Vdoc - Pub Crystal Defects and Crystalline Interfaces
Vdoc - Pub Crystal Defects and Crystalline Interfaces
W. Bollmann
Crystal Defects
and Crystalline Interfaces
Illustration on the dust cover: Dislocations and grain boundaries in stainless steel, taken with the
1500 kV electron microscope at Toulouse, France, by G. Dupouy and F. Perrier.
Tbis work is subiect to copyright. AII rights are reserved, whether the whole or part of tbe material is con-
cemed, specifically those of translation, reprinting, re~use of illustrations, broadcasting, reproduction by
photocopying machine or similar means, and storage in data banks.
Under § 54 of the German Copyright Law where copies are roade for otber tban private use, a fee is payable
to the publisher, the amount of the fee to be determined by agreement with the publisher.
<1:> by Springer-Verlag Berlin Heidelberg 1970. Library of Congress Catalog Card Number 77-124069.
Originally published by Springer-Verlag Berlin Heidelberg in 1970
Softcover reprint of the hardcover 1st edition 1970
The use of general descriptive names, trade names, trade marks, etc. in this publications, even if the fonner
are not especially identified, is not be taken as a sign that such names, as understood by the Trade Marks
and Merchandise Marks Act, may accordingly be used freely by anyone. Printed by Universitătsdruckerei
H. Stiirtz AG, Wiirzburg. Title No. 1700.
To my Mother and my Wife
Preface
It is nonnal for the preface to explain the motivation behind the
writing of the book. Since many good books dealing with the general
theory of crystal defects already exist, a new book has to be especially
justified, and here its main justification lies in its treatment of crystal-
line interfaces. About 1961, the work of the author, essentially based on
the fundamental work of Professor F. C. Frank, started to branch away
from the main flow of thought in this field and eventually led to a
general geometrical theory which is presented as a whole for the first
time in this book. Although nearly all that is presented has already
been published in different journals and symposia, it might be difficult
for the reader to follow that literature, as a new terminology and new
methods of analysis had to be developed.
Special emphasis is given to discussion and many diagrams are
included in order that a clear view of the basic concepts be obtained.
Intennediate summaries try to bring out the main points of the chapters.
Instead of specific exercises, general suggestions for them are given. The
part up to chapter 9 is considered more or less as introductory, so that
the book can be studied without specific knowledge of crystals and
crystal defects. The presentation of that part developed out of lectures
given by the author at the Swiss Federal Institute of Technology (ETH)
in Zurich. As metallurgists and mineralogists are often not familiar with
linear algebra, which is the mathematical basis of the interface theory,
the procedures involved in this technique are collected in the Appendix.
Many of the ideas contained in the book originated from observations
made on moire models and the subsequent fonnulation of these observa-
tions in mathematical terms. In order that the reader may participate
in this experience, a set of such models is presented with the book.
Now the work is finished, it is my great pleasure to express my
gratitude first to the Battelle Institute (BI) which by its financial sup-
port made this work possible, and in particular to Dr. B. D. Thomas,
fonner president of the Battelle Memorial Institute, Dr. H. Thiemann,
General Director of the Battelle Laboratories, Geneva and Dr. F. I. Mil-
ford, Director of Physical Sciences of BI, for their moral support of this
work. Much inspiration came from discussions with my friend and fonner
colleague Prof. D. G. Brandon, and I would like to express my special
VIII Preface
The central element of our study is the line defect or the dislocation.
The theory of point defects is essentially independent of the theory of
line defects, and the methods of investigation of these defects also differ.
After a brief discussion of point defects (Section 4.1) the book deals
with the individual dislocation (5-6), the interactions between two dis-
locations (7-9), the interactions between dislocations and point defects
(10) and the fonnation of dislocation networks (11). The second part
of the book treats in detail the general geometrical theory of crystalline
interfaces (12-14) which may be either boundaries between crystals of
the same nature (grain boundaries) or boundaries between crystals of
different types (phase boundaries). The applications show that this
theory, which is based on purely geometrical concepts, already leads to
significant physical infonnation, and we believe that it will make a
sound foundation for a wide field of physical research into intercrystalline
boundaries.
The summaries try to provide a general picture of a group of chapters
and, in this manner, to discuss the most important points. Occasionally
it might be advantageous to read the summary before studying the
detailed text.
As the subject matter for this book has originated from the direct
electron microscope observation of crystal defects, emphasis is placed
more on the static than on the dynamic aspects. The problems of dis-
location dynamics, i.e. the theory of plastic deformation, is mentioned
only briefly (10.2).
2. General Aspects of the Structure of Crystals
2.1. Outline
Solid state matter can be considered theoretically from different
view-points depending on the range of problems to be investigated. For
example, the elastic behaviour is generally discussed in terms of the
continuum aspect. Other properties can be explained on the basis of the
atomic arrangement. Our main interest concerns problems related to
the plastic behaviour of materials and, for dealing with those, we need a
certain knowledge of the atomic structure.
Concerning the atomic order two extreme states are distinguished,
the amorphous state, which is mainly disordered (e.g. glass) and the
essentially ordered crystalline state. In reality a certain local order
prevails even in the amorphous state. The characteristic feature of a
material in the crystalline state-i.e. of a "crystal" -is the periodicity
of its structure. A crystal consists of identical elements (building blocks)
which are joined together additively.
We may distinguish one-, two- and three-dimensional crystals.
Examples of one-dimensional crystals are chain molecules such as nylon
or silicate fibers. These molecules can, of course, assemble to form three-
dimensional arrangements; however, the bonds along the individual
fibers are much stronger than the inter-fiber bonds with the result that,
in solution nylon separates into single molecular chains (Fig. 2.1/1).
An example of a two-dimensional crystal is graphite (see Section3.8).
Within the layers the bonds are covalent and therefore very strong, while
the interlayer bonds are of the weak Van der Waals type. Other materials,
which cleave easily, such as mica, can also be understood as being made
up of stacked layers of two-dimensional crystals (Fig. 2.1/2). This concept
of the two-dimensional crystal is even useful for the understanding of
certain three-dimensional crystals (Section 3.2).
The word "crystal" usually means a three-dimensional crystal
already distinct by its external shape (e.g. the tungstane crystals of
Fig. 2.1/3). The external shape is-as is well known-a consequence of
the internal structure and most substances are crystalline without this
fact being evident from the exterior.
We have to distinguish clearly between the "model" of a crystal, its
abstract image, i.e. the ideal crystal, and the real substance, i.e. the
t*
4 2. General Aspects of the Structure of Crystals
real crystal, with all its imperfections and variations. The ideal crystal
represents only certain restricted aspects of the real material.
Before going into a discussion of certain specific crystal structures we
shall sketch some principles of crystallography. It might be thought that
the understanding of crystal defects would require a comprehensive
knowledge of the perfect crystal. However, in reality only a very limited
acquaintance with it is necessary.
(2.2-3)
(2.2-5)
(2.2-7)
The (vertical) column vectors of the 8-matrix are therefore the new
basic vectors expressed in the coordinate system of the old ones. Thus,
the transformation of the coordinates of an arbitrary vector is given by:
~i= (~1)
~2 =
(1 1 1) ({~')1)
0 1 1 (f)2 =S1{~')k. (2.2-8)
~3 0 0 1 {$')3
For example the vector ;r with the crystal coordinates (f)i= [0,1, OJ
is given in orthogonal coordinates ~i = [1, 1, 0]. Accordingly, the in-
verse matrix 8-1 describes the transition from the orthogonal to the
2.2. Mathematical Description of the Crystal 9
(1-11-10)
crystal coordinates.
8-1 = 0 (2.2-9)
o 0 1
(.;')i = (S-I)~ e (2.2-10)
(1 -1 0) (ee
i.e.
(.;'(.;')2
)1) 1
)
= 0 1 -1 2 • (2.2-11 )
(.;')3 0 0 1 e
3
All the transitions from one coordinate system to another or from one
vector system to another, described by matrices with constant terms as
above are called "linear transformations". We distinguish between
coordinate transformations and vector transformations. The formulation
of the transformation matrix 8 for a general crystal with given lattice
constants is indicated in the Appendix Ai Eq. (A 1-26).
If we need to know the length of a vector or the angle between two
vectors while those are expressed in an arbitrary coordinate system, this
information can be acquired in two ways: The first consists in trans-
forming the vectors into the orthogonal system where the required data
are obtained by means of the usual scalar products:
(2.2-12)
and
(xy)
cos (-9: x , y) = Ixllyl (2.2-13)
(where x and yare expressed in orthogonal coordinates).
The more direct method, however, makes use of the metric tensor G,
which depends on the coordinate system
G = 8 T 8. (2.2-14)
(8 T= transposed matrix of 8. See Appendix A 1.)
In the general case with x' and y' in the arbitrary coordinates, this
leads to
(2.2-15)
and
" (x'TGy')
cos ( -9: x y ) = ~1TYT . (2.2-16)
The metric tensor is always symmetrical. For a cubic crystal coordinate
system with the lattice constant a it becomes:
G= ( 0
a2 0 0)
a2 0 . (2.2-17)
o 0 a2
10 2. General Aspects of the Structure of Crystals
G~G ~ D (2.2-18)
The cubic system is a limiting case of both the tetragonal and the
rhombohedral systems. The rhombohedral system itself can be inter-
preted as a special case of the hexagonal system.
All possible translation lattices are divided into 14 Bravais lattices.
The first group of seven is identical with the seven crystal systems while
the second group is derived from the first by adding face centres, base
centres and body centres.
The internal order is specified by symmetry operations. Those are
transformations which recreate the same point arrangement as the
initial one. Translations by translation vectors are specific symmetry
operations. A special class are the point symmetry operations which leave
at least one point (i.e. the origin) fixed and which are represented by
homogenious linear transformations (see Appendix A 1). Point symmetry
operations are rotations by 360 0 (i-fold rotation axis), 180 0 (2-fold),
120 0 (3-fold), 90 0 (4-fold) and 60 0 (6-fold), inversion at a point, reflection
at plane and all the possible combinations of these operations. In this
way 32 different point groups are distinguished.
12 2. General Aspects of the Structure of Crystals
230 space groups result from the combinations of the point symmetry
operations with translations. Space symmetry elements are screw axes
and glide reflection planes. Those operations are represented by non-
homogenious linear transformations (Appendix Ai).
In the next chapter we shall discuss in greater detail several special
crystal structures.
3. Some Special Crystal Structures
3.1. Outline
In this chapter we shall discuss several common crystal structures,
sQme of which are of prime importance in metallurgy-such as the face-
centred cubic structure (austenitic steels, aluminium, nickel, copper,
silver, gold, etc.), the close packed hexagonal structure (zinc, cobalt and
others) and the body-centred cubic structure (ferritic steels, chromium,
molybdenum, tungsten, etc.). The diamond structure prevails in the
field of semiconductors, while (the hexagonal) graphite is important as
a moderator material in reactor technology and as a precipitate in
cast-iron.
Moreover, we shall consider different types of imperfections related
to the various structures.
o position A
+ {JositionB
~ position [
Fig. 3.2/1. A-, B- and C-positions in the close-packed arrangement of spheres
The position of the second layer shall now be called B. The third
layer can again be placed in two positions, namely C and A. The physical
situation differs, depending on the choice of that position. If we choose
the position C, a sequence ABC is obtained. Relative to an atom in
the B-position (in the second layer) all the direct neighbour atoms form
a centrally symmetrical arrangement (Fig. 3.2/2a). However, if we
chose position A for the third layer the sequence A BA is realized. In
this case the nearest neighbours of a B-atom are arranged III mIrror
c A
as bB
A A
Fig. 3.2/2a and b. Nearest neighbours of an atom in the B-position, a) in the
ABC-sequence, b) in the ABA-sequence
3.2. Close-Packed Structures 15
- C
B
K A
/ / / /
B
K C V V /
K B V / /
A
K A
/ / / l(
K C
/ V V
K B
V V /
K A
/ V V
/
C V" r-....V /
V" V /
K
K B
" V V
/ "" V
b
A
"
ABCABCABCAB
From a vertical cut placed through the D-D-line in Fig. 3.2/1 result
diagrams which are convenient for representing the stacking sequence
such as Fig. 3.2/3 for the face-centred cubic and Fig. 3.2/4 for the close
packed hexagonal structure. The K, Hand L-sequences are indicated
according to Fig. 3.2/ 5.
:=
,
-- A A
H B
-- ~ -- 8 H A
H B
k=:::: ..::;:: - A
<=.:: ' l,I .:::::=- H A
H B
H A
H B
H A
H B
BCABCA B C A B
a b
c
K A
H 8
c
c
A
Fig. 3.2/ 5. K-, H- and L-sequences
3.3. The Face-Centred Cubic Structure 17
[1111 [111]
[ 0 111 1111-+-__
[1/1 01/2]-+--!---'.:-+-III/
Vectors of the <11 0 )-type are even larger translation vectors; they
result from:
(1/2) [11 0] + (1/2) [11 0] = [11 0] (3·3-2)
or
[100] +[010] = [110]. (3·3-3)
The fcc-structure can also be understood as a rhombohedral structure
in which the (cubic) vectors [1/2, 0, 1/2], [1/2,1/2,0] and [0, 1/2, 1/2] are
taken as unit vectors of the rhombohedral system (Fig. 3.3/2). Expressed
III
111
Fig. 3.3/2. Rhombohedral elementary cell of the f.c.c. structure
(fa~::) = (!
-1
!-!) (~:) .
1 1 E3
(3·3-7)
The volume of the rhombohedral unit cell is one quarter of the volume
of the cubic unit cell (det(S) = 1/4). If the value of the cubic lattice
constant is a, the metric tensor of the rhombohedral coordinate system
is given by:
1/2 1/4 1/4)
G=as ( 1/4 1/2 1/4 . (3·3-8)
1/4 1/4 1/2
3.4. The NaCI-Structure 19
cr
a. cr
(1!z)(A+81
C
('Iz)(C+A I
(1J2)(8+CI
b C
Fig. 3.4/1 a and b. Elementary cell of the NaCI structure a) parallel to
individual ion layers, b) projection in the [111]-direction
2*
20 3. Some Special Crystal Structures
([3/4, 1/4, 1/4J respectively, depending on the choice of the unit vectors).
The translation vectors of the diamond structure are therefore those of
the fcc-lattice.
[0.0.01
[lh .l/z.O)
Fig. 3.5/1. Elementary cell of the diamond structure
A
Fig. 3.5/2. The diamond structure represented as the stacking of tetrahedra
22 3. Some Special Crystal Structures
[1/2,1/2,1/21---jf---!---I;.
[0.0.01
lattice (Section 2.2) for which the following translation vectors, expressed
in cubic coordinates, can be chosen (Fig. 3.6/2):
~= [1/2 1/2 1/2]
t2 = [-1/2 1/2 1/2] (3·6-1)
ia=[O 0 1].
Fig. 3.6/2. Monoatomic unit cell of the b.c.c. structure [Eq. (3.6-1)]
3.6. The Body-Centred Cubic Structure 23
1/2 -1/2
8 = ( 1/2 1/2
1/2 1/2 ~) (3·6-2)
which transforms the cubic into the oblique angled unit vectors. The
matrix 8-1
8-1 = (-~o ~ ~)
-1 1
(3·6-3)
transfomls the cubic coordinates of a given vector into its oblique angled
ones (2.2-10). The metric tensor of this new coordinate system is:
The new unit cell contains one atom and therefore its volume is
half that of the cubic unit cell (det (8) = 1/2). Its faces are {11 O} planes.
The shortest translation vectors of the bcc-structure are of the (1 /2) <111)
type and the next longer ones of the <100)-type,
8-1 = ( ~ -~ ~) . (3·6-9)
-1 0 1
24 3. Some Special Crystal Structures
Here two faces of the unit cell are {iii }-planes and one is a {11 0}-plane.
The metric tensor is:
3/4 1/4 -1/4)
G=a 2 ( 1/4 3/4 1/4. (3·6-10)
-1/4 1/4 3/4
o
--- x
Fig. 3.6/3. Monoatomic unit cell of the b.c.c. structure [Eq. (3.6-7)J
(2 0 0)
G=a 2 0 2 0 . (3·6-14)
001
The tetragonal unit cell is obviously twice as large as the bcc cell, since
it contains four atoms.
The above description demonstrates the relationship between the
body-centred and the face-centred cubic lattices. One lattice can, in
principle, originate from the other by pure deformation. This type of
phase change is called the Bain transformation after the American
metallurgist Edgar C. Bain.
which is not the case. Therefore the translation lattice of the cph-
structure is determined by a, b and c.
