consequences of this type of overlapping by filling up the available
states with electrons. Suppose we use twice as many electrons as there are
unit cells in the crystal; this number would just completely fill a band in
the absence of overlapping. With overlapping, the electron distribution
in the two-dimensional case would look as indicated in Fig. 10-14. The
first zone is partly empty, the second zone is partly filled, because there
are energy states available in the latter which lie below those at the top of
the first zone. It will be evident that under these circumstances conduction
becomes possible and the solid may behave as a metal, be it a "poor" one.
In Fig. 10-15 we have represented schematically the density of states
when overlap occurs.
267
Sec. 10-11]
ourselves to a few general remarks. 15 First of all, for not too complicated
structures such as the f.c.c., b.c.c., and hexagonal lattice, it is always
possible to choose the unit cell in such a fashion that there is one atom per
unit cell. For example, in the f.c.c. lattice one may use as translational
vectors those joining a given corner atom with atoms at the center of three
Or----,-12-r------------------------------.
3d band
~b!~d
/,
-.1
-.2
11-
il
'
'"
-.3
-.4
eu
Ni
-"i<
'<l
-.5
---- -6- -----.6
-----4--------
o __
--+- Z(EI
iiJ.Ii.Hl
faces (see Fig. 1-4a). Under these circumstances, each band can accommodate twice as many electrons as there are atoms in the lattice. It then
foHows that electronic shells which are filled in the atom will lead to
completely filled bands in the solid state (at least if T = 0). It is therefore
not difficult to understand that monovalent elements such as the alkalis,
Cu, Ag, Au are metallic because they contain a half-filled band. In the
1>
For a review, see F. V. Rayn?r, Repts. Progr. Phys., 15, 173 (1952).
268
[Chap. 10
divalent metals such as Ca, Ba, Sr, etc. there is evidently overlapping
between the energy bands associated with the valence electrons.
The zone structure of the transition elements is of considerable interest.
For example, the elements of the iron group have an incompletely filled
3d shell in the atomic state. As the atoms are brought together, the
3d level gives rise to a relatively narrow band; the 4s level broadens
much more strongly, as indicated in Fig. 10-16.1 6
As a consequence, both the 4s and 3d bands
are partly filled with electrons in these metals;
in copper the 3d band is completely filled. (The
3d band can accommodate 10 electrons per
atom because it consists actually of five
completely overlapping bands; the 4s band
contains at most two electrons.) The importance of this type of structure for the
magnetic properties will be discussed in
Fig. 10-17. Illustrating the
Chapter 19. The electronic specific heat of the
process of X-ray emission
transition. metals is abnormally high. This is a
by a metal after ionization
of K or L levels. For the consequence of the fact that the effective mass
transitions indicated, the of the 3d electrons is very high (narrow band
width of the emitted energy width). For the same reason, the 3d electrons
spectrum is equal to the
show a high paramagnetic susceptibility and a
width of the occupied
low efficiency for conducting electric current.
region in the conduction
Thus the conductivity of the transition metals
band.
is determined essentially by the 4s electrons.
10-12. The density of states and soft X-ray emission spectra
It may be mentioned here that information about the density of states
and band width may be obtained from studies of the soft X-ray emission
spectra. For example, if one ionizes the relatively sharp K or L levels in a
solid by bombardment with fast electrons, electrons from higher bands will
make transitions to the vacated levels, with emission of X-fays.. It is evident
from Fig. 10-17 that the spectrum of the emitted radiation provides information about the energy distribution of the electrons in the higher energy
bands. l ? Thus it is possible to determine the bandwidth of the upper bands,
atleastso far as they are occupied by electrons. One has found, for example,
that the conduction electrons in AI cover a range of '"'-'12 ev, in Li '"'-'4.2 ev,
and in Na '"'-'3.0 ev. This method may of course also be used to determine
16 N. F. Mott, Proc. Phys. Soc. (London), 47, 571 (1935); 49,258 (1937); 62,416
(1949).
17 For a review see, for example, H. W. B. Skinner, Repts. Prog. Phys., 5, 257 (1939);
Trans. Roy. Soc. (London), A239, 95 (1940). For recent work in this field see E. M.
Gyorgy and G. G. Harvey, Phys. Rev., 87, 861 (1952); 93,365 (1954).
Sec. 10-12]
269
the bandwidth of the upper filled band in insulators. The exact shape of the
emission spectrum also depends on the transition probabilities.
10-13. The Wigner-Seitz approximation and the cohesive energy of metals
In view of its importance, a few words may be said here about the
Wigner-Seitz approximation, which is based on the following physical
mode}.l8 Imagine a number of straight lines joining the nucleus of a
particular atom in a metal with those of its
nearest and next nearest neighbors. A set of
planes bisecting these lines perpendicularly then
defines what is known as an atomic polyhedron.
An example is given in Fig. 10-18 for a bodycentered cubic lattice. These polyhedra evidently
fill the whole space occupied by the crystal.
Confining ourselves to monovalent metals, each
of the polyhedra contains a singly charged
positive ion; one of the aims of this approximation is to obtain information about the Fig. 10-18. Atomic polybehavior of the valence electrons in the field of hedron for a body-centered
these ions. Near the center of a polyhedron,
cubic lattice.
the potential will be spherically symmetric;
in the vicinity of the boundaries of the polyhedron the field will be small.
In the Wigner-Seitz approximation it is assumed that the field is spherically
symmetric inside the whole polyhedron; also, the field is assumed to be
that of the singly charged positive ion at the center.
Consider now the wave function for an electron in the state k = O.
Then, because the wave function must be of the Bloch type, it follows that
1p = uk(r), i.e., the wave function itself must be periodic with the lattice.
One may thus require that on the boundary of the polyhedron o1p/on = 0,
where %n stands for differentiation normal to the surface of the polyhedron. For simplicity, Wigner and Seitz approximate the polyhedron
by a sphere of radius r0 such that (47T/3)r~ equals the volume of a polyhedron
and then use as a b<;lUndary condition,
(o1p/or)r=,. o = 0
(10-80)
2m
+ h2 [E -
V(r)] 1p = 0
(10-81)
for the boundary condition (10-80). Note that because V(r) represents the
18 E. Wigner and F. Seitz, Phys. Rev., 43, 804 (1933); 46,509 (1934). See also J. C.
Slater, Phys. Rev., 45, 794 (1934).
270
[Chap. 10
'/I
t
-.4
~.
-.8
2
-+r
0.39
Na
0.51
K
0.58
Cu
0.75
Ag
Au
0.88
0.95
271
Sec. 10-13]
~-----
.4
i,';
.2
.'"
"
I,;
,',;
0
.:.
-.2
"
{: --
,.
-.4
-.6
o
Fig. 10-20.
deal of progress has been made towards solving them for simple metals.
We shall discuss here a simplified theory of the cohesive energy of metals
based on the Wigner-Seitz approximation. In general, the total potential
energy of the metal is determined by the interaction of the charges within
a given polyhedron plus the interaction of the polyhedra with each other.
Suppose now that the valence electrons are distributed such that each polyhedron contains one electron. In that case the polyhedra are neutral, and
to a first approximation the interaction between them may be neglected.
Doing this, the total energy of the crystal is then given simply by the sum of
of the kinetic energy of the electrons plus the potential energy of each
electron in the field of a positive ion. Now the latter quantity is given by Eo,
represented as function of '0 in Fig. 10-10. The kinetic energy of the
electrons may be obtained to a first approximation by assuming a free
electron model for the valence electrons, which for the alkali metals is
quite good as we have seen above. In the preceding chapter we have seen
that the average energy of suc~ a system is equal to ~EF' Now, according
272
",
[Chap. to
:0
(1i2/mr~)(97T/4)2/3
(10-82)
Erohesive = -(Eo
iEF
E]) with '0 = (rO)min
(10-83)
Here Eo is the only negative quantity; EF and E] are both positive, and
as they increase the binding becomes less strong. The abpve model is, of
course, too simple and a number of corrections are required. For example
it is estimated that the Coulomb energy between the valence electrons
gives a term 0.6e 2/ro; also, account must be taken of the fact that the
electrons tend to keep away from each other, an effect which depends on
the relative spin orientations of the electrons involved. Furthermore,
there are van der Waals forces between the ions. Although it is evident
that the problem is a very complicated one, it may be of interest to indicate
the extent to which theory and experiment agree; the following comparisons are from Seitz.20
Lattice spacing
(A)
Metal
Li
Na
K
Calc.
3.50
4.51
5.82
Obs.
3.46
4.25
5.20
Sublimation energy
(kcaIJmole)
Calc.
Obs.
39
36.2
24.S
26
23
16.S
In these figures, the.minimum in the total energy versus ro curve was used
to define the theoretical lattice spacing, and the cohesive energy was
calculated for this particular value of roo Calculations of the
compressibility are also .in reasonable agreement with' experiment. For
Na the observed \lnd calculated values are, respectively, 12.3 and 12.0 X
10-12 cm 2/dyne. 21
Attempts have also been made to explain the crystal structure of
metals in terms of the electronic structure; the differences in energy
obtained for different crystal structures are in general too small to draw
unique conclusions. For certain alloy structures, however, Jones has
been able to account for structural transitions associated with particular
compositions on the basis of the band theory.22
20 F. Seitz, op. cit., p.365.
J. Bardeen, J. Chem. Phys., 6, 367, 372 (1938).
H. Jones, Proc. Roy. Soc. (London), AI44, 225 (1934); Proc. Phys. Soc. (London),
49,243 (1937); Physica, 15, 13 (1949); Phil. Mag., 41, 663 (1950).
21
22
Chap. 10]
273
REFERENCES
An elementary treatment of the band theory may be found in:
L. Brillouin, Wave Propagation in Periodic Structures; Electric Filters
and Crystal Lattices, 2d ed., Dover, New York, 1953.
A. H. Cottrell, Theoretical Structural Metallurgy, Arnold, London, 1948.
PROBLEMS
10-2. Consider an f.c.c. lattice with a cube edge a. Show that the
reciprocal lattice is b.c.c. with an edge 2/a. Also show that the reciprocal
lattice of a b.c.c. lattice is f.~.c.
'
274
[Chap. 10
10-3. Show that the volumes of a unit cell in the direct and reciprocal
lattices are the reciprocal of each other.
10-4. Suppose a beam of monochromatic X-rays is reflected by a
crystal, i.e., the beam satisfies the Bragg condition. Let So and s be unit
vectors in the direction of the incident and reflected beams. Show that
the Bragg condition is equivalent with the requirement that (s -- so)/).
must correspond to a vector in the reciprocal lattice ; J.. is the wavelength
of the X-rays.
]0-5. In Sec. ]0-10 we concluded that the band theory for cubic
crystals leads to discontinuities in the E(k) surfaces whenever k satisfies
the condition n k = 7TnZja (see expression lO-78). Show that this
condition is equivalent with that for Bragg reflection of the electrons by
the set of planes with Miller indices n1 , n z, n 3 .
10-6. Show that in the tightly bound electron approximation the
energy E(k) for b.c.c. and f.c.c. lattices are given by
E(k) = Eo - oc - 8y cos kxa cos kya cos kza
(b.c.c.)
(f.c.c.)
where 2a is the cube edge. Show also that for small values of Ikl the energy
varies proportionally with ikj2. Discuss the shape of constant energy
surfaces in k-space.
10-7. Calculate the width of the energy region occupied by electrons
in the conduction bands of Li, Na, and Al on the basis of the free electron
theory of metals, assuming that each atom contributes as many electrons
as its chemical valence. With reference to the bandwidths quoted in Sec.
10-12, what average effective mass would one have to assume in order
to obtain agreement?
10-8. Discuss the nearly free electron approximation for a onedimensional lattice.
---::-:-:--:---
..
Chapter 11
--',-
+ p(T)
(11-1)
276
[Chap. II
CONDUCTIVITY OF METALS
(11-2)
CONDUCTIVITY OF METALS
Sec. 11-2]
277
From the last three equations it then follows that the average drift velocity
in the field direction is given by
(11-7)
From (11-3) and (11-7) it then follows that the conductivity is given by
(11-8)
Suppose that under influence of an electric field Ex the electrons have
a certain average drift velocity and that at the instant t = 0 the field is
suddenly switched off. As a result of the collisions with the lattice the
average drift velocity will gradually approach zero; since the rate of
change of (l'x> by collisions alone is given by (1l-5), the decay will follow
the expression
(11-9)
where (vx(O) is the average drift velocity at t = O. Because of the
exponential form of (11-9), the quantity T is called the relaxation time.
We may note here already that with n ~ 1022 cm- 3 , expression (11-8)
requires T'_' 10-14 second in order to obtain agreement with experimental
room temperature data (see point 2, Sec. 11-1).
For the particular type of collisions postulated above, T also represents
~
,{'.l~:tJ'
..... "
278
CONDUCTIVITY OF METALS
[Chap. I I
the mean free time between collisions. This may be shown as follows:
Let pet) be the probability that t seconds after a certain collision has"
occurred, an electron has not yet collided again; P(t
dt) represents the
same quantity after (t + dt) seconds. Then
pet
+ dt) =
pet)
+ (dP/dt) dt
+ dt) =
since P
1 for
=
I' ,.
1=
f;-I/T
(11-10)
It must be emphasized, however, that the relaxation time and the mean
free time between collisions are identical only if the velocity after collision
is random. For example, if the scattering is not isotropic and Te is the
mean free time between collisions, the relaxation time can readily be
shown to be
(11-11)
T = Te/(I (cosf3
where (cos 13) is the average of the cosine of the scattering angle. 2 Thus
when nearly all collisions involve small angles, the electron has a rather
strong "memory" and it takes a relatively large number of collisions to
erase this memory, i.e., T ~ To in that case.
I
11-3. The Boltzmann transport equation
Sec. 11-3]
279
CONDUCTIVITY OF METALS
_J
Px
x
(Jt;
+ Px olJm;
+ Yot;
y + jJy otJm;
py
P. -/- Z
z
___
01
+ P. otJm
(11-14)
(11-15)
Since (11-12) and (11-15) must be equal, one is led to the following result,
obtained by expanding (11-15),
oj
oj
of
oj
oj
oj
(ojJOt)IlPhb= - - X--- Y---Z--VX--vy--V Z (11-16)
opx
OPll
opz
ax
oy
2z
where vx ' cu' v. represent the velocity components. In the steady state
there must be other processes which just balance the rate of change (11-16)
produced by fields and gradients. As we n~ted already in the preceding
section, such processes are provided by electron-lattice interactions. Thus
condition (11-13) may be written in the form
(ojJOt)flelclR
+ (ojIOI)('(l1I =
( 11-17)
where the first term is given by (11-16) and where the second term refers
to electron-lattice scattering (compare 11-6). Since the force exerted on
280
CONDUCTIVITY OF METALS
[Chap. II
F=
-e( E + ~ ., X H)
where., is the velocity vector, we may write ( 11-17) com\ined with (11-16)
in the general form
(o//Ot)('oll
---e(,E + ~c ., X H) .grad
p/
-+
(11-18)
grad r /
,,
The net difference between the above quantities integrated over dp~ dp~ dp~
determines (o//ot)ron, i.e.,
(o//ot),'OI! =
JH [f(p',r,t)P(p',p,r) -
(o"lot)
'JI
coil
-r(p,r)
(11-20)
...
Sec. 11-3]
CONDUCTIVITY OF METALS
281
>;'''....
282
CONDUCTIVITY OF METALS
[Chap. 11
fields or gradients being absent. When in the state of steady current the
average number of electrons per unit volume in the range dp", dpy dpz is
represented by
\
(11-22)
we may write immediately for the current density,
Ix
= -
2e
h3
(11-23)
such that
(11-25)
(11-26)
where the last approximation is valid for small fields so that (F - Fo) is
relatively small (physically speaking this assumption is equivalent with a
linear dependence of I on E). Making use of the fact that the energy of
the electrons is given by E = (p~
p; p;)/2m, one may write instead
of 01-26),
(11-27)
+ +
Substituting (11-27) into (11-23), one obtains for the current density,
(11-28)
We shall assume that T is a function only of the energy and not of the
direction of motion (compare 11-20). Since OFO/OE is also a function of E
alone, one may transform (11-28) into a single integral by replacing
by
v2/3 and dpx dpv dpz by 47Tp2 dp. Expressing the integrand in terms of E,
one obtains
v;
(11-29)
Now we have seen in Sec. 9-3 that OFO/OE has an appreciable value only
in an energy range of a few kT about the Fermi level EF. To a good
Sec. 11-4]
283
CONDUCTIVITY OF METALS
approximation 1' 3/27(1') under the integral sign may thus be replaced by
the quantity E:W7F in front of the integral. Furthermore,
fo'" (oF/oE) dE =
and if one substitutes
simple result,
l
EO F
-I
:,' ,,~,
(:
(11-30)
where n is the number of electrons per unit volume. It is interesting to
note that although all electrons take part in the conduction mechanism
only the relaxation time of the electrons at the Fermi level occurs in
the conductivity. The reason for this
may be explained with reference
to Fig. II-I. The full circle represents the Fermi distribution for a
Vx
two-dimensional case in the absence
of an external field. In the presence
of a field along the x-direction, the
veloCity of all electrons is shifted
by an amount ~v (the average drift
velocity), leading to the dashed circle. Fig. 11-1. Exaggerated representation
It is evident that the distribution is of the influence of an electric field on the
changed only in the vicinity of the velocity distribution for a two-dimensional crystal. The fully drawn circle
Fermi level, so that only the relaxacorresponds to the Fermi distribution in
tion time of electrons near EF is of the absence of a field; the field E;,
importance.
produces a shift AD opposite to the field
direction (dashed curve).
Note that (11-30) is essentially the
same as (11-8), except that 7 has
been replaced by 7F' Although the treatment given here was based on the
free electron approximation, a similar treatment may be given for the
band approximation. 6 The result of such a calculation is
a
n('ffe27F/m
(11-31)
284
[Chap. II
CONDUCTIVrTY OF METALS
what has been p;aid in Sec. 11-3. So far, however, we have not paid any
attention to til.: actual cause of resistivity, i.e., to the physical mechanism
which determines TF' On the other hand, it follows from the basic formula
(11-30) that features such as the temperature dependence, pressure
dependence, etc. must be hidden in the quantity TF'
Let us assume that the scattering of the electrons is isotropic; from
the discussion given at the end of Sec. 11-2 it then follows that we may
Introduce a mean free path AF between collisions for electrons at the
Fermi level by means of the relation
!I" "en".
(11-32)
where I'F is the velocity of an electron with the Fermi energy. Hence
(11-30) may then be written
(11-33)
From experimental values of (} and from a knowledge of the Fermi level
(which is determined by n) one can thus calculate A F . Results of such
calculations at OC are given for a number of monovalent metals in
Table II-I. The point of special interest is the fact that the mean free path
is of the order of several hundred Angstroms.
Table 11-1. Conductivity, Mean Free Path and Relaxation Time at OC
for Some Monovalent Metals
Metal
Li
Na
K
Cu
Ag
<1obs X
10" (esu)
1.1
2.1
1.5
5.8
6.1
Ef' (ev)
4.7
3.1
2.1
7.0
5.5
Af'
(A)
110
350
370
420
570
Tf'
in IO- u sec
0.9
3.1
4.4
2.7
4.1
Before the development of the band theory by Bloch and others, this
fact presented a great difficulty. The electrons were supposed to move in
the spaces between the ionic cores, as illustrated in Fig. 11-2, and such a
model inevitably leads to a mean free path of a few Angstroms. This model
also led to unsurmountable difficulties in explaining the temperature
dependence, pressure dependence, influence of impurities on the conductivity, etc.
In Chapter 10 we have seen, however, that the wave vector of an
electron moving in a perfectly periodic potential remains unchanged in the
absence of external fields. Thus, as a result of the wave nature of the
electrons, they can pass through a perfect crystal without suffering any
Sec. 11-5]
CONDUCTIVITY OF METALS
285
CONDUCTIVITY OF METALS
{Chap. II
The average potential :,lergy associated with the vibration is equal to half
Na
'".....o
16
Li
Ag
12
Ca
Cu
Au
Mg
4
Be
10
Atomic number
the total thermal energy, i.e., equal to kT/2 for temperatures well above the
critical temperature (j = hv/k. Hence
27T2V 2 M(x 2 )
(x 2 )
kTj2
T}>
(j
(I 1-35)
The quantity
is of particular interest for the scattering of electrons.
In order to see this, we shall first introduce the "scattering cross section"
QF associated with an atom with reference to its capability of scattering
an electron with the Fermi energy. From the definition of AF it follows
that an electron traveling over AF has unit probability C,-_'1eing scattered.
Suppose we represent the atoms by obstacles with a cross section QF
perpendicular to the direction of motion of the electron. Then QF may
be defined by the relation
(11-36)
Sec. 11-6]
CONDUCTIVITY OF METALS
287
where N is the number of atoms per unit volume. Since there is no scattering of electrons (Q J.' = 0) when the atoms are all in their equilibrium
position, one may expect that Qp is proportional to <x2 ) (both have the
dimensions of an area). Accepting this, it follows from the last two
equations that
(11-37)
Aj.' = const. M()2/T
Combining (11-37) with (11-33), we may write the conductivity in the form
a = const. M02/T
T'}>&
(11-38)
Thus a varies as T-l, in agreement with the experimental fact (3) mentioned
in Sec. 11- I. Expression (11-38) may be brought in harmony with Bloch's
theory if 0 is interpreted as the Debye rather than the Einstein temperature;
this will be done from now on.
In comparing different metals, it is more meaningful to compare
a/M()2 values than the a values themselves. The reason is that the former
quantity is a measure for the conductivity per unit amplitude of vibration
of the atoms. In Fig. 11-3 we have plotted a/ M()2 as function of atomic
number for T = 300o K. It is observed that the alkali metals and the
noble metals with one outer electron exhibit large values of this quantity,
indicating a relatively small cross section for scattering. For the divalent
metals next to them in the periodic table, alM02 is smaller by a factor
between 2 and 4; this is a consequence of the small effective number of
free electrons in these metals. Note also the low values of alM02 for the
transition metals.
As a result of the expansion of the lattice and the associated reduction
in the binding forces, () decreases slightly at high temperatures; consequently aT is not exactly constant but decreases somewhat at high temperatures. The transition metals form an exception to this rule; they
exhibit an increase of aT with increasing T which may be explained on the
basis of the band structure of these metals. 8
Matthiessen's rule. When a metal contains impurities, the field in the
vicinity of the impurities is in general different from that near the host
atoms. The impurities thus produce deviations from the periodicity of the
potential and act as scattering centers for electrons. Thus electrons in an
impure metal are scattered by impurity atoms as well as by the thermal
vibrations of the atoms. Denoting the relaxation times associated with
each of these processes by T j and TtlI' respectively, the resulting relaxation
time T is given by
(11-39)
liT = I/Ti + I/ Ttl.
because the probabilities for scattering in this simple model are additive
8
288
[Chap. II
'""
10 6 P
~.
~"
5
4
~
~
3
2
-200
-100
15
;':i ~;
Q.
'
o
Cu
25
50
--.Atomic
75
'.I;
Au
100
Au
10
",q
See. 11-6]
289
CONDUCTIVITY OF METALS
>
01-40)
!HiifH')fr
290
CONDUCTIVITY OF METALS
[Chap. 11
thermal equilibrium. For critical reviews of this subject and for the
details of the theory we must refer the reader to the literature. 4 A few
remarks will be made here in connection with the description of electronlattice interaction in terms of electron-phonon collisions.
Suppose an elastic wave of wave vector q and angular frequency Wq is
propagated through a crystal lattice. The displacement of an atom at the
lattice point r due to the vibrational mode may then be written in the form
(see Chapter 2), /
where Aq is the amplitude. For a transverse mode the displacement is
perpendicular to q; for a longitudinal mode it is parallel to q. At the
temperature T the average energy associated with this mode is given by
Planck's formula,
,. 1
liwq/[exp (liwq/kT - I)]
It is convenient to call1iwqthe energy of a "phonon," in analogy with the
liw q
licsq
k'
q + 27Tb
(11-42)
where either the upper or the lower signs should be used. The vector b is
a vector in the reciprocal lattice and for a simple cubic lattice 27Tb =
(27T/a)n, where n is a vector with integer components. For the moment we
shall assume b = 0; in this case the selection rules have a simple physical
interpretation: (11-41) expresses the conservation of energy in an electronphonon collision, the
sign corresponding to absorption, the - sign
corresponding to emission of a phonon by the electron. Similarly, (11-42)
(with b = 0) may be considered as expressing the law of conservation of
momentum in an electron-phonon collision; the momentum of the
Sec. 11-7]
291
CONDUCTIVITY OF METALS
.(
Since the absolute value of the left-hand side of this equation is ~ 1, the
electrons can interact only with such phonons for which q ~ 2k. Thus
low-energy electrons with small k-values can interact only with a fraction
of the total spectrum of vibrational modes; electrons near the Fermi level
can interact with essentially the whole spectrum of vibrations.
At temperatures far below the Oebye temperature, there are essentially
only phonons for which the wave vector q satisfies the inequality
(11-44)
(k o is Boltzmann's constant); the higher-frequency modes require too
much excitation energy. Consequently, electrons near the Fermi level can
be scattered only over small angles when the temperature is low. In fact,
according to the last two equations,
(11-45)
A few remarks may be made here about the case for which b =f:: 0
in equation (11-42); such processes are called "Umklapp-Prozesse"
("reversal processes").14 For cubic crystals they are described by
k'
q + (27T/a)n
(11-46)
292
CONDUCTIVITY OF :M'ETALS
[Chap. 11
(cgs)
(11-47)
16
Sec. 11-8]
CONDUCTIVITY OF METALS
293
<
<
<
peT) =
( 10 5 (OIT (eX
AT).Io
5
X
dx
.. "
_ 1)(1 _ e-X)
(11-48)
<
294
. [Chap. II
CONDUCTIVITY OF METALS
(11-49)
o (sp.
heat)
o(resist.)
Na
Cu
Ag
Au
Al
Pb
Ta
159
202
315~330
210~215
163~186
305~337
223
175
390
395
82~88
333
86
333
245
228
19 Work by W. 1. de Haas, 1. de Boer, and G. 1. van den Berg has been reported
in Physico 1, W9 (1934); 1, )115 (1934); 2, 453 (1935); 3, 440 (1936); 4,683 (1937).
20 D. K. C. MacDonald and K. Mendelssohn, Proc. Roy. Soc. (London), A202,
523 (1950).
295
CONDUCTIVITY OF METALS
Sec. 11-9J
200
160
.... ,
120
80
40
<
.'
, ; . -
, i,
20
40
60
lOO'K
80
-- T
Fig. 11-7. The "apparent" characteristic temperature () for Rb, as
deduced from resistivity measurements by employing (11-49). [After
MacDonald and Mendelssohn, ref. 20]
I '
K = -Qx/(dT/dx)
,",11: "
(11-50)
296
[Chap. II
CONDUCTIVITY OF METALS
K=
--f
~CvA
11-.
(11-53)
where C is the specific heat (at constant volume) of the gas per unit
volume, v is the average velocity of the molecules, and A is the mean free
path. In analogy, we may write for the conductivity associated with the
Umklapp processes,24
K"
= t
~~ C;jv;jA;j
(11-54)
, J
24
Sec. 11-9]
297
CONDUCTIVITY OF METALS
.'il
+ IIKb + IlK;
(11-55)
oc T-I
and
Ku oc T-I
for
T?> 0
(11-56)
For macroscopic crystals which are well annealed, (ii) and (iii) may usually
be neglected in this range of temperatures. For example, for NaCl at
oDe, K = 0.017 cal cm- I degree- I sec-I; assuming K = K u, we find by
using the simple expression (11-53) that Au'__ 20 A on the basis of a
specific heat of 0.45 cal cm-3 and a velocity of sound of,....,5 X I05 cm sec-I.
When processes (ii) and (iii) lead to a mean free path of the same order
as Au or smaller, they can, of course, no longer be neglected. It is obvious
that (ii) and (iii) may be expected to become important at low temperatures
and in imperfect crystals.
In considering the Umklapp processes at low temperatures (T ~ (J),
we must point out that equation (11-52) indicates that Umklapp processes
can occur only when the phonons have an energy larger than a certain
minimum value. In fact, we want at least one of the q's to be of the order
Ija, corresponding to a phonon energy ,....,ko(J (k o is Boltzmann's constant).
Peierls takes as a threshold energy ko(Jj2.22 Now the number of phonons
with this energy is proportional to lj[exp (Oj2T) - 1]. From this we
deduce that the temperature-dependences of Au and Ku at low temperatures
are essentially given by
(11-57)
Thus the Umklapp processes lead to a thermal conductivity which
298
[Chap. II
CONDUCTIVITY OF METALS
...................
~)Jin ,b~)'bbi.
s~~ttering by electrons
Fig. 11-8. The fully drawn curve represents the general theoretical
. form of the thermal conductivity of an insulator; in metals,
phonons are scattered by electrons as well (dotted curve), leading
to the dashed resultant curve. [After R. E. B. Makinson, Proc.
Cambro Phil. Soc., 34, 474 (1938)]
Sec. 11-IOJ
CONDUCTIVITY OF METALS
299
e,
- ( f - jO)jT = -eE",(aj/apx)
+ v",(aj/ax)(aTjax)
(II-58)
As long as the electric field and (aT/ax) are small, we may replacejon the
right-hand side by /0' as we did in calculating the electrical conductivity.
The thermal current density in terms of the distribution function F(pxpypz)
introduced in Sec. 11-4 is given by
(II-59)
i
300
CONDUCTIVITY OF METALS
[Chap. II
vanishes, one finds for the electronic thermal conductivity K r ' in the free
electron approximation,26
(11-60)
Here T F is again the relaxation time for electrons at the Fermi level. From
the theory of interaction of electrons with lattice vibrations one can show
that TF is proportional to T--I, so that
35
30
25
20
tK
Kr
constant
T~ (j
(11-61)
(T ~ ()
"
(11-6;:::~
Here L is called the Lorenz number;
10
the theoretical value is in rather good
agreement with experimental data in
5
the high-temperature region.
It can be shown that the existence
of a relaxation time is a sufficient
80
20
60
100
o
40
condition for the constancy of the
---+- T (OK)
Lorenz number. Experimentally one
Fig. 11-10. The thermal conductivity
finds, however, that as the temperaof two samples of sodium; sample II is
ture decreases, L decreases, indicating
purer than sample I. [After Berman
that the concept of a relaxation
and MacDonald, ref. 27]
time cannot be extended to low temperatures. At this point we may mention that, like the electrical resistivity,
the thermal resistivity associated with electrons may be considered to
consist of two parts: one due to scattering by lattice vibrations, another
due to scattering by impurities or other lattice imperfections. Denoting
these parts, respectively, by subscripts I and i, we may write, if they are
independent,
(J 1-63)
15
The last equality follows from the fact that for impurity scattering one
may always define a relaxation time (see the end of Sec. 11-3). Thus, by
plotting T/K, versus T, one can obtain liLa; from the intercept at T = 0
and Krl may be determined by subtraction.
