You are on page 1of 172

Chapter 1

Introduction

The motion of atoms in solids is restricted to small vibrations about a mean position and for most
purposes the nucleus can be considered as being at rest. As the temperature is raised, the atoms in
a solid become increasingly more mobile until their motion becomes so great that the orderly array
in the solid disintegrates and it becomes a liquid. In the liquid state the atoms are free to move
about and are in constant motion. However, they are attracted to each other and thereby give the
liquid its cohesiveness. If the process is carried far enough, the liquid can be transformed into a
gas, which is nothing more than a state in which the atoms have gained so much freedom of
motion that they no longer bother to stick together at all.
When the temperature is lowered sufficiently, a liquid freezes and a solid results. When the liquid
cools, two types of solids can form. If the liquid cools sufficiently slowly, the atoms can assume an
orderly arrangement and a crystal results. On the other hand, if the temperature is dropped
abruptly, arresting the motion of the atoms before they can reorganise themselves, then a mixed-up
structure called a glass or an amorphous solid may result. Such a glass, when reheated, does not
melt abruptly; it softens gradually and continuously until it is so soft that it is again a liquid. As
with most properties of solids, the explanation of the difference in the behavior of glasses and
crystals when heated is determined by the difference in their atomic arrangements or structures. In
a crystal, the atoms are arranged like soldiers on a parade ground, in well-defined columns and
rows. In a glassy substance, the atoms never had a chance to organise themselves into orderly
arrays. Hence in a glass it is not possible to distinguish the solid from the liquid state except by the
relative motions of the atoms. This property described by the term viscosity.
The transition between the liquid and glassy states on heating or cooling is a gradual one, whereas
the more fundamental transition between the liquid and crystalline states is more abrupt. Crystals
can form from a gas by sublimation or from a liquid by crystallisation. When a crystal forms from
a vapour or a liquid its continued growth depends on the presence of a sufficient supply of growth
units, that is, the atoms required to makeup its structure. Crystals may form and grow in a
noncrystalline solid if the solid is held at a temperature high enough to permit atoms to interchange
positions. This is the process of devitrification of a glass. A crystalline solid does not deform in a
viscous manner; there is practically no permanent deformation if the applied stresses are less than
a certain critical value, and at higher stresses the rates of plastic deformation are very different
from those of the liquid or glassy states.

Amorphous versus Crystalline Structure


An amorphous substance usually is isotropic, i.e., it exhibits the same physical and chemical
properties when tested in any direction (provided it is homogeneous and free from internal strains),
but a crystal has many properties that are directional. Directionality (anisotropy) in a crystal is
found to a greater or lesser degree, for example, in electrical and thermal conductivity, thermal
expansion, elastic constants, optical constants, and chemical reactivity of exposed surfaces.

Fig. 1.1: Single Crystal of Ice

The basic feature of all crystals is the regularity of their atomic arrangement. The first notable
feature of the regularity of crystal structures is the periodicity of their patterns. In addition to
periodicity, most crystal structures possess the property of symmetry. There are several kinds of
symmetries possible. A frequently encountered one is the symmetry of rotation. For example, a
single crystal of ice (Fig. 1.1) can be rotated about its center six successive times without repeating
the same movement (the angle of rotation each time is 60o), and each time it is rotated, the crystal
looks the same as it did before rotation. Other symmetries are also possible and, of course,
combinations of symmetries are possible. It might appear from this that a large number of different
kinds of structures are available to accommodate the extremely large number of different kinds of
solids known to exist. Actually, however, it has been amply shown that only 230 different
combinations can be formed in three dimensions.
When a solid has a crystalline structure, the atoms are arranged in repeating structures called unit
cells. The cells form a larger three dimensional array called a lattice. The unit cells can be of
different types. They each have a different name. If many crystals are growing in a melt at the
same time, the crystals will eventually meet and form grains. The junctions of the crystallites are
called grain boundaries.
The crystal structure of a material depends on the nature of the bonding. We shall, therefore, begin
with a brief discussion on the major types of bonds that exist in solid materials.

Ionic Bonding
An ionic bond is formed when one or more electrons, from the outermost shell of an atom, are
transferred to the outermost shell of another atom. In this process both the atoms acquire the
electronic configuration of a noble gas. The atom which loses the electron acquires a positive
charge and becomes a positive ion. The atom which gains the electron acquires a negative charge
and becomes a negative ion. In this way both the atoms become oppositely charged ions. The
electrostatic attraction between the oppositely charged ions forms the ionic bond.

Fig. 1.2: Section of Ionic Crystal

The bond occurs between a metal and a nonmetal. The outer shell in a nonmetallic element
contains a larger number of electrons than the outer shell in a metal. For example, the outer shell
of chlorine contains seven electrons, and there is a strong driving force to attract an electron to
make a stable group of eight. When sodium (one valence electron) reacts with chlorine, the sodium
gives up its electron to form the stable outer shell of chlorine. An aggregate of sodium and
chlorine atoms, therefore, quickly becomes an aggregate of positive sodium ions and negative
chlorine ions. The energy of such an aggregate is decreased (as a result of the electrostatic
attraction between opposite charges) if the nearest neighbours of each ion are ions of the opposite
sign. A typical arrangement of ions is shown, in two dimensions, in Fig. 1.2. Each ion has, in the
plane of the diagram, four nearest neighbours of opposite sign and is surrounded, at a greater
distance, by a ring of four ions of its own sign.
The array of atoms is in fact three dimensional being composed of layers similar to that of Fig. 1.1
and superimposed so that a line normal to the plane of the diagram passes through positive and
negative ions respectively. Fig. 1.2, therefore, represents a section of the array in any one of the
three mutually perpendicular planes. Each ion has 6 nearest neighbours of opposite sign and 12
next nearest neighbours of similar sign. Because the distance between unlike ions is less than the
distance between like ions, the electrostatic forces of attraction between the former are greater than
those of repulsion between the latter. An additional force, however, must be taken into
consideration; this is the electrostatic repulsion that occurs between the outer electron shells of the
two ions if they approach each other, even if the ions are of opposite sign. The forces of attraction
and repulsion just balance when the ions are at a definite distance apart. No positive ion in this
arrangement is uniquely associated with an individual negative ion, and it follows that the concept
of molecule does not apply in this type of structure, which constitutes a crystal.

Fig. 1.3: Covalent Bonding

Covalent or Homopolar Bonding


Elements from the central groups of the Periodic table (notably Gr IV) are not readily reduced to a
closed-shell electronic configuration, because the energy required to remove all the valence
electrons is too large. lonic bonding is, therefore, unlikely. However, it is still possible for each
atom to effectively complete its outer electron shell by sharing electrons with its neighbours. For
example, an atom of carbon has four electrons in the outer shell and four more electrons are
required to fill this shell. This can be done by sharing with four other carbon atoms, forming a very
stable configuration having the shape of a tetrahedron with the carbon atoms at the four apices.
This configuration can be extended indefinitely as a three dimensional patter, the diamond cubic
crystal structure. Covalent bonding also occurs between dissimilar atoms. Electron sharing is only
possible if the atom are so close to each other that the electrons can pass from an orbit of one atom
to an orbit of the other without becoming completely detached from either.
This electron sharing or covalent bond is the strongest of the chemical bonds. Solids so bonded are
characterised by general insolubility, great stability and very high melting and boiling points. They
do not yield ions to such solutions as they form and hence are nonconductors of electricity both in
the solid state and in solution. Because the electrical forces constituting the bond are highly
localised in the vicinity of the shared electrons the bond is highly directional and the symmetry of
the resulting crystals is likely to be lower than where ionic bonding occurs.

Metallic Bonding
Metallic bonding is an example of even unrestricted sharing of electrons. We know that metals are
very good conductors of heat and electricity, because the electrons are free to move throughout the
metal. Such freedom of motion is possible because in metallic bonding there is no way of
determining to which nucleus any particular valence electron belongs. A metal may be pictured as
a geometric array of nuclei and inner electrons (with net positive charges) and a cloud of valence
electrons (with negative charges). Fig. 1.4 shows this configuration schematically. We must bear
in mind that only the valence electrons are free to move and form the electron cloud; the electrons
of the inner shells are more strongly bound to the nuclei.

Nucleus plus Electron cloud


inner electrons
Fig. 1.4: Metallic Bonding

A common characteristic of the metallic elements is that they contain only one, two or three
electrons in the outer shell, these electrons are bonded loosely to the nucleus. Therefore when we
put a number of magnesium atom, the outer electrons leave individual atoms to enter a common
electron gas. The atoms are therefore changed to Mg2+ ions. These repeal each other but are still
held together in the block because the negative electrons are attracted to the positive ions. The end
result is that the Mg2+ ions arrange themselves in the regular pattern shown. This is the
arrangement in one plane cut through the magnesium crystal. It represents the closest possible
packing for a group of equally sized spheres. This very simple model can explain the properties
(conductivity, ductility, etc) of metals.

Types of Crystals
Crystals may be classified according to the types of bonding forces as follows:

1. Metallic Crystals
These crystals consist of positive ions immersed in a gas of negative electrons. The attraction
of the positive ions for the negative electrons holds the structure together and balances the
repulsive forces of the ions for one another and of the electrons for the other electrons. The
electrons move freely through the lattice and provide good electrical and thermal conductivity.

2. Ionic Crystals
These crystals are bound together by the electrostatic attraction between positive and negative
ions. They are combinations of strongly electronegative and electropositive elements. In NaCl,
for example, the electron affinity for chlorine atoms causes a transfer of electrons from the
electropositive sodium atoms to yield Na+ and Cl- ions.

3. Covalent Crystals
These are also known as homopolar or valence crystals. They are held together by the sharing
of electrons between the neighbouring atoms. Diamond is a typical example, in which each
carbon atom shares its four valence electrons with the four nearest neighbours and thus
completes an outer shell of eight electrons in each atom. These crystals are charecterised by
poor conductivity and great hardness.

4. Molecular Crystals
These are composed of inactive atoms or neutral molecules bound by weak van der Waals
forces. They have low melting points.
Many crystals are intermediate between these ideal types. For example, it is possible for some
interatomic bonds in a crystal to be of one type, say ionic, and others to be of another type, say
covalent; or each bond may be somewhat intermediate in type between some of those listed.

Relationship between Structure and the Position of an Element in the Periodic Table
There is a clear relationship (Table 1.1) between structure and the position of an element in the
periodic table. The classical table is arranged in the form of horizontal rows (or periods) and
vertical groups. Elements in the same group tend to have the same structure at ordinary (room)
temperature; for example, the alkali metals Li, Na, K, and Cs (Group IA) are all b.c.c.; Be, Mg,
Zn, and Cd (Group IIA and IIB) are c. p. h.; Cu, Ag, and Au (Group IB) are f. c. c. Elements to the
left in the Periodic Table show electropositive behaviour and tend to have the simple metallic
structures. Elements to the right are sometimes called the metalloids and tend to have more
complex structures. In the intermediate groups (the transition elements and the rare earths), the
process of filling inner shells while the number of electrons in the outer shells remains constant
tends to encourage the typically metallic structures. However, although these structural
relationships clearly exists in the periodic table, their interpretation in terms of detailed bonding
and band structure is still chiefly empirical.

The 8 – N Rule
The B sub-group elements, in which the outer shells are in the process being filled, are in general
less closely packed. The binding forces become more covalent between certain neighbouring
atoms, instead of a binding among positive ions and electrons. The homopolar character increases
toward the end of the horizontal periods until it is a true covalent bond in the diatomic molecules
of Group VII B, the halogens F2, Cl2, Br2 and I2.
Most of the structures in the B subgroup follow the 8 – N rule which states that each atom has 8
minus N close neighbours, where N is the number of group in the Periodic Table to which the
element belongs. For example, in carbon (Group IV B) with the diamond structure, each atom has
four neighbours. Gray tin, silicon and germanium, also of Group IV B, likewise have this diamond
cubic structure with co-ordination number 4. The heavier elements of this group tend to have more
metallic structures, while tin is b. c. t. and lead is f. c. c.
In group VB, arsenic, antimony, and bismuth have rhombohedral structures, with each atom
surrounded by three nearest neighbours. These more closely bonded atoms form layers, within the
layers the bonding is covalent, between layers it is partly metallic. Group VIB has Selenium and
Tellurium with the trigonal structures, in which the atoms are linked in chains each atom having
two neighbours. There is only one near neighbour to each atom in solid iodine, of group VIIB, for
which 8 – N = 1. The pairs of iodine atoms correspond to the diatomic molecules characteristic of
other elements of this group. The structure is orthorhombic.
The 8 – N rule does not apply to elements of group III B. For example, Gallium has orthorhombic
structure with eight atoms per unit cell and four different inter atomic distances between adjacent
atoms. Indium has a structure in which all atoms have four near neighbours: it is a tetragonal
structure that may be derived from the f. c. c. by elongating a cube edge until the axial ratio is
1.08. The failure of these elements to have a coordination number of 5 is attributed by Hume-
Rothery to incomplete ionization of the atoms.
PROBLEMS AND QUESTIONS
1. Distinguish between an amorphous and a crystalline substance. List the basic features of a
crystal.
2. Explain the nature of glassy state. How would you distinguish the solid state from the
liquid state in a glass ?
3. What is the effect of rate of cooling on the structure of the solid that results?
4. Distinguish between isotropy and anisotropy.
5. Explain the ionic and covalent bonding.
6. List the different types of bonds that occur in crystals. Discuss the important
characteristics of each type.
7. List the different types of bonds that occur in crystals. Describe the characteristics of
metallic bond.
8. Explain what is meant by covalent and metallic bonding. How do they arise? What are
the differences and similarities between those two bonds in solids?
9. Describe with examples the different types of bonds that exist between atoms in a solid
and explain how they arise.
10. Explain why a covalent bond is directional while ionic and metallic bonds are non-
directional.
11. How are atoms held in a metallic bond? Explain diagrammatically.
12. Describe the nature of metallic and compare it with ionic and covalent bonds.
13. Explain the relationship between the structure and the position of an element in the
periodic table.
14. Explain with suitable examples the 8 – N rule. Are there any exceptions to this rule?
Chapter 2
Structure of Metals

Metals are aggregates of atoms. The atomic bonding in this group of materials is metallic, and thus
nondirectional in nature. Consequently, there are no restrictions as to the number and position of
nearest neighbour atoms; this leads to relatively large number of nearest neighbours, and dense
atomic packing for most metallic crystal structures.
In metallic materials, the valence electrons lie much farther from the nucleus than do those in
nonmetallic materials. Thus the wide orbit of the valence electrons allows them to pass to regions
remote from their parent atoms. Likewise, they never associate for any appreciable period of time
with any one atom in the metal, but drift through the whole assembly as a kind of electron gas.
Only the valence electrons behave in this manner, for the electrons in the inner shells lie much
closer to the nucleus and are unable to move outside its field of influence. Consequently, as far as
the inner electrons are concerned, an atom of a metal is much the same as an isolated atom,
because these electrons are held firmly within the atom and remain in their characteristic quantum
shells.
Thus the internal structure of a metal is composed of two parts:
(a) a collection of positive ions, each consisting of the core of an atom, i.e., the nucleus and
the non-valence electrons
(b) a gas of free electrons swarming between the ions throughout the whole of the metal. The
binding force holding the metal together is provided by the attraction of the positive ions
to the negative valence electrons continually passing between them.
Introduction to Crystallography

When atoms of a metal approach each another, two opposing forces influence the internal energy.
(1) an attractive force between the electrons and the positive nuclei and
(2) a repulsive force between the positive nuclei and also between the electrons.

Repulsive Energy

Interatomic separation

Net Energy
Eo

Attractive Energy

Fig. 2.1: Potential Energy as a Function of Distance Between two Atoms

The first force tends to decrease the internal energy and the second force tends to increase it. At
some distance these two forces just balance each other and the total internal energy EO will be
minimum, corresponding to an equilibrium condition (Fig. 2.1). This equilibrium distance r O is
different for each element and is determined by measuring the distance of closest approach of
atoms in the solid state. It the atoms are visualized as spheres just touching each other at
equilibrium, then the distance between the centers of the spheres may be taken as the approximate
atomic diameter. The atomic diameter increases as the number of occupied shells increases and
decreases as the number of valence electrons increases (Table 2.1).

Table 2.1(a): Atomic Diameter and Atomic Number of Elements

Element Atomic Number Atomic Diameter (Å)

Lithium 3 3.03

Sodium 11 3.71

Potassium 19 4.62

Rubidium 37 4.87

Cesium 55 5.24

12
Structure of Metals

Table 2.1 (b): Atomic Diameter and Valence of the Elements

Element Atomic Number Atomic Diameter (Å)

Potassium 1 4.62

Calcium 2 3.93

Scandium 3 3.20

Titanium 4 2.91

Vanadium 5 2.63

Since atoms tend to assume relatively fixed positions, this gives rise to the formation of lattice
structure in the solid state. However, the atoms oscillate about fixed locations and are in dynamic
equilibrium rather than statically fixed. As already mentioned, when a metal freezes, its
component atoms assume an ordered arrangement relative to each other. In other words, there is a
regularity of atomic arrangement, an internal symmetry, i.e. metals are crystalline.
With certain very minor exceptions, which will be discussed in later chapters, all solid metals are
typically and completely crystalline. This is strongly reflected in their properties and behavior. As
a result, it is necessary that certain of the principles and many of the traditional terms and methods
of crystallography be fully understood.

Space Lattice in General


n any discussion of crystal structures the concept of space lattice is very useful. It affords a
convenient means both for describing the arrangement of atoms or molecules within a given
crystal and for comparing these arrangements among crystals of different types or in different
locations.
The essential characteristic of a crystal is the periodic nature of its internal structure. As a
consequence the atomic arrangement in a crystal may be related to a network of points in space.
This network is called a space lattice or simply a lattice. Within this lattice the surrounding of each
point is identical with the surrounding of each of the other points. In other words, each individual
point has the same number of other points as its nearest neighbors. The number, spacing and
angular relations of these nearest neighbors are the same, no matter where in the space lattice the
reference point has been taken. For example, if the simple cubic space lattice (Fig. 2.1) is
continued indefinitely in space every point in space (represented by dots) will have exactly six
nearest neighbors among the other points present. All of these nearest neighbours will be equally
distant from their own nearest neighbours, including the original reference point, and each will be
along one of a set of three rectangular reference axes originating at the reference point. This

13
Introduction to Crystallography

arrangement of points fulfills exactly the requirements stated above for a space lattice and it is in
fact the simplest three-dimensional arrangement of points that does so. In other words, if we could
move around in the lattice, we would not be able to distinguish one point from the other, for rows
and planes near each point would be identical. If we were to wander among the atoms of a solid
metal we would find the view from any lattice point exactly the same as that from any other.

Unit Cell

Lattice Points

Fig. 2.2: Eight Unit Cells of a Simple Cubic lattice

To specify a given arrangement of points in a space lattice, it is customary to identify a unit cell
with a set of coordinate axes, chosen to have an origin at one of the lattice points. In a cubic lattice,
for example, we choose three axes of equal length that are mutually perpendicular and form three
edges of a cube. Each unit cell in a space lattice is identical in size, shape and orientation to every
other unit cell.
The reference axes mayor may not be at right angles and may or may not be of equal length,
depending on the crystal considered. If all combinations of equality and inequality in inter-axial
angles and in lengths are counted, it is found that a total of only 14 kinds of lattices are possible.
This means that no more than 14 ways can be found in which points could be arranged in space so
that each point has identical surroundings. All crystal structures are based on these fourteen
arrangements.
Since the first correct formulation of the lattice concept was by Bravais they are known as the
Bravais Lattices. To describe the fourteen space lattices and the actual crystal structures derived
from them, seven different system of reference axis are required. These differ from each other
either in the angular relations among the axes or in the relative magnitudes of the units of length
used of measurement along these axes or both (Table 2.2).

14
Structure of Metals

Unit Cells and Primitive Cells


The smallest unit having the full symmetry of the crystal is called the unit cell. It is the building
block from which crystal is constructed by repetition on three dimensions. The specific unit cell
for each of the metals is defined by its parameters (Fig. 1.3) which are the edges of the unit cell a,
b, c and the angles α: (between b and c), β (between c and a) and γ (between and b).
In the literature, we often find reference to unit cells and to primitive cells. The primitive cell may
be defined as a geometrical shape, which when' repeated indefinitely in three dimensions will fill
all space and is equivalent of one atom. The unit cell differs from the primitive cell in that it is not
restricted to being the equivalent of one atom. In some cases, the two coincide, for instance, the
simple cubic cell and simple tetragonal cell are also primitive cells.
In the study of crystals, the primitive cell has limited use, because he unit cell can usually be
visualized readily where as he primitive cell cannot.

Table 2.2: Bravais Lattices and Unit Cells

Crystal Bravais lattice Geometry of cells


system Lengths Angles.
Triclinic Simple a≠b≠c α≠β ≠γ
Monoclinic Simple a≠b≠c α = β = 90o ≠ γ
Base -centered
Orthorhombic Simple a≠b≠c α = β = γ = 90o
Base -centered
Body -centered
Face -centered
Tetragonal Simple a=b≠c α = β = γ = 90o
Body -centered
Cubic Simple a=b=c α = β = γ = 90o
Body -centered
Face -centered
Trigonal Simple a=b=c α = β = γ ≠ 90o
a=b≠c α = β = 90o
Hexagonal Simple
γ = 120o

15
Introduction to Crystallography

Common Unit Cells


It is indeed very fortunate that most metals crystallise in one of the 3 basic unit cells, the body
centered cubic, the face centered cubic and the hexagonal close packed. Other materials crystallise
in these as well as the remaining 11 unit cells. We shall confine ourselves to the study of these 3
cells. However, the analysis used here can then be applied generally to the other Bravais lattices.
Y

γ a
α X
c β
Z

Fig. 2.3: Space lattice Illustrating lattice Parameters

The Body-centered Cubic Unit Cell


The body-centered cubic (bcc) unit cell is shown in Fig. 2.4. This has one atom located at each of
the corners of the cube and one atom at the body center.
Y

X
Z

Fig. 2.4: Unit Cell of the Body Centred Cubic Structure


(atoms represented by closed circles)
However there are only two atoms in a body centered cubic cell. This is because each comer atom
is shared by a total of eight adjacent unit cells and can contribute only 1/8 atom per cell. The atom

16
Structure of Metals

at the center is entirety within the cell. Thus, the total number of atoms in a body centered cubic
cell is = 1/8 x 8 (corner atoms) + 1 (body center atom) = 2.
A study of Fig. 2.4 shows that the atom at the center of the cube is collinear with each comer atom.
that is, the atoms connecting diagonally opposite corners of the cube form straight lines, each atom
touching the next in sequence. These linear arrays do not end at the corners of the unit cell, but
continue on through the crystal. The four cube diagonals constitute the close packed directions of
the body centered cubic crystal, directions that run continuously through the lattice on which the
atoms are as close spaced as possible. Further consideration of the Fig. 2.4 reveals that all atoms in
the body centered cubic lattice are equivalent. Thus the atoms at the corners of the cube of Fig. 2.4
have no special significance over those occupying center positions. Each of the latter could have
been chosen as the center of a unit cell, making all comer atoms centers of cells and all center
atoms representing the corners of the cells.

BCC Nearest Neighbors


The atoms which are closest to the reference atom are called nearest neighbors and is given by the
co-ordination number Z. In body centered cubic unit cell, the body centre atom is surrounded by
eight equidistant neighbours. These are located at the comer of the cubes. That these comer atoms
are at the shortest distance can be shown in the following way.

a
a
a
a 2 a
a

a 3

Fig. 2.5: Body Centred Cubic Nearest Neighbours

Let the sides of the unit cell (in Fig. 2.5) be of length equal to a. The length of the body diagonal is
then a 3 . The distance between any two atoms on the cube face is either a or a 2 as indicated in
the figure. Since both of these distances are greater than the distance from the body centered atom
to the comer, the nearest neighbor distance is a 3 /2. There are eight such atoms at the eight
corners of a bee unit cell. We have already seen that all atoms in this lattice are equivalent.

17
Introduction to Crystallography

Therefore, every atom of the body centered cubic structure not lying at the extension surface
possesses eight nearest neighbours and the coordination number of the lattice is eight.

Lattice Parameter of BCC Unit Cell


If the atoms are considered to be solid spheres, 'hen the size of the unit cell may be reduced to the
point where the nearest neighbor atoms just touch one another. This contact exists only for nearest
neighbors and not all atoms touch other atoms. This means that the center-to center distance
between two atoms touching one another along the body diagonal (Fig. 2.5) would-be a 3 /2. If
the atomic diameter is d, then the relation between d and the lattice parameter a is given by d =
a 3 /2 or a = 2d/ 3

BCC Density of Packing


The density of packing is defined as that fraction of the total volume of the unit cell occupied by
the nuclei and inner electron which comprise the spheres. The total volume V of the unit cell is a 3 .
The volume v of each atom is
v = 4/3 π r3 = 4/3 π (d/2)3 or v = π d3/6
and since there are two atoms per unit cell, the total occupied volume v is
d 3 d 3
v= x2 or v=
6 3
The density of packing p is then,
πd 3 /3 πd 3
ρ= or ρ=
a3 3a 3
Substituting the value of d = a 3 /2, we obtain

π (a 3/2) 3
ρ=
3a 3
or ρ = 0.681
Therefore, only 68.1 per cent of the volume of the body centered cubic unit cell is occupied by the
nuclei and the inner electrons of the atoms. The valence electrons form an electron gas that fills the
remainder of the volume.

The Face-Centered Cubic UnitCell


The face centered cubic unit cell is illustrated in Fig. 2.6 There are four atoms in a face centered
cubic unit cell. Eight atoms at the comers contribute 1/8 each and six at the center of each face that
contribute ½ each. Therefore, there are 8 x 1/8 + 6 x ½ = 4 atoms per unit cell.

18
Structure of Metals

Lattice Parameter of FCC Unit Cell


Each comer atom within the face centered cubic unit cell has as its nearest neighbors the atoms at
the centers of the adjacent face. There' are twelve such neighbors, the nearest neighbor distance
being a. Thus d = a 2 /2 or a = 2 d.

a 2 /2

a a 2
a
(a) (b)

Fig. 2.6: Face Centred Cubic Unit Cell and Atomic diameter

FCC density of Packing


There are four atoms per unit cell whose volume is V = a3. The atomic diameter is given by d =
a 2 /2. which is apparent from Fig. 2.6. (b). The occupied volume is then

(a 2 /4) 3 a3 2
v = π x4 or v = π
6 6
The density of packing

π 2
ρ= = 0.742
6
Thus 74.2 per cent of the volume is occupied in the FCC unit cell as compared to only 68.1 per
cent for the BCC.

Close Packed Hexagonal Unit Cell


The usual picture of the close packed hexagonal lattice shows two basal planes in the form of
regular hexagons with an atom at each corner of the hexagon and one atom at the center. In
addition, there are three atoms in the form of a triangle midway between the two basal planes. If
the basal plane is divided into six equilateral triangles the additional three atoms are nestled in the
center of alternate equilateral triangles.

19
Introduction to Crystallography

The hexagonal unit cell is shown in Fig. 2.7(a). A study of the drawing will reveal that the atom
positions in this structure do not constitute a space lattice, since the surroundings of the interior
atoms are not identical with those of the atoms at the cell corners. The true cell of the hexagonal
lattice is in fact only the portion re-drawn in Fig. 2.7(b). The hexagonal close packed cell may be
generated from these true unit cells shown in Fig. 2.7(b) by translation only. The hexagonal prism,
therefore, contains two whole unit cells and two halves.

a
a a

a
a
a
1200
(a) (b)
Fig. 2.7: (a) A Close Packed Hexagonal Crystal (b) A True Hexagonal Unit Cell

CPH Nearest Neighbours


The number of nearest neighbours of an atom in close packed hexagonal structures is 12 and this
fact may be verified with the aid of Fig. 2.7 (a). The central atom in the basal plane has six nearest
neighbours in that plane at the comers of the hexagon. It also has three nearest neighbours in the
plane half-way up the cell and three in the plane below. This gives a total of 12 nearest neighbours.
Since this argument is valid no matter whether the atoms in the close packed planes just above or
below the reference atom are in B or C positions, it bolds for both face centered cubic and close
paced hexagonal stacking sequences. We conclude, therefore, that the co-ordination number in
these lattices is 12.

Ideal c/a Ratio


Because of the close-packing of planes in the hcp structure, the lattice parameter c is not
independent of a in the ideal case. The value of c/a, calculated on the basis of close-packed planes
arranged in an ABA sequence is called an ideal c/a ratio.
To calculate the ideal value of c, it is convenient to imagine that three atoms of the basal plane and
one atom of the mid-plane form a tetrahedron [Fig. 2.8 (a)], in which all the atoms are in contact.

20
Structure of Metals

This tetrahedron is redrawn in a more open fashion [Fig. 2.8 (b)], the length of each side of the
equilateral triangle ABC forming the base of the tetrahedron is equal to a. Each face of the
tetrahedron is also an equilateral triangle, the length of whose sides is a. The altitude DF of the
tetrahedron is c/2, since the plane containing the three atoms entirely within the hexagonal cell in
Fig. 2.7 is located midway between the basal plane and the top of the hexagonal prism.

A a 3 /2

a E
a/2
B
(c)
(a) D
D

c/2
a
C
A c/2
a F E A
a 3/3 F
B a/2
(b) (d)

Fig. 2.8: Close packed Tetrahedron.


(a) A hard Sphere Model of the Tetrahedron (b) An Open Drawing of
this Tetrahedron (c) and (d) Calculations for determining Ideal c/a ratio

Applying Pythagorean theorem to the triangle ABE [Fig.2.8 (c)], length of the median AE of the
basal equilateral triangle

AE = a 2  (a/2) 2 = a 3 /2

Since the attitude DF of the tetrahedron intersects the median AE at point F, which is two thirds of
the distance from A to E or a/ 3 . Now from triangle ADF (Fig. 2.8 (d)

DF = c/2 = a2 – ( a 3 /3 )2 = a 2 /3

and therefore c = 2a 2 /3 or c = 1.633 a

Most material do not have c/a ratios which are equal to the ideal value of 1.633a, but some are
quite close and the manner in which they deviate from ideality is important.

21
Introduction to Crystallography

HCP Density of Packing:


Referring to Fig. 2.7 (b), we see that there are eight atoms at the comers of each cell and since each
atom is shared by eight cells there is a total of 8 x 1/8 = 1 atom per cell for the comer atoms. There
is also one atom per cell at the mid-plane. Therefore, there are two atoms per unit cell. The cell
volume (for ideal c/a = 1.633) is

a 3
V = (a) x x (c)
2

a2 3 2
= x 2a = a3 2
2 3

The atomic diameter d = a. Thus the density of packing in HCP crystals is 0.742. This is the same
density of packing that we found in the FCC structure. However, the order of stacking of the close-
packed hexagonal crystals differs from that of the FCC structures.

Comparison of the FCC and CPH Structures


Since nearest neighbour atoms touch one another in the hard sphere model, there are certain planes
of atoms where each atom is in contact with all its nearest neighbours in that plane. These planes
are referred to as close-packed planes and if the planes themselves are sketched, it may be seen
that this configuration represents the most compact arrangement of atoms in a plane (Fig. 2.9).

