0% found this document useful (0 votes)
20 views

allcrystall (1)

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
20 views

allcrystall (1)

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 115

Department of Chemistry

Dhaka University
Course: Elements of Chemical Crystallography, No: CH 442 (2 Credits)

Books Recommended:

1. G. H. Stout and L. H. Jensen, X-ray Structure Determination – A Practical


Guide, John Wiley & Sons.
2. M. F. C. Ladd and R. A. Palmer, Structure Determination by X-ray
Crystallography, Plenum Press.
3. C. Giacovazzo (edited), Fundamentals of Crystallography, International Union
of Crystallography Oxford University Press.
4. D. M. Adams, Inorganic Solids – An Introduction to Concepts in Solid-State
Structural Chemistry, John Wiley & Sons.
5. A. R. West, Solid State Chemistry and its Applications, John Wiley & Sons.

hcp unit ccp unit

1
Chapter 1: Solids
Solid: Solid is one of the three major states of matter. It is characterized by structural rigidity and
resistance to changes of shape or volume. In the solid state, attractive forces predominate. They
confine atoms, molecules and ions to fixed positions, and they limit their motions to a mere jiggling
in place.

Based on the nature of the forces that hold the component atoms, molecules, or ions together, solids
may be formally classified as ionic, covalent, metallic or molecular.

Ionic solids: An ionic solid is a solid that consists of cations and anions held together by the
electrical attraction of opposite charges (ionic bonds). Ionic solids are formed between highly
electropositive and electronegative elements when electron transfer has
occurred between the atoms. Examples are NaCl, KCl, CaS, etc.

The strength of the attractive forces depends on the charge and size of the
ions that compose the lattice and determines many of the physical
properties of the crystal.
NaCl
Covalent solids: A covalent solid is a solid that consists of atoms held
together in large networks or chains by covalent bonds. Covalent solids
are formed when elements have similar electronegativity. Sharing of
electrons and overlap of atomic orbitals take place, examples: SiC and
diamond. Because all of the bonds in this structure are equally strong,
covalent solids are often very hard and they are notoriously difficult to
melt. Diamond is the hardest natural substance and it melts at 3550oC. Diamond

Metallic solids: A metallic solid is a solid that consists of positive cores of atoms held together by a
surrounding sea of electrons. Outer electrons of metal atoms are free to move resulting in high
electrical and thermal conductivity. Example: almost all metals and their alloys, such as gold, brass,
steel.

The strength of metallic bonds varies dramatically. For example, cesium melts at 28.4°C, and
mercury is a liquid at room temperature, whereas tungsten melts at 3680°C. Metallic bonds tend to
be weakest for elements that have nearly empty (as in Cs) or nearly full (Hg) valence subshells, and
strongest for elements with approximately half-filled valence shells (as in W).

2
Electron sea model Copper

Molecular solids: Molecular solids consist of molecules held to each other by dipole-dipole
interactions, London dispersion forces, or hydrogen bonds, or any combination of these.
Because the intermolecular interactions in a molecular
solid are relatively weak compared with ionic and
covalent bonds, molecular solids tend to be soft, low
melting, and easily vaporized. Examples of molecular
solids include hydrocarbons, ice, sugar, fullerenes, sulfur
Dry ice, or frozen carbon dioxide
and solid carbon dioxide.

Properties of the major classes of solids:


Ionic solids Covalent solids Metallic solids Molecular solids
Poor conductors of Poor conductors of Good conductor of Poor conductors of
heat and electricity heat and electricity* heat and electricity heat and electricity
Relatively high High melting point Melting points Low melting point
melting point depends strongly on
electron configuration
Hard but brittle; Very hard and brittle Easily deformed Soft
shatter under stress under stress; ductile
and malleable
Relatively dense Low density Usually high density Low density
Dull surface Dull surface Lustrous Dull surface

*Many exceptions exist. For example, graphite has a relatively high electrical conductivity
within the carbon planes, and diamond has the highest thermal conductivity of any known
substance.

Again, from a structural viewpoint, solids can be divided into two distinct classes:

3
Crystalline solid: A crystalline solid is a solid material, whose Atoms, ions or molecules
constituent atoms, molecules, or ions are arranged in an orderly
repeating pattern extending in all three spatial dimensions. Some
common examples of crystalline solids are sodium chloride,
sulphur, diamond, sugar, etc. About 95% of all solids can be
described as crystalline.

Amorphous solid: An amorphous solid is a solid in which there is Atoms, ions or molecules
no long-range order or repeating pattern in the positions of its
atoms or molecules. The common examples of amorphous solids
are glass, rubber, fused silica, plastics, etc.

Crystalline solid: From a structural point of view, crystalline solids are of three distinct types:

(i) Ideal crystals: Have a very high degree of internal order, in other words,
the lattice arrays of atoms, molecules or ions persist without flaws in all
directions in the crystal. Ideal crystal does not really exist. For a ideal
crystal,
I ∞ F (where, I = intensity, F = structure factor).

(ii) Ideally imperfect crystals: Lattice arrays of objects (atoms, molecules


or ions) are not continuous throughout the crystals but are broken up into
blocks by discontinuities at the block boundaries. It is like a mosaic in
which each part is perfect but tipped slightly relative to its neighbors. An
ideally imperfect crystal is also not an accurate representation of crystals
existing in nature. For an ideally imperfect crystal,
I ∞ Fm (where, I = intensity, F = structure factor, m =1 - 2)

(iii) Imperfect crystals: Having large deviation from a highly ordered


arrangement of atoms, molecules or ions. Most of the naturally occurring
crystals are imperfect. For an imperfect crystal,
I ∞ F2 (where, I = intensity, F = structure factor)

4
Distinction between crystalline and amorphous solids
Crystalline solids Amorphous solids

1. They have a definite geometrical shape due 1. They do not have any orderly pattern of
to finite and orderly arrangement of particles in arrangement of particles and, therefore, do
three-dimensions. not have any definite geometrical shape.
2. Crystalline solids have crystal symmetry i.e., 2. Amorphous solids do not have symmetry.
when a crystalline solid is rotated about an axis,
its appearance does not change.
3. The melting point of crystalline solids is 3. Amorphous solids don't have particular
finite. melting point. They melt over a wide range of
temperature.
4. There is a sudden change in volume when it 4. There is no sudden change in volume on
melts. melting.
5. Crystalline solids have well-defined cleavage 5. Amorphous solids cleavage into uneven
planes. parts with ragged edges.
6. They are anisotropic because of these 6. They are isotropic because of these
crystalline substances show different property amorphous substances show same property in
in different direction all directions.
7. Crystalline solids exhibit an X-ray diffraction 7. They do not give any X-ray diffraction
pattern. pattern.

Isomorphism: When two or more crystalline substances have the same number of atoms and the
same shape (crystalline form), the phenomenon is called isomorphism and such substances are said
to be isomorphous substances with each other. For example, ZnSO4.7H2O and FeSO4.7H2O are
isomorphous with each other.

Polymorphism: The phenomenon by which a substance exists in two or more different crystalline
forms is called polymorphism and the different crystalline forms are said to be polymorphous with
each other. For example, As2O3 (or Sb2O3) exists in two polymorphic forms: rhombic and
octahedral. These two forms are said to be polymorphous with each other.

5
As2O3 Sb2O3

Rhombic form Octahedral form Rhombic form Octahedral form

Polymorphous forms Polymorphous forms


Isomorphous forms
Isomorphous forms

Allotropy: Different crystalline forms of the same element are called its allotropes or allotropic
forms and the phenomenon is called allotropy. For example, different crystalline forms of sulphur,
rhombic sulphur or α-sulphur and monoclinic sulphur or β-sulphur are allotropes of each other.

α-sulphur β-sulphur

6
7
Chapter 2: Crystal Lattice and Crystal Symmetry

Crystal lattice: The regular array of points, which defines Lattice points (may be atoms,
the position of atoms, molecules or ions in a crystal, is molecules or ions)
called crystal lattice. Each point must have the same
number of neighbors as every other point and the
neighbors must always be found at the same distances and
directions. All points are in the same environment.
Crystal lattice

Lattice point: The point that describes the position of an atom, molecule or ion in a crystal is
called lattice point.

Motif: A motif is a unit of pattern. In a crystal, it is an atom, a molecule, an ion or a group of


atoms, molecules or ions.

Lattice + Motif Crystal

Unit cell: The smallest repeating unit of a crystal lattice is called unit cell. It contains an integer
number of lattice points. A unit cell has one lattice point at each corner, and there may be lattice
points in the faces, and in the interior of the cell as well.

Z
c

Unit cell 
 b
Y

a
X Unit cell

The unit cell is characterized by six parameters, three axial lengths and three interaxial angles. The
lengths of the unit cell edges are designated a, b, c, and the interaxial angles α, β, γ. The angle α is
between b and c, β is between c and a, while γ is between a and b.

The entire crystal consists of a large number of unit cells adjacent to one another in all three
dimensions. The unit cell of a crystal possesses all the structural properties of the given crystal. The
unit cell can be described as a parallelepiped defined by the cell edges a, b, and c which are

1
vectors. Its volume can be calculated by the scalar triple product, V = (a · b × c, where a, b, and c

are the lattice vectors) and corresponds to the square root of the determinant of the metric tensor.

Crystal system: The crystal system is a grouping of crystal structures that are categorized
according to the axial system used to describe their lattice. Each crystal system consists of a set of
three axes in a particular geometrical arrangement. The seven unique crystal systems are:

Crystal system No. of independent Examples


parameters
Triclinic 6 K2Cr2O7, H3BO3
a ≠ b ≠ c,
α ≠ β ≠ γ ≠ 90o

Monoclinic 4 NaHCO3, S
a ≠ b ≠ c,
α = γ = 90o, β > 90o

Orthorhombic 3 KNO3, BaSO4


a ≠ b ≠ c,
α = β = γ = 90o

Tetragonal 2 SnO2, TiO2


a = b ≠ c,
α = β = γ = 90o

Cubic 1 NaCl, KCl


a = b = c,
α = β = γ = 90o

Hexagonal 2 SiO2, ZnO


a = b ≠ c,
α = β = 90o, γ =120o

Rhombohedral 2 CaSO4, KMnO4


a = b = c,
α = β = γ ≠ 90o

2
Unit cell volume: The dimensions of the unit cell of each crystal system give its volume and are
calculated as follows:
Crystal system Volume
Triclinic abc√1 - cos2α – cos2β – cos2γ + 2cosα cosβ cosγ
Monoclinic abc sinβ
Orthorhombic abc
Tetragonal a2c
Cubic a3

3√3 a2c
Hexagonal ---------
2
Rhombohedral a3√1 – 3cos2α + 2cos3α

where a, b, and c are the unit cell axes dimensions and α, β, and γ are the inclination angles of the
axes in the unit cell.

Problem: A 1,10-phenanthrolinium(1+) tetraphenylborate compound crystallizes in a monoclinic


crystal system, with a = 10.09 Å, b = 15.03 Å, c = 7.03 Å and β = 97o. Calculate the volume of its
unit cell.

Solution:
The unit cell of 1,10-phenanthrolinium(1+) tetraphenylborate compound is monoclinic with a =
10.09 Å, b = 15.03 Å, c = 7.03 Å and β = 97o.
The volume of its unit cell is (10.09 Å)(15.03 Å)(7.03 Å) sin 97o = 1058.17 Å3

Classification of unit cell:

Unit cell

Primitive unit cell Nonprimitive unit cell


Lattice points present only at the Lattice points present positions
corner positions other than corners in addition to
those at corners

Side centered unit cell Face centered unit cell Body centered (Innenzentriert)
Having additional Having additional unit cell
lattice points on the lattice points at the Having an additional
centres of two centres of its six lattice point at the
opposite faces faces centre

3
Primitive (P, H or R) Side centered (A, B or C) Face centered (F) Body centered (I)

Non primitive
A primitive cell consists of 8 lattice points at the 8 corners of the unit cell. ⅛ lattice point at 8
Each lattice point is shared equally between 8 adjacent unit cells, and corners
therefore the number of lattice points present in a primitive unit cell = ⅛ × 8
= 1.

A side-centered unit cell consists of 8 lattice points at the 8 corners, each ⅛ lattice point at 8
shared by 8 unit cells and 2 lattice points at the center of the two opposite corners
faces, each shared equally by 2 adjacent unit cells. Thus the total number of
lattice points present in side-centered unit cell ⅛×8 + ½×2 = 2.

½ lattice point at 2
faces
A face-centered unit cell consists of 8 lattice points, each shared by 8 unit ⅛ lattice point at 8
cells at 8 corners and 6 face-centered lattice points, each shared by 2 cells. corners
Thus the total number of lattice points present in face-centered unit cell ⅛×8
+ ½×6 = 4.

½ lattice point at 6
faces
A body-centered unit cell consists of 8 lattice points at the 8 corners, each ⅛ lattice point at 8
corners
shared by 8 unit cells and one lattice point at the body center, which wholly
belongs to the unit cell. Therefore, the total number of lattice points present
in the body-centered unit cell = ⅛×8 + 1 = 2.

1 atom at center

4
Bravais lattices: The great French crystallographer Auguste Bravais (1850)
first demonstrated that there could be only 14 (7 primitive and 7 non-
primitive) different types of space lattices. This means that there are only 14
different types of three-dimensional arrays whereby atoms, molecules, or
ions may pack together when forming a crystal lattice. Because of Bravais
contributions to their study, these space lattices are called Bravais lattices.
Auguste Bravais
These lattices have seven different unit cell shapes with different symmetry properties
corresponding to the seven lattice systems.

The Bravais lattices


System No. of lattices Lattice No. of lattice Nature of unit cell axes and
in this system symbols points angles
Triclinic 1 P 1 a ≠ b ≠ c, α ≠ β ≠ γ ≠ 90o
Monoclinic 2 P 1 a ≠ b ≠ c, α = γ = 90o, β > 90o
C 2
Orthorhombic 4 P 1 a ≠ b ≠ c, α = β = γ = 90o
C 2
I 2
F 4
Tetragonal 2 P 1 a = b ≠ c, α = β = γ = 90o
I 2
Cubic 3 P 1 a = b = c, α = β = γ = 90o
I 2
F 4
Hexagonal 1 P 1 a = b ≠ c, α = β = 90o, γ =120o
Rhombohedral 1 R 1 a = b = c, α = β = γ ≠ 90o

The symbol ≠ implies non-equality by reason of symmetry; accidental equality may, of course,
occur.

Not all combinations of the crystal systems and lattice centerings are unique. There are in total
7 × 6 = 42 combinations, but it can be shown that several of these are in fact equivalent to each
other. For example, the monoclinic I lattice can be described by a monoclinic C lattice by different
choice of crystal axes. Similarly, all A- or B-centered lattices can be described either by a C- or
P-centering. This reduces the number of combinations to 14 conventional Bravais lattices.

When the fourteen Bravais lattices are combined with the 32 crystallographic point groups, we
obtain the 230 space groups.

5
Unit cells of the 14 Bravais lattices
Crystal system Lattices

Triclinic

Primitive (P)

Monoclinic

Primitive (P) C-centered (C)

Orthorhombic
Primitive (P) C-centered (C) Body-centered (I) Face-centered (F)

Tetragonal

Primitive (P) Body-centered (I)

Cubic

Primitive (P) Body-centered (I) Face-centered (F)

Hexagonal

Primitive (H)

Rhombohedral

Primitive (R)

Any space lattice corresponds to one or other of the fourteen shown in figure and no other distinct
space lattices can occur. For example, there are only two distinct monoclinic lattices, described by
P and C cells, and not amenable one to the other (shown below).

6
c c c
c'

c' = c
c' b' = b c'
a b' = b b'
a b' = b =a a' = b
a' = a' a
a'
(d)
(a) (b) (c)
Figure: Monoclinic lattices: (a) reduction of a B-centered cell to a Primitive cell; (b) reduction of an
Body-centered cell to an A-centered cell; (c) reduction of an F-centered cell to a C-centered cell;
(d) reduction of a C-centered cell to a Primitive non-monoclinic cell.

Miller indices: British mineralogist W. H. Miller in 1839 developed c


the very useful system of indexing of lattice planes. The Miller plane
is the plane which intercepts the three axes of a unit cell at the points
l
a/h, b/k, c/l, where a, b and c are the unit cell vectors. Thus the
a
Miller indices are proportional to the inverses of the intercepts of the h
k
plane with the unit cell. The indices h, k and l are integers, and
b
are conventionally enclosed in round bracket as (hkl). If one or more
of the indices is zero, it simply means that the planes do not intersect that axis (i.e. the intercept is
at infinity). Negative integers are written with a bar, as in for −3. The precise meaning of this
notation depends upon a choice of lattice vectors for the crystal.

h k l = 1/1 1/∞ 1/∞ h k l = 1/1 1/1 1/∞ h k l = 1/1 1/1 1/1 h k l = 1/-1 1/∞ 1/∞
= (100) = (110) = (111) = (͞1 00)

c (200) c (020)

b b
a a
h k l = 1/½ 1/∞ 1/∞ h k l = 1/∞ 1/½ 1/∞ h k l = 1/∞ 1/∞ 1/½ h k l = 1/½ 1/½ 1/1
= (200) = (020) = (002) = (221)
Planes with different Miller indices in crystals
7
The symbol { } is used to indicate sets of planes that are equivalent; for example, the sets (100),
(010) and (001) are equivalent in cubic crystals and may be represented collectively as {100}.

The procedure for determining the Miller indices for a cubic crystal plane is as follows:
(i) Choose a plane that does not pass through the origin at (0, 0, 0).
(ii) Determine the intercepts of the plane in terms of the crystallographic x, y, and z axes for a unit
cube. These intercepts may be fractions.
(iii) Form the reciprocals of these intercepts.
(iv) Clear fractions and determine the smallest set of whole numbers that are in the same ratio as
the intercepts.
These whole numbers are the Miller indices of the crystallographic plane and are enclosed in
parentheses.

Problem 1: Determine the Miller Indices of a plane which is parallel to x-axis and cuts intercepts
of 2 and ½, respectively along y and z axes.
Solution:
(i) Intercepts  2b ½c
(ii) Division by unit translation /a =  2b/b = 2 c/2c = ½
(iii) Reciprocals 1/ ½ 2
(iv) After clearing fraction (multiply by 2) 0 1 4
Therefore, the required Miller indices of the plane are (014).

Problem 2: Determine the Miller Indices of a plane that makes intercepts of 2Å, 3 Å, 4 Å on the
co-ordinate axes of an orthorhombic crystal with a:b:c = 4:3:2.

Solution: Here the unit translations are a = 4, b = 3 and c = 2 following the same procedure:
(i) Intercepts 2 3 4
(ii) Division by unit translation 2/4 = ½ 3/3 = 1 4/2 = 2
(iii) Reciprocals 2 1 ½
(iv) After clearing fraction (multiply by 2) 4 2 1

Therefore, the Miller indices of the plane are (421).

Note that when Miller indices are of double digit, these are separated by commas e.g. (5, 12, 15).

Miller indices are also sometimes useful to calculate the perpendicular distance dhkl between
parallel planes. Consider the (hk0) planes of a rectangular lattice built from an orthorhombic unit
cell of sides of lengths a and b. We can write the following trigonometric expressions for the angle
φ shown in the illustration.
8
sinφ = dhkl/(a/h) = hdhkl/a, cosφ = dhkl/(b/k) = kdhkl/b
Then because sin2φ + cos2 φ = 1, we obtain
h2d2hkl/a2 + k2d2hkl /b2 = 1
which we can rearrange into
b/k
1/d2hkl = h2/a2 + k2/b2  d
hkl 
In three dimensions, this expression becomes a/h
b
1/d2hkl = h2/a2 + k2/b2 + l2/c2
a
Thus,
Crystal system Perpendicular distance dhkl between parallel planes

Cubic 1/d2hkl = (h2 + k2 + l2) /c2


Tetragonal 1/d2hkl = h2/a2 + k2/a2 + l2/c2
Orthorhombic 1/d2hkl = h2/a2 + k2/b2 + l2/c2
Hexagonal 1/d2hkl = 4/3a2 (h2 + hk + k2) + l2/c2
Rhombohedral 1 (h2 + k2 + l2) sin2α + 2(hk + kl + lh) (cos2α – cosα)
1/d2hkl = ---- × --------------------------------------------------------------
a2 1 + 2cos3α – 3cos2α
Monoclinic h2 l2 2hl cosβ
----- + ---- - -------------
a2 c2 ac k2
1/d2hkl = ----------------------------------- + -----
sin2β b2
Triclinic h2 k2 l2 2hk
----- sin α + ----- sin β + ----- sin γ + ------ (cosα cosβ – cosγ)
2 2 2

a2 b2 c2 ab
2kl 2lh
+ ------ (cosβ cosγ – cosα) + ------ (cosγ cosα – cosβ)
bc ca
1/d2hkl = ----------------------------------------------------------------------
1 - cos2α - cos2β - cos2γ + 2 cosα cosβ cosγ

Problem: Find the ratio of interplanar distances of planes (100), (110) and (111) in a simple cubic
lattice.

Solution: Perpendicular distance dhkl between parallel planes in a simple cubic lattice is
a
dhkl = ----------------------
√(h2 + k2 + l2)

9
a a
d100 = ---------------------- = ------------------- = a
√(12 + 02 + 02) √(1 + 0 + 0)
a a
d110 = ---------------------- = ------------------- = a/√2
√(12 + 12 + 02) √(1 + 1 + 0)
a a
d111 = ---------------------- = ------------------ = a/√3
√(12 + 12 + 12) √(1 + 1 + 1)

d100 : d110 : d111 = a : a/√2 : a/√3

Problem: In a tetragonal lattice a = b = 2.5 Å, c = 1.8 Å. Calculate the lattice spacing between the
(111) planes.

Solution: Perpendicular distance dhkl between parallel planes in a tetragonal lattice is


1
dhkl = -------------------------------
√(h2/a2 + k2/b2 + l2/c2)
1
d111 = --------------------------------------- Å = 1.26 Å
√(12/2.52 + 12/2.52 + 12/1.82)

Fractional coordinates: Coordinates of atoms expressed as Z

fractions of the unit cell lengths are called fractional c=4A

coordinates. For example, fractional coordinates of an atom


are given by (x/a, y/b, z/c) where x, y, z are coordinates in Å
1A b=4A
and a, b, c are the unit-cell repeat distance in the same 2A Y
O
direction and in the same unit. Figure demonstrates how the
a=4
fractional coordinates of O atom is (0.50, 0.25, 0). X

Fractional coordinate of O is
(0.50, 0.25, 0).

