The Crystal Structure: A U - ST
The Crystal Structure: A U - ST
The Crystal Structure: A U - ST
1.1 INTRODUCTION
Intermolecular attraction is minimum in the gaseous state and this disappears completely when
the gas is ideal. The interaction is stronger in liquids and is strongest in solids. Thermal motion
of the molecules increases or decreases by raising or lowering of temperature. The attractive
interaction between the molecules tries to keep them together and the thermal motion is opposed
to that. Hence, it is possible to change a substance from one state to another by changing its
temperature. If a liquid is allowed to cool slowly, the molecules will arrange themselves in an
orderly manner and this will finally result in a crystalline solid. If, on the other hand, cooling is
rapid, the molecules will not be able to arrange themselves in order. Rapid densification will give
a glass or an amorphous solid. It is not true that the molecules and atoms in a solid have rigidly
fixed coordinates. But they move only a small distance about their equilibrium positions. In this
book, we are concerned with crystalline solids and the word solid and crystalline solid will be used
synonymously.
What is the stable state of a given material will depend on its free energy. The stable state
will be the one that has the lowest free energy under the given conditions. Free energy A is
related to internal energy U and entropy S of the system as
A = U ST ...(1.1)
Internal energy is lowered by an orderly arrangement of the atoms, molecules or ions as that
will lead to maximum energy of interaction. But this will minimise the entropy. Since internal
energy and entropy make opposite contribution to free energy, the state of matter will be
determined by the relative contributions of U and ST to A. If interaction is strong, U is highly
negative and ST can overcome the contribution of the former only at high temperatures. Such a
substance will remain as a solid even at a relatively high temperature.
The basic feature of a crystalline solid is the regular arrangement of the atoms and molecules.
At the macro level, this translates into crystals having sharp boundaries with clear cut shapes. It
is these beautiful shapes of natural crystals that attracted human attention for ages. This beautiful
shape and colour added to their value as gems.
Early study of crystals began with the observation of their shapes and this is known as
Geometric Crystallography. The description of crystal symmetry in terms of point lattice began in
the mid-nineteenth century. This was followed by X-ray crystal structure determination following
the work of Laue and Bragg on X-ray diffraction by crystals. In the second quarter of the twentieth
century, the presence of lattice defects and their role in determining the properties of crystals
were recognised. We shall not try to follow the development of the subject in a chronological
order as the development of knowledge in an area of science does not take place in the same
2 Solid State Chemistry
logical way as one would like to see it. But only after enough knowledge gets accumulated that
a subject is put in a logical perspective. Here we shall follow the rational rather than the
chronological course of development of the subject.
Crystal lattice
It is easy to imagine a crystal as a periodic
arrangement of points as shown in figure 1.1. A point
may represent an atom or a group atoms arranged
around it in a real crystal. Let us begin with a single
point. Repeated translation of this point through a
fixed distance = (a periodic translation) will generate
a linear array of points. This movement is denoted
^
by a translation vector t1. If we add a second
^
translation t2, it will generate periodically repeating
points on a plane and this is known as a plane lattice.
^
If a third translation t3 is added, we get a three
dimensional arrangement of the points that is called Figure 1.1: A two dimensional plane lattice
a space lattice.
The lattice points are imaginary. In a real crystal, they are occupied by atoms or groups of
atoms that are arranged in a regular fashion about the lattice points. This atom or the group of
atoms is the basis and the arrangement of the imaginary points is the lattice. The real crystal is
then:
basis + lattice = crystal.
A two dimensional pattern as is usually found on a curtain cloth or a wall paper is analogous
to a two dimensional crystal lattice. We can have an array of squarely arranged points
(Figure 1.2a) or the points may be arranged along inclined lines (Figure 1.2b). We can select a
single motif and place this motif in the same way about each lattice points. This will give two
different patterns (Figure. 1.2c and 1.2d). By selecting a different motif, we may get a still different
pattern and a large number of patterns can be generated from a limited number of motifs and
lattice arrangements.