(a)

(b)

Fig. 2.9: A Close Packed Plane


(a) Each Atom is in Contact with all its Neighbours (b) There are Depressions
at the Centre due to the Spherical Shape of the Atoms

By stacking these closed packed planes over each other in certain sequences the close-packed
planes over each other in certain sequences the close-packed hexagonal lattice, or the face-

22
Structure of Metals

centered cubic lattice may be obtained. That there is more than one way of stacking close-packed
planes can be shown in the following way (Fig. 2.10).

Fig. 2.10: Close Packing of Planes


The Depressions at the Centres of the Equilateral Triangles are the
Possible Sites for the Atoms of the Next Plane

If the close-packed plane shown in Fig. 2.9 is imagined in three dimensions, there is a depression
over the center of the equilateral triangle formed by the centers of three adjacent atoms. This
depression is noticeable only when the atoms are sketched in perspective. This position is
indicated by an ‘x’ or an ‘o’ in the sketch (Fig. 2.10) which is an overhead view of the close-
packed plane. If an atom is placed in the site whose center is at x, then it is located in the place
immediately above the original place and its outline is shown by broken circles (Fig. 2.10). But
there are more sites available than those indicated by the symbol x. Between every three atoms
there is a possible site for an atom above or below the plane, but these planes are separated into
two distinct sets. If an x-site is chosen for the first atom in the plane immediately above the
reference plane, then all other x-sites are specified as shown in Fig. and the remaining set of sites
cannot be used of this plane of atoms. This remaining set of sites designated by the symbol o may
be used for other planes of atoms.
There is no preference for the choice of either x-or o-sites by the first atom setting on a single
close-packed plane but, once it is there, in a particular site, the entire plane usually follows the suit
and atoms settle in the same type of site. If the original atomic plane, which has atoms denoted by
circles of closed lines, is designated as an A-plane, the plane of atoms which are located in the x-
set of sites is designated as a B-plane and the plane of atoms located in the o-set of sites is a C-
plane.
There are many alternate methods of stacking such close-packed planes one on top of the other,
some of the possible methods of packing are ABC, ABA, ACB, ACA. Which indicates that for
ABC the A-plane is above B plane, which is above the C plane in such a way that the atoms of the
A-plane are situated in the depressions of the B-plane and the atoms of the B-plane are in the
depressions of the C-plane. The atoms of an A-plane would be directly over the atoms of any other

23
Introduction to Crystallography

A plane in the stacking sequence. The face centered cubic structure is characterized by a packing
sequence of the type ABCABC. Since the close packed planes are arbitrarily labeled with the
letters A, B or C the unit cell could equally well be represented as having an ACBA structure. The
important points to remember are that the first three lettered planes must be different and that this
three letter group must be repetitive to generate a face centered cubic structure.
For each hexagon, the basal planes at the bottom and top have their atoms directly over one
another (A-planes), while the mid-planes (B-plane) have their atoms located in the depressions of
the A-planes and since each of the basal planes and mid-planes is a close-packed plane, the order
of packing is ABABABA.
There is no basic difference in the packing obtained by stacking of spheres in the face centred
cubic or the close packed hexagonal arrangement., since both give an ideal close packed structure.
There is, however, a marked difference between the physical properties of hexagonal close packed
metals (such as cadmium, zinc and magnesium) and the face centred cubic metals (such as
aluminium, copper and nickel), which is directly related to the difference in their crystalline
structure. The most striking difference is in the number of close packed planes. In the face centred
cubic lattice there are four planes of closest packing, the octahedral planes; but in close packed
hexagonal lattice only one plane, the basal plane, is equivalent to the octahedral plane. The single
close packed plane of the hexagonal lattice create, among other things, plastic deformation
properties that are much more directional than those found in cubic crystals.

Example
Calculate the change in volume that occurs when BCC iron is heated and changes to FCC iron. At
the transformation temperature the lattice parameter of BCC iron is 2.863A and the lattice
parameter of FCC iron is 3.591A.

Answer:
Volume of BCC cell = α3 = (2.863)3 = 23.467 Å 3
Volume of FCC cell = α3 = (3.591)3 = 46.307 Å3
But the FCC unit cell contains four atoms and the BCC unit cell contains only two atoms. Two
BCC unit Cells, with a total volume of 2(23.467) = 46.934 Å3, will contain four atoms.
Thus, the volume change is
46.307  46.934
= × 100
46.934
= -1.34%
The negative sign indicates that iron contracts on heating.

24
Structure of Metals

Example 2.2
On cooling through 880oC, titanium goes through a phase change analogous to iron, except that in
this case the crystal structure changes from body centred cubic (a = 3.32 Å) to hexagonal close
packed ( a = 2.956 Å, c = 4.683 Å). What is the volume change?

Solution
The volume of the bcc is = (3.32 Å)3 = 36.6 Å3 for the two atoms per unit cell.
The volume of the hexagonal close packed is

a 3
= (a) x x (c)
2
= 35.4 Å3
Therefore, the change in volume is (35.4 - 36.6)/36.4 = - 3.3 per cent
The negative sign indicates that the volume contracts when bcc changes to hcp.

CORRELATION OF DATA ON UNIT CELLS WITH MEASUREMENT OF DENSITY


The dimensions of the unit cell can be checked with engineering data like the density. For
example, the density of copper is 8.96 g/cm3 at 20oC. We should find the same value using the
mass of copper in the volume of a unit cell. We can find mass (M) and volume (V) from
definitions. We can find the weight of an atom of copper, since each gram atomic weight of copper
(63.5 g/at. wt.) has 6.02 x 1023 (Avogadro’s number) atoms. The volume of a unit cell is merely
the lattice dimension cubed. Before we can complete the calculation, however, we must determine
the number of atoms per unit cell.
The theoretical density of a metal is calculated by using the general formula
(atoms/unit cell) (atomic mass of each atom)
ρ=
(volume of unit cell) (Avogadro' s Number)

Example: 2.3
The value of ao for copper at 20oC is = 3.61 Å. Calculate the theoretical density of copper, using
the knowledge that the cell of this element is face centred cubic.

Solution
63.5 g/ at. wt
4 atoms/unitcell 23
M 6.02 x 10 atoms/atomic weight
Density = = = 8.98 g/cm3
V (3.61)3 x 10  24 cm 3 /unit cell

25
Introduction to Crystallography

This value is quite close the density of a block of copper (8.96 g/cm3). The difference is due to the
presence of voids in the crystal structure.

Example 2.4
Determine the density of body centred cubic iron, which has a lattice parameter of 2.866A.

Solution
Atoms/cell = 2
Atomic mass = 55.85 g/g mole
Volume of unit cell = a3o = (2.866 × 10-8)3 = 23.55 × 10-24 cm3/cell
Avogadro number NA = 6.02 × 1023 atoms/g.mole
( 2)(55.85)
ρ=  24 23
= 7.879mgm-3
( 23.55  10 )(6.02  10 )

The measured density is 7.87 mgm-3. The reasons for slight discrepancy between the theoretical
and measured density has been explained in Example 2.3.

IDENTIFICATION OF PLANES OF ATOMS


Special planes and directions within metal crystal structures play an important part in plastic
deformation, hardening reactions and other aspects of metal behaviour. For example, metals
deform along the planes of atoms that are most tightly packed together. Repeated studies of planes
of atoms that are important in determining the properties of actual crystals have led to a
generalisation concerning the identification of these planes which is now usually called (the law of
rational indices or) Miller indices.

The Miller Indices of Crystallographic Planes


The Miller system of designating indices for crystallographic planes and directions is universally
accepted. In the discussion that follows, the Miller indices for cubic and hexagonal crystals will be
considered. The indices for other crystal structures are not difficult to develop but since metals
primarily have cubic and hexagonal structures, we would limit our discussion to cubic and
hexagonal crystal.
The three steps used in the determination of Miller indices of a given plane is illustrated below
with the aid of Fig. 2.11.
1. Determine the intercepts of the plane on the three crystal axes. These intercepts on the three
axes x, y and z are expressed in terms of number of axial lengths ( X = 2, Y = 3, Z = 1) from

26
Structure of Metals

the origin. If the plane under consideration passes through the origin, the origin of the
coordinate system must be moved.
2. The usual convention requires that we use the reciprocals (½, 1/3, 1) of these quantities.
3. Reduce the reciprocals to the smallest integers in the same ratio. (The reciprocals might not
always be integers as in this case. It is convenient to make integers of them with the condition
that they remain in the same ratio.
These integral reciprocals, enclosed in parentheses (326) are called Miller indices and are denoted
by h, k and I. The negative numbers should be written with a bar over the number.

b
Y

Fig. 2.11: An Imaginary Plane Showing the Intersection


with the Three Axes

It should be noted that these indices give only geometrical planes and say nothing about the
distribution or kind of atoms in the plane. On the other hand, a set of Miller indices such as (326)
describes not merely a single crystal plane but the entire array of planes parallel to the plane on
which the three-step analysis was carried out.

Identification of Planes, Examples


To gain a working knowledge of this system of identification, we shall now apply the above
principles to identify the planes indicated in the cubic cell of Fig. 2.12.
For the plane illustrated in Fig. 1.13(a) the intercept along the x-axis is identifiable, but the plane,
itself is parallel to y and z axes. This means that it never intersects the y and z axes and the value
of the y and z intercepts is infinity, represented by the symbol .

x y z axis
1   intercepts

27
Introduction to Crystallography

Performing the next steps, we find that the Miller indices of this plane is (100). The Miller indices
for the other planes (in Fig. 2.12) shown by shades can be identified in the same way.
Z Z Z

Y Y Y

X (100) X (110) X (111)

Fig. 2.12: Miller Indices of Planes in the Cubic System

An important aspect of Miller Indices for planes should be noted. Planes and their negatives are
identical (this is not the case for directions). For example, the shaded plane in Fig. 2.13. has the
indices (020) if the X, Y and Z are used but has the indices (0 2 0) if the X’, Y’ and Z’ coordinates
are used. But we are considering the same plane. Therefore, (020) = (0 2 0).

Z Z’

X X’

Fig.2.13: A Plane and Its Negative are Identical

Multiplicity of Planes
There are number of planes having the same geometrical distribution of atoms. These are called
equivalent planes. For example in the cubic system the (100), (010), (001), which are the face
planes of the cubic cell, are equivalent. These planes are alike except for their orientation, and the
entire family of the face planes may be denoted by {100} - the curly brackets indicating the family
of equivalent planes. The {100} family contains six equivalent planes (100), (010), (001), ( 1 00),
(0 1 0) and (00 1 ) in the cubic system. The number of equivalent planes is called the multiplicity
and range quite high for high index planes (Table 2.3)

28
Structure of Metals

Table 2.3: Equivalent Planes (where a =b = c only)


hkl Equivalent Planes Multiplicity
{100} (100) ( 1 00) 6
(010) (0 1 0)
(001) (00 1 )
{110} (110) ( 1 10) (101) (0 1 1) 12
(101) (1 1 0) (10 1 ) (01 1 )
(011) ( 1 1 0) (101) (0 1 1 )
{111} (111) ( 1 11) ( 1 1 1) (1 1 1) 8
(1 1 1) (111)
(11 1 ) (1 1 1 )
{h00} 6
{hh0} 12
{hhh} 8
{hk0} 24
{hkl} 48

Miller Indices for Hexagonal Crystals


In the case of hexagonal crystals, the face planes in Fig. 2.7(a) are all similar planes, just like the
six faces of the cube. The Miller Indices of these planes are: (100), (010), ( 1 10), ( 00), (0 1 0),
(1 1 0). Thus the same numbers do not exist in the set of integers that indicate the indices of the
planes that are all similar. In other words, the fact that these are all symmetrical planes is not
brought out by the Miller Indices. In order to overcome this situation, the use of a 4 co-ordinate
system in place of 3 was suggested by Miller-Bravais. Four-digit hexagonal indices are based on a
coordinate system containing four axes. Three axes correspond to close-packed directions and lie
in the basal plane of the crystal, making 120o angles with each other and are designated the a1, a2
and a3 axes. The fourth axis is normal to the basal plane and is called the C-axis.
Fig. 2.14 shows the hexagonal unit cell superimposed upon the four-axis coordinate system. It is
customary to take the unit of measurement along the a1. a2 and a3 axes as the distance between
atoms in a close-packed direction. The magnitude of this unit is indicated by the symbol a. The
unit of measurement for the C-axis is the height of the unit cell that is designated as c. Let us now
determine the Miller indices of several important close-packed hexagonal lattice planes. The
uppermost surface of the unit cell in Fig. 2.14 corresponds to the basal plane of the crystal.
Since it is parallel to the axes a1, a2 and a3, it meets them at infinity. Its C-axis intercept, however,
is equal to 1. The reciprocals of these intercepts are 1/, 1/, 1/, and 1/1. Miller indices of the

29
Introduction to Crystallography

basal plane are, therefore, (0001). The six vertical surfaces of the unit cell are known as prism
planes. Let us now consider the prism plane that forms the front face of the cell, which has the
intercepts as follows: a1 at 1. a2 at , a3 at -1 and c at . Its Miller indices are, therefore, (10 0).

C
a3

-a1
c

- a2 a2

a1
a - a3

Fig. 2.14: Hexagonal Unit Cell Superimposed upon the


4-Axes Coordinate System

Another important type of plane in the hexagonal lattice is shown in Fig. 2.14 with the intercepts
of axes at a1 at 1, a2 at , a3 at -1 and c at ½ and the Miller indices, are accordingly (10 2). In
terms of the 4 co-ordinate system, the indices of the pyramidal planes will be: (10 0), (01 0),
(1 00), ( 010), (0 10), ( 100). Thus, the same set of numbers appears for all the planes and
reveals the symmetry of the crystal.
The use of 4 co-ordinates does not change the system in any way
as the unit vectors along the three co-planar co-ordinate axes
make a closed system as in Fig. 2.15 so that a3 a2

- a1 + a2 + a3 = 0
or a3 = -(a1 + a2)
a1
Fig. 2.15: The Closed System Along the
Using this relation all calculations made on the basis of the four
3 Coplanar Coordinate Axes
co-ordinate system may readily be converted into that in the three
co-ordinate system. Since the indices of planes are mere numbers,
conversion of Miller indices to Miller-Bravais indices may be achieved by using the following
relation. Let (hkl) and (HKIL) represent the indices of the same plane in the two systems. Then:
H = h, K= k, 1=-(h + k) and L = I

30
Structure of Metals

INDICES OF DIRECTION
Certain directions in unit cells are of particular importance. Metals deform, for example, in
directions along which the atoms are in close contact. The direction of closest packing in the bcc
lattice may be described as the body diagonal while the corresponding direction in the fcc lattice is
the face diagonal. To simplify the representation any significant direction within a space lattice or
crystal structure is universally identified by means of Miller indices of direction. In general form
these resemble the Miller indices of the planes, but their derivation and significance are quite
different. In the discussion that follows, the Miller indices for cubic and hexagonal crystals will be
considered. The indices for other crystal structures are not difficult to develop.

C
Direction Indices in the Cubic Lattice
To distinguish from planes, square brackets
are used for Miller indices for directions. A
direction such as [210] can be constructed in [100]
the following manner: A line is drawn from
= [2
the origin through the point having the co- a3
ordinates x = 2, Y = 1 and z = 0 in terms of
axial lengths. This line and all lines parallel
to it are then in the given direction. The
+2
directions of the same form are sometimes [110] a2
represented by enclosing the indices of one -1
= [11 2 0]
of these directions in carets. Thus [uvw] [010] = [1210]
represents the indices for a specific a1
direction, while <uvw> represents the
Fig. 2.16: Typical Directions in HCP Cells
indices of the whole family of directions.
The procedure for finding the Miller indices for a given directions is as follows:
(i) Using a right-handed co-ordinate system, determine the co-ordinates of the two points that lie
on the direction.
(ii) Subtract the co-ordinate of the tail point from the head point to obtain the number of lattice
parameters traveled in the direction of each axis of the co-ordinate system.
(iii) It is customary to avoid fractional co-ordinates by multiplying by a suitable factor and to use
the smallest integers that will locate a point on the line.
(iv) Enclose the numbers in square brackets [ ]. If a negative sign is produced, represent the
negative sign with a bar over the number.
In the case of directions also, use of four co-ordinates will reveal the similarity and symmetry of
the hexagonal crystallography. It may be mentioned that mathematical conversion of directional
indices from one system to the other is usually more complex and more tedious than is separate
derivation of the two sets of indices from their own systems of reference axes.

31
Introduction to Crystallography

Example2.5
Determine the Miller indices of directions A, B and C in Fig. 2.17.

Z
0, 0, 1

1, 1, 1

C
B

A 0, 0, 0 Y
½, 1, 0
1, 0, 0
X

Fig. 2.17: Crystallographic Directions and


Coordinates Required for Example 1.5

Solution
Direction A
(a) Two points are 1,0,0 and 0, 0, 0
(b) 1, 0, 0 – 0, 0, 0 = 1, 0, 0
(c) No fractions to clear or integers to reduce
(d) [100]
Direction B
(a) Two points are 1,1,1 and 0, 0, 0
(b) 1, 1, 1 – 0, 0, 0 = 1, 1, 1
(c) No fractions to clear or integers to reduce
(d) [111]
Direction C
(a) Two points are 0,0,1 and ½, 1, 0
(b) 0, 0, 1 –½, , 1, 0 = -½, -1, 1
(c) 2 (-½, -1, 1) = -1, -2, 2
(d) [ 1 2 2].

32
Structure of Metals

Example 2.6
Determine the Miller-Bravais indices for directions A and B in Fig. 2.18.
C

a3

a2
B
a1
Fig. 2. 18: Hexagonal Close Packed Cell for Example 2.6.
Solution
Direction A
(a) Two points are 0,0,1 and 1, 0, 0
(b) 0, 0, 1 – 1, 0, 0 = -1, 0, 1
(c) No fractions to clear or integers to reduce
(d) [ 1 10]
Direction B
(a) Two points are 0,1,0 and 1, 0, 0
(b) 0, 1, 0 – 1, 0, 0 = -1, 1, 0
(c) No fractions to clear or integers to reduce
(d) [ 1 10].

IDENTIFICATION OF ATOMIC POSITIONS


The position of an atom within a unit cell is given by coordinates of the atom expressed. in units of
the lattice parameters a, b and c, with respect to the coordinate system x, y, and z. Fig. 2.19 shows
the coordinates for the body-center atom in the bcc unit cell. If we start at the origin and proceed
along the x-axis, we cover a distance a/2, along the y-axis, the distance is a/2 and along the z-axis

33
Introduction to Crystallography

it is a/2. If this were a non-cubic cell, we would find the position coordinates in the same manner,
except that we would use the three lattice parameters, a, b, and c, each along its own axis.

a/2 a/2
a/2

X
Fig. 2.19: Position of Body Centre Atom

The distances are positive as long as the direction is out from the origin in the positive direction
along the axis. The position coordinates are then a/2, a/2, and a/2 or, expressed in terms of the
lattice parameter as a unit of measure, they are ½, ½, ½.

Fig. 2.20: Positions of Close Packed Atoms.Hexagonal Cell

To compute the positions of the atoms in the mid-plane of the hexagonal close packed structure
(Fig. 2.20) it is necessary to determine the path from the origin to each atom. To reach atom 1

34
Structure of Metals

from the origin one must traverse a distance equal to a/3along x-axis, 2a/3 along the y-axis and c/2
along the z-axis. The position coordinates of atom 1 is then ⅓, ⅔ , ½.
It is important to remember that since a and c are not the same in length, the position coordinates
are measured in different units of length along each axis. To reach point 2 one must travel a
distance a/3 along x-axis, -a/3 along y-axis and c/2 along z-axis. The coordinates of point 2 are
therefore ⅓, -⅓, ½. By similar reasoning the position coordinates of atom 3 are -⅔ , -⅓, ½

Other Unit Cell Calculations


Atomic Radius: The atomic radius is a measure of the similarity of atoms and can be calculated
from the dimensions of the unit cell. At first, the dimensions along which the atoms are in contact
are determined. Then we merely divide the figure by the number of atomic radii present. Fig. 2.21
shows the relationship between the atomic radius and the lattice parameter in some crystals.

ao 3
r= 4
ao 2
r= 4 ao
r r= 2 r
2r r
2r
r r
r
(b) (c)
(a)

Fig. 2.21: Calculation of Atomic radius (a) FCC Crystal (b) HCP Crystal (c) BCC Crystal

Example 2.7
Determine the relationship between the atomic radius and the lattice parameter in SC, BCC, and
FCC structures.

Solution
In simple cubic structures the atoms touch along the edge of the cube. The corner atoms are
centered on the corners of the cube, so
ao = 2r (.1)
In FCC structures atoms touch along the face diagonal of the cube, which is √2 ao in length. There
are four atomic radii along this length two radii from the face-centered atom and one radius from
each corner.

35
Introduction to Crystallography

Therefore,
4r
ao = (.2)
2

In BCC structures, atoms touch along the body diagonal, which is 3 ao in length. There are two
atomic radii from the center atom and one atomic radius from each of the corner atoms on the
body diagonal. so
4r
ao = (.3)
3

Example 2.8
The atomic radius of iron is 1.24 A. Calculate the lattice parameters of BCC and FCC iron.

Answer:
4r (4 x 1.24)
For BCC iron: ao = =
3 3
= 2.86 Ǻ

4r (4 x 1.24)
For FCC iron: ao = =
2 2
= 3.51 Ǻ
Planar Density: When slip occurs under stress (plastic deformation), it takes place on the planes
on which the atoms are most closely packed. The planar density is calculated as follows:

ao 2 atoms
2
ao

ao
Fig. 2.22: Calculation of Planar Density of the Face Plane
of an FCC Unit Cell

If an atom belongs entirely to a given area, such as the atom in the center of the face in an face
centred cubic structure, we see that the trace of the atom on the plane is a circle (Fig.2.22).

36
Structure of Metals

Therefore in the area a 2o we count one atom for the center, but one-quarter atom for each of the
corners because each has a trace of only one-quarter circle on the area a 2o . The planar density is 2/
a 2o (atoms/Å2). It should be added that in all these density calculations one of the ground rule is
that a plane or a line must pass through the center of an atom or the atom is not counted in the
calculation.

Linear Density: This is an important concept because when planes slip over each other, the slip
takes place in the direction of the closest packing of atoms on the plane. We calculate the linear
density by the following convention.
A

2 atoms
ao ao 2

B
A
o
Fig. 2.23: Calculation of Linear Density on the Face Diagonal
of an FCC Unit Cell
If a line passes completely through an atom, the trace of the atom on the line is one diameter
(Fig.2. 23). In a face centred cubic face the center atom counts for one atom on the face diagonal.
The corner atoms make traces equal to only one-half diameter each on the line of length AB.
Therefore, the linear density in the AB direction [110] in a face centred cubic structure is
1  ½  ½ atoms 2 atoms
=
ao 2 Å ao 2 Å

By the same reasoning the linear density in a bcc in the [111] direction is

2 atoms
[The atoms touch along the body diagonal of a bcc].
ao 3 Å

Example 2.9
Calculate the linear and planar density for a face centred cubic structure in the [112] direction and
the (111) plane, respectively.

37
Introduction to Crystallography

Solution
When determining the linear atomic density, we must remain in the reference cell. Therefore, the
centers of the atoms cut in a face centred cubic structure are at 0, 0, 0 and ½, ½, 1. There are two
radii or one atom.
Z

d [112] ao
ao
O Y
ao
X d

0, 0, 0 r r ½, ½, 1

From the geometry of the large right triangle

(2d) 2 = (ao 2 ) 2 + (2 ao )2 or d = ao 6 / 2

1 atom 2
Therefore, linear atomic density = = (atoms/Å)
ao 6 /2 ao 6
Z

ao 2

(111)
Y

X ao 2

Number of atoms = 3 x 1 6 + 3 x 1 2 = 2 atoms

3 a0
Area of the equilateral triangle =
2
2 4
Therefore, planar atomic density = = (atoms/Å2)
3 a o2 /2 3 a o2

38
Structure of Metals

NUMBER OF ATOMS IN CUBIC STRUCTURES


We have seen that the number of atoms per unit cell in a simple cubic structure is 1, in a body
centred cubic structure is 2 and in a face centred cubic structure is 4. We shall determine the
number of atoms per square millimetre on the most common planes in each of the cubic crystals.

Simple Cubic Structures


Fig. 2.24 shows actual view and plane (100) of a simple cubic structure containing one atom each
on all the 8 corners of the cube. The plane ABCD in Fig. 2.25 (b) contains one-fourth atoms in
each of its four corners. [Trace of the atom at each corner of the plane is ¼ th of a circle].

A B A B

D C D C
a a
(a) (b)
Fig. 2.25: Plane (100) in Simple Cubic Structures

Therefore, total number of atoms contained in the plane ABCD


1
=4x =1
4

The lengths AB and AD of the plane ABCD = a mm

Area of the plane ABCD


= AB x AD = a x a = a2
and number of atoms per square mm

No. of atoms 1
= = 2
Area of plane a

39
Introduction to Crystallography

The Plane (110)


Fig. 2.25 shows actual view and plane (110) of the simple cubic structure containing one atom
each on all the 8 corners of the cube. The plane ABCD [Fig. 3.36 (b)] contains one-fourth atom in
each of its four corners.

B A B

C D C

D
(a) (b)

Fig. 2.25: Plane (110) of Simple Cubic Structures

Therefore total number of atoms contained in the plane ABCD


1
=4x =1
4

the length AD = a mm

and the length of AF = a√2 mm

Area of the plane ABCD

= AB x AD = a x a√2 = a2√2

and number of atoms per square mm

No. of atoms 1 0.707


= = 2 =
Area of plane a 2 a2

The Plane (111)


Fig. 2.26(a) and (b) shows the actual view and plane (111) of the simple cubic structure containing
one atom each on All the 8 corners of the cube. We know the plane EDG in Fig. 2.26(b) contains
one-sixth atom in each of its three corners.

40
Structure of Metals

Therefore total number of atoms contained in the plane ABC


1
=3x = 0.5
6
E E

D D G
(a) (b)

Fig. 2.26: Plane (111) of Simple Cubic Structures

The length of equilateral triangle EDG


= a√ 2
and height of perpendicular
= a √ 2 sin 600 = = a √ 2 x 0.866
 Area of the plane ABC
1
= x a √ 2 x a √ 2 x 0.866 = 0.866 a2
2

and number of atoms per square mm

No. of atoms 0.5 0.58


= = = 2
Area of plane 0.866 a 2 a

Example 2.10
The lattice constant of a unit cell of Potassium Chloride (KCL) crystal is 3.03 A. find the number
of atoms/mm2 of planes (100), (110) and (111) if KCL has simple cubic structure.

Solution
Given.
Lattice constant, a = 3.03A = 3.03 x 10-7 mm

41
Introduction to Crystallography

(100) plane
The number of atoms in the (100) plane of a simple cubic structure
1 1
= = 10.9 x 1012 Ans.
a2 (3.03 x107 ) 2

(110) plane
The number of atoms in the (110) plane of a simple cubic structure
0.707 0.707
= = 7.7 x 1012 Ans.
a 2
(3.03x10  7 ) 2

(111) Plane
The number of atoms in (111) plane of a simple cubic structure
0.58 0.58
= = = 6.3 x 1012 Ans.
a2 (3.03 x107 ) 2

Body Centred Cubic Structures

The Plane (100)


Fig.2.27 shows actual view and plane (100) of a body centred cubic structure containing one atom
on each of the 8 corners of the cube and one atom in the centre of the body. The plane ABCD in
Fig. 2.27(b) contains one-fourth atom in each of its four corners and the body centred atom will
not appear on the plane.

A A B
B

D C D C
(a) (b)
Fig. 2.27: Plane (100) of Body Centred Cubic Structure

Therefore total number of atoms contained in the plane ABCD


1
=4× =1
4

42
Structure of Metals

The lengths AB and AD of the plane ABCD


= a mm
Area of the plane ABCD
= AB × AD = a × a = a2
and number of atoms per square mm
No. of atoms 1
= =
Area of plane a 2

The Plane (110)


Fig. 2.28(a) and (b) shows actual view and plane (110) of the body centred cubic structure
containing one atom on each of the 8 corners of the cube and one atom in the centre of the body.
The plane AFGD in Fig 2.28(b) contains one-fourth atom in each of its four corners and one full
atom in the centre of the plane.

B A B

C D C
D
(a) (b)

Fig. 2.28: Plane (110) of Body Centred Cubic Structure

Therefore total number of atoms contained in the plane AFGD


1
= (4 × ) =1+1=2
4
The length AB = a mm
The length AF = a √ 2mm
Area of the plane AFGD
= AB × AD = a = a √ 2 = a2 √ 2 mm2

43
Introduction to Crystallography

and number of atoms per square mm


No. of atoms 2 1.414
= = =
Area of plane a 2 2 a2

The Plane (111)


Fig. 2.29 shows the actual view and (111) plane of the body centred cubic structure. The plane
EDG in Fig. (b) contains one-sixth atom in each of its three corners and one full atom in the centre
of the body.
E E

G
D D G
(a) (b)

Fig. 2.29: Plane (111) of Body Centred Cubic Crystal


1
Therefore total number of atoms contained in the plane EDG = ( × 3 ) + 1 = 1.5
6
Now, the length DG of equilateral triangle EDG
=a√2
and height of perpendicular
= a √ 2 sin 60o = a √ 2 × 0.866
Area of the plane EDG
1
= × a √2 × a √2 × 0.866 = 0.866 a2
2
and total number of atoms per square mm
No. of atoms 1. 5 1.732
= = 2
=
Area of plane 0.866 a a2

Example 2.11
The lattice constant of a unit cell of iron is 2.87 A. Find the number of atoms/mm2 of planes (100),
(110) and (111), if iron has BCC structure.

44
Structure of Metals

Solution
Given : Lattice constant, α =2.87 Å = 2.87 × 10-7 mm (100) plane

(100) Plane
The number of atoms in the (100) plane of BCC structure,
1 1
= 2
= -7 2
= 1.21 × 1013 Ans.
a (2.87  10 )
(110) plane
The number of atoms in the (110) plane of B.C.C structure,
1.414 1.414
= 2
= -7 2
= 1.72 × 1013 Ans.
a (2.87  10 )

(111) plane
The number of atoms in the (111) plane of B.C.C structure,
1.732 1.732
= 2
= = 2.1 × 1013 Ans.
a (2.87  10 -7 ) 2

Face Centred Cubic Structure


The plane (100)
Fig. 2.30 shows actual view and plane (100) of a face cubic structure containing one atom each on
all the 8 corners and one atom each on all the 6 faces of the cube. The plane (100) in Fig. 2.30 (b)
contains one-fourth atom in each of its four corners and one atom in the centre of the face.

A A B
B

D C D C
(a) (b)

Fig. 2.30: Plane (100) of Face Centred Cubic Structure

45
Introduction to Crystallography

There-fore total number of atoms contained in the plane


1
= (4 × )+1=2
4
the lengths AB and AD of the plane ABCD = a mm
Area of the plane ABCD
= AB × AD = a × a = a2
and number of atoms per square mm
2
=
a2

The plane (110)


Fig. 2.31 shows actual view and plane (110) of the face-centred cubic structure containing one
atom each on all the 8 corners and one atom each on all the 6 faces of the cube. We know that the
plane (110) in Fig.2.41 (b) contains one-fourth atoms in each of its four corners and half atom in
each of its two long edges AF and DG.