Crystal symmetry and their operation: A symmetry element is a geometrical entity such as a line
(or axis), a plane or a point with respect to which one or more symmetry operations may be carried
out. Among the five crystallographic symmetry elements, three are called point-symmetry elements
and two are space-symmetry elements. The rotation, mirror reflection and rotatory inversion are
point-symmetry elements and their operations through a point are point-symmetry operations, since
each leaves at least one point of the object in a fixed position.

n-Fold rotation axes: An n-fold rotation axis of symmetry is defined as a line, rotation about which
produces congruent positions (i.e. positions indistinguishable from the initial position) after rotation

10
through 2π/n. Rotation axes are described with reference to the value of n, which must of course be
integral: one-fold, two-fold ( ), three-fold (▲), four-fold (■), and six-fold ( ).

a a

c c
Two-fold rotation, 2 Three-fold rotation, 3 Four-fold rotation, 4

Note that five-fold rotation axis is not usually found in regular crystal lattice. This can be explained
by the following illustration:
Operation of 5-fold rotation axis through A
B4 A1
produces lattice points B, B1, B2, B3 and B4. B3 A2
Operation of 5-fold rotation axis through B
A B
produces lattice points A, A1, A2, A3 and A4.

If the resultant array of points is to form a lattice, it B2 A3

must be a regular array, in particular the spacing B1 A4

of points on lines parallel to AB such as B4A1, must be equal to AB or some multiple thereof. It is
evident from figure that
B4A1 = AB – AB4 cos 72o – BA1 cos 72o = AB (1- 2cos 72o) = 0.38 AB
Therefore, the array of points generated by 5-fold rotation axis through adjacent lattice points is not
regular, and consequently not itself a regular lattice.

Mirror plane: If a plane exists in a structure such that every part on one side of the plane is related
to a part on the other side as if reflected, the structure is said to possess a plane of symmetry or
mirror plane. The symbol m is used for this symmetry element.

a a

c c
Mirror plane Mirror plane (⊥r to the plane) Mirror plane (∥ to the plane)

11
n-Fold rotatory-inversion axes: A n-fold rotatory inversion axis implies that a rotation of 2π/n
(where n is 1, 2, 3, 4, or 6) followed by inversion through some point on the axis produces no
apparent change in the object or structure. These axes are symbolized as ͞n.

a a

c c
One fold inversion, ͞1 Two fold inversion, ͞2 Three-fold inversion, ͞3
Combination of the point-symmetry operations with translations gives rise to two space-symmetry
operations.

n-Fold screw axes: The combination of a rotation axis and a translation parallel to the axis
produces a screw axis. The direction of such an axis is usually along a unit cell edge and the
translation must be a subintegral fraction of the unit translation in that direction. Screw axes are
designated by an integer n and a subscript m, where n = 1, 2, 3, 4, or 6 is the multiplicity of the axis
and m is an integer less than n. Thus 21 designates a 2-fold screw axis with a translation between
successive points of ½(= m/n) of a unit translation.
(ii)
½+
a
(i)

c
A screw axis consists of a rotation followed by a translation 21 parallel to b

Coordinates of the two equivalent positions related to each other by a screw axis (21) parallel to b
are (i) x, y, z; (ii) –x, y + ½, -z.

Glide planes: The combination of a mirror plane and a translation parallel to the reflecting plane
produces a glide plane. The translation in such a plane is along an edge or face diagonal of the unit
cell and in most cases, of magnitude half the axial or diagonal length. A glide plane is designated
by a, b or c, if the translation is a/2, b/2, or c/2 and by n if (a+b)/2, (a+c)/2 or (b+c)/2 i.e. half way
along one of the face diagonal. There is only one additional type of glide plane - the diamond glide
d, characterized by a translation (a+b)/4, (a+c)/4 or (b+c)/4.
12
(ii)

a
(i)

c
A glide plane consists of a reflection followed by a Glide plane parallel to a, i.e. a glide
translation
Coordinates of the two equivalent positions related to each other by a glide, (i) x, y, z;
(ii) x + ½, y, -z.

The following tables list the graphical and typed symbols used to describe symmetry operations in
the International Tables for Crystallography.

Symbols of symmetry planes:


Graphical symbol
Nature of glide
Symmetry plane Normal to plane Parallel to plane Symbol
translation
of projection of projection

Reflection plane m None

a/2 along [100] or b/2


a, b
along [010]
Axial glide plane c/2 along z-axis or
None c
(a+b+c)/2 along [111]

Double glide plane e Along <100>

Diagonal glide (a+b)/2 or (b+c)/2 or


n
plane (net) (c+a)/2 or (a+b+c)/2

Diamond glide (a±b)/4 or (b±c)/4 or


 d
plane  (c±a)/4 or (a±b±c)/4

13
Symbols of symmetry axes:
Symmetry Graphical Translational Symbol Symmetry Graphical Translational Symbol
axis symbol axis symbol

Rotation None None 1


6-fold None 6
monad
Inversion
None ͞1 61 c/6 61
monad

(normal to None 2 62 2c/6 62


Rotation paper)
diad
(parallel to None 2 63 3c/6 63
paper)

(normal to c/2 21 64 4c/6 64


Screw diad paper)

(parallel to Either a/2 or


21 65 5c/6 65
paper) b/2

Rotation Inversion
None 3 None ͞6
triad hexad
2-fold and
c/3 31 None 2/m
Screw inversion
triads
21 and
2c/3 32 None 21/m
inversion
Inversion 4-fold and
None ͞3 None 4/m
triad inversion
Rotation 42 and
None 4 None 42/m
tetrad inversion

6-fold and
c/4 41 None 6/m
inversion
Screw 63 and
tetrads 2c/4 42 None 63/m
inversion
3c/4 43
Inversion
None ͞4
tetrad

Point groups: A point group may be defined as a set of symmetry operations the action of which
leaves at least one point unmoved: this point is taken as the origin of the reference axes for the
body, through which all symmetry elements pass. Point groups are derived by combining three
point symmetry elements - rotation, reflection and inversion. Of the 42 combinations obtained, only
32 are unique point groups. These 32 point groups are divided among the seven crystal systems.

14
System Essential Point groups Laue groups
symmetry Non-centrosymmetric Centrosymmetric
Triclinic None 1 ͞1 ͞1
Monoclinic 2 or m 2, m 2/m 2/m
Orthorhombic 222 or mm2 222, mm2 mmm mmm
Tetragonal 4 or ͞4 4, ͞4, 422, 4mm, ͞42m, 4/m, 4/mmm 4/m, 4/mmm
Trigonal 3 or ͞3 3, 32, 3m ͞3, ͞3m ͞3, ͞3m
Hexagonal 6 or ͞6 6, ͞6, 622, 6mm, ͞6m2 6/m, 6/mmm 6/m, 6/mmm
Cubic 23 23, 432, ͞43m m ͞3, m ͞3m m ͞3, m ͞3m

15
The notation is as follows:
X
Rotation axis X; Inversion axis ͞X; Rotation axis with mirror plane normal to it X/m(-----); Rotation
m
axis with diad axis (axes) normal to it X2; Rotation axis with mirror plane (planes) parallel to it
Xm; Inversion axis with diad axis (axes) normal to it ͞X2; Inversion axis with mirror plane (planes)
parallel to it ͞Xm; Rotation axis with a mirror plane normal to it and mirror planes parallel to it
X
X/mm(------ m).
m

16
Subgroups: A subgroup of a given point group is a point group of lower symmetry than the given
group, contained within it and capable of separate existence as a point group. For example, 32 is a
subgroup of ͞3m; 622 and ͞6m2 of 6/mmm, etc.

Laue groups: Of the thirty-two crystallographic point groups, eleven are centrosymmetric. These
are the Laue groups. The 32 point groups fall into these 11 diffraction symmetry groups. The Laue
groups are, then, groups of point groups that become identical when a center of symmetry is added
to those that lack it.

Note that the actual diffraction pattern (with the intensities of the reflections taken into account)
must be at least centrosymmetric from Friedel’s law (I(hkl) = I(-h-k-l)). When this centrosymmetric
requirement is combined with the actual symmetry of the crystal lattice one obtains the Laue class
or Laue symmetry of the reciprocal lattice. This symmetry is used by the data collection software,
in conjunction with systematic absences, to determine the space group of the crystal.

The Laue groups

Centrosymmetric crystal classes or Laue groups

To derive the Laue symmetry from the space group symmetry:


• Remove the lattice centering symbol (C, F, I, R)
• Remove the translational symmetry component of screw axes
• Add a centre of symmetry

Max von Laue


17
In this way the 230 space groups are reduced to the 11 Laue groups. It is the Laue group symmetry
that is determined from the symmetry of the diffraction pattern.

Comparison of molecular symmetry with crystallographic symmetry: Table lists the symmetry
properties of molecules and crystals.
Molecular symmetry Crystallographic symmetry
1. Cn axis of order n (n = 1, 2, 3, …., ∞) is 1. Cn axis of order n (n = 1, 2, 3, 4, 6) is
present. present.
2. The total number of point groups is not 2. The number of point groups is limited to 32.
limited.
3. Rotation-reflection axis is present. 3. Rotation-inversion axis is present.
4. Translational elements of symmetry are 4. Translational elements of symmetry such as
absent. screw axis and glide plane are present.
5. Schoenflies notation is used. 5. Hermann-Mauguin notation is more
commonly used.

Space groups: Combination of 32 point groups with the 14 Bravais lattices leads to 230 unique
arrangements of points in space. These are the 230 space groups that describe the only ways in
which identical objects can be arranged in an infinite lattice. The notation for space group
enumeration follows the Hermann structure, e.g. Lijk where L = lattice capital letter for three
dimensional lattice, and ijk = symmetry elements of space group for the different symmetry
directions, and i = primary, j = secondary and k = tertiary directions.

Crystal system No. of Bravais lattices No. of point groups No. of space groups
Triclinic 1 2 2
Monoclinic 2 3 13
Orthorhombic 4 3 59
Tetragonal 2 7 68
Rhombohedral (Trigonal) 1 5 25
Hexagonal 1 7 27
Cubic 3 5 36
Total 14 32 230

Triclinic: P1, P ͞1

Monoclinic: P2, P21, C2, Pm, Pc, Cm, Cc, P2/m, P21/m, C2/m, P2/c, P21/c, and C2/c.

18
Orthorhombic: P222, P2221, P21212, P212121, C2221, C222, F222, I222, I212121, Pmm2, Pmc21,
Pcc2, Pma2, Pca21, Pnc2, Pmn21, Pba2, Pna21, Pnn2, Cmm2, Cmc21, Ccc2, Amm2, Abm2, Ama2,
Aba2, Fmm2, Fdd2, Imm2, Iba2, Ima2, Pmmm, Pnnn, Pccm, Pban, Pmma, Pnna, Pmna, Pcca,
Pbam, Pccn, Pbcm, Pnnm, Pmmn, Pbcn, Pbca, Pnma, Cmcm, Cmca, Cmmm, Cccm, Cmma, Ccca,
Fmmm, Fddd, Immm, Ibam, Ibca, Imma.
And so on.

Triclinic P 1 (No. 1)

General-position diagram Symmetry-elements diagram

a a
+ +
(i)

+ +
c c
No. of equivalent positions, Z = 1
Coordinates of equivalent positions: (i) x, y, z

Triclinic P ͞1 (No. 2)

General-position diagram Symmetry-elements diagram

(ii) , - , -
a a
+ +
(i)

, - , -
+ +
c c
= Inversion monad

Origin at ͞1
No. of equivalent positions, Z = 2
Coordinates of equivalent positions: (i) x, y, z; (ii) - x, - y, - z.

19
Monoclinic P 2 (No. 3)

General-position diagram Symmetry-elements diagram

(ii) - -
b b
+ +
(i)

- -
+ +
a a
= Rotation diad, parallel to paper
Origin on 2; unique axis b
No. of equivalent positions, Z = 2
Coordinates of equivalent positions: (i) x, y, z; (ii) - x, y, - z

Monoclinic P 21 (No. 4)

General-position diagram Symmetry-elements diagram

(ii) -
b b
+ +
(i)

-
+ +
a a

= Screw diad, parallel to paper


Origin on 21; unique axis b
No. of equivalent positions, Z = 2
Coordinates of equivalent positions: (i) x, y, z; (ii) - x, y + ½, - z.

20
Monoclinic C 2 (No. 5)

General-position diagram Symmetry-elements diagram

(ii) - -
b b
+ +
(i)
(iv)
-
+
(iii)
- -
+ +
a a

= Rotation diad, parallel to paper


= Screw diad, parallel to paper
Origin on 2; unique axis b
No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) - x, y, - z; (iii) x + ½, y + ½, z;
(iv) x - ½, y + ½, - z

Monoclinic Pm (No. 6)

General-position diagram Symmetry-elements diagram

b b
+ , + + , +
(ii) (i)

+ , + + , +
a a

= Reflection plane, normal to plane of projection


Origin on plane m; unique axis b
No. of equivalent positions, Z = 2
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, z

21
Monoclinic P c (No. 7)

General-position diagram Symmetry-elements diagram

b b
½+ , + ½+ , +
(ii) (i)

½+ , + ½+ , +
a a

= Axial glide plane, normal to plane of projection

Origin on glide plane c; unique axis b


No. of equivalent positions, Z = 2
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, ½ + z

Monoclinic C m (No. 8)

General-position diagram Symmetry-elements diagram

b b
+ , + + , +
(ii) (i)

+ , +
(iv) (iii)

+ , + + , +
a a

= Reflection plane, normal to plane of projection


= Axial glide plane, normal to plane of projection
Origin on plane m; unique axis b
No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, z; (iii) ½ + x, ½ + y, z; (iv) ½ + x, ½ - y, z

22
Monoclinic C c (No. 9)

General-position diagram Symmetry-elements diagram

b b
½+ , + ½+ , +
(ii) (i)

½+ , +
(iv) (iii)

½+ , + ½+ , +
a a

= Axial glide plane, normal to plane of projection


= Diagonal glide plane (net), normal to plane of
projection

Origin on glide plane c; unique axis b


No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, ½ + z; (iii) ½ + x, ½ + y, z;
(iv) ½ + x, ½ - y, ½ + z

Monoclinic P 2/m (No. 10)

General-position diagram Symmetry-elements diagram

(iv) - , - (iii) - , -
b b
+ , + + , +
(ii) (i)

- , - - , -
+ , + + , +
a a

= Inversion monad
= Reflection plane, normal to the
plane of projection
= Rotation diad, parallel to paper
Origin at center (2/m); unique axis b
No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, z; (iii) - x, y, - z; (iv) - x, - y, - z

23
Monoclinic P 21/m (No. 11)

General-position diagram Symmetry-elements diagram

- , (iv) (ii) - - ,
b b
+ + , +
(i) (iii)

- , - - ,
+ + , +
a
a

= Inversion monad
= Reflection plane, normal to plane of projection
= Screw diad, parallel to paper
Origin at ͞1; unique axis b
No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) - x, ½ + y, - z; (iii) x, ½ - y, z; (iv) - x, - y, - z

Monoclinic C 2/m (No. 12)

General-position diagram Symmetry-elements diagram

(iv) - , - (iii) - , -
b b
+ , + + , +
(ii) (i) (viii) (vii)
-, -
+ , +
(vi) (v)
- , - - , -
+ , + + , +
a
a

= Inversion monad
= Rotation diad, parallel to paper
= Screw diad, parallel to paper
= Axial glide plane, normal to plane of projection
Origin at center (2/m); unique axis b
No. of equivalent positions, Z = 8
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, z; (iii) - x, y, - z; (iv) - x, - y, - z;
(v) ½ + x, ½ + y, z; (vi) ½ + x, ½ - y, z; (vii) ½ - x, ½ + y, - z; (viii) ½ - x, ½ - y, - z

24
Monoclinic P 2/c (No. 13)

General-position diagram Symmetry-elements diagram

(iv) - , ½ - (iii) - , ½-
b ¼ b ¼
½+ , + ½+ , +
(ii) (i)

¼ ¼

- , ½- - , ½-
¼ ¼
½+ , + ½+ , +
a
a

= Inversion monad
= Axial glide plane, normal to plane of projection
¼ = Rotation diad at z = ¼ , parallel to paper

Origin at ͞1; unique axis b


No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, ½ + z; (iii) - x, y, ½ - z;
(iv) - x, - y, - z

Monoclinic P 21/c (No. 14)

General-position diagram Symmetry-elements diagram

(iv) - , (ii) ½- - ,
b ¼ b ¼
+ ½+ , +
(i) (iii)

¼ ¼

- , ½- - ,
¼ ¼
+ ½+ , +
a
a

= Inversion monad
= Axial glide plane, normal to plane of projection
¼ = Screw diad at z = ¼ , parallel to paper
Origin at ͞1; unique axis b
No. of equivalent positions, Z = 4
Coordinates of equivalent positions: (i) x, y, z; (ii) - x, ½ + y, ½ - z; (iii) x, ½ - y, ½ + z;
(iv) - x, - y, - z

25
Monoclinic C 2/c (No. 15)

General-position diagram Symmetry-elements diagram

(iv) - , ½ - (iii) - , ½-
b ¼ b ¼
½+ , + ½+ , +
(ii) (i) (viii) (vii) ¼ ¼
- , ½-
¼ ¼
½+ , +
(vi) (v) ¼ ¼
- , ½- - , ½-
¼ ¼
½+ , + ½+ , +
a a

= Inversion monad
= Axial glide plane, normal to plane of projection
= Diagonal glide plane (net), normal to plane of
projection
¼ = Rotation diad at z = ¼ , parallel to paper
¼ = Screw diad at z = ¼, parallel to paper

Origin at ͞1 on glide plane c; unique axis b


No. of equivalent positions, Z = 8
Coordinates of equivalent positions: (i) x, y, z; (ii) x, - y, ½ + z; (iii) - x, y, ½ - z;
(iv) - x, - y, - z; (v) ½ + x, ½ + y, z; (vi) ½ + x, ½ - y, ½ + z; (vii) ½ - x, ½ + y, ½ - z;
(viii) ½ - x, ½ - y, - z

Systematic absences: It is often observed that in a data set, certain classes of reflection are
systematically absent because atoms in one motif are scattering exactly out of phase with atoms in
another. These absences are referred to as systematic absences. Systematic absences occur only
when there is some symmetry element with a translational component within the unit cell – a screw
axis, a glide plane, or a non-primitive unit cell. Thus these absences help us to identify the crystal
space lattice to be P, I, F, A, B, C in respect to the selected crystallographic axes as well as the
presence or absence of screw axes or glide planes in the crystal structure. Some important
symmetry elements and their systematic absences are shown in Table.

26
Table. Translational symmetry elements and their extinctions.
Symmetry element Affected reflection Condition for systematic absence
A-centered lattice (A) hkl k + l = 2n + 1 = odd
B-centered lattice (B) hkl h + l = 2n + 1 = odd
C-centered lattice (C) hkl h + k = 2n + 1 = odd
F-centered lattice (F) hkl h + k = 2n + 1 i.e h, k, l not all
h + l = 2n + 1 even or all odd
k + l = 2n + 1
Body-centered lattice (I) hkl h + k + l = 2n + 1 = odd
2-fold screw (21) along a h00 h = 2n+1 = odd
b 0k0 k = 2n+1 = odd
c 00l l = 2n+1 = odd
Glide planes perpendicular to a
Translation b/2 (b glide) 0kl k = 2n + 1 = odd
c/2 (c glide) 0kl l = 2n + 1 = odd
b/2 + c/2 (n glide) 0kl k + l = 2n + 1 = odd
b/4 + c/4 (d glide) 0kl k + l = 4n + 1
2 = odd
3
Glide planes perpendicular to b
Translation a/2 (b glide) h0l h = 2n + 1 = odd
c/2 (c glide) h0l l = 2n + 1 = odd
a/2 + c/2 (n glide) h0l h + l = 2n + 1 = odd
a/4 + c/4 (d glide) h0l h + l = 4n + 1
2 = odd
3
Glide planes perpendicular to c
Translation a/2 (b glide) hk0 h = 2n + 1 = odd
b/2 (c glide) hk0 k = 2n + 1 = odd
a/2 + b/2 (n glide) hk0 h + k = 2n + 1 = odd
a/4 + b/4 (d glide) hk0 h + k = 4n + 1
2 = odd
3

The following example illustrates how systematic absences arise from particular symmetry element
involving translation: e.g. c glide operation reflecting in the plane normal to the b axis.

Consider a data set that has a c glide operation reflecting in the


(ii) , +
plane normal to the b axis. The symmetry operations could be (x, y, c
+
z) and (x, -y, z + ½). If there are N atoms in the unit cell, then there (i)

are N/2 unique atoms. The summations below are over the j atoms
and run from 1 to N/2. Then,
b

27
F(hkl) = ∑ fj exp 2πi(hxj + kyj + lzj) + ∑ fj exp 2πi[hxj - kyj + l(½ + zj)]

Consider the data with k = 0. For these data the structure factors become:

F(h0l) = ∑ fj exp 2πi(hxj + lzj) + ∑ fj exp 2πi(hxj + lzj) exp2πi(l/2)


F(h0l) = ∑ fj exp 2πi(hxj + lzj) [1 + exp πil]
If l is an odd integer, then exp πil = -1 and F(h0l) = 0. If l is an even integer, then F(h0l) is not 0.
The systematic absence conditions for other symmetry operations can be derived in a similar
manner as was done above.

Some important points about systematic absences:


(i) Systematically absent means that all affected reflections must be absent.
For example, a 21 screw axis along a would lead to the absence of all h00 with h odd. A single
weak reflection with h odd rules out a screw axis !
(ii) Enantiomorphous screw axes produce the same systematic absences. Hence, pairs of
enantiomorphous space groups (e.g., P4122 and P4322) cannot be distinguished based on the
diffraction pattern alone.

(iii) Systematic absences due to unit cell centering may obscure absences due to screw axes. For
example, body (I) centering results in the absence of all reflections with (h + k + l) odd, including
h00 with h odd, 0k0 with k odd, and 00l with l odd. Hence, it is not possible to distinguish between
I222 and I212121.

Space group determination


Space group determination entails the following steps:
(i) Determination of the crystal system, lattice type and probable Laue group based on the geometry
(shape) of the unit cell.
(ii) Determination of the true Laue group based on intensities.
(iii) Assignment of screw axes based on systematic absences.
If the space group cannot be uniquely determined (enantiomorphous pairs, I222 vs. I212121), do not
despair.