The metric tensor of the hexagonal structure is
(3.7-1)
t--t-t--t--t-[ll.l. 213.1J21
each other around the c-axis, and therefore are linearly dependant. This
involves the additional condition that the sum of the first three coordinates
must be zero. The unit vector [100] in this representation becomes:
_____-1---jf-::::---t--.----~~-7 1/2
- - {11J.Z13,1121
~~---[0.0.1f21
->-+-~
[0.0.0] [21.!.1/3.01
Fig. 3.8/1. The graphite structure
28 3. Some Special Crystal Structures
A
KB \ r\ \ r\
KC 1\ \ \
K A \ 1\ \
KB \ \ \ 1\
H C 1\ \ \
K B V / /
KA / / / V
K C V V /
KB / / V
A
l;' / / V
ABCABCABCAB
the middle layer of the H-sequence being the twin boundary. This kind
of stacking fault is generally called a growth stacking fault, because it
occurs principally during crystal growth or recrystallisation. A series of
two H-sequences is a so-called deformation stacking fault (Fig. 3.9/2). On
speaking of stacking faults, deformation stacking faults are generally
meant. The configuration of this type of fault can be produced by plastic
deformation, i.e. by shifting one part of the crystal with respect to the
other.
A translation vector for shifting a layer into an equivalent neigh-
bouring position is of the (1/2)(110)-type, and for shifting it into a
stacking fault position it is of the (1/6) (112)-type. Thus, a lattice
3.9. Irregularities in the Stacking Sequence 29
C
K 8 V / /
KA / / / /
KC I V I
K 8
I / V
H A
/ / / V
H8 1\ \ 1\ 1\
K A
I / V /
KC V / V
K 8 / V /
A
/ / V /
A8CA8CA8CA8
Fig. 3.9/2. Deformation stacking fault
where the first step leads from the perfect lattice position into the
stacking fault position and the second step from this stacking fault
position into the new perfect lattice position. The sequence of these
steps cannot be chosen arbitrarily (see Section 8.2).
An important source of stacking faults is the condensation of vacancies
or interstitial atoms. The condensation of vacancies corresponds to
removing one layer followed by closure of the gap. We consider the
sequence
t
... ABCA(B)CABC .. .
... KKKKKKK ... .
If the B-Iayer is removed we obtain
... ABCA/CABC .. .
... KKHHKK .. .
i.e. the configuration of the deformation stacking fault (Fig. 3.9/2).
Is, however, a layer added at the indicated location
t
... ABCABCABC .. .
... KKKKKKK .. .
then it can occupy neither the A- nor the B-position because in these
cases the unstable AA- or BB-sequence would result. Thus, the ad-
30 3. Some Special Crystal Structures
C
K B / / / I
K A V / V ~
K C / V V
K B / I I
11 A I I I /
K B \ \ \ \
11 C \ \ \
K B I I I
K A / I I V
C V V /
ABCABCABCAB
,..
A
\
r..
A
r\ r\ 1\ r\ N \J \J
(V V vr
H B HB
HA I I HA V I I
H B ~ ~ 'J \ HB ~ N \ r\
HA V I I V HA V I / V
1. B
r\ \ \ ~ H B
1\ \ r\ 1\
1. B I I I II
~ If. A
f--
HA
I I V K C
I I V
H B
H A
~
(/
r\
JV
r\
V
r\
V
HA
HC
j - r\
V
r\
V
B
\ ~ ~
~ ~ A r\ r\ r\ )-
ABCABCABCAB ABCABCABCAB
b
A A
H D
r\ ~ ~ \l H D
H A
V V V V HA
H D
~ ~ ~ 1\ !. D
i. A
[7 lJ V V H C
If. C
V V V H B
If. B
) J J H C
H C - ~ )- I- ~ \ H B
J J )
H D ~I-
If C
H C
~ \ \ H B
B
J J~r'--
V C
A DCA DCA DCA D ABCABCABCAB
a. b
A 8
c' A'
C
8'
8~~~~~~~~~~=3~~~==~ 8
A' A'
A A
ABCABCABCABC
Fig. 3.10/1. Difference in the potential energy between the C- and the
A -position (schematic representation)
1
- ti8.
O.
3
fill 2. 8fj3
4
.2.
2.
3
1
4
1
4
·2
3
•
For interstitial atoms, two possibilities exist: the single interstitial atom
can either traverse the crystal on its own, or it can be repeatedly replaced
on its way by other atoms, in analogy with a relay. The first case is
possible, for example, in strongly anisotropic crystals such as graphite,
where an atom is highly mobile between the layers but strongly bonded
within them. In the face-centred cubic structure interstitial atoms, and
especially impurity atoms, can be located in octahedral-(1/2 [111]) -or
tetrahedral-(1/4 [111]) holes, eventually leading to distortion of the
surrounding lattice. On the other hand, there are arrangements such
38 4. Point Defects
c
Fig. 4.1/2a--c. Dumbbell configuration of interstitial atoms a) and b) in the
f.c.c. structure, c) in the b.c.c. structure
where
c= concentration of point defects,
n= number of point defects per N atoms,
EB = formation energy = work necessary to form the defect,
k= Boltzmann constant = 1.37.10-23 Watt· sec. degree-I,
T= absolute temperature.
Au ~ 0.6-0.8 eV .
For the alkali metals, the energies vary from 0.55 eV for Li to 0.26 eV
for Cs.
The formation energies for interstitial atoms are much higher:
3-6 eV. This means that in a state of thermal equilibrium only vacancies
40 4. Point Defects
5.1. Outline
Dislocations are line defects in an otherwise perfect crystal. In this
chapter the basic geometrical concepts shall be explained, such as the
path of the line, the Burgers vector (strength of the dislocation), the
relative orientation of these two, possible movements of dislocations and
their relations to Moire patterns. These latter are a convenient means for
understanding many of the basic dislocation phenomena.
/'1/J?' /'
/~/
/ 1/
,.....-
I
-;7
-r-
/'
/
a b
/ .#
.I
-t.~/
I
/
:!
I.!.
V
/ y Fig. 5.2/1. Generation of an edge
dislocation by a Volterra cut
I
c
/ /
J
L /
/1
1
1
1
1
1
1
L
1
//
~ V
b
a
Fig. 5.2/2. Generation of a screw dislocation by a Volterra cut
5.2. Definition of the Dislocation Line and the Burgers Vector 43
c R
S S ,ll
F G9 F
\ \
I I
a
1l. ' S S
F @ F
b
Fig. 5.2/3a and b. The Burgers loops as an imaging relation between the
distorted crystal C and the perfect reference lattice R. a) The loop is closed
in the distorted crystal, FS/RH-convention. b) The loop is closed in the
reference crystal
Remark
In reality, we are dealing here with three different coordinate systems, i.e.
1. the Euclidean coordinate system of the surroundings,
2. the distorted system of the crystal, determined by the real positions of
the atoms, and
3. the system of the perfect reference crystal.
44 5. The Individual Dislocation: Geometrical Basis
Fig. 5.3/1. Change of a dislocation from the edge- (1) to the screw-orienta-
tion (2)(b is drawa here in the SF/RH convention!)
7'
sgn(l.b)=,
+
right-handed screw
(5·3-3)
" - left-handed screw
Using the FS/RH convention, the right handed screw becomes positive.
This harmonises better than the negative sign with our predominantly
right-handed world.
It might be of interest to determine from a given land b the position
of the additional plane of the edge component in a general dislocation.
Instead of going through the definition of the Burgers vector one may
write
p= [lx bJ. (5·3-4)
5.4. Motion of Dislocations 47
b'J I I I I
1 11 11
I -;
I --\ I ~
I \ ! I "\ ! / I \ / /
Fig. 5.4/1. Motion of a dislocation by glide
--
--
'\. o~ ~
1"'- <""""" <==
0
" <=
,,,0 tl°
=*"
l/ ~
~
/ /
/ ,/
\
I \ I \
\ I I I I
Fig. 5.4/2. Motion of a dislocation by climb
a b
Fig. 5.4/3 a and b. Movement rule. a) Glide, b) climb
A 8 A 8 A
b
Fig. 5.5/1. a) Moire pattern, b) corresponding screw dislocations
interaction between them. There are areas, in which the crystals match
perfectly, that is where the lines cross (A), and areas of particularly bad
matching, i.e. where the two sets of lines are separated by half the line
spacing (B). This situation can be represented by the moire model
No.4. We obtain a series of light and dark bands. The light bands are
areas of good fit and the dark bands of bad fit.
4 Bollmann Crystal Defects
50 5. The Individual Dislocation: Geometrical Basis
is perpendicular and pointing outwards (accordingly, tT2 and tT3 for the
positive y- and z-axis). tT-1 is the stress on the opposite side of the cube.
Each of these stresses is split into three components corresponding to
the three directions of the coordinate axes: e.g. (tTl = O"u, O"u, 0"13)
(Fig. 6.2/1).
z
x
Fig. 6.2/1. Stress components on the surface of an "infinitesimal" cube
From the condition that the cube should be immobile the following
relation holds for the equilibrium of forces:
(i=1···3,k=1···3) (6.2-1)
and, from the equilibrium of the angular momenta, we have
(6.2-2)
This leads to 6 independent components for the stresses, forming a
symmetrical tensor:
(6.2-3)
The diagonal elements, O"ii' if they are positive, stand for tensile stresses,
otherwise for compressive stresses. Components with mixed indices O"ik are
shear stresses. Occasionally, the shear stresses are indicated by T. This
tensor defines all the stresses at one point (except for the case where
internal torques have to be taken into account (see Hirth and Lothe,
1968, p.31). Obviously all six components of the stress tensor may
depend on the location within the crystal. Elasticity theory deals with
all the possible stress functions and the relations between stress and
strain. We shall apply the results obtained by elasticity theory; for
their derivation we refer again to Hirth and Lothe (1968).
6.3. Forces Acting on a Dislocation 53
~--------a--------~
Work = Shear stress (force per unit surface) X surface X distance Ibl
W=i(l. a) . b. (6.3- 1)
This work can also be considered as the work done by a jorce on the
dislocation:
Work = force per unit length of the dislocation
X length of the dislocation
X distance moved by the dislocation
W= j.l.a. (6.3- 2)
From the comparison of these two equations it follows that the force
per unit length of the dislocation is given by:
j=i· b. (6·3-3)
54 6. Interaction between a Dislocation and its Surroundings
This force, I, is thus proportional to the shear stress 't and the absolute
value of the Burgers vector, b.
In general, the differential of the force, df, on a dislocation is given
by the formula of Peach and Koehler * :
(6·3-4)
with
b = Burgers vector,
dl = vector element in the direction of the dislocation line,
fib = stress on the plane normal to the Burgers vector, and on which b
points outwards.
The formula indicates that df is always normal to dl, i.e. the force
always acts normal to the dislocation line. It has no tangential component
in the line. The dislocation line cannot move within itself; forces in the
direction of the line are therefore meaningless. Furthermore, df is also
normal to fib' Fig. 6.3/2 shows an example of a force with a component
in the glide plane. In contrast to the example of Fig. 6.3/1 the vertical
component of the shear stress has to be taken into account. However,
due to the symmetry of the stress tensor both components are equal.
The Peach and Koehler formula also applies to climb but, as shown
by Weertman (1964), the conditions have to be defined more precisely.
Under hydrostatic pressure, P, the dislocations do not climb, even if the
temperature is sufficiently high to provide the necessary mobility of the
vacancies. It is therefore meaningful to replace the general stress, a.,. by
a." + P with P = -1/3 (au +a22 +(33 ) = hydrostatic pressure. The
external stress, defined by the Peach and Koehler formula, may be
balanced by other forces, such as "friction" or "chemical" forces, the
latter being produced, for example, by the change in the equilibrium
concentration of vacancies on movement of the dislocation.
* Privately communicated in this form by Professor J. S. Koehler.
6.4. Energy of a Dislocation 55
Fig. 6.4/1 a and b. Geometrical situation for the calculation of the energy of a
screw dislocation. a) Volume element inside the cylinder, b) evolution of the
volume element
56 6. Interaction between a Dislocation and its Surroundings
U=iTY· (6.4-4)
This is equal to the work done during the elastic deformation of a volume
element. Substitution of the above values for T and y gives:
1 Gb 2
U=Sn 2 r 2 ' (6.4-5)
The volume element is
dV=L·2nr·dr. (6.4-6)
Thus the energy content of this cylindrical volume element becomes:
"
U=U
c
+ Gb4nL
2
J~
r
=TL +
c
Gb 2 L In
4n
(~)
b' (6.4-8)
b
where ~ is the core energy.
Discussion
1. U oc L. This holds, in principle, for long straight parts of the
dislocation. We can therefore write: U =r· L, with r being the energy
per unit length.
2. r contains the term r;" the energy of the dislocation core per
unit length. We obtain a rough estimate of this energy by considering
the dislocation core, with its highly disturbed atomic arrangement, as a
6.4. Energy of a Dislocation 57
row of atoms in the liquid state. In this caSe 1;, equals the number of
atoms per unit length times the heat of fusion per atom. A closer estima-
tion is obtained from:
(6.4-9)
This leads to
Gb (1'
r= ----;;n In f +2 ) .
2
(6.4-10)
a b
Fig. 6.5/1 a and b. Extension of a dislocation loop under the action of a
shear stress T. a) Work done by the force I on the dislocation, b) work of the
line tension T (not T as erroneoulsy indicated in the drawing)
7-
Fig. 6.5/2. Frank-Read source. (The numbering refers to the time sequence)
until finally a closed loop and a separate dislocation between the obstacles
are formed. With the latter the initial configuration is reproduced.
Further dislocation loops can thus be formed by the same mechanism,
causing a glide displacement of b. The closing of the dislocation loop, i.e.
the transition from No.6 to No.7 in Fig. 6.5/2 is a dislocation reaction.
Those will be discussed in Section 7.4.
°
in cylindrical coordinates (1',6, z):
(With the angle starting on the positive x-axis.)
u(r, 0, z) = :n [0 + =i0~~)]
b [ 1 -2v cOS(20) ] (6.6-5)
v(r, 0, z} = - 2:ii2(1 -v) In(r) + 4(1 -v)
w(r, 0, z) =0.
Eq. (6.6-4) shows that for y > 0 the crystal is stretched in the direction
of the x-axis and for y < 0 it is compressed. Furthermore, the displace-
° °
ment by the Volterra cut can be seen in Eq. (6.6-5) from the difference
in u between = 0 and = 2~.
6.7. The Stress Field of a Dislocation 63
with the sence of path of the integral being linked to the line sense of
the dislocation, forming a right-handed screw.
Fig. 6.6/1. Strain field of an edge dislocation calculated from Eqs. (6.6-1)
and (6.6-4) (v = 1/3)
------
x
IZ
I
I
I
I
Z x
y
r
Z x
b
(6.7-1)
6.7. The Stress Field of a Dislocation 65
with
(6.7-2)
H
0
G (6.7-4)
0
0'32
with
Gb
0'23=0'0z= 2nr • (6.7-5)
A shear stress thus acts on the plane normal to the ()-direction (i.e. on
the plane crossing the dislocation line) in the z-direction (or on the z-
plane in the ()-direction, respectively) (Fig. 6.7/2). It is proportional to
y
the shear modulus G, and to the Burgers vector b and decreases pro-
portional to 1/r with increasing distance from the dislocation.
c) Stress Field of the Edge Dislocation (according to Section 6.6)
in Cartesian coordinates (cube edges parallel to x, y, z):
Stress tensor:
(6.7-6)
Y(XB_yB)
To b (Xl+ yl)1
X(XI_yB) (6.7-7)
Tob (XB+yl)1
0'33 =0'.. ='11(0'"" +O'yy)
y(2X2+yl)
= -To b'll (XI+y2)B
TO =G/2:n;(1-'II),
G = shear modulus,
'II = Poisson's ratio Rj 1/3.
Cube parallel to x, y, z, but components in cylindrical coordinates:
0'12=0',6 =+ -r-cos
To b 0 (6.7-9)
TO
=---2'11s1O
b . 0.
0'33 = 0',.
r
with ex being the angle between the Burgers vector and the dislocation
line.
The derivation of the stress- and strain field is given in Hirth and
Lothe (1968), p. 71.
