Sunteți pe pagina 1din 14

CHAPTER 4

4.6

CRYSTAL STRUCTURES
4.1

ATOMIC ARRANGEMENTS

4.2

LATTICES AND UNIT CELLS

4.3

4.2.1

Unit Cells in Space

4.2.2

Atomic Packing in Crystals

4.7

CRYSTALLOGRAPHIC DIRECTIONS AND PLANES


4.6.1

Coordinates of Points

4.6.2

Indices of Directions

4.6.3

Families of Directions

4.6.4

Hexagonal Crystals (MillerBravais Indices)

4.6.5

Indices of Planes

4.6.6

Families of Planes

4.6.7

Planes in Hexagonal Crystals

CRYSTALLINE MATERIALS

METALLIC CRYSTAL STRUCTURES


4.3.1

Face-Centred Cubic (FCC)


Crystal Structure

4.3.2

Body-Centred Cubic (BCC) Crystal


Structure

4.3.3

Hexagonal Close-Packed (HCP)


Structure

4.3.4

Allotropy/Polymorphism

4.4

CLOSE-PACKED CRYSTAL STRUCTURES

4.5

INTERSTITIAL POSITIONS AND SIZES

4-1

4-2

4.1 ATOMIC ARRANGEMENTS

4.2 LATTICES

AND

UNIT CELLS

The properties of a material depend not only on atomic

A lattice is a collection of points (positions in space)

bonding and forces, but also, equally important, on how

arranged in a periodic pattern so that surroundings of each

atoms pack together. There are 3 levels in which atoms

point in the lattice are identical. The points that make up

may be arranged:

the lattice are called lattice points

(Figs. 4.2-1, 4.2-2).

An

everyday example in 2-dimensions is wall-paper.


No order: there is no special positional
relationship or interaction between the
atoms; e.g. inert gases (Fig. 4.1-1a).
Short-range

order:

the

(a) No order.

specific

arrangement of atoms extends only to


an atom's nearest neighbours. Materials
exhibiting

short-range

amorphous (glassy)

order

are

(Fig. 4.1-1b).

Fig. 4.2-1 Lattices and unit cells in 2D.


(b) Short-range
order.

Long-range order: there is a special


arrangement of atoms that is repeated
throughout

the

entire

material.

Crystalline materials exhibit both shortrange

and

long-range

order.

The

repetitive pattern formed by atoms in a


crystalline solid is known as a lattice
4.1-1c).

(Fig.

Fig. 4.2-2 Lattices and unit cells in 3D.


(c) Long-range
order.

A unit cell is the smallest subdivision of a lattice that


contains all the characteristics of the entire lattice. A

Fig. 4.1-1 The levels


of atomic
arrangement.

complete crystal can be formed by translating its unit cell

4-3

4-4

along each of its edges.

The properties of a unit cell are therefore the same as those


of the whole crystal.
Each unit cell is described by lattice parameters, which
are the lengths of the cell edges and the angles between
the axes. (Figs. 4.2-1, 4.2-2)

4.2.1

Unit Cells in Space

By geometry, there are only 7 unique unit cell shapes that


may be stacked together in space. And there are only a
total of 14 possible ways in which atoms may be arranged
inside these unit cells. These 14 different unit cells are
known as Bravais lattices and they fall into one of 7 crystal
systems (Fig. 4.2-3).
One or more atoms may be associated with each lattice
point. The "group of atoms" located at each lattice point is
the basis. The actual crystal structure itself is defined by a
combination of crystal lattice and crystal basis

Crystal lattice

(Fig. 4.2-4).

crystal
basis

crystal structure
4-5

Fig. 4.2-3 14 Bravais


lattices grouped into 7
crystal systems.

4-6

4.2.2

Atomic Packing in Crystals

The properties of an entire crystal, such as the theoretical

When describing crystal structures, atoms are assumed to


be solid spheres that touch one another. This is known as
the hard sphere model. The centres of the solid spheres
coincide with the lattice points in unit cells (Fig. 4.2-5).

density, may be calculated from just one of its unit cells.


Density = Mass of unit cell

Volume of unit cell

(Number of atoms in cell)(Mass of one atom)


Volume of unit cell

= (Number of atoms in cell)(Atomic mass)

(Volume of unit cell)(Avogadro's number)

Fig. 4.2-5 Hard


sphere model of
a unit cell.