In the case of the electrical conductivity one can, even in the lowtemperature region where no relaxation time can be defined properly,
See, for example, A. H'"Wilson, op. cit., pp. 18,20],
Sec. 11-10]
301
CONDUCTIVITY OF METALS
/Hz
Kelectron.//("attice
= -rr2kgTn'iF/m ACc.
'U
R. Berman and D. K. C. MacDonald, Proc. Roy. Soc. (London), Al09, 368 (1952).
[Chap. 1 I
CONDUCTIVITY OF METALS
302
Ey = (IJc)vxHz
where r" is the average drift veiocity of the electrons. Also, the current
density may be expressed in terms of the number of electrons n per unit
volume as
t"
\ . rnnw vr;m
-IJnec
(1165)
Thus the Hall coefficient is determined essentially by the sign and density
of the charge carriers. Observed Hall coefficients for a number of metals
are given in Table 11-3. It is observed that a number of metals have
positive Hall coefficients. Qualitatively, this can be explained on the
basis of the band theory of metals, since a metal with a nearly filled band
is equivalent to a conductor in which the current is carried by positive
holes; this would change the sign of R. For further details on the Hall
effect see Chapter 13. We should mention that the same information as
obtained from Hall coefficient measurements can be obtained from the
thermoelectric force.
Table 11-3. Hall Coefficient of a Number of Metals at Room
Temperature, in volts/cm-abamp-gauss. (After Seitz, Modern
Theory o/So/ids, McGraw-Hili, New York, 1940, p. 183)
I
1012R H
.
Cu
Ag
Au
Li
Na
1012RH
--- I----~-
-5.5
8.4
Be
Zn
7.2
17.0
-25.0
Cd
AI
I0 12 RH
--
--~---
24.4"
3.3
6.0
3.0
Fe
Co
Ni
100
24
-60
Negative signs indicate electron conduction. positive signs indicate hole conduction.
,. See, for example, F. Seitz, The Modem Theory olSo/ids, McGraw-Hili, New York,
J940, p. 181.
Chap. II]
CONDUCTIVITY OF METALS
303
REFERENCES
Besides the books referred to at the end of the preceding chapter, the
following review papers may be consulted:
J. Bardeen, "Electrical Conductivity of Metals," J. Appl. Phys., 11,88 (1940).
R. Berman, "The Thermal Conductivity of Dielectric Solids at Low
Temperatures," Advances in Physics (quarterly supplement of the
Philosophical Magazine), 2, 103 (1953).
P. G. Klemens, "Thermal Conductivity of Solids at Low Temperatures,"
Encyclopedia of Physics, Springer, Berlin, 1956, vol. 14, pp. 198-276.
D. K. C. MacDonald, "Properties of Metals at Low Temperatures,"
Progress in Metal Physics, 3, 42 (1952).
')<1)' ,
D. K. C. MacDonald, "Electrical Conductivity of Metals and Alloys at
Low Temperatures," Encyclopedia of Physics, Springer, Berlin, 1956,
vol. 14, pp. 137-197.
J. L. Olsen and H. M. Rosenberg, "On the Thermal Conductivity of
Metals at Low Temperatures," Advances in Physics, 2, 28 (1953).
"Proceedings of the International Conference on Electron Transport in
Metals and Solids," Can. J. Phys. 34, Dec. 1956, No. 12A.
PROBLEMS
11-1. From the observed. electrical conductivity of copper at room
temperature, calculate the relaxation time and the mean free path for
electrons at the Fermi level on the basis of (11-30); assume one free
electron per atom. Also calculate the average drift velocity of these
electrons in a field of I volt per cm and compare the result with the average
velocity in the absence of a field.
,\.....
11-2. Show that on the basis of the classical picture of electron
scattering by rigid spheres (the atoms) and on the assumption that the
electrons obey Boltzmann statistics, the electrical conductivity should be
proportional to T-1/2. How does this compare with experiment?
11-3. Set up a simple classical theory for the thermal conductivity
K ofa metal and show that in this theory KjaT= 3(kje)2 = 2.48 X 10-13 cgs
unit, where a is the electrical conductivity. This is the Wiedemann-Franz
law. See for example the first chapters of the books by A. H. Wilson,
0p. cit., and by N. F. Mott and H. Jones, op. cit.
11-4. Show that if the ions in a metal behave as rigid spheres with
respect to electron scattering, a relaxation time can be properly defined
(see, for example, A. H. Wilson, 0p. cit., p. 8).
304
CONDUCTIVITY OF METALS
[Chap. II
,,
Chapter 12'
,-.j~
Z(E)F(E) dE
(12-1)
306
[Chap. 12
, ~.
SneE) dE = SZ(E)F(E) dE =
ji;;LI
n\
( 12-3)
where 11 is the total number of electrons per unit volume. The general
procedure of calculating neE) for given Z(E) and T therefore is this: from
(12-3) one calculates E]<' and from it
neE) may be determined by substitution into (12-1).
:::;::====Ec
f
Eg - - ----EF
Illlllllliillli
Ev
= ---------
n
C
+1
(12-4)
Z
exp [(Ev - EF)/kT]
-------:-c:-:---::--
(12-5)
It will be evident that for gap widths of the order of several electron volts,
practically all electrons in the conduction band originate from the valence
Sec. 12-2]
307
band, so that the presence of bands below the latter may be neglected. l
]n other words, we may write
(12-6)
Substituting (12-4) and (12-5) into this expression, one obtains an equation
for EF leading to:
(12-7)
Thus, in this model, the Fermi level is located exactly halfway between the
valence and conduction bands. Also, its position is independent of temperature in this approximation.
The density of electrons in the conduction band may now be found by
substituting EF from (12-7) into (12-4). Ifwe assume that the Fermi level
is more than about 4kT away from the conduction band, the term
unity in the denominators of (12-4) and (12-5) may be neglected to a good
approximation. In that case,
(12-8)
where Ey = Ec - Ev represents the width of the forbidden gap. This
result may be compared with the improved formula (12-19). The number
of holes in the valence band is, of course, equal to nco Note the occurrence
of half the gap width in the Boltzmann factor (see Problem 12-1). Clearly,
when log nc is plotted 'versus I/T, a straight line with a slope of -Eg /2k
results (see Fig. 12-6). In this connection it is of interest to note that the
conductivity of a material is given by
(12-9)
where fl represents the mob"ility of the charge carriers, (i.e., the velocity
per unit electric field); the subscripts e and h refer to electrons and holes,
respectively. In the case under discussion ne = nh = nco One speaks in
this case of intrinsic conductivity. Now we shall see in the next chapter
th;:tt flc and flh are much less strongly temperature-dependent than the
density of electrons and holes. The temperature-dependence of II in the
intrinsic region is therefore essentially given by (12-8); i.e., log II versus
I/Tyields a straight line with a slope of -- Eg /2k. We shall see below that
the same result is obtained with a more sophisticated model. Note that
the conductivities of insulators and intrinsic semiconductors increase with
increasing temperature. In contrast to this, the conductivity of metals
decreases with increasing T; the reason is that in metals the density
of charge carriers remains constant and the mobility decreases with
increasing T.
I The reader is reminded of the fact that at room temperature kT
gap width in a good insulator is several ev.
308
[Chap. 12
I~
~'
(12-10)
where Eo represents the bottom of the conduction band and Z(E) is the
I,-----Z(E)
density of the states (see Fig. 12-2). Because we expect from the results
obtained above that EF lies roughly halfway between Ev and Eo, the
Fermi function F(E) decreases strongly as one moves up in the conduction
band. In other words, to evaluate the integral (12-10) it is sufficient to
know Z(E) near the bottom of the conduction band and one may then
integrate from E = Eo to E = 00. Near the bottom of the conduction
band we have, in accordance with (10-79),
(12-11)
m:
where
is the effective mass of an electron near Ec. Hence the density
of electrons in the conduction band is
_ (4 /h3)(2 * 3/2 ~ 00 (E - Eo)1I2 dE
no 7T
me)
(E-E )/11:7'
.'
Ee e
F'
1
;.,
(12-12)
Sec. 12-3]
309
n,. =
E,;
Z(E)[1 - F(E)1 dE
boil,om
(12-14) .
mt
where
represents the effective mass of a hole near the top of the valence
band. Ifwe make the assumption that the Fermi level lies more than about
4kT above E", we may use the approximation
1 - F(E)
~ e(E-Ep )/k1'
(12-16)
Substituting the last two expressions into (12-14) and integrating from
-00 to E", one obtains in the same way as above
(12-17)
Employing the fact that nc
nIl'
In case
mt = m:, the Fermi level lies again exactly halfway between the
2 For numerical tables of integrals of the type (12-12), see J. McDougall and E. C.
Stoner., Phil. Trans., A237, 67 (1929).
.
310
[Chap. 12
top of the valence band and the bottom of the conduction band; (l2-18)
is then identical with (12-7). In general
>
and the Fermi level is
raised ~ightly as T increases. This is indicated schematically in Fig. 12-5
by t~:; '~ntrinsic Fermi level."
The density of electrons in the conduction band nc and the density of
holes in the valence band nIL may be obtained by substituting (12-18) into
(12-13). This gives
mt m:
(12-19)
where Eg represents the gap width. It is observed that the temperaturedependence is the same as in the simplified model. The temperaturedependence of nc is represented schematically by the curve labeled
"intrinsic" in Fig. 12-6. It is convenient to remember that at room
temperature
\
(12-20)
where m is the mass of a free electron. Note that the constant in front of
the exponential in (12-8) is much larger than that in (12-19). We emphasize
again that (12- I 8) and (12-19) are good approximations only if the Fermi
level is more than a few kT away from the bottom of the conduction band
and from the top of the valence band.
12-4. Models for an impurity semiconductor
,.
Sec. 12-4]
311
Ec ----..._-----'!1'---
Condo band
-------------Ec
E;...I...
_~~-l
Ev ~}$;~j@$$~
(a)
Fig. 12-3. Donor levels are indicated in (a); one of the donors
is ionized, leading to a free electron in the conduction band.
Acceptor levels are indicated in (b); one of them is ionized (i.e.,
occupied by an electron from the valence band), leading to a free
hole.
simplicity that EF lies more than a few kT below the bottom of the conduction band. In that case, the density of conduction electrons nc is given
by (12-13). This number must be equal to the density of ionized donors.
If we assume that EF lies more than a few kT above the donor levels,
the density of empty donors is equal to
... 'I .-'.'"
(12-21)
Equating (12-13) and (12-21), one obtains for the location of the Fermi
level the expression,
EF = i(Ei
(12-22)
Thus at T = 0, Ep lies exactly halfway between the donor levels and the
bottom of the conduction band. As T increases, the Fermi level drops.
This is illustrated in Fig. 12-4 for the case Ec - E; = 0.2 ev for three
3 Semiconductors in which the current is carried predominantly by electrons are
called n-type semiconductors, (n = negative); a hole conductor is referred to as a
p-type (p = positive) semiconductor.
,~ ~
312
[Chap. 12
~::ifferent
';.
values of nd. 4 Within the triangul~r region ABC the Fermi level
IS more than 2kT away from the conductlOn band and from the donor
levels; only in this region is (12-22) applicable (with an accuracy of about
8 per cent). Outside this region, the term unity in the Fermi distribution entering in (12-21) must be retained. Note that for this model, EF
falls indefinitely; in an actual case, however, the presence of the valence
Cond. band
\
-.3
-.4
200
600
400
-T('K)
band would ultimately keep the Fermi level about halfway between the
valence and conduction bands (see Fig. 12~S).
For the region in which (12-22) is applicable, the density of free
electrons in the conduction band is obtained by substituting Ep into
(12-13), leading to
(12-23)
. ,
'
~
Sec. 12-4]
313
the same way. The density of holes in the valence band, making similar
assumptions as above, is given by an expression similar to (12-23). In this
case the Fermi level lies halfway between the acceptor levels and the top
of the valence band at T = 0; as T increases, the Fermi level rises (see
Problem 12-3 and Fig. 12-5).
'I
"..-~." .,.
Condo band
..._ .....
Donors
From these results it follows that the logarithm of the density of carriers
plotted versus the reciprocal temperature should yield a straight line of
slope -!J..E/2k. However, as the temperature is increased to such values
that the intrinsic excitation becomes important, the slope changes gradually
to -Egap/2k. The reason is that the density of electrons in the filled band
is of the order of 1022 per cm3 , whereas the density of impurity centers is
usually '( 1019 per cm3 This is illustrated schematically in Fig. 12-6.
Similar curves are encountered when the logarithm of the conductivity
is plotted against liT, as we shall see in later chapters.
(ii) The above model applies to a large extent to semiconductors such
as germanium and silicon, containing trivalent or pentavalent impurities;
the former produce acceptor levels, the latter donor levels. In other cases,
such as the alkali halides containing excess metal, the density of available
levels may be larger than the number of excess electrons. In other words, "
it is possible that at T = 0 only a fraction of the' available levels is occupied.
As an extreme case, we shall assume that the density of donor electrons nd
is very small compared with the density of available levels Zi' In this case,
314
[Chap. 12
the Fermi level evidently lies below the donor levels. At any temperature
T, the number of filled "impurity" levels is equal to ,
+I -
--;-;;;---;;-.,.;':,..;;;-_ _ r-J
e(Ei-E,l/kT
Z.e (E,-E;l/kT
,
where we assumed (E; - E F) .2: few kT. As long as the temperature is low,
the density of electrons in the impurity levels is large compared with
the density of electrons nc in the
conduction band and we may write
(12-24)
,
--..ljT
Sec. 12-5]
315
Let the vacuum level (i.e., the energy of an electron at rest outside the
semiconductor) be higher than the bottom of the conduction band by an
amount X as indicated in Fig. 12-7; X is called the electron affinity of the
crystal. If x is the direction perpendicular to the surface, an electron needs
at least a momentum in the x-direction given by
(12-27)
in order to escape. As a result of thermal excitation let there be nc electrons
per unit volume in the conduction
band. If the Fermi level is assumed
to lie more than a few kT below the
Cond. band
bottom of the conduction band, the
conduction electrons have a MaxFenni level
wellian velocity distribution according to the discussion of the preceding
sections. We leave it as a problem
to the reader to show that the Fig. 12-7. lllustrating the electron
density of electrons with momenta affinity X and the work function q,
in the range dpx, dp1l' dpz is then
of a semiconductor.
equal to
_j_
(12-28)
1=
enc
(21Tm*kT)3!2
SSS(p
'II
(12-29)
The integra1ions over P1l and pz go between oo; the integration overpx
extends from
Po;0 to 00. This yields
,
1=
enc
(21Tm*kT)3!2
(12-30)
316
[Chap. 12
]n the preceding sections it was assumed that the Fermi level was
located at least a few kT below the bottom of the conduction band. In
that case the electrons in the conduction band follow closely Boltzmann
statistics, i.e., the electron gas is nondegenerate. Under certain circumstances, however, the Fermi level may enter the conduction band and the
electron gas in the conduction band may become degenerate. From the
preceding discussions it should be clear that the conditions favorable for
such a situation are the following:
(i) Relatively high donor densities (,...., 1019 per cm3)
(ii)
Sm::~_l
(iii) Lc.. density of states near the bottom of the conduction band,
i.e., small effective electronic mass (see Sec. 10-9).
When these conditions are fulfilled, the Fermi level as function of
temperature varies as indicated by the dashed curve 3 in Fig. 12-5. As T
increases from absolute zero, the donors begin to ionize and as a result of
the low density of states, the lower energy states in the conduction band
become completely filled. The position of the Fermi level relative to the
bottom of the conduction band is then given by (9-9),
/'
where nc is the density of electrons in the conduction band. As long as
EF }> kT, the electron gas is degenerate. Clearly, as the effective electronic
mass is reduced, degeneracy may occur at lower electron densities. As T
is increased further, the degeneracy is removed and the Fermi level leaves
the conduction band again.
The circumstances described here are believed to occur in InSb,
containing donor levels in concentrations of about 1018 per cm3 ; the
effective mass of the conduction electrons is probably only about m/30 in
this case.
~
REFERENCES
J. S. Blakemore, "Carrier Concentrations and Fermi Levels in Semiconductors," Elec. Commun., June 1952, pp. 131-153.
R. A. Hutner, E. S. Rittner, and F. K. DuPre, "Fermi Levels in Semiconductors," Philips Research Repts., 5, 188 (1950).
F. Seitz, The Modern Theory of Solids, McGraw-Hili, New York, 1940,
pp. 186 ff.
W. Shockley, Electrons and Holes in Semiconductors, Van Nostrand,
New York, 1950.
Chap. 12]
317
PROBLEMS
12-1. With reference to the problem discussed in Secs. 12-2 and 12-3,
consider the reaction
electron in valence band
Applying the law of mass action as used in chemical reactions, show that
the equilibrium concentration of the conduction electrons is proportional
to exp ( -~ Egapj2kT).
_" __ _
12-2. With reference to the problem discussed in Sec. 12-4, consider
the reaction
bound electron
free electron
+ empty donor
Making use of the law of mass action, answer the following questions:
(a) Assuming that at T = 0 all donor levels are filled, show that the
density of free electrons is proportional to nYz exp ( -I::iEJ2kT).
(b) Assuming that at T = 0 only a small fraction of the donor levels
is filled, show that the density of free electrons is proportional to Zi exp
(-I::iEjkT), where Z, is the density of impurity levels and I::iE is the ionization energy of the donor levels.
12-3. For an intrinsic semiconductor with a gap width of 1 ev,
calculate the position of the Fermi level at T = 0 and at T = 300, if
=
Also, calculate the density of free electrons and holes at
T = 300 and at T = 600.
m: Sm:.
nenh
2.33 X
IQ3IT3 e -E.'kT
where Eo is the gap width. Note that this holds irrespective of the presence
of donors or acceptors in the gap, as long as the condition imposed on the
Fermi level is satisfied.
318
[Chap. 12
Chapter 13
NONPOLAR SEMICONDUCTORS
13-1. Introductory remarks
Semiconductors are characterized by an electrical conductivity
(associated with the motion of electrons or holes or both) which on the
one hand is considerably smaller than that of metals, and on the other
hand, is much larger than that of "insulators." Furthermore, the conductivity increases with temperature, in contrast with the behavior of
metals at normal temperatures. The number of current carriers per unit
volume in a semiconductor is in general much smaller than the number
of atoms per unit volume. This situation is encountered, for example, in
a solid for which the forbidden energy gap between the highest normally
filled band and the conduction band is small, i.e., of the order of one
electron volt. At absolute zero such a solid is an insulator, and as the
temperature is raised, the density of free electrons and holes increases as
explained in the preceding chapter. In this case the density of free electrons
equals that of the free holes and one speaks of intrinsic semiconductors;
the properties are then characteristic of the solid itself. Semiconductor
properties may also be exhibited by solids which in the pure state are
good insulators, viz., when impurities are present which either donate
free electrons to the conduction band (donors) or free holes to the upper
filled band (acceptors); in this case one speaks of extrinsic or impurity
semiconductors. Impurity conductivity may of course be superimposed
on the intrinsic semiconductor properties of a solid.
The semiconducting elements are those appearing within the area
enclosed by the lines drawn in Table 13-1; this table represents the A
subgroup elements in a number of columns of the periodic table. Of
these, silicon and germanium have received a great deal of attention
because of their great technical importance, particularly in the field of
crystal diodes and transistors. The discussion in this chapter will be
concerned mainly with the properties of Si and Ge; the amount of
literature on this subject is so vast that the discussion is necessarily very
incomplete. A review which is up to date until the beginning of 1955
may be found in H. Y. Fan in F. Seitz and D. Turnbull (eds.), Solid State
PhysiCS, Academic Press, Ne~ York, 1955, Volume I.
319
320
[Chap. 13
NONPOLAR SEMICONDUCTORS
Table 13-1. The A Subgroups of the 3rd, 4th,5th, 6th, and 7th Columns
. of the Periodic System of Elements
IlIA
IVA
VA
VIA
BeN
Si
P
AI
Ge
Ga
In
Sn
Se
Te
TI
Pb
VIlA
\ __
Po
At
Structure. Diamond, silicon, germanium, and grey tin all have the
diamond structure represented in Fig. 13-1. Each atom is surrounded by
four others, occupying the corner points of a tetrahedron, to which it is
bound by electron pair bonds. The structure may be described by an
f.c.c. point lattice in which each lattice point corresponds to two atoms,
one located at (0,0,0) and another at (i,1-,t). The free atoms of the
elements have an outer electron configuration in which two electrons
occupy an s state and two others occupy a p state. In the solid state the
total of four outer electrons per atom is just sufficient to produce electron
pair bonds with four other atoms; in this configuration the sand p wave
functions. form hybrid wave functions giving rise to four equivalent
chemical bonds, the angle between any two of them being approximately
109.1 This type of covalent binding may be contrasted with the ionic
bonds in crystals such as the alkali halides; in the latter, the particles are
charged and the field around a given ion is spherically symmetric, i.e.,
the restriction on the coordination number is essentially of geometrical
origin. In terms of a two-dimensional picture one arrives at an electron
distribution as indicated schematically in Fig. 13-2.
One expects the electrons taking part in ,the electron pair bonds to be
rather strongly bound, i.e., one expects that a certain amount of energy
is required in order to set them free to the extent that they can move about
in the crystal. This is in agreem.ent with the fact that at very low temperatures these elements areinsulators. In terms of the energy band scheme,
I See L. Pauling, The Nature of the Chemical Bond, 2d ed., Cornell University Press,
Ithaca, 1945, p_ 81.
Sec. 13-2]
321
NONPOLAR SEMICONDUCTORS
this means that at absolute zero the electron distribution is such that a
certain number of energy bands is completely filled, the higher ones being
completely empty.
f-
..
j----
0
0
~'t--~~-..- _- . -."
0 . 0
.
..
. .
0
'~~ ~
.~
..
-----
~.
Fig. 13-2. Schematic two-dimensional representation of the electronic distribution in the diamond
structure, showing the electron
pair bonds.
Si
Ge
3.561
5.43086
5.65748
m.p.
eq
3550
1420
936
5.7
12
16
On
Egap
("K)
(ev)
C ll
C12
c ..
(10 12 dynes/em)
1800
658
362
-7
1.21
0.785
9.2
1.674
1.298
3.9
0.652
0.488
4.3
0.796
0.673
322
NONPOLAR SEMICONDUCTORS
[Chap. 13
Sec. 13-2]
NONPOLAR SEMICONDUCTORS
323
analogous to that of the hydrogen atom. 7 The difference is, however, that
the extra electron and the positive charge are embedded in a medium of
rather high dielectric constant (see Table 13-2). As a result, the radius of
the orbit covers several atomic distances and the binding energy is small.
In fact, if one employs for an estimate the simple Bohr picture modified by
taking into account the dielectric constant and the effective mass m*, one
obtains for the radius and the energy of the ground state
(13-1)
The energy Ed is measured relative to the bottom of the ionization
continuum, i.e., relative to the bottom of the conduction band. Assuming
for the moment the effective mass m* to be equal to the free electron mass,
one finds
Si
12
Ge
16
A8.5 A
6.4
E" (Bohr)
-0.0gev
-0.05
-0.05 ev
-0.01
The last column gives the experimental ionization energy of the donor
levels for doping with P, As, or Sb. 8 A detailed calculation of the ionization
energy of donors by Kittel and Mitchell gives 0.009 ev as a lower limit for
germanium and 0.03 ev as a lower limit for silicon, in good agreement with
the experimental values. 9 These calculations made use of recent information about the E(k) surfaces as revealed by cyclotron resonance experiments
(see Sec. 13-6).
Silicon and germanium may also be doped with trivalent elements such
as B, AI, Ga, and In. In these cases the added atoms are one electron short
for four electron pair bonds. Each added trivalent atom thus gives rise to a
vacant electron level slightly above the valence band. These levels are
acceptor levels because they may accept an electron from the filled band if
the electron is excited thermally. One may picture the acceptor level as a
hole describing a Bohr orbit about the impurity atom; the binding energies
are approximately equal to those for the donors. Ionization of the
acceptor level in this type of picture is equivalent to the excitation of a
valence electron into the hole. In the energy band scheme, electrons are
excited upward, holes downward (see Fig. 13-3).
From what has been said above, it is evident that ionization of donor
levels (P, As, Sb) gives rise to electronic carriers in the conduction band;
ionization of acceptor levels (8, AI, Ga, Sn) produces hole conductivity in
the valence band. In germanium at room temperature nearly all donor or
, G. Wannier, Phys. Rev., 52, 191 (1937); see also G. F. Koster and J. C. Slater,
Phys. Rev., 95,1167 (1954); 96,1208 (1954).
8 J. A. Burton, Physica, 20, 845 (1954).
C. Kittel and A. H. Mitchell, PIlys. Ret'., 96, 1488 (1954).
NONPOLAR SEMICONDUCTORS
324
[Chap. 13
Cond. band
.08
fev
"Acceptors
~
Valence band
Sec. 13-2]
NONPOLAR SEMICONDUCTORS
325
14
(1955).
16
c. J. Gallagher, Phys.
326
NONPOLAR SEMICONDUCTORS
[Chap. 13
a = nef-l
where n is the density of carriers and f-l is their mobility (drift velocity per
unit field). It is observed that measurements of aCT) provide information
only about the product n(T)f-l(T), and in general do not allow one to determine these quantities separately. However, if we assume for the moment
that the Hall coefficient for a semiconductor is given by the formula
applicable to metals we would have (see 11-65),
RII = l/nec
and
caR H = f-l
( 13-3)
327
NONPOLAR SEMICONDUCTORS
Sec: 13-3]
present a great deal of evidence (see Sec. 13-6) that this is not correct, but
in many instances the simple theory still gives rather good agreement with
the experiments. It is also assumed that the electron gas in the conduction
band is nondegenerate, and thus that it has a Maxwellian distribution.
As an example, consider a semiconductor in which the current is
carried only by electrons in the conduction band. Suppose an electric
field E", and a magnetic field Hz are applied to the materiai as indicated in
Fig. 13-5. The current density I", along the x-direction may then be obtained
from the Boltzmann transport equation in the same way as for metals.
Thus from (11-28) it foHows that
e 2 E", f<Xl 3Fo 2
1,,= --Jo
v T(E)(87T/h 3)p2dp
3
E
(13-4)
(13-5)
The last approximation is valid
only if the density of the electrons
in the conduction band is small
enough so that Fo 4:,_ I, i.e., if the
system is nondegenerate. Recognizing that 87Tp2 dp Fo/h 3 equals
the number of electrons with momentum in the range dp, it follows that
I
I
Ix
ne2E
= __'"
3kT
(V 2T)
aEx = neftE"
(13-6)
Here (V 2T) is the average value of V2T(E), the average being taken over the
Maxwellian distribution of the conduction electrons. Since 3kT = m*(v2 ),
one may also express the mobility as
(V 2T)
m* . (V2)
ft
(13-7)~
Note that if T were independent of the velocity of the electrons, this would
reduce simply to ft = er/m*, as in the simplified model discussed in Sec.
11-2. We shall return to this ~xpression in the next section. ,;'J . Ii' .
328
NONPOLAR SEMICONDUQORS
[Chap. 13
The Hall effect may be discussed by considering the case for which the
front and back faces in Fig. 13-5 are short-circuited, allowing the flow of
a current along the y-direction. An electron of velocity u'" under influence
of the magnetic field Hz will develop a velocity along the y-direction such
that
( 13-8)
(OUI[ot)H, = eu",Hz/m*e = WU'"
On the other hand, due to collisions with the lattice,
(cvy/dt)roll =
-VII /T
ev",HzT/m*c
(13-9)
WTV",
In analogy, one may thus obtain the current along the y-axis by multiplying
the integrand of (13-4) by W'T. This finally leads to
(13-10)
Thus, although the electric field is applied along the x-direction, the
resultant current has a y-component due to the magnetic field. In fact, it
is convenient to define the Hall angle On (see Fig. 13-5), where
tan On ':::' On
= Iy/I", = W
<V2T2)
-2-
(I3-11)
(V T)
If the Hall contacts ar.: not short-circuited, a field Ell is set up to counteract
the influence of the magnetic field. The Hall coefficient then becomes
1
(V 2T2)(V 2 )
(13-12)
where a has been substituted from (13-6). Note that the sign of the carrier
in the above derivation is contained in e; for electrons RH is negative,
for holes it is positive. It should be mentioned that one frequently employs
the Hall mobility flH defined in analogy with (13-3) by
fln
eaRn
(V 27 2 )
=-'-m*c <V27)
.~_
(I3-13)
"-
Sec. 13-4]
NONPOLAR SEMICONDUCTORS
329
= A/v
(13-14)
and
flH = (37T/8)fl
(13-15)
Combining this result with (c) above, one concludes that the mobility fl
should be proportional to T-312 in this case. Bardeen and Shockl ey18 find
from their calculation of A,
eT
(87T)1121i4cll
(13-16)
m* = fl = 3E12m*512(kT)312 = const. T-:-312 Ie
Here, ell is the average longitudinal elastic constant, and E1 is the shift of
the edge of the conduction band per unit dilation; the temperature. dependence of both these quantities may be neglected. For holes, El
represents the shift of the edge of the valence band per unit dilation.
Experiments indicate that El ~ 10 ev for germanium. The formula
obtained by Seitz18 is written in terms of the Debye temperature (), the
mass M of the atoms, and their number per unit volume N,
'
2l!2 X 61/3 Nl!3 eIi 2P()2M
fl =
47T5/6
m*5/2C2(kT)3/2
The constant: C has the dimensions of an energy and is of the same order of
magnitude as E1 in the Bardeen-Shockley formula; it is a measure for the
electron-phonon interaction. The mebility determined by lattice scattering
alone is usually referred to as the "lattice mobility."
18 F. Seitz, Phys. Rev., 73, 549 (1948); J. Bardeen and W. Shockley, Phys. Rev., 80,
12 (1950).
19 Compare expression (J I-II) for the relation between collision time, relaxation
330
NONPOLAR SEMICONDUCTORS
[Chap. 13
The Hall coefficient as determined by lattice scattering for semiconductors containing one or two types of carriers is given, respectively, by
377
R n = - an
8nec
nhfl~ - nefl;
.. 377
R Il = - -
8ec (n"flh
+ n"fle)
(13-17)
where the subscripts e and h in the last formula refer to electrons and holes,
respectively. The conductivity for two types of carriers is of course equal
to (nee flc
nhe flh)'
at
material.
Tr/(l -- (cos 13
where (cos 13) is the average of the cosine of the scattering angle. Making
use of the Rutherford scattering formula, Conwell and Weisskopf have
calculated an approximate expression for T with the result that
(13-18)
where E is the dielectric constant. 20 It is observed that this type of scattering
leads to a mobility which varies approximately as T3!2, in contrast with the
T-3/2 law for lattice scattering.