Unit cell
We have seen that two noncollinear translations give rise to a plane lattice and introducing a third
translation (not on the same plane) generates a space lattice. Since any line joining two lattice
points is a translation and there can be wide choice of translation, the question arises as to which
two translation should one select to describe a plane lattice. A few such combinations are shown
in figure 1.3. It is seen that they generate two dimensional units called unit cell. Combination of
^ ^ ^ ^
t 1, t 2 or t 3, t 4 leads to cells having only one lattice point per cell. These are known as primitive
^ ^
unit cell. The combination t 5, t 6 generates a double cell. There can be many more multiple cells. The
unit cell of a lattice can be primitive or multiple. A repetition of the two dimensional unit cell by
translation in two directions generates the plane lattice. This may be extended to three dimensional
lattice that may be generated by translation of a three dimensional unit cell.
The Crystal Structure 3
(a) (c)
(b)
(d)
Figure 1.2: Two different plain lattices with identical motif leading to two different patterns
t3
t4 t6
t5
t2
t1
(i) translation, (ii) rotation, (iii) translation + rotation and (iv) reflection. It was shown by Federov
and independently by Barlow that it is not possible to have arrangement of lattice points other
than than these 230 types that can repeat itself infinitely in three dimensions.
The 14 types of Bravais lattices are shown in figure 1.5.
Simple bcc
tetragonal orthorhombic
(Contd.)
The Crystal Structure 5
Rhombohedral rhombohedron a = b = c
= = 90
Monoclinic parallelopiped a b c
bc parallelopiped = = 90;
90
planes have different hkl or Miller indices. This is illustrated in figure 1.6. Here, the plane nearest
1 1 1
to the origin and cutting the b axis at , the a axis at and the c axis at is shown. This
2b 3a 2c
1 1 1
makes intercept along a axis, intercept along b axis and along c axis.
3 2 2
c
b
b
a
a
Figure 1.6: (322) set of planes
6 Solid State Chemistry
We can write
a b c
intercept 1/3 1/2 1/2
reciprocal 3 2 2
Hence this plane and a set of parallel planes separated by a distance d have the Miller
indices (322).
It should be noted that a point on a paper actually represents line of points when one
considers the three dimensional lattice. Hence on the plane of paper a line of points actually
represents a plane.
The major advantage of the Miller indices is that it permits to express interplanar distance
dhkl of a set of hkl planes in terms of lattice parameters a, b, c, , and .
For a cubic crystal
a
dhkl =
h + k2 + l2
2
e j
1.3 DIFFRACTION OF X-RAYS
In 1912, von Laue first suggested that since the lattice points in a real crystal are occupied by
atoms, the crystal lattice should act as a three-dimensional diffraction grating for X-rays. This
should happen because X-rays have wavelength of the dimension of interplanar distances in a
real crystal. Shortly after this, W.L. Bragg showed that wavelength of the X-ray undergoing
diffraction by a crystal is related to the interplanar distances by the famous Braggs equation.
2a
a
b b
a b
2a
a 2a
a a a
(100) (010) (001)
c c
c
3a
3a 3a
2a
a
a 2a 3a
b b a 2a 3a b
3a 3a
a
a a
(111) (101) (011)
reflection is also and a part of the intensity will pass though the crystal undeviated from its
path. Reflection is caused by the interaction of the electromagnetic radiation with the electrons of
the atoms in the lattice. In order that the intensity of the reflection is sufficiently strong, reflected
waves from the successive planes separated by dhkl should be in phase.
From figure 1.8, it is seen that the path difference of the waves from successive planes is
2d sin . In order that the waves travelling from successive planes are in phase the condition
n = 2dhkl sin (1.2)
should be satisfied. This is Braggs condition of reflection and is known as Braggs law.
Path diff. = BC + BD
nl = 2AB sin q
= 2d sin q
A
q
q q
C B D
2 1 d hkl
sin hkl = = (1.3)
dhkl 2
Here, we have tried to relate the magnitude of the reciprocal lattice vector to diffraction
angle and the wavelength of the X-ray. In order to see the geometric consequence of this equation,
let us imagine a sphere of radius 1/ = AO as shown in figure 1.9.
8 Solid State Chemistry
shkl
2q
A q q
q C O
Figure 1.9: Relation between reciprocal lattice point and X-ray diffraction
Let AO also be the direction of the X-ray beam incident on the crystal plane at C, the centre
of the sphere. If is the Braggs angle, the reflected beam will strike the sphere (shown as a circle
here) at the point P making an angle 2 with the passing beam. It should be noted that the angle
between the incident beam and AP is .
