You are on page 1of 20

3. PLASTIC DEFORMATION.

DISLOCATIONS AND STRENGTHENING MECHANISMS




3.1 INTRODUCTION

The objective of this theme is to establish the physical background for understanding the
plastic deformation of metals. We will see how dislocations move and interact with grain
boundaries, precipitates and other dislocations resulting in hardening.
Metallic materials are, in general, an aggregate of very small crystals, which are called
grains. Therefore, most structural materials are polycrystalline.
The required stress to initiate and produce permanent deformation, i.e. plastic
deformation, is an important concept to discuss the mechanical behaviour of materials. In
crystalline materials, this plastic flow is related directly to the presence of dislocations and to
their response to the applied stresses. For instance, the presence of dislocations explains the
low stresses required to cause plastic flow in crystals having dislocations, in comparison to
the much higher stresses necessary to initiate plastic flow in a perfect crystal, i.e. without
dislocations.
Likely, you have already dealt with dislocations as defects of the crystal lattice, and know
some elemental concept about dislocations. A material during deformation is a dynamic
system. The number and morphology of dislocations changes rapidly as plastic strain
increases, what is macroscopically manifested via a flow stress that varies with the
accumulated plastic strain.

3.2 YIELD STRENGTH OF A PERFECT CRYSTAL

Let us see that the yield strength of a perfect crystal is considerably greater than the one for a
real crystal. Assume that atomic slipping on a crystallographic plane is the cause of plastic
flow by an ideal shearing process as the figure illustrates.

a 2

Slip process in a perfect crystal
3 4
5
1
2
3 4 5
0 b
Atomic displacement x
L
a
t
t
i
c
e

e
n
e
r
g
y

High energy position
2 1
3 4
5
2 1
Low energy position
Slip
plane


b
3 4
5
2 1
a
Atomic displacement x
0
b
Atomic displacement x
S
h
e
a
r

s
t
r
e
s
s

b/2

For the atoms in the upper plane slide over those in the lower one, intrinsically strong
interatomic forces must be overcome by the applied stress. The crystal energy varies with the
relative atomic displacement across the slip plane as shown in the figure. When the atoms in
the upper plane have been displaced by one-half of their interatomic distance, b/2, the crystal
energy corresponds to a maximum. The force or shear stress required to produce the
displacement will be proportional to
dx
dU
, where U is the lattice energy. If the variation of U
with the atomic displacement is
b
x
B A x U
2
cos ) ( =
the shear stress to produce the atomic displacements is given by a function such as
b
x

2
sin
max
=
For small shear strain and stress, the relationship G = is satisfied, what means that
0

d
d
G . Then,
max
0
max
2 2
cos
2


b dx
d
b
x
b dx
d
=

=
Using
max
0
0 0
0
2 1


b
a
dx
d
a
d
d
d
d
a dx
d
a
x
dx
d
d
d
dx
d
=

=
=

Therefore,
a
Gb
b
a
d
d
G

2
2
max max
0
= =

=
where a is the spacing between the slip planes, and b the slip distance. In general,

2
max
G
b a =
So, the theoretical shear strength is estimated to be on the order of G/2. Actually, it has
been overestimated because the function (x) for a real crystal is not symmetrical in the
range 0 b/2 as represented in the above figure. A more refined approach lowers the
estimate shear stress to
30
max
G

These values are much higher than the experimental yield strength values found for single
crystals of pure metals. For instance, see the following table. The ratio
th
/
exp
is between 100
and 1000.

Metal
th
=G/30
(GPa)

exp
(MPa)
Al 0.9 0.78
Cu 1.4 0.49
Ni 2.6 3.2
-Fe 2.6 27.5
In all cases the discrepancies between the theoretical and the experimental values for the
critical shear stress are so great that there is not doubt that atomic slip, by which plastic
occurs, is accomplished by the movement of dislocations. These lattice defects were
postulated by Orowan in the 1930s to account for the mentioned discrepancies.


3.3 THE CONCEPT OF SLIP. DISLOCATION SLIP

There are two types of dislocation movements:
1. Glide or conservative motion, when the dislocation moves in the plane which contains
both its line and its Burgers vector. A dislocation able to move in this way is termed
glissile, one which can not is termed sessile.
2. Dislocation motion by climb occurs when the dislocation moves out of the glide plane,
and thus normal to the Burgers vector.