4.3 METALLIC CRYSTAL STRUCTURES


In pure metals, only one metal ion occupies each lattice

The coordination number is the number of nearest

point. Many common metals may be defined by one of 3

neighbours (touching atoms) to any atom. For materials

crystal structures: face-centred cubic (FCC), body-centred

with non-directional bonding (i.e. metals and ionic solids),

cubic (BCC) and hexagonal close-packed (HCP) crystal

the lowest energy (most stable) configuration is obtained

structures (Table 4.3-1).

when atoms pack as closely as possible, separated only by


their equilibrium bond lengths. In such crystals, the
number of nearest neighbours (i.e. coordination number)
would be as high as possible.
The atomic packing factor (APF) is the fraction of space
occupied by atoms in a unit cell.
APF = Volume of atoms in unit cell
Volume of unit cell

(Number of atoms in cell)(Volume of one atom)


Volume of unit cell
4-7

4-8

4.3.1

Face-Centred Cubic (FCC) Crystal Structure

Cubic geometry with one atom at each corner and one at


the centre of each face. (Fig. 4.3-1)

Coordination number: consider a corner atom, which


touches 4 face-centred atoms on each of the 3 mutually
perpendicular planes passing through the corner atom
itself (Fig. 4.3-2), giving CN = 12.

Fig. 4.3-2 Finding the


coordination number in
FCC structures.

Atomic packing factor: each atom at the centres of the


cube faces is shared by 2 cells; each corner atom is shared
by 8 cells (Fig. 4.3-3), such that:
Atoms per cell = 4; APF = 0.74

Fig. 4.3-1 FCC structure.

Corner atoms touch face-centred atoms (along the face


diagonal) but corner atoms do not touch one another.
Face-centred atoms also touch adjacent (but not opposite)
face-centred atoms in the midplanes of the cube.
Fig 4.3-3 Sharing of face and corner atoms in FCC structures.

Unit cell length: a = 2R 2 (where R is the atomic radius)

Examples of FCC metals: Al, Au, Ag, Cu, Ni, Pb, Pt.
4-9

4-10

4.3.2

Body-Centred Cubic (BCC) Crystal Structure

Cubic geometry with one atom at each corner and one at


the centre of the cube. (Fig. 4.3-4)

4.3.3

Hexagonal Close-Packed (HCP) Structure

Hexagonal faces at the top and bottom are linked by 6


rectangular side faces. There is one atom at each corner of
the top and bottom hexagons surrounding one atom at
the centre of each hexagon. 3 other atoms are located on
a plane midway between the hexagons (Fig. 4.3-5).

Fig. 4.3-4 BCC structure.

Atoms touch along the body diagonal.


Fig. 4.3-5 HCP structure and its smaller primitive unit cell.

a=

4R
3

Corner atoms at the top and bottom hexagons are shared


between 6 cells; the central atom in each hexagon is

CN = 8.

shared between 2 cells; the 3 atoms in the midplane


belong to only one cell.

Atoms per cell = 2; APF = 0.68


Lattice parameters:
Examples of BCC metals: Cr, Fe, W, Mo, V.

a = 2R;

c=

4
a
6

By considering the central atom in basal plane, CN =12.


Atoms per cell = 6 (or 2 per primitive unit cell); APF = 0.74
Examples of HCP metals: Co, Mg, Ti, Zn.
4-11

4-12

4.3.4

Allotropy/Polymorphism

Allotropy or polymorphism is the ability of an element or


compound to assume more than one crystal structure in
the solid state, depending on external conditions such as
temperature, pressure, magnetic and electric fields (Table 4.3-2)
The change in crystal structure is usually accompanied by
changes in properties.

Fig. 4.3-6 PZT (lead zirconate titanate)


ceramic changes its structure from
(a) cubic, to (b) tetragonal,
in response to an electric field.

Fig. 4.3-7 Use of piezoelectric effect


of PZT crystals in inkjet printer head.

Such property changes can be very useful; e.g. hardening/


softening of steel through controlled heating/cooling
9&10),

(Chps

piezoelectric transducers (Figs. 4.3-6 & 7).

Or detrimental; e.g. distortion and cracking due to sudden


changes in volume, especially in brittle ceramics, but also
in metals (Fig. 4.3-8).
4-13

Fig. 4.3-8 Tin changes from tetragonal to diamond structure below 13.2C.
The volume expansion accompanying this transformation from
soft, ductile white tin to hard, brittle grey tin causes it to disintegrate.

4-14

4.4 CLOSE-PACKED CRYSTAL STRUCTURES


Atoms in metals pack as closely as
possible, separated only by their
equilibrium bond lengths, r0, as this
gives the lowest energy (most
stable) configuration (Fig 4.4-1).
Fig. 4.4-1
Bonding energy
is higher when
atoms are away
from equilibrium
separation r0.

Close-packed planes or directions


refer to the planes or directions in a
crystal, where the atoms are in

Fig. 4.4-2 Illustration of close-packed stacking sequence.

direct contact, assuming a hard


sphere model of atomic packing.
Plastic
occurs

deformation

in

most

on

readily

metals
close-

packed planes along close-packed directions in those


planes

(Sec. 6.1).

The number and relative positions of these

planes and directions influence properties such as ductility.