The Hall coefficient and Hall mobility associated with ionic scattering
are found to be 21
(13-19)
RH = 1.93Inec,
flH = 1.93fl
(iii) Neutral impurity scattering. The scattering of charge carriers by
neutral impurities is quite similar to the scattering of electrons by hydrogen
20 E. M. Conwell and V. F. Weisskopf, Phys. Rev., 77, 388 (1950); see also W.
Shockley, Electrons and Holes, Van Nostrand, New York, 1950, pp. 258 If.; for a
quantum mechanical treatment, see H. Brooks, Phys. Rev., 83,879 (1951).
" W. Shockley, op. cit., p. 279.
Sec. 13-4]
NONPOLAR SEMICONDUCTORS
331
atoms. Thus, by suitably modifying the theory of the latter, Erginsoy has
calculated the mobility associated with this type of scattering alone. 22
He finds
). , T
m*e3
er/m* = f.l = - - -
~.
(13-20)
20Ndj3
"'1
because the probabilities for scattering are additive, each of them being
proportional to the reciprocal of the corresponding relaxation time,
._
'.
332
NONPOLAR SEMICONDUCTORS
[Chap. 13
\
-
T('K)
10
20.4
300
78
T('K)
20.4
300
78
107
10 6
55
p
55
103
10 2
.01
58
10rr______~5~8~______
.02
.04
.06
.08
.1
-1/T('K)
.02
.04
.06
.08
.1
--1/T('K)
sample 58 contains enough arsenic to make the electron gas in the conduction band degenerate over most of the temperature range. The other
samples have intermediate impurity densities.
In accordance with (13-13), the Hall mobility may be obtained from
the relation flH = CRH/P; the results are given in Fig. 13-9. It is observed
that the nearly pure sample 55 follows closely the T-3/2 law down to the
lowest temperatures. The reason for this is that neutral impurity scattering
Sec. 13-5]
333
NONPOLAR SEMICONDUCTORS
/llattice
900
300
100
1450
500
3600
1700
3000
(arbitrary temp.)
4.0 x looT-2.6
2.5 x 108T-2.3
4.9 x IO'T-1.66
1.05 x 10oT-2.33
c. C. Klick and J.
334
NONPOLAR SEMICONDUCTORS
[Chap. 13
between 1.0 and 1.5, i.e., less rapidly than predicted by the ConwellWeisskopf formula. The Erginsoy formula for neutral impurity scattering
fits their data well for an effective electron mass equal to about mj3. They
find scattering by dislocations negligible in their samples.
\
13-6. Constant-energy surfaces and effective mass in silicon and germanium.
The theory developed in Sec. 13-3 and 13-4 was based on the assumption that the energy of electrons near the bottom of the conduction band
or of holes near the top in the valence band could be represented by
1i2k 2 j2m*. This implies that constant-energy surfaces in k-space are spheres
and that m* is a constant independent of the direction of motion of the
carriers. It is presently believed that the discrepancies between theory and
experiment cited above are, at least in part, due to the fact that this
assumption is incorrect. Thus values of the effective mass calculated
indirectly from the electrical properties must be considered unreliable.
Measurements of the influence of a magnetic field on the resistivity of
single crystals of germanium also drew attention to the fact that the
constanf-energy surfaces cannot be spheres. 27
If the constant-energy surfaces are spheres, the effective mass is,
according to (10-38),
'.
However, if the energy is a function also of the direction of the wave
vector k, the effective mass is a tensor rather than a scalar, as was mentioned
in Sec. 10-4. Bya suitable choice of axes, this tensor may be diagonalized
in such a way that along the three principal axes the effective mass is
given by
(13-22)
m~ = 1i2j(d2E(k)jdkD where i = x, y, z
For example, for motion along the x-axis, the electron behaves as a particle
of effective mass 1i2/(d2Ejdk;), etc. Until recently, experimental information
about the effective mass, and hence about the curvature of constant-energy
surfaces in the k-space, could be obtained only indirectly, viz., from
experimental results for transport phenomena in which m* occurs.
However, cyclotron resonance experiments of electrons and holes have
made it possible for the first time to measure m* directly.28 In this type
of experiment, electrons in the conduction band and holes in the valence
band describe spiral orbits about the axis of a constant magnetic field H.
G. L Pearson and H. Suhl, Phys. Rev., 83, 768 (1951).
Dresselhaus, Kip, and,Kittel, Phys. Rev., 92, 827 (1953); Lax, Zeiger, Dexter, and
Rosenblum, Phys. Rev., 93, 1418 (1954); Dexter, Zeiger, and Lax, Phys. Rev., 95, 557
27
28
(1954).
335
NONPOLAR SEMICONDUCTORS
Sec. 13-6J
We =
(13-23)
eHjm*c
where r is the radius of the orbit; the plus or minus sign indicates the
opposite senses of rotation for electrons and holes. Resonant absorption
of energy from a radio-frequency electric field perpendicular to the static
~-----
---._
..
i
o
1000
,.-
2000
4000
3000
H (oersteds)
5000
6000
336
NONPOLAR SEMICONDUCTORS
[Chap. 13
it may suffice here to mention some of the results obtained for silicon
and germanium. 29 As. an example of the band structures obtained, we
give in Fig. 13-II the energy as function of the wave vector along the
(100) direction for silicon. It is observed that the minimum energy in the
conduction band does not correspond to k = 0 but that there are in all
six minima located somewhere along the six (100) axes. In the vicinity
of these minima, the constant energy surfaces are prolate ellipsoids of
. t
Electrons
~
:
:f
_1
I Eg
Forbidden band
'
-k
; "1;
Silicon:
mt
Germanium:
m t = 0.082m;
0.I9m;
m l = 0.98m
ml
1.57m
Sec. 13-6]
337
NONPOLAR SEMICONDUCTORS
The maximum energy for the valence band in both silicon and
germanium occurs for k = 0, according to the results of cyclotron
resonance experiments; furthermore, this maximum is common to two
bands which meet at k = O. The constant-energy surfaces near k = 0 for
these two 'bands are warped and are given by the expression
where A, B, and C are constants. The negative and positive roots correspond, respectively, to the highest (VI) and second highest (V2 ) valence
band. If one approximates the warped surfaces by spheres, one may
determine the average hole mass in the two bands from the experimental
values of A, B, and C. In this approximation, one obtains
Silicon:
mtl
Germanium:
mtr,
= O.l~m;
0.49m;
mt. = 0.16m
mtr. =
O.OMm
,;
We should note here that the form of expression (13-25) was indicated by
the theory of spin-orbit splitting for these crystals. ao
.
It is observed from Fig. 13-11 that there is a third valence band Va
which is separated from the VI and V2 bands as a result of spin-orbit
interaction. The maximum of the Va band lies slightly below that of the
two other bands. Near the maximum of the Va band, the constant energy
surfaces are spherical; the effective masses are:
Silicon:
mt.
Germanium:
mt
=
=
0.24m (
0.077m
: ,1
The energy difference between the top of the Va band and the common
maximum of the VI and V 2 bands has been estimated to be 0.035 ev for
Si and 0.28 ev for Ge. It will be evident that the relative hole populations
of the Va and VI' V2 bands is a function of temperature.
The energy gap. A few remarks may be made here about the consequences of the above results for the concept of the forbidden energy gap
and its experimental determination. When an electron is thermally excited
from the. valence band into the conduction band, the electron absorbs a
phonon. This process is governed by the selection rules corresponding to
conservation of momentum and energy:
=
E(k') =
k'
+ q + 27Tn
E(k) + liw
k
...
(13-26)
30 G. Dresselhaus, A. F. Kip, and C. Kittel, Phys. Rev., 95, 568 (1954); 98, 368
(1955); R. J. Elliott, Phys. Rev., 96, 266 (1954); 96,280 (1954).
338
NONPOLAR SEMICONDUCTORS
[Chap. 13
Here k' and k are, respectively, the final and the initial wave vector of the
electron; q is the wave vector of the absorbed phonon, and liwq is the
energy of the phonon; n is a vector in the reciprocal lattice, and 27Tn in
(13-26) guarantees that k' is a vector in the reduced zone. The "cheapest"
thermal excitation of an electron from the valence band to the conduction
band evidently involves a phonon of energy ;W)q = Eg where Eg is tl}.e
energy difference between the highest electronic level in the valence band
and the lowest level in the conduction band (see Fig. 13-11). Thus Eo
may be obtained from the variation of the carrier concentration with
temperature, We should note here that Eg itself is a function of temperature
(resulting from the expansion of the lattice).
Let us now consider what one measures if one determines the long
wavelength threshold for optical excitation of an electron from the valence
band into the conduction band in substances such as Si and Ge. If one
considers the optical excitation as a collision between an electron of wave
vector k and a photon of wave vector a the selection rules require
k' = k
+a
and
E(k') = E(k)
+ hv
( 13-27)
E(k') =
(13-28)
where the symbols have the same meaning as above. The presence of the
phonon momentum q thus makes it possible for the transition to be nonvertical. The
and - signs refer, respectively, to absorption and
emission of a phonon. It is of interest to recognize that at very low
at
Sec. 13-6]
339
NONPOLAR SEMICONDUCTORS
+ (ov.jot)coll =
(ejm*)Eoe iw1
V"JT
(13-29)
In the general case, however, it follows from (13-29) that the conductivity
is complex; the real part varies with frequency as
32 S. Meiboom and B. Abeles, Phys. Rev., 95, 31 (1954); I. Estermann and A. Foner,
Phys. Rev., 79, 365 (1950); G. L. Pearson and C. Herring, Physica, 20, 975 (1954).
340
NONPOLAR SEMICONDUCTORS
[Chap. 13
(IJT?>
(13-30)
where p. is the mobility of the carriers. Now if the constant-energy surfaces
are prolate spheroids, one can use
(13-30) by replacing m* by an average
50
effective mass
given by
m:v
m:v
Sec. 13-7]
NONPOLAR SEMICONDUCTORS
341
342
[Ch~p.
NONPOLAR SEMICONDUCTORS
13
where Dh is the diffusion coefficient of the holes and Ilh is the hole mobility.
There exists a fundamental relationship between the diffusion coefficient
and the mobility. According to elementary diffusion theory,
D
(13-34)
(A/3)(v)
where A is the mean free path for scattering and (v) is the average velocity
.i!' of the carriers. On the other hand, it follows from (13-7) and (13-14) that
-:"A M2
Il - m
(v
(13-35)
For thermal holes, m(v2 ) = 3kT, so that we obtain the Einstein relation,
IlID
(13-36)
e/kT
The same relationship is obtained from (13-33) by considering an equilibrium situation in which I" = 0. Under these circumstances, we should
have, according to Boltzmann,
n"
~Z
1 ..~
Sec. 13-7]
343
NONPOLAR SEMICONDUCTORS
Ix
-e Dan/ax
(we leave out the subscripts h). In (13-32), the term gh is zero for t > 0,
so that
(13-37)
an/at = D o2n/ox2 - n/T
The solution of this equation is
, n(x ,
t) =
'-
(41T Dt)1/2
e-xO/4Dt-t/T
"c~,",--'_:'"
",_j.-.'"
"-
." ,
_ ~
(13-38)
(13-39)
(DT)1/2
x
x
>:
x
x
0.737
0.716
0.705
OA8
0.48
Ty
(microsee)
144
78
16.5
9.2
3.1
T,
(microsec)
280
340
290
280
235
.,
344
[Chap. 13
NONPOLAR SEMICONDUCTORS
This shows clearly the important role played by the surface in the
recombination process in the roughly ground samples. The exact nature
of the centers at which the electron-hole recombination takes place is not .
understood. Estimates of direct recombination of electrons and holes
under photon emission indicate lifetimes of the order of one second. 3s
So far, the longest lifetimes observed are of the order of 10-3 second,
indicating that the direct recombination process is relatively unimportant.
It seems, therefore, that centers are required which act as a catalyst in
the recombination process. It is of interest to note that when a Ge crystal
is heated to higher temperatures and then quenched, the lifetime of the
carriers decreases. 37 This implies that certain types of frozen-in lattice
defects are at least in part responsible for recombination.
:f :
'i
P
As
Sb
, '. t
2.36
2.44
2.62
GaP
GaAs
GaSb
2.36
2.44
2.62
InP
InAs
InSb
2.54
2.62
2.80
Si
Ge
Sn
2.34
2.44
2.80
Sec. 13-8]
NONPOLAR SEMICONDUCTORS
345
fl.
65,OOO(T/300)-1.66cm2/volt sec.
IS ~
41
346
[Chap. 13
NONPOLAR SEMICONDUCTORS
being obtained at a rapid rate and this may be expected to continue for a
number of years to come. Some further references are given below. 43
I
REFERENCES
Proc. IRE, 40, (1952) (transistor issue); 43, (1955) (solid state issue).
"Semiconducting Materials," Proc. Reading Conference, Butterworths
Scientific Publications, London, 1951.
PROBLEMS
13-1. Calculate the distance between nearest neighbors in the germanium and silicon lattices.
13-3. Discuss how the elastic constants given in Table 13-2 can be
obtained from measurements of the velocity of elastic longitudinal and
shear waves.
13-4. On the basis of the Debye approximation, calculate the Debye
temperature for Si and Ge from the elastic constants given in Table 13-2;
compare the results with eD obtained from specific heat measurements.
13-5. From the dielectric constants given in Table 13-2 calculate the
polarizability per Si and per Ge atom in the crystalline state, assuming for
simplicity an internal field of the Lorentz type.
13-6. A germanium crystal contains 10-4 atomic per cent of arsenic;
assuming all donors are ionized, calculate the resistivity at room
temperature.
" For JnSb. see R. G. Breckenridge el al., Phys. Rev., 96, 571 (1954); for GaSb,
R. F. Blunt, W. R. Hosler, and H. P. R. Frederikse, Phys. Rev., 96, 576 (1954); and,
D. P. Detwiler, Phys. Rev., 97,1575 (1955); for AISb, R. F. Blunt, H. P. R. Frederikse,
J. H. Becker, and W. R. Hosler, Phys. Rev., 96, 578 (1954).
Chap. 13]
347
NONPOLAR SEMICONDUCTORS
'Ii
,
Chapter 14
"-,
(al
"
(bl
Sec. 14-IJ
349
,-:j
~,- ,;;
::1,
(a)
350
[Chap. 14
4>m - 4>. =
(27TjE)nde2x~
(14-4)
Thus, for a given value of the required potential energy drop, Xo varies
as n,II/2. A few examples may be given here for 4>m - 4>. = 1 eV and
= 10.
na
1015 1017 1019 per cm3
Xo
4>m - 4>. + e V =
(27T/)nde2x~
(14-5)
Sec. 14-3]
351
10
352
[Chap. 14
which they are prepared, the density of donors (or acceptors) is very
small near the metal and gradually increases to a constant value as one
moves into the semiconductor. In some instances one produces deliberately an insulating layer between the metal and the semiconductor by
chemical means.
As an idealized model for this type of system we shah consider the
following case. s An n-type semiconductor contains nd donors per cma.
The semiconductor is separated from the metal by a layer of 10-4_10-5
em thick of the same material but without donor levels. We shall further
aSStlme that any potential difference between the metal and semiconductor
essentially across the insulating layer, the field strength in the
layer being constant (see Fig. 14-4).6 Since the thickness of the layer Xo
is large compared with the mean free path for scattering of the electrons
by lattice vibrations, the electron current through the layer is due to
(I) the electric field and (2) diffusion. Let E represent the field strength
for a unit negative charge, and let [ be the electronic current density.
We may then write
(14-6)
[ = n(x)eflE - De dn/dx
ey(~
[/tuE
+ CeeEX/kT
,,~ (14-7)
Wiss. Veroffentl. Siemens Werken, 18,225 (1939). These references also take account of
the space charge region in the semiconductor.
Sec. 14-3]
353
n(x o)
n(O)e-(4>m-4>,l/kT
n(O)e-eVof kT
(14-8)
where CPm and CPs are the work functions of the metal and insulator,
respectively; the total voltage drop is then Vo = (CPm - cps)/e. We shall
assume that n(xo) is not influenced by the current flow resulting from an
external field, although this is only approximately true. The first boundary
condition leads to C = n(O) - 1/fleE. Substituting this into (14-7)
and applying the second boundary condition in the form (14-8), one
obtains for the current density,
I
fleE
n(O)eeExo/kT - n(xo)
eeExo!kT _ I
(14-9)
[fle(Vo -
1) (14-10)
(14-11)
354
[Chap. 14
the same meaning as above. The electronic current density from the
metal to the semiconductor is obtained by putting V = 0 in expression
(14-11), since for V = 0 the two currents are equal and opposite. The
resultant current is then
. _\
' i'
leV)
mn(xo)e(v)(eeV/kT -
l)
A'(eeV/kT -
I)
(14-12)
Comparing this with (14-10), it is observed that the form of the two
expressions is essentially the same. However, (v) may be considerably
smaller than pE, leading to A' ~ A. For example, at room temperature
(v) :::: 10 7 cm sec l ; for a: barrier of 10- 6 cm and a voltage drop of 1 eV,
E:::: 106 volts per cm, and with p:::: 103 cm per volt sec, we obtain
pE:::: 109 cm sec-I. It should be remarked here that V represents the
voltage across the barrier, i.e., it is equal to the applied voltage minus
the voltage drop across the bulk semiconductor.
Although the diode theory has been applied in the past to interpret
the rectifying properties of germanium and silicon diodes, a number of
observations remained unexplained. For example, according to the
above theory, the magnitude of the currents should depend strongly on
the work function of the metal because n(xo) is proportional to exp
(-eVolkT) and Vo is determined by (4),,, - 4>J. Thus, for a variation of
0.5 ev in the work function of the metals used, the currents should vary
by a factor of ,..._,10 8 . Experiments indicate variations by a factor of 10
or less for different metal points. The origin of this discrepancy will
be discussed in the next section.
14-4,.:surface states on semiconductors
",;..;;..'
4"
,;, '
Sec. 14-4]
355
material. 9 Such states, which may lie within the normally forbidden
region, may arise partly as a consequence of the sudden departure from
periodicity of the potential at the surface or in part from adsorbed atoms.
In other words, the simple band picture which one normally employs
for the bulk properties is in general not applicable close to the surface.
Thus a certain number of these surface states may be occupied without
giving rise to an excess surface charge. When the material is placed
opposite a positively charged metal, more surface states may be filled by
the induced charge.
ii'
1 ,
:;<'
<-:-"h
'>F ;,;
Condo band
E.-r------
--+------_.._- Fermi
H+---::----'+l
level
.~
:;/-<"
la)
(b)
~"'~'
356
[Chap. 14
Xo
(l4-14)
(I4-15)
We note that for very small values of n." the voltage drop Vo is very
small because a small number of extra electrons will bring the Fermi
level at the surface up to that of the bulk material. For very large values
of n." Vo becomes approximately equal to Es/e. According to Sno~kley
and Pearson, n, is of the order of 5 X 1013 cm- 2 volts-l.
From what has been said above it follows that due to the presenc~
of surface states, a layer of depleted conductivity is formed below the
surface. Under these circumstances, the space charge layer is a property
of the material itself and not particularly sensitive to the work function
of a metal which may be brought in contact with the surface. This
argument has been used by Bardeen to explain the fact that the properties
of point-contact germanium rectifiers are rather insensitive to the metal
used. 10 The presence of surface states also playa role in the interpretation
of contact potential measurements across n-p junctions. l l
14-5. The two-carrier theory of rectification
In Sec. 14-3 we have seen that if V is the applied voltage in the forward
direction the forward current of a rectifying contact should be given by
$-
1= A[e,,(V-II) -
I]
(14-16)
Sec. 14-5]
357
will become p-type. Suppose now that the semiconductor is made negative
relative to a metal in contact with it. Thi's will lead to an Increase in
the electronic current from the semiconductor into the metal but at the
same time a hole current will begin
to flow from the surface into
the semiconductor. In other words,
two types of carriers contribute to
I
I
the current. This has the effect of
p-type 1
n-type
t+--.:....::-i-----=-:.._---+- Fenni
decreasing the apparent value of r
level
in (14-16), because r is based on
electronic conductivity only. Because
of lack of space it is not possible
to discuss the quantitative aspects
of the two-carrier theory for point
contacts hereP It will be evident
that this model also explains the
Fig. 14-6. As a result of surface states,
hole injection into n-type material the valence band at the surface lies
by metals biased in the forward close to the Fermi level; this gives rise
direction, as referred to in Sec. 14-6. to a thin p-type layer near the surface of
n-type material.
. ...
358
[Chap. 14
(a)
:r:
cond ' band
(b)
Fermi
level
Fermi'
level
Valence band
~.
n to p (Fig. 14-7d). The same is true for the equilibrium electron current
Ic o' This implies that there must be a certain concentration of holes
in the n-region as well as a certain concentration of electrons in the
p-region. This is a result of the continuous thermal creation of electronhole pairs, the creation being compensated by recombination. For example, if g is the rate of production of pairs and nco and n". are the equilibrium concentrations anywhere, we must have rn,> nil = g. where ,.
is the recombination coefficient. Thus in either regio~, n,> onil II = glr.
which is constant at a given temperature. If one assumes that the ratio
g!r is independent of the donor or acceptor concentration. g!r must also
Sec. 14-6]
359
,.,...
~_f
""-,,,t
ma/cm 2
_rl~.
;'):}d
}t!!
ni
from p to n have to climb a smaller potential hill (Fig. 14-7d) and will
give rise to a current Ih exp (eV/kT). Hence
fl.') r~i$ 1}Hri (H '~:J.D:~;,>
f)
(14-] 8)
For the electron current one finds a similar expression and the total
forward current across the junction should be
(14-19)
For positive voltages applied to the n-type region the reverse current is
obtained,
I = (I + I )(1 - e-eV1k7')
(14-19a)
r
"0
eo
In Fig. ]4-8 we have represented the experimental points obtained for a,
p-n junction characteristic ;14 the fully drawn curve is the theoretical
360
[Chap. 14
;->-i-
I)
(14-24)
,\.1>,
Sec. 14-6]
361
(Fig. 14-7d). For electrons in Ge, the diffusion constant D ':::' 100 cm2/sec
and a typical lifetime is 10-4 sec. This gives a diffusion length of the order
of 0.1 cm.
The correctness of the diffusion theory for the rectifying p-n junction
has been tested further by using junctions as a photoconductive device.
For example, let photons of sufficient energy to create electron-hole
L,
Emit.
a:l
'l:)
f"',
. n.
'{'Y"
i t . ' ",(
1'l~
>"'14
'1
\ ,
n
,.
J(,~,,>._
, ,
~
p..,
-Distance
(bl
pairs be incident on the n-type region at a distance x from x _:_ O. According to what has been said above, the current response should vary as
exp (-x/L) and this has indeed been verified experimentally by Goucher
and coworkers. I5 Also, the value of L so obtained is consistent with the
one required by the rectifier equation (14-24). When the light is incident
at the junction itself, the electron and hole are separated by the strong
field at the junction and a current of one electron per absorbed photon
may be obtained.
14-7. Transistors
An n-p-n junction transistor is built up of two n-type regions separated
by a thin layer of weakly p-type material. It is mainly this type of transistor
which will be discussed below. The same reasoning applies to p-n-p
15
,.
362
[Chap. 14
(14-26)
where V is the applied voltage between base and emitter and nho is the
equilibrium concentration of holes in the emitter region. Because of
what follows it is important to realize that I" is determined by the diffusion
length of holes in the emitter region. In the same way, the electronic
current from emitter to base is determined by the diffusion length of
electrons in the base region. However, because the width of the base
region is W ~ Le, one should usc W rather than Le for the electronic
current. Hence
(l4-26a)
Sec. 14-7]
363
+
Base
Fig. 14-11.
A point-contact transistor.
'e =
(kT/el)
where I is the emitter current. For 1= 1 rna, this gives at room temperal
ture, f< = 25 ohms.
364
[Chap. 14
:-S
71
r,
rb
r,
2S
200
S x 10"
oc
0.95--0.99
POillt-COllt{/CI
125
75
2 x
2-3
IO~
"
RE'FERENCES
J. Bardeen and W. H. Brattain, Phys. Rel'., 75, 1208 (1949); Bell System
Tech. J., 28, 239 (1949).
J. S. Blakemore, A. E. De Barr, and J. B. Gunn, "Semiconductor circuit
Elements," Repts. Progr. Phys., 16, 160 (1953).
H. K. Henisch, Metal Rectifiers, Oxford, New York, (1949).
Proc. IRE., 40, (l952) (transistor issue).
W. Shockley, Electrons and Holes in Semiconductors, Van Nostrand,
New York, 1950.
H. C. Torrey and C. A. Whitmer, Crystal Rectifiers, McGraw-Hill,
New York, 1948.
Chap. 14]
365
PROBLEMS.,
14-1. Two metallic surfaces with work functions of 3 and 4 ev are
separated by a gap of 10 A. Calculate the surface charge density in
equilibrium at room temperature in terms of a number of electrons.
14-2. A metal with a work function of 3 ev is in contact with a semiconductor with an electron affinity of 1 ev; the semiconductor contains
1016 donors per cm3 close to the conduction band. Calculate the capacitance of the barrier layer per cm2 for zero applied voltage, when the
dielectric constant is 12. Do the same problem for a reverse bias of 5
volts.
14-3. Consider a block of semiconducting material with a large area
contact on one of its faces; the opposite face has a small circular point
contact of radius a. Show that the bulk or spreading resistance of the
system is r = 1/4aa, where a is the conductivity of the semiconductor.
14-4. Consider an idealized p-n junction in which the acceptor concentration is constant for x < 0, and the donor concentration is constant
for x > o. Find an expression for the barrier thickness in terms of the
acceptor and donor concentrations and the forbidden energy gap; assume
that the donor and acceptor levels lie very close, respectively, to the
conduction and valence bands. Also discuss the variation of the barrier
thickness with an applied voltage.
14-5. Repeat Problem 14-4 for a junction consisting of a p-type region
containing N acceptors per cm3 and an n-type region containing N donors
per cm3 , the two regions being separated by a transition region in which
the concentrations vary linearly with x. Assume that the transition region
is large compared with the physical barrier layer.
14-6. Consider a p-n junction with an area of 0.25 cm 2 in which the
current is carried mainly by holes. Given that for small forward voltages
the junction resistance is 800 ohms, calculate the density of holes in the
n-region if the life time of the holes in this region is 10- 4 sec and their
mobility is 1800 cm 2 volt- I sec-I.
f,)
Chapter 1!j
ELECTRONIC PROPERTIES OF
ALKALI HALIDES
In the class of solids which may be referred to as ionic semiconductors,
the alkali halides have been investigated more thoroughly than any other
group. In this respect they occupy a place similar to Si and Ge in the class
of nonpolar semiconductors. The reason for this is twofold: in the first
place large single crystals of the alkali halides may be grown with relative
ease; second, they have a simple structure. In the present chapter some
of the most outstanding electronic properties of these materials will be discussed. This discussion is necessarily very incomplete, and for further
details we must refer the reader to the bibliography at the end of this
chapter.
15-1. Optical and thermal electronic excitation in ionic crystals
In this chapter much of the dischssion will be devoted to the excitation
of the electronic system of ionic crystals. Electronic excitation may be
accomplished in various ways:
I
366
Sec. 15-1]
367
E
B
Fig. 15-1.
_r
Y atom. In general, the minima Band Q will not correspond to the same
separation. Suppose now the molecule XY is in the state B. It may then
absorb a photon hv, after which its representative point will arrive at B',
in accordance with the Franck-Condon principle (the separation between
the nuclei does not change during the transition). However, point B'
represents a highly excited vibrational state of the XY* molecule and
ultimately the system will move toward a point on the PQR curve near Q,
by energy exchange with the surroundings. In other words, some of the
optically absorbed energy is wasted in the sense that after the optical
transition a certain fraction of it is transformed into heat. The thermal
activation energy is simply given by the energy difference between Band
Q; evidently the optical activation energy is always larger than or equal
to the thermal activation energy. This was first pointed out by de Boer
and van GeeJ.2
We should note that actually the atoms in the XY molecule vibrate ,
relative to each other, even at T = O. Thus, let the molecule XY be in the
vibrational level DE" Depending on where the representative point finds
itself along the DE level, the optical excitation energy may lie anywhere in
2
368
[Chap. 15
the shaded region of Fig. 15- 1. Thus the absorption spectrum is a band
spectrum in contrast with the line spectra observed for single atoms. The
band width W will increase with increasing temperature.
The same reasoning may be applied to an electronic transition in a
solid by interpreting the coordinate r as representing the configuration of
the nuclei in the vicinity of the position in the crystal where the transition
takes place. It follows from the above discussion that when one introduces
an energy-level diagram, it is necessary to specify whether one is talking
about optical or thermal transitions, because the diagrams will in general
be different for the two cases.
To iIIustrate the role of the static and high-frequency dielectric constant
in the case of optical and thermal electronic transitions in ionic crystals,
consider the following much simplified model. 3 Suppose an electron is
trapped in an ionic crystal by all impurity of a certain radius R, the charge
of the impurity being e. Let the electron be removed optically from the
impurity to a point in the lattice far fro-m the impurity. Since only the
electronic displacements are able to follow the optical electronic transition,
the field about the impurity immediately after the transition is equal to
El = e/Eor2. After a time interval equal to a few times the period associated
with lattice vibrations has passed, the ions have adjusted themselves to the
new situation, and the field ultimately drops to E2 = e/Esr2. The process
of adjustment of the ions corresponds to the motion of the representative
point in Fig. 15-1 from B' to Q. The energy given off by the system during
this process may be estimated as follows. The energy per unit volume in
the dielectric is given by (ED)/87T. In our example E and D are both radial
vectors; hence, per unit volume there is a change in energy .equal to
l
(,
"
Sec. 15-2]
369
15-2. The upper filled band and the conduction band in ionic crystals
In the present section it will be assumed that the crystals under consideration are perfect in the sense that they are of stoichiometric composition and that they do not contain lattice defects of any kind. Although
such crystals do not actually exist, it is useful to consider the properties of
an idealized model as a starting point; the influence of lattice defects on
the electronic properties will be considered later.
Ionic crystals such as the alkali and silver halides, the oxides of the
alkaline earth metals, etc. are usually good insulators. Thus, according to
E
i
ClNa+
Cond. band
3p (occupied)
t=:::::><
3s (empty)
'iL
370
[Chap. I~
where Ec(r) is the Coulomb energy of one ion in the field of all others;
according to Sec. 5-2, Ec(r) = Ae2 /r, where A is the Madelung constant
(1.75 for the NaCl lattice). The reason for the occurrence of the Ec(r)
terms is the following. When an electron has been taken from a CI- ion,
the Coulomb interaction is lost, because the resulting atom is neutral.
Hence, to remove an electron from a Cl- ion in the lattice one requires an
energy ECI + Ec(r). In a similar way, the energy gained by putting an
electron on a Na+ ion is IN" - Ec(r) because again the Coulomb energy is
lost. Thus, as r decreases, FcCr) increases, leading to the variation of the
energy levels as indicated; for a certain value of r they cross over, and for
still smaller values of r the occupied 3p levels fall below the empty 3s levels.
Actually, the Ec(r) terms should be corrected for polarization effects.