Glide of many dislocations results in slip which is the most common manifestation of
plastic deformation in crystalline solids. Plastics deformation, or plastic flow, can be
envisaged as sliding, or successive displacements of one plane of atoms over another on
so-called slip planes. Discrete blocks of crystal between two slip planes remain
undistorted as the figure shows. Further deformation occurs either by more movement on
the existing slip planes or by the activation of new slip planes.

stress shear resolved cos cos =


Slip
direction
F
F
// F

A

A
o



Geometry of slip. Note that +90, in general


The slip planes and the slip directions in a crystal have specific crystallographic forms.
The slip planes are normally the planes with highest density of atoms, i.e. those which are
not widely spaced, and the direction of slip is one of the shortest lattice translation vectors.
Often, this direction is one in which the atoms are most closely spaced. The shortest lattice
vectors are: 0 1 1
2
1
for the fcc system, 1 1 1
2
1
for the bcc and 0 2 1 1
3
1

for the cph system. Thus, in cph solids, slip frequently occurs on the (0 0 0 1) basal plane
in directions of the type 0 1 2 1 , and in fcc metals on { } 1 1 1 planes in
0 1 1 directions.
In bcc metals the slip directions are 1 1 1 closed-packed directions, and the slip
plane can be and { } { 3 1 1 , 2 1 1 } { } 0 1 1 planes. The last are the preferred slip
planes for bcc metals at low temperatures. Note that these three families of planes have a
1 1 1 direction as a common zone axis.
A slip plane and one slip direction in it constitute a slip system. fcc crystals have four
planes with three { 1 1 1 } 0 1 1 per plane, and therefore, have 12 possible
/ { } 1 1 1 0 1 1 slip systems. The figure shows the typical slip systems for cph, fcc
and bcc crystals.

0 1 2 1 directions
(0 0 0 1) plane



Slip results in the formation of steps on the surface of the crystal, as the figure for the
geometry of slip shows. These steps give rise to the appearance of straight slip bands on
the surface of the material if its surface is carefully polished before deformation. Each band
is made up of a large number of slip steps on closely spaced parallel slip planes.


It is clear that for slip a characteristic shear stress is required. Consider a cylindrical
single crystal being deformed in tension under a uniaxial stress as is represented in one
of the above figures. The applied tensile stress is
o
A
F
= , and shear stress resolved on
the slip plane in the slip direction is cos cos = . If the applied stress required to
start deformation by slip is
c
, there exists a corresponding critical resolved shear stress
(crss) for slip, cos cos
c c
= . It is clear that the crss depends on the slip system.
It has been seen that the theoretical shear stress for slip is many time greater
than the measured one, i.e.
c
. This can be explained by the movement of the dislocations
present in a real crystal.

How dislocations move?

Consider an edge dislocation as represented in the figure. The dislocation, that can be
considered as the lineal defect created by an extra half plane of atoms inserted into the
upper portion of the crystal lattice, and terminating on a {1 0 0} plane. The means by
which relative atomic displacements are achieved by motion of this edge dislocation is
schematically represented in the following figures.






b c
b
c



d a
a d

b c
a d
b c
a d
b

Inspection of the atomic arrangements in the neighbourhood of the termination of the
extra half plane shows that, in absence of an applied stress, the atom a at the termination
is equally attracted to atoms b and c in the plane below it. Application of a stress alters
these conditions so that atom a is preferentially attracted to atom c. As a result of this,
minor atomic rearrangements occur in the neighbourhood of the dislocation core to produce
the atomic arrangement as shown in the figure. The new atomic positions are such that the
dislocation line has been translated one atomic position to the right. That this, the extra half
plane now ends at atom d. It is worthwhile emphasizing that the extra half plane has not
moved as such. Rather, the small displacements resulting from the applied shear stress
serve to relocate the position of the termination of the extra plane. With the continued
application of the stress, the dislocation continues moving, and in the process atoms laying
on the termination, i.e. on the slip plane, are displaced to the right relative to atoms on the
plane below it. The slip direction is the direction in which the dislocation line moves. The
magnitude of the slip distance corresponds to the Burgers vector b, i.e. the interatomic
spacing in the slip direction on the slip plane, and represents the relative displacement of the
atoms on this plane with respect to those on the plane immediately below it. Continued
motion of the dislocation from left to right, until it emerges on the surface of the crystal,
produces a permanent offset of magnitude b in the upper half of the crystal with respect to
the bottom one. That is, plastic deformation has resulted from the dislocation motion.