B A B
A

C D C
(b)
D
(a)
Fig. 2.31: Plane (110) of Face Centred Cubic Structure

Therefore total number of atoms contained in the plane


1 1
= (4 × )+(2× )=2
4 2
The length AD =a mm
and length AF = a √ 2 mm

46
Structure of Metals

Area of the plane


= AD × AF = a ×a √ 2 = a2 √ 2 mm2
and number of atoms per square mm
2 1.414
= =
a 2
2 a2

The plane (111)


Fig. 2.32 shows actual view and plane (111) of the face centred cubic structure containing one
atom each on all the 8 corners and one atom each on all the 6 faces of the cube. We know that the
plane in Fig 2.32(b) contains one-sixth atom in each of its three corners and half atom in each of
its three edges. E
E

D D G
(a) (b)

Fig. 2.32: Plane (111) of Face Centred Cubic Structure

Therefore total number of atoms contained in the plane


1 1
=( ×3)+( ×3)=2
6 2
The length of one side of the equilateral triangle = a √ 2
and height of perpendicular = a √ 2 sin 60o = a √ 2 × 0.866

Area of the plane


1
= ×a √ 2 × a √ 2 × 0.866 = 0.866 a2
2
and total number of atoms per square mm

No. of atoms 2 2.31


= = = 2
Area of plane 0.866a 2 a

47
Introduction to Crystallography

Example 2.12
How many atoms per square millimetre surface area are there in (100) plane, (110) plane and (111)
plane, for lead which has FCC structure and a lattice constant, α = 4.93 Å.

Solution

Given.
Lattice constant,

a = 4.93 Å = 4.93 × 10-7 mm (100) Plane

(100) Plane

The number of atoms in (100) plane of F.C.C. structure

2 2
= = = 8.2 × 1012 Ans.
a 2
(4.93  10 -7 ) 2

(110) Plane

The number of atoms in (110) plane of F.C.C. structure

1.414 1.414
= 2
= = 5.8 × 1012 Ans.
a (4.93  10 -7 ) 2

(111) Plane

We know that the number of atoms in (111) plane of F.C.C. structure

2.31 2.31
= 2
= = 9.5 × 1012 Ans.
a (4.93  10 -7 ) 2

48
Structure of Metals

PERPENDICULAR DISTANCE BETWEEN A PLANE AND THE ORIGIN


Let the Miller Indices of the plane PQR (Fig.2.33) be h, k and l. Let HL be perpendicular from the
origin of the cube to the plane PQR
Z

E F

A B
R

L Q G
Y
O
P
D C

. Fig. 2.33: Determination of Distance of a Plane from the Origin

Let a = Cubic edge or the lattice constant of the cube


d = Perpendicular distance between the origin (O) and the plane (i.e. OL),
OP, OQ and OR = Intercepts of the plane along x, y and z axis respectively, and
a, β and  = angles, which the perpendicular makes with x, y and z axis respectively.
We know that Miller Indices of a plane are the smallest integers of the reciprocals of its intercepts.
Therefore, the intercepts may also be expressed as reciprocals of Miller Indices. Or in other words,
1 1 1 a a a
OP : OQ : OR = : : = : :
h k l h k l
a a a
 OP = , OQ = and OR =
h k l
From the geometry of the right angles OPL, OQL and ORL we know that
OL h dh
cos α = =dx =
OP a a
OL k dk
cos β = =dx =
OQ a a

OL l dl
or cos  = =dx =
OR a a

49
Introduction to Crystallography

Since cos2 α + cos2 β + cos2  = 1, therefore

2 2 2
 dh   dk   dl 
  +   +   =1
 a   a  a

d2
(h2 + k2 + l2) =1
2
a

a2
 d2 =
(h 2  k 2  l 2 )

a
 or d =
h  k 2  l2
2


Note. If we draw a plane through the cube edge (H) and parallel to the plane PQR, then this
relation may be used for obtaining the distance between these two planes. In this case, the distance
HL is known as interplaner distance.

Example 2.13
Å f.c.c. crystal has an atomic radius of 1.246. What are the d200, d220 and d111 spacing ?

Solution
Given : Atomic radius, r = 1.246 Å
We know that for a f.c.c. structure, the lattice constant,
4r
a =
2

4x1.246
= = 3.52
2

and spacing with Miller Indices (h, k, l)


a
d=
h  k2  l2
2

3.52 3.52
d200 = 2
= = 1.76 Å Ans.
2
( 2)  ( 2)  ( 0) 2 2

50
Structure of Metals

Similarly
3.52 3.52
d220 = 2
= = 1.24 Å Ans.
2
( 2)  ( 2)  ( 0) 2 2.828

3.52 3.52
and d111 = 2
= = 2.03 Å Ans.
2
(1)  (1)  (1) 2 1.732

Angle Between Two Planes or Directions


Let the two planes ABCD and EFCD meet each other at an angle θ (Fig. 2.34).
A

B
D E

θ
C F

Fig. 2.34: Angle Between Two Planes

Let h1, k1 and l1 = Miller indices of plane ABCD, and h2, k2 and l2 = Miller Indices of plane EFCD.
The angle between these two planes is given by the relation:

h1 h 2  k 1 k 2  l1 l 2
cos  =
2 2 2 2
h1  k 1  l1 x h 22  k 22  l 2

Similarly, the angle () between the two directions having Miller Indices (h 1, k1 and l1) and (h2, k2
and l2) respectively is given by the relation:

h1 h 2  k 1 k 2  l1 l 2
cos  =
2 2 2 2
h1  k 1  l1 x h 22  k 22  l 2

Note: The derivation of this relation is beyond the scope of this book.

51
Introduction to Crystallography

Example 2.14
Find the angle between the two planes (111) and (1 Ī 1).

Solution
Given. planes (111) and (1Ī1).
The general expression for the angle () between two planes with Miller Indices (h 1, k1 and l1) and
(h2, k2 and l2),

h1 h 2  k 1 k 2  l1 l 2
cos  =
2 2 2 2
h1  k 1  l1 x h 22  k 22  l 2

Now substituting the values of h1, k1 and l1 = (111) and h2, k2 and l2 = (1 Ī 1),

(1 x 1)  (1 x 1)  (1 x 1)
cos  =
12  1 2  1 2 x 12  1 2  1 2

1
= = 0.3333
3

 = 700 32’ Ans.


c
Planes of a Form
In any crystal system there are equivalent lattice a
planes related by symmetry. These are called
planes of a form and the indices of one plane, a
enclosed in braces {hkl} stands for the whole set.
In general, planes of a form have the same
spacing but different Miller indices. For
example, the faces of a cube (100), (010), (001), Fig. 1. : A Simple Tetragonal Cell
(Ī00), (0Ī0) and (00Ī) are planes of a form {100}
since all of them can be generated from any one by operation of the 4-fold rotation axes
perpendicular to the cube faces.
In the tetragonal system however, only the planes (100), (010), (Ī00), and (0Ī0) belong to the form
{100}; the other two planes (001) and (00Ī), belong to the different form {001}; the first four
planes are related by a four-fold axis and the last two by a two-fold axis.

52
Structure of Metals

Planes of a Zone
A zone may be defined as a set of faces or planes in a crystal whose intersections are all parallel.
The common direction of the intersection is called the zone axis (Fig. 1. ). Such planes need not
be crystallographically equivalent, the only requirement is that they are all parallel to a line. All
directions in a crystal are zone axis, so the terms directions and zone axis are synonymous.
Zone Axis 
(1 10) (210)
(210)
(100)
(100)

Fig. 1. : All Shaded Planes shown are Planes of the Zone [001].

If the axis of the zone has indices [uvw], then any plane belongs to that zone whose indices (hkl)
satisfies the relation
hu + kv + lw = 0
A proof of this relation will be given in a later section. Any two nonparallel planes are planes of a
zone since they are both parallel to their line of intersection. If their indices are (h1k1l1) and
(h2k2l2), then the indices of their zone axis [uvw] are given by the relations
u = k1 l2 – k2 l1 v = l1 h2 – l2 h1 w = h 1 k2 – h 2 k1
The concept of zone axis is readily understood by examining an ordinary pencil. The six faces of a
pencil all form or lie in a zone because they all intersect along one direction – the pencil lead
direction – which is the zone axis for the set of faces. The number of faces in a zone is not
restricted. For example, the edges of the pencil may be shaved flat to give a 12-sided pencil, i.e.,
an additional six faces in the zone. Verbal definitions are frequently rather clumsy in relation to
the simple concepts that they seek to express. Take a piece of paper and draw on it some parallel
lines. Fold the paper along the line – and there you have a zone.
Certain important crystal planes are often referred to by name without any mention of their Miller
indices. Thus the planes of the form {111} in the cubic system are called octahedral planes. Since
these are the bounding planes of an octahedron. In the hexagonal system, the (0001) plane is called
the basal plane, the planes of the form {10 Ī 0} are called the prismatic planes, and the plane of the
form {10 Ī 1} are called pyramidal planes.

53
Introduction to Crystallography

EFFECTS OF OTHER ELEMENTS ON THE STRUCTURE OF PURE METALS


Only a few elements are widely used commercially in their pure form; pure copper in electrical
conductors is one example. Generally other elements are intentionally added to produce greater
strength or to improve corrosion resistance. Some impurity may also be present to minimize the cost
of refining. Whatever the reason we need to know the effects these elements have on structure because
structure determines properties.
When a second element is added two basically different structural changes are
possible.
1. The atoms of the new element form solid solution with the original element,
but there is still only one phase, such as a face centred cubic structure.
2. The atoms of the new element form a new second phase, usually containing
some atoms of the original element. The entire microstructure may change to
this new phase or two phases may both be present.
We may ask if there are any general principles that determine whether a given
element B will form a separate phase or a solid solution when added to another
element A. This is important in alloy development because in general a solid
solution will be more ductile and more easily shaped, while a two phase material
will usually exhibit greater hardness and strength.
Solute Atoms Solvent Atoms

(a) (b)

Fig.2.37: Types of Solid Solution


(a) Substitutional Solid Solution (b) Interstitial Solid Solution

A solid solution is formed when two metals, which are mutually soluble in the
liquid state, remain dissolved in each other in the solid state in such a way that the
elements cannot be differentiated under a microscope. There are two types of solid
solutions, substitutional and interstitial (Fig. 2.37).

54
Structure of Metals

Substitutional Solid Solution


In such solutions, some of the atoms of the solvent metal are replaced from the
lattice by the solute atoms. For example, silver atoms may substitute for gold
atoms without changing the f.c.c. structure of gold and similarly gold atoms may
substitute for silver atoms in the f.c.c lattice structure of silver.
The most important considerations in forming solid solutions are related to atomic
radii and chemical properties. The atomic volume varies from element to element.
Therefore, the substitution of solute atoms in the solvent lattice must necessarily
involve a change in parameter of the lattice. If the solute atom is larger than the
solvent atom, the change will be an expansion of the solvent lattice; if it is smaller
the change will be a contraction. Hume-Rothery and his colleagues formulated
several general rules governing the extent of solid solution. These rules, known as
Hume-Rothery rules, form a useful basis for predicting alloying behaviour. In
brief these rules are as follows:
(i) If the difference in size between the solvent and solute atoms is less than 8 per cent, the two
metals may be soluble in each other in all proportions. If the size difference is more than 15
percent solubility will be very limited.
(ii) the solute and solvent atoms should be closely related in valence. The high valence metal would
accept into solution only a small percentage of the lower valence metal, while the lower valence
metal would accept a relatively large amount of the higher valence metal into solution.
(iii) The solute and solvent atoms should be closely related in chemical activity. Generally, the farther
apart the elements are in the periodic table the greater is their chemical affinity.
(iv) The solute and solvent atoms should have similar crystal structures. Complete solid solubility of
two elements is never attained unless the elements have the same type of crystal structure.

Example 2.15
Predict whether extensive substitutional solid solubility would occur between
copper as the solvent and aluminium, nickel and chromium as the solute element.
Element Atom Radius Å Crystal Structure Electronegativity
Copper 1.28 FCC 1.9
Aluminium 1.43 FCC 1.5
Nickel 1.25 FCC 1.8
Chromium 1.25 BCC 1.6

55
Introduction to Crystallography

Answer
rx  rCu
The radius difference is =
rCu
Cu – Al + 11.7 percent
Cu – Ni - 2.3 percent
Cu – Cr -2.3 percent
All the elements fall within the required ± 15 percent radius requirement. The electronegativity values
are not vastly different. This indicates somewhat similar chemical activity for the metals.
However the atom packing with BCC chromium is not the same as in the FCC Copper. Therefore, we
would not expect extensive solid solubility between the two elements. The order of expected solid
solubility in copper based on radius difference and crystal structure, along with the observed
experimental values is as follows:

Element Observed Atomic Percent


Soluble
Nickel (best) 100
Aluminium 17
Chromium (Poorest) <1

The importance of considering atomic percentage rather than weight percentage can be seen if 17 at
percent aluminium is converted to wt. Percent.

Basis: 100 atoms


17 atom Al 83 atoms Cu

17atoms x 26.98 g/at. wt.


Weight of 17 aluminium atoms =
6.02 x 1023 atoms/at. wt.

83 atoms x 63.54 g/at. wt.


Weight of 83 copper atoms =
6.02 x 1023 atoms/at.wt

Therefore,

56
Structure of Metals

17atoms x 26.98 g/at. wt


6.02 x 1023 atoms/at.wt.
Weight Percent Al = x 100
17atoms x 26.98 g/at. wt 83 atoms x 63.54 g/at. wt.

6.02 x 1023 atoms/at.wt 6.02 x 1023 atoms/at.wt

= 8.0
Since aluminium has a low atomic weight compared to copper, the weight percent is much less than
the atomic percentage. Nickel however, is close to copper in atomic weight, and values for the atomic
and weight percentages would be very close.

It is important to note that while the Hume-Rothery rules are a very good guide to solid solubility, a
number of exceptions to these rules exist. However their accuracy is, in general, great enough to
justify their use.
Ordered and Disordered Solid Solutions
The substitutional solid solutions may be either ordered or disordered (Fig. 2.38). In disordered solid
solutions the different kinds of atoms occupy the available sites at random. In the ordered solid
solutions, atoms of one kind arrange themselves in an orderly and periodic manner on one set of
atomic sites; the atoms of the other kind do likewise on another set. Thus, in an ordered alloy, the
lattice may be regarded as being made up of two or more interpenetrating sub-lattices, each containing
different arrangements of atoms. Such a coherent atomic scheme extends over large distances, i.e. the
crystal possesses long-range order. Complete ordering occurs with concentrations of the solid solution
elements corresponding to simple atomic ratios of the components of the type AB or AB3 (For
example, AuCu and Au Cu3 in gold-copper alloys).

(a) (b)
Fig. 2.38: Ordered and Disordered Solid Solution
(a) Disordered Solution (b) Ordered Solution

Five different types of structures are usually found in binary alloys. Four of these are derived from
the cubic structures (two from the fcc and two from bcc) and one from the hcp structure. Ordering
in the fcc lattice may be described by considering the lattice as four interpenetrating simple cubic

57
Introduction to Crystallography

sub-lattices of equal size. The lattice points at the cube corners of the fcc unit cell constitute one
sub-lattice and each pair of points at the centres of the two opposite faces generate one sub-lattice
in combination with similar lattice points on the neighbouring unit cells. When one of the four sub-
lattices is occupied by a particular type of atoms, the resulting structure is known as L12. The L12
structure corresponds to the ideal composition A3B. For example, the ordered Cu3Au is generated
when one of the four sub-lattices is occupied by only gold atoms [Fig. 2.39(a)].
When two of these sub-lattices are occupied by a particular type of atoms the structure is known as
L10. This structure consists of alternate layers of A and B atoms parallel to the (001) planes. The
L10 structure [Fig. 2.39(b)] of ordered CuAu is generated when two of the sub-lattices are
occupied by gold atoms.
The bcc structure, on the other hand, may be visualised as two interpenetrating simple cubic sub-
lattices. The lattice points at the corners of the bcc unit cell constitute one sub-lattice and those at
cell centres constitute the second sub-lattice. The ordered DO3 structure [Fig. 2.40(a)] is produced
when half of the lattice sites on one of the sub-lattices is occupied by atoms of one kind. If all the
lattice sites on one of the sub-lattices is occupied by atoms of one kind the B2 (L20) structure of
ordered CuZn is produced [Fig. 2.40(b)]. Ordering in hcp structures may also be described in
terms of interpenetrating sub-lattices.

Au
Cu

(a) (b)
Fig. 2.39: Ordering in Face Centred Cubic lattice
(a) LI2 Structure (b) LI0 Structure

Cu Zn

Fe Al
(a) (b) 58

Fig. 2.40 : Ordering in Body Centred Cubic Lattice


(a) DO3 Structure (b) L2o Structure
Structure of Metals

Order does not occur in dilute solutions. This is because the mutual repulsion of similar electrostatic
charges causes the solute atoms to be isolated from each other and be surrounded by the atoms of the
solvent metal. As the concentration of the solute atoms increases, the average distance between them
diminishes and their repulsion for each other increases. Moreover, the solute atoms have attraction for
solvent atoms. This creates a situation very much like that occurring in an ionic crystal, in which order
always appears in the arrangement of the unlike atoms.

Interstitial Solid Solution


Interstitial solid solutions arise when the solute atoms are small enough to fit into the spaces between
the atoms of the solvent lattice and this leads to a very stringent size factor rule for these solutions. In
metallic structures the sizes of the interstitial holes are very small compared with the atomic diameters
of the metal atoms. The atoms of the common engineering metals are very nearly alike in size and they
do not form interstitial solid solution with each other. Only the non-metallic elements such as
hydrogen, oxygen, nitrogen, boron and carbon have atomic diameters small enough to allow any
interstitial dissolution. This can occur during solidification and in many cases, it also occurs when the
parent metal is already solid. Thus, carbon can form an interstitial solid solution with face centred
cubic iron during the solidification of steel, but it can also be absorbed by solid iron provided the latter
is heated to a temperature at which the structure is face-centred cubic. This is the basis of carburising
steels. Nitrogen is also able to dissolve interstitially in solid steel, making the nitriding process
possible; similarly the atoms of hydrogen dissolve in a number of solid metals and produce brittleness.
Typical metallic lattices have at least two different kinds of interstices that can be occupied by
additional atoms. The correlation between the basic structures, the types of voids being filled, the
degree of occupancy of such voids, and the various types and sizes of participating atoms constitute
one of the more fascinating aspects of crystal chemistry of alloy phases.

(a)

59
(b)
Introduction to Crystallography

There are two kinds of voids that occur in all closest packing. If the triangular void in a close
packed layer has a spheres directly over it, there results a void with four spheres around it [Fig.
2.41(a)]. The four spheres are arranged on the corners of a tetrahedron and such a void is called a
tetrahedral void. On the other hand, if a triangular void pointing up in one closest packed layer is
covered by a triangular void pointing down in the layer above it, then a void surrounded by six
spheres results [Fig. 2.41(b)]. These six spheres are arranged on the corners of an octahedron, and
such a void is called an octahedral void. These are the only kinds of voids that can occur in a
close-packing despite the fact that number of different close packing possible is infinite.

Fig. 2.42: Number of Voids Surrounding a Sphere in Closest packing

A sphere in a hexagonal close packed layer A is surrounded by six-triangular


voids of two kinds, B and C. When the next close packed layer above is added,
(say a B layer) than the three B voids become tetrahedral voids and three C voids
become octahedral voids. If the added layer is a C layer the C voids become
tetrahedral voids and the B voids become octahedral voids. Similarly the closest
packed layer below the A layer gives rise to three tetrahedral and three octahedral
voids. Furthermore the particular sphere in layer A being considered itself covers
a triangular void in the closest packed layer above and in the layer below the
sphere. Thus two more tetrahedral voids surrounds the sphere. This results in 2 x 3
= 6 octahedral voids and 2 x 3 + 1 + 1 = 8 tetrahedral voids surrounding the
sphere. Since the total number of spheres and voids in a closest packing is very

60
Structure of Metals

large, it is possible to determine only the average number of voids of each kind
belonging to a sphere. Each octahedral void is surrounded by six spheres and each
sphere is surrounded by six octahedral voids. The number of octahedral voids
belonging to one sphere is given by the ratio
Number of octahedral voids around a sphere 6
= =1
Number of spheres around a void 6

Each tetrahedral void is surrounded by four spheres and each sphere is surrounded
by eight voids. The number of tetrahedral voids belonging to one sphere is given
by the ratio
Number of tetrahedral voids around a sphere 8
= =2
Number of spheres around a void 4

The number of octahedral voids in a closest packing, therefore, is equal to the


number of spheres. Similarly the number of tetrahedral voids is a closest packing
is twice the number of spheres or twice the number of
octahedral voids present.

COORDINATION OF VOIDS
The voids in a closest packing can be described in
terms of the spheres that are arranged about each
void. The number of spheres surrounding a void is
called the coordination number of that void. If the
void is also represented by a sphere, then the
collection of spheres coordinating the sphere
representing the void is called a coordination Fig. 2.43: A Tetrahedron
polyhedron. Accordingly, a tetrahedral void has a
coordination number of 4. The resulting coordination polyhedron is a tetrahedron.
The tetrahedron in Fig. 2.43 can be used to determine the radius of the sphere representing the
tetrahedral void. Fig. 2.44(a) shows the tetrahedron inscribed inside a cube. The shaded diagonal
plane of the cube in Fig. 2.44(a) is shown enlarged in Fig. 2.44(b).

O 2
Q Q O

N 1

P
L L 61
M P
(a) (b)

Fig. 2.4: Determination of Sphere Representing a Tetrahedral Void


Introduction to Crystallography

It is easy to see in this figure that the right triangle LMN is similar to the right triangle LPO. It
follows, therefore, that

LM LP 2
= = .
LN LO 3

Let r be the radius of the sphere representing the void and R the radius of the spheres in a closest
packing. Then,

LM R 2
= = .
LN Rr 3

The radius of the sphere representing the tetrahedral void is obtained by rearranging the terms in.

3 2
r= R
2
= 0.225R.
N
N

1 1
M
L L M
2

(a) (b)

O O
Fig. 2.45: Determination of Sphere Representing
an Octahedral Void

An octahedral void is surrounded by six spheres, and its coordination number, therefore, is 6. If the
octahedral void is represented by a sphere, the resulting polyhedron is an octahedron, as shown in
Fig. 2.45. Only that part of each coordinating sphere that belongs to the octahedral void is shown
in Fig. 1.46(a). Since six spheres, each contributing one-sixth of its volume, coordinate the

62
Structure of Metals

octahedral void, one sphere in a closest packing belongs to the octahedral void. This octahedron
can be used to determine the radius of the sphere representing an octahedral void.
The plane passing through four coordinating spheres and the central sphere representing the void is
the square shown in Fig. 1.46(b). In the isosceles right triangle LMN
LM 2
= .
LN 1

Letting r be the radius of the sphere representing the octahedral void and R the radius of the
spheres in a closest packing, it follows that

2R  2r
= 2
2R

r= 2 R-R

= 0.414R.

63
Introduction to Crystallography

Voids in Face Centred Cubic Structure


In the face centred cubic structures the two types of interstitial voids are identified
in Fig.2.46. The larger void known as the octahedral void, are surrounded by six
atoms situated at the corners of a regular octahedron [Fig.2.46(a)]. The centres of
the voids are at the mid-points of the unit cell edge, 00½, etc, and at the centre ½,
½, ½. Smaller voids known as tetrahedral voids), are surrounded by four atoms
[Fig. 2.46(b)] with centres at positions such as ¼, ¼, ¼ etc.
a 3/4

a 2
a 2

(a) (b)
Metal Atoms Metal Atoms
Fig. 2.46: The Interstitial Voids in FCC Structure
(a) Octahedral Voids (b) Tetrahedral Voids

The maximum radius r of the sphere that fits into a tetrahedral void is 0.225R
while the maximum radius that fits into an octahedral void is equal to 0.414 R
where R represents the radius of the spheres in close packing.

(a) (b)
Metal Atoms Metal Atoms

Fig. 2.47: Interstitial Voids in Close Packed Hexagonal Structures


(a) Octahedral Voids (b) Tetrahedral Voids

64
Structure of Metals

The cph structure also has two types of interstitial voids as shown in Fig. 2.47. As in the FCC structure
the octahedral voids are larger than the tetrahedral holes.
Body Centred Cubic Packing
There are essentially two kinds of voids present in this type of packing also. One of these occurs
at the centres of the cube faces and is co-ordinated by four atoms at the corners of the face and two
atoms at the body centres of the two cubes sharing the face. Fig. 2.48(a) shows a view of the
irregular octahedron thus formed. The maximum radius of a sphere that fits into this void is
0.154R, where R is the radius of the coordinating spheres. Since there are six such spheres
coordinating each void and 18 voids coordinating each sphere, the ratio of spheres to voids is 1:3.

a 3 /2 a 3 /2 a 5 /4
a/2 a/ 2

(a) (b)

Fig. 2.48 : Interstitial Voids in Body centred Cubic Structures


(a) Octahedral Voids (b) Tetrahedral Voids
The second type of voids that occur in body centered cubic structure is in the tetrahedrally
coordinated position shown in Fig. 2.48(b). The largest sphere that can fit into this void has a
radius of 0.291R. There are 24 such voids surrounding each sphere and four spheres around each
void so that the ratio of spheres to voids is 1:6.
The tetrahedral void in a body centered cubic structure are irregular tetrahedral
and are larger than irregular octahedral voids. The irregular tetrahedral voids are
smaller than the regular octahedral voids in closest packing. This factor accounts
for the different kinds of interstitial atoms that can be accommodated in these two
types of structures.
An interstitial solid solution has essentially the crystal structure of the solvent, which is always the
element having the larger atomic diameter.
Interstitial solid solutions can form only if the electrochemical factor between the metal and the non-
metal atom is very small. The transition group metals are not strongly electropositive and so
interstitial solutions can form. With the more electropositive metals compounds having the properties

65
Introduction to Crystallography

similar to ionic compounds are formed and the amount of interstitial solubility is negligible. Interstitial
solid solutions are usually of little importance. Carbon in iron is, however, an exception and forms the
basis of hardening of steels by heat treatment.

PROBLEMS AND QUESTIONS


1. What is a space lattice? How many space lattices are possible? How many crystal
structures are possible? What is the distinction, if any, between a space lattice and a
crystal structure?
2. Illustrate, with the help of neat sketches, the difference between the unit cells of body
centred cubic, face centred cubic and hexagonal close packed crystals.
3. Show that for body centred cubic and face centred cubic crystal structures the lattice
constants are given by
a b c c = 4 r / 3 and a f c c = 4 r / 2 where r = atomic radius.

4. Calculate the density of packing for face centred cubic structure. How does this compare
with the analogous calculation for hexagonal close packed structures?
5. Show that a base centred tetragonal structure cannot be considered as a separate and
discrete arrangement of atoms. [Hint: Show that a simple tetragonal structure can be
curved out of two adjacent base centred tetragonal structure].
6. Show that the face centred tetragonal Bravais lattice is equivalent to body centred
tetragonal structure. How does the lattice parameters of a and c of the face centred
tetragonal structure compare with the a’ and c’ of the body centred tetragonal?
[Hint: Draw 2 face centred tetragonal structures side by side and see if a body centred
tetragonal structure can be constructed along the plane separating the two cells. The
heights of the two cells are the same (c = c’} and the lattice parameter of bct is equal to
half the face diagonal of fct]
7. Explain the term co-ordination number. Find the co-ordination numbers for body centred
cubic and face centred cubic structures.
8. Below are listed the atomic weight, density and atomic radius for three hypothetical
alloys. For each determine whether its crystal structure is FCC, BCC or SC and then
justify your determination.
Alloy Atomic Weight Density Atomic Radius
(g/mol) (g/cm3) (nm)

66
Structure of Metals

A 77.4 8.22 0.125


B 107.6 13.42 0.133
C 127.3 9.23 0.142

9. For the HCP crystal structure show that the ideal c/a ratio is 1.633. Assume hard sphere
model of atoms.
10. Assume hard ball model and calculate the atomic packing factor (packing density) for a
simple cubic structure.
11. Zinc has an HCP crystal structure, a c/a ratio of 1.856 and a density of 7.13 g/cm3.
Calculate the atomic radius for Zinc.
12. How many atoms/mm2 are there on the (100) and (111) planes of lead (fcc, ao = 1.75 Å )?
Ans: (100) = 8.2 x 1012 atoms/mm2, (111) = 9.5 x 1012 atoms/mm2
13. Zinc has an hcp structure. The height of the unit cell is 4.94 Å. The centers of the atoms
in the base of the unit cell are 2.665 Å apart. What is the volume of the unit cell? (b)
Would the calculated density be greater or less than the actual density of 7.135 g/cm3?
Justify your answer.
Ans: (a) 9.1 x 10-23 cm3 (b) 7.15 g/cm3
14. Titanium is bcc in its high temperature form. The radius increases 2 percent when bcc
changes to hcp during cooling. What is the percentage volume change?
15. A face centred lattice may be presented alternately as a rhombohedral lattice. (a) How do
the two lattices compare? (b) What is the axial angle? (c) What is the ratio of the unit cell
volumes? (d) Why do we prefer the fcc over the rhombohedral?

Ans: (a) afcc = arh 2 (b) 60o (c) Vfcc = 4Vrh (d) Cubic lattice is more symmetrical.
16. Calculate the linear density (atoms/Å) of atoms in the [100], [110] and [111] directions in
body centred cubic iron (ao = 2.86 Å).
17. How many atoms are there per square millimeter (i) on a (100) plane of copper; (ii) on a
(110) plane and (iii) on a (111) plane?
Ans: (i) 15.3 x 1012 Cu/mm2 (ii) 10.8 x 1012 Cu/mm2 (ii) 17.7 x 1012 Cu/mm2
18. The distance between (110) planes in a body centred cubic structure is 2.03 Å. (a) What is
the size of the unit cell ? (b) What is the radius f th atoms? (c) What might the metal be?
Ans: (a) 2.87 Å (b) 1.24 Å (c) bcc iron or Chromium.
19. Calculate the linear density (atoms/Å) of atoms in the [100], [110] and [111] directions in
face centred cubic copper (ao = 3.62 Å).

67
Introduction to Crystallography

20. Calculate the planar density of atoms (atoms/ Å2) in body centred cubic iron (ao = 2.86 Å)
in the (100), (110) and (111) planes.
21. Calculate the density of body centred cubic iron at room temperature from the atomic
radius of 1.24 Å. Compare your result with the experimental value of 7.87 g/cm3.
22. The lattice parameter of body centred cubic iron is 2.86 Å and the density is 7.87g/cm3.
Calculate the atomic weight.
23. If the atomic radius of lead is 0.175 nm, calculate the volume of the unit cell in cubic
meters.
[Ans. Vol = 1.21 x 10 -28 m3]
24. Calculate the radius of palladium atom, given that Pd has a face centred cubic structure, a
density of 12.09 g/cm3 and an atomic weight of 106.4 g.mol.
[Ans. Radius = 0.138 nm]
25. Zieconium has an HCP crystal structure and a density of 6.51 g/cm3. (a) What is the
volume of the unit cell? (b) If c/ ratio is 1.593, calculate the values of c and a.
[Ans: (a) V = 1.40 x 10 -28 m3, (b) a = 323 nm, c = 0.515 nm]
26. Calculate the size of the largest atom that can fit interstitially in a BCC structure (as a
fraction of the radius of the BCC atoms). The positions of the largest interstitial holes
have co-ordinates of the type 0. ½, ¼ in the BCC unit cell.
27. Calculate the size of the largest atom that can fit interstially into an FCC copper crystal
without destroying it. The lattice parameter for copper is 3.61 Å.
28. What are Miller indices? Of what significance or use they are? In what manner are they
determined?
29. Discuss the similarities and differences between the Miller indices of the planes and the
directions.
30. Determine the Miller indices of the family of close packed planes in FCC structure.
Determine the direction indices of the family of close packed directions in FCC crystals.
31. Determine the direction of the close packed directions in BCC structure.
32. In a cube six faces constitutre a family of planes, or plane of a form. Do the six faces of
the tetragonal cell also constitute planes of a form? If not, how many families are there?
33. We have a theoretical element of atomic weight 64.09 which is face centred cubic. An
experimental density has been determined as 8.41 g/cm3.
(a) Calculate the approximate atomic radius of Q.
(b) Why is the calculation only approximate?