Packing same-sized atoms in FCC or HCP structures gives
the smallest volume (i.e. highest density). FCC and HCP
are known as close-packed structures.
FCC and HCP differ only in the arrangements of their
close-packed planes

(Fig. 4.4-2, 4.4-3);

this difference affects

plastic deformation and ductility (Sec. 6.1).

HCP
4-15

FCC

Fig. 4.4-3 Close-packed stacking sequence and close-packed planes for HCP and FCC.
4-16

4.5 INTERSTITIAL POSITIONS

AND

SIZES

Table 4.5-1 The size and number of interstitial sites in FCC, BCC and HCP.

Lattices not completely filled with atoms. Interstices are


the holes between lattice atoms.

Crystal
structure
FCC
BCC
HCP

Size of interstitial sites, Rr


Octahedral
0.414
0.155
0.414

Tetrahedral
0.225
0.291
0.225

No. of sites per unit cell


Octahedral
4
6
6

Tetrahedral
8
12
12

Interstitial sites are classified by geometry, based on the


shape of the polyhedron formed by existing lattice atoms
surrounding a particular site. (Fig. 4.5-1 & Table 4.5-1)
The size of an interstitial site is defined by the radius ratio,
r , where r is the radius of the largest sphere that can
R
completely fill the site without straining the adjacent lattice
atoms, and R is the radius of the lattice atoms.

Tetrahedral interstitial

Octahedral interstitial

r
= 0.225
R

Cubic interstitial

r
= 0.414
R

r
= 0.732
R

Fig. 4.5-1 Interstitial sites and sizes in close-packed crystal structures.

Atoms (alloy or impurity) occupying interstitial sites


6.3.2)

(Sec.

must be larger than the size of the holes; smaller

atoms are not allowed to rattle around loose in the sites.


4-17

Fig. 4.5-2 Locations of the interstitial sites in FCC, BCC and HCP structures.
4-18

4 . 6 C RYSTALLOGRAPHIC D IRECTIONS

AND

P LANES

Some physical and mechanical material properties vary


with the direction or plane within a crystal in which they
are measured. Miller Indices provide a convenient system
of describing points, directions and planes in crystals.

4.6.2

Indices of Directions

Procedure for finding Miller indices of directions:


1. Determine the coordinates of 2 points that lie in the
direction of interest.
2. Subtract coordinates of start point from end point.

4.6.1

Coordinates of Points

3. Clear fractions and/or reduce to lowest integers.

It is customary to use the right-hand coordinate system.

4. Write the indices in square brackets [ ] without commas.

The most common orientation is to align the coordinate


axes with the edges of the unit cell, putting the origin at

Negative integer values are indicated by placing a bar


over the integer. (Fig. 4.6-3)

one corner of the cell.


Each lattice point is defined by u, v, w, which correspond
Fig. 4.6-2 Some
common directions
in a cubic unit cell.

to fractions of the lattice parameters, a, b, c.


Fig. 4.6-1 Coordinates of some points
in a unit cell. Note that coordinate
axes are not necessarily perpendicular
to one another and lattice parameters
may not be the same length. The
choice of origin is arbitrary.

Worked Example (Fig. 4.6-3)

1. Coordinates of end and start


points: 0,0,1;

1
, 1, 0.
2

2. End - start coordinates:


(0 -

1
1
), (0 - 1), (1 - 0) = - , -1, 1
2
2

3. Clear fractions:
1
2

2(- , -1, 1) = -1, -2, 2


4. Miller index: [ 1 2 2 ]
4-19

4-20

4.6.3

4.6.4

Families of Directions

Hexagonal Crystals (Miller-Bravais Indices)

Certain directions in a unit cell, e.g. the cube edges, have

Due to the hexagonal geometry, equivalent directions will

identical atomic arrangements. Their Miller indices are

not have the same set of Miller indices. To reflect the

different because of the choice of origin and axes

symmetry of hexagonal systems, a 4-axis Miller-Bravais

(Fig. 4.6-4).

These directions are said to be symmetrically equivalent.

Fig. 4.6-4 Symmetrically equivalent directions in the <100> family.

Equivalent directions that are related by symmetry may be


grouped as a family of directions. 'Members' in the same
family share the same set of Miller indices, but in different
permutations, including negative indices (Fig. 4.6-4).
The family is represented by the Miller indices of one
family member, but enclosed in angle brackets <u v w>. It
is customary to choose one of the positive indices.

coordinate system may be used instead (Fig. 4.6-5).

Fig. 4.6-5 3- and 4-axis coordinate systems for hexagonal crystals. Since only 3 axes are
required to define any 3D geometry, the extra axis, a3, in the basal plane
is actually redundant; the 3rd index, t , in the Miller-Bravais system [uvtw],
is thus a function of u and v, such that t = -(u + v).

