You are on page 1of 148

Crystal Structure and Dislocations

K. Narayan Prabhu
Department of Metallurgical & Materials Engineering
National Institute of Technology, Karnataka, Surathkal

Bonds in Materials
The number of electrons in the outermost orbit of an
element decides the tendency of an element to go into a
bond with other elements. The most inert elements are
helium and inert gases like Ne, Ar, Kr, Xe, which have
the most stable structure.
Helium has 2 electrons in its first orbit, i.e.,K shell. The
maximum number of electrons in an orbit is given by 2n 2
where n is the orbit number. If 2 electrons are present in
the first orbit, the element (He) is stable and does not
have any tendency to form bond with any other element.
Likewise, in the second orbit, element having 8 electrons
(2.22 = 8) has stable structure called as
electron octet. Ne.

Bonds in Materials
A. Primary bonds
In elements containing more than 2 orbits of
electrons, (called as K & L shells, 2 in first and 8 in
second) also, 8 electrons in the outermost orbit
seems to be the most stable structure i.e., the
electron octet.
In all other elements, there is a tendency to attain
this stable structure of electron octet by forming
atomic bonds with other elements. This tendency
gives rise to atomic bonds.
These atomic bonds have bond energies of the order
of 50 1000 kJ/mol.

A. Primary bonds
1.Metallic Bond
Atomic community bond
Positive ions surrounded by electron cloud so that at any given
moment every atom would tend to have as many electrons
surrounding it as possible. Here all atoms together tend to
satisfy the electron octet structure for each of them.
Here, all electrons are free to move in the structure.
Free electron theory explains the good conductivity
( both thermal & electrical) of metals.
Another important property of plasticity is also connected with
metallic bond.
Plasticity is due mainly to slip phenomenon and slip process is
possible due to the identical surrounding of every atomic site
in the structure

A. Primary bonds
1.Metallic Bond
The red spheres represent
metal cations (positively
charged ions), and the
black moving spheres
represent electrons.
Metallic bonds are best
characterized by the
phrase a sea or a cloud of
electrons surrounding
metal cations

A. Primary bonds

2. Covalent bond
3. Ionic bonds
The covalent bond: Two chlorine atoms with 7 electrons in the outermost orbit
come together and form the bond. Two electrons are shared between each
chlorine to form octet structure around each. Similarly, carbon atoms form
diamond structure with carbon tetrahedra.
The ionic bond: Here, sodium donates its one electron to chlorine which accepts
this to complete its octet structure. As a result of this, Na becomes a positive
ion while chlorine becomes a negative ion and the ionic attraction between the
two forms the bond.

B. Secondary Bonds
These are bonds between molecules and the bond energies
are of the order of 1- 50 kJ/mol.
In many molecules, where hydrogen takes part in the
covalent bonding, the centres of the positive and negative
charges do not coincide as happens in the water molecule. A
net negative charge develops at the oxygen end and a net
positive charge at the two hydrogen ends. Due to this, a
permanent dipole moment exists in the water molecule.
The positive end of the molecule forms a bond with the
negative end of the neighboring molecule. Hydrogen
bonding
Inert gases form solid crystals at low temperatures due to
van der Waals bond. So also giant polymer molecules at
normal temperatures form solids whose melting points can
be quite high due to this bond.

Crystalline Structure of Metals

Crystalline
Structure

Quartz
Crystal

A crystal consists of identical structural units, consisting


of one or more atoms, which are regularly arranged
with respect to each other in space. A crystal is a solid in
which the constituent atoms,molecules, or ions are
packed in a regularly ordered, repeating pattern
extending in all three spatial dimensions.

Crystals form when they undergo a process of


solidification. Under ideal conditions, the result may
be a single crystal, where all of the atoms in the solid
fit into the same crystal structure. However, generally,
many crystals form simultaneously during
solidification, leading to a polycrystalline solid.

While the cooling process usually results in the generation


of a crystalline material, under certain conditions, the fluid
may be frozen in a noncrystalline state. In most cases, this
involves cooling the fluid so rapidly that atoms cannot
travel to their lattice sites before they lose mobility. A
noncrystalline material, which has no long range order, is
called an amorphous, vitreous, or glassy material.

Crystalline structures occur in all classes of materials,


with all types of chemical bonds. Almost all metal exists
in a polycrystalline state; amorphous or single-crystal
metals must be produced synthetically, often with great
difficulty. Ionically bonded crystals can form upon
solidification of salts, either from a molten fluid or when
it condenses from a solution. Covalently bonded crystals
are also very common, notable examples being diamond,
silica, and graphite.

Polymer materials generally will form crystalline


regions, but the lengths of the molecules usually
prevents complete crystallization. Weak Van der Waals
Forces can also play a role in a crystal structure; for
example, this type of bonding loosely holds together
the hexagonal-patterned sheets in graphite.

The term "crystal" has a precise meaning within materials


science. Colloquially "crystal" refers to solid objects that
exhibit well-defined and often pleasing geometric shapes.
In this sense of the word, many types of crystals are found
in nature. The shape of these crystals is dependent on the
types of molecular bonds between the atoms to determine
the structure, as well as on the conditions under which they
formed.
Crystallography is the scientific study of crystals and
crystal formation.

Short-Range Order versus


Long-Range Order
Short-range order - The regular and predictable
arrangement of the atoms over a short distance usually one or two atom spacings.
Long-range order (LRO) - A regular repetitive
arrangement of atoms in a solid which extends over
a very large distance.

(c) 2003 Brooks/Cole Publishing / Thomson Learning

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.

(a) Photograph of a silicon single


crystal. (b) Micrograph of a
polycrystalline stainless steel
showing grains and grain
boundaries

Liquid crystal display. These materials are amorphous


in one state and undergo localized crystallization in
response to an external electric field and are widely
used in liquid crystal displays.

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Classification of materials based on the type of atomic order.