68
Structure of Metals

34. One of your associates has calculated the density of body-centred cubic iron as given
below (atomic radius = 1.24 Å and atomic weight = 55.85).
4 atoms /unit cell x 55.85 g/at.wt
6.02 x 10 23 atoms /at.wt
Density = = 68.8 g/cm3
2 (1.24 x 10 8 cm)
[ ]3
2
Since materials do not have densities this high, there must be errors. What are these errors?
(Assume that the error is not a result of having pushed the wrong buttons on the calculators.)
35. When heated to 1398oC, the structure of iron changes from face centred cubic (ao = 3.68
Å) to body cented cubic (ao = 2.926 Å). Does a given mass of iron contract or expand in
volume as the transformation from face centred to body centred cubic takes place?
Calculate the percent change in volume that takes place during this transformation.
36. Show that the hexagonal close packed unit cell is not a separate space lattice, but rather a
special case of hexagonal lattice. [Hint: Compare the equivalency, or lack there of, for an
atom in the middle layer with one in the upper or lower layers.]

69
Chapter – 3

Structure of Nonmetallic Materials

In Chapter 2 we discussed the structure of metallic materials. In this chapter we shall discuss the
structure of nonmetallic material. The atomic bonding in nonmetallic materials ranges from purely
ionic to totally covalent; many materials exhibit a combination of these two types of bonding.

IONIC CRYSTALS
Typical ionic solids contain discrete charged ions separated by regions of negligible electron
density. These ions are held together by strong electrostatic forces, which account for the large
binding energies, high melting points, hardness and incompressibility of ionic solids. The simplest
examples of ionic solids are the compounds formed by electropositive metals, such as sodium or
potassium with electronegative nonmetals such as chlorine and oxygen. Such compounds mostly
contain simple monatomic ions which pack together as if they were spheres of different sizes.
Hence the crystal structures of these substances are fairly simple, as they are determined by two
factors: the maintenance of electrical neutrality and the relative sizes of the ions.

Electrical Neutrality
If the charges on the anion and the cation are identical, the compound has the formula MX, and the
coordination number for each ion is identical to assure a proper balance of charge. Examples of
compounds having the formula MX include CsCl and NaCl. Although CsCl is a useful example of
Introduction to Crystallography

a compound structure, it does not represent any commercially important ceramics. By contrast the
NaCl structure is shared by many important ceramic materials.
If the valence of the cation is +2 and that of the anion is –1, then twice as may anions must be
present, and the formula is of the form MX2. The structure of the MX2 compound must assure that
the coordination number of the cation is twice the coordination number of the anion. For example,
each cation may have 8 anion nearest neighbors, while only 4 cations will touch each anion.
Included in the MX2 category is the perhaps the most important ceramic compound SiO2.

Ionic Radii
Ionic solids consist of cations and anions. In ionic bonding some atoms lose their outer electrons to
become cations and others gain electrons to become anions. Thus the cations are normally smaller
than the anions they bond with. The number of anions that surround a central cation in an ionic
solid is called the coordination number and corresponds to the number of nearest neighbours
surrounding a central cation. For stability as many anions as possible surround a central cation.
However the anions must make contact with the central cation.

Table 3.1 : Ionic Radii for Several Cations and Anions

Cation Ionic Radius (nm) Anion Ionic Radius (nm)

Ba2+ 0.136 Br - 0.196

Ca2+ 0.100 Cl - 0.181

Cs+ 0.170 F- 0.133

Fe2+ 0.077 I- 0.220

Fe3+ 0.069 O2- 0.140

K+ 0.138 S2- 0.184

Mg2+ 0.072

Mn2+ 0.067

Na+ 0.102

Ni2+ 0.069

Si 4+ 0.040

Ti 4+ 0.061

70
Structure of Nonmetallic Materials

The ratio of the radius of the central cation to that of the surrounding anion is called the radius
ratio. The radius ratio when the anions just touch each other and contact the central cation is called
the critical (minimum) radius ratio. The radius ratio influences both the manner of packing and the
coordination number. Interstitial atoms whose radii are slightly larger than the radius of the
interstitial site may enter that site, pushing the surrounding atoms slightly apart. However, atoms
whose radii are smaller than the radii of the hole are not allowed to fit into the interstitial site
because the central cation can rattle around in the cage of anions.
The packing geometry of one type of cation and one type of anion with the cation as the smaller
ion is a function of the ion sizes. This can be worked out from space filling geometry, when the
following conditions corresponding to a stable configuration are satisfied simultaneously.
(i) anions and cations considered as hard spheres always touch each other
(ii) anions generally will not touch but may be close enough to be in contact with one another and
(iii) as many anions as possible surround a central cation for the maximum reduction of
electrostatic energy.
When the cation is very small compared to the anion, only two anions can be neighbours to the
cation in order to satisfy the above three conditions.
From the simple geometry it can be shown that the ratio of the cation to anion radius r c/ra for the
configuration in which the three surrounding anions are touching one another and also the central
cation (Fig. 3.1) is 0.155. This triangular configuration is one of the critical configurations and the
radius ratio is said to be a critical value, because for smaller values of r c/ra not all the three anions
would touch the central cation [Fig. 3.1(b)].

(a) (b) (c)


Fig. 3.1: Triangular Coordination of Anions around a Central Cation
(a) The Critical Condition (b) Unstable Configuration (c) Stable but not Critical

This violates condition (i) above and leads to instability. Here, the only way to satisfy all three
conditions is to reduce the number of anions to two. For values of r c/ra greater than 0.155, all the
anions do not touch one another. But all three conditions are still satisfied. This situation will
prevail till the radius ratio is increased to 0.225, the next higher critical value corresponding to a

71
Introduction to Crystallography

tetrahedral (four) coordination. The possible values of coordination number and radius ratio ranges
in which they are stable are listed in Table 3.2.

Table: 3.2: Coordination Number as a Function of Radius Ratio

Coordination Number Range of Radius Ratio Configuration

2 0 – 0.155 Linear

3 0.155 – 0.225 Triangular

4 0.225 – 0.414 Tetrahedral

6 0.414 – 0.732 Octahedral

8 0.732 – 1.00 Cubic

12 1.0 FCC or HCP

The three conditions for stability is not always valid. If the directional characteristics of bonding
persist to any significant degree (as in partial covalency) the considerations based on close packing
geometry alone will fail. The nearest neighbour interactions can also affect the local packing
geometry. However, in a number of cases the coordination rules outlined above are obeyed.

Example 3.1: Find the critical radius ratio for a triangular coordination.

Solution:
In Fig. 3.2 triangle ABC is an equilateral triangle and line AD bisects angle CAB. Thus angle
DAE = 30o. Thus
AD = R + r
AE R
cos 30o = = = 0.866
AD Rr
R = 0.866(R + r)
= 0.866R + 0.866r
0.866r = R – 0.866R = R(0.134)
r Fig. 3.2 : Determination of Critical Radius for
= 0.155
R Triangular Coordination

72
Structure of Nonmetallic Materials

Example 3.2: Find the critical radius ratio for tetrahedral coordination
P

Q R
T
U
S

(a) (b)
P

Fig. 3.3: Determination of Critical Radius Ratio for Tetrahedral Coordination

Solution
At the critical condition in this case, four anions at the four corners of a tetrahedron touch one
another and also the cation at the centre of the hole between. Referring to Fig. 3.3, the normal from
the apex to the base of the tetrahedron PT is 2 a/ 3 , where a is the side of the tetrahedron. Here

a = 2 ra
The centre of the tetrahedral hole lies at three-fourths of the distance from the apex. Therefore,

rc + ra = ¾ PT = ¾ x 2 / 3 x 2 ra

= 1.225 ra
rc/ra = 0.225

If the interstitial atom becomes too large, it prefers to enter a site having a larger coordination
number. Therefore, an atom whose radius ratio is between 0.225 and 0.414 will enter a tetrahedral
site; if the radius is somewhat larger than 0.414, it will enter an octahedral site. When the atoms
have the same size, as in pure metals, the radius ratio is one and the coordination number is 12,
which is the case for metals with FCC and HCP structures.
Ionic solids cannot form close packed structures like FCC or HCP because the atoms are not of the
same size. The charge on the ions requires an alternating arrangement of anions and cations. The
type of structure is mainly determined by the packing of the larger anions which are normally the
negative ions. The structure frame work is made by anions within which the small cations (positive
ions) fit interstitially. Three common ionic crystal structures are the sodium chloride and caesium
chloride and zinc blende structures.

73
Introduction to Crystallography

Sodium Chloride Structure


The radius ratio for sodium and chloride ions is rNa /rCl = 0.97/1.81 = 0.536; the sodium ion has a
charge of + 1, while the chloride ion has a charge of – 1. Therefore, based on the charge balance
and radius ratio, each anion and cation must have a coordination number of six. The FCC
structure, with Na cations at FCC positions and Cl anions at the four octahedral sites, will satisfy
these requirements [Fig. 3.4(a)]. We can also consider this structure to be FCC with two ions – one
Na and one Cl – associated with each lattice point. Many ceramics, including MgO, CaO, and FeO
have this structure.

Cl- (a) Cl- (b)


Na+ Cs+
Fig. 3.4: Ionic Crystal Structures
(a) Unit Cell of NaCl (b) Unit Cell of CsCl

Cesium Chloride Structure


Cesium chloride (CsCl) is an ionic structure showing two interpenetrating simple cubic lattices,
one for each kind of ion [Fig. 3.4(b)]. The radius ratio, rCs /rCl = 1.67/1.81 = 0.92, dictates that a
cation is surrounded by eight anions. We can characterise the structure as a simple cubic structure
with two ions – one Cs and one Cl – associated with each lattice point. This structure is possible
when the anion and the cation have the same valence.

Example 3.3: Predict the coordination number for the ionic solids CsCl and NaCl. Use the
following ionic radii for the prediction:
Cs+ = 0.170 nm Na+ = 0.102 nm Cl- = 0.181 nm

Solution
The radius ratio for CsCl is
r(Cs  ) 0.170 nm
= = 0.94
R (Cl ) 0.181nm

74
Structure of Nonmetallic Materials

This ratio is greater than 0.732, CsCl should show cubic coordination (CN = 8), which it does.
The radius ratio for NaCl is
r(Na  ) 0.102nm
= = 0.56
R (Cl ) 0.181nm

Since this ratio is greater than 0.414 but less than 0.732, NaCl should show octahedral
coordination (CN = 6), which it does.

Example 3.4: Calculate the ionic packing factor for CsCl. Ionic radii are Cs+ = 0.170 nm and Cl- =
0.181 nm.

Solution:
The ions touch each other across the cube diagonal of the CsCl unit cell, as shown in Fig. 3.5.

Fig. 3.5: Cube Diagonal of CsCl Unit Cell

Let r = Cs+ ion and R = Cl- ion. Thus

√3a = 2r + 2R
= 2(0.170 nm + 0.181 nm)
a = 0.405 nm

75
Introduction to Crystallography

4 3 4
π r (l Cs  ion)  π R 3 (l Cl  ion)
CsCl ionic packing factor = 3 3
3
a

4 4
π (0.170 nm)3  π (0.181nm)3
= 3 3
(0.405 nm)3

= 0.68

Example 3.5: Show that MgO can have the sodium chloride crystal structure and calculate the
density of MgO. [Given: rMg = 0.66 and rO = 1.32 A]

Solution

rMg 0.66
= = 0.50
rO 1.32

Since 0.414 < 0.50 < 0.732, the coordination number is six and the NaCl structure is possible.
The atomic weights are 24.3 and 16 g/g. mole for magnesium and oxygen, respectively. The ions
touch along the edge of the cube, so

a0 = 2rMg + 2ro = 2(0.66) + (1.32) = 3.96 A

(4 Mg ions) (24.3)  (4 0 ions) (16)


ρ =
(3.96 x 10 8 cm) 3 (6.02 x 10 23 )

= 4.31 Mgm-3

Example3.6: For KCl, (a) verify that the compound may have the cesium chloride structure and
(b) calculate the packing factor for the compound. [Given: rK = 1.33 A and rCl = 1.81 A]

Answer

rk 1.33
(a) = = 0.735
rCl 1.81

Since 0.732 < 0.735 < 1.000, the coordination number is eight and the CsCl structure is likely.

76
Structure of Nonmetallic Materials

(b) The ions touch along the body diagonal of the unit cell, so

3 a 0 = 2 rK + 2 rCl = 2 (1.33) + 2 (1.81) = 6.28A

a 0 = 3.63A

4π 3 4π 3
rK (1 K ion)  rCl (1 Cl ion)
Packing factor = 3 3
a 30

4π 4π
(1.33) 3  (1.81)3
= 3 3 = 0.725
(3.63) 3

Example 3.7: Calculate the density of NaCl from its crystal structure, the ionic radii of Na+ and
Cl- ions, and the atomic masses of Na and Cl. The ionic radius of Na+ = 0.102 nm and that of Cl =
0.181 nm. The atomic mass of Na = 22.99 g/mol and that of Cl = 35.45 g/mol.

Solution
The Cl- ions in the NaCl unit cell form an FCC-type atom latice, and the Na+ ions occupy the
interstitial spaces between the Cl- ions [Fig. 3.6(a)].

(a) (b)

Fig. 3.6: (a) NaCl Unit Cell lattice Points Showing the Positions of the Ions.
(b) Octahedral Coordination of Cl-anions around a central Sodium Cation.

There is equivalent of one Cl- ion at the corners of the NaCl unit cell since 8 corners × 1/8 ion, and
there is the equivalent of three Cl- ions at the faces of the NaCl unit cell since 6 faces ½ ion = 3 Cl-
ions, making a total of four Cl- ions per NaCl unit cell. To maintain charge neutrality in the NaCl

77
Introduction to Crystallography

unit cell, there must also be the equivalent of four Na+ ions per unit cell. Thus there are four Na+
Cl- ion pairs in the NaCl unit cell.

Fig. 3.7: Cube face of NaCl Unit Cell


(Ions in contact with the cube edges)

To calculate the density of the NaCl unit cell, we shall first determine the mass of one NaCl unit
cell and then its volume. Knowing these two quantities, we can calculate the density m/V.

The mass of a NaCl unit cell is

(4Na   22.99 g/mol)  (4Cl   35.45 g/mol)


m=
6.02  10 23 atoms (ions)/mol

= 3.88 × 10-22 g

the volume of the NaCl unit cell is equal to α3, where α is the lattice constant of the NaCl unit cell.
The Cl- and Na+ ions contact each other along the cube edges of the unit cell, as shown in Fig. 3.7,
and thus
α = 2 ( rNa  R Cl  )

= 2 (0.102 nm + 0.181 nm) = 0.566 nm


= 0.566 nm × 10-7 cm/nm = 5.66 × 10-8 cm
V =a3 = 1.81 × 10-22 cm3
The density of NaCl is
m 3.88  10 22 g g
p= = = 2.14
V  22
1.81 10 cm 3
cm 3
The handbook value for the density of NaCl is 2.16 g/cm3.

78
Structure of Nonmetallic Materials

Example 3.8: Calculate the linear density of Ca2+ and O2- ions in ions per nanometer in the [110]
direction of CaO which has the NaCl structure. (Ionic radii: Ca2+ = 0.106 nm and O2- = 0.132 nm.)

Solution:
The length of the [110] distance across the base face of a unit cube is √2α, where a is the length of
a side of the cube or lattice constant [Fig. 3.8(b)].

Fig. 3.8: (a) NaCl Lattice-points Unit Cell (b) The [110] Direction

From the cube face of the NaCl unit cell (Fig. 3.7), we see that a = 2r + 2R.
Thus, for CaO,
a = 2( rCa 2  + R O 2  ) = 2 (0.106 nm + 0.132 nm) = 0.476 nm

The linear density of the O2- ions in the [110] direction is


2 2
2O 2O
ρL = = = 2.97 O2-/nm
2a 2(0.47nm)

The linear density of Ca2+ ions in the [110] direction is also 2.97Ca2+/nm if we shift the origin of
the [110] direction from (0, 0, 0) to (0, ½, 0),
Thus, there are 2.97(Ca2+ or O2-)/nm in the [110] direction.

Example 3.9: Calculate the planar density of Ca2+ and O2- ions in ions per square nanometer on
the (111) plane of CaO (NaCl structure, Ionic radii; Ca2+ = 0.106 nm and O2- = 0.132 nm.)

Solution:
If we consider, the anions (O2- ions) to be located at the DCC positions (Fig. 3.9) for the Cl- ions of
then the (111) plane contains the equivalent of two anions.

79
Introduction to Crystallography

[ 3 × 60o = 180o = ½ anion + (3 ×½) anions at each midpoint of the sides of the (111) planar
triangle of Fig. 3.9(b) = a total of 2 anions within the (111) triangle].

Anions or
cations

(a) (b)

Fig. 3.9: (a) NaCl Unit Cell lattice Points Showing the Positions of the Ions.
(b) The (111) Plane in the Unit Cell

The lattice constant for the unit cell a = 2(r + R) = 2(0.106 nm + 0.132 nm) = 0.476 nm. The
planar area A = ½ bh, where h = 3 /2 a2. Thus,

A = (½ √2α)(√ 3 /2 a) = 3 /2 a2 = √ 3 /2 (0.476 nm)2 = 0.196 nm2

The planar density for the O2- anions is

2(O 2  ions)
= 10.2 O2- ions/nm2
0.196nm 2
The planar density for the Ca2+ cations is the same if we consider the Ca2+ to be located at the
FCC lattice points of the unit cell, and thus
ρ planar (CaO) = 10.2(Ca2+ or O2-)/nm2.

Zinc Blende Structure


The zinc blende structure has the chemical formula ZnS. Although the Zn ions have a charge of
+2 and S has a charge of –2, zinc blende (ZnS) cannot have the sodium chloride structure because

rZn /rS = 0.74/1.84 = 0.402.

This radius ratio demands a coordination number of four, which in turn means that the sulfide ions
will enter tetrahedral sites in a unit cell. The FCC structure, with Zn cations at the normal lattice
points and S anions at one-half of the tetrahedral sites, can accommodate the restrictions of both
charge balance and coordination number.

80
Structure of Nonmetallic Materials

(a) (b)

Zn S Ti4+ Ba2+ O2-

Fig. 3.10: (a) Crystal Structure of Cubic Zinc Sulphide


(b) Perovskite Crystal Structure

The zinc blende structure [Fig. 3.10(a)] may be regarded as two interpenetrating(a)face centred cubic
lattices of the elements, with the corner of one located at the position of ¼, ¼, ¼ of the other.
There are four molecules of ZnS per conventional cell. About each atom there are four equally
distant atoms of the opposite kind arranged at the corners of a regular tetrahedron.
It is also possible for ceramic compounds to have more than one type of cation, BaTiO3 having
both Ba2+ and Ti4+ cations falls into this classification. This material has a perovskite crystal
structure. The unit cell of this structure is shown in Fig. 3.10(b); the Ba2+ ions are situated at all
eight corners of the cube and a single Ti4+ is at the cube centre, with O2- ions located at the centre
of each of the six faces.

Example 3.10 : Based on charge balance and radius ratio, which of the cubic structures we have
discussed would BeO most likely have? [Given: rBe = 0.35, rO = 1.32]

Solution
The radius ratio is
rBe 0.35
= = 0.265
ro 1.32

This radius ratio requires a coordination number of four. Of the cubic structures we have
examined, only the zinc blende structure, with half of the tetrahedral sites filled, satisfies both the
coordination number and charge balance requirements.

81
Introduction to Crystallography

Example 3.11: Calculate the density of zinc blends (ZnS). Assume the structure to consist of ions
and that the ionic radius of Zn2+ = 0.060 nm and that of S2- = 0.174 nm.

Solution:

Fig. 3.11: ZnS Structure showing Relationship between Lattice


Constant and radii of Sulphur and Zinc Atoms/Ions

massofunitcell
Density =
volumeofunitcell
There are four zinc ions and four sulfur ions per unit cell. Thus Mass of unit cell

(4Zn 2   65.37g/mol)  (4S 2   32.06g/mol)


=
6.02 10 23 atoms/mol
= 6.47 × 10-22 g
Volume of unit cell = α3

From Fig. 1.54,


3
α = rzn2+ + Rs2- = 0.060 nm + 0.174 nm = 0.234 nm
4
α = 5.40 × 10-8 cm
α = 1.57 × 10-22+ cm3
mass 6.47  10 22 g
Thus, Density = =  22 3
= 4.12 g/cm3
volume 1.57 10 cm
The handbook value for the density of ZnS (cubic) is 4.10 g/cm3.

82
Structure of Nonmetallic Materials

Representing Crystals in projections: Crystal Plans


The more complicated the crystal structure and the larger the unit cell, the more difficult it is to
visualize the atom or ion positions from diagrams or photographs of models, atoms or ions may be
hidden behind others and therefore not seen. Another form of representation, the crystal plane or
crystal projection, is needed, which precisely shows the atomic or ionic positions in the unit cell.
The first step is to specify the axes x, y and z from a common origin and along the sides of the cell.
The atomic or ionic positions or coordinates in the unit cell are specified as fractions of the cell
edge length in the order x, y, z. Thus in the bcc structure the coordinates of the atom/ion at the
origin is (000) and that at the centre of the cube is (½ ½ ½). As all eight corners of the cube are
equivalent positions i.e., any of the eight corners can be chosen as the origin, (000) specifies all the
corner atoms, and the two coordinates (000) and (½ ½ ½) are equal to the two atoms/ions in the
bcc unit cell. In the fcc, with four atoms/ions per unit cell, the coordinates are: (000), (½ ½ 0),
(½ 0 ½), (0 ½ ½).

(a) (b)

Fig. 3.12: Plan of (a) BCC Structure (b) CCP or HCP Structure

Fig. 3.13: Alternate Unit Cells of the Perovskite Strucutre

Crystal projections or plans are usually drawn perpendicular to the z-axis, and Fig. 3.12 are plans
of the bcc and fcc structures respectively. Note that only the z-coordinates are indicated in these
diagrams; the x and y coordinates need not be written down because they are clear from the plan.
Similarly, no z coordinates are indicated for all the corner atoms because all eight corners are
equivalent positions in the structure, as mentioned above.

83
Introduction to Crystallography

Sketching crystal plans helps one to understand the similarities and differences between structures,
in fact, it is very difficult to understand them otherwise. For example, Fig. 3.13 (a) and (b) show
how the same crystal structure (perovskite, CaTiO3). They look different because the origins of the
cell have been chosen to coincide with different atoms/ions.

COVALENT CRYSTALS
Elements from the central groups of the Periodic table are not readily ionised. The energy required
to remove all the valence electrons is too large for ionic bonding to be possible. However, it is still
possible for each atom to complete its outer shell by sharing electrons with its neighbours. Thus in
covalent bonding arrangement each atom has 8 – N neighbours.
The constituent atoms are linked by covalent bonds in one, two or three dimensions. The covalent
bonds themselves are strong, but they only confer strength and resistance to compression in the
directions in which they occur. Hence, the mechanical properties of covalent solids depend
markedly on whether the covalent bonds extend in one, two or three dimensions. Covalent solids
do not form close packed structures because the directional nature of the covalent bonds must be
maintained. The simplest covalent structure is that of diamond (and of silicon and germanium).

(a) (b)

Fig. 3.14: (a) Covalent Bonding: Atoms of Group IV each has Four Neighbours.
(b) Crystal Structure of Diamond Showing Tetrahedral Bond Arrangement:

Covalently bonded materials frequently must have complex structures in order to satisfy the
directional restraints imposed by the bonding.

Diamond Cubic Structure


Elements such as silicon, germanium, and carbon in its diamond form are bonded by four covalent
bonds and produce a tetrahedron (Fig. 3.14). The coordination number for each silicon atom is
only four, due to the nature of the covalent bonding.

84
Structure of Nonmetallic Materials

As these tetrahedral groups are combined, a large cube can be constructed [Fig. 3.14(b)]. This
large cube contains eight smaller cubes that are the size of the tetrahedral cube; however, only four
of the cubes contain tetrahedra. The large cube is the diamond cubic, or DC, unit cell. The lattice is
a special FCC structure. The atoms on the corners of the tetrahedral cubes provide atoms at each of
the regular FCC lattice points. However, four additional atoms are present within the DC unit cell
from the atoms in the center of the tetrahedral cubes.
We can describe the DC lattice as an FCC lattice with two atoms associated with each lattice point.
Therefore, there must be eight atoms per unit cell.

Example 3.11: Determine the packing factor for DC silicon.

Solution
The atoms touch along the body diagonal of the cell (Fig. 3.15). Although atoms are not present at
all locations along the body diagonal, there are voids that have the same diameter as atoms.

a0
r
2r
2r

r 2r

8r = 3 a0

Fig. 3.15: Relation between Lattice parameter and


Atomic Radius in Diamond Cubic
Consequently,

3 a0 = 8r

Packing Factor
4
(8 atoms/cell) ( π r 3 )
= 3
a 30
4
(8) ( π r 3 )
= 3 = 0.34
(8 r/ 3 ) 3

85
Introduction to Crystallography

The low pressure crystal form of carbon is graphite which differs in many respects from diamond.
It is a layer structure in which the atoms are covalently bonded to only three nearest neighbours in
the layer, and the layers themselves are only weakly bonded by van der Waals forces (Fig. 3.14).

Fig. 3.14: The Layer Structure of Graphite

Since there is one electron per atom that is not employed in the covalent bonding there is an
abundant supply of free electrons, and so graphite is a good electrical conductor whereas diamond
is an almost perfect insulator. However, the free electrons cannot cross easily from layer to layer
and so the resulting conductivity is highly anisotropic. It was believed that graphite owed its
lubricating properties to the weak van der Waals bonds between adjacent layers. It has now been
shown that lubrication is associated with adsorbed gas layers such as oxygen or organic molecules.

CRYSTALLINE POLYMERS
Polymers involve molecules rather than just atoms or ions, as with metals and ceramics. The
atomic arrangement is, therefore, more complex for polymers. We think of polymer crystallinity as
the packing of molecular chains so as to produce an ordered atomic array.
As a polymer crystallizes from the molten state, a certain amount of amorphous polymer will also
be present within the structure (to accommodate mismatch and other defects within the crystal).
For his reason most crystalline polymers are actually semicrystalline, in which the percentage of
crystalline phase can range from 40 to 90 percent.
Polymer properties are significantly affected by the degree of crystallinity. Crystalllinity tends to
increase the mechanical properties such as tensile strength and hardness while diminishing
ductility, toughness and elongation.

86
Structure of Nonmetallic Materials

The density of a crystalline polymer will be greater than an amorphous one of the same material
and molecular weight, since the chains are more densely packed together for the crystalline
structure.
The degree of crystallinity by weight may be determined from accurate density measurements
according to
ρ c (ρ s  ρ a }
% Crystallinity = x 100
ρ s (ρ c  ρ a)

Where, ρ s = the density of specimen for which the percent crystallinity is to be determined
ρ c = density of the perfectly crystalline polymer
ρ a = density of the totally amorphous polymer.

The values of ρ c and ρ a must be measured by other experimental means.

Mer Unit

Fig. 3.15: The Unit Cell of Crystalline Polyethylene

The dashed lines in Fig. 3.15 outline the unit cell for the lattice of polyethylene. Polyethylene is
obtained by joining C2H4 molecules together to produce long polymer chains that form an
orthorhombic unit cell. Some polymers, including nylon, can have several polymorphic forms.
The degree of crystallinity may range from completely amorphous to almost entirely (up to 90
percent crystalline); by way of contrast metal specimens are almost always entirely crystalline.
The degree of crystallinity of a polymer depends on the rate of cooling during solidification as
well as on the chain configuration. During crystallization upon cooling through the melting
temperature, the chains which are highly random and entangles in viscous liquid, must assume an
ordered configuration. For this to occur, sufficient time must be allowed for the chains to move
and align themselves.
On the other hand, crystallization is not easily prevented in chemically simple polymers such as
polyethylene, even for very rapid cooling rates. For linear polymers, crystallization is easily
accomplished because there are virtually no restrictions to prevent chain alignment. Any side
branches interfere with crystallization, such that branched polymers never are highly crystalline; in
fact excessive branching may prevent any crystallization whatsoever.

87
Introduction to Crystallography

For copolymers as a general rule, the more irregular and random the mer arrangement, the greater
is the tendency for the development of noncrystallinity. For alternating and block copolymers there
is some likelihood of crystallization. On the other hand, random and graft copolymers are normally
amorphous.

Example 3.12: How many carbon and hydrogen atoms are in each unit cell of crystalline
polyethylene? There are twice as many hydrogen atoms as carbon atoms in the chain. .ρ of
polyethylene= 0.95 Mgm-3.

Solution
If we let x be the number of carbon atoms, then 2x is the number of hydrogen atoms.
(x) (12 g/g. mole)  (2x) (1g/g. mole)
p=
(7.41 x 10 -8 ) (4.94 x 10 8 ) (2.55 x 10 8 ) (6.02 x 10 23 )

14 x
0.95 =
56.2

x = 3.8 ≃ 4 carbon atoms per cell

2 x = 8 hydrogen atoms per cell

To some extent, the physical properties of polymeric materials are influenced by the degree of
crystallinity. Crystalline polymers are usually stronger and more resistant to dissolution and
softening by heat.
Polymer single crystals as thin platelets and having chain folded structures may be grown from
dilute solutions. Many semicrystalline polymers form spherulites. Spherulites are considered to be
polymer analog of grains in polycrystalline metals and ceramics. Each spherulite is really
composed of many different lamellar crystals and in addition some amorphous material.
Polyethylene, polypropylene, polyvinyl chloride and nylon form a spherulite structure when they
crystallize from a melt.

PROBLEMS AND QUESTIONS


3.1 Show that the minimum cation to anion ratio for a coordination number of 4 is 0.225.
3.2 Show that the minimum cation to anion ratio for a coordination number of 6 is 0.414.