28
Chapter 3: Elementary Crystal Optics

Crystal forms: A crystal form is a set of faces which are geometrically equivalent and whose
spatial positions are related to one another according to the symmetry of the crystal. If one face of a
crystal form is defined, the point symmetry operations which specify the class to which the crystal
belongs also determine the other faces of the crystal form. To designate a crystal form (which could
imply many faces) we use the Miller index, or Miller-Bravais index notation enclosing the indices
in curly braces, i.e. {101} or {11 1}. Such notation is called a form symbol.
A simple crystal may consist of only a single crystal form. A
more complicated crystal may be a combination of several
different forms. All forms which occur in a crystal of a particular
system must be compatible with that crystal system. (a) (b)
(a) Simple and (b) complicated
The crystal forms of the five non-isometric crystal systems are the
crystals.
monohedron or pedion, parallelohedron or pinacoid, dihedron or
dome and sphenoid, disphenoid, prism, pyramid, dipyramid, trapezohedron, scalenohedron,
rhombohedron and tetrahedron. Fifteen different forms are possible within the isometric system.
These include the hexoctahedron, gyroid, hextetrahedron, diploid, and tetartoid, among others.

Thus there are a total of 47 or 48 possible forms that can be developed as the result of the 32
combinations of symmetry.
Isometric crystal forms
Name Shape No of faces Name Shape No of faces
Cube Tristetrahedron
6 12

Octahedron Hextetrahedron
8 24

Dodecahedron Deltoid
12 24
dodecahedron
Tetrahexahedron Gyroid
24 24

Trapezohedron Pyritohedron
24 12

Trisoctahedron Diploid
24 24

Hexoctahedron Tetartoid
48 12

Tetrahedron
4

1
Non-isometric crystal forms
Name Shape No of faces Name Shape No of faces
Pedion Dihexagonal
1 12
pyramid
Pinacoid
2 Rhombic dipyramid 8

Dome or Sphenoid 2 Trigonal dipyramid 6

4 Ditrigonal 12
Rhombic prism
dipyramid
3 Tetragonal 8
Trigonal prism
dipyramid
6 Ditetragonal 16
Ditrigonal prism
dipyramid
4 Hexagonal 12
Tetragonal prism
dipyramid
8 Dihexagonal 24
Ditetragonal prism
dipyramid
6 Trigonal 6
Hexagonal prism
trapezohedron
12 Tetragonal 8
Dihexagonal prism
trapezohedron
4 Hexagonal 12
Rhombic pyramid
trapezohedron
3 Tetragonal 8
Trigonal pyramid
scalenohedron
6 Hexagonal 12
Ditrigonal pyramid
scalenohedron

Tetragonal pyramid 4 Rhombohedron 6

Ditetragonal 8 Rhombic 4
pyramid disphenoid
6 Tetragonal 4
Hexagonal pyramid
disphenoid

There are two classifications with respect to crystal forms. They are: general forms and special
forms.

General forms: A general form is a form in a particular crystal class that contains faces that
intersect all crystallographic axes at different lengths. It has the form symbol {hkl}. In the
monoclinic, triclinic, and orthorhombic crystal systems, the form {111} is a general form because
in these systems faces of this form will intersect the a, b, and c axes at different lengths as the unit

2
lengths are different on each axis. In crystals of higher symmetry, where two or more of the axes
have equal length, a general form must intersect the equal length axes at different multiples of the
unit length. Thus in the tetragonal system the form {121} is a general form. In the isometric system
a general form would have to be something like {123}.

Special forms: All other forms that may be present are called special forms.

Open forms and closed forms


Open forms: Open forms are those groups of faces all related by symmetry
that do not completely enclose a volume of space. A crystal with open form
faces requires additional faces as well. Some examples of closed forms:
prisms, pyramids, pinacoids, etc.

Closed forms: Closed forms are those groups of faces all related by symmetry
that completely enclose a volume of space. It is possible for a crystal to have
faces entirely of one closed form. Some examples of closed forms:
dipyramids, octahedron, dodecahedron, etc.

There are 18 open forms and 30 closed forms.

Pinacoids: They are open forms; consist of two and only two
geometrically equivalent faces which occupy opposite sides of a crystal.
The two faces are parallel and are related to one another only by a
reflection or an inversion.

Prisms: A prism is composed of a set of 3, 4, 6, 8, or 12 geometrically equivalent faces which are


all parallel to the same axis. Prisms are given names based on the shape of their cross section.
Variants of the prism form include the trigonal prism, rhombic prism, tetragonal prism, and
hexagonal prism. All are open forms.

(i) Trigonal prism: 3 - faced form with all faces parallel to a 3 -fold rotation axis.

(ii) Rhombic prism: 4 - faced form with all faces parallel to a line that is not a symmetry element.
In the drawing, the 4 faces belong to a rhombic prism. The other faces in the model are pinacoids
(the faces on the sides belong to a side pinacoid, and the faces on the top and bottom belong to a
top/bottom pinacoid).

(iii) Tetragonal prism: 4 - faced open form with all faces parallel to a 4-fold rotation axis or . The
4 side faces in the model make up the tetragonal prism. The top and bottom faces make up the a
form called the top/bottom pinacoid.

3
(iv) Hexagonal prism: 6 - faced form with all faces parallel to a 6-fold rotation axis. The 6 vertical
faces in the drawing make up the hexagonal prism. Again the faces on top and bottom are the
top/bottom pinacoid form.

Trigonal prism Rhombic prism Tetragonal prism Hexagonal prism

Pyramids: A pyramid is a 3, 4, 6, 8 or 12 faced open form where all faces in the form meet, or
could meet if extended, at a point. The trigonal, orthorhombic, tetragonal and hexagonal crystal
systems all produce pyramids. These pyramids are named according to the shape of their cross-
section in the same way that prisms are. Thus are produced the trigonal pyramid, rhombic pyramid,
tetragonal pyramid, and hexagonal pyramid. All are open forms.

(i) Trigonal pyramid: 3-faced form where all faces are related by a 3-fold rotation axis.

(ii) Rhombic pyramid: 4-faced form where the faces are related by mirror planes.

(iii) Tetragonal pyramid: 4-faced form where the faces are related by a 4-fold axis.

(iv) Hexagonal pyramid: 6-faced form where all faces are related by a 6-fold axis.

Trigonal pyramid Rhombic pyramid Tetragonal pyramid Hexagonal pyramid

Dipyramids: Two pyramids joined base to base along a mirror plane. The upper and lower
pyramids may each have 3, 4, 6, 8, or 12 faces; the dipyramidal form therefore possesses a total of
6, 8, 12, 16, or 24 faces. The trigonal, orthorhombic, tetragonal and hexagonal crystal systems all
produce dipyramids. These dipyramids are named for the shape of their cross-section just as prisms
and pyramids are, resulting in the trigonal dipyramid, rhombic dipyramid, tetragonal dipyramid,
and hexagonal dipyramid. All are closed forms.

4
Trigonal dipyramid Rhombic dipyramid Tetragonal dipyramid Hexagonal dipyramid

Crystal zones and zone symbols: A zone is defined as a group of crystal faces that intersect in
parallel edges. The edges will all be parallel to a line and the line is called the zone axis. The
direction of the line is usually represented by a notation
similar to Miller index, but is enclosed in square bracket,
i.e. [uvw]. This notation is called the zone symbol. Note
that the indices are integer numbers arranged in the order of
The zone axis is the common crystal
the a-, b-, and c-axes [uvw] (a specified zone), <uvw> direction shared by two planes.
(general zone), similar to the case of the face index. The
index of an edge formed by two crystal faces, (h1k1l1) and (h2k2l2) can be obtained by the crossing
multiplication method, as follows:
To do so, we write the Miller index for each
h1 k1 l1 h1 k1 l1
faces twice, one face directly beneath the
other. The first and last numbers in each line
h2 k2 l2 h2 k2 l2
are discarded. Then we apply the crossing
multiplication method to determine the indices k1l2 - k2l1 l1h2 - l2h1 h1k2 - h2k1
in the zone symbol. =u =v =w

For example, the zone symbol for the (110) and (010) faces is
determined by writing 110 twice and then immediately below,
writing 010 twice. Applying the method above gives the zone
symbol for this zone as [001].

A zone symbol implies also a line that is perpendicular to


the face with the same index. In other words, [001] is a line
perpendicular to the face (001).

5
Vectorial properties of crystals: Some properties of crystals depend on direction. These are called
vectorial properties, and can be divided into two categories: continuous and discontinuous.

Continuous vectorial properties: Continuous vectorial properties depend on direction, but along
any given the direction the property is the same. Some of the continuous vectorial properties are:

(i) Hardness: In some minerals there is a difference in hardness in different directions in the
crystal. Examples: Kyanite, Biotite, Muscovite. This can become an important identifying property
and/or may lead to confusion about the hardness if one is not aware of the directional dependence.

(ii) Velocity of light (Refractive index): For all minerals except those in the isometric system, the
velocity of light is different as the light travels along different directions in the crystal. Because the
velocity of light depends on direction, the refractive index (Refractive index is defined as the
velocity of light in a vacuum divided by the velocity of light in the material.) will also depend on
direction.

(iii) Thermal conductivity: The ability of a material to conduct heat is called thermal conductivity.
Like light, heat can be conducted at different rates along different directions in crystals.

(iv) Electrical conductivity: The ability of a material to allow the passage of electrons is called
electrical conductivity, which is also directionally dependent except in isometric crystals.

(v) Thermal expansion: How much the crystal lattice expands as it is heated is referred to as
thermal expansion. Some crystals expand more in one direction than in others, thus thermal
expansion is a vectorial property.

(vi) Compressibility: Compressibility is a measure of how the lattice is reduced as atoms are pushed
closer together under pressure. Some directions in crystals may be more compressible than others.

Discontinuous vectorial properties: Discontinuous vectorial properties pertain only to certain


directions or planes within a crystal. For these kinds of properties, intermediate directions may have
no value of the property. Among the discontinuous vectorial properties are:

(i) Cleavage: Cleavage is defined as a plane within the lattice along which breakage occurs more
easily than along other directions. A cleavage direction develops along zones of weakness in the
crystal lattice. Cleavage is discontinuous because it only occurs along certain planes.

(ii) Growth rate: Growth rate is defined as the rate at which atoms can be added to the crystal. In
some directions fewer atoms must be added to the crystal than in other directions, and thus some
directions may allow for faster growth than others.

6
(iii) Solution rate: Solution rate is the rate at which a solid can be dissolved in a solvent. In this
case it depends on how tightly bonded the atoms are in the crystal structure, and this usually
depends on direction.

Cleavage, parting and fracture: The properties of cleavage, parting and fracture are all related to
the tenacity or tensile strength of a mineral, and also related to the toughness.

Cleavage: Cleavage is the splitting of a crystalline mineral along the direction of its crystal faces
where atoms have weaker bonding. This can occur only in minerals when a precise blow is given in
a particular direction. The result of cleavage is a more or less flat plane with often a silky luster.
Cleavage is a reproducible property of a mineral and can be done at any point of the cleavage
direction. The cleavage of a mineral is described according to the crystal face to which it is parallel,
as cubic cleavage (galena, halite), octahedral cleavage (fluorite), dodecahedral cleavage (shplerite),
rhombohedral cleavage (calcite), prismatic cleavage (amphibole), basal cleavage (topaz), etc.

Basal cleavage Prismatic cleavage Cubic cleavage

Parting: Parting is a break along structural planes and is parallel to a possible face, just like
cleavage. However, parting differs from cleavage in some important ways. It cannot be found in
every specimen as is true of cleavage for most every cleavable mineral. It is not absolutely
repeatable or reproducible as is cleavage down to theoretically the very atomic layers that cause
cleavage.
Parting is caused by pressures that are applied to a crystal or by twinning. The pressure breaks the
crystal on a plane of weakness.

Fracture: Fracture is an irregular break which occurs when neither cleavage nor parting is present
on a surface. Fracture may be smooth, mimicking cleavage, but more often is uneven. Various
types of fracture are recognized and some of these are diagnostic.
(i) Conchoidal: When the fracture has smooth, curved surfaces like the interior surface of a shell it
is said to be conchoidal. This is most commonly observed in such substance as glass, quartz, etc.
(ii) Fibrous or Splintery: When the mineral breaks showing splinters or fibers.
(iii) Hackly: When the mineral breaks with jagged, irregular surfaces with sharp edges.

7
(iv) Uneven or Irregular: When the mineral breaks into rough and irregular surfaces.

Conchoidal fracture Fibrous fracture Hackly fracture Irregular fracture

Crystal habit: In nature perfect crystals are rare. The faces that develop on a crystal depend on the
space available for the crystals to grow. If crystals grow into one another or in a restricted
environment, it is possible that no well-formed crystal faces will be developed. However, crystals
sometimes develop certain forms more commonly than others, although the symmetry may not be
readily apparent from these common forms. The term used to describe general shape of a crystal is
habit. Some common crystal habits are as follows:
Cubic - cube shapes; Octahedral - shaped like octahedrons; Tabular - rectangular shapes; Equant -
a term used to describe minerals that have all of their boundaries of approximately equal length;
Fibrous - elongated clusters of fibers; Acicular - long, slender crystals; Prismatic - abundance of
prism faces; Bladed - like a wedge or knife blade; Dendritic - tree-like growths; Botryoidal -
smooth bulbous shapes.

Crystal twins: Sometimes during the growth of a crystal, or if the crystal is subjected to stress or
temperature/pressure conditions different from those under which it originally formed, two or more
intergrown crystals are formed in a symmetrical fashion. These symmetrical intergrowths of
crystals are called twinned crystals. Because symmetry is added to a crystal by twinning, twining
can be defined by the symmetry operations that are involved. These include:

(i) Reflection across a mirror plane. The added mirror plane would then be called a twin plane.
(ii) Rotation about an axis or line in the crystal. The added rotation axis would then be called a twin
axis.
(iii) Inversion through a point. The added center of symmetry would then be called a twin center.
Only non-centrosymmetrical crystals can form such twins.

Twin along a {010} plane Twin along a {111} axis

8
Twinned crystals may be described as follows:
(i) Simple twins – composed of only two parts.
(ii) Multiple twins – composed of more than two orientations.
(iii) Contact twins – this occur if a definite composition plane is present.
(iv) Penetration twins – occur if two or more parts of a crystal appear to interpenetrate each other
with the surface between the parts being indefinable and irregular.
(v) Polysynthetic twins – occurs when three or more individuals are repeated alternately on the
same twinned plane. If the individuals of polysynthetic twins are thin plates, the twinning is called
lamellar e.g. plagioclase feldspars.

Simple twin Multiple twin Contact twin Penetration twin Polysynthetic twin

9
Chapter 4: X-ray Diffraction by Crystals

Generation of X-ray: X-rays are a form of electromagnetic radiation of high


penetrating power. X-rays have a wavelength in the range of 0.1 Å to 10 Å
and energies in the range of 120 eV to 120 keV. They are shorter in
wavelength than UV rays and longer than gamma rays. X-rays are also called
Röntgen radiation, after Wilhelm Conrad Röntgen, who is generally credited
as its discoverer, and who had named it X-radiation to signify an unknown W. C. Röntgen
type of radiation.

X-rays are produced in a device called an X-ray tube. It consists of an evacuated chamber with a
tungsten filament at one end of the tube, called the cathode, and a metal target at the other end,
called an anode. Electrical current is run through the tungsten filament, causing it to glow and emit
electrons. A large voltage difference is placed between the cathode and the anode, causing the
electrons to move at high velocity from the filament to the anode target. Upon striking the atoms in
the target, the electrons dislodge inner shell electrons resulting in outer shell electrons having to
jump to a lower energy shell to replace the dislodged electrons. These electronic transitions results
in the generation of X-rays. The X-rays then move through a window in the X-ray tube. These
emitted X-ray photons have energies that are equal to the difference between the upper and lower
energy levels of the electron that filled the core hole.

Schematic cross section of an X-ray tube

Thus X-rays are emitted from a target when it is bombarded with high-energy electrons. These
electrons knocked out the tightly bound electrons in the K or L electronic shell of the target. The
low-energy empty levels are filled by a falling back of the electrons from higher energy levels to
the inner levels. The energy liberated during this process is in the form of X-rays. The wavelength
of the emitted X-rays is dependent upon the energies of the two levels involved and hence
characteristics of the element.

1
M(3d) M(3p) M(3s) The selection rules for a single electron moving from
L(2p)
L(2s) one orbital to another are
l = ± 1 and
K(1s) j = 0, ± 1
K(1s) L(2s) Forbidden
Nucleus
K = K(1s) L(2p) Allowed
K(1s) M(3s) Forbidden

K = K(1s) M(3p) Allowed

X-rays due to transition from the L to the K shell are called


Kα X-rays. Kα1 and Kα2 corresponding to electrons K

originating in their different spin states (1/2, 3/2) at the L-


shell. X-rays due to transitions from the M to K-shell are Relative
intensity
called Kβ and so on. The energy difference between the M K
and K levels is larger than that between the L and K levels,
therefore, the Kβ lines appear at shorter wavelength. Note
that the Kα line is about 5 - 10 times as intense as the Kβ Wavelength, A
line.

The X-ray radiation used for most crystallographic studies should be as nearly monochromatic as
possible and should have appropriate wavelength. The Kα lines fulfill this requirement, but the
presence of the accompanying Kβ is a nuisance. Fortunately, selective filters may be found that will
remove the Kβ to any desired extent, with a relatively much smaller loss of Kα.

The ideal choice of material for an X-ray filter is a


metal whose atomic number, Z, is one less than that of K
K-edge
the anode target metal for first row transition metals
Ni
(or two less for second row transition metals). Because Relative
intensity
their absorption edges fall between the Kα and Kβ lines K
of the target material. For example, the absorption
edge of nickel metal at 1.488 Å lies between the Kα
(λ = 1.542 Å) and Kβ (λ = 1.392 Å) X-ray spectral Wavelength, A
lines of copper. Thus the effect of passing X-rays from
a copper target through a thin nickel foil is that the Kβ radiation is selectively almost completely
absorbed.

2
The thickness of the filtering material is usually chosen to reduce the intensity of the Kβ line by a
factor of 100 while reducing the intensity of the Kα line by a factor of 10 or less. The optimum
thickness, x of the filter can be determined from the mass-absorption law:
I/Io = exp{−(μ/ρ)ρx}
where (μ/ρ) is the mass absorption coefficient of the material, ρ is the density of the material, and I
and Io are the transmitted and incident X-ray intensities, respectively.

The filter CuKα line (λ = 1.5418 Å) is a close doublet n, l, j


(Kα1 = 1.5405 Å and Kα2 = 1.5443 Å) because L3 2, 1, 3/2
transitions can occur from two possible electronic L2 2, 1, 1/2

configurations (1s1/2 ← 2p1/2, 2p3/2), which differ L1 2, 0, 1/2


slightly in energy. Kα1 is twice as intense as Kα2. Thus
K2 K1
the value of 1.5418 Å is a weighted mean (2Kα1 +
Kα2)/3, the weights being derived from the relative
K 1, 0, 1/2
intensities (2:1) of the Kα1 and Kα2 lines.
Atomic levels involved in copper Kα
emission.

X-ray wavelengths of commonly used target materials


Target Kα1, Å Kα2, Å K ͞α, Å Kβ, Å Filter
Cr 2.2896 2.2935 2.2909 2.0849 V
Fe 1.9360 1.9399 1.9373 1.7567 Mn
Co 1.7889 1.7928 1.7902 1.6208 Fe
Cu 1.5405 1.5443 1.5418 1.3923 Ni
Mo 0.7093 0.7135 0.7107 0.6323 Nb
Ag 0.5594 0.5638 0.5608 0.4971 Pd

K ͞α is the intensity-weighted mean of Kα1 and Kα2.

Properties of X-rays:
(i) X-rays are highly penetrating, invisible rays,
(ii) They have a very short wavelength (about the same size as the diameter of an atom,
(iii) X-rays are electrically neutral,
(iv) X-rays cannot be deflected by electric field or magnetic field,
(v) Ionization of a gas results when an X-ray beam is passed through it,
(vi) Photographic film is blackened by X-rays,
(vii) They are absorbed (stopped) by metal and bone,
(viii) X-rays interact with matter produce photoelectric and Compton effect.

3
Diffraction of X-ray by crystals: Diffraction is the bending of a wave around objects or the
spreading after passing through a gap. Diffraction processes are most noticeable when the
obstruction or gap is about the same size as the wavelength of the impinging wave.

Diffracted Diffracted
waves waves
Diffracted waves

Diffraction effects get less obvious as the gap gets


larger
When a beam of light passes through two adjacent pinholes, it forms a
pattern of alternating bright and dark regions - the diffraction pattern,
occurs whenever the distance between the pinholes is comparable to the
wavelength of the light. The bright regions appear where light waves
reinforce each other by arriving in phase, an effect called constructive
interference. The intervening dark areas occur where light waves arrive A diffraction pattern
out of phase and cancel each other (destructive interference).

Crystals are ordered, three-dimensional arrangements of atoms with characteristic periodicities. As


the spacing between atoms is on the same order as X-ray wavelengths (1-3 Å), crystals can diffract
the radiation when the diffracted beams are in-phase. Note that the X-ray scattering units are the
electron clouds associated with the atoms in the crystal structure.

Bragg’s equation: In 1913, British Physicist Sir William


Henry Bragg and his son Sir William Lawrence Bragg
noted the similarity of diffraction to ordinary reflection and
deduced a simple equation treating diffraction as reflection
from planes in the crystal lattice. Figure shows a beam of
X-rays falling on the crystal surface. Two successive Sir W. H. Bragg Sir W. L. Bragg
planes of the crystal are shown separated by a distance d. They were awarded the Nobel Prize
Let the X-rays of wavelength λ strike the first plane at an in Physics in 1915.
angle θ. Some of the rays will be reflected at the same angle. Some of the rays will penetrate and
get reflected from the second plane.

4
Incident rays Reflected rays

Atom or ion
A
 
  d Crystal planes
C D

B
Diffraction of X-rays from crystal planes.