If the first dislocation, lying in the position of the z-axis, has right-
handed screw character with a Burgers vector b 1 and the other parallel
dislocation with b 2 is of mixed character, the force between them is
(Hirth and Lothe, 1968, p. 117):
f -- G(b1 • '1) (b 2 • '2)
2n r ur· (7.2·-3)
'2
"4 and being the unit vectors in the line sense. Eq. (7.2-3) fulfills the
invariance condition (5.2-1): F(l, b) =F( -I, -b).
I*1;l,,--r---:;xI
, yt / !
, I /-
-----~--~---~~-~-
I " " , I b //1 <Pl
I ,/ I '
-----t-~/
1/ I 'I ~
-----~~--+--~~-~Jl~
/ I
/,,
, I', 0
, ", 1
I, I, II
4 I 3 I 2 I
I I I
I I I
I
I, ~
I
I
I
.L
I
II _ _ II
I I I
I I )
I I I
1--,,-1
I
_ -
I T
I
I
I
I
0l
b
~~~ --f
Fig. 7.3/1. Force between two parallel edge dislocations [Eq. {7.3-1}] under
the condition that only glide is permitted. (The circles mark stable positions)
with their appropriate line sense and the Burgers vectors with the corres-
ponding signs, Frank's node condition states:
The sum of the Burgers vectors flowing into the node (line sense I)
equals the sum of the Burgers vectors flowing out of the node.
(d---:;S---B-
remain unchanged as long as no dislocation is cut.
---------------
Fig. 7.4/1. The proof of Frank's node condition. (Displacement of the
Burgers loop)
:2 ~
111
, __
~ __
•
Hexagonal Structure
The hexagonal structure has a preferential plane, the basic plane
(001). The smallest translation vectors are of the type [1 OOJ. A possible
reaction is as follows (Fig. 7.5/2):
[100J + [010J = [110J (7.5-7)
o o
Fig. 7.5/2. Dislocation reactions in the c.p.h. structure. 1. Stable, 2. unstable
The left hand side of the equation represents the Burgers vectors of the
two partials, which together form the total dislocation with the Burgers
vector (1/2) [110].
The b2-balance of this reaction is:
(8.2-2)
From this point of view the splitting of partial dislocations is favoured,
i.e. they repel each other. It can also be seen from Fig. 8.2/1 that the
corresponding Burgers vectors form an obtuse-angled triangle.
A 8 c [1101
[ill] /[itf)
£ 0 + £ 0
/
/
/
>
0 0 £
~4(3)+
/
(ZI r'(ZI / 11121
+A
£ £(31\ ,11111
I ,
(11 V(11 I \
0
£
+
0
C1
+
0
A 0
£
[1211 !
'[liD1
\[2ff]
a
z
• C
+B
OA
b
Fig. 8.2/1 a and b. Partial dislocations in the f.c.c. structure. Full arrows:
displacement of the part above the plane of the paper; dashed arrows:
displacement below the plane of the paper. a) projection in the [111]-
direction. (The arrows in the right-hand diagram indicate the crystallo-
graphic directions.) b) situation in space (G glide plane)
80 8. Partial Dislocations
(8.2-4)
with
1 < or. < (i-v) = correction for the deviation from the screw orientation.
The more accurate formula
r = _G_ [(b . 1) (b . I ) + [b 1X '1] . [b x
2n y 1"'1 2 2 1- v
s 's]] (8.2-5)
is derived in Hirthe and Lothe (1968) p.298. Fig. 8.2/2 shows a split
dislocation. (For the numerical values of the stacking fault energies
see remark in Section 3.10.)
partials are also called Frank partials and the glissile (1/6) 112)- <
partials are termed Shockley partials. Frank partials occur where layers
are inserted or removed (condensation of interstitial atoms or vacancies).
Several possible stacking fault configurations were discussed in
Section 3.9. The boundary lines of such configurations inside the crystal
are various types of partial dislocations.
0 A
+ B
.& C
/
/
+"\
\
\
\
all the possible steps on the (111) glide plane that are allowed for the
upper half of the crystal are indicated by solid lines. With respect to
the movement rule (5.4-1), these are the steps for which m points in
the [111 ]-direction. All these steps together form a hexagonal arrange-
ment (analogous to the graphite structure, Section 4.9). If m points in
the opposite ([111]-) direction, the dotted steps (Figs. 8.2/1 and 8.3/1)
become possible. In the first case the stacking order changes from (8.3-1)
to (8.3-2). However, if the "bottom" part is displaced and the "top"
part kept fixed, then the stacking order after the first step is
CABCACABCA (8·3-4)
KKKHHKKK.
The splitting of dislocations therefore depends on the sign of m.
Four equivalent {iii} glide planes exist in the cubic structure and
we must now determine the stacking sequence of the partial dislocations
in all these glide planes. The four {iii} planes can be joined to a tetra-
hedron, the "Thompson's tetrahedron" (Thompson, 1953). The edges of
this tetrahedron are formed by the Burgers vectors of the type 1/2 <110),
of the perfect dislocations, and the connections from the corners to the
face centres of the triangles are composed of the Burgers vectors of the
partial dislocations ofthe type 1/6 (112). In order to fulfill the conditions
for the splitting of dislocations and also to meet the movement rule,
m has to point from the respective tetrahedral surfaces towards the
interior of the tetrahedron. Otherwise the partial Burgers vectors, taken
in the correct order, would externally surround the triangle of the
tetrahedral surface. Thus, two possible tetrahedra exist for the eight
different <111) orientations of m (Figs. 8.3/2-4). The total Burgers
o o
\
\
\
\
\
\
I
I
/
I
!
o o
Fig. 8.3/2. Composition of Thompson's tetrahedra for the representation of
dislocation splitting on all four {111 }-glide planes. 'In has to point into the
tetrahedra
8.3. The Sequence of Partial Dislocations 83
x
Fig. 8.3/3. Position of Thompson's tetrahedra in the cubic elementary cell
vector b results from the addition of the partial Burgers vectors b(l)
and b(2); the first term of the vector addition, b(l) is assigned to that
partial which points in the direction of v of the.movement rule
[Eq. (5.4-1)]. I and v can always be defined in such a way that m cor-
responds to one of the four <111) directions of Fig. 8.3/4. When, on
splitting the" perfect" Burgers vector b, the partial b(l) points to the
centre of a tetrahedral face, the sequence of partial dislocations can be
determined geometrically in an unambiguous manner by means of the
Thompson tetrahedron.
Here follows a description of an algebraic procedure for deter-
mining the sequence of partials. This procedure can be useful, for ex-
ample, in connection with computer calculations (Bollmann and Perry,
1967).
We define:
the line sense I,
<
the (total) Burgers vector b [of the type (1/2) 11 0)],
the glide plane given by m, of the form <111).
6·
84 8. Partial Dislocations
o o
Fig. 8.3/4. Thompson's tetrahedron net
(8·3-9)
8.3. The Sequence of Partial Dislocations 85
The two factors on the right are the signs of the products of the co-
ordinates of c and m. Thus, with m and c given, x and y have to be
attributed to btl) and b(2) such that Eq. (8.3-9) is fulfilled.
As the vectors m and B are parallel and hence, their coordinates are
proportional, sgn (B. m) can be checked on a single coordinate, e.g.
sgn(B. m) =sgn(Ba ma)
(8·3-10)
Example
We assume a dislocation with
1=[121], b = (1/2) [11 OJ.
The screw component of this dislocation is right-handed [sgn (I· b) = (+),
Eq. (5.3-4)] and the extra plane of the edge component (p= [lxh],
Eq. (5.3-3)) is on the side [111]. This direction remains normal to the glide
plane and we choose m = [111] (Fig. 8.3/5). According to Eq. (8.3-7)
v = [m X lJ = [101]
I
-- I
I £. 1/6[11211
./
)...---- ---+----
/'
// //
x
and 1(1) is the partial dislocation which leads in the direction of v. From
Eq. (8.3-5)
c = (1/6) [112J
and from Eq. (8.3-6)
~=(1/6) [211J
y=(1/6) [121J.
Hence,sgn (f) m;) = (-) andsgn (IJ Ci) = (+) and thussgn(B.m) = (-)
• •
[Eq. (8-3-9)J, i.e. the sense of B is opposed to that of m. This condition
is fulfilled jf:
b(1)=y=(1/6) [121J
b(2)=~=(1/6) [211J.
Case f can be compared with the others, for example by placing the
arrangements f and e (Fig. 8.3/6) side by side and letting the partials
annihilate each other in a stepwise manner.
8.4. Transformation of the FCC-into the CPR Structure 87
1--1
12.--/}
liz [110J 1'-1'
\ /
\ /
\ / 1'- -
\1/6 [121J 1/6[211J / 1JJ.-0
\
\\ /
/
r; - t
/ fl-0
m~@~[mJ\ 1/6md
\ /
\ \ I
/
1~ 1. 1. --1 1 --1
/}~/} 1l.--12. 12-1l.
1'--1' l' --1' 1'-1'
1'~
IJJ.~ 0 IJJ. ~
0
r; ~
I r; =
0 fl- @
~Zl ~Z)
.Il. II
B C
K A
V I)L )/ 1)/ K B
K C ~ ~ I K A
K B k,I~ 1/1IlL-V K C
K A
'r-~V'f- V I K B
V V V
K C K A
K B --K C
K A lL~ V I-rilL K B
I V I
--K C
K B vr If vr H A
H B
ABCABCABCAB BCABCABCABC
a b
A
K C
K B
K A
-K C
K B
H A
H B
H A Fig. 8.4/1 a-c. Geometrical con-
H B KH-+-o-++-o-+-H ditions for the transformation
of the f.c.c. into the c.p.h. struc-
A ture. a), b), c) sequence of dis-
ABCABCABCABCA placements
c
becomes
.j.
ABABCABCA (8.4-2)
HHKKKKK
and
.j.
ABABABCAB (8.4-3)
HHHHKKK
etc.
The question now arises as to how the transformation actually
occurs. Electron microscopic observations show that the transformation
is always incomplete. A survey of the problems can be found in Christian
and Swann (1964).
We discuss here several possibilities for the transformation mechanism:
1. A closed loop of a partial dislocation, i.e. a circular disk of stacking
fault, is formed spontaneously and starts to extend. This leads to a
decrease in the energy of the crystal due to the increase in the stacking
fault surface with negative stacking fault energy. On the other hand,
the energy is increased by the extension of the partial dislocation. As
the energy decreases ex; r2 and increases ex; r, it will rise at the beginning,
reach a maximum value, and will then decrease. The radius of the
stacking fault disk at the energy maximum (~E = 0) is the critical
radius, re. For r>re the loop can extend spontaneously. r< may be
obtained from
(8.4-4)
with
y = absolute value of the stacking fault energy,
r = 1Ine
· . Gb 2
tensIOn ~ - 2 - .
This leads to
Gb 2
r~-
e 2 (8.4-5)
with
G ~ 7.5 .1011 dyn
cm 2 ,
a a2
b=6<112), b2 =C;'
a=3·55A,
erg
y~ 5 cm 2 •
(In order to obtain a finite negative value for y, the material has to be
undercooled before start of the phase transformation. In fact, it usually
starts at around 390°C.)
90 8. Partial Dislocations
a b c
that of the incoming ones. Hence, for example, two partials have to
move out in order to cancel the long range strain resulting from the
junction of the stacking faults. A mechanism of the type given in
Fig. 8.4/4 results from these assumptions, whereby the different steps
8 __ .--~.....~.....~...
A -+-_f'----flJ----j1-\
8 _...-_~-~-~-\
A--I7~~~~~~~~~--'_~
C --.t--tI>----fl>----tl>----tli--\
8_~-,~~-;~~~ __'_~~~~
~
[111J [110J_
[00 1] [111] ~-+--lI
?Il ~-=-~:~=;~~
[11il ~-+--lI
~ ~
Fig. 8.4/4. Phase transition mechanism by "reflection" at stacking faults
8.4 . Transformation of the FCC- into the CPH Structure 93
are thought to take place on a statistical basis. Fig. 8.4/4a shows that
at slow rates the stacking fault (2) would come to rest in an equilibrium
position at a certain distance from the stacking fault (1). Only with the
help of additional kinetic energy of the partial dislocation could new
stacking faults be formed .
The essential points of such a mechanism are:
a) The basic configuration of converging stacking faults is frequent,
since dislocations present in the cubic phase would tend to split. Rem-
nants of former stacking faults can often be observed (Figs. 8.4/ 5,
8.4/6).
Fig. 8.4/ 5. Evidence tha t the final transformation in direction C took place
across stacking faults already present in directions A a nd B
Fig. 8.4/ 7. Hexagonal grain (H) adjacent to a cubic grain (e) which is
stabilized by a framework of stacking faults
<
whereby domains in the four different 111 ) -orientations are produced
in the original cubic grains (Fig. 8.4/8).
5. A comparable mechanism was suggested by Hirth (1969), in
which it is assumed that an initially present stacking fault is cut by a
perfect dislocation, so producing a step in the stacking fault, i.e. a
partials dipole (Fig. 8.4/9). (A dislocation dipole is a pair of parallel
dislocations of opposite sign.) Under the effect of the stress field, new
dipoles would originate from the first and, as a result of the negative
stacking fault energy, they would be forced apart, thus creating new
layers of the hexagonal structure.
If, in the original assumption, a partial dislocation cuts the stacking
fault instead of the perfect dislocation, the dipoles would have to be
formed at a stacking fault. In this case the mechanism would be close
to the previous one, except that the transformation proceeds parallel
to the previously existing stacking fault, while in mechanism No . 4 the
newly formed stacking faults are parallel to the incoming stacking fault.
6. Gedwill et al. (1964) observed that cobalt whiskers (hair-like
single crystals) transform more readily in the longitudinal than in
the transverse direction. The advantage from the energetic point of
view becomes obvious if it is remembered that an oblique cut of a cylinder
96 8. Partial Dislocations
I
+1/6[1121 I
• 0-
I -116[1121
+0
I
•
I
b
j
+ •
(3)
0-
I
I (2)
+0
(1) f
+ • O-
(2)
+0
I
f (1)
.- c
+ .. O-
f (3)
+9
I
Fig. 8.4/9a-c. Dipole mechanism due to Hirth. a) Dislocation impinging on
a stracking fault, b) dipole of partials formed at the limits of the" dislocated"
stacking fault, c) formation of new partial dipoles. (The numbering sequence
reflects the time sequence)
tremity, then the shear stress produced by the line curvature (Eq. (6.5-3)]
and the increase in the extent of the stacking fault act together to
accelerate the transformation. In contrast to case 1, the transformed part
here lies outside the curved partial, and hence no critical radius has to
be surmounted. The shear stress produced by the line curvature and the
transformed surface per loop (i.e. the decrease in energy) are both larger
in the case of the longitudinal transformation than for the transverse
transformation. These considerations also hold when the shape of the
whisker is prismatic instead of cylindrical. Thus, it is to be assumed that
this kind of transformation nucleates at the surface.
9.1. Introduction
In Section 7.4 it was shown that reactions take place between dis-
locations, i.e. they form nodes and coalesce if the energy of the con-
figuration is thus lowered. At the nodes, Frank's node condition for the
Burgers vectors (continuity of Burgers vectors) has to be fulfilled
[Eq. (7.4-1)]:
(9.1-1)
(9.1-2)
The Burgers vectors are now transferred to the b-space and added: since
their sum is zero they must form a closed parallelogram configuration
(Fig. 9.1/1 d). We shall now formulate a series of duality relations which
will allow configurations in the L-space to be "translated" into those
in the b-space and "vice versa".
9.1. Introduction 99
L b
D1 An L-node corresponds to a closed b-polygon.
L b
(2) (1)
001
(2) (1)
~ #
~.
D01t ~
00'1
c
~110
(0) (2) (3)
110
d
(2) (1)
(0)
f
11
~
(2) 1 (3)
00
10
g h
"V" 01
Nothing has been said until now about the sense of the dislocation line
and of the Burgers vector. If a dislocation segment bounds an L-field,
its sense determines a sense of rotation around that field, which may be
clockwise or counter-clockwise. In an analogous manner the Burgers
vector may "flow" out of a node or into it, i.e. it can be fixed at the
node either by its tail or by its tip. The relation between the sense of
rotation and the flowing in or out of a node is remniscent of the relation
between curl and divergency in electrodynamics. In order to settle a
convention, we assume the principal sense of rotation around an L-field
to be counter-clockwise. Thus, we see from Fig. 9.1/1.g, h that, if the
dislocation segment follows this principal sense of rotation, the corre-
sponding Burgers vector flows out of the corresponding node. Therefore
we can formulate D 4 as:
counts and again not the orientation of the Burgers vector. A Burgers
vector drawn in an L-configuration serves only as an aid for visualizing
the edge- or screw character. When the L- and the b-configuration are
drawn, the line sense and the sign of the Burgers vector can be neglected.