Let us compare the means by which permanent deformation is achieved by dislocation
motion and by perfect slip. For perfect slip, every atom in the slip plane moves
simultaneously a distance b with respect to atoms in the plane below it. This concurrent
motion produces an instantaneous displacement, b, of the two portions of the crystal, and
the shear stress needed to accomplish this is given by fundamental atomic interactions as
reflected in the expression for the theoretical shear strength. If slip occurs by motion of edge
dislocations, the displacement b is performed only on one column of atoms at once; actually
this column represents a lower half atomic plane. It is only when the dislocation has moved
the entire length of the slip plane over the crystal that complete displacement of the two
halves of the crystal, by a distance b, is achieved. Indeed, if the crystal is composed of N
atomic columns, each time the dislocation moves a distance b in the slip direction, the
displacement of the upper part of the crystal with respect to the lower part is the distance
b/N. This incremental nature of the slip by dislocation motion makes that the shear stress
required to produce plastic deformation be much less than the theoretical required stress to
slip simultaneously all atoms by the distance b all at once.


Peierls stress
The resolved shear stress required to make a dislocation glide in a crystal is called
Peierls or frictional stress. It arises as a direct consequence of the periodic structure of the
crystal lattice and depends on the form of the interaction force between individual atoms, i.e.
on bonding. Peierls stress also depends on the structure of dislocation core and, for this
reason, a unique analytical expression for it can not be derived. The core is the region of the
crystal where the atom displacements with respect to the ideal sites of the lattice is greater
than b/2, or where the elastic approach can not be applied because the displacements are
large.

Slip
plane
x
y
a
A
B
b


The relative displacements between the atoms across the slip plane changes with the
distance to the dislocation line, so that interaction forces between the atoms across the slip
plane also change. If it assumes that this force is given by a periodical function
as
b
x
F
2
sin , one can obtain that the stress required to move the dislocation would be
given by the equation:
( )

=
b
w G
P

2
exp
1
2


Where w is the width of the dislocation core. For an edge dislocation, the dislocation width
may be taken as

=
1
a
w ; a is the atomic spacing across the slip plane.
P
is called the
Peierls stress or lattice friction stress. Now, it is easy to see that slip should occur most
quickly on close-packed atomic planes, which have the greatest separation distance, and in
close-packed directions for which the atomic slip distance b is a minimum. In most materials,
slip is observed to occur most promptly on close-packed planes and in closed-packed
directions as well. The Peierls stress gives a much closer approximation to the critical
resolved shear stress than the one obtained from the simple shear model for a perfect
crystal. The experimental values for the crss are greater than the Peierls stress indicating
that other strengthening mechanisms, which hinder the dislocation motion are generally
present in a real material. This can be induced by the presence of: impurities (solid solution
hardening), precipitates (precipitation hardening), dispersed particles (dispersion
strengthening), internal boundaries (boundary hardening), and defect created by
deformation or irradiation (work hardening or irradiation hardening).





3.4 DISLOCATION CLIMB

An edge dislocation usually moves in the slip plane, that is, in the plane containing the
Burger vector and the dislocation line. However, under special conditions it can move out of
its slip plane in a direction perpendicular to the slip plane. This process is called climb.
Dislocation climb occurs by the diffusion of atoms and vacancies to or away from the site of
the dislocation line. Positive climb occurs when atoms are removed from the dislocation line
by diffusion of an atom to a vacant lattice site. This produces that the dislocation moves up
one atom spacing.



a
If atoms diffuse to the dislocation and are added to the dislocation line below the extra
half-plane, the dislocation line moves down one atom spacing, then the climb is called
negative. Since climb is diffusion process, the importance of dislocation climb as a
deformation mechanism increases with temperature. Moreover, the climb rate enhances
when stresses are applied to materials at high temperatures.