88
Structure of Nonmetallic Materials

[Hint: Use the NaCl structure and assume that anions and cations are just touching along the
cube edges and across face diagonals.]
3.3 Show that the minimum cation to anion ratio for a coordination number of 8 is 0.732.
3.4. Using the ionic radii given in Table 3. , determine the coordination number expect6ed for
the following compounds.
(a) FeO, (b) CaO (c) SiC (d) PbS (e) B2O3
[ Ans. (a) NaCl (d) CsCl ]
3.5. Using the ionic radii given in Table 3. , determine the coordination number expect6ed for
the following compounds.
(a) Al2O3 (b) TiO2 (c) MgO (d) SiO2 (e)
3.6 Based on the ionic radius ratio and the necessity for charge balance, which of the cubic
structures would you expect CdS to possess?
3.7 Based on the ionic radius ratio and the necessity for charge balance, which of the cubic
structures would you expect CoO to possess?
3.8 The compound NiO has the sodium chloride crystal structure. Calculate (a) the lattice
parameter (b) density and (c) packing factor for NiO.
3.9 One of the forms of BeO has the ZnS structure at high temperature. Determine (a) the lattice
parameter, (b) the density and (c) the packing factor for the compound.
3.10 (a) Germanium has diamond cubic structure with a lattice parameter of 5.6575A. Calculate
the size of he germanium atom in the unit cell.
(b) Does this best match up with germanium’s atomic radius or ionic radius?
3.11 Calculate the fraction of [111] direction and the fraction of the (111) plane actually covered
by atoms in the diamond cubic unit cell.
3.12 Calculate the fraction of [111] direction and the fraction of the (111) plane actually covered
by sodium ions in the NaCl cubic unit cell.
3.13 Compute the atomic packing factor for the CsCl crystal structure in which the radius ratio is
0.732.
3.14 The ZnS crystal structure is one that may be generated from close packed planes of anions.
(a) Will the packing sequence for this structure be FCC or HCP? Why?
(b) Will cations fit tetrahedral or octahedral positions? Why?
(c) What fraction of positions will be occupied?
[ Ans. (a) FCC (b) Tetrahedral (c) one=half]

89
Introduction to Crystallography

3.15 Iron oxide (FeO) has the rock salt crystal structure and a density of 5.70 g/cm3.
(a) Determine the unit cell edge length.
(b) How does the result compare with the edge length as determined from radii in Table 3.
assuming that Fe 2+ and O 2- ions just touch each other along the edges.
[Ans. (a) a = 0.437 nm (b) a = 434 nm]
3.16 A hypothetical AX type of ceramic material is known to have a density of 2.65 g/cm3 and a
unit cell of cubic symmetry with a cell edge length of 0.43 nm. The atomic weights of the A
and X elements are 86.6 and 40.3 g/mol respectively. On the basis of the information, which
of the following crystal structures ia (are) possible for this material: rock salt, cesium
chloride or ZnS? Justify your choice(s).
[ Ans. Cesium Chloride]
3.17 Compute the atomic packing factor for the diamond cubic crystal structure. Assume that the
bonding atoms touch one another, that the angle between adjacent bonds is 109.5o, and that
each atom internal to the unit cell is positioned a/4 of the distance away from the two
nearest cell faces (a is the unit cell edge length).
[ Ans. APF = 0.68]

90
Chapter – 4

Symmetry and Crystal Structure

In the earlier chapters we developed an understanding of simple crystal structures by first


considering the ways in which atoms or ions could pack together and then introducing smaller
atoms or ions into the interstices between the larger ones. This is a practical approach as it not only
provides us with an immediate and straightforward understanding of the atomic/ionic
arrangements in some simple compounds, but also suggests the ways in which more complicated
compounds can be built up.
However, it is not a systematic and rigorous approach. This is because all the possibilities of
atomic arrangements in all crystal structures are not explored. The systematic approach is to
analyse and classify the geometrical characteristics of patterns to arrive at a completely general
description of all the patterns to which atoms or molecules might conform in the crystalline state.
The theory of internal structure of three dimensional crystal is basically the same as two
dimensional crystals. The structure of three dimensional crystal is, however, more complicated
because of the extra dimension. We shall begin with a discussion of the internal structure of two
dimensional crystals. Following the treatment in two dimensions we shall explore three
dimensional crystallography.
Introduction to Crystallography

TWO-DIMENSIONAL PATTERNS AND LATTICES


Let us consider the symmetry of the letters R, M, and S. R is asymmetrical. M consists of two
equal sides, each of which is a reflection or mirror image of the other; there is a mirror line of
symmetry down the centre indicated by the letter m, thus Mm. There is no mirror line in the S, but
if it is rotated 1800 about a point in its centre, an identical S appears; there is a two-fold rotation
axis usually called a diad axis at the centre of the S. This is represented by a little lens-shape at the
axis of rotation: S

Unit Cell
(a) (b) (c)
Fig. 4.1: (a) A Pattern with the Motif R (b) A Pattern with the Lattice Points Indicated
and (c) the Lattice and Unit Cell Outlined

Fig. 4.1(a) is made up of the letter R repeated indefinitely. R is asymmetric in shape and may
represent anything, a two dimensional molecule, a cluster of atoms or whatever. R is called the
motif. These motifs may be considered to be situated at or near the intersections of an (imaginary)
grid. The grid is called the lattice and the intersections are called lattice points.
The lattice points may be above, below, to one side, in the ‘middle’ of the motif – the only
requirement is that the same position with respect to the motif is chosen every time. In Fig. 4.1(b)
a position a little below the motif has been chosen as the lattice points.
Now there are an infinite number of ways in which the lattice points may be ‘joined up’ (i.e. an
infinite number of ways of drawing a lattice or grid of lines through lattice points). In practice, a
grid is usually chosen which ‘joins up’ adjacent lattice points to give the lattice as shown in Fig.
4.1(c), and a unit cell of the lattice may also be outlined. Clearly, if we know (1) the size and shape
of the unit cell and (2) the motif which each lattice point represents, including its orientation with
respect to the lattice point, we can draw the whole pattern or build up the whole structure
indefinitely.
The unit cell of the lattice and the motif, therefore, define the whole pattern or structure. This is
very simple: but observe an important consequence. Each motif is identical and, for an infinitely
extended pattern, the environment of each motif is identical. This provides us with the definition of
a lattice (which applies equally in two and three dimensions): a lattice is an array of points in space
in which the environment of each point is identical.

92
Symmetry and Crystal Structure

It should be remembered that the condition ‘the environment of each point is identical’ is
important. For example, the points in Fig. 4.2 extend infinitely. Of these only (a) and (d) constitute
a lattice; in (b) and (c) the points are certainly in a regular array, but the surroundings of each point
are not all identical.

(a) (b)

(c) (d)

Fig. 4.2: Patterns of Points. Only (a) and (d) Constitute Lattices

Fig. 4.2(a) and (d) represent two two-dimensional lattice types, named oblique and rectangular
respectively in view of the shapes of their unit cells. But in effect the rectangular lattice is just a
special case of the oblique lattice, i.e. with a 90o angle. The distinction arises from different
symmetries of the two lattices. Thus we require to extend our notions of symmetry and to classify
a series of symmetry elements. This precise knowledge of symmetry can then be applied to both
the motif and the lattice and will show that there are a limited number of patterns with different
symmetries (only seventeen) and a limited number of two-dimensional lattices (only five).

TWO-DIMENSIONAL SYMMETRY ELEMENTS


A pattern has the property of being produced by some basic design repeatedly undergoing a series
of prescribed spatial motions. A symmetry operation is an operation that can be performed either
physically or imaginatively that results in no change in the appearance of an object. In other
words, the various operations that can be performed up on an object which result in bringing it into
coincidence with its initial position are known as symmetry operations. The fundamental
symmetry operations are (i) rotation about an axis, (ii) reflection about a plane and (iii) rotation
about an axis combined with an inversion (rotary-inversion). Any grouping whose elements can be
related by repeating the same spatial operation is said to be symmetrical.

93
Introduction to Crystallography

Rotational Symmetry
If an object can be rotated about an axis and repeats itself every 90o of rotation then it is said to
have an axis of 4-fold rotational symmetry. The axis along which the rotation is performed is an
element of symmetry referred to as a rotation axis. The following types of rotational axis are
possible in crystals.

Fig. 4.3: Rotational Symmetry

1-fold rotation axis – an object that requires rotation of a full 360o in order to restore it to its
original appearance has no rotational symmetry. Since it repeats itself 1 time every 360o it is said
to have a 1-fold axis of rotational symmetry.
If an object appears identical after a rotation of 180o, that is twice in a 360o rotation, then it is said
to have a 2-fold rotation axis (360o/180 = 2). In these examples, the axes we are referring to are
imaginary lines that extend toward you perpendicular to the page. A filled oval shape represents
the point where the 2-fold rotation axis intersects the page.
Objects that repeat themselves upon rotation of 120o are said to have a 3-fold axis of rotational
symmetry (360o/120 = 3) and they will repeat three times in a 360o rotation. A filled triangle is
used to symbolize the location of 3-fold rotation axis.
Similarly an object with a 4-fold axis of rotational symmetry will repeat itself four times in a 360o
rotation (and is represented by a filled square) while an object that repeats itself six times in a 360o
rotation is said to have 6-fold symmetry (a filled hexagon is used as a symbol). No other rotation
axes are possible in crystals.

Mirror Symmetry
A mirror symmetry operation is an imaginary operation that can be performed to reproduce an
object. The operation is done by imagining that you cut the object in half, then place a mirror next
to one of the halves of the object along the cut. If the reflection in the mirror reproduces the other
half of the object, then the object is said to have a mirror symmetry. The plane of the mirror is an
element of symmetry referred to as a mirror plane and is symbolised with the letter m.

94
Symmetry and Crystal Structure

Generation of Motifs with Different Symmetries


The concept of symmetry can be developed by using an asymmetrical ‘object’ (say R). Successive
additions of different mirror lines and axes of symmetry will cause R to be repeated to form
different patterns of groups. The different patterns of Rs which are produced correspond to objects
or projections of molecules (i.e, ‘two-dimensional molecules’) with different symmetries which
are not possessed by the R alone.

(a) (b) (c)

(d) (e)

Fig. 4.4: Generation of Motifs with Different Symmetries (a) Cis-difluoroethane (b)
Ethane (c) Trans difluoroethane (d) Trialkylfluoride Ammonium and (e) Carbonate Ion

For example, in Fig. 4.4(a) ‘right-‘ and ‘left-‘ handed Rs are reflected in the ‘vertical’ mirror line
between them. This pair of Rs has the same mirror symmetry as the letter M, or the projection of
the cis-difluoroethene molecule. A reflection across a mirror line generates a new motif that has
the opposite handedness (i.e., a left hand is turned in to a right hand). It is impossible to
superimpose the motifs that are symmetric ally related by reflection if the motif movements are
confined to their common plane. Such patterns made up of non-congruent pairs (one left hand and
one right hand) is said to be enantiomorphous. Now if we add another ‘horizontal’ mirror line as in
Fig. 4.4(b). A group of four Rs – two right and two left handed – is produced. This group has the
same symmetry as the single letter O or the projection of the ethane molecule. The R may also be
repeated with a diad (two-fold rotation) axis [Fig. 4.4 (c)]. The two Rs – both right handed – have

95
Introduction to Crystallography

the same symmetry as the letter S, or the trans-difluoroethene molecule. Rotation does not change
the motif’s handedness (a pattern made by rotating a left hand about an axis will be composed of
all left hands). Such motifs are said to be congruent as distinct from pairs of motifs created by
reflection. If the operation of rotation results in a congruent set, it is called a proper rotation and
the rotation axis is called a proper rotation axis.
In Fig. 4.5 (a) and (c) [or (b) and (d)] are related by reflection through m1 and m2, by two fold
rotation or inversion through the central point. A left hand remains a left hand, and (a) and (c) are
congruent. But (a) and (b) are related by reflection only – a left hand is turned into a right hand.
The objects are enantiomorphous.

(a) (b)

(c) (d)

Fig. 4.5: Symmetry Operations in Two Dimensions

If we look back to the group of Rs in Fig. 4.4(b) we can notice that they are also related by a diad
(two-fold rotation axis) at the intersection of the mirror lines: the action of reflecting the Rs across
two perpendicular mirror lines ‘automatically’ generates the two-fold symmetry as well. This
effect, where the action of one symmetry element generates another, is quite general.

Example 4. 1: Symmetry of Rectangle


Show that a rectangle can only have two planes of mirror symmetry, one vertical and one
horizontal. It does not have mirror symmetry along the diagonal lines.

Solution
The rectangle on the left has a mirror plane [Fig. 4.6(a)] that runs vertically on the page and is
perpendicular to the page. The rectangle on the right has a mirror plane that runs horizontally and

96
Symmetry and Crystal Structure

is perpendicular to the page. The Dashed parts of the rectangles below show the part of the
rectangles that would be seen as a reflection in the mirror.

Fig. 4.6(a): Mirror Symmetry of a Rectangle

m?

Fig. 4.6(b): Mirror Symmetry of a Rectangle

The rectangle does not have mirror symmetry along the diagonal lines. This is illustrated in Fig.
4.5(b). If we cut the rectangle along a diagonal such as that labeled m? and reflected the lower half
in the mirror, then we would see what is shown by the dashed lines in the right side of the diagram.
Since this does not reproduce the original rectangle, line m? does not represent a mirror plane.

97
Introduction to Crystallography

Example 4. 2: Symmetries of an Equilateral Triangle.

Fig. 4.7: An Equilateral Triangle.

Solution
If we rotate the depicted equilateral triangle 1200 clockwise about its mid-point, the resulting
image will cover the original completely. The image occupies the same space as did the original.
All internal distances and angles have remained what they were. But all its points, except its mid-
point have been swapped with respect to each other.
Thus we can describe the just mentioned rotation by a swapping of the points indicate in Fig. 4.7.
With respect to our rotation point 3 will be replaced by 1, 2 will be replaced by 3, and 1 will be
replaced by 2. If, on the other hand, we rotate the triangle 2400 it will also cover the original. It
again occupies the same space as before. But this time: 3 will be replaced by 2, 2 will be replaced
by 1, and 1 will be replaced by 3
If we, finally rotate it 3600 we have executed an operation which is in fact equivalent to not having
it rotated at all. But this time: 3 will be replaced by 3, 2 will be replaced by 2, and 1 will be
replaced by 1. So the equilateral triangle has three rotational symmetries, one of which -- rotation
of 0 or 360 degrees - is the identical transformation, in which no points are swapped.
But the equilateral triangle has still more symmetries.
We can detect three mirror lines (reflection lines), each one going from a tip to the center of the
opposite side. When we reflect the triangle with respect to one such reflection line we swap the
two halves of the triangle divided by that line.
So the equilateral triangle has six symmetries, which are rigid transformations such that the
original is exactly covered by the image produced by that transformation: three rotational
symmetries (one of which is the identical transformation) and three reflectional symmetries.

98
Symmetry and Crystal Structure

Fig. 4.4(d) shows the R related by a triad (three-fold) axis. The projection of the tri-alkylfluoride
ammonium ion also has this same symmetry. If we add a ‘vertical’ mirror line [as in Fig. 4.4(e)]
three more (left-handed) Rs are generated, and at the same time, the R are mirror-related not just
in the vertical line but also in two lines inclined at 600 as shown. This is another example of
additional symmetry element – in this case mirror lines -- being automatically generated. This
procedure - of generating groups of R’s which represent motifs with different symmetries - may
be repeated with tetrad (four-fold) and hexad (six-fold) rotation axes of symmetry.

Example 4.3: Possibility to have five-, seven-, eight-, etc., fold rotation axes of symmetry.

Fig. 4.8: Possibility of having a Pentagon or an Octagon


as a Characteristic Motif

An array consists of motifs (with the characteristic symmetry of the array) placed adjacent to each
other. In order to fill the space continuously (a consequence of the translation periodic
requirement) the included angles of the characteristic motifs must fit together at a point such that
adjacent motifs exactly surround the point with no overlap or void.
In the case of equilateral polygon as possible motifs, only those polygons with the included angles
that are submultiples of 2π radians are acceptable where the included angles are given in terms of
the number of sides by π (1 - 2/n). The objects themselves may appear to have 5-fold or 8-fold
rotation axes. However, we can not combine objects with 5-fold and 8-fold apparent symmetry in
such a way that they will completely fill the space (Fig. 4.8). That is why 5-fold or 8-fold rotation
axes are not possible in crystals. Thus pentagons or octagons do not constitute characteristic
motifs.

99
Introduction to Crystallography

Altogether there are ten such symmetries in two dimensions: 6 unique symmetries (1, 2, 3, 4, 6,
and m) and 4 combinations (2mm, 3m, 4mm, and 6mm). They are known as plane point groups, so
called because all the symmetry elements pass through one point (i.e. at least one point is common
to all symmetry elements). The ten plane point groups are labeled with symbols which indicate the
symmetry elements present; m for one mirror plane, mm or mm2 for two mirror planes (plus diad),
2 for a diad, 3 for a triad, 3m for a triad and three mirror planes and so on.
In some cases short hand symbols are used. For example, in Fig. 4.4(b) two mirror planes intersect
in a two-fold axis. In this case, two of the elements are sufficient to define the whole, and this
particular point group is normally given the short symbol mm, rather than the full symbol 2mm or
mm2. Similarly, three mirror planes meeting in a three-fold axis [Fig. 4.4(e)] are adequately
represented by 3m. The full symbol 3 mmm is not needed because two of the m’s are redundant,
being created by the action of the three-fold axis on the other one.

THE FIVE PLANE LATTICES


We can now determine the number of possible two-dimensional or plane lattices. The important
criterion for this determination is that the lattice itself must possess the symmetry of the motif; it
may possess more symmetry elements but it cannot possess fewer. In other words, the symmetry
of the arrangement of lattice points around each lattice point must be the same as, or greater than,
the symmetry of the motif. In Fig. 4.2(d) each lattice point and half-way between there are vertical
and horizontal lines of symmetry intersecting diad axes. The motif of Fig. 4.4(b) also has this
symmetry and, therefore, it follows that a pattern with such a motif will have a rectangular lattice.
The motif of Fig. 4.4(a) has one mirror line, and a pattern with this motif will also have a
rectangular lattice -- in this case the symmetry of the lattice is greater than that of the motif. On the
other hand, the motifs with one or two mirror lines of symmetry [Figs 2.4(a) and (b)] cannot occur
in a pattern with the oblique lattice because the oblique lattice does not itself have any mirror lines
of symmetry.
This procedure can be applied to all the other motifs. For example, a motif with tetrad (four-fold)
symmetry applies to a pattern with a square lattice and a motif with triad (three-fold) symmetry
[Figs 2.4(d) and (e)] and hexad (six-fold) symmetry will apply to a pattern with an hexagonal
lattice. Altogether, five two-dimensional or plane lattices may be worked out, as shown in Fig. 4.9.
They are described by the shapes of the unit cells which are drawn between lattice points --oblique
p, rectangular p, rectangular c (which is distinguished from rectangular p by having an additional
lattice point in the centre of the cell), square p and hexagonal p.
All two-dimensional patterns must be based upon one of these five plane lattices; no others are
possible. There is no doubt that a large number of unit cell shapes are possible, but the pattern of
lattice points which they describe will be one of the five of Fig. 4.7. For example, the rectangular c
lattice may also be described as a rhombic p or diamond p lattice, depending upon which unit cell
is chosen to ‘join up’ the lattice points (Fig. 4.10). These are just two alternative descriptions of
the same arrangement of lattice points.

100
Symmetry and Crystal Structure

Oblique - p

Rectangular - p

Rectangular - C

Square - p

Hexagonal - p

Fig. 4.9: Unit Cell of Five Planer Lattices showing (right) the
Symmetry Elements present

Thus the choice of unit cell is arbitrary: any four lattice points can be joined up to form a unit cell.
Usually we choose a unit cell that is as small as possible - or ‘primitive’ (symbol p)-which does
not contain other lattice points within it. Sometimes a larger cell is more useful because the axes
joining up the sides are at 900. Examples are the rhombic or diamond lattice which is identical to

101
Introduction to Crystallography

the rectangular centred lattice described above. [An important three-dimensional case: the cubic
cell is used to describe the ccp (cubic close packed) structure in preference to the primitive
rhombohedral cell.

Fig. 4.10: The Rectangular C lattice, showing the Alternative Primitive


(Rhombic P or Diamond P) Unit Cell

There are ten point group symmetries that a motif can possess (five of which are shown in Fig.
4.8) and it may be thought that there are therefore, only ten different types of two dimensional
patterns distributed among the five plane lattices.

Fig. 4.11: Reflection and Translation can Lead to an


Operation Called Glide.

However, the combination of operations can give rise to an additional symmetry element. For
example, the combined operation of translation and reflection give rise to what is known as glide
(Fig. 4.11).

102
Symmetry and Crystal Structure

In Fig. 4.12 both the patterns have a rectangular lattice. In Fig. 4.12(a) the motif has mirror
symmetry and consists of a pair of left and right handed Rs.

(a) (b)

Fig. 4.12: Patterns with (a) Reflection Symmetry and (b) Glide Reflection Symmetry.
[The mirror lines (m) and glide lines (g) are indicated]

In Fig. 4.12(b) there is still a reflection - still pairs of right-and left-handed Rs - but one set of Rs
has been translated, or glided half a lattice spacing. This symmetry is called a reflection-glide or
simply a glide line of symmetry. It is important to remember that this glide plane is displaced
from the original mirror of the point group. The translation must be half the cell edge so that after
two operations of a glide, the motifs are returned to the same configuration as the next lattice
point. Glide lines give seven more two-dimensional patterns, giving seventeen in all - the
seventeen plane groups.

THREE DIMENSIONAL SYMMETRY ELEMENTS


The structure of three dimensional crystal is basically the same as that of two dimensional crystals.
However, the number of three-dimensional patterns are many more than the two-dimensional
patterns. Fortunately, it is not essential to work out all the three-dimensional patterns
systematically: a clear understanding of the principles that are involved, the operation and
significance of the additional symmetry elements and the main results are usually adequate.
The additional point symmetry elements required in three dimensions are centres of symmetry,
mirror planes (instead of lines) and inversion axes; the additional translation symmetry elements
are glide planes (instead of lines) and screw axes.

103
Introduction to Crystallography

THREE DIMENSIONAL SYMMETRY ELEMENTS


The structure of three dimensional crystal is basically the same as that of two dimensional crystals.
However, the number of three-dimensional patterns are many more than the two-dimensional
patterns. Fortunately, it is not essential to work out all the three-dimensional patterns
systematically: a clear understanding of the principles that are involved, the operation and
significance of the additional symmetry elements and the main results are usually adequate.
The additional point symmetry elements required in three dimensions are centres of symmetry,
mirror planes (instead of lines) and inversion axes; the additional translation symmetry elements
are glide planes (instead of lines) and screw axes.

Centre of Symmetry
Fig. 4.13 shows two motifs symmetrically related by the operation of inversion. In this operation
[Fig. 4.13(a)] lines are drawn from all points on the object through a point in the centre called a
centre of symmetry or inversion centre (a point element symbolized by the letter i). The lines each
have lengths that are equidistant from the original points. When the ends of the lines are connected
the original object is reproduced inverted from the original appearance. In Fig. 2 only a few such
lines are drawn for the small triangular face. Fig. 4.13(b) shows the object without the imaginary
lines that reproduced the object.

(a) (b)

Fig. 4.13: Inversion through a centre of Symmetry

If an object has only a centre of symmetry, we say that it has a 1-fold rotoinversion axis. Crystals
that have a centre of symmetry will exhibit the property that if you placed it on a table there will
be a face on the top of the crystal that will be parallel to the surface of the table and identical to the
face resting on the table.

104
Symmetry and Crystal Structure

Glide Planes and Screw Axes


We have seen that in two dimensions translational symmetry elements or glide lines arise. In three
dimensions also it is possible to have symmetry elements in which a translation is involved. Two
important hybrid operations containing translation elements are shown in Fig.2.14. The first is a
combination of proper rotation and a translation along the axis [Fig. 4.14(a)]. Here the symmetry
element is called a screw axis since the motif follows a spiral path as shown. The action of a
driving screw is an example of an object possessing this type of symmetry (a bolt is really a better
example than a screw since most screws taper to a point).

(a)

(b)

Fig. 4.14: Hybrid Symmetry Operations with Translation


(a) Screw Operation (b) Glide Reflection

Screw axes are noted by a number n, where the angle of rotation is 360o/n. The degree of
translation is then added as a subscript showing how far along the axis the translation is, as a
portion of the parallel lattice vector. For example, 21 is a 180o (two-fold) rotation followed by a
translation of ½ of the lattice vector. Typically the magnitude of the translation part of the screw
location is equal to the lattice translation parallel to the screw axis divided by the fold of the
rotation part of screw rotation. Thus a six-fold screw axis would translate a motif 1/6 of the lattice
translation parallel to the screw axis and rotate the motif 60o about the screw axis. Screw rotation
is a congruent operation. However, the screw axis may be left-handed or right-handed.
A combination of a translation with a reflection operation, called a glide reflection results in a
glide plane illustrated in Fig. 4.14(b). Because a glide plane combines the operation of reflection
with that of translation it occurs only in extended arrays.

105
Introduction to Crystallography

Fig. 4.15 shows an aerial view of a boat rowed by eight crews. It is obviously related to a mirror
plane, but each rower in Fig. 4.15 is the mirror image of one rowing immediately in front or
behind. Any figure is related to the next by moving one place along the boat and then reflecting
across a mirror plane.

Fig. 4.15: An Ariel View of Eight Rowers Showing a Translational Symmetry Operation;
Each Rower is Related to the next by a Combination of Translation and Reflection.

Example 4.3: The symmetry content of the crystal with rectangular sides and square ends.

A4
m
m
m m

A2
m A2

A2
A2

Fig. 4.16: Symmetry of a Crystal with Rectangular Sides and Square Ends

Solution
The crystal (Fig. 4.16) has the following symmetry elements:
(1) 1 – 4-fold rotation axis (A4)
(2) 4 – 2-fold rotation axes (A2), two cutting the faces and two cutting the edges.

106
Symmetry and Crystal Structure

(3) 5 mirror planes (m), 2 cutting across the faces, 2 cutting through the edges, and one
cutting horizontally through the centre.
(4) There is a centre of symmetry (i).
The symmetry content of this crystal is thus: i, 1 A4, 4 A2, 5m.

Combination of Symmetry Operations


Isolated objects or groups of objects may show any number of mirror planes and any kind of axis;
the symmetry of an infinite array of identical groups, such as is found in crystals, is limited by
having to pack the units together in three dimensions. In fact, in crystals there are 32 possible
combinations of symmetry elements and thus, there are 32 crystal classes. These 32 symmetry
classes can be derived from point-group theory, for they constitute the possible ways to arrange
equivalent points symmetrically around a given point in space. Accordingly they are also called
point groups. The 32 crystal classes are divided among the seven crystal systems in such a way
that each system has a certain minimum of symmetry elements as shown in Table 4.1.

Table 4. 1: Minimum Symmetry Elements in Different Crystal Systems


System Minimum Symmetry Elements
Triclinic None (only one one-fold or one-fold roto-inversion axes)
Monoclinic A single 2-fold rotation axis or a single plane
Orthorhombic Two perpendicular planes or three mutually perpendicular 2-fold axes of
rotation.
Tetragonal A single 4-fold axis of rotation or of rotation-inversion.
Rhombohedral A single 3-fold axis of rotation or of rotation-inversion.
Hexagonal A single 6-fold axis of rotation or of rotation-inversion.
Cubic Four 3-fold rotation axes (along cube diagonals)

Some crystals may possess more than the minimum symmetry elements required by the system to
which they belong, but none may have less.
The non-translational symmetry elements are combined into point group symbols that describe the
symmetry of finite groups. The addition of translation greatly increases the possibilities, so that
instead of 32 point groups 230 space groups (of which two-dimensional analogue is the plane
group) are needed to describe the symmetries of infinite arrays. We have already seen that in two-
dimension a combination of five plane lattices with the appropriate point and translation symmetry
elements gives rise to 17 possible two-dimensional patterns. The point group of a crystal describes
the translation free residue of the total symmetry of that crystal described by its space group.

107
Introduction to Crystallography

Hermann-Mauguin (International) Symbols


There are 32 possible combinations of symmetry operations that define the external symmetry of
crystals. These 32 possible combinations result in the 32 crystal classes. The Hermann-Mauguin
symbols (also called the international symbols) are used to describe the crystal classes from the
symmetry content. The procedure to derive the Hermann-Mauguin symbols will now be discussed
with the aid of some examples.

Example 4.4: Hermann-Mauguin Symbol of a Rectangular Block


The rectangular block (Fig. 4.15) has 3 (three) 2-fold rotation axes (A2), 3 mirror planes (m), and a
center of symmetry (i). The rules for deriving the Hermann-Mauguin symbol are as follows:
A2

m A2

A2
Fig. 4.15: Symmetry Symbols of a Rectangular Block

1. Write a number representing each of the unique rotation axes present. A unique rotation axis
is one that exists by itself and is not produced by another symmetry operation. In this case, all
three 2-fold axes are unique, because each is perpendicular to a different shaped face, so we
write a 2 (for 2-fold) for each axis
2 2 2
2. Next we write an "m" for each unique mirror plane. Again, a unique mirror plane is one that
is not produced by any other symmetry operation. In this example, we can tell that each
mirror is unique because each one cuts a different looking face. So, we write:
2m2m2m
3. If any of the axes are perpendicular to a mirror plane we put a slash (/) between the symbol for
the axis and the symbol for the mirror plane. In this case, each of the 2-fold axes are
perpendicular to mirror planes, so our symbol becomes:
2/m2/m2/m

108
Symmetry and Crystal Structure

Example 4.5: Symbol for a Block with a 2-fold Axis and 2 Mirror Planes

A2

m
m

Fig. 4. : Symmetry Symbols for the Rhombic Pyramidal Crystal Class

This model has one 2-fold axis and 2 mirror planes. For the 2-fold axis, we write:
2
Each of the mirror planes is unique. This is because each one cuts a different looking face. So, we
write 2 "m"s, one for each mirror plane:
2mm
The 2-fold axis is not perpendicular to a mirror plane, so we need no slashes. The final symbol is
then:
2mm
For this crystal class, the convention is to write mm2 rather than 2mm.

Example 4.6: Symbol for a Crystal with one 4-fold axis, four 2-fold axes, five mirror planes
and a centre of symmetry.
It contains 1 4-fold axis, 4 2-fold axes, 5 mirror planes, and a center of symmetry. The 4-fold axis
is unique. There are 2 2-fold axes that are perpendicular to identical faces, and 2 2-fold axes that
run through the vertical edges of the crystal. Thus there are only 2 unique 2 fold axes, because the
others are required by the 4-fold axis perpendicular to the top face. So, we write:
4 2 2
Although there are 5 mirror planes in the model, only 3 of them are unique. Two mirror planes cut
the front and side faces of the crystal, and are perpendicular to the 2-fold axes that are
perpendicular to these faces. Only one of these is unique, because the other is required by the 4-

109
Introduction to Crystallography

fold rotation axis. Another set of 2 mirror planes cuts diagonally across the top and down the
edges of the model. Only one of these is unique, because the other is generated by the 4-fold
rotation axis and the previously discussed mirror planes. The mirror plane that cuts horizontally
through the crystal and is perpendicular to the 4-fold axis is unique. Since all mirror unique mirror
planes are perpendicular to rotation axes, our final symbol becomes:
4/m2/m2/m
A4

m
m m

m A2
A2
m
A2 A2

Fig. 4.17: A Crystal with one 4-fold axis, four 2-fold axes, five
mirror planes and a centre of symmetry

THE 3-D CRYSTALS


The Bravais lattices may be thought of as being built up by stacking layers of the five plane
lattices, one on top of another (Table 4.2). The cubic and tetragonal lattices are based on the
stacking of square lattice layers; the orthorhombic P and I lattices are based on stacking of
rectangular layers; the orthorhombic C and F lattices on the stacking of the rectangular C layers;
the rhombohedral and hexagonal lattices on the stacking of hexagonal layers and the triclinic
lattices on the stacking of oblique layers.

Table 4.2: Relationship Between Plane and Bravais Lattices


Plane Lattice Bravais lattice Formed
Square Cubic and tetragonal
Rectangular Orthorhombic P and I
Rectangular C Orthorhombic C and F
Hexagonal Rhombohedral and hexagonal
Oblique Monoclinic and triclinic

110
Symmetry and Crystal Structure

These relationships between the plane and the Bravais lattices are easy to see, except perhaps for
the rhombohedral lattice. The rhombohedral unit cell has axes of equal length and with equal
angles between them. The hexagonal and the rhombohedral lattices differ in the ways in which the
hexagonal layers are stacked. In the hexagonal lattice they are stacked directly one on top of the
other. In the rhombohedral lattice they are stacked such that the next two layers of points lie above
the triangular interstices of the layer below, giving a three layer repeat.
The hexagonal and the rhombohedral stacking sequences have been met before in the stacking of
close packed layers (p.22). The hexagonal and rhombohedral lattices differ in the ways in which
the hexagonal layers are packed. The close packed hexagonal lattice corresponds to ABAB…
sequence and the rhombohedral lattice corresponds to fcc ABCABC.. sequence.
In this way, the 14 space lattices called the Bravais lattices can be obtained. These space lattices
can be broken down in to seven axial systems whose characteristics are given in Table 2.2(p. 15).

The Symmetry of the Fourteen Bravais Lattices: Crystal Systems


The unit cells of the Bravais lattices may be thought of as the building blocks of crystals. Hence
the external shape, or the observed symmetry of crystals, will be based upon the shapes and
symmetry of the Bravais lattices. We shall now describe the point symmetry of the unit cells of the
Bravais lattices.

(a) (b)

Fig. 4.18: The Point Symmetry Elements


(a) Cubic Unit Cell (b) Orthorhombic Unit Cell

The highest symmetry of all is that of the cubic system [Fig.2.18(a)]. It contains a total of nine
mirror planes: (i) three parallel to the cube faces and six parallel to the face diagonals, (ii) three
tetrad (four-fold) axes perpendicular to the three sets of cube faces, (iii) four triad (three-fold) axes
running between opposite cube corners, and (iv) six diad (two-fold) axes running between the
centres of opposite edges. This collection of symmetry elements is called point group symmetry of
the cube, because all the elements, planes and axes, passes through a popint in the centre.
Similarly, Fig. 4.18(b) shows the point group symmetry of an orthorhombic unit cell. An

111
Introduction to Crystallography

orthorhombic crystal has two-fold symmetry in three mutually perpendicular directions, the
crystallographic axes are taken parallel to the symmetry directions, so the unit cell is orthogonal.
The point group symmetry of these unit cells [Figs. 4.18(a) and 4.18(b)] is independent of whether
the cells are centred or not. All three cubic lattices, P (primitive, 1 lattice point per cell), I (body
centred, one additional lattice point at the centre of the unit cell) and F (face centred, additional
lattice points at each face of the unit cell) have the same point group symmetry; all four
orthorhombic lattices, P, I, F and C have the same point group symmetry and so on. This simple
observation leads to an important conclusion: it is not possible, from the observed symmetry of a
crystal, to tell whether the underlying Bravais lattice is centred or not. Therefore, in terms of their
point group symmetries, the Bravais lattices are grouped, according to the shapes of their unit cells
into seven crystal systems. For example, crystals with cubic P, I or F lattices belong to the cubic
system, crystals with orthorhombic P, I, F or C lattices belong to the orthorhombic system, and so
on.
However, some crystals with a hexagonal lattice (e.g. α-quartz) do not show hexagonal (hexad)
symmetry but have triad symmetry. Such crystals are assigned to the trigonal system rather than to
the hexagonal system. Hence the trigonal system includes crystals with both hexagonal and
rhombohedral Bravais lattices. There are no axes or planes of symmetry in the triclinic system.
The only symmetry that the triclinic lattice possesses (and which is possessed by all the other
lattices) is a centre of symmetry. There is, therefore, no restriction on the shape of the unit cell: all
three angles α, β, and γ have to be specified and hence the system is called triclinic.

Fig. 4.19: FCC Point Lattice Referred to Cubic


and Rhombohedral Cells

It must be emphasized that it is always possible to define a primitive unit cell. For example, the fcc
lattice (Fig. 4.19) may be referred to the primitive cell indicated by the dashed lines. The latter cell
is rhombohedral, its axial angle is 60o. Each cubic cell has 4 lattice points associated with it, each
rhombohedral cell one. Nevertheless, it is usually more convenient to use the cubic cell rather than
the rhombohedral one because the former immediately suggests the cubic symmetry which the
lattice actually possesses.

112
Symmetry and Crystal Structure

QUESTIONS AND PROBLEMS


4.1. Distinguish between a motif and a lattice point.
4.2. Define symmetry. Explain the basic symmetry operations.
4.3. Explain the symmetry of the upper case English letters: M, S and O.
4.4. Distinguish between a mirror plane and a glide plane of symmetry.
4.5. Distinguish between a proper and improper rotation axis.
4.6. Distinguish between congruent and enentiomormorphous objects.
4.7. Explain the important criteria that the lattice itself must possess at least the symmetry of the
motif.
4.8. Show, with suitable examples, that the action of one symmetry element generates another.
4.9. Identify the symmetry of an equilateral triangle.
4.10. Identify the symmetry of an square and a rectangle.
4.11. Show that it is not possible to have 5-fold rotational symmetry.
4.12. Show that it is not possible to have 8-fold rotation axis of symmetry.
4.13. Identify all plane point group symmetry.
4.14. Use a suitable example and show that the choice of the unit cell is arbitrary.
4.15. Distinguish between a screw operation and a glide reflection.
4.16. List the minimum symmetry elements in different crystal systems.
4.17. Consider a rectangular block and explain the procedure to derive the Hermann-mauguin
symbols.
4.18. Use suitable examples and show that the bravais lattices may be built up by stacking layers
of the five plane lattices.
4.19. Use suitable examples to show that it is not possible from the observed symmetry of a
crystal to tell whether the underlying Bravais lattice is centred or not.
4.20 Determine the point group symmetry of a chess board (a) without the chessmen and (b) with
the chessmen in the opening position (ignore the colour of the chessmen).
4.21. Make and examine the crystal models of NaCl, CsCl, ZnS (wurzite), and ZnS (sphalerite).
Identify the lattice and describe the motif of each structure.
4.22 Draw neat sketches to show the symmetry elements of a cube.
4.23 Identify the symmetry content of the crystal with rectangular sides and square ends.

113
Introduction to Crystallography

SOLUTION TO THE PROBLEMS


2.1. (a) A diad at the centre of the board and two diagonal mirror lines
(b) A diad only.

2.2.
Structure Bravais Lattice Motif
NaCl Cubic F (Na and Cl)
CsCl Cubic P (Cs and Cl)
ZnS (Wurzite) Hexagonal P 2 Zn + 2S
ZnS (sphalerite) Cubic F (Zn and S)

114
Chapter 5

Determination of Crystal Structure

In an earlier chapter, we presented an account of crystal structure as we understand it today. This


understanding was gained from a large amount of experimental work which confirm those views
either by direct or indirect observation and makes it possible for us to determine the atomic
arrangements and lattice parameters of different crystals. Such methods are based on the
diffraction of x-rays, electrons or neutrons. The most common method of structure determination
uses the technique of x-ray diffraction. This chapter is intended to give a basis for x-ray diffraction
procedures which are used to study the structure of metals and alloys.
Before the discovery of x-rays, crystallographers and mineralogists had accumulated a vast
knowledge about crystals by measurement of interfacial angles, chemical compositions and
physical properties. In addition the phenomenon of diffraction was well understood and it was
known that diffraction occurred whenever wave motion encountered a set of regularly spaced
scattering objects, provided that the wavelengths of the waves were of the same order of
magnitudes as the repeat distance between the scattering centers.
In 1895, Rontgen discovered x-rays. Although the nature of this radiation was unknown at that
time, the German Physicist M. von Laue argued that if crystals were composed of regularly spaced
atoms which might act as scattering centres for x-rays and if x-rays were electromagnetic radiation
of wavelengths similar in magnitude to the interatomic distance in the crystals, then it should be
possible to diffract x-rays by the crystals. In 1912 M. Von Laue and his associates passed a beam
Introduction to Crystallography

of x-rays through a thin crystal of ZnS and photographed the diffraction pattern caused by the
interference of x-rays and the crystal lattice. This experiment proved at the same time, the wave
nature of x-rays and the periodicity of arrangement of atoms within a crystal. In the same year
W.H. Bragg and W. L. Bragg analysed the Laue experiment and derived the necessary conditions
for diffraction in a more mathematical form, which laid the foundation for x-ray crystallography.

INTERACTION OF X-RAYS WITH ATOMS AND LATTICE STRUCTURE


When a beam of x-rays encounters an electron, the electric field of the beam sets it into a vibrating
or oscillatory motion. Thus the electron becomes a source of a new set of electromagnetic waves.
Since the frequency of oscillation of the electron will be the same as that of the incident beam, the
emitted radiations also have the same frequency as the incident beam. The process is thus referred
to as ‘Coherent Scattering’. The electron therefore receives a part of the incident energy and
radiates it in all directions.
When an x-ray beam encounters an atom, the interaction is limited to the extra nuclear electrons
and the nucleus does not take part in this interaction. Each electron in the atom now becomes a
source of x-rays. Although the electrons in the atom are in constant motion around the nucleus of
the atom, to a fair degree of approximation it may be assumed tat all the electrons scatter from the
center of the atom and thus the atom becomes a point source of x-rays radiating x-rays of the same
frequency in all directions. Since there exists Z number of electrons in an atom of atomic number
Z, the intensity scattered by the atom is larger than that scattered by a single electron.
When atoms spaced at regular intervals are irradiated by a beam of x-rays, each atom in the array
starts scattering the beam, and all the scattered beams have the same wavelength. Depending on
the phase difference or the path difference, the wave originating from different atoms take part in
constructive or destructive interference, leading respectively to addition or annihilation of
intensities of the individual waves. In an actual crystal the beam would be reflected from a large
number of parallel planes in the crystal and constructive interference can occur only under highly
restricted conditions.

BRAGG’S LAW FOR DIFFRACTION


We shall now determine the criteria for maximum reinforcement of x-rays reflected from atoms on
a number of equally spaced parallel planes such as exist in crystals. This reinforcement occurs
when all the diffracted waves are in phase with one another.
For this purpose, let us consider each plane of atoms in a crystal as a semi-transparent mirror; that
is each plane reflects a part of the x-ray beam and also permits part of it to pass through. When x-
rays strike a crystal, the beam is reflected not only from the atoms at the surface layers but also
from atoms underneath the surface to a considerable depth. Fig. 5.1 shows an x-ray beam that is
being reflected simultaneously from two parallel lattice planes. In the actual case, the beam would
be reflected not from two lattice planes but from a large number of parallel planes.

116
Determination of Crstal Structure

Fig. 5.1 shows the waves in the incident beam as being parallel. This is true if the radiation is
collimated before striking the sample. A line MX drawn perpendicular to the direction of
propagation of waves defines a wave-front, where all waves are in phase. A line NY is drawn
perpendicular to the diffracted beam and we have to determine the conditions under which the
waves in the diffracted beam are in phase.

Incident Normal to Diffracted


beam Surface beam

Atomic Planes

Fig. 5.1: Bragg’s law of Diffraction

The waves will be in phase at NY if the path length of the individual waves differs by an integral
number of wave-lengths. The uppermost wave has a path length equal to (NO + ON) and the next
lower wave has a path length equal to (XP + PY). The second wave travels over a greater distance
than does the first wave. The difference in the path length Δ is
Δ = (XP + PY) – (MO + ON)
If from point O, two lines are drawn perpendicular to XP and PY, respectively, at Q and R, then by
geometry
XQ = MO, RY = ON.
The path difference for these two beams is then
Δ = QP + PR.
If θ is the angle of incidence and d is the inter planar distance (the distance between two planes of
the same Miller indices), then
QP = PR = d sinθ.
Substituting in
Δ = QP + PR
we obtain Δ = 2d sin θ where d is the inter planar separation.

117
Introduction to Crystallography

Since the difference in path length must be an integral number of wave lengths for maximum
reinforcement.
nλ = 2d sin θ
This equation was fist formulated by W. L. Bragg, and is known as the Bragg’s law.

Example 5.1.

The spacing between the principal atomic planes in quartz crystal is 4.255 Å. The X-rays of
wavelength  = 1.449 Å incident on this crystal in a Bragg’s spectrometer. Find the smallest
glazing angle at which the beam will be reflected.

Solution.

d = 4.255 Å
 = 1.449 Å .

Let  = Smallest glazing angle at which the beam is reflected.

We know that as per Bragg’s Law,

n = 2d sin  = 2 x 4.255 x sin 

1 x 1.449 = 2 x 4.255 x sin 


( n = for smallest glazing angle)

1 x 1.449
 sin  = 0.17
2 x 4.255

 = 90 48’ Ans.

Order of Reflection
The value of ‘n’ in this equation is called the order of reflection, which may be any positive integer
(n = 1, 2, 3) and means that for a given value of wavelength and plane spacing ‘d’ there may be
several angles θ1, θ2, etc., at which diffraction may occur, consistent with the fact that sin θ should
not exceed unity. In a first order reflection, the scattered beam for adjacent planes such as planes I
and II in Fig. 5.1, would differ in path by one wavelength, those from alternate planes such as
planes I and III, would differ by two wavelengths etc.

118
Determination of Crstal Structure

11 th order 11 th order

5 th order 5 th order

2 nd order 2 nd order
1 st order 1 st order

Fig. 5.2: Order of Reflection

Let us now consider a simple example of application of the Bragg equation. The (110) planes of a
body centred cubic iron crystals have a separation of 1.181Ao. If these planes are irradiated with x-
rays of wave length 1.540Å, first order (n = 1) reflection will occur at an angle of
n
θ = sin-1
2d
1 x 1.54
θ = sin-1 = 40o7’
2 x 1.181
A second order reflection from (110) planes in iron is not possible with radiation of this
wavelength because the argument of the arc sin (nλ / 2d) is
2 x 1.54
= 1.302
2 x 1.181
a number greater than unity, and therefore the solution is impossible. On the other hand using a
wave length of 0.2090Å eleven orders of reflections becomes possible. The angle θ corresponding
to several of these reflections is shown in Table 5.1. Fig. 5.2 shows a schematic representation of
the same reflections. In considering the above example it is important to notice that although there
are eleven angles at which a beam of wavelength 0.209Å will be reflected with constructive
interference from (110) planes, only a very slight shift in the angle θ away from any of these
eleven values causes destructive interference and cancellation of the reflected beams.
It may be noted, that the final condition or diffraction does not contain the repeat distance of the
atoms in individual planes. Thus as long as atoms are regularly spaced within the planes, it is the
inter planer distance ‘d’ that decides the angles of diffraction. Thus if one chooses two crystals
with widely different atom spacing, even then planes with the same ‘d’ spacing in the two crystals
will diffract in the same directions. Although the x-rays are reflected by atoms, it would appear as
though x-rays are being reflected by the planes. In fact it is conventional to use the term ‘reflecting
planes’ and ‘reflected beams’ even if it actually means ‘diffracting planes’ and ‘diffracted beams’.

119
Introduction to Crystallography

It is to be noted that the incident beam, the diffracted beam and the normal to the reflecting plane
are always coplanar.

Table 5.1: Order of Reflection


Order of reflection θ, Angle of Incidence or Reflection
1 5o5’
2 10o20’
3 26o40’
4 80o

The angle between the diffracted beam and the transmitted beam is 2θ and is called the ‘diffraction
angle’, which is the one actually measured in the experiments. Since sin θ can not exceed unity,
diffraction takes place for ‘d’ values which satisfy the relation (nλ/2d) less than unity. i.e. nλ must
be less than 2d for diffraction to occur. In other words the wavelength used must be less than twice
the largest value of the interplanar spacing within the crystal. Ultraviolet rays with wavelength
500Å could not possibly be diffracted by a crystal. On the other hand if the wavelength is too
small, the values will be too small to be conveniently measured. The Bragg relation may be written
in a slightly different way to make it more convenient to handle without the botheration of having
to calculate the value of ‘n’:
λ = 2.(d/n). Sin θ
It is convenient to distinguish between indices of a reflected beam and indices of a reflecting plane
in a crystal. Indices of crystal planes never have a common factor. One is therefore free to use
indices with a common factor to express the order of reflection, n. Thus, first order, second order
and third order reflections from the planes (110) would be designated 110, 220, 330, etc. To avoid
confusion indices of reflections are written without parentheses. The above convention greatly
simplifies the interpretation of diffraction patterns, for an n order reflection from a (hkl) plane
would be designated nh, nk, nl.

Example 5.2.
For a certain BCC crystal, the (110) plane has a separations of 1.181 Å . These planes are indicated
with X-rays of wavelength 1.540 Å. How many order of Bragg’s reflections can be observed in
this case?

Solution
Given d = 1.181 Å ;  = 1.540 Å

120
Determination of Crstal Structure

We know that maximum order of the Bragg’s reflection can be observed with the maximum value
of reflection angle () equal to 900. Therefore number of orders as per Bragg’s Law,

2 d sin θ 2 x 1.181 sin 900


n=   1.53
λ 1.540

Since the value of ‘n’ can be integer only, therefore the highest permissible value of ‘n’ in this case
is 1. Ans .

Example 5.3

A beam of X-rays of  = 0.842 Å is incident on a crystal at a glancing angle of 80 35’, when the
first order Bragg’s reflection occurs. Calculate the glancing angle for third order reflection.

Solution

Given  = 0.842 Å ;  = 80 35’ when n = 1

Let 3 = Glancing angle for third order reflection, and

d = Interplaner distance.

We know that as per Bragg’s Law,

n = 2d sin 
 1 x 0.842 = 2d sin 80 35’ (i)
and 3 x 0.842 = 2d sin 3 (ii)

Dividing equation (ii) by (i),

3 x 0.842 2d sinθ 3

1 x 0.842 2d sin80 35'
sinθ 3 sinθ 3
3= 0

sin8 35' 0.15

 sin 3 = 3 x 0.15 = 0.45 or 3 = 260 45’ Ans.

121
Introduction to Crystallography

Example 5.4

X-rays from a tube undergoes first order reflection at a glancing angle of 120 from the face of a
calcite crystal. The grating space of the calcite crystal is 3.04 x 10-8 cm. Calculate the wavelength
of the X-rays. At what angle will the third order reflection take place from the crystal ?

Solution
Given :  = 120, d = 3.04 x 10-8 cm = 3.04 Å and n = 1.

Wavelength of X-rays
Let  = Wavelength of the X-rays.
We know that as per Bragg’s Law,
n = 2d sin  = 2 x 3.04 x sin 120
1x = 4.08 x 0.2079 = 1.264
or  = 1.264 Å Ans.

Third order reflection


Let 3 = Angle at which the third order reflection takes place.
We know that as per Bragg’s Law,

n = 2d sin 3 (i)
 1 x 1.264 = 2d sin 120 (ii)

and 3 x 1.264 = 2d sin 3

Dividing equation (ii) by (i),

3 x 1.264 2d sin12 0

1 x 1.264 2d sin θ 3

sinθ 3 sinθ 3
3= 0

sin12 0.2079
sin 3 = 3 x 0.2079 = 0.6237

or 3 = 380 35’ Ans.

122
Determination of Crstal Structure

FORBIDDEN REFLECTIONS
Bragg’s law is a necessary but not sufficient condition for diffraction by real crystals. It specifies
when diffraction will occur for unit cells having atoms positioned only at cell corners. However,
atoms situated at other sites (e.g., at the centres of the faces and in the interior of the unit cell as in
fcc and bcc crystals) act as extra scattering centres which can produce out of phase scattering at
certain Bragg angles. The net result is the absence of some diffracted beams which according to
Bragg’s law should be present. This is illustrated below with a simple example.

Normal to
Surface
(001)

(002) (001)
d 002

(002)
d 001

(001) (001)
(a) (b)

Fig. 5.3: Forbidden Reflections in BCC Structures

Let us now consider a body centred cubic structure as shown in Fig. 5.3 and oriented in such a way
that the (011) planes are at the proper Bragg angle. The interplanar spacing of the (001) planes is
found by substituting the proper values of h, k and l into the equation
a
d hkl =
h 2  k 2  l2
a
that is d001 = =a
(0  0  1

According to the Bragg’s law, the difference in path lengths between waves diffracted from the top
plane and the base plane of the unit cell must be an integral number of wavelengths according to
the Bragg’s law, that is
Δ = 2d001 sin θ001 = λ
But the difference in the path length between waves diffracted by the top plane and the (002) plane
is
δ = 2d002 sin θ001 = λ

123
Introduction to Crystallography

It may be noted that the angle in the above equation is the same as between (001) planes. This is
because the crystal is oriented to satisfy the Bragg condition for the (001) planes. The inter-planar
spacing of the (002) planes is
a
d002 = =a/2
( 0  0  4)

The values of Δ and δ are


Δ = 2001 sin θ001, and

δ = 2d002 sin θ001


Δ δ
or =
2 d 001 2 d 002

Thus we can write


 δ  δ
= as =
2 d 001 2 d 002 2a a
2x
2
Δ = 2δ or δ = Δ /2 or δ=λ/2 [Since Δ = λ]
The path difference between the waves diffracted from the top plane and the (002) plane is one-
half the wave length, and the waves arriving at a counter would be exactly 180o out of phase. Thus
the reflections from alternate planes annul each other, (002) canceling (001), etc. This cancellation
will not occur for the second order reflection, for then the scattering from plane (002) will be in
phase with that for (001) planes. The same is true for all even-order reflections.
The general condition for the reinforcement of diffracted beams in the bcc structure is that h + k +
l = even integer. Thus, reflections can occur from the (002) planes since the sum of Miller Indices
is 2, but not from the (001) planes whose Miller indices add up to an odd integer 1.
Note also that if the scattering power of (002) plane is not equal to that of (001) plane, there will
not be complete annihilation in the first order; the same will be true if the (002) plane is not
exactly mid-way between the (001) planes.
By similar arguments, we can show that the conditions for diffraction in the face centred cubic
structure is that the indices must be unmixed, that is
h, k, l = all even integers
h, k, l = all odd integers.
The allowed reflections for the simple, body centered and face centered cubic structures are listed
in Table 5.2. It should be noted that zero is considered an even integer. If we use the missing

124
Determination of Crstal Structure

reflections it is possible to differentiate between the diffraction patterns of simple cubic, body
centered cubic and face centered cubic crystals.

Table54.2: Allowed Reflections: Cubic Cells


hkl Simple Cubic Body Centered Cubic Face Centered Cubic
100 x
110 x x
111 x x
200 x x x
210 x
211 x x
220 x x x
221 x
300 x
310 x x
311 x x
222 x x x
320 x
321 x x
400 x x x

Example 5.5

X-rays with a wave length of 0.58 Å are used for calculating d200 in nickel. The reflection angle is
9.50. what is the size of unit cell?

Solution
Given  = 0.58 Å ;  =9.50

Let a = Size of unit cell or interatomic distance.

125
Introduction to Crystallography

We know that for a cubic crystal system (of nickel), the interplaner distance,
a
d=
h 2 k 2 l 2

a a
 d200 =  0.5a
2
2 0 0 2 2 2

and according to Bragg’s Law,


2d sin  = n
2 d200 sin 9.50 = 1 x 0.58
2 x 0.5a x 0.165 = 0.58
0.58
or a= 3.52 Å Ans.
1.165

EXPERIMENTAL DIFFRACTION METHODS


The Bragg equation sets very stringent conditions on the value of wavelength and θ for diffraction
to occur from a given crystal, and a constructive reflection should not be expected to occur every
time a monochromatic beam impinges on a crystal. Thus an arbitrary setting of a crystal in the path
of an x-ray beam of a fixed wave length, may not result in diffraction; some method of satisfying
the Bragg equation during the experiment has to be devised. One way of doing it is by varying the
wavelength to suit a given value of θ, which is illustrated as follows:
Suppose that a crystal is maintained in a fixed orientation with respect to a beam of x-rays and that
this beam is not monochromatic but contains all wavelengths longer than a given minimum value
λswl. This type of x-ray beam is called a white x-ray beam since it is analogous to white light,
which contains all the wavelengths in the visible spectrum. Although the angle of the beam is fixed
with respect to any given set of planes in the crystal and the angle θ of the Bragg law is therefore a
constant, reflection from all planes can now occur as a result of the fact that the x-ray beam is
continuous. This point can be illustrated with the aid of a simple cubic lattice.
Let the x-ray beam have a minimum wavelength of 0.5Å and let it make a 60o angle with the
surface of the crystal, which, in turn is assumed to be parallel to a set of (100) planes. In addition
let the 100 planes have a spacing of 1Å, substituting these values in the Bragg equation,

n λ = 2d Sin θ

or n λ = 2 (1) Sin 60o = 1.732

126
Determination of Crstal Structure

Thus the rays reflected from 100 planes will contain the following wavelengths:
1.732 Å for 1st order reflection
0.866 Å for 1st order reflection
0.577 Å for 1st order reflection
All other waves will suffer destructive interference, since for 4th and higher order reflections λ is
0.433 Å or less, which is less than the minimum wave length.

Normal to (0 1 2) Incident Normal to (012)


Reflected beams
beam Reflected
beam

(0 1 2) (012)

Fig. 5.4: X-ray Reflection from Planes Not Parallel to the


Surface of the Specimen

In the above examples, the reflecting planes were assumed parallel to the crystal surface. This is
not a necessary requirement for reflection, it is quite possible to obtain reflections from planes that
make any angle with the surface. Thus in Fig. 5.4, the incident beam is shown normal to the
surface and a (001) plane, while making an angle θ with the two {210} planes – (012) and (01 2).
The reflections from these two planes are shown schematically in Fig. 5.4. It may be concluded
that, when a beam of white x-rays strikes a crystal, many reflected beams will emerge from the
crystal, each reflected beam corresponding to a reflection from a particular different
crystallographic plane. Furthermore, in contrast to the incident beam that is continuous in
wavelength, each reflected beam will contain only discrete wave lengths as prescribed by the
Bragg equation. The Bragg equation can also be satisfied by varying θ to suit the given value of
wavelength or by varying λ. Depending on how this is achieved, three main diffraction Methods
have been developed.
λ variable, θ fixed: Laue Method
λ fixed, θ partly variable Rotating Crystal Method.
λ fixed, θ variable: Powder Method.
Of these diffraction techniques the powder method, also known as Debye-Scherrer method is one
of the most common methods that are used to determine both the crystal structure and the lattice
parameter of an unknown material.

127
Introduction to Crystallography

THE POWDER METHOD


The powder method, devised by Debye and Scherrer and independently by Hull in 1916, involves
diffraction of monochromatic x-rays by a powdered or a fine grained polycrystalline specimen and
can yield a great deal of structural information about the material under investigation.
Each particle or grain in the specimen is a tiny crystal oriented at random with respect to the
incident beam. There is a fair chance that a certain (hkl) plane will be correctly oriented to reflect
the incident beam. Thus a variation of θ is obtained, not by rotating a single crystal about one of its
axes but through the presence of many small crystals randomly oriented in space in the specimen.

d011= d 100 = 0.707 Å


2
d100 = 1 Å

Fig. 5.5: Relative Interplanar Spacing for Simple Cubic Lattice

The principle involved in the Debye-Scherrer method can be explained with the aid of an example.
Let us assume a crystalline structure with simple cubic lattice shown in Fig. 5.5 and that the
spacing between 100 planes equals 1Å. It can be easily shown that the inter planar spacing for the
planes of the 110 type equals that of the 100 planes divided by the square root of two and is,
therefore, 0.707Å.
In cubic crystals the distance d between any two planes of Miller indices (hkl) is given by
a
dhkl =
h 2  k 2  l2

where ‘a’ is the lattice parameter of the unit cell. Thus, all other planes in the simple cubic lattice
have a smaller spacing than that of cube or {100} planes. Now, according to the Bragg equation nλ
= 2d sin θ and if we focus our attention on 1st order reflections, where n equals 1, we have

θ = sin-1 (λ/2d).

This equation tells us that planes with the largest spacing reflect at the smallest angle θ.

128
Determination of Crstal Structure

If now it is arbitrarily assumed that the wavelength of the x-ray beam is 0.4Å, first order
reflections well occur from {100} planes (with the assumed 1 Å spacing) when

0.4 1
θ = sin-1 sin-1 = 11o30,
2(1) 5
On the other hand, 110 planes with a spacing 0.707Å reflect when
.0.4
θ = sin-1 = 16o28
2(0.707)

All other planes with larger indices reflect at still larger angles.

Fig. 5.6: Debye-Scherrer camera with the Cover Plate Removed

Fig. 5.6 shows schematically a Debye-Scherrer camera. A strip of photographic film is placed in
the camera in such a way that it covers the circumference. Holes are punched into this film strip so
that it fits over the entrance and exit inserts. The incident x-ray beam is collimated as it passes
through the entrance insert and is diffracted by the powder specimen. The beams of x-rays
diffracted from each set of planes form a cone of radiation in three dimensions. Fig. 5.7 shows
cones of diffracted radiation.
The intersection of these diffracted cones of radiation and the film are arcs as shown in Fig. 5.8.
The narrower the angle of the cone, the smaller the radius of curvature of the arc on the film strip
(for example, the radius of curvature of line 1 is less than that of line 5). The exit beam is stopped
by a beam stop which absorbs the x-radiation and prevents the dangerous beams from going
beyond the camera.

129
Introduction to Crystallography

Specimen

Incident Transmitted
Beam Beam

Fig. 5.7: Diffracted Beams in the Form of Cones

Analysis of Debye – Scherrer Film:


The angle between the diffracted beam and the transmitted beam (exit beam) is twice the Bragg
angle (2θ) [as we may see from Fig. 5.1] and is half the angle of the diffracted cone (as seen from
Fig. 5.7). If the distance to each line is measured from the centre of the exit hole in the film, (Fig.
5.8) then the distance to an arbitrary line is given by S = 2θR, where S is the distance from the exit
hole to the line, R is the interior radius of the camera and the angle θ is expressed in radians.

Beam Beam
Entrance Exit
S0
S10 S-1

Fig. 5.8: Diffraction Pattern Recorded on a Film Strip

To account for any shrinkage of the film during development, we calculate the distance S0,
between the entrance and exit holes as S0 = R π, since 1800 = π radians; and we use the ratios of S
to S0 to calculate the angle θ as
2θ S Sπ
= Or θ=
π S0 2S 0

We substitute these values of θ into Bragg’s law to find the planes causing the diffraction lines.

130
Determination of Crstal Structure

It has been mentioned earlier that in crystals of the cubic system, the distance between any two
planes of Miller indices (hkl) is given by
a
d=
(h  k 2  l 2 )
2

Now for first order reflections, the Bargg’s law may be written as λ = 2d Sin θ
a
On substituting d= in λ = 2d Sin θ
(h 2  k 2  l 2 )

λ
We obtain Sin θ = (h 2  k 2  l 2 )
2a

4a 2
or (h2 + k2 + l2) = Sin2 θ
λ2
Thus, if we know the wavelength of monochromatic radiation, we may calculate the values of h, k
and l for each reflection and determine the unit cell and lattice parameter.

Example 5.6
A diffraction pattern of a cubic crystal of lattice parameter a = 3.16 Å is obtained with a
monochromatic X-rays beam of wavelength 1.54 Å . The first line on this pattern was observed to
have  = 20.3 degrees. Determine the interplanar spacing and the Miller Indices of the reflecting
plane.

Solution.