They can always be reconstructed from the diagrams and from the
numbering scheme.
10
-11~OO
01
L b
110
10
• 00
01
I
10
.. 11<1>00
01
01Ql0
00
01
h
11
00
.. 01<:])10
11
Fig. 9.3/2. Indifferent dislocation reaction
which edge and screw orientations are present. As the energy of an edge
dislocation is 50% higher than that of the screw dislocation, it may
happen in this case that the reactions do not occur, although the b-net
has been subdivided into acute-angled triangles. It is to be noted that,
in this case, a right-handed and a left-handed screw dislocation intersect.
According to the numbering system in the L- and the b-configurations,
it can be seen that the rotation sense is inverted in the two plots.
The energy of a dislocation is not only given by the Burgers vector
and the edge and screw character, but also by its length. Dislocation
reactions may also occur if energetically unfavorable Burgers vectors are
produced, if the total energy can be lowered. In the bcc structure (iron)
small segments of <110 )-dislocations have been observed as reaction
<
products of two 1 00 )-dislocations (Hale and Brown, 1968).
104 9. Dualistic Representation of Dislocation Reactions
l b
13 12 11 10
-LIll 03 02 01 00
a b
13 12 11 10
/5lS5!
03 02 01 00
d c
ordinate system. Both, the dislocation lines and the Burgers vectors are
referred to the same coordinate system. Now, on the b-side, a starting
point (e.g. the point 00) is chosen, and according to D4, we move round
the field 00. The corresponding Burgers vectors are accordingly joined,
either by their tails or by their tips to the b-node 00. The other ends of
the Burgers vectors are then numbered according to D 3; every end
receiving the number of the corresponding neighbouring field. After-
wards the construction continues in an analogous manner by moving
round the neighbour fields one after the other, drawing the Burgers
vectors and again numbering the ends, until every L-field is attributed
to a b-node and the Burgers vector of every dislocation segment has
found its place in the b-configuration. When all the fields have been
dealt with in this way the whole b-configuration is constructed. Different
projections may be required in complicated cases.
9.3. The Prediction of Dislocation Reactions t05
l b
-
12
13
.. 13 12 11 10
~
03 02 01 00
a b
~1J •
02/ 12
~_ _.;.;.11_ _:
00 I 10
c
L b
--
05
.QL
•
.. 02 01 00
03
..
~
\
-02\
..
-(00')
05 04 03
01
..
05
04
..
02 01 00
I •
-
•
t2
01
.. ~
05 04
•
00
--
05
04
•
01 00
O} • •
~
/01 05 04 03
•
00
Since the dislocations are not joined at nodes, the b-polygon chain is not
closed.
The procedures given in this section can also be extended to partial
dislocations, as is done in the following section.
9.4. Dislocation Reactions Involving Partial Dislocations t 07
116m2]
1f6[121l~ 116[211]
22
b 02
B B
..
c c
c B
d
e
Fig. 9.4/t. Reaction of partial dislocations (triangular node)
108 9. Dualistic Representation of Dislocation Reactions
21
22 00
11
12
b
I
a
B B
c
c
d
e
Fig. 9.4/2. Reaction of partials in the hexagonal structure
9.5. Thompson's Notation of the Burgers Vectors 109
AC
[>0
C
only glide is allowed for parallel edge dislocations, then certain stable ar-
rangements exist (Section 7.3).
Dislocations which attract one another can also merge partially or
completely and form branches, i.e. they can react among themselves.
Here, the important law of continuity of the Burgers vectors, i.e. Frank's
node condition, has to be fulfilled (7.4). Besides this condition, the b2_
criterion (which, however, is only approximately valid) helps to decide
whether a dislocation reaction might take place or not. Quantitative
examples are given in Section 7.5.
Partial dislocations, representing partial displacements of a perfect
dislocation, are treated in Section 8.2. When their configurations are
considered they are seen to be limits of stacking faults. According to the
b2-criterion partial dislocations of a perfect dislocation repel each other,
but they are held together by a strip of stacking fault, the width of
which is determined by the stacking fault energy (surface tension).
The sequence of the partials in the fcc structure is not arbitrary, i.e.
the addition of the partial Burgers vectors not commutative. The sequence
can be determined geometrically by means of Thompson's tetrahedron
and the movement rule [Eq. (5.4-i)], or algebraically by means of
Eq. (8.3-13).
As an example of partial dislocations and stacking faults the phase
transition of cobalt from the fcc- to the cph-structure is discussed in
Section 8.4.
For the study of more complex dislocation reactions, the dualistic
representation was developed, whereby the Burgers vectors and the
dislocation lines are completely separated. Frank had already separated
the Burgers vector from its dislocation by representing it in the reference
lattice and not in the crystal. In principle, this idea is further developed
here. On interpreting the node condition on a geometrical basis, a strict
duality between the configurations of the dislocation lines and those of
the Burgers vectors appears (Section 9.1). The b-net represents a plan
of all the displacements on one side of the dislocation network with
respect to the other side which is considered to be fixed (Section 9.2).
The possibilities for dislocation reactions thus become directly visible
(Section 9.3). Reactions of partial dislocations can also be treated in
this manner (Section 9.4).
In order to interpret a given dislocation arrangement which has been
observed by electron microscopy, for example, the following practice
may help: A transparent paper is put on top of the photo. Then, in
every field bounded by dislocations a representative point, if possible
the central point of the field, is marked. Now all the neighbouring points
are connected across the dislocation lines. Afterwards, the configuration
is taken out of the dislocation network and rotated by 90°. It re-
112 9. Dualistic Representation of Dislocation Reactions
location will tend to align itself along the valleys. If it changes its path
across a potential hill it will be bent (Fig. 10.1/2), and such local bending
E
#//////////////1///////
1///////;;
1/'////11//11//////1//////
y
Fig. 10.1 / 2. Representation of a "kink"
10.1. Dislocations and Point Defects 115
11
to '1 is:
(10.1-1 )
A jog in a screw dislocation acts as a local edge component with the
corresponding extra plane. If the dislocation moves together with the
jog, the extra plane of the jog either grows or diminishes, and vacancies
or interstitials are thus produced. r' shall be simultaneously the direction
of movement of the jog in dislocation 1 and the total amount of the
displacement with respect to a coordinate system fixed in the crystal.
The movement of the jog can have two components, one perpendicular
to dislocation 1 (the same as r) as though it were rigidly fixed to the
10.2. Dislocation Dynamics 117
dislocation, and the other along the dislocation. The number of point
defects, N, produced on the path r' is, again according to Homstra
(1962) :
bl [r' X b 2 J
N =sgn (b 2 ) --v-- (10.1-2)
(2)
.
III! III HI
p
(I)
(a) (b)
Fig. 11.1/1. Fitting of two crystals. a) Perfect crystal divided into two parts,
b) both parts separated and crystal 2 rotated
-~
ilrHrrHl
~b(C)
~(I)
Cc) (d)
the boundary. We may illustrate this, for example, with the aid of the
moire model NO.1 (Fig. 11.2/1}. The relation is only approximate
because the Burgers vectors of dislocations can only acquire discrete
values, while x and consequently b(C) are continuously variable vectors.
Hence, if crystal 1 and 2 exchange sides the sign of the curl vector
switches, i.e. the line sense alters while the Burgers vector remains
unchanged which means that the dislocations switch their sign too.
In the next section a classification of subgrain boundaries is developed,
based on this cell structure.
124 11. Dislocation Networks-Subgrain Boundaries
Boundary Orientation of ()
special general
1- () special pure twist general pure
boundary twist boundary
Arbitrary with respect to () special partial twist general partial
boundary twist boundary
II () * special tilt boundary general tilt
boundary
* For the tilt boundary finer distinctions are needed, which will be given
in Section 11.4.
11.4. "Special" Subgrain Boundaries 125
Cf 20 40 60 eo .~
e ,
Fig. 11.4/ 1. a) Special pure twist boundary on the {111 }-plane in the f.c.c.
structure; duality between the L- and the b-net. b) Hexagonal dislocation
network in iron. (Photo: A. Mascanzoni)
126 11. Dislocation Networks- Subgrain Boundaries
O,2~
I I
Fig. 11.4/4. Bent partial twist boundary. (Bottom half is more of a twist-
while top half is more of a tilt boundary.) (Photo: M. Malone)
128 11. Dislocation Networks-Subgrain Boundaries
It was shown in Section 7.3 that two parallel edge dislocations could
form a stable array, if glide were permitted but climb excluded. We now
have the justification for this in that such arrangements are the origin of a
stable tilt boundary.
2. The orientation of the tilt boundary may be random with respect
to the cell structure. However, the boundary will not cut through the
cell walls in an arbitrary manner, but will fit into the MS-lines as a
stepped surface. Hence, such a boundary will be stepped and will consist
of different types of edge dislocation (see Fig. 12.3/4). It does not have
to be even approximately planar but can be curved and broken along
the whole length (Fig. 11.4/5).
foIA.--:F--:P'--:::iit""-7tj- - -- - - - -- - - - - - - --
~ ' Iii
[00; ;fll]
__
jI[] ~=j
----- -- -------
12
Fig. 11.5/1. Step in the b-net produced by a foreign dislocation
9 Bollmann, Clystal Defects
130 11. Dislocation Networks-Subgrain Boundaries
L b L b
a b
Fig. 11.6/1 a and b. Summary of the different types of subgrain boundaries:
a) special-and b) general boundaries. 1. Pure twist -, 2. partial twist -,
3. tilt boundaries
11. 7. The Energy of a Subgrain Boundary 131
(11.7-1)
=~+1;
1;. = elastic energy of the dislocation in its surroundings,
1; = core energy.
The number of dislocations per unit surface is:
N- 1 (c~L (11.7-3)
- h(cm)
Within a distance from the boundary which is much larger than h, the
strain fields of the dislocations will cancel out and the crystals will be
essentially undisturbed. Therefore, we choose ro = h and obtain for the
elastic energy of the boundary:
N· T,=
1
(~).~~-~ln(~)
h 4n(1-V) b
=Eo () In (1/()) (11.7-4)
=-K.)()ln().
9*
132 11. Dislocation Networks-Subgrain Boundaries
(11.7-7)
(11.7-8)
E ()
---~- [ 1-ln~-
() 1. (11.7·9)
--
Emax - (}max (}max
1.0
0.8 ./
v
~
..:;
0.6
/
/
"-
...... 0.4 /
0.2 /
II
o 0.2 0.4 0.6 0.8 1.0 1.2 1.4
(J/(Jmax-
This equation is no longer valid at large angles, (), where the spacing
between the dislocation cores becomes so small that the boundary be-
comes a high angle boundary.
250
kg/mml '-... I
r-- _I
I1
200
1 1
en
V)
'"co 150 l 1
1
1
"E I
\ I
0
I
.c
~ 100
'"
-'"
u
:>
I
I
I
'J1
50
I I
L I
o 100 200 300 400 500 600 700 800 900 °61000
Temperature
Fig. 11.8/1. Hardness of pure nickel during recrystallization
ture of the boundary may still be larger than rc whereby the grain
continues to grow. At an angle of misfit of around 20°, the specific
surface energy E~ attains its maximum value. From this point on the
138 11. Dislocation Networks-Subgrain Boundaries
L b
---- - -------
c ----
e
I 2
0-------0
1 I 2
I 2
=~
I 2
= 0+---0
2 I
2 I 2 0----0
3
I 2 ~2
4
I 2
:I :I
Fig. 11.10/1. Three-dimensional duality between subgrains and rotation
vectors e. 1. Right-handed twist boundary, 2. left-handed twist boundary,
3. partial twist boundary, 4. tilt boundary
140 11. Dislocation Networks-Subgrain Boundaries
c e
3
1
1- 1<] 2
1<]
2
2
~ 3
1- [>1
2
>+- [>1
3
4
2
2\73
5
>+- 1
f,- \72
3
6
I
i
Fig. 11.10/2. Dualistic representation of the junction of three subgrains
(Abbreviations: Tw twist boundary, Tt tilt boundary, r right-handed
rotation, lleft handed rotation). 1. All three boundaries Tw r; 2. (1,2) Tt,
(2,3) Tw 1, (3,1) Tt; 3. All three boundaries Tw 1; 4. (1,2) Tt, (2,3) Tw r,
(3,1) Tt; 5. (1,2) Tw 1, (2,3) Tt, (3,1) Tw r; 6. All three boundaries Tt
Chapter 11. Summary and Discussion 141
C-space B-space
y
4
/
V
m2 /
V
;/ ~i'-
v:
X85
/
/ 51 31 3
/ V41 13\ 55
/
/
V 25 11 5 2!j....-' ~
/
V13 5 ~~ 53
/ ~5/, ~ 11 V
13 \31 ~25 Z
~~21 'j
~ 1.":;0 - ~
.....", ml x
~ t::-- I i
.KIll r<:::- ............ I 1
'- r-.. . .
'-,-
~
other by the [1, 1J-axis. Any point from outside such a sector can be
transferred into a point inside by a rotation of 0 < 45 0. The angle
between the mirror axis and the position vector of one of the two points
related in this way is 0/2. Fig. 12.2/1 shows that the structure of the
pattern does not depend on whether a point of sector 1 is rotated into
one of sector 2 by + 0 or vice versa. The orientation of the pattern,
however, is not the same. On rotation by positive angles 0 the correspond-
ing lattice coincidence sites lie in sectors 2 and 4, and for negative angles
o in sectors 1 and 3.
In our example of the square lattices the coincidence site lattice must
be itself a square lattice, since for every point rotated over the x-axis
a similar point must be rotated over the y-axis. A measure of the coinci-
dence site lattice is given by the surface of its unit cell. When considering
whether a point within a sector (e.g. sector 2) is a possible lattice coinci-
dence site close to the origin, it must be assured that its coordinates (x, y)
do not contain common factors, since this would mean that a fraction
of the position vector would again lead to a coincidence site. In particular,
the two coordinates should not be even.
For a chosen point (x, y) of lattice 1, for example (Fig. 12.2/2), the
surface of the square of the position vector is given by:
with IX = 2 if E' is even, and IX = 1 for E' odd. E is called the" generating
function" .
For the point (4,1), E'= 17 and the centre of the square [(2,1/2) +
(-1/2, 2) = (3/2, 5/2)J is not a lattice coincidence site, as its coordinates
are not integers. Care must be taken with large coordinates since, oc-
casionally, certain lattice coincidence sites can belong to several coinci-
dence site lattices.
10 Bollmann, Crystal Defects
146 12. General Geometrical Theory of Crystalline Interfaces
o
o 0
000
o 0
0 0 0 00 0
0
o
o 0 0 0 0 0 0 & Q, co
0
0 0°0 0
0
,,° 0
0
Q
0
000 0 0 0 0
o 0 0
o
Fig. 12.2/3. Superposition of two square lattices. The picture contains four
coincidence sites
etc.}. The determinant IAI =1=0 means that both lattices shall be defined,
with respect to the number of dimensions, in the same space, i.e., for
example, a three-dimensional lattice shall not be imaged onto a two-
dimensional one.
The whole set of the lattice points may be called an "equivalence
class" or simply a "class". Hence, the two lattices are two related
classes of points [related by Eq. (12.2-4)]. Every lattice point is therefore
an "element" of one of the two classes. Lattice coincidence sites are thus
coincidences of elements of related classes. [They are not coincidences of
related elements, since by the relation in Eq. (12.2-4), a coincidence is
only possible at the origin.]
In the next section we shall generalize the notion of the coincidence
site lattice based on the above reflections.
1. [0 ,01
2. [1/5,3/51
3. [2/5,1/51
4. [3/5,4/51
5. [4/5,z;51
Alternatively, we can, starting from reel) (C(l»), find other elements e (C(l»)
of the class C(l) by adding translation vectors tel) of lattice 1 to re(l).
Hence:
(12·3-3)
If the points defined by Eqs. (12.3-2) and (12.3-3) coincide, we designate
that coincidence point by re(O) (Fig. 12.3/2):
re(lI) (C(II») =re(l) (C(l») +t(l) =re(O) (C), (12.3-4)
or
(l-A-l) re(O) = tel) (12·3-5)
(I = identity = unit transformation).
As we shall see later (Section 12.5) that all the possible translation
vectors tel) of lattice 1 form the b-Iattice, we already change the notation
here; hence t(l):= bILl (L means lattice point). Thus:
All the o-points are solutions of Eq. (12.3-6). We shall see that this
equation is the basis of the whole geometrical theory of crystalline
interfaces.