Lattice, Unit Cells, Basis, and Crystal


Structures

Lattice - A collection of points that divide


space into smaller equally sized segments.
Basis - A group of atoms associated with a
lattice point.
Unit cell - A subdivision of the lattice that still
retains the overall characteristics of the
entire lattice.
Atomic radius - The apparent radius of an
atom, typically calculated from the
dimensions of the unit cell, using closepacked directions (depends upon
coordination number).
Packing factor - The fraction of space in a unit
cell occupied by atoms.

The fourteen
types of Bravais
lattices grouped in
seven crystal
systems.

(c) 2003 Brooks/Cole Publishing / Thomson Learning

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

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Lattice parameters describe the size and shape of the


unit cell

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

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Determining the Number of Lattice Points


in Cubic Crystal Systems
Determine the number of lattice points per cell in the cubic crystal
systems. If there is only one atom located at each lattice point,
calculate the number of atoms per unit cell.
SOLUTION
In the SC unit cell: lattice point / unit cell = (8 corners)1/8
=1
In BCC unit cells: lattice point / unit cell
= (8 corners)1/8 + (1 center)(1) = 2
In FCC unit cells: lattice point / unit cell
= (8 corners)1/8 + (6 faces)(1/2) = 4
The number of atoms per unit cell would be 1, 2, and 4, for
the simple cubic, body-centered cubic, and face-centered
cubic, unit cells, respectively.

Determining the Relationship between


Atomic Radius and Lattice Parameters
Determine the relationship between the atomic radius
and the lattice parameter in SC, BCC, and FCC
structures when one atom is located at each lattice
point.

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

The relationships between the atomic radius and the Lattice


parameter in cubic systems

We find that atoms touch along the edge of the cube in


SC structures.

a0

2r

In BCC structures, atoms touch along the body diagonal.


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. There are four atomic radii along
this lengthtwo radii from the face-centered atom and
one radius from each corner, so:

a0

4r

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

Illustration of coordinations in (a) SC and (b) BCC unit cells.


Six atoms touch each atom in SC, while the eight atoms
touch each atom in the BCC unit cell.

Calculating the Packing Factor


Calculate the packing factor for the FCC cell.
Example 3.3 SOLUTION
In a FCC cell, there are four lattice points per cell; if there
is one atom per lattice point, there are also four atoms per
cell. The volume of one atom is 4r3/3 and the volume of
the unit cell is
.

3
0

4
(4 atoms/cell)( r 3 )
3
Packing Factor

a03
Since, for FCC unit cells, a 0 4r/
4
(4)( r 3 )
3
Packing Factor

3
( 4r / 2 )

0.74
18

Determining the Density of BCC Iron


Determine the density of BCC iron, which has a lattice
parameter of 0.2866 nm.
Example 3.4 SOLUTION
Atoms/cell = 2, a0 = 0.2866 nm = 2.866 10-8 cm
Atomic mass = 55.847 g/mol
Volume of unit cell =
cm3/cell

3
=
0

(2.866 10-8 cm)3 = 23.54 10-24

Avogadros number NA = 6.02 1023 atoms/mol

(number of atoms/cell)(atomic mass of iron)


Density
(volume of unit cell)(Avogadro' s number)
(2)(55.847)
3

7
.
882
g
/
cm
(23.54 10 24 )(6.02 10 23 )

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

Figure 3.16 The hexagonal close-packed (HCP) structure


(left) and its unit cell.

Points, Directions, and Planes in the


Unit Cell
Miller indices - A shorthand notation to describe
certain crystallographic directions and planes in a
material. Denoted by (),[ ] brackets. A negative
number is represented by a bar over the number.
Directions of a form - Crystallographic directions that
all have the same characteristics, although their
sense is different. Denoted by {},< > brackets.
Repeat distance - The distance from one lattice point
to the adjacent lattice point along a direction.
Linear density - The number of lattice points per unit
length along a direction.
Packing fraction - The fraction of a direction (linearpacking fraction) or a plane (planar-packing factor)
that is actually covered by atoms or ions.

Coordinates of selected points in the


unit cell. The number refers to the
distance from the origin in terms of
lattice parameters.

Determining Miller Indices of Directions


Determine the Miller indices of directions A, B,
and C in Figure.

Figure Crystallographic
directions and coordinates (for
Example 3.7).

(c) 2003 Brooks/Cole Publishing /


Thomson Learning

SOLUTION
Direction A
1. Two points are 1, 0, 0, and 0, 0, 0
2. 1, 0, 0, -0, 0, 0 = 1, 0, 0
3. No fractions to clear or integers to reduce
4. [100]
Direction B
1. Two points are 1, 1, 1 and 0, 0, 0
2. 1, 1, 1, -0, 0, 0 = 1, 1, 1
3. No fractions to clear or integers to reduce
4. [111]
Direction C
1. Two points are 0, 0, 1 and 1/2, 1, 0
2. 0, 0, 1 -1/2, 1, 0 = -1/2, -1, 1
3. 2(-1/2, -1, 1) = -1, -2, 2

4. [ 1 2 2]

(c) 2003 Brooks/Cole


Publishing / Thomson
Learning

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Figure 3.20 Equivalency of crystallographic directions of a


form in cubic systems.

Another way for characterizing equivalent directions is


by the repeat distances

Figure 3.21
Determining the
repeat distance,
linear density,
and packing
fraction for
[110] direction
in FCC copper.

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Linear density =2Repeat Distances/0.51125nm


= 3.91 lattice points/nm

Determining Miller Indices of Planes

Determine the Miller indices of planes A, B, and C in


Figure 3.22.