These rays will reinforced those reflected from the first plane if the extra distance traveled by them
(CB + BD) is equal to integral number, n, of wavelengths. That is,
nλ = CB + BD …………………………….(i)
Geometry shows that
CB = BD = ABsinθ……………………....(ii)
From (i) and (ii) it follows that
nλ = 2ABsinθ
nλ = 2dsinθ

This is known as the Bragg’s equation. When Bragg’s equation is satisfied, the reflected beams are
in phase and interference constructively. At angles of incidence other than the Bragg angle,
reflected beams are out of phase and destructive interference or cancellation occurs.

Since the maximum value of the sinθ is unity, then for a given wavelength of X-rays, there is a
lower limit to the spacing, d, that can give observable diffraction lines, viz. dmin = λ/2.

Problem: X-rays of wavelength 0.154 nm strike an aluminium crystal; the rays are reflected at an
angle of 19.3o. Assuming that n = 1, calculate the spacing, d, between the planes of aluminium
atoms in pm that is responsible for this angle of reflection.
Solution:
λ
d = --------- [ ∵ n = 1]
2sinθ
1000 pm
0.154 nm × ------------
1 nm
= --------------------------- = 233 pm
2 sin19.3o

5
Problem: Determine the expected diffraction angle for the first-order reflection from the (310) set
of planes for BCC chromium when monochromatic radiation of wavelength 0.0711 nm is used.
Given that atomic radius of Cr is 0.1249 nm.

Solution: For a BCC unit cell,


a = 4R/√3 [R = Radius of a given atom)
= (4 × 0.1249 nm)/√3 = 0.2884 nm
3a
The distance between the 310 planes
d310 = a/√(32 + 12 + 02) a
= 0.2884 nm/√10 = 0.0912 nm a
From Bragg’s equation, 2a
a
sinθ = nλ/2d310 = (1 × 0.0711 nm)/(2 × 0.0912 nm) = 0.390
∴ θ = sin-1(0.390) = 22.94o BCC Lattice

∴ 2θ = 45.88o

Reciprocal lattice: From the Bragg’s equation, sinθ = nλ/2(1/d), it is seen that sinθ is inversely
proportional to d, the interplaner spacing in the crystal lattice, i.e., structure with large d will exhibit
compressed diffraction patterns and conversely for small d. Interpretation of X-ray diffraction
patterns would facilitated if inverse relation between sinθ and d could be replaced by a direct one.
This can be achieved by constructing a reciprocal lattice based on 1/d, a quantity that varies directly
as sinθ. The reciprocal lattice can be defined as follows:

Consider normals to all possible direct lattice planes (3,1)


(2,1)
(hkl) to radiate from some point taken as the origin. (1,1) *
Terminate each normal at a point that is a distance 1/d (0,1)
* *
from this origin, where d is the perpendicular distance
*
(0,1)
b
between planes of the set (hkl). The set of points so
O a
determined constitutes the reciprocal lattice.

A two dimensional example of the process is shown in (1,1)


figure in which reciprocal lattice points corresponding
to the direct lattice planes (0,1), (1,1), (2,1) and (2,1)
(3,1)
(3,1) are designated by asterisks and assigned the
Planes in two-dimensional direct space
indices of the planes they represent.
represented by points in reciprocal space.
Note that a short axis in real space (the space of the crystal) leads to a large separation between
spots in reciprocal (diffracted) space and that a long real axis of the unit cell leads to a short

6
separation between spots. Also note that obtuse angles in the real unit cell lead to acute angles in
the reciprocal cell, and vice versa.

The relationships between the direct and reciprocal lattices in three dimensions depend on the
angles between the axes in the direct lattice.

Orthorhombic direct and reciprocal cell

a* = 1/a, a = 1/a*
c* b* = 1/b, b = 1/b*
c* = 1/c, c = 1/c*
c
b* α = β = γ = α* = β* = γ* = 90o
b
V* = 1/V = a*b*c*
o a* a V = 1/V* = abc

Monoclinic direct and reciprocal cell

a* = 1/a(sinβ), a = 1/a*sinβ*)
c* b* = 1/b, b = 1/b*
c* = 1/c(sinβ), c = 1/c*sinβ*)
c α = γ = α* = γ* = 90o
*
b

b β* = 180 - β

Z a* V* = 1/V = a*b*c* sin β*
Y o a

V = 1/V* = abc sinβ
X

Triclinic direct and reciprocal cell

a* = bc sinα /V, a = b*c* sinα* /V*


b* = ac sin β /V, b = a*c* sin β * /V*
b* c* = ab sinγ /V, c = a*b* sinγ * /V*
*
c
V* = 1/V = a*b*c* √(1 – cos2α* – cos2β* – cos2γ*
c
+ 2cos α* cosβ* cos γ*)
*
b a
V = 1/V* = abc √(1 – cos2α – cos2β – cos2γ +
Z o a
Y 2cos α cosβ cos γ)
X

7
Bragg’s law in reciprocal lattice: The most important property of the reciprocal lattice is that it
allows a simple visualization of Bragg’s law. Suppose a crystal in a beam of X-rays of wavelength
λ. The crystal is oriented in such a way that the X-ray beam is parallel to its reciprocal section a*c*
plane. A line XO is drawn in the direction of the beam and
passing through the reciprocal lattice origin O. Finally a O a*
circle of radius 1/ λ is depicted with its center C on XO and 1/λ c*
located so that O falls on its circumference (Figure). The C 

angle OPB is inscribed in a semicircle and thus is a right 


angle. ray B P
X-
Therefore,
sin OBP = sin θ = OP/OB = OP/ (2/λ) The reciprocal lattice and the sphere

sin θ = (OP/2) λ of reflection.

But since P is a reciprocal lattice point, the length of OP is by definition equal to 1/dhkl. Substitution
gives,
sin θ = 1 λ /2dhkl
1 λ = 2dhkl sin θ …………………………(1)
which is just Bragg’s law with n =1.

Sphere of reflection (Ewald sphere): A construction for considering


conditions for diffraction in terms of the reciprocal rather than the real
lattice. It is a sphere, of radius 1/λ, with the incident beam along a
diameter. The origin of the reciprocal lattice is positioned at the point
where the incident beam emerges from the sphere. Whenever a
reciprocal lattice point touches the surface of the sphere, the conditions
Paul Peter Ewald
for a diffracted (or reflected) beam are satisfied.

Limiting sphere: Since a reciprocal lattice point Sphere of Limiting sphere


must pass through the sphere of reflection in reflection
order to produce an observable reflection, only
those points where the distance from the 1/ 2/
X-ray beam
reciprocal lattice origin is less than the diameter
of this sphere can be observed. Thus a sphere of
radius equal to this diameter, 2/λ, described
about the origin of the reciprocal lattice, enclose
The sphere of reflection and the limiting sphere.

8
all of the possible reflecting points for a given wavelength. Such a sphere is known as the limiting
sphere or, since its contents vary with the radiation used, as the copper sphere or molybdenum
sphere.

Crystal density measurements: Crystal density is the physical property of a crystal, which can be
useful in interpreting and using the subsequent X-ray results. It is a function of atomic weight,
atomic radius and the structure of the aggregate. The relationship between the volume of the unit
cell, V measured in cm3 and the density dc in gcm-3 is

dc = (1.660 × FW × Z) / V Density, d = Mass / Volume


= Molar mass / Molar volume
= FW / (N × Volume of one formula unit)
= (FW × Z) / (N × Z × Volume of one formula unit)
= (FW × Z) / (N × Volume of one unit cell in Å3)
= (FW × Z) / (N × Volume of one unit cell in cm3 × 10-24)
= (FW × Z) / 6.023 × 1023 × V in cm3 × 10-24)
= (1.660 × FW × Z) / V
where FW is the formula weight of the substance concerned and Z is the number of such formula
units within the cell. Thus it is provided that the density can be measured or estimated, the above
relationship can be used to check analysis, to determine molecular weights and occasionally to
establish composition that are not readily determined by other means.

The technique most commonly used for determining the density (dobs) of small crystals is flotation
in a mixture of liquids (where the crystals are insoluble) one lighter and one heavier than the
crystal, whose proportions are adjusted until the crystal remains suspended in the medium i.e., it
neither sinks nor rises to the surface of the resulting mixture. The density of the liquid mixture
(with the same density as that of the crystal) is then found by weighing a known volume in a
specific gravity bottle or picnometer. A pair of solvents used is heptane (d = 0.684 g cm -3) and
carbon tetrachloride (d = 1.589 g cm-3), whose mixture has the advantages of being comparatively
poor solvents for many compounds and of covering a wide range of densities.

Molecular weight calculations: The standard physico-chemical methods of determining molecular


weight are not always effective; for very sparingly soluble substances, for instance, the elevation of
the boiling point or depression of the freezing point of the strongest solution obtainable may be so
small that only an approximate estimate of the molecular weight is obtained. In such circumstances,
crystallographic method is often used for more accurate determination if crystals suitable for the
determination of unit cell dimensions are available. From the unit cell dimension, the volume V of
the unit cell may be calculated. Multiplying the volume V (in cm3) by the density d (in g cm-3)
9
yields the weight of matter in the unit cell, which is the weight of either one or a small whole
number Z of molecules; Vd = Z FW, where FW is the formula weight of one molecule.

From the knowledge of approximate weight of a molecule we can find the number of molecules Z
constituting the unit cell crystal pattern. Having found Z, we can then calculate an accurate value
for FW, which is Vd/Z. The molecular weight M on the chemical scale (in Daltons) is equal to
Vd/Z × 0.6023 × 1024.

For example, the molecular weight of a dyestuff could only be estimated roughly by standard
methods to be between 400 and 500 on the chemical scale. And the rotation and Weissenberg
photographs of a small crystal of this dyestuff gave the following dimensions,
a = 13.360 Å, b = 10.440Å, c = 3.866 Å, α = 90o, β = 76o, and γ = 90o

Then, V = 13.360 Å × 10.440 Å × 3.866 Å × sin 76o


= 523.20 Å3 = 523.20 × 10-24 cm3
Suppose its density d = 1.528 g cm-3
Therefore, mass of the unit cell = 523.20 × 10-24 cm3 × 1.528 g cm-3 × 0.6023 × 1024 = 484
This is about equal to the approximate molecular weight, so that there is evidently only one
molecule in the provisional unit cell, and the precise molecular weight is, therefore, 484 on the
chemical scale.

X-ray absorption: When x-rays encounter any form of matter, they are partly transmitted and
partly absorbed. It was found experimentally that
I ∞ x, where I = intensity and x = distance.
In differential form
- dI/I = μdx, where  is linear absorption coefficient.
After integration
Ix = I0e-μx, where Ix = transmitted beam intensity, I0 = incident beam intensity.
Ix = I0e-(μ/ρ)ρx, where (μ/ρ) = mass absorption coefficient and ρ = density.

Mass absorption coefficient: The mass absorption coefficient of the substance containing more
than one element is a weighted average of the mass absorption coefficients of its constituent
elements. If w1, w2, w3 , ... are the weight fractions of elements 1, 2, 3, ... and (/)1, (/)2, (/)3,
... their mass absorption coefficients then
/ = w1(/)1 + w2(/)2 + w3(/)3 + ……
The mass absorption coefficient of an absorber is also related to its atomic number by the formula
/ = kλ3Z3, where k = a constant and Z = atomic number of absorber.

10
Characteristics of X-ray diffraction (XRD)
(i) Usually only the K-lines are useful in X-ray diffraction.
(ii) X-rays interact with the electron shells of atoms.
(iii) Both the diffraction power and the absorption for X-rays increase with the number of electrons,
i.e. samples containing heavy elements give higher diffraction intensities, but are also more
susceptible to absorption effects.
(iv) Atoms or ions with a very similar number of electrons cannot be distinguished in a crystal
structure by XRD.
(v) XRD is a method to analyze the average bulk structure of long range ordered materials.

Problem: The crystal structure of SrTiO3 is cubic, space group Pm3m with a unit cell edge a = 3.90
Å. Calculate the expected 2θ positions of the first three peaks in the diffraction pattern, if the
radiation is Cu Kα (λ = 1.54 Å).

Solution:
(hkl) = (100) (hkl) = (110)
1/d2 = (12 + 02 + 02)/(3.90 Å)2 1/d2 = (12 + 12 + 02)/(3.90 Å)2
d = 3.90 Å d = 2.76 Å
sin θ100 = 1.54 Å/{2(3.90 Å)} sin θ110 = 1.54 Å/{2(2.76 Å)}
θ 100 = 11.4° Θ110 = 16.2°
2θ100 = 22.8° 2θ100 = 32.4°

(hkl) = (111)
1/d2 = (12 + 12 + 12)/(3.90 Å)2
d = 2.25 Å
sin θ111 = 1.54 Å/{2(2.25 Å)}
θ 111 = 20.0°
2θ111 = 40.0°

Problem: What are the three largest d-spacing values for families of lattice planes (e.g. {hkl}) in
elemental gold given the lattice is face-centered cubic with a = 4.0785 Å?

Solution:
For {100} planes,
1/d100 2 = (12 + 02 + 02)/(4.0785 Å)2
d100 = 4.0785 Å

11
For {110} planes,
1/d110 2 = (12 + 12 + 02)/(4.0785 Å)2
d110 = 2.8839 Å

For {111} planes,


1/d111 2 = (12 + 12 + 12)/(4.0785 Å)2
d111 = 2.3547 Å

Problem: X rays of λ = 0.1537 nm from a Cu target are diffracted from the (111) planes of an FCC
metal. The Bragg angle is 19.2o. Calculate the Avogadro number if the density of the crystal is
2698 kg/m3 and the atomic weight 26.98.

Solution: The distance between the 111 planes


d111 = 0.1537 × 10-9 m / 2 sin(19.2o)
= 2.337 × 10-10 m
The edge length a, related to d by
a = d111√(h2 + k2 + l2), where hkl = (111) are Miller indices
a = (2.337 × 10-10) √(12 + 12 + 12) = 4.048 × 10-10 m
Mass of unit cell = 4 × 26.98 × 10-3 kg [∵ A FCC lattice contains 4 metal atoms per unit cell]

And density = 2698 kg/m3


Density = Mass/Volume
∴ 2698 kg/m3 = 4 × 26.98 × 10-3 kg / [N × (4.048 × 10-10 m)3]
∴ N = 6.030 × 1023 particles/mole

12
Chapter 5: Powder Diffraction Technique

X-ray diffraction (XRD): X-ray diffraction relies on the dual wave/particle nature of X-rays. The
electric fields of X-rays interact with the electron clouds of atoms/ions in the crystal lattice results
into the scattering of X-rays in all directions. When scattered waves traveling in the same direction
encounter one another and they undergo interference. The process of constructive interference is
diffraction. Bragg’s law gives the condition for diffraction to occur from electrons in stacks of
layers of interplanar distance dhkl, nλ = 2dhkl sinθ where θ is the diffraction angle, n is an integer,
and λ is the wavelength of the X-rays. The intensity of the diffracted beam depends on the
wavelength of the X-rays but also increases with the atomic numbers of the atoms.

Constructive (left) or destructive (right) interferences The diffraction of X-rays

In order for an X-ray to diffract, the sample must be crystalline and the spacing between atom
layers must be close to the radiation wavelength. If beams diffracted by two different layers are in
phase, constructive interference occurs and the diffraction pattern shows a peak. However, if they
are out of phase, destructive interference
occurs appear and no peak is observed.
Diffraction peaks only occur if it follows
Bragg’s Law. Since, a highly regular structure
is needed for diffraction to occur, only
crystalline solids diffract, the X-ray powder
diffraction of amorphous materials do not
depict any significant peak in diffraction
pattern.

X-ray diffraction is a laboratory-based


technique commonly used for identification of Diffractograms of crystalline, semi-crystalline and
crystalline materials and analysis of unit cell amorphous samples
dimensions. The technique was pioneered by Max von Laue in 1912. There are two primary types
of XRD analysis (i) X-ray powder diffraction and (ii) single-crystal X-ray diffraction.

1
X-ray powder diffraction: Growing a single crystal needs some 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 samples. This is commonly known as the powder method. A
powder pattern is produced when X-rays are diffracted by a sample consisting of a very large
number of randomly oriented crystallites (typically 1 to 10 micron-sized,
ideally imperfect crystals) compacted into a powder.

In powder or polycrystalline diffraction it is important to have a sample


with a smooth plane surface. If possible, we normally grind the sample
down to particles of about 0.002 mm to 0.005 mm cross section. The ideal
sample is homogeneous and the crystallites are randomly distributed.

Instrumentation for powder diffraction developed over the years, from cameras (Debye-Scherrer-
Hull, Guinier-Hägg) to sophisticated powder diffractometers (Bragg-Brentano parafocusing
geometry, flat plate or capillary samples, Guinier geometry with transmission samples). Detector
used includes photographic film; proportional, scintillation and solid state counters; position-
sensitive detectors; 2-D area detectors; image plates). More synchrotron radiation sources become
available recently, allowing high resolution and very fast data collections to be made.

Debye-Scherrer-Hull method: This method was developed as early as 1916 by Debye and Scherrer
and Hull (1917), for more than 50 years its use was almost exclusively limited to qualitative and
semi-quantitative phase analysis and macroscopic
stress measurements. The Debye-Scherrer
geometry employs a parallel beam, and the
sample has a roughly cylindrical symmetry
(either packed into a thin walled glass capillary
tube and sealed, or coated onto a thin glass fiber
which has been moistened with an adhesive,
Geometric principles of X-ray powder diffraction
such as gun Arabic, tragacanth or petroleum
measurements: Debye-Scherrer-Hull method
jelly).
(parallel beam)

The Debye-Scherrer camera consists of a hollow metal cylinder with the following components: (i)
a light proof lid, (ii) a collimator to admit a thin X-ray beam onto the sample, (iii) a beam stop to
absorb the transmitted beam, (iv) two metal clips to position and hold the film tightly against the
circumference of the camera, and (v) a sample holder mounted on a moveable magnetic disc which
can be rotated by a motor.

2
Debye-Scherrer camera

Four methods of film mounting are common in Debye-Scherrer types of cameras: (i) Straumanis,
(ii) Bradley-Jay, (iii) van Arkel and (iv) Bradley-Bragg. The Bradley-Jay, van Arkel and Bradley-
Bragg methods have their film mounted symmetrically with respect to the incident beam, whereas
the Straumanis method has the film mounted asymmetrically.

Four methods of mounting films: (a) Straumanis, (b) Bradley-Jay and (c) van Arkel and (d)
Bradley-Bragg.
Among them Straumanis method of film mounting is probably the one most commonly used,
because the derivation of θ from S is very simple. The camera is loaded with film in a dark room: a
strip of film is placed in a guillotine, trimmed to size and two holes punched into it. The film is
placed in the camera and the two metal clips adjusted to hold the film firmly in place. Finally the
collimator and beam stop are inserted, and the lid secured in position.

3
The camera is then placed against the window of an X-ray tube and exposed to X-rays. Exposure
times vary from about 20 minutes to several hours, depending on the nature of the sample, size of
camera, type of X-ray tube and tube operating conditions. In order to ensure complete randomness
of the small crystals which make up the sample, the sample holder is usually rotated during
exposure to the X-ray beam.

When X-rays are incident on a fine powder, there will always be a large number of crystallites
oriented in such a way that a given set of planes (hkl), which make an angle θ with the X-ray beam
can cause reflection to occur. These planes are tangent to the surface of a cone with a cone angle of
2θ. The beams reflected by these planes lie on the surface of a cone with a cone angle of 4θ.

Formation of diffraction cone Intersection of cones of diffracted rays with


Debye-Scherrer-Hull film
After due exposure, the film is removed from the camera in a dark room and processed. This simply
involves placing the film in developing solution, then in fixer, followed by washing and drying.
When the strip is laid flat, the intersections of the cones of diffraction with the film form pairs of
arcs around the exit and entrance holes through which the X-rays pass (Fig.). Thus although most
of the crystals in a powdered sample may not be diffracting at all, there are usually sufficient in the
correct position to contribute to the arc of diffraction seen on the film.

To obtain the d-spacings of the crystal from such a film, the separation S between pairs of
corresponding arcs (Fig.) around the exit hole is measured, and given the camera radius to be R,
then
S 4θ
------- = ------
2πR 360
4
For R = 28.65 mm (2πR = 180 mm), the measured value of S in mm is thus equal to the value of 2θ
in degrees. Thus we measure the arc diameters in mm and divide by 2 we get values of Bragg
angle, θ. The d-spacings are then be calculated from the Bragg equation, dhkl = λ/(2sinθ). Arc
diameters are usually measured by means of a travelling microscope placed over a film illuminated
with light from underneath.

The advantage with Debye-Scherrer method is that it requires a very small amount of sample, from
50 μg to a few mg depending on the size of the capillary. It is also convenient to handle air or
moisture sensitive samples as the top of the capillary can be easily sealed. The data collection time
can be made very short, it is often sufficient with a few minutes for phase identifications.

Guinier-Hägg method: Useful as is the Debye-Scherrer camera it has two important disadvantages.
The lines tend to be rather broad and there is therefore a lack of resolution of lines lying close
together, and also the white continuum in the spectrum of the X-ray tube is reflected over a wide
range of angles by each set of planes and leads to a rather dark background to the photographs. This
background obscures very weak lines.

The Guinier-Hägg camera is designed to provide high resolution


and low background with reasonable exposure times, and is based
on a focusing principle. Focusing and monochromatization of X-
ray radiation from X-ray tubes in Guinier geometry is achieved by
a bent quartz (α-SiO2) crystal in the Johansson configuration. The
curved crystal plane easily transforms the divergent beam of
The Guinier-Hägg
characteristic radiation of a fine-focus X-ray tube to convergent
camera
Kα1 X-rays, completely suppressing the Kβ and Kα2 spectral lines.

The sample holder consisted of a circular brass plate with a central hole
admitting the X-rays from the monochromator. The hole was covered by
adhesive tape onto which the sample powder was sprinkled. The sample
Sample holder
holder was rotated by a simple electric motor device.

The transmission arrangement of Guinier method is realized with the convergent primary X-ray
beam. The convergent incident beam is focused on the
internal surface of the cylindrical chamber of the
camera. A thin powder sample measured in
transmission geometry is fixed at the surface of the
chamber on an X-ray transparent substrate at the
entrance window. The monochromatic beams diffra- Schematic arrangement of a Guinier-Hägg
cted at Bragg angles 2θhkl after the primary convergent focusing camera

5
X-ray beam transmission through the sample,
are also convergent and focused onto the
same cylindrical surface. Thus if a film is
Primary beam
placed on the surface of this cylinder a
Guinier-Hägg X-ray powder diffraction pattern
powder photograph is recorded on it, and
distance along the film is directly proportional to 2θ as on a Debye-Scherrer photograph. A
Guinier-Hägg X-ray powder diffraction pattern is shown in Fig.