The general procedure for solving the equation is as follows:
1. Choice of the coordinate system.
2. Formulation of the transformation A which relates the two
lattices.
12.3. The O-Lattice 151
(12·3-7)
A=B= (
COS 0 - sin 0) . (12·3-8)
sin 0 cos 0
From this follows:
A-I = (
COS 0
sin 0) (12·3-9)
- sin 0 cos 0
and
If II-A-11 =\=0 the solution is given by Eq. (12.3-7). In our example the
determinant is:
II-A-11 =2 (i-cos 0) (12·3-11 )
As in the crystal coordinate system the two basic translation vectors are
given by the coordinates (1, 0) and (0, 1), the unit vectors of the O-lattice
are the cloumn vectors of the matrix (12.3-12), i.e.
u~O)= (1/2, -(1/2) cotan (0/2)) and u~O)= ((1/2) cotan(O/2), 1/2).
x.1!z
( \ \ ~
\ I \ ~ \ ~
-lr\ -t-I 1\ \\ ~ --
~
...-'-
\! .q= XIO) -
t-~ -- --1+ -- ....<'y-
~
l/Z
--
~
~
- ,.- 1\1 \
- 1-
--
~X
- ~
!-(
\
1\ 1\ .1 e---- ~ ~
_\
I-
)-- ~ ~
f- I \ 1\
Fig. 12.3/3. Construction of the O-lattice according to Eq. (12.3-12)
x~Ol
(xiOl) -_ (
1/2 3/2)
-3/2 1/2 b~Ll·
(biLl)
(12.3-13)
The unit vectors of the o-lattice are '140) (1/2, - 3/2) and u~O) (3/2, 1/2)
(see Fig. 12.5/1). Every second O-point is a lattice coincidence site. The
other o-points belong to the equivalence class [1/2, 1/2] * (Fig. 12.5/1).
In the case of cotan (0/2) = 4 the following equivalence classes
contain coinciding elements: [0,0], [1/2,0], [0,1/2] and [1/2,1/2],
which may be shown as an exercise.
We see that the O-lattice, in contrast to the general coincidence site
lattice, is a continuous function of the transformation A (e.g. of 0) and
that an O-lattice also exists where no coincidence site lattice is present.
A slight change of A leads to the instantaneous disappearance of a
coincidence site lattice, while the O-lattice undergoes a smooth change;
a property which is common to the MS-points (Section 11.2). It will be
shown in the next section that Eq. (12.3-6) can be understood as a
generalization of the quantized version of Frank's formula [Eq. (11.2-5)],
and that, hence, the MS-points may also be looked on as special o-points.
* The equivalence classes are indicated by boldface brackets [ ... ].
12.4. The O-Lattice and Frank's Formula 153
°
I °
° ° °
- -t,t-'*---4--_°-1._o
° ° fo '0 ° -
X(O)_ - '-t--'-_-J~'--''''-'._.-_-,-_0
2 ° ° - Xm
-1-"'- U ° ° 1
, "'- -X(2)
° ° ~ 1
- -t-Ti-r-'*_°--JLo
° ° ° ° "° -
--f-"'--1,"---1H-_-L °
° I ~ "'-,.-t-
A 0' ° ° °
8
aJ bJ
cJ dJ
Fig. 12.3/4. Application of the O-lattice to the construction of a symmetrical
and an unsymmetrical tilt boundary
b2 = OXI (12.4-1)
ba = 0
1 54 12. General Geometrical Theory of Crystalline Interfaces
where b (bt b2 ba) represents the discrete points of the b-net and (J is
placed in the z-axis. In matrix notation this equation becomes:
(12.4-2)
On the other hand, according to the procedure given in Section 12.3, for
a rotation around the z-axis, A is given by:
A= CO -sin ()
s~() cos ()
0 ~) (12.4-3)
C)
and Eq. (12.3-6)
b~L)
b~L)
= r-co,o
sin ()
0
-sin ()
1 -cos ()
0
0)C~)
o
o
x~O).
x~O)
(12.4-4)
If () is smail, cos ()~1 and sin ()~(), and Eq. (12.4-4) becomes:
0)o (X~O»)
x~O) (12.4-5)
o x~O)
Fig. 12.5/1. The b-Iattice pictured as the lattice of all the translation vectors
of lattice 1
~~~~-t---+,.L\
~~~~~~~~~~~~,~~~
-+--~~--~~~~~~~~'7_~-~~
,
\-)(
~,
Fig. 12.6/1. o-points as origins of the imaging between lattices 1 and 2
12.6. Various Aspects of the O-Points 157
which is the basic Eq. (12.)-6). This is actually the way that the O-lattice
was found.
At this stage the O-points, the generalized MS-points, were localized
but their meaning was not immediately apparent; the meaning, however,
is clear from Fig. 12.6/1 (Fig. 12.6/2 shows the complete pattern).
Fig. 12.6/2. The complete pattern of Fig. 12.6/1, showing the symmetry of
the point arrangement (same pattern as Fig. 12.2/3)
(12.6-4)
The point a:(2L) is here no longer related to a:(lL) but to another point
of lattice 1 i.e. a:(lL) +b(L) (i.e. a:(lL) +t(l»). From Eq. (12.6-4) it follows
that
(12.6-5)
and, applying Eq. (12.1-4)
_a:(0) =Ab(L) -Aa:(O). (12.6-6)
Upon multiplication of both sides by A-I, the basic Eq. (12-3-6) is
produced. Therefore, all the O-points are origins of the type described
above.
Naturally, the various meanings of the O-points indicated in this
section, such as:
Coincidences of equivalent points,
Generalization of MS-points, and
Origins for the relation between lattice 1 and lattice 2
are different manifestations of the same situation. However, which of
the various aspects has to be emphasized will depend on the problem
being treated.
(12.7-8)
and the basic equation:
i.e.
(12.7-9)
As an example, we investigate the rotation on the (110)-plane in the
bec structure, which is illustrated by moire model No. 1. The unit
vectors of the crystal coordinate system are given by the vectors
[1/2, 1/2, 1/2] and [001] in the cubic system (Fig. 12.7/1), hence
8= ( l 0)
1'2
1 • (12.7-10)
160 12. General Geometrical Theory of Crystalline Interfaces
•t xl1J2 .
10·
x~lI
1 xiO)
GO·
..ft':::"2 x iI)
': .:.':'.' ': . _ _ 1
'::: . . :::
:::,;.; . ,
• XIO)
1
50' 29'
60·
70' 32'
xli)
1
80· ': ,':'
Fig. 12.8/ 1. Comparison between calculated O-lattices and the moire models
No.1. [A : Eq. (12.7-3), A' : Eq. (12.8-9) J
11 Bollmann, Crystal Defects
162 12. General Geometrical Theory of Crystalline Interfaces
(12.8-1)
(12.8-2)
Fig. 12.8/2a and b. Point pattern for e = 50° 29' o-lattices for: a) A = rota-
tion, Eq. (12.3-8), b) A' = pure shear, Eq. (12.8-9)
This means that we choose new unit vectors in crystal 1, and starting
from these, we apply the rotation S-1 RS (in crystal coordinates) III
order to arrive at ;r(2).
(12.8-3 )
s
--=--
, ,,
,,
2
,
/
2, '
,,
,,
/
/
/
, 1
,, , /
,, , /
/
, /
/
/
.- /
0
1 en1 ;...",
~I .~I ""1
;'1 =1 0..... 1
1
""1 Lrl,
-1
0 10 20 30 40 50 60 70 80 90
Fig. 12.8/4. Determinant II - A-II. 1. Eq. (12.8-12), 2. Eq. (12.8-13)
12.8. Choice between Various Possible Transformations A 165
Oet II_A-1)
, ,,
,
/ /
... ...
...
,, /
.,.. ...
... ... ' ,
- I
~,
....
•
«>
I
~+---~--.---r---r---r---'---'---r--.
o 10 20 30 40 50 60 70 80 gil
8-
Fig. 12.8/5. Determinant II - A-II for rotation on the {112}-plane
166 12. General Geometrical Theory of Crystalline Interfaces
-1
a
U'
~
-1
b
Fig. 12.8/6. Difference in the meaning of V, Eq. (12.8-2), and V', Eq. (12.8-19)
12.8. Choice between Various Possible Transformations A 167
Fig. 12.8/7. Elements for the calculation of the Wigner-Seitz cell walls
then
(t'-t)· ~-(1/2)(t'+t). (t'-t) =0 (12.8-26)
and
[(t'·~) - (1/2) (t'. t')J - [(t·~) - (1/2) (t . t)J = O. (12.8-27)
crystal coordinate system, U~l) = [1 OOJ}. The second and the third
column vectors of U are determined in the same way, with U~l) and Uhl ) .
If two of the values I (~, t) are the same, then the point ~(2L) lies
at the same distance from two different points. (This situation corre-
sponds, for example, to the meeting point of the curves in Fig. 12.8/3).
If all the values of I (~, t) are ~ 0, then t = [OOOJ and t(l) = to' If several
t-vectors are closest to ~, t should be chosen such that U is unimodular
(det (U) = 1), as only then the two lattices are in a 1 to 1 relation.
This procedure can be programmed on a computer.
An equivalent procedure is to find the smallest value of the square
of the distance to the corner t:
(t -~)TG(t -~) (12.8-32)
where for t all the eight corners of the unit cell have to be checked.
t~ Xl (0) +t~
X2 (0) +t~
X3 (0) = b2 (L) (12.9-5)
t~ Xl (0) +t~ x 2 (0) +fs x 3 (0) = b3 (L).
The rank of the matrix T may be 3,2, 1 or 0 in the (trivial) limiting case.
In the case where the rank (T) = 3 the solutions ~(O) are given by
(12.9-6)
In crystal coordinates the b(LLvectors are of the form <100), and hence,
the column vectors of the matrix T-l are the unit vectors of the O-lattice.
170 12. General Geometrical Theory of Crystalline Interfaces
(12.9-7)
or
(12.9-8)
(i) and (k) indicate the fixed indices. t(;) and t(k) are, as already mentioned,
vectors in the reciprocal lattice 1 and v(O) is a vector in the crystal
coordinate system of lattice 1. The vector product procedure [Eq.
(12.9-9)J may be verified by formulating the equations of the planes, e.g.:
t~ vl + t~ v + t~ v
2 3 = 0
(12.9-10)
t~ vl + t~ v +t~ v
2 3 = 0.
(12.9-11)
X2(0) = A v 2 +
x 2 (P) (12.9-15)
x 3 (0) =A v3 +x3 (P)
Next we set
(12.9-20)
hence:
;r(0) = Q-l ;r' (0) . (12.9-21)
Along with Eq. (12.9-17), this gives
(RTQ-l) ;r'(0) =b'(L). (12.9-22)
This equation states that every point in the (b 1 b2 )-plane has as its
image an x-line parallel to the z-axis. The representation of the O-lines
is independent of x 3 • We can now reduce the system of equations from
three to two dimensions by simply ignoring the third component.
1:-_:_llL:1 1:1:
I
=
(12.9-23)
(12.9-25)
The solutions are
;r' (0) = T'-1 b' (L). (12.9-26)
Here we have obtained the standard form for rank (T) = 2 which re-
presents the O-lattice in the projection perpendicular to the O-lines. The
orientation of the O-lines in the orthogonal, or in the crystal coordinate
system is given by v(O) [Eq. (12.9-9)].
In the case of rank (T) = 1, the three planes represented by Eq. (12.9-5)
are parallel. In order that solutions exist, the three planes have to
coincide. Hence, only one b-component can be chosen freely, since the
others are thereby fixed. This means that, in this case, the b-subspace
is a straight line. For every b-lattice point on that line the solution of
Eq. (12.9-5) is an O-plane. All three row vectors of T are parallel and,
as such, are normal vectors of the O-plane. On the other hand, the
column vectors of T are also parallel and lie in the b-subspace, i.e. the
b-line. Each of the three equations of (12.9-5) is an equation of the
O-plane if b(L) is a lattice vector of the b-lattice and lies in the b-line, i.e.
the b-subspace.
By a procedure analogous to that used for rank (T) = 2 we can find
a standard form for rank (T) = 1. Here a normalized row vector which
represents the normal to an O-plane is denoted by v(O), and a normalized
1 74 12. General Geometrical Theory of Crystalline Interfaces
(12.9-27)
Hence:
t'g (x' (0))3 = (b' (L))3 (12.9-28)
and the standard form for rank (T) = 1 is:
'(0))3 _ (b' (L))3
(
x - t'i . (12.9-29)
We designate the O-points, O-lines and O-planes as O-elements. An
O-element is the image of one b-lattice point.
The results of this section are presented in Table 12.11/T 1. The
O-elements which are those regions where the two lattices show best
fit, are points, lines, or planes depending on the rank (T). O-planes are,
for example, twin planes. The (trivial) limiting case of rank (T) = 0
signifies that the transformation A is equal to I i.e. we are dealing with
a perfect crystal containing no interface.
In the case of rank (T) = 2 we can interpret the vector v(O), which
fixes the orientation of the O-line, as an invariant eigenvector of the
transformation A. An eigenvector maintains its orientation during the
transformation but may change its length by a factor A.
A;r =A;r. (12.9-30)
If, in addition, it conserves its length we call it an invariant eigenvector
(A = 1), i.e.
(12.9-3 1)
We write
(12.9-3 2)
(i.e. we refer to the end of the transformation ;r(0) instead of to the
initial ;r). Thus:
(12.9-33)
and, along with Eq. (12.9-31) this gives:
(l-A-l) ;r(0) =0 (12.9-34)
which is the homogeneous part of the basic equation.
12.10. The Behaviour of the O-Lattice upon Translation of Lattice 2 175
(12.10-1)
(12.10-2)
and
~(Od) =A-1 ~(Od) _A-1 d(2) +b(L) (12.10-3)
( l - A - l ) ~(Od) =b(d) (12.10-4)
with
b(d) = b(L) + d(b) = b(L) _ A -1 d(2) (12.10-5)
d(b) is the displacement of the b-lattice. Hence:
d(b) = - A -1 d(2). (12.10-6)
If ~(Od) is decomposed as follows
~(Od) =~(O) +d(O) (12.10-7)
then:
(12.10-8)
and hence:
(12.10-9)
or
(12.10-10)
176 12. General Geometrical Theory of Crystalline Interfaces
This latter relation is valid in any case for rank (T) = 3 and will have to
be rechecked for the other cases (Section 12.11). In summary, if lattice 2
is translated by d(2) , then the b-lattice is displaced by d(b) , with
(12.10-11)
b2
b',
-,
,,
iiT ,!...
a ,
I xiI)
2'
10' +t-
r
b L
60' ,
:t'
C -1 0 1 2
SO'29 \\
\ ~
d -I 0 1 2
~\\
e
\ .
.J t';------.- .........
. .
'
No O-Lines
., .
, ,
, ~-----',-..,...
<,
f
J'~.
. o
,
, ,
,
Fig. 12. 11 / 1. Comparison of the calculated O-lattice-translation with the
moire model NO.1
12.11. Translation of the O-Lattice for the Case of Rank (T) <3 179
lattice 1, cannot be imaged onto one another by the pure rotation, i.e.
there are no coincidences of equivalent points i.e. no O-lines.
An analogous situation appears in the case of rank (T) = 1. Here the
translation d(b) of the b-Iattice has to be split into one component parallel
to the O-line and another perpendicular to it. The translation of the b-
lattices provokes also here a corresponding translation or disappearance
of the O-lattice.
Now we are in the position to solve the important problem which
arises if the b-subspace contains few b-Iattice points or none at all. In
this case, the homogeneous transformation A could produce an O-lattice
consisting of only a few O-elements. The O-lattice would be "diluted".
On the other hand, the b-vectors between the b-Iattice points (i.e. the
Burgers vectors of the dislocations between the O-elements) would
become very large. In reality, however, a "dense" O-lattice will appear
with elementary b-vectors. For this reason, O-elements will also have to
originate from b-Iattice points in the close surroundings of the b-subspace.
We project these b-Iattice points perpendicularly onto the b-subspace,
where the component b(LIi) determines the position of the O-element,
and the component b(L.l) corresponds to a translation of lattice 2 with
respect to lattice 1. The O-lattice is determined here by a nonhomo-
geneous transformation where the homogeneous part is the same for all
O-elements, while the translation may be different for each O-element.