For directions in hexagonal system, first find Miller indices


(Sec. 4.6-2),

then convert to Miller-Bravais indices as follows:

Miller to Miller-Bravais:

Miller-Bravais to Miller:

[u v w] [uvtw]

[uvtw] [u v w]

u = n (2u - v)

u = u - t

belong to the same family of <100> directions. Face

3
v = n (2v - u)
3
t = - n (u + v)
3

diagonals form a different family, <110>, and body

w = nw

In cubic systems, the directions along the cube edges

diagonals yet another family, <111>.

v = v - t
w = w

where n is the integer required to clear fractions and/or reduce to lowest integers
4-21

4-22

4.6.5

4.6.6

Indices of Planes

Procedure for finding Miller indices of planes:

Families of Planes

Planes that contain identical atomic arrangements in a unit

1. Find the coordinates of the points where the plane


intercepts the 3 axes. If the plane passes through the
origin, shift the origin of the coordinate system.

cell and are related by symmetry are equivalent.


A family of equivalent planes share the same set of Miller
indices. The family is expressed by enclosing the Miller
indices of one member in braces, {h k l}.

2. Take reciprocals of these intercepts.


3. Clear fractions but do not reduce to lowest integers.
4. Write the indices in round brackets ( ) without commas.
Negative integer values are indicated by placing a bar
over the integer. (Fig. 4.6-6)

Fig. 4.6-7 The family of {110} planes contains (110) (101) (011) ( 11 0 ) ( 10 1 ) ( 01 1 ).

Worked example
(Fig. 4.6-6)

1. Plane passes through origin, so


shift axes to x, y, z.
Intercepts: x = -1, y = 1, z =

2. Reciprocals:

1
1
1
= -1, = 1, = 0
y
x
z

3. No fractions to clear.
4. Miller indices: ( 1 10 ).

Fig. 4.6-8 Some planes in a cubic unit cell.


4-23

4-24

4.6.7

4.7 CRYSTALLINE MATERIALS

Planes in Hexagonal Crystals

In order that equivalent planes have the same indices, the


In a single crystal, the same orientation and alignment of

4-axis Miller-Bravais coordinate system may be used.

unit cells is maintained throughout the entire crystal.


Miller-Bravais indices for planes are determined using the
(no

Most materials are polycrystalline: their structures are

formulae required), except that 4 intercepts are found,

composed of many small crystals (grains) with identical

giving indices of the form (h k i l).

structures but different orientations (Fig. 4.7-1).

same procedure as in 3-index Miller systems

(Sec. 4.6-1)

Worked Example (Fig. 4.6-9)

Miller Indices
of planes

(100)
(1 1 0 )
(0 1 0)

Reciprocals of intercepts for


4-axis Miller-Bravais system
a1

a2

a3

1
1
0

0
-1
-1

-1
0
1

0
0
0

Miller-Bravais Indices
of planes

(10 1 0 )
(1 1 00 )
( 0 1 10 )

Fig. 4.7-1 The solidification of a polycrystalline material.

Grain boundaries are the

interfaces where grains of


different

orientations

though the planes are equivalent. The 4-index Miller-

(Fig. 4.7-2).

Single crystals do

Bravais notation yields similar indices for equivalent planes.

not contain grain boundaries.

Note that with only 3 axes, Miller indices are different even

4-25

meet

4-26

The absence of grain boundaries in single crystals impart

Therefore, grain boundaries are attacked more aggressively

unique properties but such crystals are extremely difficult

by chemical etchants. Under a microscope, the more

to grow, requiring carefully controlled conditions, which

deeply etched grain boundaries scatter more light, and

are expensive. Single crystals are essential to some

appear darker

applications; e.g. semiconductors and jet turbine blades,

This is the principle behind metallography.

(Figs. 4.7-3/4),

thus revealing the microstructure.

piezoelectric transducers.
Besides the absence of grain boundaries, single crystals
exhibit directionality in properties, such as magnetism,
electrical conductivity and elastic modulus and creep
resistance, which depend on the crystallographic direction
of measurement. This directionality is called anisotropy.
The

random

orientation

of

individual

grains

in

polycrystalline material means that measured properties


are

independent

of

crystallographic

direction.

Such

materials are said to be isotropic.

Fig. 4.7-3 Observation of grains and grain boundaries in stainless steel sample.
Note that different orientations of the grains result in differences in reflection.

Grain boundaries are disordered regions of atomic


mismatch

where

atoms

equilibrium positions
coordination numbers

are

(Fig. 4.4-1)

displaced

from

their

and there are improper

(Sec. 4.2.2)

across the boundaries.

Atoms at grain boundaries hence possess higher energy.


This

interfacial

energy

makes

grain

boundaries

preferential sites for chemical reactions and other chemical


changes.

Fig. 4.7-4 Observed microstructure in 2-D and the underlying 3-D structure.
4-27

4-28

S-ar putea să vă placă și