"",Thus, when an electron is excited optically from a certain Cl- ion to a
remote Na-i ion, one should subtract from (15-2) the polarization energy
around the neutral CI and Na atoms, using the high-frequency dielectric
constant EO' If the excitation is thermal, the polarization energy must be
calculated on the basis of the static dielectric constant Ea' Here again the
thermal excitation energy will be smaller than the optical excitation energy.
Hence one arrives at two possible electron level schemes: an optical one
and a thermal one.
When the lattice parameter r reaches values such that the wave functions
of neighboring ions of equal sign begin to overlap, the discrete atomic
levels broaden into bands (see Chapter 10). The actually observed lattice
parameter r = a determines the distance between the bands as well as the
widths of the bands, as indicated on the right in Fig. 15-2.
One thus arrives at the conclus'ion that the upper filled band in NaCI is
associated with the occupied 3p levels of the CI- ions, the empty band
corresponding to the unoccupied 35 levels of the Na+ ions. Similar
Sec.15-2}
371
identifications may be made in other ionic crystals. Actually, the conduction band in ionic crystals probably"corresponds to an ionization
continuum, i.e., other empty bands will overlap with the one identified
above.
Information about the width of the upper filled band may be obtained
from soft X-ray emission spectra, as explained in Sec. 10-12. Results of
such studies for a number of ionic and semi-ionic crystals are given in .
Table 15-1. 5 It is observed that the band width for the alkali halides and
Table 15-1. The Width of the Upper Filled Band in ev
for a Number of Solids
LiF
NaF
KF
2.1
1.7
1.5.
LiBr
NaBr
KBr
RbBr
AgBr
1.2
0.75
0.55
0.45
1.1
Li,O
CaO
SrO
BaO
12.8
10.8
9.2
8.4
silver halides is of the order of I ev; for the oxides it is of the order of
JO ev. From this one might expect the effective mass of ~ hole to be larger
in the halides than in the oxides.
The energy difference between an electron at rest in vacuum and an
electron at the bottom of the conduction band is called the electron
affinity X of the crystal (see Fig. 15-2), For alkali halides r. is probably of
the order of 0.5 ev or less.
15-3. The ultraviolet spectrum of the alkali halides; excitons
A great deal of experimental information about the electronic structure
of ionic crystals has been obtained from optical absorption measurements. 6
As an example we give in Fig. 15-3 the absorption spectrum of KBr.7 The
alkali halides are transparent in the visible region of the spectrum, and the
absorption spectrum associated with electronic transitions lies entirely in
the ultraviolet. It consists of a number of absorption peaks which are best
resolved at low temperatures. In the vicinity of the peaks, the absorption
coefficient is of the order of 10 6 per cm, so that thin evaporated layers are
used in these experiments. The high energy region is very difficult to
investigate experimentally so that virtually nothing is known about the
5 N. F. Mott and R. W. Gurney, op. cit., pr. 75, 79.
-,
For a review, see R. W. Pohl, Physik. z., 39, 36 (1938); E. G. Schneider and H. M.
O'Bryan, Phys. Rev., 51, 293 (1937); L. Apker and E. Taft, Phys. Rev., 81, 698 (1951).
New studies, with emphasis on the range from 950 A to 1700 A are presently being
carried out at Cornell University by Hartman, Siegfried, and Nelson.
, R. Hilsch and R. W. Pohl, Z. PhYSik, 59,817 (1929).
,72
[Chap. 15
short wavelength tail of the absorption region. The tail on the longwavelength side is at least in part due to imperfections, as will be explained
further in Sec. 15-5; it is therefore
temperature-sensitive.
It is important to note that when
photons are absorbed in the long wavelength tailor in the first absorption peak,
no photoconductivity results. This indicates that the first absorption peak does
not give rise to an electronic transitior.
from the filled band into the conduction
band. It is believed, therefore, that the
first absorption peak gives rise to an
excited state of the halogen ions, i.e., an
electron from the filled band is raised to
a level below the conduction band. This
1"---_
220 ml' situation may be compared with the excited
160
180
200
_A
states of an atom. Complete ionization
of an atom may then be compared
Fig. 15-3. Absorption spectrum
with
the transition of an electron from
of pure KBr. [After Bilsch and
the filled band into the conduction band.
Pohl, ref. 7)
It is for this reason that the energy
band scheme of a perfect ionic crystal contains a number of narrow
"exciton bands" below the conduction band, as indicated in Fig. 15-4.
One may also look at this by saying that when a photon corresponding
to the first absorption peak is absorbed by an electron in the filled band,
the excited electron is still bound to some extent by the Coulomb field
produced by the hole it left behind. This will be further illustrated in the
... ;'.~xt section. The combination of an electron in an excited state and the
~ssociated hole is called an exciton; the unit as a whole is neutral. It has
been suggested that an exciton may be thought of as resulting from an
electronic transition from a negative ion to a nearest neighbor positive ion.
A transition from the filled band to the conduction band in this type of
picture then corresponds to an electron transfer from a negative ion to a
faraway positive ion. This is probably an oversimplification, although it
illustrates the electron-hole interaction.
At present it is not known where the ultraviolet absorption spectrum
goes over from exciton bands into the ionization continuum. A rough
estimate may be made on the basis of the following simplified model.
Consider the exciton as an electron and a hole revolving about each other
as a result of Coulomb interaction. The energy of this system may be
estimated from the analogy with a hydrogen atom in the ground state.
Two modifications are required: (1) the mass of the electron is approximately equal to that of the hole; (2) the field between the two particles is
I \
Sec. 15-3]
373
ev
LiCI
NaCI
KCl
RbCl
CsCl
1430
1580
1620
1660
1620
8.6
7.8
7.6
7.4
7.6
Salt
ev
NaBr
KBr
RbBr
CsBr
1900
1890
1930
1870
6.5
6.5
6.4
6.6
Salt
ev
.~
KI
RbI
2200
2230
5.6
5.5
374
[Chap. 15
by Frenkel 8 and by Slater and Shockley 9 the possibility of exciton propagation was related to the overlapping of the excited-state wave functions
on neighboring atoms. More recently, Heller and Marcus have shown that
even if the overlap is small, the propagation of excitons may be good as a
result of the dipole coupling between the excited ion and a neighboring
identical atom .in the ground state_Io Thus an exciton may be represented
as a neutral "particle" with a certain effective mass m*, its motion being
characterized by a wave vector k. If the exciton is produced by the absorption of a photon, the initial wave vector of the exciton will be the same as
that of the incident photon (conservation of momentum). In this case, the
dipole moment of the exciton will be perpendicular to the direction of
propagation. As the exciton propagates, it may interact with lattice
vibrations or imperfections, and scattering results. Excitons for which the
dipole moment is parallel to the direction of propagation are also possible
and the two types can be converted into each other by scattering processes.
The optical lifetime T of an exciton, i.e., the average period elapsing between
the production of the exciton and the instant that the electron and hole
recombine under emission of a photon, is probably of the order of 10-8
second (corresponding to the emission of dipole radiation). However,
long before the optical lifetime is up, the exciton may give off its
energy to lattice imperfections. For example, an exciton may transfer
its energy to an F center, thereby raising the trapped electron into
the conduction band. Excitons may also transfer their energy to
an electron in the filled band in the vicinity of a negative ion vacancy,
leading to the production of an F center. These matters will be further
discussed in Sec. 15-9. According to Marcus and Heller, the effective
mass of an exciton is given by
m*
m(31T/4n)I/3/".Inaae
(15-3)
where n is the number of ions per unit volume, m is the free electron mass,
a. is the effective Bohr radius of the excited atom, and In is the oscillator
strength connecting the ground state and the excited state." When In. ~ 1,
as it is in the alkali halides, it follows from (15-3) that m* ~ m. Hence
at room temperature the velocity of an exciton is approximately equal to
that of a thermal electron, i.e., V ~ 10 7 cm/sec. During the optical lifetime
the total path covered by an exciton is thus of the order of 10 7 X 10-8
~ 0.1 cm. As a result of the scattering by phonons and imperfections, the
path is actually curled up as in any type of Brownian motion. If one
8 J. Frenkel, Phys. Rev., 37, 17, 1276 (1931); Phys. Z. Sowjetunion, 9, 158 (1936);
see also G. H. Wannier, Phys. Rev., 52, 191 (1937).
9 J. C. Slater and W. Shockley. Phys. Rev., 50, 705 (1936).
). W. R. Heller and A. Marcus. Phys. Rev . 84, 809 (1951).
Sec. 15-3]
375
assumes a mean free path for scattering A. -::: 10 Angstroms, one finds for
the mean square displacement
(r2)
= 2DT = iA.VT,-...J
10-8 cm2
(15-4)
Thus, during the optical life, the excitons may on the average undergo a
displacement of approximately 10-4 cm relative to point of origin.
15-4. Illustration of electron-hole interaction in single ions
In view of the importance of the concept of an exciton, it maybe useful
to consider the following example of the Coulomb interaction between an
electron and a hole. To remove the 3s electron in a free sodium atom, the
ionization energy (5.12 ev) is required. To remove a second electron,
another 47 ev is required, as illustrated in Fig. 15-5a. At first sight one
0------
-5.1--jl.- - 3s
E(ev)
-,T 3s . ':
~..,
-47-_'_-2p
(al
(bl
,.;
1../,,:1
376
[Chap. 15
+
+
+
+
_g_B
+
+
A
-A OAA
+
Condo band
B
+ BOB +
B
+
+
A....L.
represented by a short bar, and it is understood that this implies that the
level is localized in the vicinity of a particular vacancy. Each positive ion
vacancy gives rise to six of these levels because there are six Cl- ions
surrounding the vacancy. To a lesser extent, similar considerations hold
for the next-nearest Cl- ions; the levels for the 3p electrons on these ions
will be much closer to the filled band.
It may be noted that if in some way or other a free hole should be
created in the filled band, the hole may be trapped at one of the A levels
mentioned above. This is not surprising, because a position where a
positive ion is missing would be a favorable site for a positive hole to
reside. The trapped hole may then be represented as a Cl atom neighboring
a positive ion vacancy. Actually, the hole will probably be shared by the
six surrounding halogen ions, because they are all equivalent. Trapped
holes of this kind are called V centers; these will be discussed in Sec. 15-12.
The situation around a negative ion vacancy may be discussed in a
similar way. Suppose by some means one had created a free electron in the
conduction band. A likely place for this electron to get trapped would be
a negative ion vacancy; the latter has an effective positive charge and thus
attracts the electron. In this case the electron is shared by six surrounding
Na+ ions and the resulting center is called an F center. In the energy level
Sec. 15-5]
377
scheme this means that the empty 3s states of the Na+ ions surrounding a
negative ion vacancy (8 in Fig. 15-6) do not lie in the conduction band but
below it.
Summarizing, we may say, that positive ion vacancies give rise to occupied electronic levels above the filled band; negative ion vacancies give
rise to unoccupied sta,tes below the conduction band. It will be evident
that as a result of these changes in the energy level scheme, new absorption
bands will arise, the extent of the absorption being proportional to the
density of lattice defects. More complicated lattice defects, such as pairs,
triplets, etc., will also change the absorption spectrum. Usually the new
absorption bands lie in the tail of the first fundamental absorption band.
Thus, although changes in the tail may be observed, for example, by
variations in temperature, it is difficult to resolve the new bands. At low
temperatures, however, one has observed the so-called (1. band, which is
believed to be associated with the presence of single negative-ion
vacancies. 11
15-6. Nonstoichiometric crystals containing excess metal
A great deal of fundamental information about semiconducting ionic
crystals has been obtained by studying the properties of nonstoichiometric
crystals, i.e., crystals containing an excess of one of their constituents.
For example, when an alkali halide crystal is heated in the vapor of its
metallic constituent, an excess of metal is incorporated in the crystal.
Some properties resulting from the excess metal will now be discussed.
F centers. In the first place, crystals heated in the metal vapor and
quenched to room temperature show an absorption band in the visible or
ultraviolet, whereas the original crystals were transparent in that region.
This absorption band is called the F band (the German word for color is'
Farbe). As an example we show in Fig. 15-7 the Fband in KBr at various
temperatures according to Mollwo. 12 The width of the band increases
and its position shifts to lower energies when the crystals are heated. At
room temperature the position of the F band peak in the alkali halides is
as given in Table 15-3. It is interesting to note that according to Mollw012
the F-absorption frequency VEt' is related to the shortest interionic distance
a by the approximate expression
(15-5)
11 Delbecq, Pringsheim, and Yuster, J. Chern. Phys., 19, 574 (1951); 20,746 (1952);
see also W. Martienssen. Z. Physik, 131,488 (1952); W. Martienssen and'R. W. Pohl,
Z. Physik, 133, 153 (1952).
10 E. Mollwo, Z. Physik, 85, 56, 62 (1933); the fact that 1', is essentially determined
hy a and only very slightly by the dielectric constant of the material may be explained
on the basis of calculations by L. Pincher!e, Proc. Phys. Soc. (London), 64, 648 (1951).
378
[Chap. 15
1017 (n +
n 2)2 KIllilX H per cm'3
2
Abs.
1.0
( 15-6)
4t.
5.0
3.1
2.7
NaF
NaCI
NaBr
3.6
2.7
2.3
KF
KCI
KBr
KI
2.6
2.2
2.0
1.8
RbCI
RbBr
Rbi
2.0
1.8
1.6
CsCI
2.0
13 See R. Hilschand R. W. Pohl,Z. Physik, 68, 721 (1931); F. Seitz. Revs. Mod. Phys.,
18, 384 (1946); 26, 7 (1954).
14 F. G. Kleinschrod. Ann. PhYSik, 27, 97 (1936).
15 See, for example, F. Seitz, Revs. M?d. Phys . 26, 7 (1954).
Sec. 15-6]
379
'
i 'I
,)"
"~,:,"'.-'
10
10 17
ir..
'"
10 16
,./.
_ nv
1016
1
1.6
10 17
/~
1.4
1.2
-+-
1.0
lOOOIT
Fig.
,... ..1
(II F )
F center
kT
(l5-7)
where 1> is the e~ergy required to take an atom from the vapor and incorporate it as an F center in th.e crystal. In Fig. 15-9 we h.ave plotted nF1nv
as function of T-1 as given by Rogener for KBr and KC[.l6 From the
slopes it follows that
for K in KBr
1>
= -0.25 ev
1>
~-=
-0.10 ev
Note that in both cases 1> is negative, i.e., energy is released by taking an
atom from the vapor into the crystal. It is observed that n F > n~ for these
crystals.
,. H, Rogener, A 1111, Physik, 29, 386 (1937),
380
[Chap. 15
:;(: :
-
Sec. 15-6]
38 I
T-
382
[Chap. 15
KCI
RbCI
KBr
KI
1.8 Jl
Sec.. 15-6]
383
because otherwise the electron will not remain in the excited state but
will be further excited thermally into the conduction band. The luminescence of F centers in additively colored alkali halides has r 'ently been
observed by Botden, van Doorn, and Haven. 26 The crystals were,; ,diated
with light in the F band and luminescence in the infrared was ot ::rved at
200 K and at nOK. In Fig. 15-12 we
reproduce the emission spectra at
9
F
20oK. The energy of the emitted
photons at 20 K are given in ev below,
together with the corresponding
A
'e 6
~
photon energy for F absorption at
:.c:
room temperature. It is observed
that the absorption frequency is
3
nearly twice the emission frequency,
/" \
,
I
and this difference would actually be
F'
/
---------.. . -!!
larger if the absorption energies were
2.5
3
2
referred to 20oK. This illustrates
1.5 ev
again clearly the importance of the
Fig. 15-13. Formation of the F' band
Franck-Condon principle in ionic
at the expense of the F band, by irracrystals. After the optical excitation diation of an additively colored crystal
has taken place, the ions in the vicinity "'ith F light at 173K. [After Rilsch
and Pohl, ref. 271
of the excited F center will adjust
themselves to the new charge distribution, thereby giving off energy (corresponding to the representative
point moving from 8' to Q in Fig. 15-1). When the electron returns to
the ground state after this ionic displacement has taken place, an infrared
quantum is. emitted.
,i
c
.
KCI
RbCI
KBr
KI
Absorption
Emission
2.2
2.0
2.0
1.8
1.25
1.10
0.96
0.85
15-7. The transformation of F centers into F' centers and vice versa
When an additively colored crystal containing F centers is irradiated
with light in the F band, a new band appears at the long-wavelength .>ide
of the F band. The new band grows at the expense of the F band and is
called the F' band. For example in Fig. 15-13, curve A represents the
26 Th. P. J. Botden, C. Z. van Doorn, and Y. Haven. Philips Research Rep/s . 9, 469
(1954).
384
[Chap. 15
Sec. 15-7]
385
The rather sudden drop in the quantum yield in the beginning of the
irradiation with F light at temperatures below 1400 K is explained as
follows: an F center corresponds to an electron trapped at a negative ion
vacancy, so that for large distances the electron moves in an electric field
-e2 /Eor 2 , where EO is the high-frequency dielectric constant. 29 As in a
hydrogen atom, there must therefore be a number of excited states below
the ionization continuum, Le., below the conduction band. Mott assumes
that the absorption of an F quantum raises the electron from the ground
...L!.-F'
+hVFF--'-
F--
F--Empty
state to the first excited state, which is close to but not in the conduction
band (Fig. IS-II). At temperatures above 1400 K the thermal lattice
vibrations are intense enough to provide the additional energy required
to raise the electron from the excited state into the conduction band, but
at low temperatures the probability for the electron to fall back to the
ground state takes over. Hence, at low temperatures, absorption of an
F quantum does not liberate the electron, leading to the drop in quantum
yield of the reaction 2F -+ F'.
The decrease in quantum yield as the time of irradiation increases is a
result of the increase in the number of negative ion vacancies; this increases
the probability for an electron in the conduction band to be trapped by a
negative ion vacancy.
Another set of interesting data has been obtained by Pick employing an
additively colored KCl crystal in which 80 per cent of the F centers had
been transformed into F' centers. In such crystals one can study the
reverse process viz., the transformation of F' centers into F centers by
irradiating with F' light. In Fig. 15-16 we reproduce Markham's representation of Pick's data for the number of F centers formed as function of the
number of F' photons absorbed. 28 It is observed that at low temperatures
up to about 90 0 K two F centers are formed per absorbed F' photon, at
least in the beginning of the irradiation. This is in agreement with the model
of an F' center discussed above: an F' center from which an electron is
released is transformed into an F center; the free electron captured by a
negative ion vacancy produces the second F center. The drop in the
386
[Chap. 15
for a free electron to form an F center from a negative ion vacancy decreases.
The decrease of the quantum yield as the irradiation proceeds is a consequence of the increase in the number of F centers and the decrease in the
number of negative ion vacancies.
15-8. Photoconductivity in crystals containing excess metal
We have seen above that absorption of an F photon by an F center
produces a free electron if the temperature is not too low. Thus when a
crystal containing F centers is irradiated with F light and at the same time
an electric field is applied to the crystal, a photocurrent is observed.
However, if only the electrons are mobile. a space charge will soon be
built up in the crystal, thereby lowering the field and the current. The space
charge may be neutralized by electrons entering the crystal from the anode.
or by electrolytic conduction in the crystal. If this is not the case, space
charge difficulties may be avoided by employing low light intensities for a
short period. Before discussing briefly some classical experiments by
Pohl's group it may be useful to make some general remarks on the
Sec. 15-8]
387
(15-8)
unless the electron has arrived at the
anode before being trapped. Suppose
the electron is liberated at a distance
Xo from the anode, and let L be the
distance between the electrodes (Fig.
15-17). The charge passing through
.....- - L
( 15-9)
= rJ N "e/2
(15-100)'
1t seems that for alkali halides saturation occurs only for crystals that are
too thin to be used in photoconductivity experiments. Saturation has been
observed, however, in the silver halides. in zinc sulfide, and in diamond.
388
[Chap. 15
:--A
40
J:I.
I
I
I
20
F.light
B
I
I
I
I
....X
E
::s
<.>
WL[l
~
t
F-light
I
I
I
I
I
30'C
'"'0
....
C---t
Dark
SO'C
ts;
'_t,
40
_\
'J:)-" -_,:"
. ',,-
125'C
20
10
15
20
25 sec
Fig. 15-18. Time dependence of the photocurrent in NaCl containing 8 x 10 15 excess Na atoms per em', at various temperatures_
[After Glaser and Lehfeldt, ref. 30]
'0 G. Glaser and W. Lehfeldt, Nachr. Akad. Wiss. Gottingell, 2, 91 (1936); see also
R. W. Pohl, Physik. z., 39,36 (1938).
"c:
. .
Sec. 15-8]
389
interval A. This is a true primary current for which relation (15-9) holds.
Thus the quantity 'Y/x/E should be independent of the field strength, in
agreement with the experiments. From what has been said in the preceding sectioo one expects F' centers to be formed during the irradiation
with F light. That this is indeed the case may be seen from the intervals C,
which also indicate a larger production of F' centers at lower temperatures,
,
1+-")
","-
10- 9
..~--
> 10- 10
..
10- 11
10-12
...
Ii
10- 13
-200
-150
-100
-
Fig. 15-19.
-50
OC
Temp.
KCI crystal containing 2.7 x 1016 F cente'rs per cm'. [After Pohl,
ref. 301
increase in the effective displacement above room temperature is a consequence of thermal excitation of electrons from F' centers. The region
below _. 180C probably corresponds to photoelectrons liberated from'
colloidal sodium particles.
Another result of great interest which confirms the conclusions of the
preceding section is represented in Fig. 15-20 where x/ E for KCl has been
plotted as function of the density of F centers n F for a fixed temperature
390
[Chap. 15
10- 11
Fig. 15-20.
t x/E
0,
(m 2 /volts)
I
,J,"
\.-----
traps for free electrons. We leave it to the reader to show that from Fig.
15-20 one may estimate a capture cross section of about 1O-lL lO-15 cm2 .
15-9. The photoelectric effect in alkali halides
Although the photoelectric effect of insulators has not been studied
very thoroughly, some recent experiments by Apker and Taft31 on alkali
halides show that such investigations can provide useful information about
the behavior of imperfections in crystals. As in the study of this effect
from metals, one can measure the number of photoelectrons emitted per
incident quantum as function of the frequency of the light employed, and
the energy distribution of the emitted electrons. In pure alkali halides, the
energy required to produce a free electron in the conduction band is of
the order of 8 ev, and in order to observe the photoelectric effect, photons
of an energy of this order of magnitude are required. On the other hand,
if F centers are present in the crystal, i.e., electrons trapped in levels about
2 ev below the conduction band, one expects an appreciably lower
threshold frequency. That this is indeed the case may be seen from Fig.
) 5-21, representing the photoelectric yield in electrons per quantum for
potassium iodide containing F centers. For hv:::: 2.3 ev, about 10-8
31
For a review of this work see the article by Apker and Taft in W. Shockley (ed.),
Sec. 15-9]
391
electron is emitted per incident quantum, but the curve rises rapidly to a
plateau with a yield of 10-4 . The sample ~~ntained about 1019 F centers
per cm3 produced by electron bombardment. The rise beyond the plateau
is interpreted as follows. The first fundamental absorption peak of KJ
occurs at hv = 5.66 ev at room temperature. Thus photons of 5 ev
3000 K
--+- hv lev)
392
[Chap. 15
are formed. Thus, when a crystal which initially contains no color centers
is irradiated with light in the first fundamental absorption band, the
excitons produce F centers which may thereupon be ionized' by other
excitons. One thus expects a build up of the F-center concentration as
function of time, and associated with this, an increase in the photoemission
current. This is shown in Fig. 15-22 according to Apker and Taft. 33 In
the same figure, the decay of the photoemission resulting from heating the
crystal is represented.
:
...,
I')
<lI
!::::>
()
....00
..<::
p..
t
o
10
30
20
-Time
40 min
I
r"
Sec. 15-10)
393
.5
2 ev
394
[Chap. 15
hole. These holes are most likely to be found near a positive ion vacancy
where they can be trapped. The optical absorption associated with a
trapped hole may be, for example, the transition of an electron from the
filled band into the hole. A hole trapped at a positive ion vacancy is called
a VI center. There is good evidence to believe, however, that the dominant
peak observed by Mollwo is not of this simple type. The reason for this is
the following. According to Mollwo's experiments the saturation density
+
+
+
Ol
~
Vz
+
+
+
+
"[fJV:1
~F
+LJ+~+
- r=olVi
L_j -
~~~@~C!J:
+
Sec. 15-13)
395
not have sufficient momentum to displace ions and therefore lose their
energy by producing free electrons and holes, excitons, and phonons.
It is evident that this will give rise to trapped electrons as well as trapped
holes. Hence; color centers of both the F type and V type are formed. In
contrast with the additively colored crystals, the color centers in X-ray
irradiated crystals are not permanent. They can be bleached by irradiation
with light or by heating, because,ultimately the excited electrons and holes
will recombine. Without going into details, it will be evident that studies
of the coloration, photoconductivity,
'
and bleaching at various temperatures
K
~j' nJ '
F
provide information about the proHI ~f;. ' ~ :~'~
perties of the color centers and their
interaction. As an example, we give
in Fig. 15-25 the absorption spectrum
induced in KCl by irradiation with
X-rays at 20C. 39
It is of interest to note that measurements of the change in density of
5
4
2 ev
3
6
alkali halides during X-ray irradiation
show that the lattice starts to expand Fig. 15-25. The F and V bands proas soon as the irradiation begins. It duced in KCI by irradiation with X-rays
thus seems that during the irradiation at room temperature. [After Dorendorf
and Pick, ref. 39J
vacancies are formed. 40 Eventually
the expansion saturates, a typical
value being 5 X 10-5 cm for a crystal with dimensions of I cm. At present
tbe interpretation of the production of vacancies upon X-ray irradiation is
still rather speculative, although it seems very likely that dislocations play
an important role in the process. 41
,I
REFERENCES
J. H. de Boer, Electron Emission and Adsorption Phenomena, Cambridge,
London, 1935.
I
N. F. Mott and R. W. Gurney, Electronic Processes in Ionic Crystals,
Oxford, New York, 1940.
R. W. Pohl, Proc. Phys. Soc. (London), 49 (extra part), 3 (1937); Physik. Z.,
39, 36 (l938).
F. Seitz, Revs. Mod. Phys., 18, 384 (1946); 26, 7 (1954). (These papers
constitute the most extensive review of the properties of alkali halides.)
H, Dorendorf and H. Pick, Z. Physik, 128,106 (1950).
K. Sakaguchi and T. Suita, Technol. Repts. Osaka Ulliv., 2, 177 (1952).
41 See J. J. Markham, Phys. Rev., 88, 500 (1952); also, F. Seitz, Revs. Mod. Phys.,
26, 7 (I 954).
39
40
396
[Chap. 15
PROBLEMS
15-1. Assume that the first characteristic absorption peak for KCl
(observed at 7.6 ev) is due to the transfer of an electron from a Cl- ion to
a neighboring K + ion. Calculate the energy required for this process on
the assumption that the positions of the nuclei remain fixed and that there
is no polarization. Compare the answer with the observed value, and from
this calculate the polarization energy.
15-2. Assume that the second characteristic absorption peak for KCI
(observed at 9.4 ev) is due to the transfer of an electron from a Cl- ion to
a distant K+ ion. Assuming that the nuclei remain fixed and that there is
no polarization, calculate the energy required for the transfer. Compare
the result with the observed value and calculate the polarization energy.
15-3. (a) Show that for a monatomic gas at a temperature T the free
energy is given by
... F = - NvkT[Jog (27TmkT/h2)3/2
+I
- log nv]
i1
where nv is the number of atoms per unit volume and Nv is the total
number of atoms in the system; m is the mass per atom.
(b) What is the configurational entropy of an alkali }lalide crystal
containing N ion pairs and nF F ~enters, relative to the crystal without
F centers?
(c) Consider an alkali halide crystal containing N ion pairs in equilibrium with the vapor of the alkali metal at a temperature T. The vapor
contains nv atoms per cm 3 ; the crystal contains nF F centers per cm 3 . Set
up an expression for the change in the free energy !1F of the system crystal
plus vapor if one atom is transferred from the vapor to the crystal;
neglect thermal entropy changes. Show that because !1F = 0 in equilibrium,
27TmkT) -3/2
nF'""
Nn v ( - e-~/kT
h2 -
for
n F'~
~ N
where cP is the energy required to take an atom from the metal vapor into
the crystal, thereby forming an F center. Compare the result with equation
(15-7).
15-4. Calculate the paramagnetic susceptibility of KCl at room temperature containing 5 X 1017 F centers per cm3 and compare this with the
diamagnetic susceptibility. Do the same for liquid air and for liquid
helium temperatures.
15-5. From the data given in Fig. 15-20, estimate the cross section for
capture of an electron by an F center and compare the result with that
stated at the end of Sec. 15-8.
Chap'. 15]
397
Chapter 16
LUMINESCENCE
16-1. General remarks
When a substance absorbs energy in some form or other, a fraction
of the absorbed energy may be re-emitted in the form of electromagnetic
radiation in the visible or near-visible region of the spectrum. This
phenomenon is called luminescence, with the understanding that this
term does not include the emission of black-body radiation, which obeys
the laws of Kirchhoff and Wien. Luminescent solids are usually referred
to as phosphors.
Luminescence is a process which involves at least two steps: the
excitation of the electronic system of the solid and the subsequent emission
of photons. These steps mayor may not be separated by intermediate
processes. This will be further discussed in the next sections. Excitation
may be achieved by bombardment with photons (photoluminescence),
with electrons (cathodoluminescence), or with other particles. Luminescence can also be induced as the result of a chemical reaction (chemiluminescence) or by the application of an electric field (electroluminescence).
When one speaks of fluorescence, one usually has in mind the emission
of light during excitation; the emission of light after the excitation has
ceased is then referred to as phosphorescence or afterglow. These definitions are not very exact since strictly speaking there is always a time lag
between a particular excitation and the corresponding emission of a
photon, even in a free atom. In fact, the lifetime of an atom in an excited
state for which the return to the ground state is accompanied by dipole
radiation is ,._, 1O-I! second. For forbidden transitions, involving quadrupole or higher-order radiation, the lifetimes may be 10-4 secolld or longer.
One frequently takes the decay time of ,._,10-8 second as the demarcation
line between fluorescence and phosphorescence. 1 Some authors define
fluorescence as the emission of light for which the decay time is temperatureindependent, and phosphorescence as the temperature-dependent part. 2
In many cases the latter definition is equivalent to the former, but there
are exceptions.
1 See, for example, G. F. J. Garlick, Luminescent Materials, Oxford, New York,
1949, p. I.
2 See, for example, F. A. Kroger, Some Aspects of the Lumillescence of Solids,
Elsevier, New York, 1948, p. 36.