We consider the case of rank(T) =2 and split b(L) as follows
(12.11-5)
with
b(LIi) = b(L) - (Cb(L)) c (12.11-6)
c is the unit vector normal to the b-plane (= b-subspace)
b(LJ.) = (Cb(L)) c. . (12.11-7)
Now the basic equation becomes
(l-A -1) x(O) = b(L) - (cb) (L)C (12.11-8)
and the translation
d(bJ.) = _b(LJ.) = _ (Cb(L)) C = - A - l d(2). (12.11-9)
The change of sign: d(bJ.) = _b(LJ.) arises because the b-Iattice point is
translated from outside into the b-subspace, which is considered to be
fixed; hence, d(bJ.) points towards the b-plane, while b(LJ.) points away
from the b-plane (Fig. 12.11/2a).
Thus:
(12.11-10)
12·
180 12. General Geometrical Theory of Crystalline Interfaces
b
Fig. 12.11/2a and b. Projection of the b-lattice points onto the b-subspace.
a) Perpendicular projection, b) projection along the O-lines in a general case
Pi"
rt-
Rank(T) O-Elements Determination of the b-Subspace Determination of the Determination of
(nD = n-dimen- O-Elements b-Subspace 0-Elements if b(L) o·
;;
sional) does not lie in the o>+>
b-Subspace rt-
::r
(p
2 O-line (1 D) Vector product of row b-plane (2D) Normal to b-plane Normal projection of ~
vectors of T +partic- =vector product of b(L) onto the b-plane ri'
(p
ular solution column vectors of T
8'
...,
O-plane (2D) Normal to O-plane b-line (1 D) Column vector of Normal projection of st
b(L) on b-line (p
= row vectors of T = b-line
T + particular solution D
(f1
(p
0 O-space (3 D) Total x-space b-point (OD) Point b [0, 0, OJ Translation into b-point o>+>
(perfect crystal) (lattices 1 and 2 ::0
coincide perfectly) po
;;
:>;'
-:l
1\
w
....
....00
182 12. General Geometrical Theory of Crystalline Interfaces
(12.12-2)
(12.12-3)
The imaging relation between the b-space and the crystal space is
given by
b(~) = (l-A-l) ~:= T~ (12.12-4)
and
(12.12-5)
(12.12-6)
(12.12-7)
In matrix form:
(12.12-8)
Chapter 12. Summary and Discussion 183
with
(b - b(L)f = bT _ b(L)T. (12.12-9)
From (12.12-8) follows that
bTGb -bTGb +b(L)TGb +bTGb(L) _b(L)TGb(L) =0. (12.12-10)
The first two terms cancel one another. Since all the terms are scalars
they are equal to their transpose. Because (AB)T =BT AT:
bTGb(L) = [bTGb(L)V =b(L)TGTb =b(L)TGb. (12.12-11)
Ib (x)12 = Ib (x _x(O))12
which leads to Eq. (12.12-13) and is equivalent to an imaging of Wigner-
Seitz walls of the b-lattice onto the O-lattice (Section 12.12).
Chapter 12. Summary and Discussion 185
The equation for the cell walls is plausible but not completely sure.
Since the intersections of the interface with the cell walls represent the
dislocations (as was already shown in Section 11.2) it is possible that
elastic anisotropy has to be considered. Here we have chosen X(2) as
variable x. Once the choice of X(1), or a point between X(2) and X(1) is
made, the orientation of the cell wall may vary, but it must always
bisect neighbouring O-elements.
For given structures and orientations of the two crystal lattices, all
the elements of the O-lattice are determined with the aid of the indications
given in this chapter. The applications are dealt with in the in following
chapter.
Suggestions for Problems
There is no lack of possible problems in this chapter. The calculations
of Sections 12.7 and 12.8 can be carried out explicitly on given examples,
such as moire models Nos. 1, 2, and 3. The transformations which relate
nearest neighbours can be determined. The calculated O-lattices can
be quantitatively compared with the moire models. Furthermore, the
examples of rotation of moire model No.1 ({ 11 O}-plane, bcc structure)
by 50° 29' and 70° 32' can be reduced to the standard form. Similar
possibilities exist with moire model No.2, for () =34° 03' and 62° 58'.
(The necessary indications are to be found in Table 13· 3fT 1.)
In addition, with respect to the translation and the cell structure, a
sufficient number of examples can be found in connection with the moire
models (Sections 12.10 to 12.12): From the behaviour upon translation
by () = 60° of moire model No.1 ({ 11 O} bcc), for example, it can be shown
that the coordinate system of the O-lattice is left-handed (I TI < 0).
However, care must be taken when using the moire models, since
translation operations carried out by means of sheet a (lattice 1) cor-
respond to d(1) =d(b) instead of d(2) (d(1) = -A -1 d(2)). Rotation, on the
other hand, is done with sheet b (lattice 2).
13. Applications of the O-Lattice Theory
13.1. Outline
Up till now the two adjoining crystals have been idealized as two
interpenetrating point lattices. The O-lattice with its cell structure
introduces an ordering into the interpenetrating lattices. The next step
is the choice of the boundary. Once the boundary is chosen, the atoms
will occupy, on the one side of the boundary the positions of crystal 1,
and on the other side those of crystal 2. The boundary itself is a remnant
of the O-lattice structure.
It shall be emphasized once again that the O-lattice is a description
of those positions where the two crystals fit in an optimum manner. The
cell walls, on the other hand, are the positions of worst fit, i.e. the posi-
tions where the correlation between nearest lattice points in the two
lattices changes between two O-elements (origins!).
A crystal boundary will be placed, as far as possible, through O-ele-
ments. However, not all the possible boundaries are equivalent, since
the energy might vary. The problems related to the boundary energy
will have to be discussed.
Often the problem is not only to find some boundary for some given
relative orientation of the two crystals, but also to first determine the
crystal orientation such that the best possible boundary can be drawn.
Here, neither the relative orientation of the crystals nor that of the
boundary is known at the beginning. We require a criterion for the
optimum boundary, which is basically the criterion for the minimum
boundary energy. In most cases it is not possible to calculate accurately
enough the boundary energy. It will be shown that in certain cases an
energy minimum can be determined without knowledge of the value of
the energy.
We distinguish roughly two ranges of problems according to the
dimensions of the O-lattice unit cell compared to those of the crystal
lattice.
2. the O-lattice unit cell and the crystal lattice unit cell are of similar
dimensions.
13.1. Outline 187
In the first case the two crystals deviate only slightly from one an-
other. The crystal boundaries are formed by dislocation networks where
the intersections of the boundary with the O-lattice cell walls (Sec-
tion 12.12) are the dislocation lines. On choosing the crystal boundary,
a surface, which may eventually be stepped, is traced in the b-lattice;
this surface contains all those b-lattice points which correspond to the
O-elements intersected by the boundary. These selected b-lattice points
form the b-net of the dislocation network in the boundary. The Burgers
vectors are attributed to the dislocation lines by the duality relations
of Sections 9.1 and 11.2.
No optimum boundaries exist for subgrain boundaries, i.e. low angle
boundaries between identical crystals. Here the optimum state is the
perfect crystal, since the minimum energy of a subgrain boundary is at
() =0 (Section 11.3). However, optimum dislocation networks are
possible in phase boundaries. An example will be discussed in Section 13 .2.
The second example of when the dimensions of the O-lattice- and
the crystal unit cells are similar, is the case of high angle- or the general
phase-boundaries. Here the conventional notion of the dislocation
loses its meaning since, if such a boundary is considered as a dislocation
network, the dislocation cores would overlap. Nevertheless, optimum
boundaries are also possible, but the emphasis has to be placed on the
pattern of the lattice points: whether this pattern is periodic or not, and
in the former case whether the period is large or small. It is to be ex-
pected that, for a small period, small groups of atoms come together
and thus give rise to an energy minimum in the sense that any deviation
from that boundary orientation, as well as from the relative orientation
of the crystals, causes an increase in the boundary energy. The problems
arising in connection with the pattern of lattice points will be treated
in Section 13.3.
Fig. 13.1/1 shows an electron micrograph of a grain boundary in
stainless steel taken with the 1.5 MeV microscope at Toulouse. The
boundary shows the type of features as are excepted from the O-lattice
theory. Further electron microscope observations have been reported
and discussed by: Gleiter (1969), Ishida et al. (1969) and Levy (1969).
Calculations on grain boundary energy were done by Fletcher (1967).
2 OOO'&'
Fig. 13.2/ 1. Cryptoperthite (feldspar, moonstone) m monoclinic (homogeneous)
t triclinic (twinned)
colours. For this reason it is often used as a gem stone under the name
" moonstone" .
Chemically speaking, this moonstone is essentially (K, Na) (AI, Sils0 s
and its structure is given in Fig. 13.2/2. Because of the fact that in the
-!SA Of( 00
O!,i---,!,.---!;i-3-!;-i---C'l!;-i oSiAl
Fig. 13.2/ 2. The structure of feldspar
190 13. Applications of the O-Lattice Theory
Table 13.2/T 1
Structure la (A)
Lattice 1 8.561 (6) 12.996(2) 7.193(4) 90° 116° l' 90°
monoclinic
("Ortho-
clase" Or 92)
Lattice 2 8.135(3) 12.788(3) 7.154(2) 94° 14' 116° 31' 87° 43'
triclinic
(Low albite
Kodarma)
First of all, an attempt was made to fit the monoclinic and the
triclinic phases together. The monoclinic phase was taken as lattice 1.
The transformation A was formulated according to Section 12.7. In
order to determine the matrices 8(1) and 8(2) the a-axis was placed along
the x-axis and the c-axis in the (xz)-plane. (For the procedure see
Eqs. (A 1-17) to (A1-26).]
For the determination of the optimum boundary, a rotation R of
only a few degrees is required, since the two phases are very similar.
The complete rotation can be expressed by three separate rotations
around the X-, y- and z-axes. As long as the rotations are small enough,
they can be superposed as polar vectors, i.e.
(13. 2-1)
-1
3 !px
already been mentioned, the energy is not known. For energy calculations
to be carried out, values of material constants would be needed, the
accuracy of which would be doubtful. It is preferable to adhere strictly
to the geometrical approach. From the considerations of subgrain
boundaries, we know that the energy first increases with increasing
dislocation density (Section 11.7). Therefore, we replace the energy
function by a purely geometrical parameter, which we can assume to
vary monotonically with the energy in the range of interest. It does not
necessarily have to be proportional to the energy. We only require that
it increases when the energy increases, and vice versa.
For a series of parallel dislocations of spacing d and with an absolute
value of the Burgers vector b, we replace the energy by the parameter
(13·2-3)
(13.2-4)
This function has a positive definite value, and remains continuous even
when one of the dislocation series disappears, i.e. when one of the d;
becomes infinite.
192 13. Applications of the O-Lattice Theory
xlO)
-3
~ and Fa follow from the cyclic permutation of the indices. d2 is given by:
d2 - _I ~2 (0) m~Ol
X Im~Oll
1-__
m~Ol
1_1 (0)
~2 X~3
(0)1
(13·2-6)
and:
Ib~L)12 Ib~L)12 'Im~OlI2
(d 2)2 = Im~Ol xmls°ll2 (13· 2-7)
and together with Eq. (13.2-5):
p. _ Ib~L)12Im~OlI2 + Ib~L)12Im~OlI2
1- Im~Ol x mls0ll2
(13·2-8)
and, correspondingly Fa and Fa.
13.2. Example of an Optimum Phase Boundary 193
The orientation of the crystal lattices (<Px <Py <Pz), for which the smallest
of the above parameters exhibited a minimum, was understood as being
that orientation for which the O-lattice furnishes the optimum boundary.
The orientation of the boundary was given by the two x~O)-vectors which
were contained in the expression defining that parameter. ~ proved to
be the smallest of the parameters, and its value is given as a function
of the angle <Pi in Fig. 13.2/5. The surprising result is that this parameter
~z
3
2
, ,,
,,
,,
-1 ~~ ______ ~-Y
-1 3
!Px
Fig. 1 3.2/5. Geometrical parameter P 1 for the monoc1inic-tric1inic adaptation
(coordinates as in Fig. 13.2/3)
exhibits two minima. The value of ~ at the minima is about 26 units.
The function is continuous on the surface ITI = 0, and the minima lie
very close to that surface, i.e. the O-lattice of the optimum boundary
is practically a line lattice. The orientations of the crystals and of the
boundary determined above are completely different from the measured
values.
It is to be noted that the triclinic phase is twinned. According to
F. Laves this twinning is an averaging process which takes place in the
triclinic phase in order to achieve a pseudo-monoclinic character. The
twinning of the triclinic phase can be explained by one of two different
twin laws, the "Albite law" (twin axis b*) and the "Pericline law"
(twin axis b). (The details are found in the indicated article.) Accordingly
the matrix 8(2) was averaged, so that now two monoclinic phases had
to be fitted together.
t3 Bollmann, Crystal Defects
194 13. Applications of the O-Lattice Theory
-1 +-Y--r--~-Y
-1
<f;-space and, in addition, the value of ~ is 3 units, i.e. about eight times
smaller than in the case of Fig. 13.2/3 which indicates that here also
the boundary energy is markedly smaller than in the first case. Fig. 13.7
shows the calculated optimum boundary. The calculated dislocation
network might well represent the period in the boundary, but not the
details. The dislocations will split into partials since a Burgers vector
of 13 Awill probably not be stable. In addition, the boundary is disturbed
by the twinning but is nevertheless stable. The comparison between the
c·alculated and the measured values shows good agreement (Fig. 13.2/8).
The differences in the averaging process carried out according to the two
twin laws do not appear to be significant. However, in order to reduce
the twin boundary surface to a minimum, the Albite twinning law is
preferred in this case.
The calculation brings to light another property of the optimum
boundary. The optimum transformation A (<f;~opt)) has three eigenvectors
wjth the corresponding eigenvalues Ai. The optimum boundary contains
those two eigenvectors, with the eigenvalues Ai, which show the smallest
deviation from 1. In the boundary, the crystal experiences no rotation
only extension, and this extension is minimal.
The result of the calculation is surprising if we consider under what
assumptions it was obtained. The only numerical data used was the two
sets of six lattice constants each of the two crystals (Table 13.2/T1); no
13.2. Example of an Optimum Phase Boundary 195
Phase boundary
Lattice (1 )'\attice (2)
b3~
Front view 1 )---t1 12, -+--,f--+--+-~
1 1 1
1 I
I
1-
I I I I i
I I -0--0--0--0--0-
I I I I I
Top view I I I I I
11 I I I
I I I I 1020
- 0 - - 0 - - 0 - -0 --a-
+-Hu;:J"
t I I I I
1I I I I
I I
I 1
i
I I I •
Q2
I 11 _~o_o~ ~09~~0_0? :001 000
I I I I
250 ~
I-----i
Fig. 13.2/8A-H. Comparison between the calculated and the measured data.
Relative positions of the unit cells, measured rotation: 55-60'; calculated:
56'. A: measured eigenvector (EV) [3,0, 1J; B: calculated EV [2.56,0,1];
C: measured EV [1,0, 6J; D: calculated EV [1,0, 11.8J; E to F: range of
the measured normals to the phase boundary (6, 0, 1)-(8, 0, 1); G: calculated
normal to the boundary (12.6, 0, 1); H: calculated position of the boundary
[1,0,11.8J
13*
196 13. Applications of the O-Lattice Theory
liz
-~~-~-
0 0 0
0 0
0 0 0
0 0
0 0 0
a li1 b
c d
Its reduced O-lattice can in this case also contain a maximum of 100
points. If, however, the period of the pattern is smaller, then the reduced
O-lattice contains less than 100 points. It is even possible that it consists
of only very few points (Fig. 13.3/1); occasionally even of a single point.
Starting from a highly periodic pattern consisting, for example, of two
pattern elements, and changing the crystal orientation slightly so that
the pattern is no longer periodic, the reduced O-lattice will consist of
100 points, but these will now be arranged in two small groups around
the original two points. However, if the observed area of the O-lattice is
extended to infinity, these small groups of points will extend, fuse and
finally be distributed homogeneously over the whole unit cell of lattice 1,
which houses the reduced O-lattice. If we consider the orientation of the
crystal as variable, and limit the O-lattice to a finite range, then the
fact that the reduced O-lattice consists of small groups of points can be
considered as an indication that the orientation of the crystals is close
to a periodic pattern.
