You are on page 1of 24

FPEM1013 Introduction to Engineering Materials

Chapter 1 Structures of Solids


1.1 Fundamental Concepts of Matters
Solid materials may be classified according to the regularity with which atoms or ions
are arranged with respect to one another. A crystalline material is one in which the
atoms are situated in a repeating or periodic array over large atomic distances; that
is, long-range order exists, such that upon solidification, the atoms will position
themselves in a repetitive three-dimensional pattern, in which each atom is bonded
to its nearest-neighbor atoms. All metals, many ceramic materials, and certain
polymers form crystalline structures under normal solidification conditions. For
those that do not crystallize, this long-range atomic order is absent; these are called
noncrystalline or amorphous materials.

Some of the properties of crystalline solids depend on the crystal structure of the
material, the manner in which atoms, ions, or molecules are spatially arranged. There
is an extremely large number of different crystal structures all having long range
atomic order; these vary from relatively simple structures for metals to exceedingly
complex ones, as displayed by some of the ceramic and polymeric materials. The
present discussion deals with several common metallic crystal structures.

When describing crystalline structures, atoms (or ions) are thought of as being solid
spheres having well-defined diameters. This is termed the atomic hard sphere model
in which spheres representing
nearest-neighbor atoms touch one
another. An example of the hard
sphere model for the atomic
arrangement found in some of the
common elemental metals is
displayed in Figure 1-1c. In this
particular case all the atoms are
identical. Sometimes the term lattice
is used in the context of crystal
structures; in this sense “lattice”
means a three-dimensional array of
points coinciding with atom
positions (or sphere centers).

Figure 1-1 For the face centered


cubic crystal structure, (a) a hard
sphere unit cell representation, (b) a
reduced-sphere unit cell, and (c) an
aggregate of many atoms.

Chapter 1 1|Page
FPEM1013 Introduction to Engineering Materials
1.2 Short-Range Order versus Long-Range Order
In different states of matter, we can find four types of atomic or ionic arrangements
(Figure 1-2).

No Order In monoatomic gases, such as argon (Ar) or plasma


created in a fluorescent tubelight, atoms or ions have no orderly arrangement.

Figure 1-2 Levels of atomic arrangements in materials: (a) Inert monoatomic


gases have no regular ordering of atoms. (b,c) Some materials, including water vapor,
nitrogen gas, amorphous silicon, and silicate glass, have short-range order. (d) Metals,
alloys, many ceramics and some polymers have regular ordering of atoms ions that
extends through the material.

Short-Range Order (SRO) A material displays short-range order


(SRO) if the special arrangement of the atoms extends only to the atom’s nearest
neighbors. Each water molecule in steam has short-range order due to the covalent
bonds between the hydrogen and oxygen atoms; that is, each oxygen atom is joined
to two hydrogen atoms, forming an angle of 104.5° between the bonds. There is no
long-range order, however, because the water molecules in steam have no special
arrangement with respect to each other’s position.

A similar situation exists in materials known as inorganic glasses, i.e. tetrahedral


structure in silica that satisfies the requirement that four oxygen ions be bonded to
each silicon ion [Figure 1-3(a)]. As will be discussed later, in a glass, individual
tetrahedral units are joined together in a random manner. These tetrahedra may
share corners, edges, or faces. Thus, beyond the basic unit of a (SiO4)4- tetrahedron,
there is no periodicity in their arrangement. In contrast, in quartz or other forms of
crystalline silica, the (SiO4)4- tetrahedra are indeed connected in different periodic

Chapter 1 2|Page
FPEM1013 Introduction to Engineering Materials
arrangements. Many polymers also display short-range atomic arrangements that
closely resemble the silicate glass structure. Polyethylene is composed of chains of
carbon atoms, with two hydrogen atoms attached to each carbon. Because carbon has
a valence of four and the carbon and hydrogen atoms are bonded covalently, a
tetrahedral structure is again produced [Figure 1-3(b)]. Tetrahedral units can be
joined in a random manner to produce polymer chains.

Figure 1-3 (a) Basic Si-O tetrahedron in


silicate glass. (b) Tetrahedral arrangement of C-
H bonds in polyethylene.