Figure 3.22
Crystallographic planes
and intercepts (for
Example 3.8)

(c) 2003 Brooks/Cole Publishing /


Thomson Learning

SOLUTION
Plane A
1. x = 1, y = 1, z = 1
2.1/x = 1, 1/y = 1,1 /z = 1
3. No fractions to clear
4. (111)
Plane B

(c) 2003
Brooks/Cole
Publishing /
Thomson
Learning

1. The plane never intercepts the z axis, so x = 1, y = 2, and


z=
2.1/x = 1, 1/y =1/2, 1/z = 0
3. Clear fractions:
1/x = 2, 1/y = 1, 1/z = 0
4. (210)
Plane C
1. We must move the origin, since the plane passes through
0, 0, 0. Lets move the origin one lattice parameter in the
, y = -1, and z =
y-direction. Then, x =
2.1/x = 0, 1/y = 1, 1/z = 0
3. No fractions to clear.

4. (0 1 0)

Calculating the Planar Density and Packing


Fraction
Calculate the planar density and planar packing
fraction for the (010) and (020) planes in simple cubic
polonium, which has a lattice parameter of 0.334 nm.
Figure 3.23 The
planer densities of
the (010) and
(020) planes in SC
unit cells are not
identical (for
Example 3.9).
(c) 2003 Brooks/Cole Publishing /
Thomson Learning

SOLUTION
The total atoms on each face is one. The planar density
is:

atom per face


1 atom per face
Planar density (010)

2
area of face
(0.334)
2

14

8.96 atoms/nm 8.96 10 atoms/cm


The planar packing fraction is given by:

area of atoms per face


(1 atom) (r )
Packing fraction (010)

2
area of face
(a 0)

r 2

0.79
2
( 2r )

However, no atoms are centered on the (020) planes.


Therefore, the planar density and the planar packing
fraction are both zero. The (010) and (020) planes are
not equivalent!

Drawing Direction and Plane


Draw (a) the [1 2 1] direction and (b) the
cubic unit cell.

[ 2 10] plane in a

Figure 3.24
Construction
of a (a)
direction and
(b) plane
within a unit
cell (for
Example 3.10)

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

SOLUTION
a. Because we know that we will need to move in the
negative y-direction, lets locate the origin at 0, +1, 0.
The tail of the direction will be located at this new
origin. A second point on the direction can be
determined by moving +1 in the x-direction, 2 in the ydirection, and +1 in the z direction [Figure 3.24(a)].
b. To draw in the [ 2 10] plane, first take reciprocals of
the indices to obtain the intercepts, that is:
x = 1/-2 = -1/2 y = 1/1 = 1 z = 1/0 =

Since the x-intercept is in a negative direction, and we


wish to draw the plane within the unit cell, lets move
the origin +1 in the x-direction to 1, 0, 0. Then we can
locate the x-intercept at 1/2 and the y-intercept at +1.
The plane will be parallel to the z-axis [Figure 3.24(b)].

Figure 3.25 Miller-Bravais


indices are obtained for
crystallographic planes in
HCP unit cells by using a
four-axis coordinate system.
The planes labeled A and B
and the direction labeled C
and D are those discussed in
Example 3.11.

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

Determining the Miller-Bravais Indices for


Planes and Directions
Determine the Miller-Bravais indices for planes A and B
and directions C and D in Figure 3.25.

Figure 3.25 Miller-Bravais


indices are obtained for
crystallographic planes in
HCP unit cells by using a
four-axis coordinate
system. The planes labeled
A and B and the direction
labeled C and D are those
discussed in Example 3.11.

(c) 2003 Brooks/Cole Publishing /


Thomson Learning

SOLUTION
Plane A
1. a1 = a2 = a3 = , c = 1
2. 1/a1 = 1/a2 = 1/a3 = 0, 1/c = 1
3. No fractions to clear
4. (0001)
Plane B
1. a1 = 1, a2 = 1, a3 = -1/2, c = 1
2. 1/a1 = 1, 1/a2 = 1, 1/a3 = -2, 1/c = 1
3. No fractions to clear
(11 21)
4.

(c) 2003 Brooks/Cole


Publishing / Thomson
Learning

Direction C
1. Two points are 0, 0, 1 and 1, 0, 0.
2. 0, 0, 1, -1, 0, 0 = 1, 0, 1
3. No fractions to clear or integers to reduce.
[ 1 01] or [2113]
4.

Example 3.11 SOLUTION (Continued)


Direction D
1. Two points are 0, 1, 0 and 1, 0, 0.
2. 0, 1, 0, -1, 0, 0 = -1, 1, 0
3. No fractions to clear or integers to reduce.
4. [ 1 10] or [ 1 100]

h = (2h-k)/3
k= (2k+h)/3
i= -(h+k)/3
l= l
(c) 2003 Brooks/Cole
Publishing / Thomson
Learning

Defects in Crystals
Point defect, such as vacancy, interstitial
atom, impurity
Line defect, called dislocation
Planar imperfection, such as grain
boundaries
Volume or bulk imperfections, such as
voids/inclusions

Point Defects

Point defects - Imperfections, such as vacancies, that are located


typically at one (in some cases a few) sites in the crystal.
Extended defects - Defects that involve several atoms/ions and
thus occur over a finite volume of the crystalline material (e.g.,
dislocations, stacking faults, etc.).
Vacancy - An atom or an ion missing from its regular
crystallographic site.
Interstitial defect - A point defect produced when an atom is
placed into the crystal at a site that is normally not a lattice point.
Substitutional defect - A point defect produced when an atom is
removed from a regular lattice point and replaced with a different
atom, usually of a different size.

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Figure 4.1 Point defects: (a) vacancy, (b) interstitial atom, (c) small
substitutional atom, (d) large substitutional atom, (e) Frenkel defect,
(f) Schottky defect. All of these defects disrupt the perfect
arrangement of the surrounding atoms.

Point Defects in a Single Crystal Lattice


Extra atom

Foreign
atom

Missing atom

Extra atom

Schematic illustration of types of defects in a singlecrystal lattice: self-interstitial, vacancy, interstitial,


and substitutional.