The search for periodic patterns (e.g. such as may be produced by the
moire models) can be performed by means of a computer in the following
manner: First, 10 X 10 O-points are calculated in crystal coordinates,
according to Eq. (12.8-5):
where xiO) are the column vectors of the matrix (/- U-lS-IR-lStl . Of
the internal coordinates of the O-points, only the first digits after the
decimal point are extracted; e.g. of the coordinates (3.725, - 2.533), the
internal coordinates are (0.725,0.467). Now a 10 X 10 matrix M(j, K) is
defined in which, at the beginning, all the elements are zero. Then, the
element M (7, 4) is given the value 1. The same procedure is repeated
for all the 100 calculated O-points. If, in the course of the calculation,
M(7, 4) appears again, its value remains 1. The value of an element
only indicates whether it is occupied or not, but not by how many points.
The addition of the values of the matrix elements furnishes the number N
of occupied positions.
10 10
N = z:, z:, M(j, K).
J~1 K~l
(13.3-2)
8
20 25 , 30 35 I 45
I
M <=> i:o N
"" ;;; ~ on 0
'"
.....
..... ~ ,., ~
i1::l ""
""
~-------2~0------~25------~30~-----3~
5 ----~--------4~5-- 8
Fig. 13.3/ 2. Number of pattern elements N(fJ) for rotation on the {11 O}-pla ne
(compare with Fig. 12.8/4)
N'
=m
N
' (13-3-3)
N' is the number of crystal units per period of the pattern. The crystal
unit is the volume (the surface resp.) of the unit cell of lattice 1.
The exact values of N, N' and ITI can be determined by means of a
generating function £ in an analogous manner t o that given in Sec-
tion 12.2. E can be generalized as follows:
£' = Iml ~ +m2 u 2 12
(13·3-4)
= [m~I~ 12 +m~ lu212 +2ml m2(~ u 2)]
E = (1/oc) E'. (13·3-5)
Ul , u 2 are the unit vectors of the crystal unit [column vectors of the
S-matrix, Sections 2.2, 12.7, Eq. (A 1-26) J.
~,m2 are the coordinates of the considered point of lattice 1 (according
to Fig. 12.2/1).
oc = 1 or 2 depending on whether E' is odd or even.
200 13. Applications of the O-Lattice Theory
3. From that value in the table which lies closest to the () calculated
by (13.3-6), ITI is taken, and an approximate value of N = E . ITI is
determined. As the table was calculated only in steps of 0.1°, the cal-
culated N might not be an exact integer, but the nearest integral value
will be the correct one. The exact ITI is obtained from N/E=N/N'.
From the coordinates of ~}O), the internal coordinates of the O-points
and hence the reduced O-lattice can be determined.
4. In a second computer programme, on the basis of the exact
()-values, ITI, the complete matrix M(j, K), and the exact coordinates
of the unit vectors of the O-lattice in crystal- and in orthogonal co-
ordinates were calculated. The matrix M(j, K) gave a direct image of
the reduced O-lattice.
The difference between E and N' shows up if ITI =0 (and hence
N' = (0) while E remains finite. It must be stressed that E is a measure
of the coincidence site lattice and N' a measure of the O-lattice. The
coincidence site lattice which represents the period of the pattern is
exclusively determined by the pattern itself, while the O-lattice depends
on the correlation A between the two lattices. All the different O-lattices
of a given configuration contain the same coincidence site lattice
(Fig. 12.8/2). In the case of ITI = 0, a series of coincidence sites lie within
the same O-element. They are not separated by "dislocations". N' makes
this distinction, E does not. Therefore, N' is a more realistic indication
of the energy of a configuration.
If the pattern is represented by different transformations A, then the
total period (the coincidence site lattice) is independent of the actual
value of A. However, the transformation which correlates the nearest
13.3. The Periodicity of the Pattern of Lattice Points 201
Table 13.3/T 1
No.
1° I
/I-A-l/
IN
N'
I m Im
i 2
Equivalence classes of o-points
(reduced O-lattice)
20 20
10 10
10 10
- 20 20
b - JI~IIIIII_I;~IIIIII:III~IIIIIII~Fllllllli~IIIIIII;)~1111''';10
Fig. 13.3/3. a) Pattern with N=1, {110}-rotation, () = 38° 56.6' (13 .3/T1,
No. 33), cell walls calculated according to Eq. (12.12-13). b) Corresponding
pattern from moire model No.1
13.3. The Periodicity of the Pattern of Lattice Points 203
20 20
10 10
- 10 10
- 20 20
a
3~1 11'111;~III'II:III~II I'III~I~'I'IIII,I~IIII'I'i~I"1'1 ' ;10
1-
-t
20
L 20
10
· 10 ~ 10
I
!
-20 20
Fig. 13.3/4a and b. Two further examples of periodic patterns on the {112}-
plane. a) Table 13.3/ T 1, No. 81 and b) No. 89
204 13. Applications of the O-Lattice Theory
18 34 ° 12.4' 0.345946 64 185 13 4 (t): (0, 0) (0, 1/8) ... (0, 7/8);
(1/8,0) ... (1/8,7/8); (2/8,0) ...
19 39° 52.2' 0.400000 2 5 3 1 (a)
20 39° 58.0' 0.467153 64 137 11 4 (f)
21 41° 6.8' 0.493151 36 73 8 3 (c)
22 43° 36.2' 0.551724 16 29 5 2 (c)
(V2/2
Rotation on {110}-plane,
U=I
S =
1/2 ~) G=e/
1/2
4 1/~)
23 10° 6.0' 0.031008 1 4 129 1 16 -7 (b)
24 11 ° 32.4' 0.040404 4 99 14 -6 (g) : (0,0) (0, 1/4) (0,2/4) (0, 3/4)
25 13° 26.6' 0.054794 4 73 12 -5 (b)
26 16° 6.0' 0.078431 4 51 10 -4 (g)
27 20° 3.0' 0.121212 4 33 8 -3 (b)
28 22° 50.4' 0.156863 8 51 14 -5 (h) : (0, 0) (0, 4/8) ; (1/4, 3/8)
(1/4, 7/8) ; (2/4, 2/8) (2/4, 6/8);
(3/4, 1/8) (3/4, 1/8)
29 26°31.6' 0.210526 4 1 19 6 1-2 (g)
l! = (
-1
1
~)
30 29° 42.0' I 0.255474 35 137 16 -5 (0,0) (1/35, 1/7) (2/35,2/7) ..
(6/35,6/7); (7/35, 0)
(8/35, 1/7) ...
31 31 ° 35.6' 0.222222 6 27 10 -3 (0, 0) (1/6, 0) (2/6, 0) ...
32 33° 43.2' 0.186916 20 107 14 -4 (0,0); (1/20,2/5) (3/20,1/5)
(5/20, 0); (2/20,4/5)
(4/20, 3/5) ...
33 38° 56.6' 0.111111 1 9 4 -1 (0,0)
34 44° 0.2' 0.052632 3 57 14 -3 (i): (0, 0) (1/3, 0) (2/3, 0)
35 45° 58.6' 0.033898 2 59 10 -2 (0,0) (1/2,0)
36 47° 41.2' 0.019608 3 153 16 -3 (0, 0) (1/3, 0) (2/3, 0)
37 50° 28.6' 0.0 <Xl (11) 6 -1
38 53° 35.6' -0.016260 2 123 14 -2 (0,0) (1/2, 0)
39 55° 52.6' -0.024390 1 41 8 -1 (0,0)
40 58° 59.6' -0.030303 1 33 10 -1 (0,0)
41 61° 1.0' -0.030928 3 97 12 -1 (0, 0) (1/3, 2/3) (2/3, 1/3)
42 62° 26.4' -0.029851 2 67 14 -1 (0,0) (1/2, 0)
43 83° 29.4' -0.028249 5 177 16 -1 (0,0) (1/5,3/5) (2/5,1/5)
(3/5, 4/5) (4/5, 2/5)
44 70° 31.6' 0 <Xl (3) 2 0
45 79° 2.6' 0.074380 9 121 12 +1 (0, 0) (1/9, 7/9) (2/9, 5/9) ... ;
(5/9,8/9) (6/9,6/9) ...
46 80° 37.8' 0.093023 4 43 10 +1 (0, 0) (1/4, 1/2) (2/4, 0)
(3/4, 1/2)
47 82° 56.8' 0.122807 7 57 8 +1 (0,0) (1/7, 5/7) (2/7,3/7)
(3/7,1/7) (4/7,6/7) (5/7,4/7)
I I I I (6/7, 2/7)
13.3. The Periodicity of the Pattern of Lattice Points 205
Rotation on {111}-plane,
U=I
u= (_~ ~)
71 (0,0) (0,2/12) (0,4/12) ...
17024.4'1-0.091603112113118121
(0,10/12); (1/2, 1/12)
(1/2,3/12) ...
206 13. Applications of the O-Lattice Theory
No.1 e
I ml I m21 Equivalence classes of O-points
(reduced O-Lattice)
u=(_~ ~)
82 52° 15.4' -0.193548 6 31 5 4 (0,0) (0, 1/3) (0,2/3); (1/2,0)
(1/2,1/3) (1/2,2/3)
83 57° 7.2' -0.114286 4 35 2 6 (0,0) (0,1/2) (1/2,0) (1/2, 1/2)
84 62° 56.0' O. CXl (33)
u= (_~ ~)
85 66° 6.0' -0.052829 1~3 ~57
86 66° 18.2' -0.056075 6 107 4 10 (0,0) (1/6, 5/6) (2/6,4/6) ...
(5/6),1/6)
87 68° 27.8' -0.088608 7 79 5 12 (0,0) (1/7,6/7) (2/7, 5/7) ...
(6/7, 1/7)
88 72° 35.6' -0.142857 11 77 5 6 (0,0) (1/11, 10/11)
(2/11,9/11) ... (10/11, 1/11)
89 78° 27.8' -0.200000 1 5 2 (0,0)
90 85° 8.4' -0.237288 14 59 2 6 (0,0) (1/14, 13/14)
(2/14,12/14) ...
(13/14,1/14)
91 87° 13.2' -0.242718 25 103 7 12 (0,0) (0, 1/5) (0,2/5) ...
(0,4/5); (1/5,0) (1/5, 1/5) ...
(1/5,4/5); ... (4/5,4/5)
(13.4-3)
(13.4-4)
with T = (1 _A-I) and:
(13.4-5)
2. Another possible way to conserve the pattern consists in the dis-
placement of the primary O-point (which is a coincidence site, and thus
coincides with a lattice point of lattice 1) into another lattice point of
lattice 1. The coincidence site thus appears at a new lattice point of
lattice 1. The pattern as a whole is conserved but is displaced by a
translation vector of lattice 1. We term the lattice of all the displace-
ments of this type as the "elementary pattern shift lattice" or DSE-
lattice, indicated by the shifts d(2-SE), d(b-SE) and d(O-SE). The trans-
lation vectors of lattice 1 are, by definition, the b(L)-vectors. Hence:
(13.4-6)
with
d(b-SE) =T b(L) (13.4-7)
and
d(2-SE) = _Ad(b-SE) = -ATb(L). (13.4-8)
208 13. Applications of the O-Lattice Theory
0) b) c)
d) (0 sn
........ ·X . ..X ..
. ................. ....... ,
(Fig. 13.4/2).
20 20
10 10
- 10 · 10
- 20 · 20
-3~" 'I' ~';I~ "'I' I:',1~ "'1" I~f~ "I"",'~"'I'" i~IIIIII ';io
a
Fig. 13.4/2a-c. Moire pattern corresponding to Fig. 13.4/1
20 20
10 10
- 10 - 10
- 20 -20
-31~ '''1' ~ ;I~ "'I' I:',1~ '''1'' ~f~'1111 ",I~'1111 II ;I~"'I' II ;10
b
13.5. Dislocation Networks in General Crystalline Interfaces 211
20 20
10 • 10
o
o -+-· ~
- 10 10
20 20
while the other O-lattice shall be that which deviates slightly from the
optimum pattern given by A':
(13·5-2)
x(O) and x' (0) are images of the same b(L)! Eliminating b(L), we obtain:
x' (0) = (T' -1T) x(O) : = Bx(O). (13·5-3)
Here the transformation B between two O-lattices is analogous to the
transformation A between two crystal lattices.
In analogy to the original O-lattice, the new O-lattice is termed the
"O-lattice of second order" or "02-1attice". In order to formulate the
new basic equation we have to know which vectors now correspond to
the original b(L)-vectors. They have to be those translations of the 0-
lattice which conserve the pattern as a whole i.e. the d(O -SC)-vectors.
Thus, the basic equation becomes
(13·5-4)
It is seen that, if A' =A then B =1, and (I-B-l) acquires rank 0, so
that here no secondary dislocation network can exist.
Before we calculate the cell walls explicitely, we have to determine
the Burgers vectors which are attributed to the cell walls. In a primary
14'
212 13. Applications of the O-Lattice Theory
O-lattice these are the differences between the b(L)-vectors, i.e. the
translation vectors of lattice 1, which are identical with the d(b-R)_
vectors. By analogy, the new Burgers vectors are chosen as the d(b-SC)_
vectors, which contain the d(b-R) (= b(L))-vectors as a subset. The
conceptual problems involved with the Burgers vectors in boundaries
(displacement of lattice 2 or lattice 1) will be discussed in Section 14.2.
Thus, considering the d(b-SC)-vectors attributed to the ;1:(02)-vectors,
we obtain, according to Eq. (13.4-2):
Td(O-SC) =d(b-SC)
(13·5-5)
which, together with Eq. (13.5-4) gives:
T (1 -B-1) ;1:(02) = Td(O -SC) =d(b-SC) (13·5-6)
with
(13·5-7)
Hence, the new basic equation is:
T(02) ;1:(02) =d(b-SC). (13·5-8)
Every vector d(O-SC) is related on the one hand to an ;1:(02), and on the
other to d(b-SC), in such a way that, to every cell wall, a Burgers vector
can be attributed according to the duality relations of Section 9.1.
The cell wall between the element ;1:(02) = [OOOJ and a neighbouring
;1:(02)-element is analogous to Eq. (12.12-13)
[d(b-SC)T GT(02)J ;1:-(1/2) [d(b-SC)T Gd(b-SC)J =0. (13.5-9)
I --+-+-+-+-< C)I I_
.,'
;/r
Ir '--Y-+-+-I- H
,/
100 1.0
(14.2-1)
(14.2-2)
(14.2-11 )
(Attention has to be paid to the sequence of 0' and 0, in order that the
sense of the loop is preserved!)
The result here is:
A -1 (~(2LF) _ ~(2LS)) = b' (L) _ b(L) . (14.2-12)
The steps in lattice 2 are equivalent to those in lattice 1, insofar as both
lattices are coupled through the transformation A.
In the case of rank (T) < 3, the following holds:
(14.2-13)
which has to be taken into account in the loop steps (14.2-11).
218 14. Completion of the Linear O-Lattice Theory
Fig. 14.2/1 b
It has been shown above that the definition of the Burgers vectors
obtained from the duality relation between the 0- and the b-lattices is
equivalent to the original definition by the Burgers loop, if translation
vectors related by A in both lattices are taken as being" identical".
The kinematic definition of the Burgers vector (movement rule,
Section 5A) can also be extended to general crystalline interfaces. For
the case of a single dislocation in a perfect crystal it was assumed that
the dislocation could be displaced by arbitrary amounts in the direc-
tion v . In the present context, we make restrictions to this assumption.
The dislocation shall be displaced by one O-lattice cell period, such
that an O-element after the displacement takes the position of a neigh-
bouring element. The total pattern before and after the displacement is
identical with respect to structure and position.
This is achieved either by keeping lattice 1 fixed and displacing
lattice 2by one of its minimum translation vectors [d(2-R) Eq. (13.2-3)
to (13 .2-5) J, or vice versa by keeping lattice 2 fixed and shifting lattice 1
by its corresponding translation vector d(l -R) = d(b -R). If m [of the
movement rule Eq. (5A-1)J points into lattice 2, then d(2-R) is the
translation vector. If the direction and size of v are left unchanged
then m can only change its sign, and thus point into crystal 1 (which
now becomes the shifted part) , if band l simultaneously change their
sign. In addition, b (i.e. d(R)) changes its orientation, and eventually its
14.2. Dislocations and Burgers Vectors in Crystalline Interfaces 219
(14.2-14)
where
d(l-R) = b' (L) _ b(L) . (14.2-15)
Thus, also here the Burgers vector is the coordinate difference of minimum
translation vectors in both lattices which are coupled through A (mini-
mum, because it was assumed that the O-element is shifted into the
neighbouring position).