The powder diffractometer: The tedious steps involved in the photographic method, namely the
long duration of exposure, developing the film in dark room and problems such as the shrinkage of
film and the measurement of relative intensities, have been overcome by the diffractometer
technique. Most modern X-ray diffraction laboratories rely on automated powder diffractometers.
The essential components of a typical X-ray powder diffractometer used in a materials analysis
laboratory: (i) a source of X-rays, usually a sealed X-ray tube, (ii) a goniometer, which provides
precise mechanical motions of the tube, sample, and detector, (iii) an X-ray detector, and (iv)
electronics for counting detector pulses in synchronization with the positions of the goniometer.

Diffractometers can be operated both in transmission and in reflection configurations. The


reflection one is more common. The powder sample is filled in a small disc-like container and its
surface carefully flattened. The disc is put on one axis of the diffractometer and tilted by an angle θ
while a detector (scintillation counter) rotates around it on an arm at twice this angle. This
configuration is known under the name Bragg-Brentano theta-2 theta. Another configuration is the
Bragg-Brentano theta-theta configuration in which the sample is stationary while the X-ray tube
and the detector are rotated around it. The angle formed between the tube and the detector is 2theta.
This configuration is most convenient for loose powders.

A powder diffractometer Modern Bragg-Brentano laboratory diffractometer in


reflection geometry

6
In a diffractometer of Bragg-Brentano theta-2 theta geometry, the powder sample in the form of a
thin circular disc of about 15 mm diameter 1-2 mm thickness is placed on a holder standing
vertically. The holder can be rotated about an axis perpendicular to the table. The X-ray beam is
allowed to fall on the sample making an angle with the surface. An X-ray photon counter rotates in
a circle around the sample detecting all the reflected beams in turn. For better focusing of the
reflected beam on the detector, the sample is rotated at a speed half of that of the detector. The
detector converts the intensity into current and the diffraction angles are plotted as 2θ against the
current on a strip-chart recorder.

A typical diffraction pattern is shown in Fig. Such a pattern is usually referred to as a


diffractometer trace. A linear intensity scale was used in the recording of the trace. Note that the
instrument has automatically printed out a scale of 2θ.

A powder diffractogram

As a powder sample consists of a multitude of single crystals of the same crystal form, which are
randomly oriented, the powder diffractogram consists of the overlay of an infinite number of single
crystal diffractograms with random orientation. Thus, a powder diffractogram, represented in two
dimensions, contains rings concentric to the incident beam. The usual final diffractogram is a radial
cut through those rings resulting in a plot of the measured intensity vs. the diffraction angle.

Indexing the powder patterns: Identifying every d value in the powder pattern of a pure solid
substance with a set of Miller indices (hkl) is known as indexing of a powder pattern. This is done
using the equations that relate the lattice parameters for the crystal a, b, c and α, β, γ with dhkl.
Indexing is simple for cubic crystals and not difficult for tetragonal and hexagonal crystals, but may

7
be quite difficult for crystals of low symmetry like those with monoclinic or triclinic unit cells. In
the cubic system the relationship between the d-spacing and the lattice parameter a is given by
a = dhkl √(h2 + k2 + l2) ……………………………………(1)
where, the integers h, k, l are the Miller indices of the set of lattice planes under consideration.
Bragg’s Law tells us the location of a peak with indices hkl. θ is related to the interplanar spacing,
dhkl , as follows:
λ = 2dhkl sinθ
1/dhkl2 = 4 sin2θ/λ2 ……………………………………..(2)

Combining equation (1) and (2)


(h2 + k2 + l2)/a2 = 4 sin2θ/λ2
or sin2θ = (λ2/4a2)(h2 + k2 + l2) ………………………….(3)
where λ and a are constants, hence λ2/4a2 is constant. Thus sin2θ is proportional to h2 + k2 + l2 i.e.,
planes with higher Miller indices will diffract at higher values of θ or vice versa.

For the primitive cubic unit cell, all integer values of the indices h, k, l are possible. Table shows
the values of hkl in order of increasing value of (h2 + k2 + l2) and therefore increasing sinθ values.
One value in the sequence, 7, is missing; this is because there are no possible integral values for
(h2 + k2 + l2) = 7. There are also other higher missing values where (h2 + k2 + l2) cannot be an
integer: 15, 23, 28, etc.

Table: Values of (h2 + k2 + l2)


hkl 100 110 111 200 210 211 220 300 = 221
(h2 + k2 + l2) 1 2 3 4 5 6 8 9

Taking equation (3), if we plot the intensity of diffraction of the powder pattern of a primitive cubic
system against sin2θhkl we would get six equispaced lines with the 7th, 15th, 23rd, etc. missing.
Consequently it is easy to identify a primitive cubic system and by inspection to assign indices to
each of the reflections. The unit cell dimension a is equal to the spacing of the 100 planes. An
example of Debye-Scherrer X-ray powder pattern of a cubic crystal with a primitive unit cell is
shown in Fig.
2 = 180o

2 = 0o

220 310 222 311 300, 221 211 200 110 100 111 210

300, 221 311 222 310 220 210 111 100 110 200 211

Powder diffraction pattern from a cubic crystal with a primitive unit cell

8
The pattern of observed lines for non-primitive body-centered and face-centered cubic lattices is
different from primitive. The differences arise because the centering leads to destructive
interference for some reflections and the missing reflections are known as systematic absences
(Fig.).

The reflection from the 200


plane is exactly out of phase
with the 100 reflection and
destructive interference occurs
and no 100 reflection is
observed.

Figs. show indexed Debye-Scherrer X-ray powder diffraction patterns from cubic materials with F
and I lattices respectively. For an I lattice only reflections for which h + k + l are even are present,
i.e. those for which h2 + k2 + l2 = 2, 4, 6, 8, 10, 12, 14, 16 ….. For an F lattice h, k, and l must be
either all even or all odd for a reflection to be present. The only lines with h 2 + k2 + l2 = 3, 4, 8, 11,
12, 16, ….. will be present.
2 = 180o

2 = 0o

220 222 310 211 110 200

(a)

310 222 220 200 110 211


2 = 180o

2 = 0o

220 222 311 200 111

(b)

311 222 220 111 200

Powder diffraction pattern from a cubic crystal (a) for a body-centered unit cell (I), and (b) for a
face-centered unit cell (F).

9
Table: Some values of (h2 + k2 + l2) for cubic crystals

Forbidden Primitive, P Body-centered, I Face-centered, F Corresponding


numbers hkl values
1 100
2 2 110
3 3 111
4 4 4 200
5 210
6 6 211
7 -
8 8 8 220
9 221, 300
10 10 310
11 11 311
12 12 12 222
13 320
14 14 321
15 -
16 16 16 400
17 322, 410
18 330, 411
19 331
20 420

We saw earlier that for a cubic crystal we could derive the relationship
λ2
sin2θ = ------ (h2 + k2 + l2)
4a2
The possible values of h2 + k2 + l2 are listed in Table for each of the cubic lattices. Using these
pieces of information we can see that if the observed sin2θ values for a pattern are in the ratio
1:2:3:4:5:6:8…., then the unit cell is likely to be primitive cubic, and the common factor is λ/4a2.

A face centered cubic unit cell can also be recognized: If the first two lines have a common factor,
A, then dividing all the observed sin2θ values by A gives a series of numbers, 3, 4, 8, 11, 12, 16,
…., and A is equal to λ/4a2.

A body centered cubic system gives the values of sin2θ in the ratio 1:2:3:4:5:6:7:8….with the
values 7 and 15 apparently not missing, but now the common factor is 2λ2/4a2.

10
Problem 1: Aluminium powder gives a diffraction pattern that yields the following eight largest
d-spacings: 2.338, 2.024, 1.431, 1.221, 1.169, 1.0124, 0.9289 and 0.9055 Å. Aluminium has a
cubic close packed structure and its atomic weight is 26.98 and λ = 1.5405 Å. Index the diffraction
data and calculate the density of aluminium.

Solution: The Bragg equation, λ = 2dhkl sinθ can be used to obtain sin, sinθ = λ/2dhkl. The ccp
lattice is an F type lattice and the only reflections observed are those with all even or all odd
indices. Thus the only values of sin2 in sin2 = A(h2 + k2 + l2) that are allowed are 3A, 4A, 8A,
11A, 12A, 16A, 19A and 20A for the first eight reflections.

Insert the values into a table and compute sin and sin2. Since the lowest value of sin2 is 3A and
the next is 4A the first entry in the Calcd. sin2 column is (0.10854/3)×4 etc.

dhkl/Å sinθ sin2θ Calcd. sin2θ hkl


2.338 0.32945 0.10854 111
2.024 0.38056 0.14482 0.14472 200
1.431 0.53826 0.28972 0.28944 220
1.221 0.63084 0.39795 0.39798 311
1.169 0.65890 0.43414 0.43416 222
1.0124 0.76082 0.57884 0.57888 400
0.9289 0.82921 0.68758 0.68742 331
0.9055 0.85063 0.72358 0.72360 420

The reflections have now been indexed.

Calculation of a
For the first reflection (for which h2 + k2 + l2 = 3)
sin2 = 3A = 3(λ2/4a2)
a2 = 3λ2/4sin2
a = 4.04946 Å = 4.04946 × 10-8 cm

Calculation of the density of aluminium

a3 = 66.40356 Å3 = 66.40356 × 10-24 cm3


If the density of aluminium is  (g cm-3), the mass of the unit cell is  × 66.40356 × 10-24 g
The unit cell of aluminium contains 4 atoms.
The weight of one aluminium atom is 26.98/(6.022 × 1023) = 4.48024 × 10-23 and the weight of four
atoms (the content of the unit cell) is 179.209 × 10-24.

11
∴  × 66.40356 × 10-24 = 179.209 × 10-24

or  = 2.6988 g cm-3

Problem 2: The X-ray powder diffraction pattern of AgCl obtained using radiation of wavelength
1.54 Å is shown below. The peaks are labelled with 2θ values.

On the basis that the structure is cubic and of either the NaCl or CsCl type, answer each of the
following:
(i) Index the first six reflections,
(ii) Calculate the unit cell parameter, and
(iii) Calculate the density of AgCl (Assume the following atomic weights: Ag, 107.868; Cl, 35.453;
and Avogadro’s number is 6.022 × 1023).

Solution: Since  values are available, sin2 values can be calculated and inserted in a Table.

2  sin2θ 2
Calcd. sin 
27.80 13.90 0.0577
32.20 16.10 0.0769 0.07693
46.20 23.10 0.1539 0.1539
54.80 27.40 0.2118 0.2116
57.45 28.73 0.2310 0.2308
67.45 33.73 0.3083 0.3077

From sin2 = A(h2 + k2 + l2) the possible values are:

12
(i) for a face centred lattice 3A, 4A , 8A, 11A, 12A and 16A
(ii) for a primitive lattice 1A, 2A, 3A, 4A, 5A and 6A
The second option is not possible as the first 2 are not in the ratio of 1:2. To test the first option,
divide the first by 3 and multiply the result by 4, 8 etc.

Density of AgCl
Since sin2 = (λ2/4a2)(h2 + k2 + l2)
a2 = (λ2/4 sin2)(h2 + k2 + l2) = [(1.54 Å)2/(4×0.3083)] (16 + 0 + 0) using the largest (most accurate)
2
or a2 = 30.7692 Å2
∴ a = 5.547 Ǻ

Formula weight of unit cell = 4 AgCl = 573.284 g. This is the weight of 4 moles of AgCl.
The weight of 4 molecules is 573.284 / (6.022 × 1023)
∴ Density = 573.284 g mol-1 / (6.022 × 1023 mol-1 )(5.547 × 10-8 cm)3 [1 Å = 10-8 cm]

= 5.580 g cm-3

Intensity of the lines: Intensity of the diffracted lines depends on the scattering power of the atoms
that occupy a Miller plane. Higher the number of electrons in the atom, higher is its scattering
power. This is so because the scattering of the electromagnetic radiation takes place by interaction
with electrons. Mathematically, the intensity of a peak Ihkl is given by:
Ihkl ∞ ‫׀‬Fhkl‫׀‬2
where Fhkl is the structure factor. Again the structure factor,
Fhkl, of a reflection, hkl, is dependent on the type of atoms and
their positions (x, y, z) in the unit cell.
Fhkl = ∑fi e2πi(hxi+ kyi+ lzi)
i
where fi is the scattering factor for atom i and is related to its
atomic number. Fig. shows the atomic scattering factors, f for
H, C, O, Cl-, Ca2+ and Fe2+, plotted against sinθ/λ.

Intensities are normally measured by diffractometry, as peak heights or peak areas at slow scanning
speeds. It is difficult to obtain quantitative intensities from films unless a microdensitometer is also
available.

X-ray powder pattern is a crystal’s fingerprint: Each crystalline substance has a unique
repeating three-dimensional array of atoms and produces a unique X-ray diffraction pattern. The
number of observed peaks is related to the symmetry of the unit cell (higher symmetry generally

13
means fewer peaks). The d-spacings of the observed peaks are related to the repeating distances
between planes of atoms in the structure. And finally, the intensities of the peaks are related to what
kinds of atoms are in the repeating planes. The scattering intensities for X-rays are directly related
to the number of electrons in the atom. Hence, light atoms scatter X-rays weakly, while heavy
atoms scatter X-rays more effectively. These three features of a diffraction pattern: the number of
peaks, the positions of the peaks, and the intensities of the peaks, define a unique, fingerprint X-ray
powder pattern for every crystalline material. For example, the X-ray diffraction patterns of the
isostructural compounds NaCl and KCl can be compared in Fig. Notice that the two patterns are
fairly similar. The relative intensities of the peaks differ, but the primary difference is that the
diffraction peaks for KCl appear at lower angles, higher d-spacings, than those for NaCl due to the
larger unit cell of KCl.

Intensity and its relation with the atomic arrangement: The relation between the intensity of
diffracted X-rays and the arrangement of atoms in a crystal can be best illustrated by considering
the Debye-Scherrer X-ray powder diffraction patterns of two similar face-centered cubic (fcc)
binary crystals, KCl and NaCl, shown in Fig.

X-ray power patterns (a) KCl and (b) NaCl

The first thing that strikes us is that the angles θ (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. Thus for instance, (110) is systematically absent but (111) may be observed. Fig. shows that
(111) planes have Na+ ions lying on the planes and Cl- ions midway between the planes. The Na+
and Cl- ions scatter exactly 180o out of phase with each other for these planes, but since they have
different scattering powers the destructive interference that occurs is only partial. Similar effects

14
are expected in case of KCl. However, K+ and Cl- ions are
isoelectronic and therefore have identical scattering power.
(111)
Since the radiation scattered by these planes are out of phase,
the reflected waves will get completely annihilated and the
(111)
reflections from (111), (311) etc. will not be seen.
(111)
(111)
Na+ Cl-

(111) planes in NaCl

Effect of crystal size on the powder pattern: The ideal size for powder diffraction depends on the
relative perfection of the polycrystalline material, but usually lies in the range 10-3 to 5×10-5 cm.
When the crystallites exceed the ideal size, there will be too few of them to satisfy the assumption
that they lie in all possible orientations. As a
result the lines on a powder photograph can
be seen to be spotty. At the onset of the
effect the lines are continuous but contain
Crystallites Crystallites Crystallites
more intense spots corresponding to rays
10-3 - 5×10-5 cm > 10-3 cm < 2×10-5 cm
reflected from the largest particles present. diameter diameter diameter
Debye-Scherrer X-ray powder diffraction pattern

If the crystallites are smaller than the ideal size (< 2000 Å diameter), the number of parallel planes
available is too small for a sharp diffraction maximum to build up, and the lines in a powder
photograph become broadened (Fig.). This is because if there is not a corresponding plane that can
cause destructive interference, then the intensity is only partially cancelled out. Hence one can
observe intensities close to Bragg angles.

(a) (b)
Fig. (a) Hypothetical Bragg peak and (b) the experimental observed Bragg peak showing the effect
of the crystallite size

As the crystallite size decreases, the X-ray peaks broaden, so it is possible to obtain a measure of
the crystallite size. The Scherrer equation uses this concept to provide an approximation of
crystallite size using the position and broadening of the peak, such that
15

τ = ---------
β cosθ
where τ is the mean size of the ordered domain or crystallite; K is a dimensionless shape factor with
typical values close to unity; λ is the incident radiation wavelength; β is the full-width at half-
maximum (FWHM) of the peak; and θ is angular position of the peak. The Scherrer equation can
only be applied to crystallites with an average size less than 0.1 μm, i.e. nanoparticles; for
crystallites much larger than this value, the Bragg peak broadening is too small to quantify
accurately with this method.

High temperature powder diffraction: Several commercial instruments are available for
recording powder patterns at high temperatures. Some are diffractometers fitted with a small
furnace around the sample. The powder pattern is recorded in exactly the same way as for room
temperature operations. Inert, refractory construction materials such as tungsten and iridium are
used and very high temperatures, e.g. 2000oC, are available.

Film methods are also available and a particularly elegant one is the Guinier-Lenne camera which
operates up to ~1200oC. A thin powdered sample, mounted on fine platinum gauze is suspended in
the middle of a small furnace. A quartz crystal monochromator is used to provide a convergent X-
ray beam and the diffraction pattern is recorded on film by the focusing method. The sample and
furnace may be programmed to heat or cool
at a certain rate and a continuous X-ray
photograph is taken. The film is rectangular
and bent to lie on the focusing circle
(cylinder) of the camera. A schematic
photograph showing the polymorphic
changes that occur on heating Li2ZnSiO4 is
shown in Fig. The horizontal axis is 2θ or d-
spacing, as usual, and the vertical axis is Schematic high temperature Guinier powder
temperature. photograph of Li2ZnSiO4

High temperature powder diffraction is invaluable for identifying and studying phases that exist
only at high temperatures. A more technical application of high temperature powder diffraction is
in the measurement of coefficients of thermal expansion, data which for some materials may be
very difficult to obtain using conventional dilatometry.

16
Comparison of diffractometry with film methods

Feature Diffractometry Film method (Focusing camera)


Exposure time Usually 30 min 10 min – 1 hr
Accuracy of 2θ values Good – very good Good – very good
Intensities Very good Poor - fair
Peak shape Very good Poor - fair
Comparison of different samples Poor (clumsy) Excellent
Resolution of closely spaced lines Good - excellent Excellent
Amount of sample required 0.05 – 2 g ~ 1 mg
Storage and retrieval of results Clumsy, unless done by Easy
computer

Uses of powder diffraction: In general, powder diffraction data are unsuitable for solving crystal
structures. Some advances have recently been made using the Rietveld method. However, this is far
from trivial and it works best in relatively simple cases. It is very difficult to be sure that the unit
cell is correct as the reflections overlap and are difficult to resolve from one another. However,
some important advantages and uses of powder diffraction are
(i) The need to grow crystals is eliminated.
(ii) A powder diffraction pattern can be recorded very rapidly and the technique is non-destructive.
(iii) With special equipment very small samples may be used (1-2 mg).
(iv) A powder diffraction pattern may be used as a fingerprint. It is often superior to an infrared
spectrum in this respect.
(v) It can be used for the qualitative, and often the quantitative, determination of the crystalline
components of a powder mixture.
(vi) Powder diffractometry provides an easy and fast method for the detection of crystal
polymorphs (Polymorphs are different crystal forms of the same substance).

17
18
The first thing that strikes us is that the angles θ (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.

(111)

(111)

(111)
(111)
Na+ Cl-

19
Chapter 6: Single Crystal and Data Collection

Solid state reaction: Solids do not usually react together at room temperature over normal
timescales and it is necessary to heat them to much higher temperatures, often 1000 to 5000oC, in
order for reaction to occur at an appreciable rate. Many factors are usually involved in a particular
solid-state reaction. These are mainly:
(i) the area of contact between the reacting solids and hence their surface areas,
(ii) the rate of nucleation of the product phase, and
(iii) the rates of diffusion of ions through the various phases and specially, through the product
phase.

Why are solid state reactions difficult?

Solid state reactions take place with great difficulty and only at high temperatures it is because of,
(i) differences in the structure between reactants and products, and
(ii) the large amount of reorganization that involved in forming the product: bond must be broken,
reformed and atoms must move towards each other for recombination to form desired product.

To explain the mechanism of solid state reaction, let us take an example of the reaction of MgO and
Al2O3, in a 1:1 molar ratio, to form spinel MgAl2O4. Generally, MgAl2O4 has a structure, which
shows some similarities and differences to the structures of both MgO and Al2O3. Both MgO and
MgAl2O4 have a cubic closed packed array of oxides ions. On the other hand, in Al2O3 the oxide
ions are in distorted hexagonal close packing. In both Al2O3 and MgAl2O4, Al3+ ions occupy
octahedral sides, whereas the Mg2+ ions are octahedral in MgO, but tetrahedral in MgAl2O4 spinel.

The reacting ions Mg2+ and Al3+ are normally trapped on their appropriate lattice sites and it is
difficult to go into empty sites. Thus, only at very high temperatures such ions have sufficient
thermal energy to enable them to jump out of their normal lattice sites and diffuse through the
crystals. Nucleation of MgAl2O4 probably involves some reorganization of the oxide ions at the site
of the potential nucleus together with the interchange of Mg2+ and Al3+ ions across the interface
between the MgO and Al2O3 crystals.

Wagner reaction mechanism: The mechanism of reaction between MgO and Al2O3 involving the
counter diffusion of Mg2+ and Al3+ ions through the product layer followed by further reaction at
two reactant-product interfaces. This mechanism is known as Wagner mechanism. The detailed
mechanism is shown in Fig.

1
MgAl2O4 product layer

Heat Mg2+
MgO Al2O3 MgO Al2O3
treatment
Al3+

Original interface New reactant- Original Another new


product interface interface reactant-product
interface
Wagner reaction mechanism

Initially, two crystals MgO and Al2O3 are in intimate contact across one shared face. Reaction only
occurs at contact points and gets nucleation of MgAl2O4. At this stage, there are two reaction
interfaces: one that between MgO and MgAl2O4 and another between MgAl2O4 and Al2O3 layers.
In order for further reaction to occur and the MgAl2O4 layer to grow thicker, counter diffusion of
Mg2+ and Al3+ ions must pass or cross right through the existing MgAl2O4 product layer to the new
reaction interfaces. However, the diffusion rates are extremely slow at normal temperature.