The Effect of Temperature on


Vacancy Concentrations
Calculate the concentration of vacancies in copper at room
temperature (25oC). What temperature will be needed to heat
treat copper such that the concentration of vacancies produced
will be 1000 times more than the equilibrium concentration of
vacancies at room temperature? Assume that 20,000 cal are
required to produce a mole of vacancies in copper.
Example 4.1 SOLUTION
The lattice parameter of FCC copper is 0.36151 nm. The basis
is 1, therefore, the number of copper atoms, or lattice points,
per cm3 is:

4 atoms/cell
22
3
n

8
.
47

10
copper
atoms/cm
8
3
(3.6151 10 cm)

Example 4.1 SOLUTION (Continued)


At room temperature, T = 25 + 273 = 298 K:

Q
n n exp

RT

8.47 10

atoms
. exp
3
cm

22

1.815 10 vacancies/cm

cal

20,000

mol

cal
1.987
298K
mol K

We could do this by heating the copper to a temperature at


which this number of vacancies forms:
11

n 1.815 10
22

Q
n exp

RT
o

(8.47 10 ) exp(20,000 /(1.987 T )), T 102 C

Vacancy Concentrations in Iron

Determine the number of vacancies needed for a BCC iron


crystal to have a density of 7.87 g/cm3. The lattice parameter
of the iron is 2.866 10-8 cm.
Example 4.2 SOLUTION
The expected theoretical density of iron can be calculated from
the lattice parameter and the atomic mass.

SOLUTION (Continued)
Lets calculate the number of iron atoms and vacancies
that would be present in each unit cell for the required
density of 7.87 g/cm3:

Or, there should be 2.00 1.9971 = 0.0029 vacancies per


unit cell. The number of vacancies per cm3 is:

Sites for Carbon in Iron


In FCC iron, carbon atoms are located at octahedral sites at
the center of each edge of the unit cell (1/2, 0, 0) and at
the center of the unit cell (1/2, 1/2, 1/2). In BCC iron,
carbon atoms enter tetrahedral sites, such as 1/4, 1/2, 0.
The lattice parameter is 0.3571 nm for FCC iron and 0.2866
nm for BCC iron. Assume that carbon atoms have a radius
of 0.071 nm. (1) Would we expect a greater distortion of
the crystal by an interstitial carbon atom in FCC or BCC
iron? (2) What would be the atomic percentage of carbon in
each type of iron if all the interstitial sites were filled?

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

Figure (a) The location of the , , 0 interstitial site in BCC


metals, showing the arrangement of the normal atoms and
the interstitial atom (b) , 0, 0 site in FCC metals, (for
Example 4.3). (c) Edge centers and cube centers are some of
the interstitial sites in the FCC structure (Example 4.3).

SOLUTION
1. We could calculate the size of the interstitial site at the
1/4, 1/2, 0 location with the help of Figure (a). The radius
RBCC of the iron atom is:

From Figure 4.2(a), we find that:

For FCC iron, the interstitial site such as the 1/2, 0, 0


lies along 100 directions. Thus, the radius of the iron
atom and the radius of the interstitial site are [Figure
4.2(b)]:

SOLUTION (Continued)

The interstitial site in the BCC iron is smaller than the


interstitial site in the FCC iron. Although both are smaller
than the carbon atom, carbon distorts the BCC crystal
structure more than the FCC crystal. As a result, fewer
carbon atoms are expected to enter interstitial positions in
BCC iron than in FCC iron.

SOLUTION (Continued)
2. We can find a total of 24 interstitial sites of the type 1/4,
1/2, 0; however, since each site is located at a face of the
unit cell, only half of each site belongs uniquely to a single
cell. Thus:
(24 sites)(1/2) = 12 interstitial sites per unit cell
Atomic percentage of carbon in BCC iron would be:

In FCC iron, the number of octahedral interstitial sites is:


(12 edges) (1/4) + 1 center = 4 interstitial sites per unit cell
Atomic percentage of carbon in BCC iron would be:

Other Point Defects


Interstitialcy - A point defect caused when a normal
atom occupies an interstitial site in the crystal.
Frenkel defect - A pair of point defects produced when
an ion moves to create an interstitial site, leaving behind
a vacancy.
Schottky defect - A point defect in ionically bonded
materials. In order to maintain a neutral charge, a
stoichiometric number of cation and anion vacancies
must form.
Krger-Vink notation - A system used to indicate point
defects in materials. The main body of the notation
indicates the type of defect or the element involved.

(c) 2003 Brooks/Cole Publishing / Thomson


Learning

Figure 4.3 When a divalent cation replaces a monovalent


cation, a second monovalent cation must also be removed,
creating a vacancy.

Dislocations
Dislocation - A line imperfection in a crystalline material.
Screw dislocation - A dislocation produced by skewing a
crystal so that one atomic plane produces a spiral ramp
about the dislocation.
Edge dislocation - A dislocation introduced into the
crystal by adding an extra half plane of atoms.
Mixed dislocation - A dislocation that contains partly
edge components and partly screw components.
Slip - Deformation of a metallic material by the
movement of dislocations through the crystal.

Slip Direction Direction in which the dislocation moves


Slip Plane
Slip System

(c) 2003 Brooks/Cole Publishing / Thomson Learning

Figure 4.4 the perfect crystal (a) is cut and sheared one atom
spacing, (b) and (c). The line along which shearing occurs is a
screw dislocation. A Burgers vector b is required to close a loop
of equal atom spacings around the screw dislocation.

Figure 4.5 The perfect crystal in (a) is cut and an extra


plane of atoms is inserted (b). The bottom edge of the
extra plane is an edge dislocation (c). A Burgers
vector b is required to close a loop of equal atom
spacings around the edge dislocation. (Adapted from
J.D. Verhoeven, Fundamentals of Physical Metallurgy,
Wiley, 1975.)