J ogs and kinks
A complete row of atoms is not removed or added to the extra plane of the edge
dislocations. In practice, single vacancies or small vacancy clusters diffuse out or to
dislocation line. The effect is that a short portion of the dislocation line climbs up or down
resulting in the formation of two steps at the dislocation line, called jogs. Both positive and
negative climbs proceed by the nucleation and motion of jogs. They are sources and sinks for
vacancies. Actually, a jog is a step in the dislocation line that moves it from one atomic slip
plane to another slip plane. Thus, the dislocation line may lie in several slip planes
simultaneously.

A jog in an edge dislocation (left) and a screw dislocation (right).

a
a
b
b
Steps which displace the dislocation line in the same atomic slip plane are called kinks.

A kink in an edge dislocation (left) and a screw dislocation (right).
The formation of jogs results in an increase of the critical shear stress required to
produce dislocation slip. Jogs are produced extensively during plastic deformation by the
intersection of dislocations, and are present even in well-annealed crystals. There is a
number of jogs per unit length of dislocation in thermodynamic equilibrium (thermal jogs)
given by

=
kT
E
n n
j
o j
exp
eV E
j
1 is the energy associated to a jog resulting from the increase in dislocation length,
and in principle can not climb. However, a small edge component or jog on a screw
dislocation will provide a site for the start of climb.

3.5 PLASTIC STRAIN DUE TO DISLOCATION MOTION

Now, we are going to quantify the amount of plastic strain produced by the dislocation
movements. Both mechanisms of dislocation motion, slip and climb, produce plastic
deformation. The relation between plastic strain and the applied stress depends on many
factors such as temperature, strain rate, microstructure, impurity content etc, and it is in
general very complicated to obtain in terms of an analytical equation. However, a simple
relationship between the plastic strain and the dislocation density in the material there exists.
This is because when dislocations move, the atoms on two sites adjacent across the plane of
motion are displaced relative to each other by the Burgers vector b of the dislocation. Let us
derive the relationship between plastic shear strain and dislocation density for a slip process.
Recall that the dislocation density in a crystal is defined as the total length of
dislocations per unit volume. It is quoted in m
-2
or cm
-2
. Thus,
V
l
= , where V is the volume
of crystal containing a dislocation line length l. An alternative definition for is the number of
dislocation intersecting a unit area. These two density values are the same if all dislocations
are parallel, but for completely random arrangement the volume density is twice the surface
density. The typical values of for well annealed metals are in the range 10
10
10
12
m
-2
, and
for heavy cold-rolled metals between 10
14
10
15
m
-2
.


x
i
D
h
l
d
i
i
Plastic deformation D produced by the dislocation slip. x
i
represents the displacement
of dislocation i.


Consider a crystal of volume V=hld containing straight edge dislocations. Under a high
enough applied shear stress acting on their slip plane in the direction of their burgers vector
b, as shown in the figure, the dislocations will glide, positive ones to right, negative ones to
left. Therefore the top surface of the crystal is displaced plastically relative to the bottom
surface as demonstrated in the figure. If a dislocation moves completely along the slip plane
the distance d, it contributes with a distance b to the total displacement D. Since b is very
small in comparison to d and h, the contribution made by a dislocation that moves a distance
x
i
may be taken as
d
b
x
i
. If N dislocations move, the total displacement will be
i
N
x
d
b
D

=
1

and the macroscopic plastic shear stress would be

= =
N
i
x
hd
b
h
D
1

If the average distance moved by a dislocation is x N x x
N
x
N
i
N
i
= =

1 1
1
, and
the dislocation density of mobile dislocations in the crystal is
hd
N
hld
Nl
m
= = , the plastic
shear strain can be expressed by
m
x b x N
hd
b
h
D
= = = tan
The strain rate would be
m
v b
dt
d

= = &
where
dt
x d
v = is the average dislocation velocity. The same relationship holds for screw and
mixed dislocation.
It should be noted that in the above equations
m
represents density of mobile
dislocations. Dislocations that do not move, i.e. sessile dislocations, would not contribute to
the plastic flow. Nevertheless, it is frequent to see the above equations applied the total
density of dislocations,
tot
, instead to mobile dislocation density. It can be demonstrated that
tot tot m
tot tot m
v b v b
x b x b


= =
= =
&

tot tot
x and represent the mean values of the displacement and velocity measured relative
to the total population of dislocations. Demonstrate the above.