Long-Range Order (LRO) Most metals and alloys, semiconductors,


ceramics, and some polymers have a crystalline structure in which the atoms or ions
display long-range order (LRO); the special atomic arrangement extends over much
larger length scales ≥100 nm. The atoms or ions in these materials form a regular
repetitive, grid-like pattern, in three dimensions. We refer to these materials as
crystalline materials. If a crystalline material consists of only one large crystal, we
refer to it as a single crystal. Single crystals are useful in many electronic and optical
applications. For example, computer chips are made from silicon in the form of large
(up to 12 inch diameter) single crystals [Figure 1-4(a)]. Similarly, many useful
optoelectronic devices are made from crystals of lithium niobate (LiNbO3). Single
crystals can also be made as thin films and used for many electronic and other
applications. Certain types of turbine blades may also be made from single crystals of
nickel-based superalloys. A polycrystalline material is composed of many small
crystals with varying orientations in space. These smaller crystals are known as
grains. The borders between crystals, where the crystals are in misalignment, are
known as grain boundaries. Figure 1-4(b) shows the microstructure of a
polycrystalline stainless steel material.

Chapter 1 3|Page
FPEM1013 Introduction to Engineering Materials

Figure 1-4 (a) Photograph of a silicon single crystal. (b) Micrograph of a


polycrystalline stainless steel showing grains and grain boundaries

Figure 1-5 shows a summary of classification of materials based on the type of atomic
order.

Figure 1-5 Classification of materials based on the type of atomic order.

Chapter 1 4|Page
FPEM1013 Introduction to Engineering Materials
1.3 Unit Cell
Our goal in choosing a unit cell for a crystal structure is to find the single repeat unit
that, when duplicated and translated, reproduces the entire crystal structure. For
example, imagine the crystal as a three-dimensional puzzle for which each piece of
the puzzle is exactly the same. If we know what one puzzle piece looks like, we know
what the entire puzzle looks like, and we don’t have to put the entire puzzle together
to solve it. We just need one piece! To understand the unit cell concept, we start with
the crystal. Figure 1-6(a) depicts a hypothetical two-dimensional crystal that consists
of atoms all of the same type.

Next, we add a grid that mimics the symmetry of the arrangements of atoms. There is
an infinite number of possibilities for the grid, but by convention, we usually choose
the simplest. For the square array of atoms shown in Figure 1-6(a), we choose a
square grid as is shown in Figure 1-6(b). Next, we select the repeat unit of the grid,
which is also known as the unit cell. This is the unit that, when duplicated and
translated by integer multiples of the axial lengths of the unit cell, recreates the entire
crystal. The unit cell is shown in Figure 1-6(c); note that for each unit cell, there is
only one quarter of an atom at each corner in two dimensions. We will always draw
full circles to represent atoms, but it is understood that only the fraction of the atom
that is contained inside the unit cell contributes to the total number of atoms per unit
cell.

Thus, there is 1/4 atom / corner * 4 corners = 1 atom per unit cell, as shown in Figure
1-6(c). It is also important to note that, if there is an atom at one corner of a unit cell,
there must be an atom at every corner of the unit cell in order to maintain the
translational symmetry. Each unit cell has its own origin, as shown in Figure 1-6(c).

Figure 1-6 The unit cell. (a) A two-dimensional crystal. (b) The crystal with an
overlay of a grid that reflects the symmetry of the crystal. (c) The repeat unit of the
grid known as the unit cell. Each unit cell has its own origin.

Chapter 1 5|Page
FPEM1013 Introduction to Engineering Materials
1.4 Metallic Crystal Structures
The atomic bonding in this group of materials is metallic and thus non-directional in
nature. Consequently, there are minimal restrictions as to the number and position of
nearest-neighbor atoms; this leads to relatively large numbers of nearest neighbors
and dense atomic packings for most metallic crystal structures. Also, for metals, using
the hard sphere model for the crystal structure, each sphere represents an ion core.
Table 1-1 presents the atomic radii for a number of metals. Three relatively simple
crystal structures are found for most of the common metals: face-centered cubic,
body-centered cubic, and hexagonal close-packed.