Figure 4.6 A mixed dislocation. The


screw dislocation at the front face
of the crystal gradually changes to
an edge dislocation at the side of
the crystal. (Adapted from W.T.
Read, Dislocations in Crystals.
McGraw-Hill, 1953.)

Figure Schematic of slip line, slip plane, and slip (Burgers)


vector for (a) an edge dislocation and (b) for a screw
dislocation. (Adapted from J.D. Verhoeven, Fundamentals
of Physical Metallurgy, Wiley, 1975.)

Figure 4.7 Schematic of slip line, slip plane, and slip


(Burgers) vector for (a) an edge dislocation and (b) for a
screw dislocation. (Adapted from J.D. Verhoeven,
Fundamentals of Physical Metallurgy, Wiley, 1975.)

The Burgers vector of a dislocation is a crystal vector,


specified by Miller indices, that quantifies the difference
between the distorted lattice around the dislocation and
the perfect lattice. Equivalently, the Burgers vector
denotes the direction and magnitude of the atomic
displacement that occurs when a dislocation moves.
Trace around the end
of the dislocation
plane to form a closed
loop. Record the
number of lattice
vectors travelled along
each side of the loop

In a perfect lattice, trace out the same path, moving the


same number of lattice vectors along each direction as
before. This loop will not be complete, and the closure
failure is the Burgers vector

Dislocations in 3D
In three dimensions, the nature of a dislocation as a line defect
becomes apparent. The dislocation line runs along the core of
the dislocation, where the distortion with respect to the perfect
lattice is greatest.
There are two types of three-dimensional dislocation. An edge
dislocation has its Burgers vector perpendicular to the
dislocation line. Edge dislocations are easiest to visualise as an
extra half-plane of atoms. A screw dislocation is more complex
- the Burgers vector is parallel to the dislocation line. Mixed
dislocations also exist, where the Burgers vector is at some
acute angle to the dislocation line.

EDGE
DISLOCATION

SCREW
DISLOCATION

EDGE DISLOCATION

SCREW DISLOCATION

When a dislocation moves under an applied shear


stress:
individual atoms move in directions parallel to the
Burgers vector
the dislocation moves in a direction perpendicular to
the dislocation line.
An edge dislocation therefore moves in the direction of
the Burgers vector, whereas a screw dislocation moves
in a direction perpendicular to the Burgers vector.

The screw dislocation 'unzips' the lattice as it moves


through it, creating a 'screw' or helical arrangement of
atoms around the core.
The ease of dislocation glide is partly determined by the
degree of distortion (with respect to the perfect lattice)
around the dislocation core. When the distortion is spread
over a large area, the dislocation is easy to move. Such
dislocations are known as wide dislocations, and exist in
ductile metals.
The width of a dislocation (W) gives a measure of the
degree of disruption a dislocation creates with respect to
the perfect lattice.

When w is several atomic spacings in dimension, the


dislocation is wide;
if W is of the order of one or two atomic spacings, it is
narrow.
Dislocation glide occurs most easily in wide dislocations
- these are found in simple metals with simple closepacked crystal structures, hence these materials are
ductile .
Ceramics, for example, tend to have narrow
dislocations, and are hard and brittle as a result.

During slip, a dislocation moves from one set of


surroundings to an identical set of surroundings.
The Peierls-Nabbaro stress is required to move the
dislocation from one equilibrium location to another.

c. exp( kd / b)
is the shear stress required to move the dislocation, d
is the interplanar distance between the adjacent slip
planes, b is the Burgers Vector
The Dislocation moves in slip systems that require the
least expenditure of energy.

he stress required to cause the dislocation increases exponentially


with the length of the Burgers Vector. Thus the slip direction
should have small repeat distance or high linear density. The Close
packed planes in metals satisfy this condition.
he stress required to cause the dislocation to move decreases
exponentially with the interplanar spacing of the slip planes. Slip
occurs between planes that are far apart.
3. Dislocations do not move easily in materials which have covalent
bonds.
4. Ceramics are also resistant to slip. Movement of a dislocation
disrupts the charge balance around cations and anions, requiring
that bonds between anions and cations are broken. During slip,
ions with a like charge must pass close together causing repulsion.
Further b is larger than in metals. Brittle failure will occur before
the dislocation can move.

Example 4.8
Burgers Vector Calculation
Calculate the length of the Burgers vector in copper.
(c) 2003 Brooks/Cole Publishing / Thomson
Learning

Figure 4.10
(a) Burgers
vector for FCC
copper. (b)
The atom
locations on a
(110) plane in
a BCC unit cell
(for example
4.8 and 4.9,
respectively)

The magnitude of Burgers Vector is the repeat distance between


the adjacent atoms along the highest atomic density direction

Example 4.8 SOLUTION


Copper has an FCC crystal structure. The lattice
parameter of copper (Cu) is 0.36151 nm. The closepacked directions, or the directions of the Burgers vector,
are of the form 110 . The repeat distance along the 110
directions is one-half the face diagonal, since lattice points
are located at corners and centers of faces [Figure
4.10(a)].

The length of the Burgers vector, or the repeat distance, is:


b = 1/2(0.51125 nm) = 0.25563 nm

Identification of Preferred Slip Planes


The planar density of the (112) plane in BCC iron is 9.94 1014
atoms/cm2. Calculate (1) the planar density of the (110) plane
and (2) the interplanar spacings for both the (112) and (110)
planes. On which plane would slip normally occur?
(c) 2003 Brooks/Cole Publishing / Thomson
Learning

Figure 4.10
(a) Burgers
vector for FCC
copper. (b)
The atom
locations on a
(110) plane in
a BCC unit cell
(for example
4.8 and 4.9,
respectively)

Example 4.9 SOLUTION


1. The planar density is:

2. The interplanar spacings are:

The planar density and interplanar spacing of the (110)


plane are larger than those for the (112) plane; therefore,
the (110) plane would be the preferred slip plane.

Observing Dislocations
Etch pits - Tiny holes created at areas where dislocations
meet the surface. These are used to examine the
presence and number density of dislocations.
Slip line - A visible line produced at the surface of a
metallic material by the presence of several thousand
dislocations.
Slip band - Collection of many slip lines, often easily
visible.

Significance of Dislocations
Plastic deformation refers to irreversible deformation or
change in shape that occurs when the force or stress
that caused it is removed.
Elastic deformation - Deformation that is fully recovered
when the stress causing it is removed.
Dislocation density - The total length of dislocation line
per cubic centimeter in a material.

An annealed material has a dislocation density of approximately


108 to 1010 /cm2, while cold
worked material has a density of approximately 1016 /cm2.

Dislocation motion along a crystallographic direction is


called glide or slip. There must be a local shear stress in
an appropriate direction on the dislocation for glide to
occur. Dislocation glide allows plastic deformation to
occur at a much lower stress than would be required to
move a whole plane of atoms past another.
The stress required to cause slip by moving entire planes
past one another, and the stress required to cause slip by
dislocation motion can be estimated. The stress required
for slip is much lower when the mechanism of slip is
dislocation motion. Slip does occur by dislocation motion.

ANIMATION OF SLIP BY DISLOCATION GLIDE

ANIMATION OF SLIP BY MOVEMENT OF


WHOLE LATTICE PLANES

Movement of an Edge Dislocation

Movement of an edge dislocation across the crystal lattice under


a shear stress. Dislocations help explain why the actual strength
of metals is much lower than that predicted by theory.
Requires lower shear stress to cause slip on a plane than in a perfect
lattice:
1. Earthworm moves forward through a hump that starts at the tail and moves
toward the head;
2. Moving a large carpet by forming a hump at one end and moving it to the
other end. The force required to move a carpet is much less than that required
to slide the carpet along the floor.

Metals - Slip & deformation

1. Slip on atomic planes without dislocations


2. Slip with dislocation low stress process

A sessile dislocation is one for which the


burgers vector is not in the direction of slip so it
wont move when a force is applied. Glissile
dislocations will move under the application of
a force their burgers vector is in the direction of
slip. Sessile dislocations can form by
multiple vacancies moving to the same area to
produce a dislocation.

Metals - deformation
Metals are malleable (can be flattened) and ductile

(can be drawn into wires) because of the way the


metal cations and electrons can "flow" around each
other, without breaking the crystal structure.
Due to the above, metals cannot be fractured without
undergoing extensive deformation. The amount of
energy needed to fracture them (called toughness) is
very high compared to ceramics. This energy is a sum
of energy for plastic deformation and energy needed
to create new fracture surface, the proportions of each
being approx. 100:1 respectively.

Metals - Slip & deformation


Plastic deformation occurs at stresses much smaller
than the theoretical strength of perfect crystals
considering the force needed to break all the bonds
in the slip plane.
The presence of line defects called as dislocations
enable the metallic crystals to deform at low stresses
since these dislocations move by breaking a small
fraction of the bonds in the slip plane at a time.

Slip & deformation in metals


Deformation in metals
occurs mainly by slip.
On the left, is a
schematic sketch of slip
in metals.
On the right, is a photo
of deformed single
crystal of zinc.

Section 4.6
Schmids Law
Schmids law -The relationship between shear stress, the
applied stress, and the orientation of the slip system
that is, cos cos
Critical resolved shear stress - The shear stress required
to cause a dislocation to move and cause slip.

(c)2003 Brooks/Cole, a division of Thomson Learning, Inc. Thomson Learning is a


trademark used herein under license.

Figure 4.14 (a) A resolved shear stress is produced on a slip


system. (Note: ( + ) does not have to be 90.) (b)
Movement of dislocations on the slip system deforms the
material. (c) Resolving the force.

Slip & deformation in metals


Critical resolved shear
stress(CRSS) for slip:
Slip begins when the shearing
stress on the slip plane in the slip
direction reaches a threshold
value called CRSS. From figure,
CRSS = P cos = P cos cos
A / cos
A
Note when = 0, = 90o
and when = 0, = 90o.
In such cases, CRSS is zero and
slip cannot take place.

Calculation of Resolved Shear Stress


Apply the Schmids law for a situation in which the single
crystal is at an orientation so that the slip plane is
perpendicular to the applied tensile stress.
(c) 2003 Brooks/Cole Publishing
/ Thomson Learning

When the slip plane is


perpendicular to the
applied stress , the angle
is 90 and no shear
stress is resolved.

SOLUTION
Suppose the slip plane is perpendicular to the
applied stress , as in Figure 4.15. Then, = 0o, =
90o, cos = 0, and therefore r = 0. As noted before, the
angles f and l can but do not always add up to 90o. Even
if the applied stress s is enormous, no resolved shear
stress develops along the slip direction and the
dislocation cannot move. Slip cannot occur if the slip
system is oriented so that either or is 90o.

Influence of Crystal Structure


Critical Resolved Shear Stress
Shear stress required for breaking enough metallic
bonds in order for slip to occur.
Number of Slip Systems
Cross-slip - A change in the slip system of a
dislocation.

Cross slip only occurs for screw dislocations, and involves

the dislocation moving off one slip plane onto


another slip plane in the same family of planes. No
cross slip can occur in HCP because the slip planes are parallel
and not intersecting. HCP metals hence tend to remain brittle.

Ductility of HCP Metal Single Crystals and


Polycrystalline Materials

A single crystal of magnesium (Mg), which has a HCP crystal


structure, can be stretched into a ribbon-like shape four to six
times its original length. However, polycrystalline Mg and other
metals with a HCP structure show limited ductilities. Use the
values of critical resolved shear stress for metals with different
crystal structures and the nature of deformation in
polycrystalline materials to explain this observation.