398
LUMINESCENCE
Sec. 16-1]
399
400
LUMINESCENCE
[Chap. 16
the moment let us assume that the luminescence is associated with the
presence of activator atoms. The incorporation of an activator atom in a
crystalline solid will in general give rise to localized energy levels in the
normally forbidden energy gaps. These localized levels may be classified
into two categories: (i) levels which belong to the activator atoms themselves and (ii) levels belonging to host atoms which are under the perturbing
influence of the activators. The levels of group (ii) may be associated
with host atoms in the immediate vicinity of the impurity atoms, but they
may also be associated with lattice
defects (e.g., vacancies) whose existence is tied up with the incorporation
,r - + - "
of
the activator. For example, if
A-'Mn4+- ions were incorporated on sites
normally occupied by Zn2+ in a ZnS
G-+lattice, there may be localized levels
associated with the Mn4+- ion, levels
associated with the S2- and Zn 2+
(a)
(b)
ions in the vicinity of the MnH
Fig. 16-1. The ground state G and an
ion, and levels associated with ions
excited state A of a luminescence in the vicinity of a positive ion
center. In (a) excitation takes place by
vacancy (produced as a result of
direct absorption of a photon hi'".
the
presence of the Mn4+- ion to
In (b) excitation is achieved by capture
compensate for the excess positive
of a hole at G and of an electron at A.
charge).
I n terms of the energy band picture of Fig. 16-1 let G and A be two
levels corresponding to one of the categories (i) and (ii) mentioned above.
In the ground state, level G is occupied by an electron and A is empty;
in the excited state the reverse is true. The excitation from G to A may
be accomplished in at least three ways:
(a) It is possible that an incident photon of the proper frequency is
absorbed directly by the electron in level G, whereupon it arrives in A
(sel: Fig. 16-1 a). As a result of lattice vibrations the absorption will
correspond to a band centering about a certain frequency VII"
(b) The excitation process may also involve the diffusion of an exciton
(see Sec. 15-3). Suppose, for example, that in some part of the crystal an
exciton is produced; since the exciton may diffuse about in the crystal, it
may reach a center such as AG, whereupon it may give off its energy to
the center, resulting in excitation of the electron. This consideration is of
importance, since it provides a mechanism whereby energy can be transferred from the exciting source to the impurities via the host crystal. In
other words, the exciton mechanism makes it possible for the activators
to receive more energy than they ought to on the basis of their relative
concentration in the lattice.
LUMINESCENCE
Sec. 16-2]
401
(c) The excitation process may also involve the motion of free electrons
and holes. For example, let electron-hole p-airs be created somewhere in
the crystal, as for example, by bombardment with photons or electrons.
If the center'l4G is in its ground state, the level G may capture a hole
from the valence band and A may trap an electron from the conduction
band. In this way, excitation of the center has been achieved, as indicated
in (Fig. 16-lb). Evidently this type of excitation process should be
associated with conductivity, ~n contrast with processes (a) and (b).
E
.F:
'---'.
~--~...--'.
G
-q
(a)
402
LUMINESCENCE
[Chap. 16
403
LUMINESCENCE
Sec. 16-31
intensity of luminescence after excitation has ceased can readily be set up.
let the instant at which the exciting source is removed be denoted by
t = O. Suppose at any instant t the number of electrons in excited states
such as A in "Fig. 16-1 is given by n(t). Let us assume that the probability
for an electron in A to return to the ground state G \s 1IT per second, and
that such a transition is associated with the emission of a photon. If the
center is well screened from its environment, the average lifetime T of the
excited state is independent of temperature and of the number of other
--:,' _,'
excited centers. Hence the intensity of luminescence l(t), i.e., the number of photons
emitted per unit time, is given by
E
I(t)
= -(dnldt) = nIT
(16-2)
J. T. Randall and M, H. F. Wilkins, Proc. Roy. Soc. (Londoll), AI84, 379 (1945).
404
LUMlNESCENCE
[Chap. 16
"
net)
no/(noa.t
+ I)
and
let)
rwM(noa.t
+ 1)2
( 16-7)
const./t(f3kT+ 1)
(16-9)
LUMINESCENCE
Sec. 16-3]
405
<
1= -(dn/dt)
(n/To)e- E /H '
(n) =
log no
-I
(16- 10)
OTO' To
Since the temperature To has been chosen such that kTo <{ E, the lower
limit of integration may be replaced by zero. One thus obtains
1= (n o/To)e- E/k1' exp [-(I/OTo) f.T e- E/k1' dT]
,0
(16-11)
..
"
406
LUMINESCENCE
[Chap. 16
When "pure" alkali halides are irradiated with X-rays one observes in
the dark a faint luminescence. The decay of the phosphorescence may be
followed with sensitive equipment for several hours and has been studied
by a number of investigators. 8 The interpretation of these experiments is
somewhat difficult in view of the possible role played by small amounts
of unknown impurities. On the other hand, alkali halides activated with
thallium freqm:ntly exhibit high efficiencies for luminescence; studies of
these materials have provided a fairly detailed uQ.derstanding of their
luminescent properties. Although it
is impossible to enter into a detailed
5
discussion of this subject here, some
's 4
of the most important features may
E3
be mentioned, in particular those of
::.:
KCI :TI (we shall adopt this notation
2
to indicate the host crystal and the
activator).
1800
2000
2200
2400
2800
LUMINESCENCE
Sec. 16-4]
407
same in all alkali halide host crystals, the maximum shift being I ev.l0
For KCI :TI the bands measured at r<fom temperature occur at the
following energies and wavelengths.
....
(ev) ........ .
A(A) ........ .
Transition .. .
(-.
4.9
2470
150
5.9
2060
'c
6.3
1960
IS" - .. IP 1
->- 3Pl
Emission spectra. The two principal emission bands of KCI :TI center
around 3050 A and 4750 A. The former has been identified with the
transition 3P 1 --+ ISO; the 4750 emission is due to IP 1 --+ ISO' An example
of a glow curve obtained by Johnson and Williams is given in Fig. 16-5. 14
For a summary of data, see, for example, G. F. J. Gallick, op. cit., p. 50.
F. Seitz, J. Chern. Phys., 6, 150 (1938).
.
12 For a brief review and references to this work see F. E. Williams, "Solid State
Luminescence," Advances in Electronics, 5,137 (1953).
13 F. E. Williams, J. Chern. Phys., 19,457 (1951):
If P. D. Johnson and F. E. Williams, J. Chern. Phys., 21, 125 (1953). 'J, . ,
10
11
408
LUMINESCENCE
[Chap. 16
From the two peaks they conclude the existence bf two metastable levels
with activation energies of 0.35 ev and 0.72 ev; these must probably be
ascribed to the 3Po and 3 P 2 states. From these and other detailed studies
Johnson and Williams suggest the energy diagram as function of the radial
'1,'
-.6
-.4
-.2
_q
409
LUMINESCENCE
Sec. 16-4]
~ejf~e
+ {J(I
- e)J = ejfe
+ ({3j~)(l -
e)J
where {Jj~ is the ratio of the capture cross section of a photon of given
wavelength by a lattice atom and by a TI+ ion. Evidently, the ratio {3j~
will be a function of the wavelength of the exciting radiation; it also
depends on temperature. Thus,
1]
= e
+ ({3j~)(l
- c)
(16-12)
.,'.,
410
LUMINESCENCE
[Chap. 16
Sec. 16-5]
LUMINESCENCE
411
a reasonable rate of recrystallization, one requires under normal circumstances temperatures of the order of 1200(;. However, if one adds a flux.
temperatures of, say, 800C may be sufficient, and for practical reasons
this is usualI)' done. For example, for activation of ZnS with monovalent
metals such as Cu, Ag, Au, salts like NaCI or CaCI 2 are found to be
suitable fluxes. It seems that the important role played by the flux material
in the preparation of sulfide phosphors was never fully realized until some
years ago. In fact, Kroger and collaborators have shown that the
incorporation of activators such as Ag, Cu, Au, Li, Na in zinc sulfide is
governed by the so-called principle of charge compensation, which will
now be explained. 20
First consider what may occur when zinc sulfide is fired with another
divalent sulfide, say MnS. Since the valences of the metal atoms are the
same and since their radii do not differ too much (Mn2+ is approximately
10 per cent larger than Zn2+), one may expect a substitutional mixed
crystal to be formed in which Mn2+ ions have replaced Zn2+ ions. Consider
now, however, a solid solution of ZnS and Ag 2 S in which Ag+ ions occupy
positions normally occupied by Zn2+ ions. Tn order to conserve charge,
the crystal must contain one sulfur vacancy for each two silver ions
incorporated in the lattice. Since the creation of a vacancy requires a good
deal of energy, the amount of silver incorporated in the lattice will be
strongly limited. Similarly, a mixed crystal of ZnS and ZnCI 2 , in which
Cl- ions occupy lattice sites normally occupied by S2- ions, must contain
positive ion vacancies. The formation of vacancies may be avoided,
however, ifin the case of ZnS :Ag a monovalent negative ion is incorporated
for each Ag+ ion. For example,
412
LUMJNESCENCE \
[Chap. 16
!
Sec. 16-5]
LUMINESCENCE
413
[Chap. 16
LUMINESCENCE
414
action of the field taking the place of the action of thermal vibrations.
For further details of the present situation we refer the reader to the
literature. 24
The Destriau effect. The emission of light by a phosphor resulting
solely from the action of an electric field applied to a suspension of luminescent particles in an insulator was first discovered by Destriau. 25 In this
case one may speak of intrinsic electro luminescence, since the effect does
not involve previous photo-excitation, nor the injection of charge carriers
from an external source. An electroluminescent cell is usually made in
the form of a parallel-plate capacitor of which at least one of the conducting plates is transparent. In order to transfer power to the dielectric
consisting of the luminescent powder embedded in an insulator, alternating voltages or pulses must be used. For sinusoidal voltages the average
brightness B increases rapidly with increasing amplitude. Several empirical
formulas have been introduced to describe the observed brightness versus
voltage curves, for example,
B
aV n exp (-h/V)
(I 6-13)
where a, h, and n are constants. The curve shown in Fig. 16-9 has been
obtained by Roberts 26 for a copper-activated zinc sulfoselenide phosphor
embedded as a powder in a variety of dielectric materials. If one assumes
that the luminescent particles are spheres, one can show that the local
field 2 in the phosphor is given by the expression
E2 =
3lE
-----=----21
2 -
11(2 -
(16-14)
1)
LUMINESCENCE
Sec. 16-6]
415
10
o
-+ E2
(volts/micron)
REFERENCES
G. F. 1. Garlick, Luminescent Materials, Oxford, New York, 1949.
F. A. Kroger, Some Aspects of the Luminescence of Solids, Elsevier,
New York, 1948.
H. W. Leverenz, Luminescence of Solids, Wiley, New York, 1950.
P. Pringsheim and M. Vogel, Luminescence of LiqUids and Solids, Interscience, New York, 1943.
F. E. Williams, "Solid State Luminescence," Advances in Electronics.
Academic Press, New York, 1953, Vol. 5, p. 137.
27
'7
416
LUMINESCENCE
[Chap. 16
PROBLEMS
16-1. Suppose an X-ray tube is operated at 60 kv and 10 mao Assume
that energywise 2 per cent of the electric energy is transformed into X-rays.
The luminescence efficiency of a good phosphor under X-ray excitation is
,_,10 per cent. Estimate the number of photons emitted by the phosphor
for an excitation energy ,....,5 ev; take into account a reasonable geometry
for the coupling between the X-ray source and the phosphor.
16-2. With the most intense light sources one can obtain a beam of
photons with energies ~ 3 ev corresponding to ,..._.1019 per cm2 per second
incident on a phosphor. Suppose the phosphor contains lOIS luminescence
centers per cm3 . Assuming that the host lattice does not absorb the incident photons, estimate the penetration depth of the photons and from
it the average number of primary photons available per activator per
second. Does this explain why saturation effects have not been observed
for photoluminescence? Assume a lifetime of an excited center of 100s
second.
16-3. Suppose a cathode-ray tube operating at 25 kv delivers 2.l()4
watts per cm 2 to a phosphor screen: Calculate the penetration depth of
the electrons from the simplified Bethe formula [Ann. Physik, 5, 325
(1930)], x = 2j47TNZe4, where E is the energy of the incid~nt electrons,
N is the number of atoms per unit volume, and Z is the average number
of electrons per atom. Assume further that the primary electrons expend
approximately 30 ev per excitation of a luminescence center (this includes
losses of several kinds). Estimate the number of excitations available per
center per second if the density of the latter is 1018 per cm3 Show that this
may lead to saturation effects, in contrast with the photoluminescence
in the previous problem.
]6-4. It was noted in this chapter that the concentration-dependence
of the luminescence of KCl :Tl can be described by formula (16-12) with
.Z ':::' 70. What does this imply for the lower limit of the distance between
two TI+ ions required to prevent quenching? .
16-5. Suppose that the decay of a luminescent material may be
described by a bimolecular mechanism of the type dnjdt = -rxn 2 If
at t = 0 the exciting source is switched off and the luminescent intensity
is then 10 , show that the time required for the intensity to reach half its
initial intensity is given by (Vl - I)j(/orx)1I2. (Note thatt l / 2 depends on 10)'
For the same phosphor discuss the build-up of the intensity ofluminescence
.~'!
.~ H
under constant illumination.
Chap. 16]
417
LUMINESCENCE
16-6. Give a proof of expression (16-14) for the local field in spherical
particles of a luminescent material embedded in a homogeneous dielectric.
16-7. FQr spherical particles of dielectric constant
El
and resistivity
p embedded'in an insulator, show that the field in the particles leads the
field in the insulator by an angle rp such that tan rp = 2/EIPV, where v is
the frequency of the applied field.
r ,
Chapter 17
;:(1
.. ", h
418
419
Sec. 17-IJ
~hree
categories of electrons
This may be illustrated by considering as an example the energy distribution of the electrons emitted by silver upon bombardment with primary
electrons of 155 ev, as shown in Fig. 17-1 according to Rudberg. 3 (Such
I
o
50
100
150
-E(ev)
Fig. 17-1.
E. Rudberg. Proc. Roy. Soc. (London). A127, III (1930): Phys. Rev., 4, 764 (1934).
420
[Chap. 17
_-
Sec. 17-2]
421
'"
8
7
,-
6
~
4
3
2
1
.........
~
"-
/
I
'\.
"'-"-
"
~ ~o'c
~ i'--- r--
---r-r--
II
1.0
3.0
2.0
4.0
5.0
-+ Epo (kev)
Fig. 17-3. Secondary yield '0 versus primary energy (kev) for a'
single crystal of MgO. The upper curve refers to room temperature,
the lower one to 740 C (see Sec. 17-9). [After Johnson and McKay.
ref. 5]
1.2
1.0
h
1
t
.8
.6
r; ~
V
II
"~
~ ~20'C
~t-I
.4
.2
1.0
2.0
'c
3.0
4.0
5.0
-Epo (kev)
Fig. 17-4. Secondary yield t5 versus primary energy in kev for a
germanium single crystal. The upper curve refers to room temperature, the lower one to 525C. [After Johnson and McKay, ref. 6]
"
",
422
[Chap. 17
to the surface; the same is true for the examples given in Figs. 17-3,
J 7-4, J 7-5. We note that the maximum yield of metals is of the order of
unity, the largest value having been observed for platinum (l.8).7 The
intrinsic semiconductors Ge and Si also have a maximum yield of about
unity; according to Johnson and McKay, the yield of Ge is independent
2.0
1.8
1.6
~
1.4
1.2
t-- t--
r-- t--
1/
1.0
.8
.6
400
800
-Epo
1200
1600
2000
(ev)
Substances
Ag
AI
Cu
Fe
Pt
Ge
<1m
ED'" (ev)
1.5
800
1.0
300
1.3
1.3
Si
NaCI
600
350
800
400
250
600
MgO
21,
1100
1.8
1.1
1.1
Sec. 17-3]
423
(a) The primaries, as they penetrate the sDlid, mDve alDng straight
lines alDng the directiDn Df incidence; this assumptiDn thus neglects
.<l !
elastically and inelastically reflected primaries.
(b) The primaries are incident perpendicular to. the surface.
(c) The energy IDSS Df the primaries per unit path length is given by
WhiddingtDn's law 9
dEp(x)
A
(17-2)
---where A is a CDnstant characteristic of the material.
(d) The number ofsecDndaries produced in layer dx by a single primary
is proportional to dEp/dx, i.e.,
"",'
1 dEl)
(17-3)
n(x) = - - . Ee
dx
where Ee represents the average excitation energy required to produce a
secDndary.
(e) The probability fDr a secondary prDduced at a depth x to escape
from the surface is determined by an exponential absorption law,
(17-4)
424
[Chap. 17
;0 -
E;(x) =
5
...~
I
'it 3
I
Jt
./
,2
.4
.6
.8
l--""V
2Ax
1.0
. n(x)
= ( 2"
. fEe(x" _
X)1/2 (17-6)
t5 = 1(0) f n(x) dx
= 1(0) ~
(17-7)
fEr
Thus, for low primary energies the secondary yield should rise proportionally to E pO' The other extreme case to be considered corresponds to
a primary range very large compared with the range of the secondaries.
Under these circumstances one is essentially interested in the function
n(x) for very small values of x/x", because the function I(x) in (17-1)
decreases strongly for values of x> x" = 1/a.. From Fig. 17-6 it is
evident that the production of secondaries as a function of depth may then
be considered as approximately constant over the range of x-values of
425
Sec. 17-3]
interest. J n fact one may then employ in 3ccordance with (17-5) and
(17-6),
\
n(x)
n(O)
= _._-
,E)o
(J
(J
for
proportional to E"
proportional to E)-I
x"<{x,,
for
x,,}>x s
I"
A)l/:!. I
(
="2
y2 = lJ(x" - x)
expression (I 7-9) may be rewritten in the form,
(J =
10
2A)I/21
v=.
- j(O)-=,' (, ell" dl'
( -.IJ(
,
0
{I 7-9)
426
[Chap. 17
Writing ctX p = E!oct/2A = Z2, the last expression gives for the yield as
function of the primary energy.
/) =
(17-10)
.0
Evm
2A)1/2
0.92 ( --;
(17-11)
tJ
F(z)
om = F(O.92)
[J 2A
1.85F Evo
IX ]
1.85F
[O.92E po ]
Evm
( 17-12)
11
Sec. 17-4]
427
Both'theoretical curves show considerable deviation from the experimental one for primary energies> Epm. In this connection it is interesting
to note that for magnesium oxide, which is an insulator, the yield decreases
as E;;oI, in agreement with the theoretically predicted behavior (Fig.
17_3).12 Germanium (Fig. 17-4), on the other hand, shows deviations
similar to those for metals. It has been suggested that the deviations
from the Epr.,t law for high primary energies result from the presence of
inelastically reflected primaries. The influence of such primaries on the
yield is twofold:
(i) They increase the yield simply because they are collected as
emitted electrons.
(ii) They may increase the yield because they may produce more
secondaries over a depth equal to the range of the secondaries"
than does a primary that penetrates deeply into the solid. The
reason is that their path may pass twice or more through the
same region of the solid close to the surfa,ce.
I' A. J. Dekker, Phys. Rev., 94, 1179 (1954).
428
[Chap. 17
13
Sec. 17-5)
429
per primary electron inside the material. Then, if x", is the mean depth
of origin and exp (-~xm) is the probability for escape, the yield for
perpendicular incidence (B = 0) is
given by'",
708
1.25
.75
In (rJ o/t5 o)
~xm = I _ cos B
[/
~V
1.00
Hence
(17-13)
.50
~V
\ro8
-- ,
40 8 .........
...........
I
I lIP'b
60 .Ii
o R
constant
where A and ~ are the material constants introduced in Sec. 17-3: This
relationship is in very good agreement with his measurements on nickel,
nickel carbide, and lithium. Similarly, his theory permits establishment
of a relation between the maximum yield 15 m ~nd the angle of incidence,
leading to
rJ m (cos B)ItZ = constant
1. This value was found for nickel by A. Becker, Ann. Physik, 2, 249 (1929).
J. L. H. Jonker, Philips Research Repts., 6, 372 (1951).
,,'
16
430
[Chap. 17
This relation is in reasonable agreement with his results for the materials
mentioned above. Finally, it is interesting that, according to his calculations, a universal yield curve is obtained by plotting b versus const.
1)0 cos (), whereby the constant is chosen in such a manner that the
maxima of"the curves for different values of () coincide. His experiments
bear out the fact that such a universal curve for different ()'s indeed exists.
However, there is the same discrepancy between the experimental and
the theoretical curves as discussed in the preceding section. On the other
hand, it seems that Jonker has established an important experimental fact.
17-6. Baroody's theory of secondary emission for metals
Although the elementary theories provide a certain amount of insight
into the phenomenon of secondary emission, it does not allow one to
discuss many details. For example, one simply speaks about numbers of
secondaries without paying attention to their energies. Also, the exponential absorption law for secondaries actually hides a very complicated
mechanism by which the secondaries lose energy on their way to the surface.
,Baroody in employing a Fermi model for the conduction electrons
in metals, improved the situation considerably for this group ofmaterialsP
His theory shows, among other things, that the secondary yield for
metals with high work function is larger than for those of low work
function. This had been found experimentally but could not be explained
by the elementary theory. In fact, one would expect metals of high work
function to have a low yield because it is difficult for secondaries to
escape in that case. It may be noted that Kadyschewitsch used a similar
model as Baroody, but his calculations are complicated. The essential
points of Baroody's theory will now be discussed.
It is assumed that the metal is at absolute zero so that in the momentum
space all electrons lie within a sphere of radius PF about the origin;
all higher states are empty. This assumption does not restrict the application of the results, because no temperature effect of the yield has
been detected for metals. The velocity of the primary electrons is assumed
to be very high relative to that of the conduction electrons. Consider
then, as represented in Fig. 17-9, the collision between a primary and a
conduction electron. The instant at which the distance between the two
electrons is b will be denoted by t = o. Assuming the conduction electron
to be at rest and the primary to move along a straight line, the component
of the Coulomb force between the two particles perpendicular to the
primary path is at any instant t given by
(17-16)
17
431
Sec. 17-6]
V.__!----!t=~--~~3~
~. = J+oo
p
00
Fdt
2e
bu
(17-17)
"
432
[Chap. 17
produced per unit primary path length with a momentum> PPF varies
inversely as the primary energy. The assumption of a Whiddington law
for the primaries, as used in the elementary theory discussed previously,
is in agreement with this result.
Equation (17-18) gives the production of secondaries for a particular
value Ep of the primary. Now Ep is a function of the path length covered
by the primary inside the solid. Denoting the depth below the surface by
x. Baroody employs the same relationship as that used in the older theories,
E;(x) = E;o - ax
(17-19)
The constant a here is equal to Aj2 used in the preceding sections. From
(17-19) and by differentiating (17-18) one thus obtains for the number of
secondaries produced per primary in a slab dx and with a momentum
between f-lPF and {f-l + df-l)PF,
2BE,Y/f-l df-l dx
(17-20)
,---------
EF
+ cp
or
ft~
1 + CPjEF
(17-22)
Sec. 17-6]
433
1.6
"1'
1.4
Au
~
_-
'lou
T~O
,:Mo- Pd
Co
~- Cd,Zr
im 1.2
1.0
.8
Pt
Ag
-~
Jb
!i}.;
:AI
Ba
Ca,K
.6
oLi
3
"
I-'PF has on the average a probability {fJ - 1-'0)/1-' to escape from the
surface (see Problem 17-7). One thus obtains finally for the secondary yield,
d=
rx"
-x/Ld
e
x
roo fJ - fJo d
F Jo (E;o _ ax)I/2 Jl'o (fJ2 _ 1)2 fJ
2BE1I2
aL
(17-23)
,
434
[Chap. 17
several metals. 20 The solid line is the one drawn by McKay, the dashed one
represents (0.354112 and is matched at thorium. We emphasize that the
dependence on 4> does not enter through the escape mechanism but
rather through the expression for the production of secondaries (17-20),
which contains the factor E}J2. We must also remark that if for a given
metal the work function is lowered, for example, by a monolayer on the
surface, the yield, of course, increases.
Space does not permit discussion of this model any further, but
it may be noted that from (17-23) the energy distribution of the secondaries may also be obtained; the agreement with experiment is good.
17-7. Wave-mechanical theory of the production of secondaries
e2
I
V(R,r) = -,
-,.R - r E
(17-24)
21
Sec. 17-7]
435
V(R,r) =
IR e rl exp [ - IR A rl]
'tw
~'C'~~2;)
with
According to the usual procedure of time dependent perturbation theory,
the coefficients are given by
I
a,
Il K
,(t)
f
iii JoJr
= -
ft
,}Ih
e-i(K"R)w
Til
Hl n; f!il Ii
,(r)
E
IR _ rl
"
ei(K'R)w
(r) e-i(E-E')I'h dr dR dt
Til
(I 7-28)
where the notations dr and dR refer to integrations over the volume of,
the crystal. The integration over R becomes
.
436
[Chap. 17
Ia". .(1)1
2 4
167T e
2q4
(E' _ E)2
( 17-29)
""-',
07-30)
y.
(17-31)
ei(h)u,,(r) .
with b representing 2
1T
1= Lcb(k)
u,,(r) =
.!
blb.b.
cb(k)eib' r
Sexp [i(q + k -
k'
+ b). rJ dr
(17-32)
+k -
k'
+b =
0 or
+k +b =
K'
+ k'
(17-33)
Sec. 17-7]
437
-?
4m 2e4 K'
K;,4(q2
+ .1.2)2 1'1
dO'
(17-34)
----e;'
where Ep is the primary energy and ). = ;,2A2j2m c:::' 40 ev for A c:::' 108
em-I. Note that the production increases as Ep decreases, in accordance
with the ideas employed in the elementary theory. We also observe that
(17-35) remains finite even for E' equal to the Fermi energy E F; this is a
consequence of the screened potential. Another quantity of interest
is the energy loss of the primaries per unit path length. It turns out that
in first approximation this quantity is given by a Bethe-type law,
p 7TM.e
dE
(Ep)
--c:::'--log
4
dx
Ep
eE).
(17-36)
438
[Chap. 17
,Aj
23
See, for example, A. 1. Dekker and A. van del' Ziel, Phys. ReI'., 86. 755 (1952).
Sec. 17-8]
439
440
[Chap. 17
Eo - Noc(T)
(17-38)
Also, the mean free path for collisions with lattice vibrations for electrons
of several ev of energy is proportional to the energy times a function of
temperature. We may therefore write for the mean free path,
A(E,T) = AoEf(T)
(17-39)
25
Sec. 17-9]
441
(17-40)
moti~
.. ,,-, - 07-41)
where the averages must be taken over the N", collisions. Now, according
to (17-38) we may write
(E2)av = Eg
+ a.2(N2)av -
2a.Eo(N)av
<x2 )av =
const. N m [f(T)]2
[f(T)J2
= const. - ( ocT)
(17-42)
-,...., - - - -
(17-43)
For ionic crystals, the mean free path for lattice scattering is given bY~,5
I
or
/<T) =
(2n"
I)
(17-44)
where n" = [exp (hv/kT) - 1]-1 and l' is the frequency of the optical
442
[Chap. 17
hv \
-dE A = :-----:-
2n,. + I
dx
(17-45)
Thus for ionic crystals, expression (17-43) may be written in the form
61
T ~
{}2
[2n"2
2nvl
+ 1] 1/2
+1
for
xp:;>-
Xs
( 17-46)
This result has been applied to explain the variation in yield as function
of temperature for magnesium oxide single crystals. 2 ,i For Tl = 1013K
and T2 = 298K, Johnson and McKay observed an average ratio ()1/()2
= 0.78 for primary energies above 2 kev. 5 Now, from optical absorption
measurements it follows that for MgO, hv = 1300 k. 25 Employing this
value, one obtains from (17-46) for the same ratio, 01 /6 2 = 0.76, in
good agreement with the experimental value. It thus seems that the simple
model used above gives a satisfactory explanation of the temperature
effect in MgO. For nonpolar crystals a similar calculation may be carried
out, starting from (17-43). However, no experimental data are available
to check the theory further. It may be noted that a more general theory 27
based on the Boltzmann transport equation leads to the same result as
obtained here.
It must be emphasized that as the temperature is raised the average
energy loss suffered by the secondary per collision decreases. For MgO,
for example,
~
(298K) = 0.108 ev
IX
(I0l3K)
0.063 ev
The decrease of the yield with increasing temperature is thus a consequence of the reduction in the mean free path and of the fact that the
path of the secondaries is curled up. In fact, if the secondaries moved
in straight lines toward the surface, there would be no temperature effect,
because dEldx is temperature-independent for electrons of a few ev
energy. (For thermal electrons this is not true).
For low primary energies, corresponding to the rising part of the
yield curve, the influence of temperature on the yield is very slight,
because most secondaries are then produced close to the surface and the
energy losses resulting from scattering become less important.
17-10. The possible influence of donor levels on the secondary yield of
insulators
Sec. 17-10]
443
REFERENCES
H. Bruining, Physics and Applications of Secondary Electron Emission.
McGraw-Hili, New York, 1954.
28
444
[Chap. 17
17-1. At first sight it may seem strange that the "range" of a primary
electron may be smaller than that of a secondary, because the energy of
the former is always larger than that of the latter. From the definitions of
x and Xs used in the theory of secondary emission, explain that this
difficulty actually does not exist.
J)
Chap. 17]
445
Chapter 18
(ii) paramagnetism
I..
(18-1 )
Sec. 18-1]
447
B = H
+ 47T M
(18-2)
= ftH
+ 47TX
Jlf:;mom !}l()Qt
(18-3)
This relation is the analogue of the expression for the dielectric constant
when X represents the ratio of the electric moment per unit volume and
the applied electric field.
It is convenient to normalize the potential energy of a dipole fJ. in a
magnetic field H in such a way that
(18-4)
by
the volume
by
fdE = -- f: M dH = -
f~l XH d~ =
-hH2
(18-5)
,
448
(Chap. 18
(18-6)'
= IS/e
where I is the current and S is the area of the loop. The dipole direction
is perpendicular to the plane of the loop. Employing this relation, let us
consider the magnetic dipole moment associated with an electron describing
a circular orbit of radius r, the angular velocity of the electron being Woo
The loop current in this case is4 -ew o/27T so that, according to (18-6),
the magnetic dipole moment associated with the electron orbit is
(18-7)
It is of interest to relate the magnetic dipole moment to the angular
momentum of the electron, which in this case is m(l)or2 According to
(18-7) we have
ft = -(e/2mc) X angular momentum
"I
(18-8)
The minus sign indicates that the dipole moment points in a direction
opposite to the vector representing the angular momentum. Relation
(18-8) is valid for any electron orbit, as will be shown in Sec. 18-7; it is
not valid, however, for the spin of an electron or nucleus, as we shall
see below.
The use of quantum numbers.5 A few remarks may be made here to
refresh the reader's memory on the use of quantum numbers in the theory
of atoms.
r,"
I
(a) The principal quantum number n determines the energy of the
orbit; it can accept only the integer values n = 1, 2, 3, .... The corresponding electronic shells are called the K, L, M, N, .,. shells.
0, 1, 2, '" , (n - 1)
(18-9)
3 For a general proof. see, for example, R. Becker, Theorie del' Elektrizitiit, Teubner,
Leipzig, 1933, Vol. 2, p. 96. See also Problem 18-1 for a particularly simple example.