But, after an appropriate heat treatment, the crystals have partially reacted to form a layer of
MgAl2O4 at the interface. In order to maintain charge balance, for every three Mg2+ ions, which
diffuse to the right-hand interface, two Al3+ ions must diffuse to the left-hand interface. The
reactions that occur at these two interfaces may be written as:

(i) Interface MgO/MgAl2O4

2Al3+ - 3Mg2+ + 4MgO → MgAl2O4

(ii) Interface MgAl2O4/Al2O3

3Mg2+ - 2Al3+ + 4Al2O3 → 3MgAl2O4

Overall reaction:

4MgO + 4Al2O3 → 4MgAl2O4

So, it can be seen that reactions (ii) gives three times as much spinel product as reaction (i); and
hence right-hand interface should grow or move at three times faster rate than that of left-hand
interface. This mechanism has been experimentally tested and found to be satisfactory.

Crystal growth: Crystals grow from, among other things, supersaturated solutions, supercooled
melts and vapours. The formation of a crystal may be considered in two steps:

(i) Nucleation: The initial process that occurs in the formation of a crystal from a solution, a liquid,
or a vapour, in which a small number of ions, atoms, or molecules become arranged in a pattern

2
characteristic of a crystalline solid, forming a site upon which additional particles are deposited as
the crystal grows.

e.g. a melt Nucleus in a melt

(ii) Growth of a nucleus to a crystal: As the nucleus attracts further atoms, they take up positions
on its faces in accordance with its three dimensional periodicity. In this way new lattice planes are
formed. The growth of the nucleus, and then of the crystal, is characterized by a parallel
displacement of its faces. The rate of this displacement is called the rate of crystal growth, and is a
characteristic, anisotropic property of a crystal.

Atoms adhere to the nucleus Growth of a new layer on the The formation of a
faces of a nucleus macrocrystal

Methods of crystal growth: Based on the phase transformation process, crystal growth techniques
are classified into three basic methods, melt growth, solution growth and vapour growth. The flow
chart of Fig. gives an overview of various techniques commonly used for crystal growth.

Crystal growth

Melt growth techniques Solution growth methods Vapour phase growth methods
(i) Czochralski (CZ) (i) Growth from aqueous solution (i) Direct synthesis (DS)
(ii) Bridgman (ii) Travelling heater (THM) (ii) Physical vapour transport (PVT)
(iii) Floating zone (FZ) (iii) Solute solution diffusion (SSD) (iii) Chemical vapour transport (CVT)
(iv) Verneuil (iv) Solvent evaporation (SE) (iv) Solid phase reaction
(v) Temperature difference under
controlled vapour pressure (TDM-CVP)
(vi) Hydrothermal synthesis
]

The suitability of a particular technique for a given crystal is guided by material properties like
thermal expansion, melt viscosity, thermal conductivities and chemical stabilities of melt/solid,
requirement of size and quality etc.
3
Large number of the references is available with detailed description of the growth techniques, only
a brief description of few techniques are given below.

Melt growth techniques: Crystal growth from melt is undoubtedly the most popular method for
growing single crystals at large scale. More than half of the technological crystals are nowadays
obtained by this technique, including elemental semiconductors, metals, oxides, halogenides and
chalcogenides.

Czochralski (CZ) technique: This technique was first developed by Jan Czochralski, a Polish
Chemist, in 1918. Particular advantage of this technique lies in the convenience with which growth
conditions viz. temperature gradients, growth orientation and the ambient can be modified to suit
requirements dictated by the material properties.

In this technique, the starting charge, which could be pre-synthesized phase in powder form, or a
stoichiometric mixture of constituents or polycrystalline chunks, is taken in a crucible and melted.
The schematic of Czochralski growth is shown in Fig.

Czochralski method for crystal growth

A seed crystal, mounted on a rod, is placed in contact with the melt at the center of the crucible.
The melt temperature is so adjusted as to establish thermal equilibrium between the rotating seed
and the melt. The seed is thereafter pulled in a controlled way ensuring that pull rate is smaller or
equal to the rate of melt crystallization. As the seed is withdrawn, the melt solidifies on the surface
of the seed to give a rod-shaped crystal in the same crystallographic orientation as the original seed.
The melt and growing crystal are usually rotated counterclockwise during pulling in order to
maintain a constant temperature, melt uniformity, etc. This process is normally performed in an
inert atmosphere, such as argon, and in an inert chamber, such as quartz.

The method is widely used for the growth of crystals of semiconducting materials, Si, Ge, GaAs,
etc. It has also been used to produce laser generator materials such as Ca(NbO3)2 doped with
neodymium.
4
Solution growth methods: Solution growth techniques are often applied to fabricate high-quality
single crystals which cannot be grown from their own melts. Depending on the particular crystal
class aqueous solutions or flux melts (high-temperature solutions) are usually employed.

Hydrothermal synthesis: Hydrothermal synthesis includes the various techniques of crystallizing


substances from high-temperature aqueous solutions at high vapor pressures. Advantages of the
hydrothermal synthesis method comprise the ability to synthesize crystals of substances which are
unstable near the melting point, and the ability to synthesize large crystals of high quality. Also,
materials which have a high vapour pressure near their melting points can also be grown by the
hydrothermal method. Disadvantages of the method include the need of expensive autoclaves and
the inability to monitor crystals in the process of their growth.

The crystal growth is performed in an


apparatus consisting of a steel pressure vessel
called an autoclave, in which the reaction
mixture is placed along with an appropriate
amount of water. A temperature gradient is
maintained between the opposite ends of the
growth chamber. At the hotter end the nutrient
solute dissolves, while at the cooler end it is Equipment for Hydrothermal crystal growth
deposited on a seed crystal, growing the desired crystal.
For the growth of single crystals by hydrothermal methods it is often necessary to add a
minerallizer. A minerallizer is any compound added to the aqueous solution that speeds up its
crystallization. It usually operates by increasing the solubility of the solute through the formation of
soluble species that would not usually be present in the water.

Hydrothermal synthesis is commonly used to grow synthetic quartz, gems and other single crystals
with commercial value.

Vapour phase growth methods: This involves vapour to solid phase transformation. Vapour-grown
crystals can have considerably lower defect densities than crystal grown with other methods.

Chemical vapor transport (CVT) method: A variety of processes of crystal growth proceeds via the
gas phase. To date the chemical vapor transport method developed by H. Schäfer (1971) to be an
important and versatile preparative method of solid state chemistry.

Basically, the method consists of a tube, usually of silica glass, which contains the reactant(s), A, at
one end and which is sealed, either under vacuum or more usually with an atmosphere of a gaseous
transporting agent, B. The tube is placed inside a furnace such that a temperature gradient exists
5
inside the tube; there is a temperature change of 50oC along the length of the tube. Materials A and
B react together to form gaseous AB which subsequently decomposes at the other end of the tube to
redeposit crystals of A (Fig.).

The transport is caused by different temperatures and


Source Glass tube Sink
therefore changed equilibrium position in source and
sink. It is common to characterize the volatilization A B (gas)
(source) and the deposition temperature (sink) with T 1
T2 T1
and T2, respectively, T1 representing the lower
temperature. The transport direction results from the A A B (gas) A
sign of the reaction enthalpy of the transport
Equilibrium: A(s) + B(g) ⇌ AB(g)
reaction based on Le Chatelier’s principle.
Therefore, exothermic transport reactions always Single vapour phase transport experiment for
transport to the zone of higher temperature - from the growth of single crystals of substance A.
T1 to T2 (T1 → T2), endothermic reactions transfer the solid to the cooler zone - from T2 to T1
(T2 → T1).

This method may be used for the synthesis of new compounds, for the growth of single crystals and
for the purification of a compound. Pure and crystalline species of various solids could be made with
the help of chemical transport reactions: metals, metalloid, and intermetallic phases as well as halides,
chalcogen halides, chalcogens, pnictides and many others.

Graphite intercalation compounds (GICs): Graphite intercalation compounds are complex


materials having a formula CXm where the ion Xn+ or Xn− is inserted (intercalated) between the
oppositely charged carbon layers. Typically m is much less than 1. These materials are deeply
colored solids that exhibit a range of electrical and redox properties of potential applications.

Structurally, intercalation compounds of graphite can be grouped in three categories:


(i) Covalent compounds, which result from the attack of strong oxidizing agents such as nitric acid,
fluorine, or permanganate on graphite. Then, the aromatic planarity of the graphite is destroyed and
a buckled sp3-hybridized layer is formed.
(ii) Lamellar compounds, generated by the attack of moderately strong oxidants or reductants. The
aromaticity of graphite is maintained, and more importantly the conductivity is enhanced.
(iii) Residue compounds, arising from the decomposition of lamellar compounds by thermolysis or
in vacuo treatment.

6
Most of the intercalation reactions are reversible, when the structure and planarity of the carbon
layers are essentially unaffected by intercalation. According to the nature of intercalants and the
reversibility of the process, a chemical classification of GIC has also been suggested.

(i) The intercalant participates with the graphite in an electron exchange process, and the
intercalation is formally reversible. Alkali metals or bromine are representative examples of such
species.
Partial
Graphite + K → C8K → C24K → C36K → C48K → C60K
vacuum
Graphite + Br2 → C8Br

The electronic structure of graphite is modified by intercalation of potassium atoms since partial
electron transfer from potassium takes place, and the resulting polar structure may be represented as
C8-K+. Thus, in C8K, the graphite behaves as an electron acceptor. In contrast, in C8Br, graphite
acts as an electron donor to the halogen and the ionic formula C8+Br- may be attributed to graphite
bromide.

(ii) The intercalation occurs with an irreversible loss of part of the intercalating agent. Hence the
reaction of graphite with sulfuric acid gives the blue graphite.
Electrolysis
Graphite + H2SO4 → C24+(HSO4)-.2H2SO4 + H2
Blue graphite
(iii) The intercalating agent forms one or more true covalent bonds with the graphite. Examples are
provided by the formation of graphite oxide or graphite fluorides.
25oC
Graphite + HF/F2 → C3.6F to C4.0F
Black graphite fluoride

450oC
Graphite + HF/F2 → CF0.68 to CF
White graphite fluoride

Among the lamellar derivatives of graphite, alkali-graphite intercalation compounds have been
widely studied. In these compounds, alkali metals are sandwiched between pairs of carbon rings. If
all such sites were occupied, the stoichiometry C2M would result, but in the usual formula C8M
only one quarter is occupied, in an ordered fashion. Fig. shows this disposition in which the
graphite layers are superposed in projection and the alkali atoms are intercalated between the
carbon rings.

7
Typical structure of potassium graphite, C8K

Potassium-graphite having the composition C8K is the most commonly used alkali-graphite
intercalation compound. In this substance all carbon layers are separated by a layer of potassium
atoms. Several stages are also possible and the well-defined intercalation compounds C12nM (stage
n > 1) can be formed. Thus, in C24K, C36K, C48K and C60K the alkali metal is intercalated in each
second, third, fourth, and fifth interlayer spacing, respectively.

Bronze Steel blue Blue Black Black


C8 M C24M, n = 2 C36M, n = 3 C48M, n = 4 C60M, n = 5

Layer arrangements and stoichiometries

The stoichiometry C8M is observed for M = K, Rb and Cs. For smaller ions M = Li+, Sr2+, Ba2+,
Eu2+, Yb3+, and Ca2+, the limiting stoichiometry is C6M.

The alkali metal-graphites are extremely reactive, pyrophoric, ignite in air, and are very sensitive to
moisture, reacting explosively with water. The compounds, however, are stable in dry ethereal
solvents (THF, DME, or diethyl ether) or pentane under argon for at least 24 hour at room
temperature. The electrical conductivity of these compounds is greater than that of α-graphite. C8K
is a superconductor with a very low critical temperature Tc = 0.14 K.

Preparation of potassium graphite: Fresh, clean pieces of potassium (5.0 g) are added to finely
divided graphite (12.5 g) with stirring for 20-30 min under argon at 150-160oC. The reaction is very
exothermic, and the potassium is added at a slow enough rate to ensure complete reaction before
addition of more metal. When the potassium melts, the mixture is vigorously stirred for 10-15 min
to afford C8K as a fine bronze-coloured powder.

8
Additionally, C8K, C24K, C8Rb, and C8Cs can also be generated at room temperature using metal
transfer reactions in hydrocarbon solvents.

Preparation of thin films: A thin film is a layer of material ranging from fractions of a nanometer
to several micrometers in thickness. The controlled synthesis of materials as thin films is a
fundamental step in many applications. Advances in thin film deposition techniques during the 20th
century have enabled a wide range of technological breakthroughs in areas such as magnetic
recording media, electronic semiconductor devices, LEDs, optical coatings, hard coatings on
cutting tools, and for both energy generation (e.g. thin film solar cells) and storage (thin-film
batteries). It is also being applied to pharmaceuticals, via thin-film drug delivery.

Various methods are used to prepare thin films; they fall into two main groups, chemical (including
electrochemical methods) and physical methods. In both cases, the techniques are based in the
formation of vapor of the material to be deposited, so that the vapor is condensed on the substrate
surface as a thin film. Usually the process must be performed in vacuum or in controlled
atmosphere, to avoid interaction between vapor and air.

Among chemical and electrochemical methods the most important are electrolytic deposition,
electroless deposition, anodic oxidation and chemical vapor deposition. On the other hand, the most
important physical methods for the preparation of thin film are cathode sputtering and vacuum
evaporation.

Cathod sputtering: The apparatus used for this technique is outlined in Fig. Basically, it consists of
a bell jar which contains a reduced pressure 10-1 to 10-2 torr of an inert gas, argon or xenon. The gas
is subjected to a potential drop of several - +
v
kilovolts. A glow discharge is thereby created in
the inert gas from which positive ions are
accelerated towards the cathode (target). These Cathode
Inert gas
high energy ions remove material from the cathode inlet
Ar+
which then condenses on the surroundings, Substrate
(Anode)
including the substrates to be coated, which are
placed in a suitable position relative to the cathode.
The mechanism of sputtering, or removal of
material from the cathode, appears to involve the Exhaust

transfer of momentum from the gaseous ions to the Cathode sputtering equipment for thin film
cathode in such a way that atoms or ions are then deposition
ejected from the cathode.

9
Single crystal: A single crystal or monocrystalline solid is a material in which the crystal lattice of
the entire sample is continuous and unbroken to the edges of the sample, with no grain boundaries.
The absence of the defects associated with grain boundaries can give monocrystals unique
properties, particularly mechanical, optical and electrical, which can also be anisotropic, depending
on the type of crystallographic structure.

Diffractometry: A single-crystal diffractometer consists of an X-ray source, an X-ray detector, a


goniostat that orients the crystal so that a chosen X-ray diffracted beam can be received by the
detector and a computer that controls goniostat and detector movements and performs the
mathematical operations required positioning the crystal and detector in the desired orientations.
The detector is usually of the scintillation counter type.

The modern diffractometer is often called a four-circle diffractometer (Fig.) because it has four arcs
ω, χ, φ and 2θ that can be used to adjust the orientation of the crystal and counter so as to bring any
desired plane into a reflecting position and detect this reflection. These are normally divided into
the crystal orienter, which contains the φ and χ circles, and the base, which contains ω and 2θ. The
two base circles, ω and 2θ, are mounted about a common axis and can either be adjusted
independently or be geared together so that their movements are correlated. Although φ, ω and 2θ
are true circles, χ can be either a full circle or a segment of one with a setting range only a little
more than 90o. X-ray diffractometer is built on the basis of the proposition of Eulerian cradle.

Fig. Typical four-circle diffractometer

Single crystal X-ray diffraction: Single-crystal X-ray diffraction is a non-destructive analytical


technique which provides detailed information about the internal lattice of crystalline substances,
including unit cell dimensions, bond-lengths, bond-angles, and details of site-ordering. Directly
related is single-crystal refinement, where the data generated from the X-ray analysis is interpreted
and refined to obtain the crystal structure.

10
The samples are unfractured and optically clear single crystals. Their size should be between 0.1
and 0.2 millimeters in the three directions of space. They are normally selected using an optical
microscope (x 40) equipped with a polarizing attachment and observing if light extinguishes
regularly every 90º when turning the stage of the microscope.

A selected crystal is fixed on the tip of a thin glass fiber using epoxy or Crystal
cement, or in a loop including specific oil, which fits into the goniometer head Glass fiber

in the diffractometer. The crystal is then aligned along the beam direction.
Brass pin

Crystals can be sensitive to light, air or moisture, or susceptible to loss of Crystal


crystallization solvent. If so, a special treatment is required. For example, they Glass capillary
can be mounted inside sealed glass capillaries or the data collection can be
Brass pin
performed at low temperature.

Once the crystal is mounted on the diffractometer, a preliminary rotational image is often collected
to screen the sample quality and to select parameters for later steps. An automatic collection routine
can then be used to collect a preliminary set of frames for determination of the unit cell. Reflections
from these frames are auto-indexed to select the reduced primitive cell and calculate the orientation
matrix which relates the unit cell to the actual crystal position within the beam. The primitive unit
cell is refined using least-squares and then converted to the appropriate crystal system and Bravias
lattice. This new cell is also refined using least-squares to determine the final orientation matrix for
the sample.

After the refined cell and orientation matrix have been determined, intensity data is collected. Data
is typically collected between 4° and 60° 2θ for molybdenum radiation. A complete data collection
may require anywhere between 6-24 hours, depending on the specimen and the diffractometer.
Exposure times of 10-30 seconds per frame for a hemisphere of data will require total run times of
6-13 hours. Older diffractometers with non-CCD detectors may require 4-5 days for a complete
collection run.

After the data have been collected, corrections for instrumental factors, polarization effects, X-ray
absorption and (potentially) crystal decomposition must be applied to the entire data set. This
integration process also reduces the raw frame data to a smaller set of individual integrated
intensities.

The intensity of X-rays in a diffraction pattern depending only upon the crystal structure is referred
to as called the structure factor:
N
F(hkl) = ∑ fj [exp 2πi(hxj + kyj + lzj)]
j=1
11
where h, k and l are the indices of the diffraction planes (Bragg reflections), N is the number of
atoms in the cell and (xj, yj, zj) are the fractional coordinates of the jth atom with scattering factor
fj. Each structure factor represents a diffracted beam which has an amplitude |F(hkl)| and a relative
phase φ(hkl). The crystal structure can be obtained from the diffraction pattern if the electron
density function is calculated at every point in a single unit cell:
1
ρ(x, y, z) = ---- ∑ ∑ ∑ |F(hkl)| cos[2π(hxj + kyj + lzj) – φ(hkl)]
V h k l
where the summation is over all values of h, k and l and V is the volume of the unit cell. Since
X-rays are diffracted from the whole crystal, the calculation yields the contents of the unit cell
averaged over the whole crystal. In practice, the calculation of the electron density produces maps.
The maxima on these maps represent the position of the atoms in the cell.

The structure factors are reciprocal space vectors whereas the electron density is from the real
space. The diffraction pattern is the Fourier transform of the electron density and the electron
density is the inverse Fourier transform of the diffraction pattern. The measured intensities of a
diffraction pattern enable the determination of only the structure factor amplitudes but not their
phases. The calculation of the electron density is not then obtained directly from experimental
measurements and the phases must be obtained by other methods. This is the so called phase
problem. The most usual methods to overcome the phase problem are direct methods and methods
based on the Patterson function. The former are the most important in chemical crystallography and
the latter are currently applied when some heavy atoms are present. The phases are obtained
approximately and have to be improved. With the calculated phases and structure factors
amplitudes, a first electron density map is calculated, also approximate, from which the atomic
positions will be obtained.

The next step is the completion of the structure by Fourier synthesis and refinement of the
structural parameters to optimize the fitting between the observed and calculated intensities in the
diffraction pattern. The refinement cycles include positional atomic parameters and anisotropic
vibration parameters. Finally, the hydrogen atom positions, if present, are determined or calculated.
The structural refinement is evaluated from the agreement between the calculated and the measured
structure factors. The refinement is considered finished when the following essential conditions are
fulfilled:
(i) The agreement factors are small enough.
(ii) The structural model is chemically appropriate.
(iii) The estimated standard deviations of all geometrical parameters are as small as possible.
(iv) The peaks remaining in the electron density map are as small as possible.

12
Once the structure is determined and refined, several geometrical parameters such as bond lengths,
bond angles, torsion angles, π-stakings and hydrogen-bond are evaluated and appropriate tables and
graphics representing the structure are prepared. A standard file (CIF: crystal information file)
containing all the information of the structure is created and can be used to evaluate their quality
and possible problems.

The steps involved in a crystal structure determination are summarized in the flow chart.

Selection of a
suitable crystal

Mounting a crystal

Crystal system and


unit cell dimensions

Bravais lattice selection

Full data set


collection

Data reduction

Space group selection

Construction of an
electron density map

Location of
atom positions

Structure refinement

Preparation of figures, tables, etc.

Interpretion and
communication of results

13
Chapter 7: Structures of Solids

Close packing: A close packing is a way of arranging equidimensional objects in space so that the
available space is filled very effectively. Such an arrangement is achieved when each object is in
actual contact with the maximum number of like objects.

In metallic crystals, the atoms are assumed to be hard, incompressible spheres with identical size.
They then packed in such a way that energy is minimum in closest packings and that the spheres
have maximum number of nearest neighbors. If the constituents are not spherical, or there are
several sizes of atoms in the basis, or if the hard sphere approximation are not valid, less closely
packed structures can result. The name close packed refers to the packing efficiency of 74.05 %.
The two most efficient packing arrangements are the hexagonal closest-packed (hcp) structure and
the cubic closest-packed (ccp) structure.

Close-packed structures: Structures having least waste of space – with each sphere having its
geometrically maximal number of neighbors. The most efficient way of packing spheres in two
dimensions is as shown in Fig. 1. Each sphere, e.g., A, is surrounded by, and is in contact with, six
others. By regular repetition, infinite sheets called close packed layers are formed. The coordination
number of six is the maximum possible for a planar arrangement of contacting equal sized spheres;
while in a non close-packed layer, each sphere has a coordination number of four only (Fig. 2).