Given : a = 3.16 Å ;  = 1.54 Å ; n=1;  = 20.30

We know that as per Bragg’s Law.

2d sin  = n

2d sin 20.30 = 1 x 1.54 = 1.54

1.54 1.54
 d= 0
  2.22 Å
2 sin 20.3 2 x 0.3469

131
Introduction to Crystallography

We also know that for a cubic crystal system, the interplanar distance,

a
d =
h 2 k 2  l 2

3.16
2.22 =
h 2 k 2 l 2

h 2  k 2 l 2 = 3.16/2.22 = 1.423

This value is very close to 2 (i.e. 1.414)

 h 2  k 2 l 2 = 2 or h2 + k2 + l2 = 2

This above equation may be solved by assuming any one of the indices (say l) equal to zero.

 h2 + k2 = 2

This is only possible if h = 1, k = 1.

Similarly, if we assume h = 0 or k = 0, we get the values as k = 1, l = 1 and h = 1, l =1


respectively. Thus the Miller Indices of the reflecting planes are (110), (011) or (101). Ans.

Example 5.7.

An X-ray powder pattern of molybdenum gives the following measurements :


Distance between lines 1-1’, S1 = 80.96 mm
Distance between lines 2-2’, S2 = 117.16 mm
Distance between lines 3 -3’,S3 = 147.2 mm
The wavelength of the X-ray beam used is 1.54 Å. If the radius of the powder camera is 57.3 mm,
determine the crystal structure of molybdenum and its lattice constant.

Solution.

Given :

S1 = 80.96 mm ; S2 = 117.16 mm, S3 = 147.2 mm, R = 57.3 mm and  = 1.54 Å.

132
Determination of Crstal Structure

(I) Crystal Structure

We know that the Bragg’s angle,

S
 = (degrees)
4

80.96
For line 1-1’, 1 = = 20.240
4

sin 1 = sin 20.240 = 0.346

and sin2 1 = ( 0.346)2 = 0.120

117.16
For line 2-2’, 2 = = 29.290
4

sin 2 = sin 29.290 = 0.489

and sin2 2 = (0.489)2 = 0.239

147.2
For line 3.3’, 3 = = 34.80
4

sin 3 = sin 34.80 = 0.599

and sin2 3 = (0.599)2 = 0.359

Now if we examine the values of sin2  for all the three cases, then we find that the values are in
the ratio 1: 2: 3. Since sin 2  is proportional to h2 + k2 + l2 values, therefore we can say that h2 + k2
+ l2 values for the given diffraction pattern are in ratio of 1: 2: 3. This ratio corresponds to simple
cubic (SC) structure. But we know from our experience that no pure metal has simple cubic
structure. Therefore we assume that h2 + k2 + l2 (or sin2 ) values in the ratio 2: 4: 4. This ratio
corresponds to body centered cubic (BCC) structure. Thus molybelenum has BCC structure.

(ii) Lattice constant

We know that for first reflection,

h2 + k2 + l2 = 2 and sin2  = 0.120

133
Introduction to Crystallography

1
and h2 + k2 + l2 = . sin2 
C
1
2= x 0.120
C

 C = 0120/2 = 0.060

We also know that the constant (C)

λ2 (1.58) 2
0.60 = 2

4a 4a 2

 a = 1.58/(2 x 0.06 ) = 3.23 Å Ans.

A very important application of the powder method is based upon the face that each crystalline
material has its own characteristic set of inter planar spacing. Thus while copper, silver and gold
all have the same crystal structure (face centred cubic), the unit cells of these three metals are
different in each case. Since each crystalline material has its own characteristic Bragg angles, it is
possible to identify unknown crystalline phases in metals with the aid of their Bragg reflections.
For this purpose a card file system has been published which lists approximately a thousand
elements and crystalline compounds with not only the Bragg angle of each important Debye-
Scherrer diffraction line, but also its relevant strength or intensity.
The identification of an unknown crystalline phase in a metal can be made by matching powder
pattern Bragg angles and reflected intensities of the unknown substance with the proper card of the
index. The method is quite analogous to a fingerprint identification system and constitutes an
important method of qualitative chemical analysis.

THE X-RAY DIFFRACTOMETER


The photographic methods of recording diffraction patterns are now supplemented in almost all x-
ray laboratories by the X-ray diffractometer, The x-ray diffractometer is a device that measures the
intensity of the x-ray reflections from a crystal with an electronic device, such as Geiger
Countertube or ionization chamber.
Fig. 5.9 shows the elementary parts of the diffractometer – a crystalline specimen, a parallel beam
of x-rays, and a Geiger countertube. The specimen S in the form of a flat plate is mounted so that
rotations about the axis labeled O are possible; this axis is perpendicular to the plane of the paper.
The counter is mounted on a movable carriage which may also be rotated about the O axis. The

134
Determination of Crstal Structure

carriage and the specimen are mechanically coupled so that a rotation of the specimen through θ is
accompanied by a 2θ rotation of the counter.

Fig. 5.9: Schematic Diagram of an X-Ray Diffractometer

This keeps the intensity measuring device at the proper angle during the rotation of the crystal so
that it can pick up each Bragg reflection as it occurs. The intensity measuring device is connected
through a suitable amplification system to a chart recorder, where the intensity of reflection is
recorded on a chart by a pen. In this manner one obtains an accurate plot of intensity against Bragg
angles. A typical x-ray diffractometer plot is shown in Fig. 5.10.
Intensity (Relative)

Diffraction Angle 2θ
Fig. 5.10: Diffraction pattern for Polycrystalline Iron

The high intensity peaks in the figure result when the Bragg condition is satisfied by some set of
crystallographic plane. The X-ray diffractometer is capable of measuring the intensities of the
Bragg reflections with great accuracy. Thus both qualitative and quantitative chemical analyses
can be made by this method.

135
Introduction to Crystallography

DETERMINATION OF CRYSTAL STRUCTURES


The basis of the methods employed for the determination of crystal structure is the fact that in the
x-ray diffraction patterns, the positions of diffraction lines depend on the crystal structure and the
line intensities depend on the location and kind of atoms in the unit cell. It is therefore possible to
solve for the structure from the diffraction pattern, and the actual procedure adopted is one of trial
and error. A metallurgist generally works with known structures and so should learn enough to be
able to index diffraction patterns of substances of known structures.
The basic principles involved in structure determination will now be illustrated with a pattern of
cubic substance and finding its lattice parameter. In this example, a radiation of wavelength 1.54 Å
was used and reflection lines were observed as shown in Table 5.3. The sin θ and sin2 θ values are
calculated and listed in the table as shown.

λ2
Since = A = constant,
4a 2

1
then we have h2 + k2 + l2 = Sin2 θ
A
Table 5.3: Analysis of the Pattern of a Cubic Substance
Line θ Sin θ Sin2θ S= h2 + k2 + l2 λ2/4a2
1 26.05 0.4392 0.1925 3 0.0642
2 30.50 0.5075 0.2575 4 0.0644
3 45.95 0.7187 0.5166 8 0.0646
4 57.50 0.8433 0.7113 11 0.0647
5 61.80 0.8813 0. 7767 12 0.0647

The quantity (h2 + k2 + l2) must be an integer because h, k, l are all integers, and only those planes
which conform to the conditions expressed in Table 4.2 lead to diffraction maxima. We now
proceed to determine the structure by trial and error, by checking the values of sin 2 θ against the
allowed values of h, k, l. The first assumption is to assign a value of h 2 + k2 + l2 = 1 to line 1, 2 to
line 2, 3 to line 3, etc. This does not give a constant value in this case and this means that the
structure is not simple cubic.
A second assumption is then made and a value of h2 + k2 + l2 = 2 to line 1, 4 to line 2 etc. are
assigned using the values of h, k, l for body centred cubic crystals, listed in Table 5.2. For this
particular case, this assumption also is not correct, since the values of sin 2 θ / (h2 + k2 + l2) is not a
constant.

136
Determination of Crstal Structure

The third assumption is to assign a value of 3 to the first line, 4 to second line, 8 to third line and
we find that this gives a reasonably constant value. The structure, therefore is face centred cubic
since the reflections that occur correspond to those allowed diffraction maxima of Table 5.2 for
face centred structure.
Referring to Table 5.3, it may be noted that there is a systematic error in the values of ‘a’. This
error decreases as θ increases and therefore we select a value of ‘a’ for the highest angle namely,
61.8 in this case, as being the most accurate of those listed.
This analysis of the pattern is thus indicative of a face centred cubic structure. If we substitute the
value of λ = 1.54 Å into the equation λ2/4a2 = constant, we find the lattice parameter to be 3.027 Å.
In general powder patterns of cubic substances can usually be distinguished at a glance from those
of non-cubic substance since the patterns of non-cubic crystals normally contains many more lines.
In addition, the Bravais lattice can usually be identified by inspection: there is an almost regular
sequence of lines in simple cubic and body centred cubic patterns but the former contains almost
twice as many lines, while a face centred cubic pattern is characterised by the alternate presence of
a pair of lines and a single line. The procedure followed in the indexing of patterns of hexagonal
crystals will be discussed in a later section (chemical analysis by x-ray techniques).

QUESTIONS AND PROBLEMS


5.1. Derive the Bragg’s law for diffraction? Explain how this is used in the study of metal
crystals. Name the experimental x-ray diffraction methods.
5.2 What are forbidden reflections? Show that, an x-ray beam of a given wave-length, oriented
to reflect (first order) constructively from cube planes with a separation equal to the lattice
constant a in a BCC crystal lattice, suffers destructive interference for the cube planes
having an actual separation a/2.
5.3 The (111) plane of a cubic crystal is inclined at 26o to an x-ray beam. If the interplanar
spacing is 1.506 Å, compute the x-ray wave-length that will give a first order reflection at
this angle.
5.4 How many orders of reflection are possible if the interplanar spacing is 1.506 Å and the
wave-length is 0.724 Å? At what angle will they occur?
5.5 Calculate the angles for first order reflection (n-1) from the (100) and (110) planes for a
simple cubic lattice of side 3 Å when the wave-length is 1.0 Å. If the lattice were BCC
would you expect to find diffracted beams at the same angles?
5.6 Draw neat sketches and explain the analysis of crystal structures by the Debye-Scherrer
method

137
Introduction to Crystallography

5.7 A BCC crystal is used to measure the wavelength of some x-rays. The diffraction angle is
20.2o for reflections from (110) planes. What is the wavelength? The lattice constant of the
crystal is 3.15 Å.
5.8 An x-ray powder pattern of a cubic crystal obtained using a diffractometer with a chart
recorder shows strong peaks at 2θ values of 90000’, 1120 00’, and 1200 00’. If an x-ray
tube is used which produces radiation of λ = 2 Å, calculate the lattice parameter of this
crystal and state whether this BCC or FCC.
5.9 X-rays with a wave-length of 1.54Å are used for calculating the spacing of (200) planes in
aluminium. The reflection angle is 22.4o. What is the size of the unit cell?
5.10 A powder pattern obtained for lead (FCC) with radiation of wavelength 1.54 Å. The (222)
reflection is observed at Bragg angle θ = 32o.
What is the lattice parameter of lead?
What is the lowest index reflection observed in the powder pattern?
5.11. Using a diffractometer and a radiation of wavelength 1.54 Å, a reflection from an FCC
crystal is observed when 2θ is 1210.
What are the indices (hkl) of this reflection?
What is the spacing between planes in the crystal from which this reflection originates?
(Hint: Assume that this reflection is from the lowest index plane for FCC. Then check
whether a reflection from the next higher-index plane is possible.)
5.12 From a powder camera of diameter 115.6 mm, using an X-ray beam of wavelength 1.54 Å,
the following S values in mm were obtained:
86, 100, 148, 180, 188, 232 and 272. Find out the lattice constant and structure of the
material.
5.13 Calculate the Bragg angels for Mo-Kα radiation (λ= 7.12 ×1012m) diffracted from the (110)
and (200) planes of tungsten is BCC. with a lattice parameter of 316 × 10-12m. Would you
expect a diffracted beam from the (100) or (111) plane ?
5.14 A powder photograph of a certain metal, taken with radiation of wavelength parameter of
154 × 10-12m. exhibits diffraction lines corresponding to the following Bragg angels:
20014’, 29023’, 36050’, 43058’, 50045’, 58044’, 66035’, 78034’.
Assuming that the metal is cubic, determine whether the lattice is FCC or BCC and
calculate the unit cell parameter. By comparing this value with tabulated values for metals
try to identify the metal used.

138
Determination of Crstal Structure

5.15 The powder pattern of aluminium made with Cu Kα radiations contains ten lines whose
sin2θ values are: 0.1118, 0.1487, 0.294, 0.403, 0.439, 0.583, 0.691, 0.727, 0.872, 0.981.
Index these lines and calculate the accurate value of the lattice parameter.
5.16 In each of the following sections the powder patterns of elements is represented by the
observed sin2θ values for the first seven or eight lines on the pattern made with Cu kα
radiations. In each case index the lines, find the crystal system, Bravais lattice and
approximate lattice parameter(s).
a 0.0806 0.0975 0.1122 0.2100 0.2260 0.2740 0.3050 0.3210
b 0.0603 0.1610 0.2210 0.3220 0.3830 0.4840 0.5450 0.6450
c 0.1202 0.2380 0.3570 0.4750 0.5930 0.07110 0.8300 --
d 0.0768 0.0876 0.0913 0.1645 0.2310 0.2740 0.3080 0.3190

5.17 The following data wee obtained from the Debye-Scherrer pattern of a simple cubic
substance made with Cu Kα radiations. The values given are sin 2θ values with h2 + k2 + l2
values given in brackets: 0.9114(38), 0.9761(41) and 0.9980(42). Determine the lattice
parameter of the material accurate to four significant figures by graphical extrapolation.
5.18 If a Bragg 41.31o angle is observed for first-order diffraction from the {110} plane of body
centered cubic niobium using copper Kαl radiation (λ = 0.1541 nm), what is the interplanar
spacing of the {110} planes in this metal?
5.19 Consider the Bragg equation with respect to first order reflections from {100} planes of a
bcc metal. By how much of a wavelength do the reflected pathlengths differ for two
adjacent parallel (100) planes, if the interplanar spacing, d. is taken as a instead of α / 2?
Does this explain why the {100}plane is not listed as a reflecting bcc plane? Explain.
5.20 Determine powder pattern, S, values for the first four reflecting planes, listed in Appendix
C, if the specimen is a gold powder. Assume that Cu Kαl radiation (λ = 0.1541 nm) is used
and that the lattice parameter of the gold crystal is 0.4078 nm.
5.21 The metal iridium has an FCC crystal structure. If the angle of diffraction for the (220) set
of planes occurs at 69.22o (first-order reflection) when monochromatic x-radiation having a
wavelength of 0.1542 nm is used, compute (a) the interplanar spacing for this set of planes,
and (b) the atomic radius for an iridium atom.
5.22 The metal rubidium has a BCC crystal structure. If the angle of diffraction for the (321) set
of planes occurs at 27.00o (first-order reflection) when monochromatic x-radiation having a
wavelength of 0.0711 nm is used, compute (a) the interplanar spacing for this set of planes,
and (b) the atomic radius for the rubidium atom.

139
Introduction to Crystallography

5.23 For which set of crystallographic planes will a first-order diffraction peak occur at a
diffraction angle of 46.21o for BCC iron when monochromatic radiation having a
wavelength of 0.0711 nm is used?
5.24 If a Bragg 41.31o angle is observed for first-order diffraction from the {110} plane of body
centered cubic niobium using copper Kl radiation (λ = 0.1541 nm), what is the interplanar
spacing of the {110} planes in this metal?
5.25 Derive a relation analogous to Bragg’s law of diffraction but giving the angle Ø as a
function of the interplanar spacing and the wavelength of the X rays. Here Ø is the angle at
which complete cancellation should occur.
5.26 X-ray powder pattern of a cubic crystal obtained using a diffractometer with a chart
recorder shows strong peaks at values of 2θ of 90o 00’, 112o 00’, and 120o 00’. If an X-ray
tube is used which produces radiation of λ = 2Å, calculate the lattice parameter of this
crystal and state whether it is body-centered or face-centered cubic.
5.27 Tin exists in two forms – grey tin and white tin. The unit cell of white tin is simple
tetragonal with the lattice parameters α = 5.83 Å and c = 3.18 Å. Calculate the Bragg angles
θ at which the first six diffraction lines appear. The distance between any two planes of
indices (hkl) in the tetragonal system is

1
d=
[(h  k ) /a 2 ]  (l 2 / c 2 )
2 2

and a copper target x-ray tube is used in this experiment.

140
Chapter - 6

The Stereographic Projection

The angular relationship among crystal faces, crystal edges, zones and crystallographic symmetry
elements cannot be accurately displayed by perspective drawings and if they are stated precisely in
mathematical terms, they are often difficult to comprehend and to manipulate. But frequently we
are more interested in these angular relationships than in any other aspect of the crystal. Therefore,
we need a kind of drawing on which the angles between planes can be accurately measured and
which will permit graphical solution of problems involving such angles. The stereographic
projection permits the mapping of crystallographic planes and directions in two dimensions in a
convenient and straightforward manner. The real value of the method is attained when it is
possible to visualise crystallographic features directly in terms of their stereographic projections.
The stereographic projection is much used for the following purposes:
(i) Studying preferred orientations of polycrystalline materials (pole figures).
(ii) Reorienting crystals preparatory to cutting specified crystal faces
(iii) Determining the crystallographic indices of surface markings such as slip lines, deformation
bands etc.
(iv) Solving crystallographic problems involved in precipitation, solid state transformations, etc.
The purpose of this section is to concentrate on the geometrical correspondence between
crystallographic planes and directions and their stereographic projections. In each case, a sketch of
Introduction to Crystallography

a certain crystallographic feature, in terms of its location in the unit cell, is compared with its
corresponding stereographic projection.

REFERENCE SPHERE AND ITS STEREOGRAPHIC PROJECTION


The nature of stereographic projection of a crystal is easily understood if the crystal is assumed to
be very small and to be located exactly at the centre of a sphere. Crystal planes within the crystal
can be represented by:
(i) A set of plane normals radiating from some one point within the crystal. This is because the
orientation of any plane in a crystal can also be represented by the inclination of the normal to
that plane relative to some reference plane. If a reference sphere is now described about this
point, the plane normals will intersect the surface of the sphere in a set of points called poles.
The poles of the {100} planes of a cubic crystal is shown in Fig.6.1(a). The array of poles on
the sphere, forming a pole figure, represents the orientation of the crystal planes without, of
course, indicating the size and shape of the crystal planes.

P1 K

P2
L
B
A

C
D

N
(a) M (b)

Fig.6.1: (a) {100} Planes of a Cubic Crystal (b) Angle between Two Planes

(ii) By the trace the extended plane makes in the surface of the sphere, as illustrated in Fig.6.1(b).
In this figure the trace ABCDA represents the plane whose pole is P1. The crystal is assumed
to be so small that all crystal planes pass through the centre of the sphere. The trace ABCDA
is, therefore, a great circle (a circle of maximum diameter) on the sphere. A plane not passing
through the centre will intersect the sphere in a small circle.
The angle α between two planes is evidently equal to the angle between their great circles or to the
angle between their normals [Fig.6.1(b)]. But this angle, in degrees, can also be measured on the
surface of the sphere along the great circle KLMNK connecting the poles P1 and P2 of the two

142
The Stereographic Projection

planes, if this circle is divided into 360 equal parts. The measurement of an angle has thus been
transferred from the planes themselves to the surface of the reference sphere.
However, in practice it is usually more convenient to measure angles on a flat sheet of paper rather
than on the surface of a sphere.. The stereographic projection is one of the methods by which the
sphere may be mapped without distorting the angular relations between the planes or the poles.

N'

Projection plane

W' Light
W
C source
E
E'

Reference
S
sphere

Basic circle
S'

Fig.6.2: The Stereographic projection

The stereographic projection is made by placing a plane of projection normal to the end of any
chosen diameter of the sphere and using the other end of that diameter as the point of projection. In
Fig.6.2 the projection plane is normal to the diameter AB, and the projection is made from the
point B. If the sphere is transparent, the markings on the surface of the sphere will be projected as
shadows on the plane of projection. The pattern made by the shadows is a stereographic projection.
In Fig.6.2 the point P′ is the stereographic projection of pole P. . This is obtained by drawing the
line BP and producing it until it meets the projection plane. Alternately stated, the stereographic

143
Introduction to Crystallography

projection of the pole P is the shadow cast by P on the projection plane when a light source is
placed at B. The observer, incidentally, views the projection from the side opposite the light
source. The distance of the projection plane from the sphere is immaterial; a change in the distance
merely changes the magnification and does not alter the geometrical relation.
The plane NESW is normal to AB and passes through the center C (Fig.6.2). It, therefore, cuts the
sphere in half and its trace in the sphere is a great circle. This great circle projects to form the basic
circle N′ E′S′W′ on the projection, and all poles on the left-hand hemisphere will project within
the basic circle shown in the figure.

N'

W' W
A B
E
P' E' P P' C
W' E'
S

S'

Fig.6.3: Stereographic Projection of Circles


(a) Projection of a Great Circle (b) Projection of a Small Circle

Poles on the right-hand hemisphere will project outside this basic circle, and those near B will
have projections lying at very large distances from the center. If we wish to plot such poles, we
move the point of projection to A and the projection plane to B and distinguish the new set of
points so formed by minus signs, the previous set (projected from B) being marked with plus
signs. It is, thus, possible to represent the whole sphere within the basic circle if the two
projections are superimposed, the one for the left-hand hemisphere constructed with the light
source at B (Fig.6.2) and the other for the right hand hemisphere with the light source at A and the
screen on the right (at B). Note that the movement of the projection plane along AB or its
extension merely alters the magnification; we usually make it tangent to the sphere. But we can
also make it pass through the center of the sphere, for example, in which case the basic circle
becomes identical with the great circle NESW.

144
The Stereographic Projection

A lattice plane in a crystal is several steps removed from its stereographic projection, and it may
be worth-while at this stage to summarise these steps:
(i) The plane C is represented by its normal CP.
(ii) The normal CP is represented by its pole P (its intersection with the reference sphere).
(iii) The pole P is represented by its stereographic projection P′.
A student should have enough familiarity with the stereographic projection, so as to be able
mentally to omit these intermediate steps and refer to the projected point P′ as the pole of the plane
C or, even more directly, as the plane C itself.

Projection of Great and Small Circles


In general, a great circle projects within the basic circle as an arc of a circle. However, the great
circles that pass through the points A and B (Fig.6.3) will project as straight lines through the
center of the projection. This is the case when the projection is a circle of infinite radius. Projected
great circles always cut the basic circle in diametrically opposite points, since the locus of a great
circle on the sphere is a set of diametrically opposite points. Thus, the great circle ANBS projects
to form a straight line (N′S′) passing diametrically through the basic circle on the projection. The
great circle AWBE will project to form a straight line W′E′.
That this is true will be seen from the fact that the great circle and its projection are in fact lines of
intersection of a plane with the sphere and projection plane respectively. If the great circle is
graduated in degrees, its projection will be a scale of stereographically projected degree points and
will be useful for reading angular distances on the projection.
A small circle inscribed about a point such as P (Fig.6.3) that lies on the great circle AWBE will
cut the great circle at two points, each of which is φo from P. The point P will project to P′. The
bundle of projection lines for the small circle will form an elliptical cone with its apex at S, and the
cone will intersect the plane in a true circle of which the centre is on the line W′E′, either inside or
outside the basic circle. The point P′ will not be in the centre of the area of this projected circle but
will lie on line W′E′ at a point located at an equal number of projected degrees (φo in this case)
from all points of the projected circle. The centre of the projected circle is at C. If the radius of the
small circle is increased, it finally becomes a great circle. Since this great circle does not pass
through point N, its projection will not be a straight line, but will be a circle having its centre on an
extension of W′E′; it will cut the line W′E′ at the point 90 stereographic degrees from P′.

Properties of Stereographic Projection


The most important properties of the stereographic projection are:
(i) Angular truth is conserved, i. e., the angle between lines on the surface of a sphere is equal to
the angle between the projections of these two lines.

145
Introduction to Crystallography

(ii) Circles on the surface of the sphere projects as circles on the plane of projection (circles
project as circles). This is true for both great circles and small circles.
However, the centres of the small circles on the sphere do not project to the centre of the area
of the projected circles but will be displaced radially an amount corresponding to equal
angular distances from the centre to all points on the circumference.
Great circles on the sphere appear on the projections as circles cutting the basic circle at two
diametrically opposite points; a great circle lying in a plane perpendicular to the projection
plane becomes a diameter on the projection while the great circles in inclined positions on the
sphere may be made to coincide with one of the meridians of the Wulff net.
These two properties in particular make the stereographic projection the most useful and
appropriate for representing the angular relationships in three dimensions. Another important
application of stereographic projection is that of representing the symmetry relations between the
faces of the crystal. On a well formed crystal all the faces that are related to one another by the
symmetry of the crystal structure are developed to an equal extent and the shape of the crystal
therefore, reveals its true symmetry. Accidents of crystal growth (such as unequal access of the
crystallizing liquid or interference by adjacent crystals) sometimes impede the growth of the
crystal in certain directions in such a way as to musk the true symmetry. These accidents do not,
however, affect the angles between the faces. Thus the poles representing the planes in the crystal
are entirely uninfluenced by their relative sizes, and the symmetry of the arrangements of these
points on the surface of the sphere reveals the true symmetry of the crystal, whether or not it be
well formed. This symmetry – can be recognized in the stereographic of these points.

Projection of Some Important Planes and Directions of a Cubic Lattice


The stereographic projection is a two-dimensional drawing of three-dimensional data. The
geometry of all crystallographic planes and directions is accordingly reduced by one dimension.
Planes are plotted as great circle lines, and directions are plotted as points. Also the normal to
plane completely describes the orientation of a plane.
The stereographic projection of the (100), (110) and (111) planes are shown in the three parts of
Fig.6.4. It can be seen that the stereographic projection of each plane can be represented either by
a great circle or by a point showing the directions in space that is normal to the plane (Fig.6.4).

Directions that Lie in a Plane


The positions of crystallographic directions that lie in a particular plane of a crystal can be shown
on the stereographic projections. Thus, in a body centred cubic crystal one of the more important
planes is {110}, and in each of these planes there are two close-packed <111> directions. The two
that lie in the (101) plane are shown in Fig.6.5 where appear as dots lying on a great circle
representing the (101) plane.

146
The Stereographic Projection

Fig.6.4: Stereographic Projection of Some Important Planes of Cubic Crystal

Fig.6.5: The (101) Plane and Two <111> Directions in Cubic System

147
Introduction to Crystallography

Planes of a Zone
Those planes that mutually intersect along a common direction form the planes of a zone, and the
line of intersection is called the zone axis. The [111] direction (Fig.6.6) is a zone axis.

Fig.6.6: The three {110} Planes of the Cubic System that belong to the Zone
of Planes, the Zone Axis of which is [111] Direction.

There are three {110} planes that pass through the [111] direction. There are also a number of
other planes that have the same zone axis. If only the poles of these planes are plotted, all of the
poles fall on the great circle (Fig.6.7) representing the stereographic projection of the (111) plane.

(011)

(132) (111) pole


(121)
( 231)
(110)
(321)
(211) (312)
(101) ( 213)
(112) (123)
(011)

Fig.6. 7: Stereographic Projection of the Planes that have [111] as their Zone Axis.

148
The Stereographic Projection

THE WULFF NET


The device most useful in solving problems involving the stereographic projection is the Wulff net
(Fig.6.8). It is the projection of a sphere ruled with parallels of latitude and longitude lines on a
plane parallel to the north-south axis of the sphere. On a Wulff net all meridians (the longitude
lines) are great circles connecting the north and south poles of the net. The equator is also a great
circle. All other latitude lines on a Wulff net are small circles extending from side to side.

Fig.6.8: Wulff Net Drawn to 2o Intervals

These nets are available in various sizes, one of 18-cm diameter giving an accuracy of about one
degree, which is satisfactory for most problems; to obtain greater precision, either a larger net or
mathematical calculation must be used.
In solving problems with the Wulff net, it is customary to cover it with a piece of tracing paper. A
common pin is then driven through the paper and into the exact centre of the net so that it is free to
rotate with respect to the net. The paper thus mounted serves as the work sheet on which
crystallographic data are plotted with the basic circle of the same diameter as that of the Wulff net.

149
Introduction to Crystallography

Measurement of Angle
Poles Lie on a Great Circle: We have already seen (Fig.6.1) that the angle between two crystal
planes could be measured on the surface of the sphere along the great circle connecting the poles
of the two planes. This measurement can also be carried out on the stereographic projection if, and
only if, the projected poles lie on a great circle. In Fig.6.9, for example, the angle between the
planes A and B or C and D can be measured directly, simply by counting the number of degrees
separating them along the great circle on which they lie. Note that the angle C-D equals the angle
E-F, there being the same difference in latitude between C and D as between E and F.

Fig.6.9: Rotation of Poles about NS Axis of Projection

Poles Not on a Great Circle: If the two poles do not lie on a great circle, then the projection is
rotated relative to the Wulff net until they do lie on a great circle, where the desired angle
measurement can then be made. Fig.6.10(a) is a projection of the two poles P1 and P2 [Fig.6.1(b)],
and the angle between them is found by the rotation illustrated in Fig.6.10(b). This rotation of the
projection is equivalent to rotation of the poles on latitude circles of a sphere whose north-south
axis is perpendicular to the projection plane.
Intersection of Great Circles: We have seen that a plane may also be represented by its trace in
the reference sphere. This trace becomes a great circle in the stereographic projection. Every point
on this great circle is 90o from the pole of the plane. Therefore, the great circle may be found by
rotating the projection until the pole falls on the equator of the underlying Wulff net and tracing
that meridian which cuts the equator 90o from the pole [Fig.3.11(a)]. If this is done for two poles,
as in Fig.6.11(b), the angle between the corresponding planes may be found from the angle of
intersection of the two great circles corresponding to these poles. This method of angle
measurement is not, however, very accurate.

150
The Stereographic Projection

Fig.6.10(a): Stereographic Projection of Poles P1 and P2 of Fig.6.1

Fig. 310(b): Rotation of Projection to Put Poles on the Same


Great Circle on the Wulff Net.

151
Introduction to Crystallography

Fig.6.11(a): Method of Finding the Trace of a Pole


(the Pole P in Fig.6.10)

Fig.6.11(b): Measurement of an Angle between Two Poles by


Measurement of Angle of Intersection of the Corresponding Traces

152
The Stereographic Projection

Rotation of Poles around the Axes


The rotation of poles about an axis normal to the projection is accomplished simply by rotation of
the projection around the center of the Wulff net. On the other hand, rotation about an axis lying in
the plane of the projection is performed in two steps: (i) by rotating the axis about the center of the
Wulff net until it coincides with the north-south axis (if it does not already do so), and (ii) moving
the poles involved along their respective latitude circles the required number of degrees.

Fig.6.12: Rotation of Poles about NS Axis of Projection

Suppose it is required to rotate the poles A1 and B1 (Fig.6.12) by 60o about the NS axis, the
direction of motion being from W to E on the projection. Then A1 moves to A2 along its latitude
circle as shown. B1, however, can rotate only 40o before finding itself at the edge of the projection:
we must then imagine it to move 20o in from the edge to the point B1 on the other side of the
projection, staying always on its own latitude circle. The final position of this pole on the positive
side of the projection is at B2 diametrically opposite B′1.
Rotation about an axis inclined to the plane of projection is accomplished by compounding
rotations about axes lying in and perpendicular to the projection plane. In this case, the given axis
must first be rotated into coincidence with one or the other of the two latter axes, the given rotation
performed, and the axis then rotated back to its original position. Any movement of the given axis
must be accompanied by a similar movement of all the poles on the projection.

153
Introduction to Crystallography

For example, we may be required to rotate A1 about B1 by 40o in a clockwise direction (Fig.6.13).
In (a) the pole to be rotated A1 and the rotation axis B1 are shown in their initial position. In (b) the
projection has been rotated to bring B1 to the equator of a Wulff net.

A2 40o
o
A1 48
N
o
A4 48 A3

B3 48o
B1 B2

(a)

N N

A1 A4
A1

W E W E
B1
B1
D
C

S S
(b) (c)

Fig.6.13: Rotation of a Pole about an Inclined Axis

154
The Stereographic Projection

A rotation of 48o about the NS axis of the net brings B1 to the point B2 at the center of the net; at
the same time A1 must go to A2 along a parallel of latitude. The rotation axis is now perpendicular
to the projection plane, and the required rotation of 40o brings A2 to A3 along a circular path
centered on B2.
The operations which brought B1 to B2 must now be reversed in order to return B2 to its original
position. Accordingly, B2 is brought to B3 and A3 to A4, by a 48o reverse rotation about the NS axis
of the net. In (c) the projection has been rotated back to its initial position, construction lines have
been omitted, and only the initial and final positions of the rotated pole are shown. During its
rotation about B1, A1 moves along the small circle shown. This circle is centered at C on the
projection and not at its projected center B1. To find C we use the fact that all points on the circle
must lie at equal angular distances from B1; in this case, measurement on a Wulff net shows that
both A1 and A4 are 76o from B1. Accordingly, we locate any other point, such as D, which is 76o
from B1, and knowing three points on the required circle, we can locate its center C.

Fig.6.14: Standard projection of Cubic Crystal on (001) Plane

Standard Projection of Crystals


A standard projection (Fig.6.14) shows the relative orientation of all the important planes in the
crystal. Such a projection is made by selecting some important crystal plane of low indices as the
plane of projection and projecting the poles of various crystal planes onto the selected plane. The
construction of a standard projection of a crystal requires a knowledge of the inter-planar angles
for all the principal planes of the crystal. A set of values applicable to all crystals in the cubic

155
Introduction to Crystallography

system is given in Table 6.1, but those for crystals of other systems depend on the particular axial
ratios involved and must be calculated by using appropriate equations.
We have seen that the poles of planes of a zone will all lie on the same great circle (Fig.6.6) on the
projection, and the axis of the zone will be at 90o from this great circle. Furthermore, important
planes usually belong to more than one zone and their poles are, therefore, located at the
intersection of zone circles. It is also helpful to remember that important directions, which in the
cubic system are normal to planes of the same indices, are usually the axes of important zones.
Fig.6.14 shows the principal poles of a cubic crystal projected on the (001) plane of the crystal.
This is properly called a standard (001) projection of a cubic crystal. The location of the {100}
cube poles follows immediately. The {110} poles must lie at 45o from {100} poles (Table 6.1),
which are themselves 90o apart. In this way we locate (011), for example, on the great circle
joining (001) and (010) and at 45o from each. After all the {110} poles are plotted, we can find the
{111} poles at the intersection of zone circles.
Inspection of a crystal model or the use of the zone relation will show that (111), for example,
belongs to both the zone [ 1 01] and the zone [0 1 1]. The pole of (111) is thus located at the
intersection of the zone circle through ( 1 00), (0 1 1), and (010). This location may be checked by
measurement of its angular distance from (010) or (100), which should be 54.7o.

Representation of Symmetry on the Stereographic Projection


It may be noted that each of the basic crystallographic directions (Fig.6.14) is represented by a
characteristic symbol. For the {100} poles, this is a square, signifying that these poles correspond
to four-fold symmetry axes. If the crystal is rotated 90o about any of these directions, it will be
returned to an orientation exactly equivalent to original orientation. In a 360o rotation about a
{100} pole, the crystal reproduces its original orientation four times. In the same fashion, a <111>
direction correspond to a three-fold symmetry axis, and these directions are indicated in the
stereographic projection by triangles. Finally the two-fold symmetry of the <110> directions is
indicated by the use of small ellipses to designate their positions in the stereographic projection.

Miller Indices of a Given Pole


It is sometimes necessary to determine the Miller indices of a given pole on a crystal projection,
for example the pole A in Fig.6.15(a), which applies to a cubic crystal. If a detailed standard
projection is available, the projection with the unknown pole can be superimposed on it and its
indices will be disclosed by its coincidence with one of the known poles on the standard.
Alternatively, the method illustrated in Fig.6.15(b) may be used.
The pole A defines a direction in space, normal to the plane (hkl) whose indices are required, and
this direction makes angles ρ, σ, τ with the coordinate axes a, b, c. These angles are measured
along the great circles indicated by turning the plot to the proper orientation on a stereographic net.

156
The Stereographic Projection

Table 6-1 Angles between Planes in Cubic Crystals


HKL hkl
100 100 0.00 90.00
110 45.00 90.00
111 54.74
210 26.56 63.43 90.00
211 35.26 65.90
221 48.19 70.53
310 18.43 71.56 90.00
311 25.24 72.45
320 33.69 56.31 90.00
321 36.70 57.69 74.50
322 43.31 60.98
331 46.51 76.74
332 50.24 64.76
410 14.04 75.96 90.00
411 19.47 76.37
110 0.00 60.00 90.00

110 111 35.26 90.00


210 18.43 50.77 71.56
211 30.00 54.74 73.22 90.00
221 19.47 45.00 76.37 90.00
310 26.56 47.87 63.43 77.08
311 31.48 64.76 90.00
320 11.31 53.96 66.91 78.69
321 19.11 40.89 55.46 67.79 79.11
322 30.96 46.69 80.12 90.00
331 13.26 49.54 71.07 90.00
332 25.24 41.08 81.33 90.00
410 30.6 46.69 59.04 80.12
411 33.56 60.00 70.53 90.00

111 111 0.00 70.53


210 39.23 75.04
211 19.47 61.87 90.00
221 15.79 54.74 78.90
310 43.09 68.58
311 29.50 58.52 79.98
320 36.81 80.78
321 22.21 51.89 72.02 90.00
322 11.42 65.16 81.95
331 22.00 48.53 82.39
332 10.02 6050 75.75
410 45.56 65.16
411 35.26 57.02 74.21

157
Introduction to Crystallography

Let the perpendicular distance between the origin and the (hkl) plane nearest the origin be d
[Fig.6.15], and let the direction cosines of the line A be p, q, r.

Fig.6.15: Determination of Miller Indices of a Pole

The plane will then have intercepts a/h, b/k, and c/l, where h k l are the Miller indices.
Therefore
d d d
p = cos ρ = , q = cos σ = , r = cos τ = .
a/h b/k c/l
From which it follows that
h : k : l = pa : qb : rc.
For the cubic system
H:k:l = p:q:r

PROBLEMS AND QUESTIONS


6.1. What is a stereographic projection? Explain, with neat sketches, the measurement of
angles between two planes of a crystal by using a stereographic projection.

6.2 Do the following planes all belong to the same zone: (110), (312) and (132) ? If so, what is
the zone axis? Give the indices of any other plane belonging to this zone.

6.3. Show that the planes (110) , (121) and (321) belong top the zone [111].
6.4. On a sheet of tracing paper plot all six {110} poles using a standard (100) projection, that
is the basic circle is the (100) plane.
6.5. On a sheet of stracing paper plot all four {111} poles and all {100} poles.

158
The Stereographic Projection

6.6. Combine 3 and 4 to make a standard (100) projection showing all {111}, {110} and
{100} poles.
6.7 What is a standard projection? How could you identify the Miller indices of a plane with
the help of a standard projection? How wou8ld you identify the Miller indices of a plane
when the standard projection is not available?
6.8 Explain the terms great circle and small circle. How do they appear on a stereographic
projection? Explain with neat sketches.
6.9. What is a pole? Show that a pole represents the orientation of the corresponding plane.
6.10 What is a Wulff net? Explain with a neat sketch the measurement of angles between two
planes of a crystal by using a Wulff net.
6.11 Construct a Wulff net, 18 cm in diameter and graduated at 30o intervals, by the use of
compass, dividers, and straightedge only. Show all construction lines.
In some of the following problems, the coordinates of a point on a stereographic
projection are given in terms of its latitude and longitude, measured from the center of the
projection. Thus, the N pole is 90oN, 0oE, the E pole is 0oN, 90oE, etc.
6.12. Plane A is represented on a stereographic projection by a great circle passing through the
N and S poles and the point 0oN, 70oW. The pole of plane B is located at 30oN, 50oW.
(a) Find the angle between the two planes.
(b) Draw the great circle of plane B and show that the stereographic projection is angle-
true by measuring with a protractor the angle between the great circles of A and B.
6.13. Pole A, whose coordinates are 20oN, 50oE, is to be rotated about the axes described
below. In each case, find the coordinates of the final position of pole A and show the path
traced out during its rotation.
(a) 100o rotation about the NS axis, counterclockwise looking from N to S.
(b) 60o rotation about an axis normal to the plane of projection, clockwise to the
observer.
(c) 60o rotation about an inclined axis B. whose coordinates are 10oS, 30oW, clockwise
to the observer.
6.14. Draw a standard (111) projection of a cubic crystal, showing all poles of the form {100},
{110}, {111}and the important zone circles between them. Compare with Fig.6.13.
6.15. On a standard (001) projection of a cubic crystal, in the orientation of Fig.6.13, the pole
of a certain plane has coordinates 53.3oS, 26.6oE. What are its Miller indices? Verify your
answer by comparison of measured angles with those given in Table 6.1.

159
Chapter – 7

The Reciprocal Lattice

The Bragg law is a very powerful tool and is all that is needed for an understanding of a great
many applications of x-ray diffraction. However, it is useful under certain conditions to transform
a Bravais lattice to what is known as a reciprocal lattice. For example, the Bragg law is totally
unable to explain the diffraction effects involving diffuse scattering at non-Bragg angles, and these
effects demand a more general theory of diffraction for their explanation. The reciprocal lattice
provides the framework for such a theory. This powerful concept was introduced into the field of
diffraction by the German physicist Ewald in 1921 and has since become an indispensable tool in
the solution of many problems.
The reciprocal-lattice theory of diffraction, being general, is applicable to all diffraction
phenomena from the simplest to the most intricate. Familiarity with the reciprocal lattice will,
therefore, not only provide the student with the necessary key to complex diffraction effects but
will deepen his understanding of even the simplest.

Vector Multiplication
The reciprocal lattice is best formulated in terms of vectors. We shall, therefore, first review a few
theorems of vector algebra, namely, those involving the multiplication of vector quantities. The
scalar product (or dot product) of two vectors* a and b, written a.b, is a scalar quantity equal in
Introduction to Crystallography

magnitude to the product of the absolute values of the two vectors and the cosine of the angle α
between them, or
a .b = αb cos α. b

α
a

Fig. 7.1: Scalar Product of Two Vectors

Geometrically Fig. 7.1 shows that the scalar product of two vectors may be regarded as the product
of the length of one vector and the projection of the other upon the first. If one of the vectors, say
a, is a unit vector (a vector of unit length), then a.b gives immediately the length of the projection
of b on a. The scalar product of sums or differences of vectors is formed simply by term-by-term
multiplication:
(a + b)(c – d) = (a .c) – (a .d) + (b .c) – (b .d).
The order of multiplication is of no importance: i.e.
a .b = b.a

c=axb
b

a
Fig. 7.2: Vector Product of Two Vectors

The vector product (or cross product) of two vectors a and b, written a × b, is a vector c at right
angles to the plane of a and b, and equal in magnitude to the product of the absolute values of the
two vectors and the sine of the angle α between them, or
c = a × b,
c = ab sin α.

162
The Reciprocal lattice

The magnitude of c is simply the area of the parallelogram constructed on a and b, as suggested by
Fig. 7.2. The direction of c is that in which a right-hand screw would move if rotated in such a way
as to bring a into b. It follows from this that the direction of the vector product c is reversed if the
order of multiplication is reversed, or that
a × b = - (b × a).

The Reciprocal Lattice


Corresponding to any crystal lattice. we can construct a reciprocal lattice, so called because many
of its properties are reciprocal to those of the crystal lattice. Let the crystal lattice have a unit cell
defined by the vectors a1, a2 and a3. These vectors will be the edges of a unit cell and will have
lengths a1, a2 and a3 respectively.
The corresponding reciprocal lattice has a unit cell defined by the vectors b1, b2 and b3, where
1
b1 = (a2 × a3), (7.1)
V
1
b2 = (a3 × a1), (7.2)
V
1
b3 = (a1 × a2), (7.3)
V
and V is the volume of the crystal unit cell. A vector to any corner in this lattice is
H = hb1 + kb2 + lb3,. where h, k and l are integers.
This way of defining the vectors b1, b2, b3 in terms of the vectors a1, a2, a3 gives the reciprocal
lattice certain useful properties which we will now investigate.
Consider the general triclinic unit cell shown in Fig. 7.3. The reciprocal-lattice axis b3 is,
according to Eqn. (7.3), normal to the plane of a1 and a2, as shown. Its length is given by
a1  a 2
b3 =
V
(area of parallelogram OACB)
=
(area of parallelogram OACB) ( height of cell)

1 1
= = .
OP d 001

since OP, the projection of a3 on b3, is equal to the height of the cell, which in turn is simply the
spacing d of the (001) planes of the crystal lattice. Similarly, we find that the reciprocal lattice axes
b1 and b2 are normal to the (100) and (010) planes, respectively, of the crystal lattice, and are equal
in length to the reciprocals of the spacing of these planes.

163
Introduction to Crystallography

b3

a3
P
a2 B C

O a1 A

Fig. 7.3: Location of the Reciprocal lattice Vector b3

By extension, similar relations are found for all the planes of the crystal lattice. The whole
reciprocal lattice is built up by repeated translations of the unit cell by the vectors b1, b2, b3. This
produces an array of points each of which is labeled with its coordinates in terms of the basic
vectors. Thus, the point at the end of the b1 vector is labeled 100, that at the end of the b2 vector
010, etc. It is thus seen that there is a one to one correspondence between points in the reciprocal
space and planes in the Bravais lattice and that for each Bravais lattice there is a corresponding
reciprocal lattice.
This extended reciprocal lattice has the following properties:
(1) A vector Hhkl drawn from the origin of the reciprocal lattice to any point in it having
coordinates hkl is perpendicular to the plane in the crystal lattice whose Miller indices are
hkl. This vector is given in terms of its coordinates by the expression
Hhkl = hb1 + kb2 + lb3.
(2) The length of the vector Hhkl is equal to the reciprocal of the spacing d of the (hkl) planes,
or
1
Hhkl = .
d hkl

The important thing to note about these relations is that the reciprocal-lattice array of points
completely describes the crystal, in the sense that each reciprocal-lattice point is related to a set of
planes in the crystal and represents the orientation and spacing of that set of planes.
Before proving these general relations, we might consider particular example of the reciprocal
lattice as shown in Fig. 7.4 cubic crystals. The reciprocal lattice is drawn from any convenient
origin, not necessarily that of the crystal lattice, and to any convenient scale of reciprocal
angstroms. The Eqns. (7.1) through (7.3) take on a very simple form for any crystal whose unit cell
is based on mutually perpendicular vectors, i.e., cubic, tetragonal, or orthorhombic. For such
crystals, b1, b2 and b3 are parallel, respectively, to a1, a2 and a3, while b1, b2 and b3 are simply the
reciprocals of a1, a2 and a3. In Fig. 7.4 four cells of the reciprocal lattice are shown, together with

164
The Reciprocal lattice

two H vectors. By means of the scales shown, it may be verified that each H vector is equal in
length to the reciprocal of the spacing of the corresponding planes and normal to them.
Note that reciprocal lattice points such as nh, nk, nl, where n is an integer, correspond to planes
parallel to (hkl) and having l/n their spacing. Thus, H220 is perpendicular to (220) planes and
therefore parallel to H110, since (110) and (220) are parallel, but H220 is twice as long as H110 since
the (220) planes have half the spacing of the (110) planes.
1A
a2 (010)

(100)
(110)

a1
a3

Fig. 7.4: The Reciprocal Lattice of a Cubic Crystal which has a1 = 4A.
The axes a3 and b3 are Normal to the Plane of the paper.

Other useful relations between the crystal and reciprocal vectors follow from Eqns. (7.1) through
(7.3). Since b3, for example, is normal to both a1 and a2, its dot product with either one of these
vectors is zero, or
b3. a1 = b3 . a2 = 0.
The dot product of b3 and a3, however, is unity, since (Fig. 7-3)
b3.a3 = (b3) (projection of a3 on b3)
1
=( ) (OP)
OP
= 1.
In general,
am . bn = 1, if m = n, (7.4)
= 0, if m ≠ n. (7.5)

Example 7.1: Show that reciprocal lattice vector Hhkl is normal to the plane (hkl).

Solution
Let ABC (Fig. 7.5) be part of the plane nearest the origin in the set (hkl). Then, from the definition
of Miller indices, the vectors from the origin to the points A, B and C are a1/h, a2/k and a3/l,
respectively.

165
Introduction to Crystallography

a3

H
C
(hkl)
a3/l
a2
N
B
n
a2/k
A
O a1
a1/h

Fig. 7.5: Relation Between Reciprocal lattice Vector H and Crystal Plane (hkl)

Consider the vector AB, that is, a vector drawn from A to B, lying in the plane (hkl). Since
a1 a
+ AB = 2 ,
h k
Then
a a
AB = 2 - 1 .
k h
Forming the dot product of H and AB, we have
a 2 a1
H. AB = (hb1 + kb2 + lb3). ( - ).
k h
Evaluating this with the aid of Eqs. (7.4) and (7.5), we find
H . AB = 1 – 1 =0.
Since this product is zero, H must be normal to AB. Similarly, it may be shown that H is normal to
AC. Since H is normal to two vectors in the plane (hkl), it is normal to the plane itself.

Example 7.2: Show that the reciprocal lattice vector Hhkl is the reciprocal of spacing of the hkl
planes, dhkl

Solution
To prove the reciprocal relation between H and d, let n be a unit vector (Fig. 7.5) in the direction
of H, i.e., normal to (hkl). Then

166
The Reciprocal lattice

a1
d = ON = . n.
h
But
H
n= .
H
Therefore
a1 H
d= .
h H
a 1 (hb1  kb 2  lb 3 ) 1
= . = .
h H H

Example 7.3: Reciprocity of Face Centred and Body centred Lattices

a3

a 1fcc
a fcc a2
2

a fcc
3

a1

Fig. 7.6: Basic Vectors of an Elementary FCC Cell

The basic vectors a 1f c c , a f2 c c and a f3 c c , of the elementary cell (Fig. 7.6) are given in terms of the
vectors a, b, c of the face centered cell by

a 2  a3 a 1  a3 a1  a2
a 1f c c = a f2 c c = a f3 c c =
2 2 2

In a similar way, the basic vectors a 1b c c , a b2 c c and a 3b c c of the elementary cell of a body centered
lattice are given in terms of the basic vectors of the multiple cell by (Fig. 7.7):

- a1  a 2  a 3 a1  a 2  a 3 a1  a 2  a 3
a 1b c c = a b2 c c = a 3b c c =
2 2 2

167
Introduction to Crystallography

The unit cell vectors of the reciprocal lattice of the face centred lattice are given by

a1  a 3 a1  a 2
b1f c c = x )/ ( a 1f c c , a f2 c c , a f3 c c )
2 2

1
= [ a 1 x a 1 x a 2   a 1 x a 1 x a 3 
4( a 1fcc , a fcc
2 , a 3fcc )

1
= [ a 1 x a 1  a1 x a 2  a 3 x a 1  a 3 x a 2 ]
4( a 1fcc , a fcc
2 , a 3fcc )

1
= [ a1 x a 2  a 3 x a1  a 3 x a 2 ]
4( a 1fcc , a fcc
2 , a 3fcc )

a3

a b2 c c a 1b c c
a2
a b3 c c
a1

Fig. 7.7: Basic Vectors of an Elementary BCC Cell

Now ( a 1fcc , a fcc fcc fcc fcc fcc


2 , a 3 ) = a1 . ( a 2 x a3 )

a2  a3 a  a1 a1  a 2
= ( ). ( 3 x )
2 2 2
1
= [a 2  a 3 ).{a 3 x (a 1  a 2 )  a 1 x (a 1  a 2 )}]
8

1
= [(a 2  a 3 ).{a 3 x a 1  a 3 x a 2  a 1 x a 1  a 1 x a 2 )]
8

168
The Reciprocal lattice

1
= [a 2  a 3 ).(a 3 x a 1  a 3 x a 2  a 1 x a 2 )]
8

=
1
[a 2 .(a 3 x a 1 )  a 2 . (a 3 x a 2 )  a 2 . (a 1 x a 2 )  a 3 .(a 3 x a 1 )  a 3 . (a 3 x a 2 )  a 3 . (a 1 x a 2 ) ]
8

1
= [a 2 .(a 3 x a 1 )  a 3 .(a 1 x a 2 ) ]
8
1
= (a 1 , a 2 , a 3  .(a 1 , a 2 , a 3 )
8
1
= (a 1 , a 2 , a 3 )
4

Or 4 ( a 1fcc , a fcc fcc


2 , a 3 ) = (a 1 , a 2 , a 3 )

1
Therefore, b1f c c = [ a1 x a 2  a 3 x a1  a 3 x a 2 ]
4( a 1fcc , a fcc
2 , a 3fcc )

1
= [ a1 x a 2  a 3 x a 1  a 3 x a 2 ]
( a1 , a 2 , a 3 )

1
= [ a1 x a 2  a 3 x a1  a 2 x a3 ]
( a1 , a 2 , a 3 )

a1 x a 2 a2 x a3 a 3 x a1
= - +
( a1 , a 2 , a 3 ) ( a1 , a 2 , a 3 ) ( a1 , a 2 , a 3 )

a2 x a3 a1 x a 2 a 3 x a1
=- + +
( a1 , a 2 , a 3 ) ( a1 , a 2 , a 3 ) ( a1 , a 2 , a 3 )

= -b1 + b2 + b3
-2b1  2b 2  2b 3
This may also be written as =
2
This relation shows that the reciprocal lattice of a face-centered lattice is a body centered lattice
whose multiple cell is defined by 2b1, 2b2, 2b3.

169
Introduction to Crystallography

Example 7.4: Reciprocity of Volumes of Direct and Reciprocal Lattices


Let
The unit cell vectors in the direct lattice be a1, a2 and a3
The unit cell vectors in the reciprocal lattice be b1, b2 and b3
The volume of the direct lattice = V and
The volume of the reciprocal lattice be V*.

Then V* = b1. (b2 x b3)

a 2 x a 2 a 3 x a1 a 1x a 2
= .( x )
V V V
(a 2 x a 3 ) [ (a 3 x a 1 ) x (a 1 x a 2 )
=
V3
(a 2 x a 3 ) [{ (a 3 x a 1 ) . a 2 } a 1 - (a 3 x a 1 ) . a 1 } a 2 ]
=
V3
[ Since A x B x C = (A.C)B – (A.B)C
and A = (a3 x a1), B = a1 and C = a2]
(a 2 x a 3 ) (a 3 x a 1 . a 2 ) a 1
=
V3

V2 1
= =
V3 V
So the volume of direct lattice and reciprocal lattice are inverse to each other.

Used purely as a geometrical tool, the reciprocal lattice is of considerable help in the solution of
many problems in crystal geometry. Consider, for example, the relation between the planes of a
zone and the axis of that zone. Since the planes of a zone are all parallel to one line, the zone axis,
their normal must be coplanar. This means that planes of a zone are represented, in the reciprocal
lattice, by a set of points lying on a plane passing through the origin of the reciprocal lattice. If the
plane (hkl) belongs to the zone whose axis is [uvw], then the normal to (hkl), namely, H, must be
perpendicular to [uvw]. Express the zone axis as a vector in the crystal lattice and H as a vector in
the reciprocal lattice:
Zone axis = ua1 + va2 + wa3,
= hb1 + kb2 + lb3.
If these two vectors are perpendicular, their dot product must be zero:

170
The Reciprocal lattice

(ua1 + va2 + wa3) . (hb1 + kb2 + lb3) = 0,


hu + kv + lw = 0
Diffraction and the Reciprocal Lattice
Transforamation of the Bravais lattice into a reciprocal lattice allows a simple geometrical
representation of the Bragg diffraction law. This can be shown by considering how x-rays
scattered by the atom O at the origin of the crystal lattice (Fig. 7.6) are affected by those scattered
by any other atom A. The coordinates of A with respect to the origin are pa1, qa2 and ra3, where p,
q and r are integers. Thus,
OA = pa1 + qa2 + ra3.
Let the incident x-rays have a wavelength λ, and let the incident and diffracted beams be
represented by the unit vectors So and S, respectively. So, S and OA are, in general, not coplanar.

(S – So)

So

(S – So) O θ

n
-S

v
So u
A

Fig. 7.6: X-ray Scattering by Atoms at O and A.

To determine the conditions under which diffraction will occur, we must determine the phase
difference between the rays scattered by the atoms O and A. The lines Ou and Ov (Fig. 7.7) are
wave fronts perpendicular to the incident beam So and the diffracted beam S, respectively. Let δ be
the path difference for rays scattered by O and A. then
δ = uA + Av
= Om + On
= So. OA + (-S). OA
= - OA. (S – So)

171
Introduction to Crystallography


The corresponding phase difference is given by multiplying this phase difference by
λ
2π
Φ=
λ
S - S0
= - 2π ( ). OA. (7.6)
λ
The scattered waves will be in phase and maximum reinforcement will occur if the phase
difference is 2π, which will occur if the path difference is equal to the wave length λ, but in
general there will be a phase difference φ given by 2π/λ times the path difference. We have also
seen that if the atoms are in a crystalline array, the diffraction is equivalent to a reflection from a
crystal plane.
The unit vectors S and So are at equal angles to the reflecting plane; therefore, S – So is a vector
parallel to the vector Hhkl that represents the reflecting plane in the reciprocal space (Fig. 7.7).
Diffraction can now be related to the reciprocal lattice by expressing the vector (S – S0)/λ as a
vector in that lattice. Let
S - S0
= hb1 + kb2 + lb3.
λ
This is now in the form of a vector in reciprocal space but, at this point, no particular significance
is attached to the parameters h, k and l. They are continuously variable and may assume any
values, integral or no integral. Equation (7.6) now becomes
φ = 2π (hb1 + kb2 + lb3).(pa1 + qa2 + ra3)
= 2π (hp + kq + lr).
Diffraction will occur only if reinforcement occurs, and this requires that φ be an integral multiple
of 2π. This can happen only when (hp + kq + lr) = n. Therefore, the condition for diffraction is that
the vector (S – So)/λ end on a point in the reciprocal lattice, or that
S - S0
= H = hb1 + kb2 + lb3 (7.7)
λ
where h, k, and l are now restricted to integral values.
The Bragg law can be shown to be equivalent to the condition just stated. As shown in Fig. 7.6, the
vector (S - S0) bisects the angle between the incident beam S0 and the diffracted beam S. The
diffracted beam S can, therefore, be considered as being reflected from a set of planes
perpendicular to (S - S0). In fact, Eqn. (7.7) states that (S - S0) is parallel to H, which is in turn
perpendicular to the planes (hkl). Let θ be the angle between S (or S0) and these planes. Then,
since S and S0 are unit vectors,
(S - S0) = 2 sin θ.

172
The Reciprocal lattice

Therefore
2 sin θ S - S0 1
= =H= ,
λ λ d
or
λ = 2d sin θ.

Ewald’s Sphere of Reflection


Ewald showed that a simple geometrical construction in reciprocal space gives the condition of
reinforcement of the scattered beams in a way that proves to be extremely useful in diffraction
work. It expresses the condition for diffraction in terms of a sphere in the reciprocal space. This
sphere is known as the reflection sphere or the Ewald’s sphere.

P (hkl)

O
So
λH C
1
λ

Sphere of
reflection

Fig. 7.7: The Ewald Construction.


Section containing the incident and the reflected beam vectors

The sphere is erected with the vector representing the incident beam as a radius. This vector is
So/λ; therefore, it is drawn with a length 1/λ because So is a unit vector. The terminal point O of
this vector is taken as the origin of the reciprocal lattice, drawn to the same scale as the vector
S0/λ. A sphere of radius 1/λ is drawn about C, the initial point of the incident-beam vector. Then
the condition for diffraction from the (hkl) planes is that the point hkl in the reciprocal lattice
(point P in Fig. 7.7) touch the surface of the sphere, and the direction of the diffracted beam vector
S/λ is found by joining C to P. When this condition is fulfilled, the vector OP equals both Hhkl and
(S - S0)/λ, thus satisfying Eq. (7.7). Since diffraction depends on a reciprocal-lattice points
touching the surface of the sphere drawn about C, this sphere is known as the “sphere of
reflection.”
Our initial assumption that p, q and r are integers apparently excludes all crystals except those
having only one atom per cell, located at the cell corners. For if the unit cell contains more than

173
Introduction to Crystallography

one atom, then the vector OA from the origin to “any atom” in the crystal may have non integral
coordinates. However, the presence of these additional atoms in the unit cell affects only the
intensities of the difffracted beams, not their directions, and it is only the diffraction directions
which are predicted by the Ewald construction. Stated in another way, the reciprocal lattice
depends only on the shape and size of the unit cell of the crystal lattice and not at all on the
arrangement of atoms within that cell.

X-ray Diffraction Methods


The orientation and unit cell dimensions of a crystal determine the arrangement of the reciprocal
lattice points. The condition of diffraction is that the sphere of reflection touches a reciprocal
lattice point. The common methods of x-ray diffraction are differentiated by the methods used for
bringing reciprocal-lattice points into contact with the surface of the sphere of reflection. The
radius of the sphere may be varied by varying the incident wavelength (Laue method), or the
position of the reciprocal lattice may be varied by changes in the orientation of the crystal
(rotating-crystal and powder methods). However, a detailed discussion of how these are actually
accomplished is beyond the scope of this course.

PROBLEMS AND QUESTIONS


7.1 Define reciprocal lattice? Consider any crystal lattice and explain how a reciprocal lattice is
drawn. What are its important properties?
7.2 Explain with suitable examples, how reciprocal lattice helps the study of crystal geometry.
7.3 If the axis of zone has indices [uvw], then any plane belongs to this zone whose indices
(hkl) satisfy the relation hu + kv + lw = o – prove.
7.4 Derive the conditions of diffraction in the reciprocal lattice. How are these conditions
satisfied in the various x-ray diffraction methods?
7.5 What is the Ewald’s sphere of reflection? How is it erected? How does it help the study of
crystals?
7.6 Show that the spacing dhkl of the plane (hkl) is he reciprocal of the length of the reciprocal
lattice vector rhkl.
7.7 Prove the reciprocity of the volumes of direct and reciprocal lattice.
7.8 Show that the reciprocal lattice vector Hhkl is the reciprocal of spacing of the hkl planes, dhkl
7.9 Show that reciprocal lattice vector Hhkl is normal to the plane (hkl).
7.10 Explain the terms zone and zone axis. Do the following planes belong to the same zone:
(110), (311), (132)? If so what is the zone axis?

174

You might also like