Unless otherwise specified, the electronic charge will be represented by -e.
o A clear account may be found in G. Herzberg, Atomic Spectra and Atomic Structure,
Dover, New York, 1944.
449
Sec. 18-2]
1i[1(I + 1)]112
Electrons
as~ciated
(18- 10)
'//E
~----- hVI(I+ 1)
------~,.
s= -1/2
'
=+1/2
g~BH
,,
1-1
.~
(18-11 )
F or example, a p electron has the possible components of angular momentum along the direction of a magnetic field Ii, 0, -Ii. Consequently, the
possible magnetic moment components along the direction of an applied
magnetic field are (see Fig. 18-1)
-eli/2mc, 0,
+eli/2mc
20
.;'
450
[Chap. 18
On the basis of (18-8) one thus expects that the electron spin will give
rise to a component of half a Bohr magneton. It must be emphasized,
however, that for the spin, relation (18-8) is not valid. In fact the magnetic
moment component fl ... of the spin along an external field is given by
(18-12)
flu = g(e/2mc)(1i/2)
"
g= I
J(J
+ I) + S(S + 1) 2J(J
1)
L(L
+ I)
,
(18-14)
(i) The electron spins add to give the maximum possible S consistent
with the Pauli principle.
For a derivation see G. Herzberg, op. cit., p. 109.
Sec. 18-2)
451
(ii) The orbital momenta combine to give the maximum value for L
that is consistent with (i).
(iii) For an incompletely filled shell, we have
"J =
J= L
10-24 erg/oersted
(18-15)
'f
I, I
0), etc.
Iqq:
452
[Chap. I
Consider a loop current with its associated magnetic field. When one
attempts to change the magnetic flux enclosed by the loop by applying
an external field H, a current is induced in such a direction that the
magnetic field resulting from the induced current counteracts the field H.
Suppose now that the electrical resistance of the loop is zero; the induced
current will then persist as long as the external field is present. Such a
situation is realized in the loop current associHand WL
ated with the motion of an electron in an
atom. It is also approached in superconductors. Consequently, any atomic orbit
will produce a negative contribution to the
magnetic susceptibility.
The Larmor precession. Let us now consider the influence of a magnetic field on the
motion of an electron in an atom quantitatively. With reference to Fig. 18-3, we shall
assume an arbitrary direction for the angular
momentum vector Ma relative to the magnetic
field H.
The magnetic dipole moment is in accordance with (18-8)
(dJdt)M"
(18-16)
The magnetic field produces a torquefJ.X H
on the dipole, so that, according to Newtonian
mechanics, we may write
=
fJ. X H = -(eJ2mc)M" X H
(18-17)
Sec. 18-3]
453
From what has been said above, it follows that under the influence of
an external field, the plane of the orbit is nut stationary, but precesses
about H. As a result of the charge of the electron, the precession produces
an induced m&gnetic moment with a component opposite to that of H.
In fact, in accordance with (18-8), this component is equal to
(18-19)
where (p)2 is the mean square radius of the projection of the orbit on a
plane perpendicular to H. When this treatment is extended to a solid
containing N atoms per cm3 , each atom containing Z electrons, one
obtains for the diamagnetic susceptibility defined as the induced moment
per cm3 per gauss,
(18-20)
l
Here it has been assumed that the charge distribution of the atoms is
spherically symmetric, so that (r2) = i(pj2 represents the mean square
distance of the electrons from the nucleus. 8 The diamagnetic susceptibility
is thus determined essentially by the charge distribution in the atoms.
Note that X is negative. With (r2) ~ 10-16 cm2 and with N ~ 5.1022 cm-3 ,
one obtains X ~ 1O- 7Z ~ 10-6
Experimental values for the molar diamagnetic susceptibility of a
number of ions in solids are given in Table 18-1. 9 It should be emphasized
that the susceptibility of ions is determined to some extent by their
environment and the values are therefore approximate. Note the increase
in the absolute magnitude of X.u with the number of electrons per ion.
The reader may compare this table with that for the polarizabilities of
these ions (Table 6-1).
,(
Table 18-1. The Molar Diamagnetic Susceptibility x 10' for a
Number of Ions
Li+
Na+
K+
Rb+
Cs+
-0.7
-6.1
-14.6
-22.0
-35.0
Mg2+
Ca H
8rH
BaH
-4.3
-10.7
-18.0
-29.0
FClBr1-
-9.4
-24.2
-34.5
-50.6
(x 2) = (y") = (Z2);
furthermore
(r2) = (x')
= (x')
+ (yO).
G. W. Brindley and F. E. Hoare, Trans. Faraday Soc., 33, 268 (1937); Proc. Phys.
Soc. (Londoll), 49, 619 (1937).
454
[Chap. 18
.\
between the dipoles is weak, so that the field in which a given dipole finds
itself is equal to the applied field H. We shall assume in this section that
the magnetic field is constant or varies very slowly with time. In the
classical theory, the dipoles are assumed to be freely rotating. Hence the
resulting magnetic moment M per unit volume can be calculated in
exactly the same way as the polarization P for a dipolar gas. Thus,
according to the Langevin-Debye theory (see Sec. 6-3) we find
M
(18-21)
N p..L(pH/kT)
where
MJ
J, (J -
(18-23)
See, for example, F. Seitz, Modern Theory of Solids, McGraw-Hill, New York,
1940, p. 583.
11 L. Pa uling, J. Chem. Phys., 4, 673 (1936).
. , q; I'
Sec. 18-4]
455
'",
M = N
--...:.J-..,.+-:J;----------
(18-24)
2 exp (MJgPBH/kT)
-J
x=
M/H
Ng2J(J
+ 1)p~/3kT
{I 8-25)
This result is identical with the classical result (l8-22) because the total
magnetic moment PJ associated with J is given by
, pi
g2J(J + l)p~
(18-26)
= g[J(J +
(18-27)
1)11/2
2J + 1
[(2J + 1)x]
-----ucoth
2J
-
1
( x)
2J coth 2J
(18-29)
456
[Chap. 18
12
Dy
10
.8
Pefl'
6
Pr Nd
Cp
atom was represented by the sum of the applied field and the field due to
the polarization of the surroundings. On the other hand, in the derivation
of the magnetic susceptibility above, the field acting on a dipole in a
paramagnetic solid was assumed to be equal to the applied field H. The
justification for this is the following: the order of magnitude of the internal
field is given by H + y M = H(l + YX), where y ~ 4. Hence the fractional
error made in neglecting the internal field correction is of the order of X.
As we have seen above, this is small for paramagnetic materials. For the
electrical case, the susceptibility is PIE = ( - 1)/47T, and the internal
field cannot be neglected in solids or liquids, since ( - 1) is not small
compared with unity.
It should finally be mentioned that there exists also a temperatureindependent paramagnetic contribution to the susceptibility at low
temperatures. This is called van Vleck paramagnetism. 'For its theoretical
treatment we refer to van Vleck, Theory oj Electric and MagnetiC
Susceptibilities, Oxford, New York, 1932.
Sec. 18-5]
457
\
The rare earth ions. The theory
\
\
4
\
outlined in the preceding section
\
\
describes the behavior of most of the
\
\
rare earth salts quite well. This may 2
\
\
be seen from Fig. 18-4, where the full
\
curve represents the effective number
22
24
20
26
of Bohr magnetons calculated by van
-z
Vleck from expression (18-27); the
J values and g were obtained from Fig. 18-5. The effective moment in
Hund's rules and from Lande's Bohr magnetons for the iron group as
function of the number of electrons Z
formula, as outlined in Sec. 18-2. in the ions. The dashed curve repre. The vertical lines correspond to ob- sents the values calculated from (18-27);
served values of Perr, obtained from the full curve refers to the "spin-only"
measurements of the temperature formula (18-30). The vertical lines
dependence of X (see equation 18-25). represent the ranges of experimental
values. (After Bates, Modern Magnetism,
The ions Sm3+ and Eu3+ evidently
Cambridge, 1951, p. 152]
do not obey the simple theory.
However, it has been shown by van Vleck and Frank13 that these' discrepancies can be explained satisfactorily if one considers the special
situation with regard to the energy levels of these ions.
The iron group ions. If one calculates the effective number of Bohr
magnetons for the ions of the iron group from expression (18-27), the
results do not agree at all with the experimental values obtained from the
Curie law. This may be seen from Fig. 18-5 where the vertical lines
represent experimental values and the dashed curve represents (18-27).
However, if one assumes that only the electron spins contribute to the
magnetization, i.e., if one replaces (18-27) by
Peff = 2[S(S
+ 1W/
. (18-30)
one obtains quite good agreement with experiment (full curve in Fig.
18-5). Thus the iron group ions behave as if the orbital magnetic moment
,3 A. Frank, Phys. Rev., 39, 119 (1932).
458
[Chap. 18
does not contribute at all. One speaks in this case of quenching of the
orbital momentum. The quenching is not necessarily complete; it may be
partial. Stoner suggested the following explanation for the different
behavior of the rare earth and iron groups in this respect:14 In the solid
state, the paramagnetic ions find themselves in strong electric fields
produced by neighboring diamagnetic ions. In the iron group, the paramagnetic 3d electrons are the outermost electrons and these are therefore
fully exposed to the crystalline field. Consequently, the orbital motion is
locked into the field of the neighbors and cannot orient itself in an external
magnetic fieldP The electron spin has no direct interaction with the
electrostatic field and thus orients itself freely in an external magnetic field.
In the rare earth group, on the other hand, the paramagnetic 4f electrons
lie relatively deep inside the ions, because the outer electrons occupy 5s
and 5p levels. The screening of the 4f electrons from the crystalline field thus
leaves the orbits of the 4f electrons practically the same as in the free ion.
Further experimental evidence for the idea of quenchi.ng of the orbital
momentum in the iron group salts has been obtained from studies of the
anisotropy of the susceptibility in single crystals. The crystalline fields
distort the orbits in particular directions and thus the magnetic field
associated with these orbits has directional properties. The spin magnetic
moment orients itself along the resultant of the external field plus the
field associated with the orbits, and anisotropy results.
]n connection with expression (18-28) it may be noted that at low
temperatures saturation effects are observed which are described accurately
by the Brillouin function ;16 for the iron salts one must, of cou~, use S
rather than J in expression (18-28).
I;
Sec. 18-7)
459
F = -eE - (e/c)v X H
(l8-31)
The spin of the electron will be neglected in this section. Introducing the
vector potential A by means of the relation H = curl A, it can be shown
that (18-31) is equivalent with the following expression for the total energy
(the Hamiltonian) :18
e)2 +V
1 ( p+-A
.Ye=-
( 18-32)
2m
= -k.vH;
A~
= ixH;
Az
= 0
From this we may draw two important conclusions. First, if the electron
motion were associated with a permanent magnetic dipole moment (.L.
this should give rise to a term -(.L' H = -- J.1. z H in the Hamiltonian, in
accordance with (18-4). Thus the second term on the right may be
identified with - J.1. z H, so that
fl.
= --(e/2mc)(xp" -- yp,.)
(18-34)
M,,=rXp
( 18-35)
460
[Chap. 18
where (p2) represents the mean of the squares of the radii of the projections
-NZ(e2 /4mc2)(p2)
-NZ(e2/6mc2)(r2)
where (r2) represents the mean square distance of the electrons relative to
the nucleus. It is observed that this result is identical with (18-20).
18--8. The principle of adiabatic demagnetization19
Because of its importance in obtaining temperatures below IK, we
may briefly indicate the principle of adiabatic demagnetization. The
working substance in this process is a paramagnetic salt. In Fig. 18-6, let
the curve OAB represent the entropy of the system as function of T, in
the absence of an external field. Suppose now that at the temperature T H
T
.~ ,9 This method ,was (irst suggested by P. Debye, Ann. Physik, 81, 1154 (1926) and by
W. F. Giaugue, J. Am. Chern. Soc., ~9, 1864 (1927).
Sec. 18-8]
461
where the last equality follows from one of the Maxwell relations. For
an isothermal process (B to C) one may thus write
dS=(aa~)HdH
,or
S=SH~O+JoH(~~)HdH
(18-38)
= N ft~p:tfH/3kT
so that then
S
S}/~o -
Nft1p;tfH2/6kT~
(18-39)
literature. 2o
REFERENCES
L. F. Bates, Modern Magnetism. 3d ed., Cambridge, London, 1951.
See, for example, N. Kiirti and F, Simon. Proc. Roy. Soc. (Londoll), A149, 152
( 1935).
462
PROBLEMS
[Chap. 18
OJ = wo + eH/2mc = OJo +
when the Larmor frequency w L
(I)L
'
~ Wo'
Chap. 18]
463
oT
H,
~
t
Chapter 19
\
\
I
FERROMAGNETISM,
ANTIFERROMAGNETISM, AND
FERRIMAGNETISM
Ferromagnetism
19-1. Introductory remarks
In ferromagnetic materials the magnetization versus magnetic field
relationship exhibits hysteresis similar to that encountered in Chapter 8
for the relationship between P and E in ferroelectric materials. Of the
elements, only Fe. Ni, Co, Gd, and Dy are ferromagnetic, although there
are a relatively large number of ferromagnetic alloys and oxides (see Table
19-2). Above a critical temperature Of' ksown as the ferromagnetic Curie
temperature. the spontaneous magnetization vanishes and the material
becomes paramagnetic. Well above the Curie temperature the susceptibility follows the Curie-Weiss law,
x=
C/(T - 0)
(19-1 )
Sec. J9-1]
465
FERROMAGNETISM
12r---+---+---+-~~-4~---+---+--~
8
4
0
-4
-8
-12
-16
-20
-.08
-.04
-
.04
.08
H(gaU88)
that for this particular case a very weak field (of the order of 10-2 gauss)
is sufficient to produce a magnetization M = B/47T = 103 gausses. It
should be mentioned that the coercive field for bulk materials may be
several orders of magnitude larger than for the example in Fig. 19-1.
Assuming atomic dipoles of the order of one Bohr magneton (,.._,1O-20 cgs
unit) one verifies readily that values of M of the order of 103 gausses
require essentially a parallel alignment of all the atomic dipoles in the
specimen; hence the saturation of the magnetization in that region. By
way of contrast this may be compared with a paramagnetic solid which in
the same field of 10-2 gauss would give a magnetization M ~ N f-t~H/kT ~
10-6 gauss at room temperature; this is smaller by a factor of 10 9 . Note
2
466
FERROMAGNETISM
[Chap. 19
Hm= H+ yM
Ngp,nlBAx)
(19-3)
...
Sec. 19-2]
467
FERROMAGNETISM
T>8r
_x
Jt will be evident that there must exist a relation between the Curie
temperature 8/ and the molecular field constant y; in fact, one expects
Of to increase with y because the tendency for parallel alignment increases
as y becomes larger. In order to establish this relationship, we make use
of the fact that for x ~ I (near the origin in Fig. 19-2), the Brillouin
function is approximately given by
BAx) ~ (J
+ l)x/3J
(19-6)
Hence, the tangent of curve (19-3) at the origin has a slope equal to
NgflB(J
1)/3. Putting this equal to the slope of curve (19-5) for T = Of'
one obtains
(19-7)
is
, Although the origin in Fig. 19-2 is also a point of intersection, it can be shown that
the free energy of the state with nonvanishing M value is smaller than that for M = 0,
i.e., the latter is unstable.
468
FERROMAGNETJSM
[Chap. 19
saturation, we may employ the approximation (19-6) for BAx), and (19-3)
becomes
(19-8)
M = NgP,B(J + 1)xf3
where x is given by (19-4). Solving for M/H after substituting x into
(19-8) one obtains readily the Curie-Weiss law
x=
(19-9)
M/H= Cf(T- 0)
where C = N p,2/3k and 0 = yN p,2/3k = yc. Note that the value obtained
here for 0 is identical with that obtained for Of from expression (19-7).
In other words, the Weiss theory does not distinguish between the paraand ferromagnetic Curie temperatures.
19-3. Comparison of the Weiss theory with experiment
(19-10)
BAx)
xkT/yNg2p,jP = xT(J
+ 1)/3JO,
(19-11)
469
FERROMAGNETISM
Sec. 19-3]
.8
~
......
.6
Fe
.4
'"
Ni
xCo
f ,,'''_
.2
.4
.6
.8
1.0
T/9{
i.e., the ratio between the magnetic moment and the angular momentum;
for the electron spin g = 2, for the orbital motion g = I. Results of such
experime~ts are given in Table 19-1; they show that the magnetization is
largely due to the electron spins. 5 __ _
Table 19-1. The Magnetomechanical Ratio g for Some Ferromagnetics"
u
g
"
Fe
1.93
1.93
FeaO. (magnetite)
Cu.MnAI (Heusler alloy)
2.00
Co
1.87
Ni
1.92
78~'-;; Ni, 22 % Fe (permalloy)
1.91
a For references to the original literature, see c. KITTEL. Introduction 10
Solid Slale PhysiCS. Wiley. New York. 1953. p. 168.
470
[Chap. 19
FERROMAGNETISM
The effective number of Bohr magnetons per atom. From the saturation
magnetization at T = 0 and the number of atoms per unit volume, one
can calculate the effective number of Bohr magnetons neff per atom.
Values of neff obtained in this way are given in Table 19-2, together with
the ferromagnetic Curie temperature ()f and the spontaneous magnetization.
It is observed that although each atom has an integral number of electrons,
the values of neff are all nonintegral. The reader may at this point be
reminded that for the single ions the number of unpaired 3d electrons is
determined by the total number of 3d electrons in accordance with Hund's
rules as follows: 6
Total number of 3d electrons:
0 I 2 3 4 5 6 7 8 9 10
Number of unpaired 3d electrons: 0 I 2 3 4 5 4 3 2 I 0
Thus for iron, which has 6 electrons in the 3d shell in the ionic state;
one expects on this basis four Bohr magnetons (5 with an "up" spin and
I with a "down" spin). We see from Table 19-2, however, that neff is 2.2.
Table 19-2. Saturation Magnetization, Ferromagnetic Curie Point, and the
Effective Number of Bohr Magnetons per Atom. a For the mixed oxides nerr
is calculated per molecule MOFe 20 3 where M is the divalent metal ion.
Solid
Fe
Co
Ni
Gd
Msat (cgs)
OaK
0, CK)
"e1! (OCK)
1707
1400
485
1752
1446
510
1980
. ..
675
(580)
1043
1400
631
289
105
630
603
506
318
533
745
587
336
2.221
1.716
0.606
7.10
...
Dy
...
MnBi
Cu 2MnAI
CU2Mnin
MnAs
MnB
Mn,N
MnSb
CrTe
CrO,
600
MnOFe~03
358
458
FeOFe.O.
CoOFe 20 3
NiOFe.O.
CuOFe.O.
MgOFe 2O.
430
500
670
147
183
710
240
(600)
870
...
...
...
...
...
...
...
...
...
.. ,
...
240
290
143
'.
Msat (cgs)
room temp.
...
...
KlTTEL~
..
783
848
793
863
728
583
Physics~
...
3.52
(4.0)
(4.0)
3.40
. ..
0.24
3.53
2.39
2.07
5.0
4.2
3.3
2.3
1.3
1.1
Wiley. New York.
Sec. 19-3]
FERROMAGNETISM
471
This discrepancy is not surprising if one recognizes that in the solid the
atomic levels are broadened into bands and that the simple atomic picture
cannot be valid.
Thus M04.t7 and SlaterS explain the nonintegral values for neff on the
basis of a wide 4s band overlapping with a narrow 3d band (Fig. 10-16).
In general, there is on the average a certain fraction of the total number
of 3d and 4s electrons in each band.
l/x --~For example, the fact that iron has
ncH'= 2.2 indicates that in the 3dband
there are 5 electron spins parallel and
2.8 antiparallel. Hence, of the total
of 8 electrons, 7.8 reside on the
average in the 3d band and 0.2 in
the 4s band.
. 1,
The paramagnetic region. Com-T
prehensive experimental studies of
the behavior of Fe, Co, and Ni
above the Curie points have been Fig. 19-4. Schematic representation of
made by Sucksmith and Pearce 9 the behavior of the ferromagnetic metals
above the Curie point; the slight
and by Fallot.lO According to the curvature leads to the distinction
Curie-Weiss law (19-9), a plot of between the ferromagnetic and paraI/x versus T should yield a straight
magnetic Curie points.
line, the intercept along the T-axis
being equal to O. The experiments show that this law is indeed
satisfied with considerable accuracy except in the region close to
the Curie point. In fact, for all three metals there occurs a concave
upward curvature near the Curie point, which leads to the distinction
between the ferrom:lgnetic and paramagnetic Curie temperatures Of and
0, respectively. This behavior is indicated schematically in Fig 19-4. To
illustrate this point, we give here Of and 0 in degrees Kelvin for these
metals. According to Stoner the observed curvature near the Curie point
is consistent with his theory of ferromagnetism based on the collective
electron treatment.n
0,
(j
Fe
Co
Ni
1043
1093
1393
1428
631
650
472
[Chap. 19
FERROMAGNETISM
From what has been said in the preceding section one may conclude
that, apart from certain details, the Weiss field describes the observations
satisfactorily. So far, however, we have not touched upon the problem
of the origin of this field. We shall limit the discussion here to one interpretation, viz., that given by Heisenberg; references to other interpretations
are given below.
First of all, a rough estimate of the required molecular field Hm may
be made as follows. The energy of a given atomic dipole in this field
should be of the order of k8, i.e.,
(19-12)
For a Curie temperature f) ~ 1000 0 K this gives Hut ~ 10 7 gausses. From
this one concludes immediately that the internal field is not due to a
simple dipole-dipole interaction between neighbors, because such fields
would be of the order IlH/a 3 ~ 103 gausses. It may be pointed out here
that in the case of ferroelectric materials the situation is quite different,
because atomic electric dipoles are larger than magnetic ones by a factor
of about 100;12 thus, at least in principle, the molecular field in ferroelectrics may be due to dipole-dipole interaction. We may also point out
that in the ferromagnetic solids the molecular field constant y = Hm/M ~
10 7/103 ~ 104 , which is orders of magnitudes larger than the Lorentz
factor 417/3 which one might expect for a simple model based on dipoledipole interaction.
In 1928 Heisenberg showed that the large molecular field may be
explained in terms of the so-called exchange interaction between. the
electrons_l3 The principle of this explanation may be illustrated by
considering the hydrogen molecule. Let the nuclei be denoted by
a and b, the atomic wave functions by "PI! and "Pb' the electrons by
I and 2. The interaction potential between the two atoms is then,
in a self-explanatory notation,
~ ....
(19-13)
The reader familiar with the elementary Heitler-London theory of
chemical binding knows that the energy of the system may be written
in the form
(I9-14)
E=KJe
12
13
0.92
Sec. 19-4]
473
FERROMAGNETISM
when~
" (19-15)
The plus sign in (19-14) refers to the nonmagnetic state of the molecule
in which the two electronic spins are antiparalleJ. The minus sign corresponds to the case in which the two spins are parallel, i.e., to the magnetic
state. It is evident from (19-14) that the
magnetic state is stable only if Ie is positive, because then (K - Ie) < (K + Ie).
It can be shownl4 that (19-14) may
be written in a more convenient form
which contains the relative orientation
",
of the two spins, viz.,
-
E = const. - 21eSI S2
.~,'~-
--
(19-16)
Co
3.64
Ni
3.94
Cr
2.60
Mn
2.94
Gd
3.1
,. See, for example, F. Seitz, The Modern Theory of Solids, McGraw-Hili, New York,
1940, p. 612.
,. H. Bethe, Handbuch der Physik, Vol. 24/2.
,. J. C. Slater, Phys. Rev., 36, 57 (1930).
474
[Chap. 19
FERROMAGNET]SM
Note that Cr and Mn are not ferromagnetic. One might raise the question
here whether an element with uncompensated spins, which itself is not
ferromagnetic because the rablro value is not favorable, may be combined
with another nonferromagnetic element to form a compound for which
the r"blro value is suitable for ferromagnetism. That this seems indeed
possible is illustrated by the fact that for example MnAs and MnSb are
both ferromagnetic; the lattice constants of these compounds are, respectively, 2.85 and 2.89 A, as compared with 2.58 A for pure Mn. The ferromagnetism of the other alloys given in Table 19-2 can presumably be
explained in a similar manner. We may also mention here that the Curie
point may be shifted by applying high pressures. l7
Because of the importance of the exchange integral, one would like
to relate it to the Weiss constant y and to the ferromagnetic Curie temperature. Although this is a very complicated problem, an approximate
relationship between Ie and y may be found by a simplified procedure
suggested by Stoner. IS We shall assume that the exchange integral is
negligible except for nearest neighbors and that its value is Ie for all
neighboring pairs. In accordance with (19-17) we may then write for the
exchange energy of a given atom i with its neighbors,
V = -2Ie ~ Si' Sj
(19-18)
where the summation is over the nearest neighbors of atom i. The essential
assumption of Stoner is that the instantaneous values of the neighboring
spins may be replaced by their time averages. Thus, if there are z nearest
neighbors, we have
I
'--(,';
'""
(19-19)
Assuming that the magnetization M is along the z-direction, we may write
(19-20)
According to (19-19) and (19-20),
-2zIeSziMjgNflB
(19-21)
18
475
FERROMAGNETISM
Sec. 19-4]
rel~tion
1)/3k
between Of and Je ,
(19-24)
We mentioned in Sec. 19-1 that in order to explain the fact that a piece
of ferromagnetic material may exist in the nonmagnetized state, whereas a
weak magnetic field may produce saturation magnetization in the same
specimen, Weiss introduced the domain hypothesis. Each domain is
spontaneously magnetized, the magnetization being appropriate to the
temperature T of the specimen. The over-all magnetization is given by the
sum "Of the domain vector.s, and thus may vanish under certain circumstances; an example is given in Fig. 19-6a. Magnetization of a specimen
may occur either by the growth of one domain at the expense of another,
i.e., by the motion of domain walls (Fig. 19-6b), or by rotation of domains
(Fig. 19-6c). A representative magnetization curve is given in Fig. 19-7,
indicating the predominant processes in the different regions. We may
note here that originally it was thought that the well-known Barkhausen
jumps were due to the rotation of a complete domain and that the size
of the Barkhausen discontinuities was a measure of the size of the domains.
W. Opechowski, Physica, 4,181 (1937); 6,1112 (1939).
..
P. R. Weiss, Phys. Rel'., 74, 1493 (1948).
21 F. Bloch, Z. Physik, 61, 206 (1930).
22 See the papers by C. Kittel, C. Zener and R. R. Heikes, J. C. Slater, E. P.
Wohlfarth, and J. H. van Vleck held at the Washingt6n Conference on Magnetism,
Revs. Mod. Phys., 25 (1953); see also J. H. van Vleck, Ret's. Mod. Phys., 17, 27 (1945) .
19
20
~.
476
FERROMAGNETISM
[Chap. 19
'--.
-H
Y
....------71
"Hard"
/'directWn
/
/
Wall motion
Nonmagnetized
Domain rotation
leI
(a)
Fig. 19-6. The domain structure (a) corresponds to the nonmagnetized state; (b) represents magnetization due to wall motion;
in (c) the magnetization is due to rotation of the domain vectors
from an \'easy" to a "hard" direction (see Sec. 19-6).
I
II
I
r
surface of the specimen; since there are strong local magnetic fields near
the domain boundaries, the particles collect there and the domains may
be observed under a microscope.
23 H. J. Williams and W. Shockley. Phys. Rev., 75, 178 (1949).
2. F. Bitter, Phys. Rev., 38, 1903 (1931) .
'
477
FERROMAGNETISM
Sec. 19-5]
~~~
N S N S
I
I
I
I
I
I
,"
I
I
I
t I, ~ I, t I, ~
I
,
I
I
,
,
t'I tI II t II t, ,'t
I
I : :
S..._.,..._.,..._....
N S N
(el
(al
Fig. 19-8.
478
FERROMAGNETISM
[Chap. 19
determined by the anisotropy of the crystal, i.e., by the fact that ferromagnetic materials have "easy" and "hard" directions of magnetization.
For example, from the magnetization curves represented in Fig. 19-9 one
sees that in iron, which is cubic, the easy directions of magnetization are
the cube edges. 25 In 'nickel, which is also cubic, the easy directions of
magnetization are the body diagonals. In cobalt the hexagonal axis of the
crystal is the only preferred direction; thus in a cobalt crystal with
prominent domains magnetized along the hexagonal axis, the closure
18
15
12
t
9
~/
L L
:5
.....
[1001
[1111
r
Iron
ISC
.6
Hjloo
Fig. 19-9. Magnetization curves at 18e for a single crystal of iron
for different directions of the field relative to the crystal axes.
[After Piety, ref. 25]
Sec. 19-6J
479
FERROMAGNETISM
Thus, for a cubic crystal, let Cl I , oc 2 , and OC 3 represent the direction cosines
of the magnetization referred to the cubic crystal axes. Because of the
cubic symmetry, the anisotropy energy should be an even power of each
Cl; furtherdrore, it should be invariant for interchange between the oc's.
The lowest-order combination satisfying these conditions is (oci + oc~ + ocn
but since this is identically equal to unity, it does not enter in the anisotropy'
effects. The next order combination is (ociocI + ocioc~ + oc~ocn; although
this term by itself represents the experimental results for iron and nickel
reasonably well, one usually adds one more term, viz., ocioc~oc~. Thus for
cubic crystals the anisotropy energy may be written as
( 19-27)
when higher terms are neglected. The constants KI and K2 can be determined from experiment; for iron at room temperature, )
K2 = 1.5
>(
10'; ergs/cm 3
(19-28)
For crystals with a single preferred axis, such as cobalt, the anisotropy
energy may be written in the form
(19-29)
where 4> is the angle between the magnetization and the easy axis; higher
order terms are usually neglected. For cobalt at room temperature,
K2 = 1.0
10 6 ergs/cm 3
(19-30)
2" J,
480
FERROMAGNETISM
[Chap. 19
1
,
Fig. 19-10.
(19-31)
Hence the energy decreases when N increases. This raises the question:
Why does not the wall become infinitely thick? It is at this point that
the influence of the anisotropy energy must be considered. Since the spins
within the wall are nearly all directed away from the easy axes, one expects
an anisotropy energy which is approximately proportional to the thickness
of the wall. This has the effect of limiting the wall thickness, as may be
seen from the following arguments. Let us consider a wall of 1 cm 2 area,
the thickness being Na, where a is the lattice constant. The total wall
energy per cm z may then be written in the form
a
2'
2"
aeJ:
+ aau
.(.
...
(19-32)
,i
Sec. 19-7]
FERROMAGNETISM
481
c::::'
300
or
c::::'
1000 A
where t is the thickness of the wall. We may note that the domain walls
in ferroelectric materials are only a few Angstroms thick, as we have seen
in Chapter 8.
The total energy per cm2 of a Bloch wall may be estimated by substituting for N from (19-34) into (19-33). This gives
= 2S4>o(J,Xja)1/2
(19-35)
which for iron turns out to be of the order of I erg per cm 2 We should
emphasize that the above treatment is rather crude; for example, due to
the anisotropy, tile angle between successive spins is not constant
throughout the Bloch wall.
We may mention here that there exists a critical size of ferromagnetic
particles below which the single domain configuration is more stable than
a multidomain structure; the critical size is determined by the anisotropy,
the shape of the particles and the intensity of the magnetization. For
spherical iron particles the critical radius is of the order of 10-6 cm.
The calculations are given in C. Kittel, Revs. Mod. Phys., 21, 541 (1949).
Similar calculations have been carried out for the critical single domain
size offerrites by Morrish and YU. 29 For a more detailed and mo~ecomplete
treatment of the energy considerations entering in the discussion of domain
formation we must refer the reader to Kittel's paper,
19-8. Coercive force and hysteresis
482
FERRor..iAG NETISM
[Chap. 19
~
A
Position of wall
;;
!
. \
Fig. 19-11. Schematic representation of the energy of a ferromagnetic specimen as function of the position of a domain
wall.
Bloch wall. Such motions may be vizualized with the aid of a potential
curve, as indicated in Fig. 19-11; the curve represents the energy of a
Bloch wall as function of its position in the crystal. The variations in
energy are a consequence of local strains, impurities, lattice defects, etc.
In the absence of an external field, the wall will be in a position corresponding to an energy minimum, say in A. Application of the field will
modify this curve and unless the field is large enough to help the wall
climb across a maximum such as B, only a small reversible wall displacement will result. For larger fields, the wall displacement may be large,
but irreversible; this corresponds to the region A B in Fig. 19-7. The
domain rotations occurring in the region Be of Fig. 19-7 take place when
the applied magnetic field does not coincide with an easy direction of
magnetization, i.e., work must be done against the anisotropy forces.
The above qualitative picture explains the fact that the coercive force
increases with an increased intensity of local internal strains. The
observation that alloys containing a precipitated phase are magnetically
hard (high He) is also consistent with this picture. The quantitative
aspects are, however, quite complicated. 30
.0
For details, see R. Becker, Physik. z., 33, 905 (1932); M. Kersten, Grundlagen
einer Theorie der ferromagnetischen Hysteresis lind Koerzitivkraft, Edwards, Ann Arbor,
(1943) L. Neel, Ann. univ. Grenoble, 22, 299 (1946); E. C. Stoner and E. P. Wohlfarth,
Phil. Trans., A240, 599 (1948).
..
483
Sec. 19-9]
AntiferromagnL tIsm
\
.v
'_
484
ANTIFERROMAGNETISM
[Chap. 19
The most direct experimental evidence for the basic picture of antiferromagnetism has been obtained from neutron diffraction experiments. 35
When neutrons are incident on a crystal they are scattered by the atomic
nuclei but also by the interaction between the neutron spin and paramagnetic ions which may be present. Consequently, the ordered antiferromagnetic state gives rise to "extra" diffraction lines just as one
observes extra X-ray diffraction lines for ordered alloys. The intensity of
these extra lines decreases as the temperature increases because the antiferromagnetic order diminishes. Above the antiferromagnetic temperature
the extra lines disappear. An example has been given already in Fig. 1-17
for MnO.
i
. l
(19-36)
= H - f3 M a - rxMb
where H is the applied field, and M" and Mb represent the magnetization
of the A and B lattices; rx and f3 are positive Weiss constants. We shall
consider two temperature regions:
(i) T > TN' When the temperature is above the Neel temperature, we
are far away from saturation, and the magnetization of the A lattice may
be written
M" = (NfJ,2f3kT)H"
with
/-,2 =
/-,}Jg2J(J
+ I)
(19-37)
Sec. 19-10]
485
ANTI FERROMAGNETISM
c
T+O
(19-40)
-9
_T
I/X
TIC
ferro
I/x
= (T-(J)/C
alltilerro
I-Ix
= (T
+ (J)/C
(19-41)
(ii) The region be/ow the Neel temperature. At the NeeI temperature TN
itself, one is still sufficiently far away from saturation effects to employ the
equations given above for Ma and M b Thus in the absence of an applied
magnetic field we may write for T = TN in accordance with (19-37),
or
+ N fl2J3kTfI')rx]M + (Nfl2J3kT.'I.)f3M = 0
(Nfl2/3kTN )fJM" + [I + (Nfl2/3k-T_y)rx]Mb = 0
[(I
Similarly,
II
( 19-42)
(19-43)
486
ANTIFERROMAGNETlS~
[Chap. 19
The last two equations have a nonvanishing solution for Ma and Mb only
if the determinant of their coefficients vanishes. Making use of the fact
that 2N fl2/3k = C one finds that
TN = C(P -_ oc)/2
(19-44)
+
TN/ e =
({J - oc)/({J
+ oc)
\
(19-45)
Compound
Crystal
structure
MnF2
FeF2
CoF2
NiF.
MnO,
MnO
MnS
FeO
CoO
rutile
rutile
rutile
rutile
rutile
NaCl
NaCl
NaCl
NaCI
Cation lattice
structure
b.c.
b.c.
b.c.
b.c.
b.c.
tetragonal
tetragonal
tetragonal
tetragonal
tetragonal
f.c.c.
f.c.c.
f.c.c.
f.c.c.
TN
("K)
72
79
38'
73
84
122
165
198
292
()CK)
X.tx T N
113
117
53
116
316
610
528
570
280
0.76
0.72
...
...
0.94
0.67
0.82
0.79
...
ANTlFERROMAGNETISM
Sec. 19-10]
thems~lves.
487
l.
(bl
Fig. 19-14. Illustrating the calculation of Xi' as described in the
text, for an antiferromagnetic arrangement of dipoles.
In the present case, the field tends to line up the dipoles along the field
direction, but as a result of the tendency for the A and B dipoles to remain
antiparallel, a compromise is obtained in which the dipoles make a certain
angle cp with the original spin direction. To calculate the susceptibility
Xi for this case, we proceed as follows. Consider one of the dipoles B as
made up of two unit poles, as indicated in Fig. 19-14b. The forces on the
positive pole are Hand -f3M,,, as indicated; the forces on the negative
pole are equal but of opposite sign. In equilibrium, the resultant forces
should lie along the line joining the poles, so that for small angles cp we
must have
,
t
2f3Ma cp = H
I
Since Mil = M,), the total magnetization along the external field direction
is equal to
so that
'
Xi =
liP
(19-46)
488
ANTIFERROMAGNETISM
"- \
[Chap. 19
The reader will have noticed that nowhere in the above derivation for
X.1 did we introduce an argument that explicitly referred to the existence
of a natural spin direction; we considered only the balance between the
force produced by the external field and the
exchange force between nearest neighbors.
A simple way in.which the existence of a
natural spin direction might be introduced
is indicated in Fig. 19-15 for one of the B
dipoles. It is assumed that there exists a
constant field Han which by itself tends to
keep the dipole in the "natural" spin
Fig. 19-15. The resultant of the
three forces shown should coin- direction, i.e., Han is an anisotropy field.
cide with the line joining the two If one then considers the equilibrium of
poles, as described in the text.
forces in the presence of an external field
H l_ Han' one must require that the
resultant of H, Han> and the exchange force -f3Ma lie in the dipole
direction. We leave it as a problem for the reader to show that, as long
as the angle cp is small, one finds in this case
1
X -----.L f3
Han/2Ma
(19-47)
A few remarks may be made here about the nature of the interaction
in antiferromagnetics of the NaCl structure, such as MnO. From the
neutron diffraction experiments by Shull and Smart35 on MnO one
3. J. W. Stout and M. Griffel, J. Chem. Phys., 18, 1455 (1950).
Sec. 19-11]
ANTIFERROMAGNETISM
489
concludes (see Fig. 19-17) that the stronges~ negative interaction for a
given Mn2+ ion does not come from its nearest Mn2+ neighbors but from
those Mn2+ ions which are at a distance y2 times as far. In fact, the
lot.
x
negative interaction takes place between those Mn2+ ions which are
separated by an 0 2- ion such that the angle Mn2+ - 0 2- - Mn2 + is 180.
Since the overlap between the 3d electrons of these Mn2+ ions is negligible
o Mn2+
0
02 -
...
490
ANTIFERROMAGNETIS~
[Chap. 19
'>
Anderson38 and van Vleck ;39 in simple terms, the nature of this type of
interaction may be understood qualitatively as follows: the description of
manganese oxide as a completely divalent ionic compound of the type
Mn2+02- is inadequate in the sense that one should include in the wave
functions terms corresponding to Mn+ and 0- ions; we may call Mn+Oan excited state of Mn2+02-. The electron configuration of the ground
state (Mn2+02-) for the two types of ions involved may be represented by
\
The 0- ion has evidently a resulting spin which has the same direction as
that of the Mn+ ion to which the electron has been added. Suppose now
that on the right-hand side of the 0- ion another Mn2+ ion is located; as
a result of the spin of the 0- ion the magnetic moment of this Mn2+ ion
will have a tendency to be lined up antiparallel to that of the 0- ion if
the interaction between these ions is antiferromagnetic (negative exchange
integral, i.e., not too large separation between the ions). Hence, on this
assumption one obtains an antiparallel alignment between the two Mn2+
ions as a result of the presence of the 0 2- ion between them. The angle
of 180 is particularly suitable for this type of interaction because of the
dumbbell shape of the 2p wave function involved. This type of interaction
may also play an important role in the antiferromagnetic interactions in
I
ferrites, as we shall see below.
0
Ferrimagnetism
!I
3' P. W.
39
FERRIMAGNETISM
Sec. 19-11]
491
the trade name Ferroxcube. 40 The most important of these are the MnZn
ferrites (Ferroxcube IV). The doc resistivity of ferrites is 104 to 1011 times
as large as that of iron. Thus in transformer cores they can be used up
to much hi~her frequencies than iron.
19-12. The structure of ferrites 41
The physical properlies offerrites are intimately related to the structure
of these solids. They belong to the large class of compounds which have
the spinel structure (after the mineral spinel, MgAI204). The oxygen ions,
with a radius of 1.32 A form, to a good approximation, a close-packed
cubic structure. The unit cell contains 32 oxygen ions, 16 Fe3 + ions, and
8 divalent metal ions. The total of 24 metal ions, ranging in radius
between 0.4 and I A, are distributed amongst eight tetrahedral interstices
(surrounded by four 0 2- ions) and sixteen octahedral interstices (surrounded by six 0 2- ions). The distribution of the metal ions is very
important for an understanding of the magnetic properties of these
materials; the following distributions may occur.
(i) In the "normal" spinel structure of a ferrite the 8 divalent metal
ions occupy tetrahedral positions; the 16 trivalent iron ions occupy
octahedral positions. We shall follow a usual notation for this structure:
Me2+[Fe~+]04 the brackets around the Fe3+ ions indicating that they
occupy octahedral sites.
(ii) In the "inverse" spinel structure of a ferrite, the divalent Me2+
ions occupy octahedral sites; the Fe3+ ions are distributed in equal
numbers over the tetrahedral and octahedral sites. The arrangement may
thus be represented by Fe3+[Fe3+Me2+J04.
(iii) In the intermediate case we have arrangements of the type
.0
the
the
the
for
See J. J. Went and E. W. Gorter, Philips Tech. Rev., 13, 181 (1952); J. J. Went,
G. W. Rathenau, E. W. Gorter, and G:W. Oosterhout, Philips Tech. Rev., 13, 194 (1952) .
., For a comprehensive review, see E. W. Gorter, Philips Research Repts., 9, 295
(1954); also F. C. Romeyn, "Physical and Crystallographic Properties of Some Spinels,"
Thesis, Leiden, 1953 .
.. E. W. Gorter, Philips Research Repts., 9, 321 (1954).
',. .
,' .
.v
492
FERRI MAGNETISM
[Chap. 19
'I
Since the ferrites are essentially ionic compounds, one would expect
8
7
6
iJ.B
Mn
Fe
Co
Lio,5 + FeO,5
Ni
2
1
Mg
.2
,6
.4
Composition
.8
1.0
ZnFe204
Zinc ferrite and cadmium ferrite, which are known to have the normal
spinel structure, are paramagnetic, All other known simple ferrites which
are ferromagnetic have the inverted spinel structure, and it thus seems
that ferromagnetism is associated with the inverted structure, The rather
peculiar magnetic properties may further be illustrated by noting that
according to Fig. 19-18 the replacement of paramagnetic ions such as
Fe2+, C02+, Mn'H by the diamagnetic Zn'l.+ ions leads to an increase in
the saturation magnetization, at least for small zinc concentrations.
We may also mention that when one plots the reciprocal of the
susceptibility versus temperature above the Curie point, one frequently
obtains a concave curvature towards the T-axis, rather than a straight line
predicted by the normal ~urie-Weiss law.
FERRIMAGNETISM
Sec. 19-14]
493
xMa
+ (2 -
X)Mb
(19-49)
We shall first consider the paramagnetic region above the Curie point.
.3 L. NeeJ, Ann. phys., 3, 137 (1948).
,}
:,
rChap. 19
FERRI MAGNETISM
494
(J
XO
T- ()
(19-53)
-=-=-+---Xmole
Cm
where
1
- = (y/4)[2x(2 - x) - ax2
Xo
(J
fJ(2 - xy>']
+ a) x)(2 + a + fJ)
() = tyC m x(2 -
(2 - x)(1
+ fJ)]2
lal
(19-54)
I
Mb = NgSflBBS(gSflBHb/ kT ) l
(19-55)
, '------..-
Sec. 19-14]
495
FERRI MAGNETISM
------~-~.
zinc ferrite (which has the normal spinel structure). The other divalent
ions Mn 2+, Ni 2+, etc. occupy octahedral sites, and the Fe3+ ions are
distributed over the remaining tetrahedral and octahedral sites. Thus the
mixed zinc ferrites satisfy the representation
For low zinc concentration there are a sufficient number of Fe3 + ions in the
A sites to Cause all the magnetic moments in the B sites to remain paraIleI
(due to AB interaction). Hence for low zinc concentrations the saturation
magnetization will increase with increasing Zn2+ concentration (because
M A decreases relative to Mb)' In fact, the slope of the magnetization
versus composition (x) should be such that for x = I the intercept should
give 10,uB' This is represented by the dashed straight lines in the figure.
The fact that the actual magnetization falls below these curves is a result
of the continually reduced AB interaction; the BB interactions then take
over, favoring an antiparallel alignment of the B atoms. Finally, for x = I,
we have the pure zinc ferrite with vanishing saturation moment.
.. E. W. Gorter, Philips Research Repts., 9, 321 (1954).
4. C. Zener, Phys. Rev., 81, 440 (1951); 82,403 (1951).
~.
to
496
[Chap. 19
FERRI MAGNETISM
REFERENCES
EI~evier,
19-3. Show that equation (19-14) can actually be written in the form
(19-16). See for example F. Seitz, Modern Theory of Solids, McGraw-Hill,
New York, 1940, p. 612.
Chap. 19]
497
FERRI MAGNETISM
19-4. Give a discussion of the collective electron theory of ferromagnetism. See, for example, J. C. Slater, Quantum Theory of Matter,
McGraw-Hill, New York, 1951, Chap. 14, Appendix 22.
N
:.
-,
I
Chapter 20
Paramagnetic Relaxation
j
(20-1)
(Me - M)/T
+ Mo cos (wt -
9')
(20-3)
Sec.lO-J]
499
'l..8He
(20-4)
where X"/X' = tan cpo For low frequencies X" = 0 and x' =--' Xs' The
frequency-dependence of X' is called dispersion, in analogy with the optical
case; X' is referred to as the high-frequency susceptibility, for obvious
reasons. The quantity X" determines the absorption of energy by the
specimen. In fact, making use of (20-2) and (20-3), one finds for the
absorption per second per unit volume,3
A
= (W/27T)
cp M dH =
(w/2)X" H~
(20-5)
(20-6)
it follows that
(20-7)
x* =
where
X' - iX"
is the complex susceptibility. We may note in passing that x' and X" are
related to each other by the so called Kramers-Kronig relations 4 (see
Problem 20-3). Also, both x' and X" are functions of H, as well as of
frequency.
So far, the description has been completely phenomenological. The
task of the theory of paramagnetic relaxation is twofold:
(a) The quantities '/ and X" should be related to the relaxation times
mentioned above.
(b) The relaxation times must find an interpretation based on
the properties of the magnetic atoms and the lattice in which they are
incorporated.
20-2. Relaxation mechanisms
In order to get an insight into the problem, consider a system of free
magnetic dipoles, oriented at random. Suppose the dipoles have no
C. 1. Gorter, Physica, 3,503 (1936).
.,
The reader is reminded that by replacing in the "normal" thermodynamic expressions the pressure by M and the volume by H, one obtains expressions appropriate to
the magnetic case.
, H. A. Kramers, AlIi cOI(f{r.fis., Como. 545 (1927); R. Kronig. J. Opl. Soc. Amer.
12,547 (1926).
2
. ~>
;".:.
500
[Chap. 20
Sec. 20-2]
501
(20-9)
Although these equations describe the observations qualitatively, better
agreement is obtained with a thermodynamic theory developed by Casimir
and DuPre.l The basis of this theory is that the relaxation time associated
with the spin-lattice interaction is so long compared to the spin-spin
relaxation time that the spin system can be considered to be always in
thermodynamic equilibrium. Thus the spin system is treated as a thermodynamic system, separate from but in energy contact with the lattice.
The spin system has its own specific heat, temperature, etc. In contrast
with this, the previous theories mentioned above considered the individual
spins in energy contact with the lattice. That the temperature Ts of the
spin system is not necessarily the same as that of the lattice may be seen
as follows. Consider a system of spins of t in thermal equilibrium with
the lattice in an external field H at a temperature T. According to
Boltzmann statistics we then have Np/Na = exp (2P B H/kT), where Nil and
Na refer, respectively, to the number of spins parallel and antiparallel to H.
9 C. J. Gorter and R. Kronig, Physica, 3, 1009 (1936); R. Kronig, Physica. 5, 65
(1938).
10 H. B. G. Casimir and F. K. DuPre, Physica, 5, 507 (1938).
502
[Chap. 20
When H is suddenly increased to H', the temperature of the lattice remaining T, it takes some time for the ratio Np/Na to adjust itself to the new
field. However, substituting H' for H in the Boltzmann distribution above,
one can define a certain temperature T. such that the instantaneous
populations satisfy the Boltzmann expression; in this case, Ts would be
the spin temperature. In the above example Ts > T immediately after the
increase of the field; as time goes on, Ts approaches T. In an oscillating
field, the difference () = Ts - T will also oscillate, the amplitude of the
oscillation becoming smaller as the heat contact between the spin system
and the lattice becomes better. As long as () is not too large, the heat
transferred from the lattice to the spin system during a short time interval
dt may be written
dQ = -oc() dt .
(20-10)
where the quantity oc may be called the coefficient of heat contact between
the spin system and the lattice. On the other hand, the first law of thermodynamics for the spin system may be written in the form
(20-1 I)
where ClI and C111 are the specific heats at constant Hand M, respectively.
For a field of the type (20-6) we may write
H(t)
M(t)
+ Hoe
Me + Moiw'
",co
()
()
iw1
H,
oeiWI
so that
(20-12)
: '.
(20-13)
By eliminating 00 from the last two equations one obtains for the complex
susceptibility,
x*
Mo = (OM)
oH,
Ho
l'
I
1
+ (iw/oc)C
(20-14)
lIl
+ (iwjoc)Cf[
x."
(20-15)
x"
x.
where the relaxation time
].1X
7
)l"h{J;"'-
(20-16)
Sec. 20-3]
503
heat contact between the spin system and the lattice. Note that the highfrequency susceptibility X' contains, besides the Debye function (20-9), a
constant part equal to C JIICII' A typical example of a set of dispersion
~ .6~----+---~~~~---+---=~~~==~
-><
t
1
2
3
4
.2
H=800
H=1600
H=2400
H=3200
O~----~--
. .1
______
.2
______L -____
.5
1.0
________
2.0
5.0XI06
Frequency (cps)
1 H=4OO
2 H=1600
3 H=3200
_.-'"
_.-- --
.2
.5,
1.0
2.0
5.0
lOX 106
Frequency (cps)
,.'
"':-'",.
504
[Chap. 20
"See, for example, L. J. F. Broer, Physica, 10, 801 (1943); A. Wright, Phys. ReP.,
76, 1826 (1949); J. H. van Vleck, Phys. Rev., 74, 1168 (1948).
,
",
Sec. 20-4]
505
spin to change its axis of precession will be of the order of T .. ~ l/(J) I.' That
this gives indeed the order of magnitude cim be seen from the table below.
Table 20-1. Some Spin-Spin Relaxation Times and Internal Fields"
Solid'
H, (gausses)
I
I
Fe(NH,)(SO')2' 12H 2 0
Gd 2(SO,) .. 8H 2 O
Cu(NH,).(SO')26H20
CuSO .. 5H 2 O
'T~ exp
450
1380
200
370
1.1
0.57
9.1
6.7
l/wL c.. lc
(in 10~lO sec)
1.2
0.4
2.7
1.5
a'Reproduced with kind permission of the Physical Society. London. from A. H. Cooke. Nepts. Progr.
Ph),s .. 13. 276 (1950).
_........--.-'~
/'
(20-18)
(20-19)
(compare 18-11).
The magnetic moment fA. associated with Ma is given by
(20-20)
in analogy with (18-12). Here M'IJ is the proton mass and g is the inverse
of the gyromagnetic ratio. The maximum component of fL along an
applied field H is thus equal to
(20-21)
506
[Chap. 20
-rr-----1=3/2
gJln H
-3/2
- - ' - - - - - - - -1/2
,:::
J:>;l
t -------1/2
------3/2
Sec. 20-6]
507
--'1-'-
Hy == Hz = 0
(20-24)
H.,= Hocoswt;
H y = Hosinwt;
Hz=O
left
Hx = Ho cos wI;
Hz = 0
(20-25)
If
It must be realized that the excess number in the lower level is usuallYI
i
508
[Chap. 20
very small indeed; for protons for example g = 5.58 and one finds with
H -:::: 104 gausses at room temperature, Np/Na -:::: 1 7 X 10- 6
Hz = He;
Ho
< He
(20-27)
Consider first the influence of a field H alone, on a single nuclear dipole tL.
In accordance with (18-17) we may write
(20-28)
, I
i.e., the field alone simply leads to a Larmor precession of tL about H.
Adding the effect of all dipoles per unit volume, we may write for the
rate of change of M due to the field alone,
==
yM X H
(20-29)
"':f'
,I
(20-30)
Besides the influence of the field, two other sources contribute to the
rate of change of M, viz.,
(i) The spin-lattice interaction
(ii) The spin-spin interaction
Their influence will now be considered. Suppose that Me represents the
magnetization along the z-direction if the system is in thermal equilibrium
when only the constant field If.e is applied. When this field is suddenly
switched off, the magnetization will gradua!ly approach zero. Similarly,
when the field is suddenly switched on, a certain time interval is required
to obtain the equilibrium value Me. During this build-up, a certain
fraction of the dipoles must flip over from an antiparallel to a parallel
orientation relative to the field. Since this process requires a change in
energy of the spin syste~? the build-up time is determined by the heat
Sec. 20-7]
509
contact between the spin system' and tht:_ lattice. Thus, as a result of the
spin-lattice interaction the rate of change of the z-componem of M is
assumed to be given by
...
(oMZ/ot)'1
-(Mz
Mc)h
(20-31)
where the subscripts sf refer to the spin-lattice interaction; the characteristic time Tl is the spin-lattice relaxation time. Combining (20-31) with the
z-component of equation (20-29) we obtain for the total rate of change
of M.,
dMz/dl = y[-M",Ho sin wI - MyHo cos wI]
+ (Me -
Mz)h (20-32)
This is one of the Bloch equations; the two others provide expressions
for the rate of change of the transverse components M", and My. In order
to set up the expressions for Mx and My, it is important to realize that if
there were a completely random distribution of the x and y components
of the nuclear dipoles, M x and M" would be zero. In other words, one is
interested in the lifetime associated with a certain Mx or My value in the
absence of an applied oscillating field. Now, consider two neighboring
identical dipoles i and j. Since both are precessing about He' then j will
produce an oscillating field of the Larmor frequency at the position of i
and vice versa. Consequently, transitions may take place in which i and j
simultaneously reverse their orientation (spin exchange), thus limiting the
lifetime of each state. Since the interaction energy !}.E is of the order of
~;'/r3, the lifetime T2 as given by the Heisenberg uncertainty principle is
T2 ~ Il/!}.E = Ilr 3 / ~;.. The characteristic time T2 was introduced by Bloch
as the spin-spin relaxation time. He thus assumed that the rate of change
of M", and My as determined by the spin-spin interaction is given by
(oMx/ot)ss = -M,JT2
and
(20-33)
From a different point of view, one may argue that a given dipole sees,
besides the applied field, an internal field Hi""'" ~n/r3 produced by its
neighbors. Thus one expects a spread in the precession frequencies !}.wL
where, according to the res~nance condition (20-23),
!}.wL ~ g~nHdll ~ g~~/Ily3
(20-34)
(20-35)
(20-36)
510
[Chap. 20
Since the constant field and the oscillating field are applied, respectively,
in the z and x-directions, one is particularly interested in solutions of the
Bloch equations (20-32), (20-35), and (20-36) for M z and M.,. Without
\
giving the mathematical details here, one ob!ains 18
(20-37)
(20-38)
where Xe is the static susceptibility given by the Curie law. From the
definition of the complex susceptibility it follows (see 20-4) that
(20-39)
(20-41)
The reader is reminded that the absorption of radio frequency energy is
determined by X", as expressed by formula (20-5).
In discussing the results obtained, one may distinguish between two
cases;
(i) The amplitude Ho of the oscillating field is so small that y2R~'Tl'T2 ~ 1.
In this case M z is simply equal to XeRe, i.e., equal to the static equilibrium value. This means that the spin-lattice relaxation is rapid enough to
maintain a Boltzmann distribution of the population in the various energy
levels, notwithstanding the fact that, because of radio frequency absorption,
an excess is thrown from lower to higher levels. For this case, the frequencydependence of x' and X" is represented in Fig. 205. Note that the half
width of the absorption line under these circumstances is determined by
the spin-spin relaxation time 7 2 ; it is, in fact, equal to 1/72 in terms of an'
angular frequency.
(ii) When y2HJ71'T 2 is not negligible compared to unity, M z < XeHe
and both X' and X" are reduced'in magnitude. In this case one speaks of
saturation of the spin system; the spin-lattice relaxation is not able to
maintain a Boltzmann distribution of the populations in the energy levels
under these circumstances. In other words, the spin temperature increases
beyond the lattice temperature as a result of the rapid rate of absorption
18
See F. Bloch, Phys. Rev., 70, ..460 (1946); G. E. Pake, Am. J. Phys., 18,438 (1950).
"
iV,,,
Sec. 20-7]
511
-4
-3 -2 -1
4
_(wL-wl1"2
Fig. '20-5. The real (X') and imaginary (X") part of the complex
susceptibility as function Of(WL - WP2, pertaining to the case of
negligible saturation.
20
512
[Chap. 20
21
22
23
Sec. 20-8]
513
moments, and are a very efficient medium for establishing heat contact
between the nuclear spins and their surrou-ndings.
The spin-spin relaxation time also depends on T e , as is illustrated in
Fig. 20-6. L-et us consider a solid at very low temperature where T e is long
because atomic jumps are rare. The spin-spin relaxation time will then
have some small limiting value, say
10-6 sec. As the temperature is raised, 10- 3 . . . - - - - - - - - - - - - . . , .
T2 wi!! remain constant, and so will
the line width in accordance with
(20-41), until Te has been reduced to 10-1
a value of the same order as T2' As
the temperature is increased further,
the number of spin exchanges per
10- 5
unit time decreases, since atoms are
nearest neighbors for a time T c' which
is smaller than the lifetime of the
spin states. Hence T2 begins to in10-3
10- 15
10- 11
crease and will continue to do so as Te
decreases; in the region of To values
Fig. 20-6. 7"1 and 7"2 as function of the
where both T1 and T2 increase with correlation time 7",; 7"1 is given by
decreasing T c , the values of T1 and T2 (20-43). [After Bloembergen, Purcell,
are approximately equal.
and Pound, ref. 17)
20-9. Some applications to solid state physics
Nuclear magnetic resonance experiments have become a powerful tool
in studying the physical properties of solids. Although it is not possible
to go into details here, a few examples may be given here to illustrate this.
(i) Structural studies.24 The width and structure of a resonance
absorption line are influenced by the magnetic interaction between the
dipoles. Since. this interaction is determined by the relative positions of
the nuclei, the width and shape of the lines provide information about
the structure of solids. The simplest case is encountered in solids where
the nuclei occur in single pairs, so that the effective magnetic field at the
position of a given nucleus is determined by the applied magnetic field He
plus the internal field produced by its partner. With reference to Fig. 20-7
let the vector r, joining two such nuclei, make an angle () with the applied
field He in the z-direction. In order to calculate the effective magnetic
field at the site of nucleus b we shall start from the classical formula for
the field produced by a dipole !Joa at a point r:
(20-44)
514
[Chap. 20
(ii) Molecular rotation in solids. 27 In the liquid and gaseous states one
usually deals with narrow absorption lines, which are generally well
resolved; this is a result of the fact that in these cases 'T e is small, resulting
in a large value of'T2 (and 7 1 ), In solids on the other hand, 'Te is large,
'T2 is small and therefore the bands are broad. In certain solids, however,
G. E. Pake, J. Chern. Phys., 16, 327 (1948) .
H. E. Petch, D. W. L. Smellie and G. M. Volkotf, Phys. Rev., 84, 602 (1951);
Can.J. Phys., 30, 270(1952); G. M. Volkotf, Can.J. Phys.,31, 820(1953); H. E. Petch,
N. G. Crana, and G. M. Volkotf, Can. J. Phys., 31,837 (1953).
27 H. S. Gutowski and G. E. Pake, J. Chern. Phys., 18, 162 (1950).
'" .~
{ ',.,
.\
Sec. 20-9]
515
'"
aba.
:!
1
l __
.------.
o
-5
-10
10 gauss
oH
..
..
100
_.
200
-T(OK)
of vacancies in the sqdium lattice for reasons explained above. From the
transition temperature and the slope of the curve, Gutowski arrives at an
activation energy for the self-diffusion of 9.5 1.5 kca1. 28 According to
an analysis by Norberg and Slichter, the diffusion coefficient and its
temperature dependence determined from nuclear resonance experiments
516
[Chap. 20
Nucleus
Magnetic
moment
gl
neutron
HI
LF
Na 23
AP7
Cu 3
Cu"
CI35
1/2
~1.9135
1/2
3/2
3/2
5/2
3/2
3/2
3/2
2.7935
3.2571
2.2178
3.6419
2.2266
2.3850
0.8222
Resonant frequency
for H = 10' gausses
in megacycles/sec.
29.1
42.6
16.5
11.3
11.1
11.3
12.1
4.2
--
29 R. E. Norberg and C. P. Slichter, Phys. Rev., 83, 1074 (1951); see also R. E.
Norberg, Phys. Rev., 86, 745 (1952).
30 H. C. Torrey, Phys. Rev., 92, 962 (1953).
31 W. D. Knight, Phys. Rev., 76,1259 (1949).
32 Townes, Herring and Knight, Phys. Rev., 77,852 (1950).
33 N. Bloembergen and T. J. Rowland, Acta Metallurgica, 1, 731 (1953).
Sec. 20-11]
517
= g(e/2mc)Hc
(20-46)
Since the electron mass is ,....._, I 03 times smaller than M p' the resonance
frequencies for the same field are ,....._,103 higher than for nuclear resonance.
For a free electron g = 2.0023, and in that case WL = 2.8026H megacycles
per second when H is expressed in gausses. Paramagnetic resonance was
first observed by Zavoisky on the paramagnetic salt CuCi 2 2H 20.34
Studies of paramagnetic resonance in crystalline solids have provided a
great deal of accurate information about the crystalline electric fields.
A summary of this has been given in a paper by Bleany and Stevens. 3S
Other investigations have been concerned with free radicals, trapped
electrons, conduction electrons in metals, and excited molecules. We shall
confine ourselves to some remarks iIi connection with color centers and
donor levels.
We have seen in Sec. 15-6 that an F center in an alkali halide crystal
is considered an electron trapped at a negative ion vacancy, the electron
being shared by the six surrounding positive ions. Such electrons may be
expected to exhibit the electron spin resonance phenomenon and this is
indeed the case. The absorption lines are quite broad. For example, in
KCI colored additively with excess potassium, one observes a resonance
line of 54 gausses wide (one usually employs a fixed frequency of the
transverse a-c field and sweeps the constant part He slowly through
resonance) and a g factor 0(1.995. 36 Now, when one calculates the width
of the line on the basis of dipolar interaction between randomly distributed
F centers, the width would be only 0.1 gauss. However, attempts to ascribe
the observed line width to the interaction between the F center electron
and the surrounding nuclei K39 and K41 have been successful. For example,
from the fact that K39 has a nuclear magnetic moment of 0.391O,un and
K41 of -0.2145, one would expect on the basis of this notion that replacing
K39 by K41 should produce a narrower line. One finds for K41CI (containing
3. E. Zavoisky, J. Phys. U.S.S.R., 9, 211, 245, 447 (1945).
518
[Chap. 20
99.2 per cent K41) irradiated with X-rays, a line width of 36 gausses. If
only the immediately neighboring K ions were the source of interaction,
the line width would have been 31 gausses; presumably the interaction
with the next shell of chlorine ions also contributes to some extent to
the line width.
We also mentioned in Sec. 15-6 that from the observed g factor and
line width one has concluded that the F center electron is not accurately
described by a pure s-state wave function; the wave function also contains
components with non vanishing orbital momentum.
Electron spin resonance lines have also been observed in n- and p-type
silicon. 37 The lines exhibit a hyperfine structure resulting from the interaction between the electron spin and the nuclear spin of the atom to
which it belongs. In general, for a nuclear spin I, one obtains (21 + 1)
lines; the number of lines observed is in agreement with this rule. At high
donor concentrations the lines become narrow and the splitting disappears;
this is a result of the ionization of the donor levels, the remaining line
being attributed to the conduction electrons.
20-12. Ferromagnetic resonance and relaxation
In principle, ferromagnetic resonance experiments are very similar to
nuclear and electron spin resonance experiments. A specimen of the
material, usually in the form of the thin disk, is placed in a microwave
cavity so that the specimen is acted upon by an oscillating magnetic field
of angular frequency wand small amplitude H Q At the same time, a
relatively strong d-c field He is applied parallel to the disk, so that the
magnetization is saturated. The magnetization vector M .. may be considered as precessing about He' and for a fixed frequency w, He may be
varied such that the precession frequency equals w; energy is then
absorbed from the microwave field.
Ferromagnetic resonance was first observed by Griffiths.3s At first sight
one is tempted to interpret the results on the basis of the resonance
condition (20-46) for paramagnetic resonance. However, one then obtains
values for g which are much larger than the free electron value g ,__ 2.
It was shown by KitteP9 that for a sample in the form of a disk with He
parallel to the disk, the resonance condition is given by
(20-47)
(8= magnetic induction). When this formula is used, the g values obtained
31 See, for example, A. M. Portis, A. F. Kip, C. Kittel, and W. H. Brattain, Phy.l.
Rev., 90, 988 (1953).
.
3' J. H. E. Griffiths, Nature, 158,670 (1946).
3. C. Kittel, Phys. Rev., 71, 270 (1947): 73,155 (1948).
Sec.20-121
519
are close to the free electron value. As an example, we give in Fig. 20- 10
the ferromagnetic absorption line for nicKel ferrite, measured at 24,000
megacycles/sec.
Jn gene~l, the theory of ferromagnetic resonance is in good agreement with experiments; however, the
explanation for the very large line widths
(,.._,100 gausses) which are observed is
still in doubt. 40 Since the line width
is determined by relaxation effects
(compare 20-41 ),' this difficulty has
stimulated studies of ferromagnetic
relaxation; references on this topic
may be found in E. Abrahams,
Advances in Electronics and Electron
PhysiCS, 7, 47 (1955).
For antiferromagnetic solids, the
7.6
6.8
7.2
resonance frequencies lie just beyond
...
H
(kilogauss)
the limit of the experimentally accessible
region. The reason for this may be Fig. 20-10. The ferromagnetic resofound in Kittel's theory of antiferro- nance line in Ni-ferrite at 24,000
magnetic resonance. 41 On the basis of megacycles/sec. {After Yager, Galt,
a two sublattice model, he finds that Merrit, and Wood, Phys. Rev., 80,
744 (1950)
below the Curie point the resonance
frequency is a doublet determined by
(20-48)
where (' = g(e/2mc), H is the applied field, He! is the anisotropy field for
one sublattice, and Hrnf is the molecular field. Thus for MnF 2 for which
H A ':::: 9000 gausses and Hmf ~ 10 6 gausses, one obtains Wo ':::: 10 cm-l.42
The anti ferromagnetic doublet has, however, been observed for
CuCI 2 '2H 20, which has an antiferromagnetic Curie point of 4'3K and
thus a relatively weak molecular field. 43
20-13. Frequency-dependence of the initial permeability in ferrites
Because of the great interest in ferromagnetic insulators, such as the
ferrites, for high-frequency applications, extensive investigations are being
made of the high-frequency behavior of these materials. In particular, one
'" C. Kittel, J. phys. rud., 12, 291 (1951); J. H. Van Vleck, Physica. 17, 234 (1951).
<I C. Kittel, Phys. Rev., 82, 565 (1951) .
.. F. Keffer, Phys. Rev., 87, 608 (1952) .
3 Ubbink, Poulis, Gerritsen, and Gorter, Physica, 18, 361 (1952); for the theory.
see J. Ubbink, Physica. 19.9 (1953) .
.
"
520
[Chap. 20
24
20
,,'-I
16
12
8
4
0
-2
1 2
5 10
100
1000
10,000
!mc/sec~..
Sec. 20-13]
521
an internal field Hi. 45 Thus the electron_spins precess about Hi with the
Larmor frequency within each crystallite of the polycrystalline material.
When an alternating magnetic field of the Larmor frequency is applied,
ferromagri~tic resonance occurs when this field has a component perpendicular to Hi' This has been called "natural" resonance, in contrast with
the "induced" resonance obtained with an applied static magnetic field. 46
More recently, Rado et al. carried out some interesting experiments
of the same type on "ferramic A," which is a sintered mixture of several
oxides, but containing mainly magnesium ferrite.47 The curves obtained
for (/1/ - I) and fl" are given in Fig. 20-11. It is observed that in this case
two resonances occur, one at about 50 megacycles/sec and another in the
vicinity of 1000 megacycles/sec. When they carried out the same experiment with small particles ('"'-'0.5 fl) embedded in wax, they observed that
the 50-megacycle resonance was absent. Also, they had shown previously
that particles of this size behave essentially as single-domain particles.
They conclude from these results that the 50-megacycle resonance is
associated with domain-wall displacements, and that the 100-megacycle
resonance is due to domain rotations. The theory of these phenomena is
still in a state of flux and will not be discussed here. 48
,
REFERENCES
E. Abrahams, "Relaxation Processes in Ferromagnetism," Admnces in
Electronics and Electron PhYSiCS, 7, 47 (1955).
A. H. Cooke, "Paramagnetic Relaxation Effects," Rep/s. Progr. Phys.,
13, 276 (1950).
K. K. Darrow, "Magnetic Resonance," Bell System Tech. J., 32, 74,
384 (1953).
C. J. Gorter, Paramagnetic Relaxation, Elsevier, New York, 1947.
522
[Chap. 20
PROBLEMS
20-1. Consider a series arrangement of a self-inductance L and a
resistance R; show that the conductance G of the system as function of
frequency is given by
I.
G(w) =
_ i G(O)wT
G(O)
2
W T2
/1
2
W T2
with T = L/R. Note that this expression is similar to that for X according
to the Debye equations (20-9).
20-2. Show that at low frequencies the paramagnetic absorption is
one order of magnitude more sensitive than the dispersion.
20-3. As mentioned in Sec. 20-1, X' and X" are not independent of each
other; in fact if one of them is known for all angular frequencies w, the
other may be calculated from one of the Kramers relations:
.' w __ 2 (C/O wX"(w) dw
X ( 0) - -)0
(2
i)
7T
W
Wo
and
"
X (w o) =
2 ~oo woX'(w) dw
1T' 0
(2
W
2)
Wo
(a) Show that the Debye equations (20-9) satisfy these relations; do
the same for the Casimir-Dupre equations (20-15) and (20-16).
(b) Give a proof of the Kramers relations by following these hints:
Apply a magnetic field in the form of a delta function
H(t)
bet)
= .!_
I'
00
7T .0
cos wt dw
Chap. 20]
523
.!_
ex; (x' cos wI
7T ()
M(t) =
Now, for t < 0 we must have M(t) = 0; also cos wI is an even and sin w(
is an odd function of I. Hence, we must require for I > 0 that
.10
f{l)
~(t) =
He
+ Hoe
iw
and
PUP
= (P"p)Ho=O
(O:ap)
,uH
Hoe i ,"'
l'
where Ho ~ He' Set up the equations for (oNplot) and (?N,Jot) appropriate to the field H(t). Calculate the magnetization Mo corresponding
to the a-c field and show that
X = MolHo
Xstatic(l
+ iWt)-l
with
T = liP
APPENDIX
A. Thermodynamic conditions for equilibrium
+ p dV
(A-I)
(A-2)
(A-4)
Here F is called the Helmholtz free energy or, as in this book, simply the
free energy. Note that in this case F must be a minimum when equilibrium
has been reached.
(c) Systems held at constant pressure and temperature. In the physics
of solids this is the most frequently occurring case. It follows from
(A-I) that
dG = deE p V - TS) ~ 0 for constant p, T
(A-5)
525
APPENDIX
526
\,
(B-1)
where E is the total energy of the particle, i.e., in our case the kinetic
energy. The general solution is
'P(x)
Aeil.:x
+ Be- ib
with
k2 = (2m/1i2)E
for
x = 0
and for
x = L
The first condition yields A = -B; this leaves only solutions of the type
sin kx. Applying the second boundary condition, one singles out only
those solutions for which
l'
sin kL
or
kn
The solutions
(B-2)
IS
Note that lik n represents the momentum of the particle. The energy
spectrum evidently consists of discrete levels, the separation depending
on L2 and n 2 The constant C in (B-2) may be obtained from the requirement that for a particle known to be in the state 'P", the probability to
be found anywhere between x = 0 and x = L must be equal to unity, i.e.,
(B-4)
..
-~
.r' .
527
APPENDIX
(B-5)
+ n2 + nz2)
(B-6)
ll
p2L2j1i 27T2 = n;
+ n; + 11; _
R2
(B-7)
where the factor t arises from the fact that the integers are positive. For
each set of integers n x , ny, n z there is one wave function, i.e., one state;
the spin is not included in this case. Note that (B-7) may be interpreted
as follows; divide the momentum space into cells of h3 ; each cell then
corresponds to a possible wave function per unit volume.
C. Indistinguishable particles and the Pauli principle
, j
02
+ ox~ +
2m)
Ii'!. E Vl(X I ,X 2 ) ~= 0
(C-2)
and
Vlb(X 1)Vla(X 2 )
(C-3)
APPENDIX
528
The former describes the situation in which particle I is in state 'Pa and
particle 2 is in state 'Pb; the latter corresponds to particle 1 in 'Pb and
particle 2 in 'Pa' Note that both solutions correspond to the same energy
E = "
Eb From the mathematical point of view any linear combination of the solutions (C-3) is a satisfactory solution of (C-2). From the
point of view of physics, however, there are only two acceptable linear
combinations, viz.,
and
\.-
+ 'Pb(Xl )'1pix 2)
'1Ps y m =
'Pixl )'Pb(X 2)
'Panti =
---
(C-4)
(C-5)
(C-6)
1/
either
'P1,2 = "1'2,1
or
"1'1,2
-"1'2,1
\:~ (C-7)
(C-8)
Note that (C-4) and (C-5) satisfy, respectively, (C-7) and (C-8); it can be
shown that (C-4) and (C-5) are the only solutions with these properties.
In nature there are two types of particles: those for which the twoparticle wave function is always symmetric and those for which the twoparticle wave function is always antisymmetric. To which group a
particular type of particles belongs must be decided from experiment.
Electrons, protons, and neutrons require antisymmetric wave functions.
Particles described by antisymmetric wave functions have the following
fundamental peculiarity: from (C-5) it follows that if 'Pil =--= 'Ph' i.e., if both
APPENDIX
529
particles are in the same state, 1fJ,,"ti vanishes, i.e., such a situation does
not exist. By extending the above treatment to many particles, one
arrives at the following conclusion:
In a sstem of particles described by antisymmetric wave functions,
such as electrons, only one particle can occupy anyone "state." This is
the Pauli exclusion principle.
The word "state" must be amended here in the following sense; the
complete wave function of an electron does not contain only the spatial
coordinates x,y,z but also the spin, which can accept two possible values.
Thus if the spin is included in the wave functions 1fJa and 1fJb' the wording
of the conclusion is correct. If a state is considered to be described by its
spacial coordinates only, the Pauli principle should read that no more
than two electrons can occupy a given state. Particles obeying the Pauli
exclusion principle give rise to Fermi-Dirac statistics. Particles described
by symmetric wave functions give rise to Bose-Einstein statistics, and for
them no limitation exists on the number of particles occupying a given state.
D. Fer.mi statistics
Consider a system of particles for which the possible wave functions
(states) and energy levels are known. Let the energy levels be denoted by
1' 2, ... , i and let the number of possible states (including the spin)
corresponding to these energy levels be denoted by Zl' Z2' ... ,Zi' ....
The interaction between the particles is assumed to be weak, so that the
total energy of the system E is equal to the sum of the separate energies
of the particles. The fundamental problem of statistical mechanics is this:
given the total number of particles N and the total energy E, what is the
most probable population n1 , n 2 , . , ni , ... of the energy levels? Evidently,
1: n,
1
and
~ nif. i
,
(D-1)
Also, the levels of interest for the problem are only those below the value
E; this limits the total number of possible states involved to
ZtOI"1
Zl
+ Z2 + ... + Z; + ... + Z R
(0-2)
(0-3)
530
""'1
APPENDIX
What is the probability P(n1' n 2 , ... ) that the populations in the energy
levels are nl , n 2 , ... ? Consider Zi boxes and n i indistinguishable balls and
assume that each box can con tam either one or no ball (Zi ~ n i). The
probability for a specific distribution (say box 1 empty, box 2 filled, box 3
filled etc.) is evidently p7l,. However, there are in general many ways Wi
in which the balls can be distributed, viz., just as many ways as there are
possible arrangements of n i occupied and (Zi - n;) empty boxes. Hence
Wi =
Zil![nil(Zi - ni )!]
p"'W;
p"'Zil![n;l(Z; - n;)l]
p N W 1W 2
... ) =
(D-4)
pNW
Note that r" is a constant and that the most probable state of affairs is
determined by the maximum value of W. One thus has to find that set
of values nl' n 2 , n 3 , .. , for which W obtains its maximum value. It is more
convenient, however, to maximize log W. By applying Stirling's theorem,
assuming all quantities involved to be ;?> 1, one may write
log W = b log W; =
1.
1.
b log W
l: [-log n i
+ log (Z; -
n i)] On;
= 0
(D-6)
However, the variations on i are not independent of each other, but should
satisfy the following auxiliary conditions derived from (D-I):
oN = l: on i = 0
oE = L
and
o;
On;
(D-7)
olog W -
(X
L On"1.
f./ L. "
o.
P
On1 = 0
(D-8)
(X -
(3ok = 0
(X -
{Joj = 0
531
APPENDIX
This is always possible, because we have two equations from which fJ and
IX may be found. Now the variations r5"'i are independent except for two
of them (because there are two auxiliary conditions). If we consider bn k
and bn j as the dependent ones, it is evident that (0-8) can be satisfied
only if for all values of i,
log [(Z; - ni)/n,] Hence
;'
ni
= Z;/(e HfJ;
Q( -
fJfi = 0
I) =~ ZiF(fi)
(D-9)
(D-IO)
bE= oQ - p oV
(E-1)
where p 0 V is the work done by the system. On the other hand, if the
total energy is E, we may write
oE = :E
"
fj
i'Jn i
+ :E, n
i r5fi
(E-2)
It must be emphasized that any changes Ofi in the energy levels are
possible only if the volume changes; this follows from the discussion in
appendix B. Hence the last term in (E-2) may be written
1: n, bE
i"
: OE.
1: n _' b V = - P i'J V
; ' oV
(E-3)
oQ = ~,
Ei
On;
(E-4)
532
APPENDIX
.~
.
b log W =
7 an
(log W) bn i
1
Hence, {3 bQ must also be a complete differential. We know from thermodynamics that bQ itself is not a complete differential, but that liT is an
integrating factor. Therefore {3 = IjkT, where k is a constant, and instead
of (E-5) we may write
k b log W
oS or
k log W
+ const.
(E-6)
This is the famous Boltzmann relation between the entropy S and log W.
The value of k must be obtained by comparison with experiment, and
turns out to be Boltzmann's constant k = 1.38 X 10- 16 erg degree-I.
The above Wand entropy are associated with the distribution of
energy and in this volume are written W(h and Sth' The subscripts stand
for "thermal" and distinguish them from Wrf and Srf, which refer to configurational or mixing entropy which results from possible arrangements
of particles in space.
I ..
'I
INDEX
a-band, 377
Absorption, infrared, 56
theory, 154 f{.
Acceptor levels, 310, 323
Acoustical branch, 55
Activation energy, diffusion, 70, 152, 172
Activators, luminescence, 399, 409
Adiabatic demagnetization, 460
Adsorption, 228
Alkali halides, bandstructure, 369
dielectric constant, 145
diffusion, 168, 171
electron mobility, 393
F centers, 377 ff.
index of refraction, 145, 158
infrared absorption, 147
ionic conductivity, 175 ff.
ionic radii, 126
lattice energy, 122
molecules, J31
optical absorption, 366 ff.
photoconductivity, 386 ff.
photoelectric effect, 390 ff.
thallium activated, 406 ff.
transport numbers, 181
vacancies, 160", 167, 173
Alkali metals, band structure, 267
overlapping bands, 265, 268
photoemission, 234
Alkaline earth oxides, diffusion, 169
lattice energy, 122
Allotropy, 62
Alloys, band theory, 108
electrical resistivity, 288
Hume Rothery phases, 107
interstitial, 104
Jones' theory, 272
neutron diffraction, 21
ordered, 109 ff.
phase diagrams, 116
substitutional, ]04
Amorphous, 2
Anisotropy, 27, 31, 82
magnetization, 478
533
534
INDEX
Brass, 109
Bravais lattices, 4, 6, 7
,
Brillouin function, magnetism, 455, 467
Brillouin zones, 48, 246, 255, 263
Burgers vector, 85
CARRIER injection, 341, 415
Casimire-DuPre' theory, spin-lattice, 501
Cathodoluminescence, 398, 417
Cavity field, 159, 196
Center of symmetry, 185
Cesium chloride structure, 123, 127
Charge compensation, principle of, 410 ff.
Classification of solids, 23, 25, 27
Clausius-Mosotti formula, 144, 197
Closure domains, magnetism, 477
Coactivators, luminescence, 411
Coercive force, 185, 476, 481
Cohesive energy, metals, 269 ff.
ionic crystals, 117
Collisions, electron-phonon, 289
phonon-phonOn, 296
Colloids, F centers, 392
Color centers, 377 ff.
magnetic resonance, 51 7
models, 394
X-rays, 394
Compound unit cell,S
Compressibility, 31, 33, 120
Conduction band, 251
Conductivity, see Electrical or Thermal
conductivity;
ionic, 160 ff.
Configurational, coordinate, 408
entropy, 63, 106, 161
Constant energy surfaces, 262, 334
Contact potential, 230
Conwell-Weisskopf scatteri ng formula,
330
Coordination number, 60
Covalent bands, 320
Creep, metals, 83
Crystals, directions, 9
growth, 31, 210, 324
planes, 9
Cubic, body-centered, 5, 7, 30, 60
face-centered, 5, 7, 30, 60
simple, 7, 8
system, 8
Cubic lattice, band theory, 260
Curie constant, 186, 193
Curie law, paramagnetic, 455
I
j
INDEX
Dislocations (cont.):
recombination of, 90
screw, 90
stress fieldSl 91 ff.
Dispersion, 155
Dissociation energy, 24, 123, 124
Domains, ferroelectric, 184, 207
ferromagnetic, 475 ff.
Donor levels, 310, 322
Double hysteresis loop, ferroelectrics, 206
Drift velocity, 276
Dushman-Richardson equation, 221
535
Electrons (cont.):
mean free path, metals, 283
mobilities, 333
range in solids, 428, 445
relaxation time, 276, 284, 328
transverse mass, 336
Electron scattering, ions, 330
neutral impurities, 331
phonons, 289, 329
Electrostriction, 186
Energy bands, allowed and forbidden,
239 ff., 245
overlapping of, 265
silicon, 336
sodium, 261
Energy levels, metals, 213
Entropy, 525, 532
configurational, 63, 106, 161
thermal, 63, 161
Etch pits, dislocations, 98
Exchange integral, 473
Expansion coefficient, 33, 197
Extinction coefficient, 156
FACE-centered cubic lattice, 5, 7, 30, t>l
F centers, 377 ff.
coagulation, 392
photoconductivity, 388
F' centers, 383 ff.
Fermi-Dirac statistics, 214, 529
Fermi energy, 214, 215, 224, 231
Fermi level, insulator, 306, 309
Fermi temperature, 219
Ferrimagnetism, 490 ff.
NceJ's theory, 493 ff.
Ferrites, 491
initial permeability, 519
Ferroelectric domains, 207
F crroelectrics, ]84 ff.
thermodynamic theory, 20]
Ferromagnetism, anisotropy energy, 478
Curie temperature, 464, 470
Curie-Weiss law, 464, 471
domains, 464, 475 ff.
g-factor, 469
Heisenberg theory, 472 ff.
paramagnetic Curie point, 471
resonance, 518
Ferroxcube, 491
Fick's law, 70, 169
Field emission, 227
First-order transition, 186, 204, 206
Floquet's theorem, 240
536
INDEX
Fluorescence, 398
";
Flux density, 133
Forbidden transitions, 398
Franck-Condon principle, 367, 401
Frank-Read source, dislocations, 99
Free electrons, 211
diamagnetism, 219, 462
effective number of, 250
energy distribution, 213
paramagnetism, 217
specific heat, 216
Frenkel defects, 67, 163
Fundamenta! ,absorption, ionic crystals,
373
GALVANOMAGNETIC effects, 304
Germanium, constant energy surfaces, 334
crystal growth, 324
electrical properties, 331 ff.
infrared absorption, 339
lattice properties, 320 ff.
lifetime of carriers, 343
physical constants, 311
secondary emission, 421
Gibbs' free energy, 526
Glow curve, luminescence, 405, 407
'Goldschmidt radii, ions, 126
Gorter-Kronig theory, paramagnetic relaxation, 523
GOllY balance, 462
Grain boundaries, 1,98
Grey tin, 320
Gruneisen constant, 33
Gudden-Pohl effect, luminescence, 413
Gyromagnetic ratio, 450
Barnett method, 469
Einstein-de Haas method, 469
HALL effect, alkali halides, 393
metals, 301
semiconductors, 326 ff.
Hall mobility, 328, 333
Hamiltonian, electrons, 459
Harmonic oscillator, 34 ff., 51, 101
Helmholtz free energy, 525
Hexagonal close-packed structure, 61
Hexagonal system, 8
Hole, 252
density, 309
effective mass, 337
lifetime, 343
mobility, 333
,,.
Homopolar bonds, 26
Hume-Rothery alloy phases, 107
Hund's rules, 450, 470
Hydration energy, 164
Hydrogen bonds, 187, 189
Hysteresis, 184,465,481
double loop, 206
IMAGE force, dislocation, 95
electrons, 225
ions, 228
Impurity semiconductors, 310 ff., 319
Index of refraction, alkali halides, 145
complex, 156
metals, 237
Indistinguishability of particles, 527
Infrared absorption, ionic crystals, 56,
147
germanium, silicon, 339
Initial permeability, 519
Injection of carriers, 341, 415
Insulators, band scheme, 251
dielectric properties, 133
electron distribution, 305 ff.
secondary emission, 440
thermal conductivity, 295 ff.
Interaction, dislocations, 93
exchange, 472
super exchange, 488
Interatomic forces, 23 ff., 119, 121, 128
Intermetallic compounds, 344 ff.
Internal field, 141, 194, 195, 199
Interstitial atoms, 63, 67
Intrinsic semiconductors, 308 ff., 319
Inverse spinel, 491
Inversion center, 6
Ionic conductivity, 160, 175
Ionic crystals, 25 ff.
defects, 160
infrared absorption, 56, 147
lattice energy, 117, 122
Ionic polarizability, 134, 137
Ionic radii, 124, 126, 127
Ionization energy, 123, 128,229
Ion mobility, 180, 183
Isotropic, 27
JUNCTION, rectifier, 357 ff.
transistor, 361
K-band, color centers, 381
Killers, luminescence, 399, 417
INDEX
Kirkendall effect, 76
Knight shift, nuclear resonance, 516
Kronig-Kramers relations, 499, 522
Kronig-Penney. model, 243 ff.
LAGRANGE, undetermined multipliers,
530
Landau diamagnetism, free electrons, 219
Lande formula, 450
Langevin function, 139, 193
Larmor precession, 452
Lattice defects, ionic crystals, 160, 166,
375
metals, 62, 85
Lattice, direct, 253
reciprocal, 254
Lattice energy, ionic crystals, 117, 122,
130
Lattice vibrations, 32 ff.; see also Vibrational modes
Debye model, 41
Einstein model, 36, 65, 161
LCAO method, 257
Lenz's law, 451
Lifetime, carriers, 341
Liquid crystals, 2
Long-distance order, 3, III
Longitudinal mass, 336
Lorentz field, 142, 199
Lorentz force equations. 280, 302
Lorenz number, 300
Loss factor, 149
Luminescence, 398 ff.
carrier injection, 415
coactivators, 411
decay mechanisms, 402 ff.
F centers, 382
glow curves, 405
killers, 399, 417
quenching, 409
sulfide phosphors, 410
thallium-activated KCI, 406 ff.
MADELUNG constant, 118, 131
Magnetic, induction, 447
permanent dipoles, 448
permeability, 447
susceptibility, 446
Magnetism, '!ee Dia-, Para-, Ferro-,
Anti/erro- and Ferrimagnetism
Magnetite, 491)
Magnetization, 446
11.
537
Magnetization (cont.):
easy and hard directions, 476, 478
Magnetization curve, ferromagnetic, 465
Magnetocaloric effect, 463
Magnetoresistance, 304
Matthiessen's rule, 275, 287
Metals, atomic radii, 60
band structure, 251
cohesive energy. 26, 269 ff.
electrical conductivity, 275 ff.
electron mean free path, 284
free electron theory, 211 ff.
Hall effect, 301
mutual solubility, 105
photoelectric emission, 232
rectifying properties, 348
secondary emission, 422, 430 ff.
structure, 60
thermal conductivity, 299 ff.
thermionic emission, 220
Metastable levels, 403, 408
Miller indices, 8
Minority carriers, 341
Mobility, electrons, 307, 329
ionic, 180. 183
Molecular field, ferromagnetism, 466 ff.
Molecular rotation, 514
Monoclinic system, 8
Mosaic structure, 97
Mott-Schottky theory of rectification, 351
NEARLY free electron approximation,
260,262
Neel temperature, 483
Neutron diffraction, 21, J87, J 89; 20 J, 488
Nondegenerate level, 257
Nonstochiometric composition, 26, 377 ff.
n-type conductivity, 311
Nuclear magnetic moments, 516
Nuclear magnetic resonance, 505 ff.
Nuclear magnet on, 451
OHM'S law, 177,275
One-electron approximation, 238
Optical absorption, 154
alkali halides, 366 ff.
thallium-activated KCI, 406
Optical branch, 55
Optical transitions, band scheme, 337
Order-disorder transitions, 109 ff.
Bragg-Williams theory, III
Ordered alloys, electrical conductivity,
288
538
INDEX
J
INDEX
Schottky, (COlli,):
defects, 65, 163
effect, thermionic emission, 226
Scintillation c~nter. 417
Screened potential, 435
I,:
Screw dislocation, 90
Secondary electrons, 418
Secondary emission, 418 ff.
. ",
escape mechanism, 438
, temperature effect, 440
theory, 423 ff.
universal yield curves, 425
Second-order transition, 186, 202, 206
Seebeck effect, 304
Segregation coefficient, 324
Seignette salt, 187
Semiconductors, electron distribution,
305 ff,
electronic degeneracy, 316
energy gap, 337 ff,
Hall effect, 326 ff.
impurity, 3 to ff.
intrinsic, 251, 308
nonpolar, 319 ff,
II-type, 311
p-type; 311
surface states, 354 ff.
thermionic emission, 314 [[,
work function, 315
Shear modulus, 79
Shear stress, 79
critical resolved, 82
Short-distance order, 3, 114
Silicon, constant energy surfaces. 337
crystal growth, 324
electrical properties, 331 ff.
energy bands, 336
infrared absorption, 339
lattice properties, 320 ff,
physical constants, 321
secondary emission, 422
Silver halides, 129
Simple cubic lattice, 8
Slip, 82, 83
Smakula's formula, color centers, 378
Sommerfeld, model of metals, 212
theory of conductivity. 281
Space lattices, 6, 7
Specific heat, 32 ff.
Born cut-off, 45
Debye theory, 41, 57
.f
Dulong and Petit, 34
I
"
539
540
INDEX
123,