Table 1-1 Atomic Radii and Crystal Structures for 16 Metals

1.4.1 The Face-Centered Cubic Crystal Structure


The crystal structure found for many metals has a unit cell of cubic geometry, with
atoms located at each of the corners and the centers of all the cube faces. It is aptly
called the face-centered cubic (FCC) crystal structure. Some of the familiar metals
having this crystal structure are copper, aluminum, silver, and gold (see also Table 1-
1). Figure 1-1a shows a hard sphere model for the FCC unit cell, whereas in Figure 1-
1b the atom centers are represented by small circles to provide a better perspective
of atom positions. The aggregate of atoms in Figure 1-1c represents a section of
crystal consisting of many FCC unit cells. These spheres or ion cores touch one
another across a face diagonal; the cube edge length a and the atomic radius R are
related through

a = 2R√2

For the FCC crystal structure, each corner atom is shared among eight unit cells,
whereas a face-centered atom belongs to only two. Therefore, one-eighth of each of
the eight corner atoms and one-half of each of the six face atoms, or a total of four
whole atoms, may be assigned to a given unit cell. This is depicted in Figure 1-1a,
where only sphere portions are represented within the confines of the cube. The cell
comprises the volume of the cube, which is generated from the centers of the corner

Chapter 1 6|Page
FPEM1013 Introduction to Engineering Materials
atoms as shown in the figure. Corner and face positions are really equivalent; that is,
translation of the cube corner from an original corner atom to the center of a face
atom will not alter the cell structure.

Two other important characteristics of a crystal structure are the coordination


number and the atomic packing factor (APF). For metals, each atom has the same
number of nearest-neighbor or touching atoms, which is the coordination number.
For face-centered cubics, the coordination number is 12. This may be confirmed by
examination of Figure 1-1a; the front face atom has four corner nearest-neighbor
atoms surrounding it, four face atoms that are in contact from behind, and four other
equivalent face atoms residing in the next unit cell to the front, which is not shown.

The APF is the sum of the sphere volumes of all atoms within a unit cell (assuming the
atomic hard sphere model) divided by the unit cell volume—that is

(number of atoms/cell)(volume of each atom)


Packing factor =
volume of unit cell

For the FCC structure, the atomic packing factor is 0.74, which is the maximum
packing possible for spheres all having the same diameter. Metals typically have
relatively large atomic packing factors to maximize the shielding provided by the free
electron cloud.

Chapter 1 7|Page
FPEM1013 Introduction to Engineering Materials
1.4.2 The Body-Centered Cubic Crystal Structure
Another common metallic crystal structure also has a cubic unit cell with atoms
located at all eight corners and a single atom at the cube center. This is called a body-
centered cubic (BCC) crystal structure. A collection of spheres depicting this crystal
structure is shown in Figure 1-7c, whereas Figures 1-7a and 1-7b are diagrams of BCC
unit cells with the atoms represented by hard sphere and reduced-sphere models,
respectively. Center and corner atoms touch one another along cube diagonals, and
unit cell length a and atomic radius R are related through

4𝑅𝑅
𝑎𝑎 =
√3

Figure 1-7 For the body-centered cubic crystal structure, (a) a hard sphere unit
cell representation, (b) a reduced-sphere unit cell, and (c) an aggregate of many
atoms.

Chromium, iron, tungsten, as well as several other metals listed in Table 1-1 exhibit a
BCC structure. Two atoms are associated with each BCC unit cell: the equivalent of
one atom from the eight corners, each of which is shared among eight unit cells, and
the single center atom, which is wholly contained within its cell. In addition, corner
and center atom positions are equivalent. The coordination number for the BCC
crystal structure is 8; each center atom has as nearest neighbors its eight corner
atoms. Since the coordination number is less for BCC than FCC, so also is the atomic
packing factor for BCC lower—0.68 versus 0.74.

Chapter 1 8|Page
FPEM1013 Introduction to Engineering Materials
1.4.3 The Hexagonal Close-Packed Crystal Structure
Not all metals have unit cells with cubic symmetry; the final common metallic crystal
structure to be discussed has a unit cell that is hexagonal. Figure 1-8a shows a
reduced-sphere unit cell for this structure, which is termed hexagonal closepacked
(HCP); an assemblage of several HCP unit cells is presented in Figure 1-8b. The top
and bottom faces of the unit cell consist of six atoms that form regular hexagons and
surround a single atom in the center. Another plane that provides three additional
atoms to the unit cell is situated between the top and bottom planes. The atoms in
this midplane have as nearest neighbors atoms in both of the adjacent two planes.
The equivalent of six atoms is contained in each unit cell; one-sixth of each of the 12
top and bottom face corner atoms, one-half of each of the 2 center face atoms, and all
3 midplane interior atoms. If a and c represent, respectively, the short and long unit
cell dimensions of Figure 1-8a, the c/a ratio should be 1.633; however, for some HCP
metals this ratio deviates from the ideal value. The coordination number and the
atomic packing factor for the HCP crystal structure are the same as for FCC: 12 and
0.74, respectively. The HCP metals include cadmium, magnesium, titanium, and zinc;
some of these are listed in Table 1-1.

Figure 1-8 For the hexagonal close-packed crystal structure, (a) a reduced-sphere
unit cell (a and c represent the short and long edge lengths, respectively), and (b) an
aggregate of many atoms.

Chapter 1 9|Page
FPEM1013 Introduction to Engineering Materials
1.5 Lattice Parameters and Interaxial Angles
The lattice parameters are the axial lengths or dimensions of the unit cell and are
denoted by convention as a, b, and c. The angles between the axial lengths, known as
the interaxial angles, are denoted by the Greek letters α, β, and γ. By convention, α is
the angle between the lengths b and c, β is the angle between a and c, and γ is the
angle between a and b, as shown in Figure 1-9. (Notice that for each combination,
there is a letter a, b, and c whether it be written in Greek or Roman letters.)

Figure 1-9 Definition of the lattice parameters and their use in cubic,
orthorhombic, and hexagonal crystal systems.

In a cubic crystal system, only the length of one of the sides of the cube need be
specified (it is sometimes designated a0). The length is often given in nanometers (nm)
or angstrom (Å) units, where

1 nanometer (nm) = 10-9 m = 10-7 cm = 10 Å


1 angstrom (Å) = 0.1 nm = 10-10 m = 10-8 cm

The lattice parameters and interaxial angles for the unit cells of the seven crystal
systems are presented in Table 1-2.

Chapter 1 10 | P a g e
FPEM1013 Introduction to Engineering Materials
Table 1-2 Characteristics of the seven crystal systems

To fully define a unit cell, the lattice parameters or ratios between the axial lengths,
interaxial angles, and atomic coordinates must be specified. In specifying atomic
coordinates, whole atoms are placed in the unit cell. The coordinates are specified as
fractions of the axial lengths.

Number of Atoms per Unit Cell Each unit cell contains a specific
number of lattice points. When counting the number of lattice points belonging to
each unit cell, we must recognize that, like atoms, lattice points may be shared by
more than one unit cell. A lattice point at a corner of one unit cell is shared by seven
adjacent unit cells (thus a total of eight cells); only one-eighth of each corner belongs
to one particular cell. Thus, the number of lattice points from all corner positions in
one unit cell is

1
lattice point 8 corners 1 lattice point
�8 �� �=
corner cell unit cell

Corners contribute 1/8 of a point, faces contribute 1/2, and body-centered positions
contribute a whole point [Figure 1-10(a)].

Chapter 1 11 | P a g e
FPEM1013 Introduction to Engineering Materials

Figure 1-10 (a) Illustration showing sharing of face and corner atoms. (b) The
models for simple cubic (SC), body-centered cubic (BCC), and face-centered cubic
(FCC) unit cells, assuming only one atom per lattice point.

The number of atoms per unit cell is the product of the number of atoms per lattice
point and the number of lattice points per unit cell. The structures of simple cubic
(SC), body-centered cubic (BCC), and face-centered cubic (FCC) unit cells (with one
atom located at each lattice point) are shown in Figure 1-10(b).

Chapter 1 12 | P a g e
FPEM1013 Introduction to Engineering Materials
Atomic Radius versus Lattice Parameter Directions in the unit cell
along which atoms are in continuous contact are close-packed directions. In simple
structures, particularly those with only one atom per lattice point, we use these
directions to calculate the relationship between the apparent size of the atom and the
size of the unit cell. By geometrically determining the length of the direction relative
to the lattice parameters, and then adding the number of atomic radii along this
direction, we can determine the desired relationship.

If we refer to Figure 1-11, we find that atoms touch along the edge of the cube in SC
structures. The corner atoms are centered on the corners of the cube, so

a0 = 2r

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

4r
a0 =
√3

In FCC structures, atoms touch along the face diagonal of the cube, which is in length.
There are four atomic radii along this length—two radii from the face centered atom
and one radius from each corner, so

4r
a0 =
√2

Figure 1-11 The relationships between the atomic radius and the lattice parameter
in cubic systems.