Since 2 sin hk l/ = 1/dhk l, we can see that 1/dhk l = OP. Thus, the point P is the reciprocal
lattice point for the set of planes from which the X-ray beam is reflected. Also, since the Braggs
diffraction conditions are satisfied, the diffracted beam will touch the sphere (circle in the picture)
at point P. The same will be the case with other set of parallel planes except that they will strike
the sphere at some other point which are the reciprocal lattice points for the respective set of
planes. The three-dimensional sphere is called the sphere of reflection or Ewald sphere.
2q
X-ray
O
2q
Diffracted beam
special techniques and is not always easy. However, it is possible to get important structural
information by recording the X-ray diffraction pattern of the powdered polycrystalline samples.
This is commonly known as the powder method.
Synthetic chemists prepare many solid materials in the laboratory. Before proceeding fur-
ther, one would like to know whether the desired structure has been formed. This is done by
recording the X-ray powder diffraction pattern of the material and comparing this with that of
the known pattern. Extensive powder diffraction data have been compiled in the ASTM X-Ray
Data Files that makes such comparison possible. Moreover, the powder diffraction patterns can
provide important structural information such as the type of Bravais lattice, size of the unit cell
and the space group. In the case of a simple crystal, it is even possible to determine the coordi-
nates of all the atoms by analyzing these patterns. No surprise that this method has become a
useful tool of the solid state chemists. If a set of Miller planes satisfy the Braggs condition, the
reflected beam will emerge making an angle 2 with the undeflected beam. Since the crystals are
randomly oriented, the same Miller planes in another crystal may satisfy Braggs condition, but
the deflected beam at angle 2 may have a different direction as shown in the figure 1.10. Since
the number of crystals is very large, all kind of orientations are possible and the diffracted X-ray
will form a cone with angle 4.
intensity of the lines is not very satisfactory and it is time consuming. For these reasons, this
method is hardly used in recent times and has been replaced by the automatic X-ray powder
diffractometer. We shall not discuss here the different sources of error while recording the
X-ray diffraction pattern using the Debye-Scherrer camera as it is rarely used nowadays.
Diffraction
cones
Film
(a)
(b)
Figure 1.11: Diffraction from a powdered sample using Debye-Scherrer camera: (a) diffracted
cones and (b) the part of the cones as pairs of lines in the uncoiled film
80 70 60 50 40 30 20
2q
which is used for finding a. Knowing a, c can be found out using equation (1.8). Powder patterns
of crystals of low symmetry are difficult to index and will not be considered here.
12 Solid State Chemistry
321 222
400 311
410 310
330 221
411 300
331
420 220 211 210 200 111 110 100
20.00
15.00
d=0
10.00
a,
5.00
0
5.00 10.00 15.00 20.00
d,
111
200
220
311
222
400
331
420
422
511
333
440
531
600
442
620
533
622
444
711
511
640
(b)
(a)
200
220
222
400
420
422
440
600
442
620
622
444
640
642
Figure 1.14: X-ray powder patterns (a) KCl and (b) NaCl
The first thing that strikes us is that the angles (given by the position of the lines) due to
the same set of Miller planes are slightly larger in NaCl. This is because the NaCl unit cell is
slightly smaller than that of KCl. The more striking difference is the absence of certain lines like
(111), (311), (511) etc. in KCl although these lines are present in powder pattern of NaCl. From
systematic absence rules for fcc crystals, we know that the lines with mixed indices should be
absent for both NaCl and KCl. This is indeed so. Further, for NaCl we see that the successive
orders of (111) planes (these are 111, 222, 333, 444 etc.) are alternately weak and strong. For
example, the reflection from (111) is weak and that from (222) is strong and so on. We know that
in the binary fcc compounds like NaCl or KCl, the (111) planes are alternately occupied by Na+
(or K+) and Cl ions. If the scattered waves from two or more such planes containing only sodium
are in phase and intensify the reflection, the planes containing sodium will be out of phase with
planes containing chlorine and will interfere destructively thus diminishing the intensity. This is
the reason why (111) reflection in NaCl is weak and (222) reflection is strong. For KCl, the
alternate (111) planes have potassium and chlorine. These two ions have the same number of
electrons and identical scattering power. So far as X-ray is concerned, K+ and Cl are identical.