Now, let us see the strain induced by dislocation climb under an external tensile load as
shown schematically in the following figure.

h
d
h
Plastic tensile strain produced by climb of the edge dislocations

When and edge dislocation climbs, a portion of extra plane of width b is inserted into, or
removed from, the crystal in the plane over which the dislocation moves along. As in the case
of slip, if a dislocation moves along a distance x
i
, it contributes
d
b
x
i
to the displacement h
of the external surface of the crystal. Now, it is easy to see that the total plastic tensile strain
parallel to the Burgers vector b of the dislocations is given by
m
N
i
x b x N
hd
b
x
hd
b
h
h
= = =

=

1

and for the strain rate,
m
v b = &
The same relations also hold for mixed dislocation, except that b is then the magnitude of the
edge component of the burgers vector.



3.6 DEFORMATION OF SINGLE CRYSTALS AND POLYCRYSTALS

There are important differences between plastic deformation in single crystals and
polycrystalline materials. In a polycrystalline material, the individual crystals, or grains, have
different orientations, and the resolved shear stress for dislocation slip, cos cos = ,
varies from grain to grain.

Single crystals
The initial elastic strain is caused by the simple stretching of bonds. Hooke's Law applies
to this region. At the yield point, stage I begins. The crystal will extend considerably at almost
constant stress. This is called easy glide, and is caused by slip on one slip system (the
primary slip system).


Stage I
Stage II
Stage III
Shear stress
R
e
s
o
l
v
e
d

s
h
e
a
r

s
t
r
e
s
s

Plot of the resolved shear stress
R
acting on one slip system against the shear
strain for a single crystal of a cubic metal

The geometry of the crystal changes as slip proceeds. The Schmid factor changes for each
slip system, and slip may begin on a second slip system when the resolved shear stress on
the secondary slip plane reaches the critical value to move dislocations. This occurs as its
Schmid factor is equal to that of the primary slip system. In this stage of deformation, known
as stage II, dislocations are gliding on two slip systems, and they can interact in ways that
inhibit further glide. Consequently, the crystal becomes more difficult to extend, that is, work
hardening occurs. The stress / strain ratio in stage II may be constant. Stage III corresponds
to extension at high stresses, where the applied force becomes sufficient to overcome the
obstacles, so the slope of the graph becomes progressively less steep. The work hardening
saturates. Stage III ends with the failure of the crystal.


Polycrystals
Polycrystals are composed of many grains with different relative crystallographic
orientation. If the material is untextured, the grains are randomly oriented. When the bulk
material is deformed, each individual grain undergoes slip. The stress at which slip begins in
each grain depends on its orientation with respect to the tensile axis, following the Schmids
law. The shape change in a plastically deforming grain may be constrained by neighbouring
grains that have not yet reached their yield point. In addition, the grain boundaries, being
regions of considerable atomic misfit, act as strong barriers to dislocation motion. Also, the
internal stresses around piled-up groups of dislocations at grain boundaries, that have
yielded, may cause dislocation sources operating in the neighbouring grains. Thus the
macroscopic yield stress at which all grains yield depends on grain size. Finally, a grain in a
polycrystal is not free to deform plastically like a single crystal, because it must remain in
contact with, and accommodate the shape changes of its neighbour grain. The inability to
meet this condition leads to a small strain at failure. The result of all these factors is that the
stress-strain curve for a polycrystalline material is different than the one for a single crystal
and does not exhibit distinct stages characteristic of the single crystals.
So that, depending on grain orientation respect to the direction of the load applied to the
material, some grains yield first, just when the resolved shear stress reach a critical value,
and then other grains will follow progressively as the applied load increases. When a c.c.p.
metal exhibits stages I, II and III in the single crystal stress-strain curve, these stages will not
be seen in a stress-strain curve from a polycrystalline sample of the same material: each
individual grain goes through these stages at the same resolved shear stress and hence at
different applied stress. The only well-defined feature in the stress-strain curve is the yield
point, at which plastic deformation begins across the whole sample.
Each grain has a different Schmid factor, m=cos cos. Then, the yield stress would be
given by the called Schmids law:
m
c