1 1
6 2
A 4 A 2
5 3
4 3

Fig. 1. Two-dimensional close packing Fig. 2. Two-dimensional non-close packing

The most efficient way of packing spheres in three dimensions is to stack close packed layers on
top of each other to give close packed structures. There are two simple ways in which this may be
done, resulting in hexagonal close-packed (hcp) and cubic close-packed (ccp) structures. These are
derived as follows: the most efficient way for two close packed layers A and B to be in contact is
for each sphere of one layer to rest in a hollow between three spheres in the other layer (Fig. 3).
Any B (blue-violet) sphere is, therefore, seated between three A (gray solid) spheres, and vice
versa.
The third layer can be placed upon the second in two different ways such that:

1
(i) Each sphere in the third layer is directly above a corresponding sphere in the first layer, then the
following sequence arises
...ABABABAB…
This is known as hexagonal close packing of spheres and is abbreviated as hcp (Fig. 4).

A A

A
A
A B
A B
B
B
A B A A
A A

Fig. 3. Close packing of spheres Fig. 4 (a). Hexagonal close Fig. 4 (b). Composite view of
in two layers. packing (ABA..) of spheres. hexagonal close packing (ABA..)
of spheres.

(ii) The third layer of spheres rests in the other set of hollows identified in Fig. 5, so that it is not
directly above either of the preceding layers; if the position of the third layer is called C, then a
sequence
…ABCABCABC….
arises. This is called cubic close packing (ccp) of spheres (Fig. 6). The unit cell of this structure is a
cube, which contains 14 spheres. An closer study of ABC ABC… arrangement shows that, in
addition to the 8 spheres placed at the 8 corners of a cube, there is also one sphere at the center of
each of the six faces of the cube and hence the structure is also called face-centered cubic (fcc)
structure.

A B C
A
B
B A A B
A C B
B
C A B A B
A
C C
A A B
B A
A A
A C A

Fig. 5. Cubic close packing of Fig. 6 (a). Cubic close Fig. 6 (b). Composite view
spheres in three layers. packing (ABCA..) of of cubic close packing
spheres. (ABCA..) of spheres.

The unit cells of hexagonal close-packed and cubic close-packed structures are given in Figs. 4(b)
and 6(b).

In both close-packed arrangements every sphere has a coordination number of twelve, namely six
spheres arranged about it in the form of a planar hexagon, three more spheres above and three

2
below, as in Fig. 7. Both types of closest packing are equally economical of space, 74.05% being
occupied.

(a) (b)
Fig. 7. (a) Hexagonal close packed (hcp) structure and (b) Cubic close packed (ccp) structure
Here is another way of looking at this difference:

Octahedral and tetrahedral sites in fcc/ccp structure O

O
O-sites: Centre of edges and body center ⇒ 4 O-sites O

O
T T
T-sites: Division of cell into 8 minicubes. Each minicube has 4 T
O
T
O

anions, which make up the tetrahedron. T-sites are at the center O


T T
O O
of each minicube ⇒ overall 8 T sites (4 T+ and 4 T- sites). T
O
T

O O
Thus there are twice as many T sites as O sites i.e. O-sites: N and
O
T-sites: 2N (with N being the number of formula units per cell; O = Octahedral site
N(fcc) = 4). T = Tetrahedral site

3
Packing fraction or packing efficiency (P.E.): The packing efficiency is the fraction of the unit
cell actually occupied by the spheres (atoms that are usually spherical). It must always be less than
100% because it is impossible to pack spheres without having some empty space between them.
Volume occupied by all the spheres in unit cell
Packing efficiency = ----------------------------------------------------------- × 100
Total volume of the unit cell

Calculation of packing efficiency


(i) In hcp structure:
The hcp unit cell has 6 atoms. The volume of 6 atoms = (6 × 4/3 πr3) = 8 πr3
Let a be the side length of its base and c be its height. Then,
a = 2r and c/a = 1.633
2√2 4√2 Atoms per unit cell =
∴ c = ------ a = ------ r (r = the radius of the spherical atom) ¹/6×12 + ¹/2×2 + 3 = 6
√3 √3
∴ The volume of the hexagonal unit cell = 3 × a2c sin120o
4√2 √3
= 3 × (2r)2 × ------ (r) × ----- = 24√2 r3
√3 2

8πr3 √2 π
∴ Relative density of packing in hcp unit cell = --------- = ------ = 0.7405
24√2 r3 6
∴ Packing efficiency of a hcp structure = 0.7405 × 100 = 74.05%

[Calculation of c/a ratio for an ideal hexagonal close packed structure: Let c be the height of
the unit cell and a be its edge. The three body atoms lie in a horizontal plane at c/2 from the ortho-
centers of alternate equilateral triangles at the top or base of the hexagonal cell. These three atoms
just rest on the three atoms at the corners of the triangles.

In ΔABY,
AY
cos 30o = ------
AB
a√3
o
AY = a cos30 = -------
2
(AZ)2 = (AX)2 + (ZX)2 ………………(i)
In ΔAXZ,
2 2 a√3 a Bottom of hcp structure
AX = --- AY = --- ------ = ----- ………(ii)
3 3 2 √3

4
c
ZX = ---- …………………………….(iii)
2
From (i), (ii) and (iii), we get
a2 c2
a2 = ---- + -----
3 4
c2 a2
2
----- = a - -----
4 3
c2 8
----- = ----
a2 3
c √8 2√2
----- = ---- = ------]
a √3 √3

(ii) In ccp structure:


The number of atoms present in a face-centered cubic unit cell is 4
The volume of 4 atoms = (4 × 4/3 πr3)
In a face-centered cubic unit cell r (the radius of the spherical atom) and a
(the edge length of the cube) are related as: a = 4r/√2

4r a
Volume of the cube = a3 = (-----)3 r
√2
(4 × 4/3 πr3) √2 π 2r
a
2a
∴ Relative density of packing in fcc cubic unit cell = -------------- = ------
r
4r 6
3
(----)
2 a = 4r
√2
= 0.7405
∴ Packing efficiency of a ccp structure = 0.7405 × 100 = 74.05%

Void space: A vacant space not occupied by the constituent atoms (spheres) in the unit cell is
called void space.
% void space = 100 – packing efficiency

5
Comparison between hexagonal close packed (hcp) and cubic close packed (ccp) structures
Hexagonal close packed (hcp) structure Cubic close packed (ccp) structure

1. In hcp, the top and bottom three are directly 1. In ccp, they are staggered.
above one another.
2. Layer ordering may be described as 2. Layer ordering may be described as
ABABA. ABCABC.
3. In the hcp, there are never any atoms over 3. These tiny channels are absent in ccp.
the right-pointing holes. Thus, if we look
directly down on the structure we can see tiny
channels down throughout the hcp structure.
4. The hcp structures are less ductile but 4. The ccp structures are more likely to be
stronger. ductile than hcp.
5. Each sphere has twelve nearest-neighbors. 5. Each sphere has twelve nearest-neighbors.
6. Packing efficiency is 74.05% 6. Packing efficiency is 74.05%
7. The arrangement contains tetrahedral and 7. The arrangement contains tetrahedral and
octahedral interstitial sites. octahedral interstitial sites.

Hexagonal close packed (hcp) structure Cubic close packed (ccp) structure

There is another possible arrangement of packing of spheres called body centered cubic (bcc)
arrangement which will be only formed when the spheres in the first layer (marked A) of cubic
closest packing undergo slight opening. Due to this, none of these will remain in contact with each
other. The second layer of the spheres (marked B) may be built up on top of the first layer in such a

6
way that each sphere of the second layer has been found to be in contact with four spheres of the
layers below it. If the successive building of the other layer is carried on the third layer will be
same as the first layer i.e., on the top of A. If this pattern of building layers is repeated infinitely, an
arrangement shown in Fig. 9 would be obtained.

A A A
B B B B
A A A A A
B B
A A A

Arrangement of spheres in three layers Unit cell of body-centered Body-centered cubic


to get body-centered cubic structure. cubic structure. arrangement
Fig. 9. Body centered cubic structure

Only 68.0% of the total volume is actually occupied by the spheres in this arrangement, where one
sphere is present at each corner and also at the center of each unit cube. Hence, each sphere has
been found to be in contact with eight other spheres so that the coordination number in this type of
arrangement would be eight.

Positions inside the cube are expressed by fractional numbers. Thus, the bcc structure has atoms at
positions (0,0,0). (1,0, 0), (0,1,0), (0,0,1), (1,1,0), (0,1,1), (1,0,1), (1,1,1) and (½,½,½).

Packing efficiency for a body centered cubic (bcc) structure

The number of atoms present in a body centered cubic unit cell is 2


The volume of two atoms = 2 × 4/3 πr3 = 8/3 πr3
Atoms along the body diagonal touch each other. The length of the
diagonal is √3 a. Thus in a body-centered cubic unit cell r (the radius
of the spherical atom) and a (the edge length of the cube) are related
as: a = 4r/√3 Atoms per unit cell = (⅛×8) + 1
=2
4r
Volume of the cube = a3 = (-----)3
√3
(8/3 πr3)
∴ Relative density of packing in bcc cubic unit cell = ----------
4r
(----)3
√3
√3 π
= ------ = 0.6802
8
∴ Packing efficiency of a bcc structure = 0.6802 × 100 = 68.02%

7
Problem 1: Calculate the efficiency of packing (%) in case of a metal crystal for simple cubic.

Solution: In a simple cubic lattice the atoms are located only on the corners of the cube.
Let a = edge length of the cube
Let r = radius of each spheres
The relation between radius and edge, a = 2r
The volume of the cubic unit cell = a3 = (2r)3 = 8r3
Number of atoms in unit cell = 8 × 1/8 = 1
e Volume of the occupied space = (4/3)πr3
Volume of one atom (4/3)πr3 π
Packing efficiency = -------------------------------- = ---------- × 100 = ----- × 100 = 52.36%
Volume of cubic unit cell 8r3 6

Metal crystal structure periodic table

8
Problem 2: A crystal adopts bcc type arrangement. If the edge length of the unit cell is 288.5 pm
and density is 7.2 g cm-3. What will be the value of Avogadro’s number if molar mass of the crystal
is 52 g mol-1.

Solution:
Z×M
d = ---------- [d = density, Z = the number of atoms per unit cell]
N × a3
Z×M
N = ---------
a3 × d
2 × 52 g mol-1
N = ---------------------------------------- = 6.015 × 1023 mol-1
(288.5 × 10-10 cm)3 × 7.2 g cm-3

Problem 3: Platinum has a density of 21.5 g cm-3 and a molar mass of 195.1. How many platinum
atoms are in a unit cell, if the metal crystallizes in the cubic system with a unit cell edge length of
391.4 pm?

Solution: Let Z = the number of atoms per unit cell. The density of the unit cell equals the mass of
the unit cell divided by the volume of the unit cell.
Z × 195.1 g mol-1
-3
21.5 g cm = ---------------------------------------------------
(6.022 × 1023 mol-1) × (391.4 × 10-10 cm)3
∴ Z = 3.98 ≈ 4

Thus the structure is face centered cubic, with 4 atoms per unit cell.

Problem 4: What is the fractional volume change in iron as it transforms from bcc to fcc at 910oC?
Assume that a(bcc) = 0.2910 nm and a(fcc) = 0.3647 nm.

Solution: The volume (V) per atom is equal to volume per unit cell / atoms per unit cell.
For bcc Fe, V(bcc) per atom = a3(bcc)/2. Similarly for fcc Fe, V(fcc) per atom = a3(fcc)/4.
Substituting, V(bcc)/atom = (0.2910)3/2 = 0.01232 nm3; V(fcc)/atom = (0.3647)3/4 = 0.01213 nm3.
Therefore, ∆V/V = (0.01213 - 0.01232)/0.01232 = - 0.0154 or -1.54%, a result consistent with the
abrupt contraction observed when bcc Fe transforms to fcc Fe.

Interstitial sites: The packing fraction in both fcc and hcp cells is 74.05%, leaving 25.95% of the
volume unfilled. The unfilled lattice sites (interstices) between the atoms in a cell are called
interstitial sites or vacancies. The shape and relative size of these sites is important in controlling
the position of additional atoms. In both fcc and hcp cells most of the space within these atoms lies
within two different sites known as tetrahedral sites and octahedral sites.

9
Tetrahedral sites (T sites): Any sphere resting in the hollow formed by three spheres in an adjacent
layer forms a tetrahedral site; the centers of the four spheres are at the vertices of a regular
tetrahedron (Fig. 10).

Tetrahedral site

A A

Fig. 10. Tetrahedral sites

The size of the site is considerably less than that of the surrounding spheres: r/R = 0.225, where R =
radius of the spheres forming the closest packed layers, r = radius of the largest sphere which can
fit into a tetrahedral site without distorting the structure. This is why a tetrahedral hole created by
ions of radius R cannot accommodate any ion with radius greater than 0.225 R. Two tetrahedral
sites are associated with each atom in a close packed layer, one above and one below the packing
atom.
D
C
T-
T+

B
A

T-sites: It lies between 4 spheres. T+ (up) and T- (down) refer to the orientation of arrow indicated
above: T+ : (1 1 1) direction; T-: (͞1 ͞1 ͞1) direction

Octahedral sites (O sites): These sites are found at the center of regular octahedron, the vertices of
which are defined by two sets of three spheres in adjacent layers. They are illustrating in Fig. 11.
This site is formed in both hcp and ccp systems. An octahedral site is considerably larger than a
tetrahedral site, and has r/R = 0.414. Thus an octahedral hole cannot accommodate any ion larger
than 0.414 R. There is only one octahedral site for an atom in the unit cell.

Octahedral site

A
B B

A A

Fig. 11. Octahedral sites


10
Note that there are 8 tetrahedral sites and 4 octahedral sites in the ccp structure. Coordinates of
tetrahedral sites are:
T+ sites: ¾ ¼ ¼, ¼ ¾ ¼, ¼ ¼ ¾, ¾ ¾ ¾; T- sites: ¼ ¼ ¼, ¾ ¾ ¼, ¼ ¾ ¾, ¾ ¼ ¾

While the coordinates of octahedral sites are: ½ 0 0, 0 ½ 0, 0 0 ½ , ½ ½ ½

On the other hand, the hcp structure has 12 tetrahedral sites and 6 octahedral sites available.

Structure of ionic solids: Packing arrangements other than hcp, ccp, and bcc become possible if
the sizes of the atoms are appreciably different as in compounds. Ionic solids are built up from ions
of opposite charge, usually different in size. The packing of ions of different size can be understood
in most cases by considering staking of the larger ions in hexagonal or cubic close packed
structures in which the smaller ions (usually the cations) occupy tetrahedral or octahedral sites in
required numbers to maintain stoichiometry and electrical neutrality. The relative sizes of the ions
are thus a key factor in determining the structure of an ionic crystal.

The figure below shows some of the most common structures (fluorite, halite, zinc blende) as well
as a rather rare one (Li3Bi) that derive from the fcc lattice. From the hcp lattice, we can make the
NiAs and wurzite structures, which are the hexagonal relatives of NaCl and zinc blende,
respectively.

Fig. 12. Structures derived from the fcc lattice

Radius ratio (Rr) and Coordination number (C.N.): The ratio of radius of cation to that of
anion is called the radius ratio. Mathematically, Rr = rc+/ra- , where, Rr is called the radius ratio, and
rc+ and ra- are the ionic radii of the cation and anion of the ionic crystal c+a- respectively.

The coordination number of cation and shape of an ionic crystal is determined by the radius ratio of
cation and anions constituting the crystal. This is called the radius ratio rule. The radius ratio at

11
which the anions just touch one another as well as the central cation is called the limiting radius
ratio. If the radius ratio of an ionic crystal falls below the limiting radius ratio, then the structure is
unstable (Fig. 13).

Anion Cation Anion Cation Anion Cation

Stable structure, Limiting case of stable Unstable structure,


r+/r- = 0.414 – 0.732 structure, r+/r- = 0.155 – 0.225 r+/r- = 0.000 – 0.155

Fig. 13. Stability of c+a- ionic crystal in which the cation c+ has a coordination number equal to 3.

The radius ratio values for various coordination numbers and their corresponding shapes are given
in Table 1.

Table: Radius ratio value for different coordination numbers of a cation.

Radius ratio Diagram Maximum coordination Shape of the ionic crystal


value, Rr number of cation
0 – 0.155 2 Linear

0.155 – 0.225 3 Trigonal planar

0.225 – 0.414 4 Tetrahedral

0.414 – 0.732 4 Square planar

0.414 – 0.732 6 Octahedral

0.732 – 1.000 8 Body-centered cubic (bcc)

1.000 12 Cubic close packed (ccp) and


hexagonal close packed
(hcp) structure
[Note: The radius ratio gives an indication of the likely coordination number of a compound; the
higher the ratio, the greater the coordination number.]

Calculation of limiting radius ratio (r+/r-) for two common geometries: The limiting radius ratio
(r+/r-) values for the cations of coordination number 2, 3, 4, 6, 8 and 12 can be calculated. However,
calculations of (r+/r-) values for the tetrahedral, octahedral and body-centered cubic type
geometrical arrangements are given below:

12
(i) Limiting radius ratio in a tetrahedral lattice (C.N. = 4): To provide the tetrahedron
coordination geometry around the cation, the cation is placed at the center of a cube, and four
anions are placed at the four alternate corners of the cube. At the limiting condition, the two anions
mutually touch along the face diagonal while each anion touches the cation along the body
diagonal. Thus in the limiting condition, we get:
length of the face diagonal of the cube = 2r-, i.e. √2 a = 2r-
Anion
and, ½ of the length of the body diagonal = r+ + r-
2a
i.e. ½(√3 a) = r+ + r-
where, a = edge length of the cube. + Cation
3a
Thus from the above equations we get:
r+ + r- √3 a
---------- = --------
2r- 2√2 a
r+ + r- √3
---------- = ------
r- √2
r+ √6
or ----- = ------ - 1 = 1.225 – 1 = 0.225
r- 2

(ii) Limiting radius ratio in an octahedral lattice (C.N. = 6): The limiting situation in an
octahedral lattice is shown in Fig. 1 where four anions are placed at the corners of the square and
the cation is placed at the center. One anion remains below the plane another anion remains above
the plane along with the axis (C4) passing perpendicularly to the square plane through the centre of
the square (these two axial anions are not shown in the figure). At the limiting condition we get:
edge length of the square = 2r-, i.e. a = 2r-
and, length of the diagonal = 2(r+ + r-) = √2 a
where, a = edge length of the square. a
+ Cation
From the above equations, we get:
2(r+ + r-) √2 a Anion
------------ = -------
2r- a A cross section through an octahedral site
r+
or ----- = √2 – 1 = 1.414 – 1 = 0.414
r-

(iii) Limiting radius ratio in a body-centered cubic (bcc) lattice (C.N. = 8): In a body-centered
type cubic lattice, the anions are at the corners of the cubes and the cation is placed at the center of
each cube. In the limiting condition (Fig.), the spherical anions will just mutually touch each other

13
along the edge; and along the body diagonal the cation will remain just in contact with the two
anions on opposite sides. Thus in the condition we have:
edge length of the cube = 2r-, i.e., a = 2r-
and length of the body diagonal of the cube = 2(r+ + r-) = √3 a
Anion
where a = edge length of the cube 2a
Thus a
+ Cation
2(r+ + r-) √3 a 3a
----------- = -------
2r- a
r+
----- = √3 – 1 = 0.732
r-

Note that the correlation between structure and radius ratio of ionic compounds is based on the
assumption that maximum stability is attained when every ion is surrounded by maximum number
of ions of opposite charge. However, there is a practical limitation of applying this assumption to
high coordination number. Let us consider salts of the type MX: the coordination numbers of Mn+
and Xn- ions must be equal. As the coordination number increases (radius increases) more and more
Xn- ion may be packed around each Mn+ ion; but then each such Xn- should also have as many Mn+
ions surrounding them. So the gain through electrostatic attraction will be largely cancelled by an
increase in electrostatic repulsion. Thus the radius ratio provides a rough guide to crystal structure
from a consideration of the immediate neighbourhood of a particular ion, but it does not take
account of the details of a three dimensional arrangement.

Coordination number of an ion: In an ionic crystal, the arrangement of ions is such that there is
more electrostatic attraction between the oppositely charged ions than there is electrostatic
repulsion between the similarly charged ions. It is for this reason that each ion in an ionic crystal
prefers the greatest possible number of oppositely charged ions in its surroundings as its neighbors.
The number of nearest oppositely charged ions by which a given ion is surrounded is called the
coordination number of the given ion (Fig. 14). This is denoted by C.N.

Anion Anion
Anion
Cation Cation

Cation

Cation surrounded tetrahedrally Cation surrounded octahedrally by Cation surrounded by 8


by 4 anions (C.N. of cation = 4) 6 anions (C.N. of cation = 6) anions (C.N. of cation = 8)

Fig. 14. Concept of coordination number of 4, 6 and 8

14
Comparison of cell properties of some crystal structure
Crystal sc bcc fcc hcp
specification
Coordination
6 8 12 12
numbers
Nearest neighbor a√3 a√2
2r = a 2r = ------ 2r = ------ 2r = a
distance
2 2
Lattice constant a = 2r a = 4r/√3 a = 4r/√2 a = 2r
Number of atoms
n = ⅛×8 = 1 n = ⅛×8 + 1 = 2 n = ⅛×8 + 3 = 4 n=6
per unit cell
Number of lattice
1 2 4 6
points
Volume of all the
4 4 4 4
atoms in a unit V = 1 × ---- πr3 V = 2 × ---- πr3 V = 4 × ---- πr3 V = 6 × ---- πr3
cell 3 3 3 3

Volume of unit 4r 3 4r 3
V = a3 = (2r)3 V = a3 = ---- V = a3 = ---- V = a3 = (2r)3
cell
√3 √2
Atomic packing π √3 π √2 π √2 π
fraction ---- = 0.52 ------ = 0.68 ------ = 0.74 ------ = 0.74
6 8 6 6

Structures of ionic crystals of AB type: These are of five main AB structure types: NaCl, CsCl,
ZnS (zinc blende), ZnS (wurtzite) and NiAs, each of which is found in a large number of
compounds.

Sodium chloride (NaCl): As the radius of Na+ ion is 0.95 Å and that of Cl- ion is 1.81 Å, the radius
ratio is 0.95/1.81 = 0.524. This value lies between 0.414-0.732 thereby revealing that the
coordination number should be 6 or 4 and the corresponding shape either octahedral or square
planar. However, X-ray crystallography shows that this has octahedral structure. Each Na+ ion is
surrounded by six Cl- ions at the corner of a regular octahedron and also each Cl- ion is surrounded
by six Na+ ions (Fig. 15). Thus the coordination number of both Na+ and Cl- ions is six.