Since the radiation scattered by these planes are out of phase, the reflected waves will get completely
annihilated and the reflections from (111), (333) etc. will not be seen.
X-ray scattering power of an atom f is proportional to the number of electrons in the atom.
One need not consider the scattering power of all the infinite number of atoms in a crystal. It is
enough to consider the atoms of one unit cell. We must know the location of the atoms in a unit
cell. This will enable us to determine what is called the structure factor (Fhkl) for a reflection hkl.
Fhkl = f1 e 2i( hx1 + ky1 + lz1 ) +. . . + f n e 2i( hxn + kyn + lzn ) (1.10)
where N is the number of atoms in the unit cell and xnynzn defines the coordinate of an atom.
Taking the position of a cation as the origin, NaCl crystal will have in its unit cell Na+ at 000,
11 1 1 11 111 1 1 1 1
0 , 0 and 0 and Cl ions at , 00 , 0 and 00 . Putting these coordinates
22 2 2 22 222 2 2 2 2
and simplifying, we get
Since ein is +1 when n is even and 1 when n is odd, the Miller indices hkl with all even
or all odd will make the terms within the second square brackets in equation (1.11) equal to 4.
With mixed hkl, this part of the equation becomes zero. So the structure factor for NaCl is
Perfection in the crystal will minimize the free energy by making the internal energy UL more
negative, and the contribution from entropy ST will try to minimize the same by making the
crystal more imperfect thus increasing S. The equilibrium structure at any temperature other than
0 K will have some disorder or imperfection. We see that the presence of defects in a crystal is
a thermodynamic requirement for stability. The main defects in a crystal are of three types: point
defects, dislocations and grain boundaries.
+ + + + + +
+ + + + +
+
+ + + + + +
+ + + + +
+ + + + + +
( a) (b )
Figure 1.15: (a) Frenkel defect and (b) Schottky defect in an ionic crystal
The Crystal Structure 17
positive charges and will be represented as VO. If it captures one electron, its effective charge
will be unipositive and the symbol for it is VO . If it captures a second electron, its effective charge
will be zero and the symbol is VOX , although its true charge is 2. Similarly, the charge of other
types of point defects should be included in the symbol. Thus, if a bivalent metal ion goes to a
normally unoccupied interstitial site, the effective charge of the interstitial site becomes +2 be-
cause originally this site did not have any charge. In this case, however, the real charge too is +2.
The inclusion of the effective charge in the notation of the point defects is important as complete
balance of effecticve charge will have to be maintained in writing balanced equations for defect
reactions.
MM + OO l VM + VO + MM + OO
The regular lattice sites on the right hand side means that the formation of Schottky defect
requires that new lattice sites are created. Cancelling common terms, we write
O l VM
+ VO
where O stands for a perfect lattice. Here we have assumed that the doubly charged vacancies
predominate. It is obvious that the doubly charged vacancies can capture or release electrons and
can either become singly charged or neutral. If these species are to be included in the equation,
the corresponding electrons or holes are also to be added.
A Frenkel defect is created by transferring a regular ion to an interstitial site. Suppose the
interstitial ion is a cation as is generally the case, we can write
MM l VM + Mi
assuming that the doubly charged interstitials and vacancies predominate.
The Crystal Structure 19
We have considered the important types of point defect. Next we shall look into the concen-
tration of the thermodynamic defects under specific conditions. The equilibrium concentration of
the inherent thermodynamic defects is dependent on temperature. It is possible to calculate this
concentration at any given temperature provided the energy required for their formation is known.