SOLUTION
From Table 4-2, we note that for HCP metals such as Mg,
the critical resolved shear stress is low (50100 psi). We
also note that slip in HCP metals will occur readily on the
basal planethe primary slip plane. When a single crystal
is deformed, assuming the basal plane is suitably
oriented with applied stress, a very large deformation can
occur. This explains why single crystal Mg can be
stretched into a ribbon four to six times the original size.
When we have a polycrystalline Mg, the
deformation is not as simple. Each crystal must deform
such that the strain developed in any one crystal is
accommodated by its neighbors. In HCP metals, there are
no intersecting slip systems, thus dislocations cannot
glide over from one slip plane in one crystal (grain) onto
another slip plane in a neighboring crystal. As a result,
polycrystalline HCP metals such as Mg show limited
ductility.

Importance of Defects
Effect on Mechanical Properties via Control of the Slip
Process
Strain Hardening
Solid-Solution Strengthening
Grain-Size Strengthening
Effects on Electrical, Optical, and Magnetic Properties

(c)2003 Brooks/Cole, a division of Thomson Learning, Inc. Thomson Learning is a trademark used herein under license.

Figure 4.22 If the dislocation at point A moves to the


left, it is blocked by the point defect. If the dislocation
moves to the right, it interacts with the disturbed lattice
near the second dislocation at point B. If the dislocation
moves farther to the right, it is blocked by a grain
boundary.

Design/Materials Selection for


a Stable Structure
We would like to produce a bracket to hold ceramic bricks
in place in a heat-treating furnace. The bracket should be
strong, should possess some ductility so that it bends
rather than fractures if overloaded, and should maintain
most of its strength up to 600oC. Design the material for
this bracket, considering the various crystal imperfections
as the strengthening mechanism.
SOLUTION
In order to serve up to 600oC, the bracket should not
be produced from a polymer material. Instead, a
metal or ceramic would be considered.

SOLUTION (Continued)
In order to have some ductility, dislocations must move
and cause slip. Because slip in ceramics is difficult, the
bracket should be produced from a metallic material.
We might add carbon to the iron as interstitial
atoms or substitute vanadium atoms for iron atoms at
normal lattice points. These point defects continue to
interfere with dislocation movement and help to keep the
strength stable.
Of course, other design requirements may be
important as well. For example, the steel bracket may
deteriorate by oxidation or may react with the ceramic
brick.

Etching to Reveal Dislocation Motion


Movement of dislocations in
LiF.
A are sessile dislocations &
B are glissile dislocations.
When a stress is applied the
glissile dislocation slides, hence,
when it is etched again it leaves
a new trace where as the old
trace gets wider, but not deeper.
Sessile dislocations do not move
and the etch pits get deeper and
wider on repeated etching.

Strain-Hardening Mechanisms
Frank-Read source - A pinned dislocation that,
under an applied stress, produces additional
dislocations. This mechanism is at least partly
responsible for strain hardening.

Figure 7.5 The Frank-Read


source can generate
dislocations. (a) A
dislocation is pinned at its
ends by lattice defects. (b)
As the dislocation continues
to move, the dislocation
bows, eventually bending
back on itself. (c) finally the
dislocation loop forms, and
(d) a new dislocation is
created. (e) Electron
micrograph of a Frank-Read
source (330,000). (Adapted
from Brittain, J., Climb
Sources in Beta PrimeNiAl, Metallurgical
Transactions, Vol. 6A, April
1975.)

Frank-Read source

Slip & deformation in metals


Bauschinger effect: Dislocations pile up on slip planes at
barriers in the crystal. The pile-ups produce a back
stress which opposes the applied stress on the slip plane.
In the figure, the crystal is strained to point O,
unloaded, and then reloaded in the direction opposite to
the original slip direction. Note that on reloading, the
crystal yields at a lower shear stress than earlier. This
lowering of the yield stress is called Bauschinger effect.

Two Plastic Deformation Mechanisms for


Single Crystals
a)Slipping & b)Twinning
a) Slip: Permanent deformation (also called plastic
deformation) of a single crystal subjected to a shear
stress: (a) structure before deformation; and (b)
permanent deformation by slip. The size of the b/a ratio
influences the magnitude of the shear stress required to
cause slip.

Two Plastic Deformation Mechanisms for


Single Crystals
b) Twinning : (a) Permanent deformation of a single
crystal under a tensile load. Note that the slip planes
tend to align themselves in the direction of the pulling
force. This behaviour can be simulated using a deck of
cards with a rubber band around them. (b) Twinning in
a single crystal in tension.

Slipping
Slipping of 1 atom plane over an adjacent
plane.
Require certain amount of shear stress to
undergo plastic (permanent) deformation.
The figure shows slip
lines in copper at a
magnification of 500X
in an optical microscope.

Twinning
Portion of the crystal forms a mirror image of itself
across the plane of twinning.
Usually occur in hcp and bcc metals by plastic
deformation; occur in fcc by annealing.

Twinning
Microstructures of twins: a) Neumann bands in iron.
b) mechanical twins in zinc produced by polishing.
c) annealing twins in gold silver alloy.

Slip Lines and Slip Bands


Schematic illustration of
slip lines and slip bands in
a single crystal (grain)
subjected to a shear stress.
A slip band consists of a
number of slip planes. The
crystal at the center of the
upper illustration is an
individual grain surrounded
by other grains.
Slip System
Combination of a slip
plane & slip direction.
Slip system >= 5, Ductile

Body-Centered Cubic Crystal Structure

The body-centered cubic (bcc) crystal structure: (a) hard-ball


model; (b) unit cell; and (c) single crystal with many unit cells.
Source: W. G. Moffatt, et al., The Structure and Properties of
Materials, Vol. 1, John Wiley & Sons, 1976.

Slip System of BCC


In Body-centered-cubic (BCC):
48 possible slip systems
Probability is high for externally applied shear
stress to cause slip
However b/a is high, so shear stress required is
high
Metal with bcc structure generally have good
strength & moderate ductility.
Eg., molybdenum,and tungsten

Face-Centered Cubic Crystal Structure

The face-centered cubic (fcc) crystal structure: (a) hard-ball


model; (b) unit cell; and (c) single crystal with many unit cells.
Source: W. G. Moffatt, et al., The Structure and Properties of
Materials, Vol. 1, John Wiley & Sons, 1976.

Slip System of FCC


In face-centered-cubic(FCC):
12 Slip systems
Probability for slip is moderate
b/a is relatively low, so required shear stress is
low
FCC Metal generally have moderate strength
& good ductility
Eg.Aluminum,Copper,gold,and silver

Hexagonal Closed-Packed Crystal Structure


The hexagonal closepacked (hcp) crystal
structure:
(a) unit cell; and
(b) single crystal with
many unit cells.
Source: W. G.
Moffatt, et al., The
Structure and
Properties of
Materials, Vol. 1,
John Wiley & Sons,
1976.

Slip System of HCP


Hexagonal-close-packed (hcp):
3 slip systems
Low probability of slip
However more slip systems become available
at high temperature
HCP metal generally brittle at room
temperature
e.g., beryllium, magnesium, and zinc

Summary of Slip systems


BCC

FCC

HCP

Slip System

48
Truly close
packed plane
does not exist

12

3
(more at high
temperature)

Probability for Slip

High

Moderate

Low

Ductility (>5)

Moderate

Good

Brittle

b/a

High

Low

N/A

Shear Stress

High

Low

Moderate

Strength

High

Moderate

Moderate

Example

Cr, W, Mo, Ta
Alpha Iron
(<912C or
>1394C)

Cu, Al, Au, Ag,


Gamma Iron
(912C ~
1394C)

Ti, Zr, Mg,


Zn

(Allotropy or
polymorphism)

Ionic ceramic crystal

Alternate positive(black) & negative(white) ions

Slip is difficult in ionic crystals


of ceramics

Ionic structure of
alternate anions and
cations

One interionic distance


movement along a plane
causes fracture as like
ions face each other.

Imperfection In Metal Structure


(General)
Actual strength of metal is 1 to 2 orders of magnitude lower
than theoretical calculated strength level
This is due to defects & imperfection in the crystal structure
( particularly, dislocations as illustrated)
Structure Sensitive: Mechanical & electrical properties of
metal, such as yield, fracture strength & electrical
conductivity, adversely affected by those defects
Structure Insensitive: Physical & chemical properties, such
as melting point, specific heat, coefficient of thermal
expansion & elastic constants are not sensitive to those
defects

Strain (Work) Hardening


Dislocation can become entangled & interfere with each other
(Example: moving two humps at different angles across a carpet
with the two humps interfering with each others movement, making it
more difficult to move the carpet)

Impeded by barriers, such as grain boundaries, impurities &


inclusions in the materials
Entanglement and impediment increase the shear stress
required for slip
The effect of an increase in shear stress increases the strength
of metal at ambient temperature
Strengthening wire by reducing its cross-section by drawing it through
a die
Producing the head on a bolt by forging it
Producing sheet metal for automobile bodies and aircraft fuselages by
rolling

Solidification of Polycrystalline Structure


Schematic illustration
of the stages during
solidification of molten
metal; each small
square represents a unit
cell. (a) Nucleation of
crystals at random sites
in the molten metal;
note that the
crystallographic
orientation of each site
is different. (b) and (c)
Growth of crystals as
solidification continues.
(d) Solidified metal,
showing individual
grains and grain
boundaries; note the
different angles at
which neighbouring
grains meet each other.
Source: W. Rosenhain.

Crystals

Grain Boundary

Grain or Crystalline
Structure

Plastic Deformation of Polycrystalline Metal


Deformation in each grain takes places by the
mechanisms for single crystals
The greater the deformation, the stronger the
metal becomes
~ Because of GB entanglement of dislocation

The smaller the grain sizes, the higher the


strength for metals
~ Because they have a large grain-boundary surface
area per unit volume of metal

Plastic deformation of Polycrystalline Metal


(Anisotropy)
Anisotropy (As a result of plastic deformation, the
grain have elongated in one direction & contracted
in the other. Its properties in vertical direction are
different from horizontal direction.)
Preferred orientation
Mechanical fibering
Anisotropy influences mechanical & physical
properties (increase in strength and decrease in
ductility)

Summary of Plastic Deformation


Plastic deformation causes increase in strength,
decrease in ductility & causes anisotropic
behavior
Plastic deformation can be reversed and properties
of the metal can be brought back to their original
level, by heating the piece in a specific
temperature range for a period of time (annealing)

Deformation, Strain Hardening, and


Annealing
Three Main Topics for Discussion
Cold Working or Strain Hardening
Hot Working
Annealing

By controlling the processes of deformation and


heat treatment, we are able to process materials
into usable shapes while improving and
controlling properties
Topics pertain mainly to metals, particularly
ductile metals

Cold Working
Development of Strain Hardening
A specimen is stressed beyond the yield strength
before the stress is removed
Now the specimen has a higher yield strength and
tensile strength, but lower ductility
By repeating the procedure, the strength continues to
increase and the ductility continues to decrease until
the alloy becomes very brittle

Mechanism of Cold Working


Dislocation Multiplication
Beyond the yield strength, the
dislocations begin to move causing slip
(plastic deformation)
Eventually, a dislocation moving on its
slip plane encounters obstacles which
pin the ends of the dislocation line
As we continue to apply the stress, the
dislocation attempts to move by
bowing in the center until a dislocation
loop forms
More dislocations we have, the more likely they are to
interfere with one another, and the stronger the metal
becomes

You might also like