Coordination Number The coordination number is the number of


atoms touching a particular atom, or the number of nearest neighbors for that
particular atom. This is one indication of how tightly and efficiently atoms are packed
together. For ionic solids, the coordination number of cations is defined as the
number of nearest anions. The coordination number of anions is the number of
nearest cations.

Chapter 1 13 | P a g e
FPEM1013 Introduction to Engineering Materials
In cubic structures containing only one atom per lattice point, atoms have a
coordination number related to the lattice structure. By inspecting the unit cells in
Figure 1-12, we see that each atom in the SC structure has a coordination number of
six, while each atom in the BCC structure has eight nearest neighbors. Each atom in
the FCC structure has a coordination number of twelve, which is the maximum.

Figure 1-12 Illustration of the coordination number in (a) SC and (b) BCC unit cells.
Six atoms touch each atom in SC, while eight atoms touch each atom in the BCC unit
cell.

The FCC arrangement represents a close-packed structure (CP) (i.e., the packing
fraction is the highest possible with atoms of one size). The SC and BCC structures are
relatively open. Metals with only metallic bonding are packed as efficiently as possible.
Metals with mixed bonding, such as iron, may have unit cells with less than the
maximum packing factor. No commonly encountered engineering metals or alloys
have the SC structure, although this structure is found in ceramic materials.

Density The theoretical density of a material can be calculated


using the properties of the crystal structure. The general formula is

(number of atoms/cell)(atomic mass)


Density ρ =
(volume of unit cell)(Avogadro constant)

If a material is ionic and consists of different types of atoms or ions, this formula will
have to be modified to reflect these differences.

Chapter 1 14 | P a g e
FPEM1013 Introduction to Engineering Materials
1.5.1 Points, Directions, and Planes in the Unit Cell
Coordinates of PointsWe can locate certain points, such as atom positions, in the
lattice or unit cell by constructing the right-handed coordinate system in Figure 1-13.
Distance is measured in terms of the number of lattice parameters we must move in
each of the x, y, and z coordinates to get from the origin to the point in question. The
coordinates are written as the three distances, with commas separating the numbers.

Figure 1-13 Coordinates of selected points in the unit cell. The number refers to the
distance from the origin in terms of lattice parameters.

1.5.2 Directions in the Unit Cell


Certain directions in the unit cell are of particular importance. Miller indices for
directions are the shorthand notation used to describe these directions. The
procedure for finding the Miller indices for directions is as follows:
• Using a right-handed coordinate system, determine the coordinates of two
points that lie on the direction.
• Subtract the coordinates of the “tail” point from the coordinates of the “head”
point to obtain the number of lattice parameters traveled in the direction of
each axis of the coordinate system.
• Clear fractions and or reduce the results obtained from the subtraction to
lowest integers.
• Enclose the numbers in square brackets [ ]. If a negative sign is produced,
represent the negative sign with a bar over the number.

Chapter 1 15 | P a g e
FPEM1013 Introduction to Engineering Materials
Example below illustrates a way of determining the Miller indices of directions.
Determine the Miller indices of directions A, B, and C in Figure 1-14.

SOLUTION
Direction A
• Two points are 1, 0, 0, and 0, 0, 0
• 1, 0, 0 - 0, 0, 0 = 1, 0, 0
• No fractions to clear or integers to reduce
• [100]

Direction B
• Two points are 1, 1, 1 and 0, 0, 0
• 1, 1, 1 - 0, 0, 0 = 1, 1, 1
• No fractions to clear or integers to reduce
• [111]

Figure 1-14 Crystallographic directions and coordinates

Direction C
• Two points are 0, 0, 1 and 1/2, 1, 0
• 0, 0, 1-1/2, 1,0 = -1/2, -1, 1
• 2 (-1/2 , -1, 1) = -1, -2, 2
• [1� 2� 2]

Significance of Crystallographic Directions Crystallographic directions are


used to indicate a particular orientation of a single crystal or of an oriented
polycrystalline material. Knowing how to describe these can be useful in many
applications. Metals deform more easily, for example, in directions along which atoms
are in closest contact. Another real-world example is the dependence of the magnetic
properties of iron and other magnetic materials on the crystallographic directions. It
is much easier to magnetize iron in the [100] direction compared to the [111] or [110]
directions. This is why the grains in Fe-Si steels used in magnetic applications (e.g.,
transformer cores) are oriented in the [100] or equivalent directions.