Here we shall derive the equations that relate the defect concentration to temperature.
a f a
S = k N ln N N n ln N n n ln n f (1.18)
It is now possible to write
a
A = n kT N ln N N n ln N n n ln nf a f (1.19)
Differentiating A with respect to n and equating to zero (condition of equilibrium concen-
tration of defects), we get
= kT
n
a f a
N n ln N n n ln n f
This will give
N n
= kT ln
n
and
n FG IJ
N n
= exp
kT H K (1.20)
20 Solid State Chemistry
Since N >> n,
n FG IJ
N
= exp
H kT K
If EV is the energy necessary to create one mole of vacancy, then
EV = Avogadros number and
n E FG IJ
N
= exp V
RT H K (1.21)
Thus, knowing the energy of fromation of vacancies, it is possible to calculate the number
of vacancies per mole of metal atoms at a given temperature.
There will be ns anion vacancy since stoichiometry is MX. The number of ways in which ns
anion vacancies and (N ns) anions can be distributed among N anionic sites is also equal to w
given by equation (1.22). The total probability W = w.w. The configurational entropy is
N!
S = 2k ln
bN n g ! n !
s s
(1.23)
which means
b g b g
S = 2k N ln N N n s ln N n s n s ln ns (1.24)
We can now write
b g b
A = ns 2kT N ln N N ns ln N ns ns ln ns g (1.25)
In order to minimize A, we differentiate the above equation with respect to ns and equate
that to zero. After rearranging the result, we get
ns FG IJ
H 2kT K
s
= exp (1.26)
N
If Es is the energy of formation of one mole of Schottky defect in the crystal MX, then
ns FG E IJ
H 2RT K
s
= exp (1.27)
N
The expression for Schottky defect concentration for crystal of composition MX2 can be
similarly deduced.
The Crystal Structure 21
Let the total number of interstitial sites be N*. The number of ways in which nf interstitial
atoms and (N* nf) unoccupied interstitial sites can arrange gives the probability
N *!
eN * n j ! n !
w* = (1.29)
f f
Solving
F AI
GH n JK f
= 0 (1.32)
V ,T
we get
n 2f F I
( N n f ) ( N * n f )
= exp GH kT JK f
1.6.6 Dislocations
Dislocations also known as line defects and are of two kinds. They are edge dislocation and screw
dislocation.
An edge dislocation is what would happen if a half plane is inserted into a crystal. The
dislocation extends along a line perpendicular to the plane of the paper. The structure is distorted
near the dislocation and several lattice constants are to be covered before this distortion disappears.
Edge dislocation can move along the crystal under a shear force. Suppose force is applied in
22 Solid State Chemistry
opposite directions at the two ends of a crystal. Instead of separating it into two parts that would
need a large number of bonds to be broken, atoms on one side can just move a short distance.
This will make the dislocation move along the crystal in the direction of the shear force. Movement
of the edge dislocations in the crystal are responsible for their plastic property. Edge dislocation
may form accidentally during crystal growth or may be produced by bending a crystal. Dislocation
density in a common crystal is of the order 106/cm2.
Edge dislocation are given the symbol as shown in figure 1.16 where the dislocation line
is perpendicular to the plane of the paper. Dislocations can be identified by taking electron
microscope pictures at magnification 105 or higher where the lattice planes appear as straight
lines. Dislocations show up as disruption in these straight lines.
The other type of line defect is known as screw dislocation. Screw dislocation results from
movement of one part of a crystal relative to another. The distance of movement is less than a
lattice constant and hence the coordination number of the atoms near the screw dislocation does
not change. The screw dislocation only leads to a distortion of the bonds in its vicinity. If one
moves from lattice point to lattice point around the dislocation, one will move up as if in a spiral
staircase. The name screw dislocation is derived from this (see figure 1.17). Dislocations severely
affect the mechanical properties of solids.
SUGGESTED READING
1. Azaroff, L.V. Introduction to Solids, TMH edition, Tata McGraw-Hill, New Delhi.
2. Azaroff, L.V. and Buerger, M.J. The Powder Method in X-ray Crystallography, McGraw-Hill, 1958.
3. Epifanov, G.I. Solid State Physics, MIR, Moscow, 1979.
4. Hannay, N.B. Solid State Chemistry, Prentice-Hall, New Jersey, 1967.
5. Keer, H.V. Principles of the Solid State, Wiley Eastern, New Delhi, 1993.
6. Kittel, C. Introduction to Solid State, 3rd edition, John Wiley, 1966.
7. Megaw, H.D. Crystal Structure: A Working Approach, Saunders College Publ., 1973.