= =
y
cos cos
This equation is generalised to
c
M =
y

Where M=m
-1
is known as the Taylor factor. The average value of the Schmid factor over the
whole sample can be calculated, assuming that the grains are randomly oriented. The
resulting Taylor factor is M3 for cubic close packed metals. This gives
y
approximately
equal to 3
c
.
3.7 STRENGTHENING OF MATERIALS

We will start with the primary mechanisms by which the flow strength of crystalline solids
increases by restricting the dislocation mobility. A material can contain various types of
obstacles or barriers, that either alone or in combination, hinder the dislocation motion.
These obstacles are: other dislocations, internal boundaries (grain boundaries, interfaces,
cell boundaries ) solute atoms, point defects, vacancy clusters, second phase particles
etc.
Most high-strength alloys are hardened by more than one mechanism. Although the basic
physics of strengthening is well known, accurate quantification of the incremental strength
provided by each mechanism is difficult and further when they operate in combination. One
should consider that the intrinsic strength of a crystalline material is given by the Peierls
stress being determined by the crystal structure and the bonding characteristics.

Stress to bend a dislocation
Remember that the dislocation energy per unit length is
2
Gb E =
where is a parameter that depends on the dislocation type, and G the shear modulus.
Usually 0.5.
When a straight segment of dislocation is bent, its energy increases due to the length
change. A dislocation has associated a line tension, T, which is analogous to the surface
tension of the liquid, that is, the energy required to increase the dislocation length in a unit
length. It is indeed the dislocation energy per unit length. One can calculate the stress that
has to be applied to bend a dislocation line from the following drawing

The total force acting on the element of line of a bent dislocation is
dl
d R
d/2
d/2
T
T
T
x
T
y
T
y

Td
d
T T
y
= )
2
sin( 2 2
An applied shear stress acting on the slip plane,
o
, would be required to maintain in
equilibrium the dislocation. It would be given by the condition:
R
Gb
bR
T
bRd bdl Td
o o o

= = = =
R is the radius of curvature,
o
corresponds to the shear stress resolved in the slip direction,
i.e. in the direction of b in the case of edge dislocations, and bdl is the area swept by the
element of dislocation line.
o
represents the stress required to maintained the dislocation line
bent to a radius R.

Obstacle strength
Consider a slip plane containing an array of unspecified obstacles, as shown in the
figure. The array is characterized by a mean inter-obstacle spacing on the slip plane a
distance L. In response to an applied stress, the dislocation segments are pinned by the
obstacles and bowed between two adjacent obstacles to a radius R. This radius, in
equilibrium, would be given by
0 0
2

GB Gb
R =


L
applied rss
b
bL area swept by the
dislocation segment
T
s
T
r

K
r


L
L
If the resistance offered by each obstacle is represented by a force K acting at a point on the
dislocation line, then the equilibrium condition at an obstacle is satisfied if
2
cos
2
cos 2
2

Gb T K =
Each obstacle resists the forward force bL acting on a dislocation segment of initial length L.
The dislocation segment breaks away from the obstacles when this force overcomes the
maximum resisting force of the obstacles, K
max
. Then, the strength of an obstacle,
ob
,would
be obtained by the condition
2
cos
max
max
c
ob ob
L
Gb
bL
K
K bL

= =
The obstacle strength is characterized by the critical angle
c
at which the breakaway of the
dislocation segment occurs.
It should be noted that actually L is the effective obstacle spacing. That is, if all obstacles
have the capacity for pinning dislocations, i.e. if they are strong obstacles, L would be the
average obstacle spacing in the slip plane. For very strong obstacles,
c
0, and the obstacle
strength is
L
Gb
ob
for very strong obstacles.
For weak obstacles, the effective obstacle spacing does not correspond to the real
average spacing between to adjacent obstacles but to
2
cos
C
L
L

=
according to an estimation obtaining applying some approaches. Thus, for weak obstacles it
is obtained
2
3
2
cos

c
ob
L
Gb
for weak obstacles





3.8 WORK HARDENING
When the dislocation slip was analysed, the interaction of mobile dislocations with other
dislocations was neglected. However, this interaction during dislocation slip produce changes
in the distribution and density of the dislocations. That is, dislocations themselves are
obstacles to dislocation motion. Thus, the glide resistance for dislocations increases. This
results in work hardening.
Consider the interaction of one mobile dislocation gliding on a slip plane with dislocations
that lie on a different slip plane, as shown in the figure. When the mobile dislocation runs into
the immobile dislocations, it is pinned by the immobile ones until the applied resolved shear
stress overcomes the obstacle strength,
ob
, given by
L
Gb
ob
=
where is a factor giving the obstacle strength in this case.