Chapter 1 16 | P a g e
FPEM1013 Introduction to Engineering Materials
1.5.3 Planes in the Unit Cell
Certain planes of atoms in a crystal also carry particular significance. For example,
metals deform along planes of atoms that are most tightly packed together. The
surface energy of different faces of a crystal depends upon the particular
crystallographic planes. This becomes important in crystal growth. In thin film
growth of certain electronic materials (e.g., Si or GaAs), we need to be sure the
substrate is oriented in such a way that the thin film can grow on a particular
crystallographic plane.

Miller indices are used as a shorthand notation to identify these important planes, as
described in the following procedure.
• Identify the points at which the plane intercepts the x, y, and z coordinates in
terms of the number of lattice parameters. If the plane passes through the
origin, the origin of the coordinate system must be moved to that of an
adjacent unit cell.
• Take reciprocals of these intercepts.
• Clear fractions but do not reduce to lowest integers.
• Enclose the resulting numbers in parentheses (). Again, negative numbers
should be written with a bar over the number.

The following example shows how Miller indices of planes can be obtained.
Determine the Miller indices of planes A, B, and C in Figure 1-15.

Figure 1-15 Crystallographic planes and intercepts.

SOLUTION
Plane A
• x = 1, y = 1, z = 1
1 1 1
• x = 1, y = 1, z = 1
• No fractions to clear
• (111)

Chapter 1 17 | P a g e
FPEM1013 Introduction to Engineering Materials
Plane B
• The plane never intercepts the z axis, so x = 1, y = 2, and z = ∞
1 1 1 1
• x = 1, y = 2, z = 0
1 1 1
• Clear fractions: x = 2, y = 1, z = 0
• (210)

Plane C
• We must move the origin, since the plane passes through 0, 0, 0. Let’s move
the origin one lattice parameter in the y-direction. Then, x = ∞, y = -1, and z
= ∞.
1 1 1
• x = 0, x = -1, z = 0
• No fractions to clear.
• (01� 0)

Chapter 1 18 | P a g e
FPEM1013 Introduction to Engineering Materials
Close-Packed Planes and Directions In examining the relationship
between atomic radius and lattice parameter, we looked for close-packed directions,
where atoms are in continuous contact. We can now assign Miller indices to these
close-packed directions, as shown in Table 1-3.

Table 1-3 Close-packed planes and directions

You may remember from the discussion on metallic crystal structures that both face-
centered cubic and hexagonal close-packed crystal structures have atomic packing
factors of 0.74, which is the most efficient packing of equal-sized spheres or atoms. In
addition to unit cell representations, these two crystal structures may be described
in terms of close-packed planes of atoms (i.e., planes having a maximum atom or
sphere-packing density); a portion of one such plane is illustrated in Figure 1-16a.
Both crystal structures may be generated by the stacking of these close-packed planes
on top of one another; the difference between the two structures lies in the stacking
sequence. Let the centers of all the atoms in one close-packed plane be labeled A.
Associated with this plane are two sets of equivalent triangular depressions formed
by three adjacent atoms, into which the next close-packed plane of atoms may rest.
Those having the triangle vertex pointing up are arbitrarily designated as B positions,
while the remaining depressions are those with the down vertices, which are marked
C in Figure 1-16a.

Chapter 1 19 | P a g e
FPEM1013 Introduction to Engineering Materials

Figure 1-16 (a) A portion of a close-packed plane of atoms; A, B, and C positions are
indicated. (b) The AB stacking sequence for close packed atomic planes.

A second close-packed plane may be positioned with the centers of its atoms over
either B or C sites; at this point both are equivalent. Suppose that the B positions are
arbitrarily chosen; the stacking sequence is termed AB, which is illustrated in Figure
1-16b. The real distinction between FCC and HCP lies in where the third close-packed
layer is positioned. For HCP, the centers of this layer are aligned directly above the
original A positions. This stacking sequence, ABABAB . . . , is repeated over and over.
These close-packed planes for HCP are (0001)-type planes, and the correspondence
between this and the unit cell representation is shown in Figure 1-17.

For the face-centered crystal structure, the centers of the third plane are situated over
the C sites of the first plane (Figure 1-18a). This yields an ABCABCABC . . . stacking
sequence; that is, the atomic alignment repeats every third plane. It is more difficult
to correlate the stacking of close-packed planes to the FCC unit cell. However, this
relationship is demonstrated in Figure 1-18b. These planes are of the (111) type; an
FCC unit cell is outlined on the upper left-hand front face of Figure 1-18b, in order to
provide a perspective.

Chapter 1 20 | P a g e
FPEM1013 Introduction to Engineering Materials

Figure 1-17 Close-packed plane stacking sequence for hexagonal close-packed.

Figure 1-18 (a) Close-packed stacking sequence for face-centered cubic. (b) A
corner has been removed to show the relation between the stacking of close-packed
planes of atoms and the FCC crystal structure; the heavy triangle outlines a (111)
plane.

Chapter 1 21 | P a g e
FPEM1013 Introduction to Engineering Materials
1.6 Single Crystals
For a crystalline solid, when the periodic and repeated arrangement of atoms is
perfect or extends throughout the entirety of the specimen without interruption, the
result is a single crystal. All unit cells interlock in the same way and have the same
orientation. Single crystals exist in nature, but they may also be produced artificially.
They are ordinarily difficult to grow, because the environment must be carefully
controlled. If the extremities of a single crystal are permitted to grow without any
external constraint, the crystal will assume a regular geometric shape having flat
faces, as with some of the gem stones; the shape is indicative of the crystal structure.
A photograph of a garnet single crystal is shown in Figure 1-19. Within the past few
years, single crystals have become extremely important in many of our modern
technologies, in particular electronic microcircuits, which employ single crystals of
silicon and other semiconductors.

Figure 1-19 Photograph of a garnet single crystal that was found in Tongbei, Fujian
Province, China.

1.7 Polycrystalline Materials


Most crystalline solids are composed of a collection of many small crystals or grains;
such materials are termed polycrystalline. Various stages in the solidification of a
polycrystalline specimen are represented schematically in Figure 1-20. Initially, small
crystals or nuclei form at various positions. These have random crystallographic
orientations, as indicated by the square grids. The small grains grow by the successive
addition from the surrounding liquid of atoms to the structure of each. The
extremities of adjacent grains impinge on one another as the solidification process
approaches completion. As indicated in Figure 1-20, the crystallographic orientation
varies from grain to grain. Also, there exists some atomic mismatch within the region
where two grains meet; this area, called a grain boundary.

Chapter 1 22 | P a g e
FPEM1013 Introduction to Engineering Materials

Figure 1-20 Schematic diagrams of the various stages in the solidification of a


polycrystalline material; the square grids depict unit cells. (a) Small crystallite nuclei.
(b) Growth of the crystallites; the obstruction of some grains that are adjacent to one
another is also shown. (c) Upon completion of solidification, grains having irregular
shapes have formed. (d) The grain structure as it would appear under the microscope;
dark lines are the grain boundaries.

1.8 Non-crystalline Solids


It has been mentioned that non-crystalline solids lack a systematic and regular
arrangement of atoms over relatively large atomic distances. Sometimes such
materials are also called amorphous (meaning literally without form), or supercooled
liquids, inasmuch as their atomic structure resembles that of a liquid. An amorphous
condition may be illustrated by comparison of the crystalline and non-crystalline
structures of the ceramic compound silicon dioxide (SiO2), which may exist in both
states. Figures 1-21a and 1-21b present two-dimensional schematic diagrams for
both structures of SiO2. Even though each silicon ion bonds to three oxygen ions for
both states, beyond this, the structure is much more disordered and irregular for the
non-crystalline structure. Whether a crystalline or amorphous solid forms depends
on the ease with which a random atomic structure in the liquid can transform to an
ordered state during solidification. Amorphous materials, therefore, are

Chapter 1 23 | P a g e
FPEM1013 Introduction to Engineering Materials
characterized by atomic or molecular structures that are relatively complex and
become ordered only with some difficulty. Furthermore, rapidly cooling through the
freezing temperature favors the formation of a non-crystalline solid, since little time
is allowed for the ordering process. Metals normally form crystalline solids, but some
ceramic materials are crystalline, whereas others, the inorganic glasses, are
amorphous. Polymers may be completely non-crystalline and semi-crystalline
consisting of varying degrees of crystallinity.

Figures 1-21 Two-dimensional schemes of the structure of (a) crystalline silicon


dioxide and (b) non-crystalline silicon dioxide.

Chapter 1 24 | P a g e

You might also like