After having seen that the notion of the dislocation can be extended
to general crystalline interfaces, these interfaces can also be interpreted
as "dislocation networks". (An exception is the perfect twin boundary
which is formed by an O-plane.) We distinguish on the one hand physical-
and geometrical (or mathematical) dislocations and on the other hand
primary and secondary dislocations.
We understand here as a dislocation line the intersection of the crystal
boundary with a lattice cell wall. It is a physical dislocation if the O-lattice
cell is large compared to the unit cell of the crystal lattice. In this case,
the relaxation which starts from the O-elements acts such that con-
ventional dislocations, with cores and strain fields, are formed at the
intersection of the boundary with the cell walls.
However, if the O-lattice cell is of similar size to the crystal unit cell,
then no physical dislocation can be formed any more, but, by definition,
the intersection of the boundary with a cell wall is also here a "dis-
location" with the invariance and continuity properties of the Burgers
vector (the latter follows from the existence of a connected b-net.) The
" dislocation" is now a geometrical delimitation of the different O-element
regions, and as such it is called a geometrical dislocation.
The distinction between primary and secondary dislocations can be
seen best from the behaviour of the pattern on translation of one crystal
lattice by a moving dislocation. Is the pattern preserved with respect
to structure and position, the translation is caused by a (perfect) primary
dislocation, be it physical or geometrical. If, however, the pattern is
preserved only with respect to its structure but is allowed to change its
position, then the translation is caused by a secondary dislocatioll. As
was shown in Section 13.4, the Burgers vectors of the primary dislocations
form a subset of the Burgers vectors of the secondary dislocations (the
d(b -R) correspond to the primary Burgers vectors and these are included
in the d(b-SC) of the secondary Burgers vectors).
Secondary dislocations are formed if a minimum energy pattern has
to be preserved. Each one of these patterns has its own DSC-Iattice
(Section 13.4) and thus its own set of secondary Burgers vectors. The
220 14. Completion of the Linear O-Lattice Theory
Boundary energy
Relative orientation
Type of dislocations:
of crystals
Geom. primary
, I
Phys. primary
Phys. secondary ,
5
1 ,
5
2
I--X-+--X---t
r--X~ S
, 1 '
3 ,
S
4
I--X--+-X---t
I
(X Periodic substnatum)
Perfect crystal
(14.2-18)
(14·3-1)
(14·3-2)
is determined. ;1:(0) is then given either by
;1:(0) (b(L)) = ;1:(2) (z (b(L))) (14·3-3)
or by
;1:(0) (b(L)) =;1:(1) (z(b(L)) + b(L)) . (14·3-4)
We set: (z)
X(l) =X(l) (14·3-5)
X(l) (z + biLl) = X(l) + biLl (14·3-6)
and
X(2) (z) = AX(I) (14·3 -7)
which again leads to the basic equation.
An example will now be discussed of a dislocation which ends at a
grain boundary. The boundary is assumed to be the one formed by a
rotation of 36° 52.2' in the cubic {100}-plane. The pattern is given in
Fig. 13.4/1 a with the number of pattern elements N=2 (coincidence
site and "cross"). Its period is N' = 5 crystal units . Again we let two
lattices interpenetrate, one containing an edge dislocation, the other
being perfect.
2I
xl') 0
0
./ ,
Xl21
,
xl 11
o
o
o·
o
o
Fig. 14.3/ 1. Calculated pattern for the case of a dislocation ending at a grain
boundary (original pattern Fig. 13.4/ 1 a)
14.3. Non-Linear Problems 223
a) b)
t DISLOCATION
W LATTICE 1
c) d) e)
o x(ZJ
/1
o o_xl1J
o 0 1
o
o
equation becomes:
R(~+u(~)) - [~+b(L) +U(~+b(L»)J =0. (14.3-11 )
a) b)
f DISLOCATION
47
~LATTICE 2
I" LATTICE 1
c) e)
1. Coordinate Systems
All quantitative indications of positions are referred to a coordinate
system, which is determined by an origin and three linearly independent
basic vectors u 1 , U 2 , U:J, such that every point in the space is determined
by the position vector ;x:
(A 1-1)
The vectors are indicated by small, bold face letters, the coordinates in
ordinary small letters. As is done in tensor calculations, we distinguish
between lower and upper indices and apply Einstein's summation
convention
3
x' u i : = L xi u •. (A 1-2)
.=1
If, in a term or a product, the same index appears above and below, it
represents the summation over this index. The meaning of the upper and
the lower indices will be discussed below.
In principle, any arbitrary coordinate system may be chosen. In
our case, however, two types are especially useful:
a) The first is the orthogonal coordinate system with three mutually
perpendicular basic vectors of the same unit length (of 1 A, or the length
of the cube edge in a cubic crystal system). The unit vectors are arranged
230 Appendix
IX = -t (b, c) (A 1-8)
(A 1-10)
(Ai-H)
(A 1-12)
C=AB. (A 1-13)
In index notation:
(A 1-14)
Sometimes it is convenient to denote the summing indices by Greek
letters as in Eq. (A1-14). For example the element:
(Ai-iS)
232 Appendix
i.e.
Eq. (A 1-16)
(A1-16)
This means that the scalar product of the row- or line-vector of the first
matrix and the column vector of the second matrix is the element at the
crossing point in the resultant matrix. The matrix product is usually not
commutative (An =l=BA). If the first matrix consists of only one row
(row vector), then the resulting matrix consists also of one row. Cor-
respondingly, if the second matrix is a column vector, the result is also
a column vector.
Remark
We combine here matrix- and tensor notation. In the matrix notation the
position of the elements in the matrix is determined by the sequence of the
indices and the matrix multiplication is non-commutative. In the tensor
notation, however, the coefficients form a set, where every element is iden-
tified by its indices but has no fixed position in a matrix. Hence, in the tensor
notation the indices can be placed on top of each others and the multiplica-
tion of the coefficients is communtative.
In the combined matrix-tensor notation Eq. (AI-H) should be written:
(A1-14a)
Then the transpose of C: C T = BT AT is:
(A1-14 b)
In tensor notation
(A1-14 c)
is the same as Eq. (AI-14) as only coefficients are multiplied and added
according to the summation convention and not rows an columns of matrices.
In this book, where the indices are placed vertically above each others the
upper index has to be considered as the first.
Our task now is to determine the coefficients of the S-matrix
[Eq. (A 1-11)]. If the c-axis is placed on the z-axis, then
sl=c (Ai-H)
and
(A1-i8)
A 1. Matrix Calculation Procedures 233
s~=o. (A1-19)
(A 1-20)
and the definition (A 1-8).
For example:
(u ,l U3') = a . c . cos fJ = 3 = SI
3 • S3
SI 3 . C (A 1-21)
i.e.
s~=a. cosfJ. (A 1-22)
(A 1-23)
i.e.
(A 1-24)
and
1
SI . fJ •
=a· sIn (A 1-25)
In this way, the whole S-matrix is obtained. Due to the lack of space, we
write the column vectors below, instead of beside one another.
a s~=a.sinfJ
s~=o
s~ =a· cosfJ
b s~ = (b/sin fJ) (cos y -cos ex cos fJ)
s~ = (b/sin fJ) [2 cos ex cos fJ cos y +sin2 ex sin 2 fJ
(A 1-26)
- cos2 ex cos 2 fJ - cos 2 y]1
s~=b· cos ex
c sl=o
s:=o
s:=c
The column vectors of the S-matrix are thus the unit vectors of the
crystal coordinate system expressed in orthogonal coordinates.
The basic vectors of any crystal system can be expressed in this
form by specifying the lattice constants. Summarizing: The S-matrix
234 Appendix
By using the Greek letter ~', ~'. we stress the fact that they represent
the same vector in different coordinate systems. The transition between
the coordinates is given by:
{A 1-28)
(A 1-29)
e'=8-1 e (A 1-30)
[The calculation of the inverse matrix S-l will be given in Eq. (A1-36),
(A1-37).J
It is important to know, in every case, whether one is dealing with a
coordinate transformation or with a vector transformation.
A 1. Matrix Calculation Procedures 235
,+,
3+,
"'-
"'-
~ /' / / ' / / / 6-
4-
""
'- ~ /' --'
2+ " ~ /' ~ /' /' 5-
/' '---- /' '---- /' /.
~ all an a33 + a 12 a 23 a 31 + a 13 a Z1 a 3Z
'----v----' '----v---'
(1) (2) (3)
- a13 aZZ all - a1z aZl all - all 0Zl an
'---v------' '---v----' ~.~~~
(A 1-33)
(A 1-34)
In order to calculate the element a~31) of the inverse matrix, the sub-
determinant of the element a32 (the mirror image across the main dia-
gonal) is calculated and divided by the determinant of A.
-1
"'. . .
( •" ' ., a23(-1) )
. .X., (A1-37)
"
1= (°10 0°1°0) .
1 (A 1-39)
'*'
time an inverse matrix has been determined.
An inverse matrix only exists if the determinant IAI 0, as shown
by Eq. (A1-36).
By means of the inverse matrix, for example, Eq. (A1-29) may be
solved for ;r:
;r' =S;r. (A 1-40)
Both sides are multiplied from the left by S-I:
S-I;r' = S-1 S;r =I;r =;r (A1-41)
which is the same as Eq. (A 1-31).
Another application of the inverse matrix is the determination of the
basic vectors of the reciprocal lattice U'i, expressed in orthogonal co-
A 1. Matrix Calculation Procedures 237
ordinates. On the one hand the U'i and u; are related by Eq. (A 1-5):
(A 1-42)
(u; = general unit vectors, u i = orthogonal unit vectors). On the other
hand:
(A1-43)
According to the matrix multiplication convention a row vector of the
first matrix is multiplied as a scalar product with a column vector of the
second matrix. Here, the column vectors of 8 are the vectors u~, hence
the row vectors of 8-1 must be the U'i. Thus, the vector products of
Eq. (A1-3) can be replaced by the matrix inversion in order to deter-
mine the vectors u' i .
5. Metrics
The metric tensor is defined in the following way:
G =gik = (u; u~). (A 1-44)
The terms of G are thus the scalar products of the basic vectors. Since
the scalar products are commutative, the metric tensor is always sym-
metrical.
(A1-45)
In the orthogonal system, G is, according to Eq. (A 1-20), given by
G =gik = (Ui Uk) =~ik =1. (A 1-46)
G can be determined from the 8-matrix by:
G=8 T 8 (A 1-47)
with 8T the transposed matrix of 8. (The transposed matrix is the mirror
image with respect to the main diagonal.)
(A 1-48)
The metric tensor is a characteristic of the coordinate system. It helps to
determine metric properties such as absolute length and angles in general
coordinate systems. The scalar product of two vectors is:
I;}:I 'Iyl cos ¢ =;}:T Gy = X· gik yk (A 1-49)
and, as a special case, the square of the absolute length is
1;}:12 =;}:T G;}: = Xi gik Xk (A 1-50)
and the angle ¢ between the vectors;}: and y is given by:
xTGy xi gik yk
cos ¢ = ~ = [(xi gik xk)'-'-;(~y'o--·g-ik-y---'k-:-:)]"! . (Ai-51)
238 Appendix
a
Expansion E = ( 0 b 0 .
0 0) (A 1-59)
o 0 c
I
b) A convenient representation of a "otation in the right-handed
screw sense by the angle () around an axis given by the unit vector
[Ct, clI ' caJ (c. = direction cosines) is the following:
c.,,= ( ~3 -; -::).
(A 1-60)
-Cll '1. 0
In this way, the vector c required for the rotation Eq. (A 1-60) is deter-
mined. Further: cos 0 = va and sin 0 = (1 - v=)§, so that now the data
needed for the formulation of the rotation matrix are all known.
c) A pure shear in the direction of the y-axis, with the (xy)-plane
invariant, is described by:
Shear C = (1 0 0)
0 1 (J
001
• (A 1-64)
8. Transformation of Transformations
It may occur, for example, that a pure shear has to be represented
in such a way that the orientation of the shear direction and the shear
plane are given in space. We may start from a special representation
of the type of Eq. (A 1-64), and then rotate the whole transfonnation.
A 1. Matrix Calculation Procedures 241
9. Addition of Transformations
The addition of two transformations A +B means the addition of
transformed vectors 00' originating from the same vector 00
Coo=(A+B) 00 =Aoo+Boo. (A 1-74)
The matrices are added element by element
C'k =a'k +bik · (A1-75)
10. Invariant Elements of a Transformation
By a non-degenerate linear transformation, the vector space as a
whole is transformed into itself. In the case of a homogeneous trans-
formation the origin stays unchanged oo[OOOJ -+ 00' [OOOJ. We may ask
whether further unchanged elements exist. Vectors with invariant
orientation (while their length may change), are called eigenvectors. They
are determined by the following condition:
(A 1-76)
The numerical parameter A is called the eigenvalue. Its meaning is the
ratio of the length of the eigenvector after and before the transformation
~ (A 1-77)
A= 1:Ili .
The equation for the eigenvectors is:
(A-U)oo=O. (A 1-7S)
where 0 means the zero vector [000]. This equation has non-trivial
solutions only when:
IA-UI=O. (A 1-79)
Eq. (A 1-79) is an equation of the third degree for A. The (real) solutions A
are introduced into Eq. (A1-7S), from which the eigenvectors 00 can be
calculated.
If we refer to the end vector 00' instead of to the starting vector 00,
and denote this end vector by 00(0), Eq. (A 1-7S) becomes:
(l-AA-I) 00(0) =0. (Ai-SO)
A2. Moire-Models 243
A2. Moire-Models
Four pairs of moire models and a scale M are joined to the book.
The model 1 corresponds to a {11 0}-face in the body-centred cubic (bcc)
structure. The point arrangement is approximately hexagonal. The
length of the cube edge is chosen as 2 mm. Model 2 represents a {11 2}-
face and model 3 a {1 OO}-face in the same structure. Model 4 consists
of parallel lines also with 2 mm spacing. Fig. A2/1 shows the dimensional
situation.
Every model consists of a sheet a and a sheet b. Sheet a may be
placed below the scale and its position fixed by means of the straight
scales. The zero position on all four linear scales corresponds to the
exact centering. Sheet b can be placed on top of the scales, and its
position is determined with respect to rotation. As moire figures are
extremely sensitive to changes in relative position, a high precision is
needed, especially if calculated patterns are to be compared quantitatively
with the models.
It was shown in Section 12.10 that, with respect to translation a
shift d(S) is equivalent to one by d(l) ( = d(b») , where d(l) = - A -1 d(lI) •
Instead of executing the rotation and the translation with sheet b, the
translation can be done with sheet a in coordinates of crystal 1 and the
rotation alone with sheet b. The angle of rotation is best fixed with all
four marks simultaneously on the circular segments. The coordinate
244
.-.-.-....
••••••••••- •••••
•••••
Appendix
-1.·.·.·.·.·
-·~r·!·!·! .....
•••••••••• •••••
••••• 1312· • • • •
-
I fil
•••••••
•••••••
- .......
•••••••
-1• • • • • • •
11 I
Fig. A2/1. Dimensions of the moire models (the unit corresponds to 2 milli-
meters)
origin in this case is the centre of the scale sheet, i.e. also the centre
of sheet b.
The upper left-hand comer of the given lattices are cut such that the
sheets can always be placed on top of each other in the same orientation.
In this way eventual irregularities in the drawing or distortions due to
the photographic reproduction may cancel out optimally.
In order to add an expansion to the rotation, the aspect of perspective
can be applied by keeping the sheets apart when viewing them. The
movement rule [Eq. (5.4-1)] for edge dislocations can be checked in this
way with model 4. Expansion combined with rotation and translation
can also be controlled quantitatively by mounting the sheets on glass
plates and photographing them above a luminous table by means of a
reflex camera.
The models are transpcuent in view of their possIble use as demon-
stration material in lectures, where they can be projected. Here a view-
graph is recommended, in which the sheets M and a can be placed
under- and the sheet b on top of the transparent plastic foil used for
writing. In this way the displacements of sheet b do not affect the lower
sheets. An epidiascope should not be used for projection, as the models
are likely to be burnt.
A3. The Direct Observation of Crystal Defects 245
Special References
Section:
2.2-2.3 International tables for X-ray crystallography. Birmingham, Eng-
land: K ynoch Press 1962.
4.1 Billington, D. S.: Radiation damage in ·solids. Proceedings of the
Internat. School of Physics. "Enrico Fermi" XVIII Course,
Italian Physical Society. New York: Academic Press 1962.
- Crawford, J. H., Jr.: Radiation damage in solids. New Jersey:
Princeton University Press 1961.
248 References