1
2
3
Slip plane
Slip direction
Immobile dislocations
mobile dislocation
L
If the materials has a dislocation density distributed in a network of mean spacing L, then

Gb
L
Gb
L
ob
ob
=

=
=
2
1

That is, plastic flow will occur for an applied resolved shear stress equals to the above
obstacle strength
ob
. This means that during plastic flow, the flow stress have to be
Gb =
Furthermore, assuming that each dislocation moves an average distance x before being
stopped by other dislocations, from the relationship between plastic strain andx , we obtain
the following equation
n
x
b
G
Gb
x b
x b

2
1

=
= =

That describes the relationship between resolved shear stress and plastic shear stress. Here
n=0.5.
The above equations are changed to apply the model to polycrystalline materials
uniaxially tested. In these cases, the tensile stress is related to the resolved shear stress
through the Taylor factor M. So that, now
n n
K
x
b
MG =

=
2
1

A relation between tensile stress and tensile strain like the Hollomon equation is obtained.
Actually, the above equation gives us the dislocation contribution to the flow stress, i.e.
the dislocation strengthening. In addition, the lattice resistance to dislocation motion also
acts, so that the Peierls stress, or intrinsic lattice strength is also contributing to flow stress.
Therefore, the flow stress would be expressed by
Gb
o
+ =
or
n
o
x
b
G
2
1

+ =
Usually, it is used 4 . 0 for bcc and cph metals, and 2 . 0 for fcc metals.
When tensile stress and tensile strain is used, the above relationship are expressed by
MGb
o
+ =
n
o
n
o
K
x
b
MG + =

+ =
2
1

Thus, a relationship similar to the Ludwik equation used to describe the relationship between
flow stress and tensile strain is obtained.



3.9 GRAIN BOUNDARY STRENGTHENING

Grain boundaries in a polycrystalline material are also barriers to dislocation motion. Actually,
geometrical considerations suggest that planar defects like internal boundaries can be
stronger obstacles to dislocations than linear defects or point defects because the
intersection of a slip plane with a boundary is line instead of a point. This should provide a
resistance to motion stronger than an isolated point obstacle because the dislocation pinning
is along the entire length of the dislocation line that intersects the boundary. Grain boundaries
are particularly very effective barrier to dislocation motion since block the way the adjacent
grain. The change of the crystallographic orientation across the grain boundary cuts off the
slip plane and does not allow the passage of a dislocation to the adjacent grain.
Tensile tests performed on polycrystalline materials in which the grain size is the only
variable parameter in the sample, it has been established that yield strength and grain size
are correlate by the called Hall-Petch law
n
y o y
d k

+ =
where k
y
is a material constant, n0.5 and
o
is the intrinsic lattice strength. This relationship
can be obtained from the dislocation pileup theory (see Hirth and Lothe, Theory of
dislocations 2
nd
ed, p. 788.





3.10 SOLID SOLUTION STRENGTHENING

Solute atoms in a crystal lattice increase the yield strength of most metals as a result of the
interactions between moving dislocations and solute atoms which are on or nearby the slip
plane of the dislocations. Since dislocation stress field is a long-range one, solute atoms
located both on the slip plane, and above or below it, interact with dislocations. For instance,
substitutional impurity atom that is smaller than the host atoms exerts tensile strain on the
surrounding crystal lattice. On the contrary, a larger impurity atom produces compressive
strains, as shown in the figures. These substitutional atoms tend to diffuse to the dislocation
and segregate around it lowering the total strain energy of the crystal since some of the
lattice strain around the dislocation is cancelled. That is, the dislocation is localization of
minimum energy, and a higher critical resolved shear stress will be required to move out from
this position.






The solution strengthening can be very strong as the following figure shows for Cu-Ni
alloys.

You might also like