It crystallizes as cubic close packed (ccp) array of Cl- ions, with Na+ ions occupying all the
octahedral sites. The stoichiometry of NaCl crystal is 1:1.

15
Cl- Na+

Fig. 15. Crystal structure of NaCl. Each ion has six nearest neighbors, with octahedral geometry

Many other compounds such as NaF, KCl, KBr, AgCl, AgBr, CaS, CaO, LiF, LiCl, LiI, etc, adopt
the NaCl structure. A number of inter-atomic distances may be calculated for any material with a
NaCl structure using the lattice parameter, a.
Na - Cl = a/2 = 0.5 a;
Na - Na = Cl - Cl = a/√2 ≈ 0.707 × a

Cesium chloride (CsCl): In CsCl the radius ratio (rCs+/rCl- = 1.69 Å/ 1.81 Å) is 0.933. Since this
value lies between 0.732 and 1.000, the coordination number of the ions is 8 in a cubic
arrangement, i.e., each Cs+ ion is surrounded by eight Cl- ions and similarly each Cl- ion is
surrounded by eight Cs+ ions. The coordination is, therefore, 8:8.

The arrangement of the ions in this crystal is body-centered cubic type, i.e., the unit cell of its
structure has one ion at the center and the oppositely charged ions at the corners of the cube.
Conventionally, the Cl- ions are regarded taking the positions at the 8 corners of the cube and Cs+
ions are present at the center of each cube (Fig. 16).

Cl-

Cs+

Fig. 16. Structure of CsCl

Note that the structure of CsCl is not strictly a body-centered cubic. In body-centered cubic
arrangement the atom at the center is identical to those at the corners. This structure is actually
found in metals, but in CsCl the ions at the corners are Cl- then there will be Cs+ at the body center.
As a result, the structure of CsCl must be described as a body-centered cubic type and not body
centered cubic.

16
CsCl type structure is found in those crystals in which the cations are comparable in size to anions.
The inter-atomic distance may be calculated for any material with a CsCl structure from the cell
lattice constant, a.
Cs-Cl = √3/2 a ≈ 0.866 a
Cs-Cs = Cl-Cl = a

Examples of some compounds, which have CsCl structure, are CsBr, CsI, TlCl, TlBr, etc. As the
coordination number of CsCl is higher than that of NaCl, it is expected that CsCl is more stable
than NaCl.

Zinc sulphide (zinc blende or sphalerite): As the radius ratio of ZnS (0.70 Å/1.84 Å) is 0.40; this
reveals that the coordination number is 4 and it is having a tetrahedral structure. Hence each Zn2+
ion is being surrounded by four S2- ions, which are deposed towards the four corners of a regular
octahedron. Also, each S2- is being surrounded by four Zn2+ ions (Fig. 17).

Conventionally, in zinc blende structure the S2- ions are fcc type of arrangement and Zn2+ ions are
occupying half of the 8 alternative tetrahedral sites while the remaining tetrahedral sites are empty.
The stoichiometry of the compound is 1:1.

S2-

Zn2+

Fig. 17. Structure of zinc blende (ZnS)

Examples of some other compounds, which have zinc blende structure, are: CuCl, CuBr, AgI, BeS
etc. A number of inter-atomic distances may be calculated for any material with a zinc blende unit
cell using the lattice parameter, a.
Zn-S = (a√3)/4 ≈ 0.422 a
Zn-Zn = S-S = a/√2 ≈ 0.707a

Zinc sulphide (wurtzite): Wurtzite is another polymorph of zinc sulphide. The unit cell of the
wurtzite structure is shown in Fig. 18. It is composed of an hcp array of S2- ions with every
alternate tetrahedral holes occupied by Zn2+ ions. Each Zn2+ ion is tetrahedrally coordinated by four
S2- ions and vice versa. The stoichiometry of the compound is 1:1.

17
Tetrahedral
arrangement of 4S2-
ions round Zn2+ ion Tetrahedral
arrangement of 4Zn2+
ions round S2- ion

Zn2+ S2-

Fig. 18. Crystal structure of wurtzite (ZnS)

The main difference between zinc blende and wurtzite structures is that the former is having ccp
type of packing for S2- ions while the latter is having hcp type of packing for S2- ions. However, the
coordination number of Zn2+ and S2- in either case is 4:4. Other compounds having wurtzite
structure are ZnO, AlN, CdS, BeO, etc. A number of inter atomic distances may be calculated for
any material with a wurtzite cell using the lattice parameters, a and c.
Zn-S = √(3/8) a ≈ 0.612 a = (3/8) c = 0.375 c
Zn-Zn = S-S = a = 1.632 c

Note that the Madelung constants (1.638 for zinc blende and 1.641 for wurtzite) for the two
structures are very much comparable. The energy difference between the structures is also small.
Because of this fact, some compounds crystallized in any form of the two may be changed into the
other.

Nickel arsenide (NiAs): It is a rare structure and is the hcp analogue of the very common NaCl
lattice. The anions (As3-) build an expanded and distorted hexagonal-close-packed array and the
cations (Ni3+) fill all octahedral holes (Fig. 19). Thus Ni3+ ion centers are in octahedral geometry
while each As3- ion resides at the center of a trigonal prism constituted by the Ni3+ ion. This
structure type is typical for AB-compounds, which contain polarizable anions and cations.

Fig. 19. Structure of nickel arsenide (NiAs)

This structure is also adopted by NiS, FeS, and other sulphides. It is common where there are soft
anions and soft cations, suggesting a degree of covalency.

18
Structures of ionic crystals of AB2 type

Calcium fluoride (fluorite): The fluorite structure is shown in Fig. 20. The fluorite structure is
found when the radius ratio is 0.73 or above. In the fluorite structure, each Ca2+ (rCa+ = 0.99 Å) ion
is surrounded by eight F- (rF- = 1.36 Å) ions at the corners of a body centered cubic arrangement
while each F- ion is surrounded by four Ca2+ ions arranged tetrahedrally. Thus the coordination
numbers of Ca2+ and F- are 8 and 4 respectively. Ca2+ ions in fluorite structure are arranged in face-
centered cubic close packed (ccp) type of arrangement. As there are two tetrahedral sites available
for every Ca2+ ion, the F- ions occupy all the tetrahedral sites. Thus the stoichiometry of the
compound is 1:2.

Other examples of crystals having fluorite type structure are: SrF2, BaF2, BaCl2, CdF2 etc.

Fig. 20. Structure of CaF2 (fluorite)

Antifluorite is a mineral with a crystal structure identical with that of fluorite but with the positions
of the cations and anions reversed (Fig. 21). Lithium oxide, Li2O has an antifluorite structure. The
O-2 ions constitute a cubic close packed lattice (fcc structure) and the Li+ ions occupy all the
tetrahedral voids. Thus the stoichiometry of the compound is 2:1. Each oxide ion, O-2 ion (rO2- =
1.40 Å) is in contact with 8 Li+ ions and each Li+ (rLi+ = 0.59 Å) ions having contact with 4 oxide
ion. Therefore, Li2O has 4:8 coordination.

Fig. 21. Structure of Li2O (antifluorite)

Examples are Na2O, K2O, K2S, Na2S, Rb2O, Rb2S.

19
Titanium dioxide (rutile): The unit cell of the crystal of TiO2 is given in Fig. 22. In it, Ti4+ ions are
present at all eight corners and at the body center of a distorted cube. The cube is distorted because
one of the axes of the cube is shorter than the other by 30%. Therefore we can say that Ti4+ ions
have a body centered distorted cubic arrangement. Oxide ions occupy positions of three fold
coordination.

Fig. 22. Structures of TiO2 (rutile)

Since the radius ratio (rTi4+/rO2- = 0.745 Å/1.260 Å = 0.591) in this crystal lies between 0.441 and
0.732, the coordination number of Ti4+ ion is six, i.e., each Ti4+ ion is octahedrally surrounded by
six O2- ions placed at the corners of an octahedron. On the other hand, each O2- ion is surrounded
by three Ti4+ ions placed at the corners of a triangle. Thus TiO2 has a 6:3 coordination. Examples of
other crystals having TiO2 type structure are NiF2, CoF2, SnO2, PbO2 etc.
]

As in the case of fluorite structure, in rutile we can have also the antirutile structure (Fig. 23). The
antirutile crystal structure is the converse of the more familiar rutile structure. In it, the nitride
anions occupy the eight corners and at the body center of a distorted cube and the titanium cations
occupy positions of three fold coordination. Example, ɛ-Ti2N.

Fig. 23. Structure of ɛ-Ti2N (antirutile)

20
Comparison of the characteristic features of the common basic ionic crystals
Crystal Characteristics of arrangement of C.N. Examples
the ions Cation Anion
NaCl In a face-centered lattice of the 6 (Oh) 6(Oh) NH4Cl, NH4Br,
anions, all the Oh sites are NH4I, AgF,
occupied by the cations. AgCl, AgBr,
MgO, CaO,
TiO, FeO, NiO
CsCl Both the cations and anions 8 8 CsBr, CsI,
occupy the cubic holes formed by (cubically) (cubically) CsCN, CaS
the opposite ions
ZnS (zinc The anions form a fcc lattice in 4 (Td) 4 (Td) BeS, CuCl,
blende) which half of the Td sites are CuBr, CuI, AgI,
occupied by the cations HgS
ZnS The anions have a hcp lattice in 4 (Td) 4 (Td) ZnO, CdS, BeO
(wurtzite) which half of the tetrahedral sites
are occupied by the cations
NiAs A hcp array of As atoms with - -
nickel atoms in all octahedral sites
CaF2 The cations form a fcc lattice in 8 4 (Td) SrF2, BaF2,
(fluorite) which all the Td sites are occupied (cubically) BaCl2, SrCl2,
by the anions CdF2, HgF2
Li2O The positions of the cations anions 4 (Td) 8 K2O, Na2O,
(Antifluorite) are reversed with respect to those (cubically) K2S, Na2S
of the fluorite structure
TiO2 (Rutile) The cations lie in a body-centered 6 (Oh) 3 MgF2, CoF2,
tetragonal geometry (triangular) NiF2, SnO2,
MnO2, etc.
ɛ-Ti2N The positions of the cations anions 3 6 (Oh) -
(Antirutile) are reversed with respect to the (triangular)
rutile structure

21
Mixed metal oxide structures

Perovskite: The structure is named after the mineral CaTiO3. The perovskite structure is adopted by
many oxides with the general formula ABO3 and is very versatile having many useful technological
applications such as ferroelectrics, catalysts, sensors and superconductors. It is generally
represented as: AIIBIVO3, AIIIBIIIO3, AIBVO3. Examples: AIIBIVO3 (A = Ca, Sr or Ba; B = Ti, Zr, Ge
or Sn); AIIIBIIIO3 (A = La; B = Al, Ti, Cr or Mn); AIBVO3 (A = Na or K; B = Nb or Ta). Some
mixed fluorides (ABF3) such as KZnF3 adopt the perovskite structure.

Fig. 24. Crystal structure of CaTiO3

In CaTiO3 structure, the Ca2+ and O2- ions are known to form a cubic close-packed arrangement, in
which the smaller Ti4+ ion occupies the octahedral holes formed exclusively by the O2- ions (Fig.
24). In the unit cell, 8 corners of a cube are occupied by Ca2+ ions, O2- ions are placed at the centers
of the 6 faces of the cube, and Ti4+ is placed at the center of the cube. Contribution of Ca2+ ions is 1
(= 8×⅛), contribution of O2- ions is 3 (= 6×½) and contribution of Ti4+ is 1 to the unit cell. It leads
to stoichiometry CaTiO3. Here each Ca2+ ion surrounded by twelve O2- ions and each Ti4+ ion
surrounded by six O2- ions.

Normal spinel and inverse spinel: The spinel minerals have the generic formula AB2O4, taking
their name from the naturally occurring mineral spinel, MgAl2O4. It is usually described as a cubic
close-packed array of O2- ions with the A and B cations occupying ⅛th of the tetrahedral sites and
½ of the octahedral sites, i.e. [A]tet[B2]octO4 . The occupancy of the interstitial sites results in a face-
centered cubic unit cell which is 2×2×2 times that of the basic ccp oxygen array, and therefore there
are 32 O2- ions in the spinel unit cell. The unit cell contents are therefore A8B16O32. The large
number of ions in the unit cell makes this a difficult structure to illustrate. However, Fig. 25(a)
shows one way of illustrating the regular cation distribution.

22
One of the characteristic of the spinel structure is its flexibility in the range of cations and cation
charge combinations (A and B may have different types of charge combinations such as: AIIB2IIIO4
(A = Mg, Zn, Cd, Mn, Fe; B = Al, Cr), AIVB2IIO4 (TiZn2O4) and AVIB2IO4 (A = Mo; B = Na)) it
will accept, making it a structure adopted by over a hundred compounds, many of them are
important minerals or important commercial magnetic oxides.

The spinel structure shown by MgAl2O4 is complex, which contains 32 O2- ions arranged in almost
perfect cubic closest packing in which eight of the 64 tetrahedral interstices are filled by Mg2+ and
sixteen of the 32 octahedral interstices are filled with the Al3+ ions.

(a) (b)
Fig. 25. (a) Structure of AB2O4 spinel, and (b) crystal structure of MgAl2O4

Many sulphides and halides also have the normal spinel structure and different charge combinations
are possible, namely: ZnAl2S4, Cu2SnS4, Li2NiF4.

There are also inverse spinels, where half of the B ions occupy tetrahedral sites, leaving the
remaining B ions and all the A ions in octahedral sites, i.e. [B]tet[A,B]octO4. Thus, here A has a
strong tendency to occupy the octahedral sites. This generally takes place in transition metal oxides.
In general the site selection is governed by the relative change of crystal field stabilization energy
(CFSE) in occupying the octahedral and tetrahedral sites by A. If in placing A in the octahedral
sites instead of the tetrahedral sites, it brings more CFSE, thus the inverse spinel structure is
favoured, provided in shifting B from the octahedral site to the tetrahedral site, the CFSE change is
negligible. This condition is generally attained when A is not a d5 or d10 system and B bears a d5 or
d10 configuration because in the weak field (which favours the high spin complexes) provided by
the O2- ions, the d5 and d10 systems do not have any CFSE in both the octahedral and tetrahedral
fields. Therefore, the condition to attain the inverse spinel structure is, “gain of CFSE in shifting A
from a tetrahedral site to an octahedral site > loss of CFSE in shifting B from an octahedral site to a
tetrahedral site”.

23
As an example, we can consider magnetite, Fe3O4. This compound contains one Fe2+ and two Fe3+
ions per formula unit, so we could formulate it as a normal spinel, Fe2+(Fe3+)2O4, or as an inverse
spinel, Fe3+(Fe2+Fe3+)O4. Which one would have the lowest energy?

d-orbital energy diagram for Fe2+

First we consider the crystal field energy of the Fe2+ ion, which is d6. Comparing the tetrahedral and
high spin octahedral diagrams, we find that the CFSE in an octahedral field of O2- ions is
[(4)(⅖) − (2)(⅗)]Δo – P = 0.4Δo − P. In the tetrahedral field, the CFSE is [(3)(⅗) − (3)(⅖)]Δt – P =
0.6Δt − P. Since Δo is about 2.25 times larger than Δt, the octahedral arrangement has a larger CFSE
and is preferred for Fe2+.

d-orbital energy diagram for Fe3+

In contrast, it is easy to show that Fe3+, which is d5, would have a CFSE of zero in either the
octahedral or tetrahedral geometry. This means that Fe2+ has a preference for the octahedral site,
but Fe3+ has no preference. Consequently, we place Fe2+ on octahedral sites and Fe3O4 is an inverse
spinel, Fe3+(Fe2+Fe3+)O4.

Ferrites (MIIFe2O4 = Fe, Mg, Mn, Co, Ni, Zn) can be normal or inverse spinels, or mixed spinels,
depending on the CFSE of the MII ion. Based on their CFSE, Fe2+, Co2+, and Ni2+ all have a strong
preference for the octahedral site, so those compounds are all inverse spinels. ZnFe2O4 is a normal
spinel because the small Zn2+ ion (d10) fits more easily into the tetrahedral site than Fe3+ (d5), and
both ions have zero CFSE. MgFe2O4 and MnFe2O4, in which all ions have zero CFSE and no site
preference, are mixed spinels. Chromite spinels, MIICr2O4, are always normal spinels because the
d3 Cr3+ ion has a strong preference for the octahedral site.

Note that besides the CFSE, many other factors may contribute to determine the spinel or inverse
spinel structure, but the prediction from CFSE calculation is in good agreement with the
experimental fact.
24
Ilmenite: The general formula of both perovskite and ilmenite is ABO3. However, the ilmenite
structure occurs when the A ion in an ABO3 compound is too small to form the perovskite
structure. For example, when the A ion radius is less than about 1.0 Å, the ilmenite arrangement
sometimes occurs. It is named after the mineral FeTiO3. This structure is closely related to that of
the minerals corundum and hematite and may be described as a hexagonal close-packed array of
O2- ions with Fe2+ and Ti4+ ions each occupying one-third of the octahedrally coordinated
interstices. Thus, both Fe2+ and Ti4+ ions are 6-coordinated by
O2- ions (Fig. 26). Having different type of charge
combinations, the structure can be represented as: AIIBIVO3,
AIBVO3. Examples: AIIBIVO3 (A = Fe, Cd, Mg, Co; B = Ti),
AIBVO3 (A = Li, B = Nb).

Note that some compounds may have a perovskite structure in


one temperature range and an ilmenite structure in other
temperature ranges. CdTiO3, for example, has a perovskite
structure at elevated temperatures but has an ilmenite structure Fig. 26. Structure of FeTiO3
at lower temperatures.

Problem 5: One crystalline form of metallic iron is body centered cubic, in which the metallic
radius is 124 pm. This form of iron is called α-iron. Calculate the density (d) of α-iron.

Solution: Recall that a body centered unit cell contains 2 atoms. Let a = the unit cell edge, in cm.
2 × 55.85 g mol-1
d = --------------------------------
(6.022 × 1023 mol-1) × a3
Again for a body centered cubic unit cell the atoms touch along the body diagonal. This body
diagonal has a length equal to 4 times the radius (r) of an iron atom.
i.e., 4r = √3 a
4 × 1.24 × 10-8 cm = √3 a
a = 2.86 × 10-8 cm
2 × 55.85 g mol-1
Then d = ------------------------------------------------- = 7.93 g cm-3
(6.022 × 1023 mol-1) × (2.86 × 10-8 cm)3

25
Problem 6: Analysis shows that a metal oxide has the empirical formula M0.96O. Calculate the
percentage of M2+ and M3+ ions in the crystal.

Solution: Let the M2+ ion in the crystal be x and M3+ = 0.96 - x
Since total charge on the compound must be zero
2x + 3(0.96 – x) - 2 = 0
x = 0.88
% of M2+ = (0.88/0.96) × 100 = 91.67
% of M3+ = 100 - 91.67 = 8.33

Problem 7: KCl crystallizes in the same type of lattice as does NaCl. Given that rNa+/rCl- = 0.50 and
rNa+/rK+ = 0.70, calculate the ratio of the side of the unit cell for KCl to that of for NaCl.

Solution: For NaCl type structure, r+ + r- = a/2, where a is the edge length of the unit cell. So,
rK+ 0.5
----- + 1 ------ + 1
aKCl rK+ + rCl- rCl- 0.7
------- = ------------- = ----------- = ------------ = 1.143
aNaCl rNa+ + rCl- rNa+ 0.5 + 1
----- + 1
rCl-

Problem 8: Why does KCl adopt the rock salt structure in spite of a radius ratio greater than 0.732?
Given, rK+ in C.N. 6 = 1.38 Å, rCs+ in C.N. 8 = 1.51 Å, rCl- = 1.81 Å. Madelung constant for KCl and
CsCl structures 1.748 and 1.763 respectively.

Solution: The electrostatic potential energy, the main contributor to lattice energy for any structure,
is directly proportional to the Madelung constant (A) for that structure and inversely proportional to
the interionic distance d(= r+ + r-). Here,
d(KCl structure) = (1.38 + 1.81) Å = 3.19 Å
d(CsCl structure) = (1.51 + 1.81) Å = 3.32 Å

The ratio of the lattice enthalpies will be roughly equal to the ratio of the respective electrostatic
potential energies, i.e. the respective A/d terms. Hence,
UKCl AKCl dCsCl 1.748 × 3.32
-------- = -------- × -------- = ----------------- = 1.03
UCsCl ACsCl dKCl 1.763 × 3.19

This shows that the lattice for the KCl structure would be marginally more stable and favour this
structure over the CsCl structure.

26
Problem 9: How many ions of each are contained in the face-centered unit cell of NaCl? The
specific gravity of NaCl is 2.165, what is the cell edge length of the unit cell.

Solution: The unit cell of ccp contains 4 each of sodium and chloride ions. The atomic masses for
Na and Cl are 23.0 and 35.5 respectively. Assume the cell edge length to be a, then we have
4 × (23.0 + 35.5) g mol-1
------------------------------- = 2.165 g cm-3
(6.023 × 1023 mol-1) × a3
Solving for ‘a’ gives
a = 5.64 × 10-8 cm
= 564 pm

Problem 10: Li2O forms in the anti-fluorite structure (anions in fcc positions; Li+ ions occupy all
the tetrahedral sites).
(i) Compute the value of the lattice constant. Given, rLi+ = 0.59 Å, rO2- = 1.40 Å.
(ii) Calculate the density of Li2O.
(iii) What is the maximum radius of cation which can be accommodated in the vacant interstice of
the anion array?

Solution:
(i) (√3/4)a = rLi+ + rO2- = 0.59 Å + 1.40 Å = 1.99 Å
⸫ a = 1.99 Å × (4/√3) = 4.5957 Å

(ii) density, d = (8×6.94 + 4×16) g / (6.02×1023) (4.5957×10-8 cm)3


= 119.52 g / (6.02×1023) (97.06329×10-24 cm3)
= 119.52 g / (584.32×10-1 cm3)
= 2.045 g/cm3

(iii) √3a - (4rLi+ + 2rO2-)


= √3×4.5957 Å - (4×0.59 Å + 2×1.40 Å) = 2.79 Å
⸫ Radius = 1.395 Å
a - 2rO2- = 4.5957 Å - 2×1.40 Å =1.7957 Å
⸫ Radius = 0.898 Å

27

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy