You are on page 1of 80

Dislocation Theory for Engineers:

Worked Examples
Dislocation Theory
for Engineers:
Worked Exalllpies

Roy Faulkner
and
John Martin
Book 744
First published in 2000 by
10M Communications Ltd
1 Carlton House Terrace
London SW1 Y 5DB

© 10M Communications Ltd 2000


All rights reserved

10M Communications Ltd


is a wholly-owned subsidiary of
The Institute of Materials

ISBN 1-86125-124-6

Typeset in the UK by
Keyset Composition, Colchester

Printed and bound in the UK at


The Alden Press, Oxford
Acknowledgements
To my wife (Linda) for putting up with my enthusiasm for things which she regards
with lower priority. (RGF)

I am indebted to my colleagues at the Oxford University Department of Materials


who have provided a number of the numerical problems quoted in the text. I am also
most grateful to Professor Brian Cantor FREng for the facilities provided for me in
the Department, and to Peter Danckwerts for his efficient dealing with editorial
matters. (JWM)

v
Contents
Introduction IX

1 Dislocations in Pure Materials 1


The stress required for plastic yield in a single crystal 1
Example 1.1 Identification of active slip system 1
The relative energies of defects formed by vacancy condensation 4
Example 1.2 Frank loop and prismatic loop 4
Example 1.3 Frank loop and tetrahedral defect 5
The effect of point defects on plastic flow 5
Examp le 1.4 Chemical stress 6
The effect of other dislocations on plastic flow 7
Example 1.5 Dislocation density 7
Example 1.6 Average plastic shear strain 7
Example 1.7 Forest model of flow stress 8
Example 1.8 A parabolic stress-strain relationship 9
States of work hardening 10
Example 1.9 Model of Stage I work hardening 11
Example 1.10 Neutron irradiation hardening 12
Dislocation/grain boundary interactions-the Hall-Petch equation 14
Example 1.11The Hall-Petch relationship (1) 15
Example 1.12 The Hall-Petch relationship (2) 16

2 Dislocation Interactions in Single Phase Materials 19


Segregation of solute atoms to dislocations 19
Example 2.1 The maximum temperature of segregation 19
Example 2.2 Solute interaction with an edge dislocation 20
Example 2.3 The rate of strain ageing 21
Dislocation dynamics of yielding 22
Example 2.4 Upper and lower yield points 23
Strengthening by substitutional solute atoms 25
Example 2.5 The CRSS for a crystal containing a random array of
point obstacles 25
Example 2.6 Solute hardening in steel 28

vii
Contents
3 Dislocation Interactions in Two-Phase Alloys 31
Particle shearing by dislocations 31
Example 3.1 Chemical strengthening 32
Example 3.2 Stacking-fault strengthening 33
Example 3.3 Coherency strengthening 34
Example 3.4 Order hardening 34
Particle by-pass by dislocations 35
Example 3.5 Effect of particle spacing on yield stress 35
Example 3.6 Calculation of critical interparticle spacing 36
Combination of strengthening processes 37
Example 3.7 Combining grain-size, solute and particle strengthening 37

4 Effects at Elevated Temperatures 41


Recovery and recrystallisa tion 41
Example 4.1 The driving force for sub grain growth 42
Example 4.2 Energy released during recrystallisation 43
Dislocation climb 45
Example 4.3 Rate of annihilation of dislocation loops 45
Dislocation creep 47
Example 4.4 Recovery creep model 49
Example 4.5 Power law creep calculation 50
Example 4.6 Use of the Monkman-Grant equation 51

5 Dislocations in Fracture Processes 55


Ductile fracture 55
Example 5.1 Calculation of ductile fracture strain in a steel 56
The ductile-brittle transition 58
Example 5.2 The effect of dislocation density on the ductile-brittle
transition 58
Fa tigue fracture 60
Example 5.3 S-N curve prediction 64

viii
Introduction
Our application of elementary dislocation theory to the behaviour of crystalline
materials will be made systematically, as follows:
1. Pure materials, and the effect of lattice defects upon the behaviour of
dislocations, considering in turn point defects (vacancies), line defects (i.e.
other dislocations) and planar defects (grain boundaries).
2. Solid solutions, considering the effects of the segregation of solutes to
dislocations, and also the interaction of moving dislocations with dispersed
solute atoms.
3. Two-phase materials, considering the interaction of dislocations with second-
phase particles.
4. Elevated temperature effects, including recovery and recrystallization, as well
as dislocation creep.
5. Fracture processes including ductile fracture processes, the effect of disloca-
tions at crack tips upon the ductile-brittle transition behaviour, as well as
fa tigue fracture properties.
We were commissioned to approach the subject in the form of a set of 'Worked
Examples' in order to complement the book of Worked Examples in Dislocations by
M. 1. Whelan, which was published by the Institute of Materials in 1990.1 That book
provides the essential basic concepts of dislocation theory, and the aim of the
present text is to apply that elementary theory to a number of problems in materials
science which should be of interest and value to the student reader.
The level is appropriate for that of the undergraduate student, and other, equally
appropriate introductory texts on the theory of dislocations in addition to that of
Whelan would be those by D. Hull and D. J. Bacon (Introduction to Dislocations, 3rd
edition, Pergamon, 1995) and by 1. Weertman and 1. R. Weertman (Elementary
Dislocation Theory, our; 1992).
There are several advanced texts on dislocation theory to which the graduate
reader can turn. These include 1. P. Hirth and 1. Lothe (Theory of Dislocations 2nd
edition, Krieger Publishing Co., 1992) and F. R. N. Nabarro (Theory of Crystal
Dislocations, Dover, 1987).
Finally, we include a list comprising a Glossary of the symbols employed in our
text, and a list of the principal equations we have quoted without derivation:
b Burger's vector
d grain size

ix
Introduction
D diffusion coefficient
E Young's modulus
E; dislocation core energy per unit length
f volume fraction
F obstacle strength
G shear modulus
k Boltzmann's constant
ky slope of Hall-Petch plot
KJ stress intensity (Mode I)
iii Taylor factor
M mobility
N, number of obstacles per unit area
N; number of obstacles per unit volume
ro inner cut-off radius from the dislocation
t time
T line tension of dislocation
w dislocation width

e tensile strain
B strain rate
l' shear strain
I'SF stacking-fault energy
v Poisson's ratio
p dislocation density
U normal stress
e work hardening rate
'T shear stress
n atomic volume

EQUATIONS
1. Elastic field of a screw dislocation
U()z = Uz() = Gb/27Tr
2. Elastic stress field of an edge dislocation lying along the z direction with its
Burger's vector b lying parallel to x:
Uxx = - Dy(3x2 + y2)/(X2 + y2)2
Uyy = Dy(x2 + y2)/(X2 + y2)2
Uzz = v( Uxx + Uyy)
uxy = Dx(x2 - y2)/(X2 + y2)2
Uxz = uyz =0

x
Introduction
where D = Gb/27T(1 - v), G is the shear modulus and v is Poisson's ratio.
3. Force per unit length acting on a dislocation (Peach-Koehler equation) = Tb.

2
4. Energy per unit. Iengt h 0f . diISIocation
a screw . = E e + -4-
Gb In- R
7T ro

Where Rand ro are the outer and inner cut-off radii respectively, and E; is the core
energy per unit length.
5. Energy per unit length of an edge dislocation:

Ed =
Gb
2

E; + 47T(1'- v) In
(R)ro
where E; is the core energy per unit length, Rand ro the outer and inner cut-off
radius from the dislocation respectively.

xi
1
Dislocations in Pure Materials

We will first consider the application of the Peach-Koehler formula, which is derived
in Whelan,! Ch. 5.

THE STRESS REQUIRED FOR PLASTIC YIELD IN


A SINGLE CRYSTAL

EXAMPLE 1.1 IDENTIFICATION OF ACTIVE SLIP SYSTEM

An fcc metal crystal has its initial tensile axis orientation [123J. Find which slip system
will operate initially.

When an axial tensile stress is applied to the crystal, plastic flow will take place
(J"

when the glide force per unit length Fg experienced by the dislocations is sufficient
to overcome the lattice friction. The Peach-Koehler equation gives:

where b is the dislocation Burger's vector and 1"c is the critical value of the shear
stress on the slip plane resolved in the slip direction.
Figure 1.1 illustrates the relation between the slip plane, slip direction and the axis
of tension for a cylindrical crystal of area A under an axial load P. The normal stress

Slip direction

Fig. 1.1 Slip plane (shaded), and slip direction in a cylindrical crystal.

1
Dislocation Theory for Engineers: Worked Examples

Fig. 1.2 Showing the relationship between the position of the tensile axis [123] and
the operating slip system (111)[101].

is = PIA. If the angle between the slip plane normal and the tensile axis is <p, and
(J"

the angle between the slip direction and the tensile axis is A, then 'T c may be readily
obtained. The component of the axial force P acting in the slip direction is P cos A;
the area of the slip plane is A/cos <p, so:

'T c = (J" cos Acos <p (1.1)

The operative slip system will be that for which the value of r in equation (1.1) is a
maximum. The values of cos A and cos <p for each of the twelve slip systems of the
fcc structure can be readily calculated by substitution in the 'cosine formula' for
angles between plane normals in cubic crystals:

fcc crystals slip on {111]<110>, and the three <110> directions lying in each of the
four {111}planes may readily be written down, since the Weiss Zone Law states that
the direction [UVW] lies in the plane (hkl) if

hU+kV+IW= 0

If a table of planes and directions is constructed, with the corresponding cos A and
cos <p values, it will be seen that the operative slip system will be (111)[101], the value
of cos A cos <p being 0.4665.
Figure 1.2 illustrates on a stereogram the relationship between the position of the

2
Dislocations in Pure Materials
w2

Fig. 1.3 Showing the operating slip systems for an fcc crystal as each stereographic
triangle in turn contains the tensile axis.

tensile axis [123], which lies in the stereographic triangle (001-011-111), and the
operating slip plane and direction which we have just calculated. We can recognise
that this plane and direction can be identified from the selected triangle by:

(a) reflecting (111) in the side opposite to this pole, to give (111), and similarly
(b) reflecting [all] in the side of the triangle opposite to [011], to give [101].

This process of 'reflecting' can be applied to whichever stereographic triangle the


tensile axis happens to lie in, enabling one to predict by inspection the operating slip
system in each case.
This general situation is illustrated in Fig. 1.3, in which the four {111} slip planes
are labelled A, B, C and D and the six <110> slip directions are labelled in Roman
characters, I, II, III, IV, V and VI. Each triangle is marked with the operative slip
system using the above nomenclature, if the tensile axis lies within it. Thus, for the
example we have calculated above (Fig. 1.2), the tensile axis lies in the triangle
labelled B IV, which means that plane B (111) and direction IV [101] are
operative.
Figure 1.3 illustrates the operative slip systems for every triangle containing the
tensile axis, and its validity can be checked by carrying out in each case the
'reflection' process described above.

3
Dislocation Theory for Engineers: Worked Examples

THE RELATIVE ENERGIES OF DEFECTS FORMED BY


VACANCY CONDENSATION

EXAMPLE 1.2 FRANK Loop AND PRISMATIC Loop

Explain how vacancies iii an fcc metal can condense to form prismatic Frank
dislocation loops, and how the Frank loop can be converted into a prismatic loop with
unit Burger's vector. Find the critical size of a circular Frank loop in copper at whicn
conversion to a prismatic loop is energetically favourable. [Assume that the elastic
strain energy of dislocation line is given by !Gb2 where G is the shear modulus
(45 GPa for copper), r is the radius of the loop and b the magnitude of the Burger's
vector. The lattice parameter of copper is 360 pm and the stacking fault energy is
40mJm-2.j

A detailed discussion of the processes of condensation of vacancies in crystals will


be found in Modern Physical Metallurgy by R. E. Smallman (Butterworth, 1985).2
Excess vacancies in a fcc crystal can lower their energy by agglomerating on {111}
planes. The crystal planes can then collapse, enclosing a stacking fault within a Frank
loop of dislocation of Burger's vector a/3 <111>, as illustrated in Fig. 1.4. The loop
will increase in diameter as more vacancies condense out on the close-packed
plane.
To unfault the Frank loop, a Shockley partial must pass across it and react with
the Frank loop. On the (111) plane this would result in:

a/3 [111] + a/6 [112] --+ a/2[110]

i.e. the dislocation loop acquires unit Burger's vector whose direction is not in the
slip plane of the loop. If the stacking fault energy is not very low, then it becomes
energetically favourable for the faulted Frank loop to transform in this way when the

A-------------------------

i
C-------------------------
8---
[111]
A---

C----

8-------------------------
A-------------------------
Fig.l.4 A loop of Frank dislocation of Burger's vector a/3[111] enclosing a stacking
fault.

4
Dislocations in Pure Materials
loop achieves a critical diameter such that the energy of the unfaulted loop
(bounded by a dislocation of greater Burger's vector) is less than that of the
original.
Energy of the faulted Frank loop of radius r = 2'ITr(iGb2) + 'ITr 'Y SF where 'Y SF is
the stacking fault energy and b = a/3[111]. The energy of the unfaulted prismatic
loop = 2'ITr(iGb2), where b = a/2[110]. Equating these energies and solving for r, we
have

i.e. r = [45.109 (! 360 .10-


2 24
) - (i3602 .10-24)J/0.04 = 24.3 nm.

EXAMPLE1.3 FRANKLoop ANDTETRAHEDRAL


DEFECT

In gold the clustering of vacancies leads to a lattice defect of tetrahedral shape with
stacking faults on its faces and stair-rod dislocations (al6<110» along each edge. If
the size of the defect tetrahedron is large, a triangular-shaped Frank sessile loop will
have a lower energy. Calculate the maximum size of the tetrahedron that can form, if
the stacking fault energy of gold is 30 mJ m ", the lattice parameter is 408 pm and the
shear modulus 42 GPa.

If the side of the tetrahedron is I, the area of each face 0.43312


The total energy of the tetrahedral defect = 4 X 0.43312. 'YSF + 61.iGb2, where
b = a/6<110>
The total energy of the triangular Frank loop = 31. !Gb2 + 0.43312. 'Y SF
The maximum size of the tetrahedron can be found by equating these energies:

whence I = 42.109.4082.10-24/3.9.0.03 == 59.8 nm.

THE EFFECT OF POINT DEFECTS ON PLASTIC FLOW

It is recognised that the hardening produced by point defects introduced by


quenching or irradiation is of two types:
(i) source hardening due to the pinning of dislocations by point defects in the
form of coarsely spaced jogs;
(ii) a general lattice hardening which persists after initial yielding, which will
change the value of TO, the lattice friction stress (equation 1.5).

5
Dislocation Theory for Engineers: Worked Examples
Vacaflcy supersaturation stress
When an edge dislocation encounters a high supersaturation of vacancies, it will act
as a vacancy sink and the dislocation will climb. The dislocation may be thought of
as being influenced by an effective stress somewhat analogous to osmotic pressure,
known as the chemical stress.

EXAMPLE 1.4 CHEMICAL STRESS

Estimate the chemical stress which would be exerted on the dislocations in copper by
each of the following operations:
(a) quench from 1000°C to 150°C
(b) 10% plastic strain at room temperature
The energy offormation of a vacancy may be taken as 1 eV, and the Burger's vector
b = 0.25 nm.

The change in free energy dU when dn vacancies are added to the system at
temperature T may be written:

dU/dn = Ef+ kTln(n/N)


= -kTlnco+kTlnc
= kTln(c/co)

where E, is the energy of formation of a vacancy, c the actual and Cothe equilibrium
concentration of vacancies at temperature T. This may be rewritten as

dU/dV = Energy/volume = stress, (J"c = (kT/b3) [In(c/co)] (1.2)

where dV is the volume associated with dn vacancies and b3 is the volume of one
vacancy.

(a) Copper quenched from 1273 K to 423 K


The equilibrium concentration of vacancies at 1273 K will be given by

C2= exp[Ef/1273k]

At 423 K the equilibrium vacancy concentration is given by C1= [Ef/423k]. Since


In(c2/c1) = (Ef/k)[423-1 -1273-1], from equation (1.2), the chemical stress «(J"c)
acting at 423 K is therefore given by

(J"c = Eflb3 [1 - (423/1273)]

6
Dislocations in Pure Materials
Taking Ef = 1 eV (=1.6 X 10-19 J), and b = 0.25 nm we obtain:

(Tc = 6.84 GPa

(b) Plastically deformed copper


During cold work, vacancies are produced at dislocation jogs ansmg from the
intersection of dislocations. The vacancy concentration c depends on the amount of
plastic strain 8 according to the approximate relation c = 10-48.
In the present case therefore c = 10-5•
At 300 K, Co = 1.69.10-17•
Substitution in equation (1.2) gives (Tc = 7.33 GPa.

THE EFFECT OF OTHER DISLOCATIONS ON PLASTIC FLOW

EXAMPLE 1.5 DISLOCATION DENSITY

Define the term 'dislocation density' and give typical values for Si, annealed Al and
heavily cold worked Cu. Calculate the average dislocation spacing in a crystal with a
dislocation density of 1012 lines m:".

The density of dislocations in a crystal may be defined as the total length of


dislocation contained within unit volume (metres per cubic metre). For a random
array of dislocations, an equivalent way of defining the density would be to express
it as the number of dislocations intersecting unit area, (m "),
In carefully prepared crystals of Si the dislocation density (p) ~ 106 - 108 m -2.
In annealed AI, p ~ 1010 - 1012 m-2 and in cold worked Cu the density can rise to
10 - 1016 m ".
15

Dislocation spacing. If p is the number of dislocations intersecting unit area, then


p1l2 is the number per unit length. Their spacing is therefore p -112, giving, for
p = 1012 m -2, a spacing of 1 J.Lm.

EXAMPLE 1.6 AVERAGE PLASTIC SHEAR STRAIN

Express the average plastic shear strain (y), and the shear strain rate in terms of the
number of mobile dislocations (N), the Burger's vector of the dislocations (b) and the
area of the slip plane that is swept when they move with an average velocity v.

7
Dislocation Theory for Engineers: Worked Examples
Consider a unit volume of crystal; it contains a slip plane of unit area. If a single
dislocation sweeps over the slip plane, the displacement of the element will be b, so
the average shear strain will also be b.
If p dislocations sweep the plane, then the average shear strain is Nb. If (say) a
fraction A of the slip plane is swept, it follows that
'Y = ANb (1.3)
If the average dislocation velocity is v, it follows that the average shear strain
rate = A Nbv.

Models of work hardening - simple 'forest' model

The dislocations in an undeformed crystal may be regarded as existing in the


form of a network within each grain. Dislocations cannot end within a crystal:
they either extend to the surface of the crystal, to a grain boundary or to
other dislocations (forming a dislocation node). The shear stress required to
cause a dislocation to glide on its slip plane will depend on the interaction
the gliding dislocation has with those other dislocations which intersect the
slip plane and which might be regarded as 'trees in a forest'.

EXAMPLE 1.7 FOREST MODEL OF FLOW STRESS

On a simple 'forest' model, calculate the flow stress for dislocation glide in a crystal
if the dislocation density is p.

A relationship may be derived in two ways:

1. Dimensional analysis
If the dislocations have Burger's vector b, an applied shear stress r will give rise to
a force per unit length on a dislocation of magnitude ,.b (Peach-Koehler equation,
Whelan, p. 51).1 The force of resistance to dislocation glide arises from interaction
with other dislocations and it may be written as exb2 (Whelan, p. 55),1 where ex is a
constant dependent upon the dislocation geometry. The balance of these forces
implies that r b.(X

The flow stress will depend upon the shear modulus, G, the Burger's vector, b, and
the dislocation density p. The simplest dimensionless expression involving these
parameters may be written:

where C and m are constants.

8
Dislocations in Pure Materials
!,
The requirement that T cc b indicates that m = giving the simplest expression for
the relationship between flow stress and dislocation density to be:

T = CGbpl!2

2. Simple cubic network


If it is assumed that the 'forest' dislocations of density p are arranged on a simple
cubic network, then p1l2 is the number of dislocations per unit length, and p-1I2 is
therefore the dislocation spacing. The applied shear stress causes the glide
dislocation to bend to a radius of curvature R, where

Tb = TIR

Taking T as the line tension of the dislocation, T = !Gb2, this leads to the familiar
relationship
T = !GbIR (1.4)

If, at yield, we assume that the glide dislocation bows between the 'forest'
dislocations, the maximum shear stress will correspond to the dislocation being
bowed into a semicircle of diameter equal to the forest dislocation spacing, i.e.

T = 2Gblp-1I2

If To is the lattice friction stress, then in general we can write the flow stress as:

T = TO + Gbo'" (1.5)
where C is a constant.

EXAMPLE 1.8 A PARABOLIC STRESS-STRAIN RELATIONSHIP

Show that a parabolic stress-strain relationship is obtained if it is assumed that the


flow stress is that necessary to move a dislocation in the stress field of those
dislocations surrounding it.

The first theory of work hardening, due to G. 1. Taylor, assumes that dislocations
created during straining become stuck after travelling an average distance L. If the
mean distance between the dislocations is l, the stress necessary to move a
dislocation is given by

Gb
T=---- (1.6)
81T(1 - v) 1

9
Dislocation Theory for Engineers: Worked Examples
r

" III
y

Fig.1.5 A typical shear stress-shear strain curve for an fcc single crystal oriented
initially for single slip.

If we assume that N square dislocation loops of side L are emitted per unit volume,
then from equation (1.3):

(1.7)

We have already observed that the dislocation spacing may be taken as the
reciprocal square root of the density, i.e.

l == (4LN)-1I2 (1.8)

So combining the above equations we obtain:

G (~b)1I2
'T == 81T(1 - v) L

This is the required relationship, although single crystals do not in general exhibit
this behaviour.

STATES OF WORK HARDENING

Figure 1.5 shows the form of a typical shear stress-shear strain curve for an fcc single
crystal oriented initially for single slip. The curve has three distinct stages: Stage I
commences at the critical stress 'To after an initial elastic deformation, and this stage
of low work-hardening rate is known as 'easy glide'. Stage II is characterised by an
increased, approximately linear rate of work-hardening, and Stage III is accom-
panied by a reduction in the rate of work-hardening.

10
Dislocations in Pure Materials

EXAMPLE 1.9 MODEL OF STAGE I WORK HARDENING

Derive a simple expression for the work hardening rate in Stage I of the stress-strain
curve for single crystals ('easy glide') assuming that it arises from the accumulation of
primary dislocation loops in the slip plane.

The model (due to Seeger et al.)3 assumes:


N dislocation sources/unit volume
n dislocation loops are produced by each source at stress l'

the dislocation loops are square, of side L


the slip line spacing is d( ~L)

A stress increment 8,- produces an increase 8n in the number of loops, which gives
rise to a strain increment 88 equal to b X N X area of slip plane swept, i.e.

(1.9)

When the slip plane is filled with loops, the number of loops per unit area must be
l1L2.
The number of slip planes per unit volume must be l/d, therefore

(1.10)

So 88 = b 8n/d (1.11)

The increase in the number of loops, 8n, gives rise to an increase in the back stress,
8,- B, and this will be given by:

(1.12)

Loop generation will cease when 81' = 81'B'


Equations (1.11) and (1.12) give the work hardening rate, 8/:

8/ = 8,-/88 = (G/27r)(d/L) (1.13)

Substitution of typical values of d and L gives


81~ 10-4 G, which is of the correct order of magnitude.

11
Dislocation Theory for Engineers: Worked Examples

The behaviour of polycrystalline aggregates (The Taylor Factor)

A number of theories have been developed to predict the stress-strain curve


of a polycrystalline aggregate from the behaviour of single crystals. Taylor
adopted the von Mises criterion that five independent slip systems are
required to cause a general shape change in a body, and in a given material
he assumed that those five systems which operate in each grain are those
which require the least work. He equated the work done by the macroscopic
stress (0") and strain (de) in the polycrystal to the work done by the five slip
systems in the individual grains. By a numerical treatment for a range of
crystal orientations, he related the critical shear stress for slip (To) to 0" by:

m is known as the Taylor factor, and its value depend on the crystal structure
of the material. An example of a calculation using this factor now follows.

EXAMPLE 1.10 NEUTRON IRRADIATION HARDENING

During neutron irradiation an assortment of dislocation loops are created in all


materials by the Frenkel pairs which emanate from the primary knock-ons and
subsequent cascades produced by the incident neutrons. The increased dislocation
loop density provides an additional set of barriers to glide dislocation movement and
hence an increase in the yield strength of the material. This can be modelled using
analogous arguments to those for Orowan strengthening (see section 3.2). Odette"
has proposed the following equation for yield strength increase:

Where Iii is the Taylor factor, relating the shear stresses on a slip plane in a single
crystal to the applied stress in a polycrystal, f is a random array efficiency factor
assuming that the obstacles are perfectly hard, and is usually about 0.8, a is the
obstacle strength (=1 in the case of precipitates), G is the shear modulus, b is the
glide dislocation Burger's vector, n is the loop density and d is their average
diameter.
In this approach the inter-obstacle spacing in the Orowan equation (3.2) has been
replaced by the term (nd)-1I2, which is equivalent to the inter-loop spacing. The a

12
Dislocations in Pure Materials
obstacle strength term should be emphasised because the strength of loops seems to
be somewhat less than that of hard particles. Odette" has estimated its value to be
about 0.4, with precipitates equal to 1 and network dislocations, very small loops
«1 nm) and bubbles about 0.2.

An austenitic steel with lattice parameter 0.356 nm is neutron irradiated at 495°C to a


dose of 14 dpa. The following microstructural parameters were measured after the
irradiation. If the yield stress in the pre-irradiated condition is 150 MPa, calculate the
yield stress after irradiation. The yield stress measured by the small specimen punch
technique was 738 MPa. Comment on this value in relation to your calculated value.

Network dislocation density = 1.3 X 1014 m ?


Small faulted loop density = 6 X 1O" m:'
Average diameter of small faulted loops = 0.8 nm
Large faulted loop density = 22 X 1O" m-3
Large faulted loop average diameter = 1.5 nm
Void density = 1.67 X 1~1 m-3
Mean void diameter = 19.3 nm
Taylor factor = 3.1
Random array efficiency factor = 0.8
Shear modulus of steel = 81 GPa

Answer

~O"small = 3.1 X 0.8 X 0.2 X 81 X 109 X (0.356 X 10-9/Y2)


(6 X 1021 X 0.8 X 10-9)112 Pa = 7 MPa
dO"large = 3.1 X 0.8 X 0.4 X 81 X 10 X (0.356 X 10-9/Y2)
9

(22 X 1021 X 1.5 X 10-9)112 Pa = 116.3 MPa


~O"voids = 3.1 X 0.8 X 0.2 X 81 X 109 X (0.356 X 10-9/Y2)
(1.67 X 1021 X 19.3 X 10-9)1I2Pa = 57.5 MPa
~O"disln = 3.1 X 0.8 X 0.2 X 81 X 109 X (0.356 X 10-9/Y2)
(1.3 X 1014)112 Pa = 115.4 MPa
O"y = 150 + 116.3 + 57.5 + 115.4 = 446.2 MPa

The larger measured value is probably due to the inadequacies of the small
specimen punch test. This is a shearing test rather than a pure tensile test and friction
forces play a large role in controlling the deformation so that larger stresses than
expected are sometimes measured. Also, the lattice friction stress has not been
included in the calculation of 0"y' This could bring the total yield stress up to values
close to that observed.

13
Dislocation Theory for Engineers: Worked Examples

DISLOCATION/GRAIN BOUNDARY INTERACTIONS - THE


HALL-PETCH EQUATION

We will make two approaches to this process, the first considering the effect of the
general increase in dislocation density at yield, and the second assuming these
dislocations are in the form of a pile-up.

DISLOCATION FOREST ApPROACH

One simple derivation of the Hall-Petch relation starts from equation (1.5) - the
relationship between the flow stress and the dislocation density. When a polycrystal
of grain size d undergoes plastic yield, dislocations are assumed to be injected into
the grains by the grain boundaries, and the yield stress will be given by equation
(1.5). If a length m of dislocations is produced per unit area of grain boundary at
yield, then each grain will generate a total length of dislocation = m X 47'i(d/2)2. This
length will, on average, be shared between two grains, so the dislocation density
generated at yield will be given by:

_ ~m4TI(~r
p- 4TI(~r/3
= 3m/d

Substituting in equation (1.5), we obtain:

T = To + CGb(3m)1I2d-1I2

i.e. (1.14)

which is the Hall-Petch equation. The lattice friction stress, To, can be equated with
the Peierls stress of the material (jF

2G (-27'iW)
«»= (l_v)exp -b-·

where v is Poisson's ratio and W is the dislocation width (it is usually assumed
w = lOb).

14
Dislocations in Pure Materials

EXAMPLE 1.11 THE HALL-PETCH RELATIONSHIP (1)

The lower yield stress, ay, of annealed iron (grain size 16 grains mm=) is 100 MPa,
and 250 MPa for a specimen with small grain size (4096 grains mm=). Determine the
lower yield stress of iron with grain size 250 grains mm >.

Tensile stresses will be related to the grain size by a similar expression to equation
(1.14), which refers to shear stresses, by introducing an orientation factor m relating
the applied tensile stress 0" to the shear stress, T, i.e. 0" = tin,
Let us first express the grain size in terms of d:

grains mm"" grains mm " d (mm) d-1I2 O"y (MPa)


16 4 0.25 2 100
250 15.8 0.0632 3.98 ?
4096 64 0.0156 8 250

Substitution in equation (1.14) in terms of the tensile stress gives:

250 = 0"0 + 8ky


100 = 0"0 + Zk;
:.150 = 6ky, so k; = 25 and 0"0 = 50

For a grain size of 250 mm ? we obtain O"y = 149.4 MPa which is the answer
required.

DISLOCATION PILE-UP ApPROACH

The principle of grain boundary strengthening is outlined in Fig. 1.6. If the average
grain size is d, a number n of dislocations pile-up at the grain boundary over a
distance d/2 as a result of the applied resolved shear stress, Ts, operating a source of
dislocations at the grain centre, Sl.
Now the local stress at the head of the pile-up, acting on the lead dislocation, is
n times the applied stress at the source. When this local stress exceeds a critical value
Tc (the grain boundary shear strength), the blocked dislocations are able to glide past

the grain boundary. Thus

Eshelby et al? have shown that

n = 'TrTs(d/2)/Gb (1.15)

15
Dislocation Theory for Engineers: Worked Examples
Grain
boundary

Fig. 1.6 Dislocation model for the propagation of yielding past a grain boundary.

But 'Ts is equal to the applied stress (r) minus the lattice friction stress ("0).
So by substituting for n and for "s, we can write

(1.16)

At yield, ,. = "y, so equation (1.16) can be rearranged to give:

,.y _- "0+ (2GbTc)1I2d


-- 1T -112

which is the Hall-Petch equation (1.14) with

k; = (2Gb,. c/1T )112

EXAMPLE 1.12 THE HALL-PETeH RELATIONSHIP (2)

The following tensile data for the lower yield point for 70-30 brass were obtained at
20°C. 64.7 MPa for a grain size of 100 uni and 188 MPa for a grain size of 6.25 J.Lm.
Calculate the grain boundary shear strength Tc at 20°C given:
Shear Modulus of brass = 48.3 GPa
Dislocation Burger's vector in brass = 0.255 nm

We will assume an orientation (Taylor) factor (m) of 3.1 for the fcc material, so
equation (1.14) becomes:

16
Dislocations in Pure Materials
Substituting the given data, we obtain:

64.27 = c, + 3.1ky X 100


188 = a, + 3.1ky X 400

Subtracting gives

123.8 = 930ky

So k; = 123.8/930 = (2GbTc/7r)1I2

Giving Tc = 2.26 Pa which is the answer required.

17
2
Dislocation Interactions in Single
Phase Materials

The behaviour of dislocations in crystals containing solute elements may be affected


in two ways. If the solute atoms are able to diffuse in the lattice, the energy of
interaction between the strain field of the dislocation and that of the solute atom
may lead to the solute atoms segregating to the dislocations, leading to possible
locking and yield point phenomena. Only a very small atomic fraction of solute is
required to cause this effect which may raise the yield stress (i.e. the stress to initiate
plastic flow in a previously undeformed crystal) but have little influence on the flow
stress (the stress required to cause continuing plastic flow in a crystal).
In concentrated solid solutions, randomly dispersed solute atoms can act as point
obstacles to dislocation movement, and this can lead to a raising of the flow stress
of the crystal. We will consider these two situations in turn.

SEGREGATION OF SOLUTE ATOMS TO DISLOCATIONS

EXAMPLE 2.1 THE MAXIMUM TEMPERATURE OF SEGREGATION

Assuming the atomic concentration of solute, c, around a dislocation line at


temperature T can be described by the expression

c = Co exp(WlkT)

where Co is the atomic concentration of the solute in the dislocation-free crystal, k is the
Boltzmann constant and W is the binding energy, estimate the temperature below
which yield points would be observed if the maximum binding energy is 0.5 eV and
Co = 10-5•

Strong yielding behaviour is expected up to temperatures at which a condensed


atmosphere of solute exists at the dislocation. This corresponds to a value of c of

19
Dislocation Theory for Engineers: Worked Examples
unity, we may therefore substitute in the equation:

1 = 10-5 exp(0.5/8.62 X 10-5 X T)


i.e. T = 0.5 X 105/(8.62 X 11.51) = 504K = 231°C.

This is the required temperature, below which yield points would be observed.

EXAMPLE 2.2 SOLUTE INTERACTION WITH AN EDGE DISLOCATION

An edge dislocation lies along the z direction and its Burger's vector b is parallel to x.
The stress field of the dislocation is

0"xx = - Dy(3x2 + y2)/(X2 + y2)2


O"yy = Dytx? + y2)/(X2 + y2)2
O"zz = v(<Txx + <Tyy)
<Txy = Dx(x2 - y2)/(X2 + y2)2
O"xz = O"yZ =0

where D = Gb/2'1T(1 - v), G is the shear modulus and v is Poisson's ratio. Derive an
expression for the pressure near the dislocation and sketch contour maps of its
distribution. Explain the physical importance of the pressure in relation to solution
hardening.

y
y

x
(a) (b)

Fig 2.1(a) Showing an edge dislocation lying along the z direction, with its Burger's
vector parallel to x. (b) Showing lines of equal pressure (full lines) and lines of flow
(broken lines) for solute atoms migrating to an edge dislocation. The arrows denote
the direction of flow.

20
Dislocation Interactions in Single Phase Materials
The hydrostatic pressure, p, is defined by:

and this may be expressed in polar co-ordinates (Whelanl p. 30) as

p = 2(1 + v)D sin 8/3r


As shown in Fig. 2.1(a), r = 2R sin 8
.'. sin 81r = 1/2R
so the pressure is constant round a circle of radius R, and contours of constant
pressure may be drawn as shown in Fig. 2.1(b).

Solution hardening
The interaction energy between a solute atom of misfit volume dv and the
dislocation will be approximately p dv.

If the radius of the solute atom is r(l + s) instead of r, I1v = 30s (where 0 is the
atomic volume)
The interaction energy (Eint) ::::::3p Os
For maximum interaction, R::::::b, and 8 = 90°:

2(1 + v)D sin 8


p = 3r

= 2(1 + v)D
3 X2b

Substituting for D, we obtain:

GOs(l + v)
Einteraction = 21T(1 - v)

Substituting typical values for the parameters gives an interaction energy of the
order 0.1 eV This is therefore a relatively weak interaction.

EXAMPLE 2.3 THE RATE OF STRAIN AGEING

Assuming the interaction energy of a solute atom at a radial distance r from a


dislocation is proportional to r ", show that the increase in concentration of solute at
the dislocation core varies as (Dt/T)2/3 where D is the bulk diffusion coefficient of the
solute, T the temperature, and t the ageing time. How may this relation be tested
experimentally?

21
Dislocation Theory for Engineers: Worked Examples
In Fig. 2.1(b), the pressure contours are surfaces of constant interaction energy,
hence the flow of solute atoms takes place down the pressure gradient, i.e. along
another set of circles as indicated as dotted lines in the figure.
The small routes are exhausted first, then only the flow lines with larger radii are
still active. In the strain field of a dislocation, a solute atom is subjected to a force
F = - grad Einh controlling a drift velocity, v, given by the Einstein equation

v = DFlkT (2.1)

If Eint ex: 11r, we can put Eint = -Air, where A is a constant. The mean gradient of Eint
along a flow line of radius r is of the order dEin/dr, therefore F = Air.
In time t, all solute atoms within a radius r = ro will reach the dislocation, so, from
equation (2.1), the mean speed of a solute atom round the line is given by

dr/dt = DAlr kT

f
ro DA
rdr=-dt
o kT
3DA
:.r5= kT t

The concentration, c, of solute at the dislocation will be given by

C = Co + ('ITY61
b2) Co

= Co + co( 'ITlb2()3DAtlkT)2/3 (2.2)

which is the required relationship.


An internal friction method may be used to measure the fraction of carbon atoms
remaining (in random solution) in iron specimens after ageing them by various
amounts. It is found that this amount varies linearly with t2/3 as required in
equation (2.2).

DISLOCATION DYNAMICS OF YIELDING

One of the reasons why iron containing carbon or nitrogen shows very marked yield
point effects is because there are strong interactions between these solute atoms and
the dislocations. The distortions caused by these atoms, which occupy the octahedral
interstices, are non-centrosymmetrical- having both a hydrostatic and shear nature
- thus locking both screw as well as edge dislocations, so that the mobile dislocation
density is very low.

22
Dislocation Interactions in Single Phase Materials
When such an alloy yields plastically the density of mobile dislocations is
suddenly increased, and the yield point phenomenon can be explained by
considering the dynamics of the dislocation motion.
The strain-rate 8 is given by

(2.3)

where 8e and 8p are the elastic and plastic strain rates respectively. We can write (see
section 1.6)
(2.4)

where Pm is the mobile dislocation density, b is the Burger's vector, and v the
dislocation velocity. If pre-existing dislocations are firmly locked, plastic flow at the
yield point may be controlled by the multiplication of new dislocations. The density
of mobile dislocations, Pm, may be expressed as

Pm =fp (2.5)

where f is a constant and P is the total dislocation density. p will be a function of the
plastic strain, 8p' i.e.

P = Po + C8p (2.6)

where C is a constant, and Po is the dislocation density prior to straining.


For a given value of Ep, if Pm increases, equation (2.4) indicates that v must
decrease. The dislocation velocity will depend upon the applied shear stress, and we
may write empirically:

(2.7)

where "0 and m are constants. If work hardening is taken into account, the effective
stress should be used in equation (2.7), i.e.

(2.8)

where e is the work-hardening rate.

EXAMPLE 2.4 UPPER AND LOWER YIELD POINTS

Derive an expression for the form of stress-strain curve of material which exhibits a
yield point due to dislocation multiplication. What are the values of the upper yield
point and lower yield point, and upon what parameters does the magnitude of the yield

23
Dislocation Theory for Engineers: Worked Examples
r

Fig. 2.2 The form of equation (2.9).

drop depend?

We may substitute equations (2.5), (2.6) and (2.8) into (2.4) to obtain:

which can be re-arranged to give:

(2.9)

Initially, as a specimen is strained at a constant strain rate, B, most of the strain


produced is elastic, since Bp ~ Be, and the stress follows the elastic line (Fig. 2.2). As
a result of dislocation multiplication and the increase in stress, the contribution from
Bp increases until the whole of the strain rate is due to plastic flow.
Further increase in strain leads to a decrease in stress, due to the decrease in
dislocation velocity (the second term in the above equation).

The upper yield point


The upper yield point thus corresponds to the stress at which 8p = 0, giving:

_ 8• p ) 11m

'TllYP - 'To ( Pofb (2.10)

The stress decreases until the work-hardening term (i.e. the first term in the equation
to the curve) becomes more important.

24
Dislocation Interactions in Single Phase Materials
The lower yield point
This is when dr/ds == O.
We can rearrange equation (2.9):

_
'T - as + 'To (Bp- ) 11m ( 1 ) 11m
(2.11)
p fb Po + Ce.,

So by differentiation and putting dr/ds = 0, a value of sp at the lower yield point


('TlYp) is obtained. Substitution of this value into equation (2.9) will give 'Tlyp"
Thus yield points are produced which are more pronounced the smaller the value
of m. m is ----2 for Ge at high temperature, ----25 for LiF, "-'35 for Fe-Si alloys and
----200 for fcc metals. Yield points are therefore not observed in pure fcc crystals,
except in whiskers or dislocation-free crystals (Po = 0). The size of the yield drop will
increase with decreasing work-hardening rate (8) and increasing value of C, which
describes the multiplication rate of the mobile dislocations.

STRENGTHENING BY SUBSTITUTIONAL SOLUTE ATOMS

We have seen that dislocation pinning by solute atoms is normally associated with
dilute solid solutions, but the solute atoms segregate preferentially to the disloca-
tions. The yield stress is thus raised, but after yielding has occurred, the (solute-free)
mobile dislocations will encounter few solute atoms (since the alloy is dilute), so
little increase in the flow stress over that for the pure metal is to be expected. This
explains the yield point phenomena observed, for example, in steels.
In order, therefore, to induce significant solute hardening, we would require a
more concentrated solid solution and the increase in strength will arise from the
interaction between the glide dislocations and the individual solute atoms they
encounter upon the glide plane. Each solute atom may be regarded as a point
obstacle to the gliding dislocations, and a relationship may be derived between the
yield stress of such an alloy and the force F exerted by each point obstacle upon the
dislocations.

EXAMPLE 2.5 THE CRSS FOR A CRYSTALCONTAININGA RANDOM ARRAY


OF POINT OBSTACLES

Show that if there are N, obstacles per unit area of slip plane, each of strength F, the
flow stress T is given by

where G is the shear modulus and b the Burger's vector.

25
Dislocation Theory for Engineers: Worked Examples
T T

Fig. 2.3 Showing a dislocation held up at an array of point obstacles.

We will be using the symbol 'T for shear stress and 'Tc for the CRSS, based on the
model illustrated in Fig. 2.3, which shows a dislocation interacting with the solute
atoms, which are treated as point obstacles of identical strength lying in the slip
plane. The glide dislocation bows out between the particles as shown and, when the
included angle <t>between the two arms of the dislocation reaches a certain critical
value, the dislocation breaks away from the obstacle. We shall refer to this value of
<t>as the breaking angle, and at this critical point the obstacle strength, F, is related
to the dislocation line tension, T, by

F == 2T cos!<t> (2.12)

When the breaking angle <t>== 0, the particle behaves as an impenetrable obstacle,
while for values of <t>> 0, the obstacle can be sheared by the glide dislocation, with
the required shearing force equal to F.
The shear stress needed to cause the dislocation to break away from the obstacle
may be related to the obstacle strength F as follows. The applied shear stress, 7,
causes the dislocation of Burger's vector b to bow into an arc of radius of curvature
R, where

rrb == TIR (2.13)

From Fig. 2.3 it can be seen that

2R sin e == A
so
rrb == (2Tsin e)/A

But it can also be seen that e == 90° - !<t>,hence from (2.12) we have

vb == FIA

i.e. rrb A == F (2.14)

26
Dislocation Interactions in Single Phase Materials
B'

,,-
,
,,
/

,-
,- /
,;-

Fig.2.4 The Friedel process for dislocations interacting with point obstacles.

It is important to determine how A, the effective obstacle spacing along the


dislocation, depends upon the applied stress. Friedel approaches this by assuming
that during the yielding process the dislocation takes up a steady-state configuration,
namely that each time a dislocation breaks through one obstacle (B, in Fig. 2.4) it
meets one, and only one, other obstacle (B' in Fig. 2.4) as it bows out to the
configuration compatible with the stress applied (equation 2.13).
Each time an obstacle is sheared, an area A of the slip plane (shown shaded in
Fig. 2.4) is swept out by the dislocation. On average, the value of A is thus inversely
proportional to the number of obstacles per unit area of slip plane (Ns). The
so-called 'square lattice spacing', L, between the obstacles is given by

and

In Fig. 2.4 the area swept out per obstacle sheared will be given approximately -by
to; so
L2 = hs: (2.15)

From the property of a circle (Fig. 2.4),

A2 = 2hR, for h -e; R (2.16)

Eliminating h from equation (2.15) and (2.16), we obtain

i.e. (AIL)2 = sec!<f>= 2TIF (2.17)

27
Dislocation Theory for Engineers: Worked Examples
The effective spacing of the obstacles, A, thus depends upon the value of the
breaking angle, <p, and hence upon F. Because the dislocation has some flexibility,
the number of obstacles it touches per unit length increases as the obstacles become
stronger, hence the smaller the value of A. This decrease in A more than offsets the
increase in the 'square lattice' spacing of the obstacles.
Substituting for A in equation (2.14) we obtain a value for the yield stress,

(2.18)

Alternatively, this may be expressed in terms of the breaking angle through equation
(2.12) as

To obtain the required expression, equation (2.18) may be squared and substitution
made for T and for L. T is often taken as 4Gb2, where G is the shear modulus, and
L = N;1I2 giving

(2.19)

This is the required expression.


Insofar as N, is proportional to the solute content of the alloy, equation (2.19)
indicates that the yield stress should be proportional to the square root of the solute
content of the alloy. This relationship is not always observed in practice.
The magnitude of F is sensitive to atomic size differences between solvent and
solute, and also to differences in elastic properties. Fleischer" has defined a mismatch
parameter, e, which combines both of these contributions. This leads to a definition
by Mott and Nabarro? such that

_ Gbe 2/3
'T - 2A' c (2.20)

where A' is the average dislocation line length. This contrasts with equation (2.19)
which implies a c1l2 relationship with 'T since N, o: c. The experimental data available
make it difficult to distinguish between these two relationships: often a simple linear
relationship is assumed.

EXAMPLE 2.6 SOLUTE HARDENING IN STEEL

Plain carbon steels are alloyed with so-called microalloying elements like Nb, W, V
and Ta. Much of the purpose of this is to provide precipitation strengthening.
However, all of these elements contribute to solution strengthening. Assuming that 5
atomic percent of each of these elements remain in solution in a steel intended for a

28
Dislocation Interactions in Single Phase Materials
railway line application, calculate the contribution to the yield strength of the steel for
(a) Nb, (b) W, (c) Ta, and (d) V.

Data:
Atomic radii
ex-ferrite = 0.1241 nm
Nb = 0.142nm
W = 0.1368nm
Ta = 0.143nm
V = 0.131nm
Average dislocation line length = 50 nm
Shear modulus of iron = 8.1 X 104 MPa
Burger's vector = 2 X 10-10 m

Solution
Calculate the misfits for each element.

= 0.142 - 0.1241 = 0 144 Nb


e 0.1241 .

= 0.1368 - 0.1241 = 0 124 W


e 0.1241 .

= 0.143 - 0.1241 = 0 154""-:


e 0.1241 . ra

= 0.131 - 0.1241 = 0055 V


e 0.1241 .

Hence the yield stress contribution for solution strengthening is

8.1 X 1010 X 2 X 10-10 X (0.05)2/3 X e


'T = 2 X 5 X 10-8 Pa

This gives values as follows:


Nb 'T = 3.16 MPa
W 'T = 2.7MPa
Ta 'T = 3.4 MPa
V 'T = 1.2MPa

29
3
Dislocation Interactions in
Two-Phase Alloys

.Particle hardening

The effect of dispersed precipitates upon the yield stress can be assessed
by treating the microstructure once more as an array of point obstacles.
When a shear stress is applied, the glide dislocations bow out between the
particles as shown in Fig. 2.3. When the breaking angle <p = 0, the particle
behaves as an impenetrable obstacle, while for value of <p > 0, the particles
can be sheared by the glide dislocations, with the required shearing force
equal to F. We will consider these two types of interaction separately.

PARTICLE SHEARING BY DISLOCATIONS (<p > 0)

Application of the point obstacle theory leads to an expression for the shear yield
stress given by:

(2.19)

Precipitate particles can impede the motion of dislocations through a variety of


interaction mechanisms. Those for which theories for the value of F (or 4» have been
developed include:
(i) chemical strengthening, which arises from the energy required to create an
additional particle/matrix interface when the particle is sheared by the
dislocation;
(ii) stacking-fault strengthening, which occurs when there is a difference between
the stacking-fault energy of the particle and that of the matrix when these are
either both fcc or hcp in structure;

31
Dislocation Theory for Engineers: Worked Examples
(iii) modulus hardening, which arises from differences between the elastic moduli of
the matrix and particle;
(iv) coherency strengthening, which arises from the elastic coherency strains
surrounding a particle that does not fit the matrix exactly;
(v) order strengthening, which is due the additional work required to create an
antiphase boundary in the case of dislocations passing through precipitates
which have an ordered lattice.
We will, by the use of simple dislocation models, derive expressions for F in
equation (2.19) for several different dislocation-particle interactions, and hence
estimate of the yield stress for these situations.

EXAMPLE 3.1 CHEMICAL STRENGTHENING

Estimate the chemical strengthening arising from the cutting of a volume fraction f of
precipitate particles of diameter 2r by a screw dislocation of Burger's vector b.
Figure 3.1 shows a section of a particle in the slip plane of a screw dislocation. The
slipped particle (above the slip plane) is shown as a dotted line. When the
dislocation advances by Sx, the surface area increases by 2b Sx. If the specific surface
energy of the precipitate-matrix interface is "Is, then the force F exerted by the
particle on the dislocation is given by:

The number of particles/unit area of slip plane (Ns) may be estimated as follows. The
number/unit volume (Nv) is the volume fraction, f, divided by the volume of one
particle, i.e.

But

--- Dislocation
----',l: ----------------,~
\ I
\ I
\ I
\ I

Above slip~ / Below slip


plane -. ",""
",/ plane
---- -
••....•...

Fig. 3.1 Particle in slip plane.

32
Dislocation Interactions in Two-Phase Alloys
where p is the probability that a plane will intersect a single particle placed randomly
within a unit cube. Since, of the various positions of the cross-sectional plane, only
those positions existing over the length 2r would lead to the plane intersecting the
single sphere, the probability of intersection is equal to 2r, so,

Giving

(3.1)

Substitution in equation (2.19) leads to:

1" = ( 12f )112!. ",3/2


c 7TGb r IS

The theory thus predicts that, for a fixed volume fraction, f, the critical resolved
shear stress decreases as the particle size increases, contrary to what is normally
observed in age-hardening. It therefore appears that chemical strengthening is not
an important mechanism, except perhaps for particles of very small size.

EXAMPLE 3.2 STACKING-FAULT STRENGTHENING

Derive a simple expression for the shear yield stress of a crystal hardened by a volume
fraction f of particles of diameter 2r, if there is a difference in stacking-fault energy
~~SF between the lattice of the matrix and that of the particle (assume that the SFE of
the particle is < that of the matrix).

If a dislocation in the matrix dissociates into two partial dislocations with an


equilibrium separation Wm, when it moves into a particle of lower SFE it will split
further into a ribbon of width wp- This interaction will lead to a force being exerted
by the dislocation on the particle, whose maximum value will be:

Where I is the length of dislocation inside the particle at the critical breaking
condition. I will be a function of the particle radius, say I = or, where ex is a constant'
of the order 1.5.
Again equation (3.1),
Dislocation Theory for Engineers: Worked Examples
so substitution in equation (2.19) leads to:

= ( 3fr )112 (u Ll"lSF)3/2


Tc 2'ITG b2
This model thus leads to a variation of the CRSS with Ll"l3J} and (f. r)1I2.
Experimental data due to Gerold and Hartmann" on aged AI-Ag single crystals
show good agreement with this model.

EXAMPLE 3.3 COHERENCY STRENGTHENING

Estimate the magnitude of F (and hence 'fe) arising from coherency strains associated
with precipitates, if the shear strain in the matrix at a distance R from the precipitate
(of radius r) is given by er'/R3, where e is the misfit parameter of the coherent
p recip itate.

The maximum strain and stress occur at the particle interface, i.e. at R = r, and we
can take the appropriate stress as approximately Ge. From the Peach-Koehler
equation, we can write the magnitude of the force per unit length experienced by a
dislocation of Burger's vector b as it approaches a particle as Geb.
This force acts effectively over a length of dislocation equal to the particle
diameter, so we can write:
F:::::::2Gerb
or
F:::::::kGerb

where k is a numerical constant which a more accurate treatment gives as between


3 and 4. Substituting into equation (2.19), and replacing N, as before (equation 3.1),
we obtain:
Tc = k3/2 Ge3/2(rlb )112(3/2'IT )112 t'"
The increment of CRSS due to the presence of the particles is thus predicted to be
proportional to t'" and also to r1l2. Experimental data indicate that these
proportionalities are observed in practice, although the above expression over-
estimates the actual values of Tc. It should be noted that the (rib) and f terms are
both related to composition, c. The above equation can therefore be used to
demonstrate a c1l2 dependence on Te.

EXAMPLE 3.4 ORDER HARDENING

Estimate the maximum value of F (and hence 'fe) arising [rom the [ormation of
an antiphase boundary (apb) of energy 'Yapb in an ordered spherical particle of
radius r.

34
Dislocation Interactions in Two-Phase Alloys
The passage of a dislocation creates an apb in the particle, and an estimate of the
maximum force exerted by a spherical particle may be obtained by assuming that the
dislocation lies along a diameter, in which case

F = 2rYapb
Making the same substitutions as in the previous examples we obtain:

= (12 fr )112~~~2b
'T c 'ITO b2

This equation accounts for the increase in 'T c in ageing in terms of the increase in r
for a given volume fraction. Order hardening is a very important hardening
mechanism in nickel alloys in which large volume fractions of the ordered ~'-phase
(Ni3AI, Ni3Ti, or Ni3(AI, Ti)) may be present. A complication which commonly arises
in this class of alloy is that the dislocations travel in pairs - the first dislocation
creates an apb and the second restores the order. The approximation in the Friedel
statistics that the volume fraction of precipitates is small is not applicable to many
nickel-based superalloys, where the volume fraction of v' precipitates can exceed 0.7
in some alloys.
The strengthening behaviour of cubic ~' precipitates in such alloys is not modelled
well by the above or similar equations, where there is evidence that the yield stress
decreases with increasing precipitate sizes.

PARTICLE BY-PASS BY DISLOCATIONS - OROWAN


HARDENING (<p = 0)

As the magnitude of F increases, the breaking angle <p decreases, and eventually
becomes zero. For larger values of F the dislocation bypasses the particle rather than
shears it.

EXAMPLE 3.5 EFFECT OF PARTICLE SPACING ON YIELD STRESS

Illustrate graphically the variation in the tensile yield stress (O"y) with particle spacing
(L) for the case of particle by-pass by dislocations, in the case of aluminium.
[Lattice parameter 0.404 nm, shear modulus 26.2 GPa.j

When dislocation by-pass takes place, the force exerted by the particle is balanced
by the line tension of the two arms of the dislocation on each side of the particle.
Therefore F = 2T = Gb2, and substitution in equation (2.19) leads to:

(3.2)

35
Dislocation Theory for Engineers: Worked Examples
500~----------------------------------

m 400r------+----------------------------
a,
~
~ 300r-------~~------------------------
en
~
W 200r-----------~----------------------
"0
Q)
~ 100r---------------~------------------

50 100 300 700 1000 2000


Particle spacing - nm

Fig. 3.2 Orowan plot.

where L is the square lattice spacing of the particles (assuming that the particle
diameter is negligible in comparison with L).
For the case of glide dislocations in aluminium, b = alV2 = 0.29 nm, giving:

r, = 26.2 X 109 X 0.29 X 10-91L = 7.61L (Pa)

Let us calculate Tc for a series of values of L:

L (nrn) 50 100 300 700 1000 2000

'Tc (MPa) 152 76 25.3 10.9 7.6 3.8

<Ty (MPa) 456 228 75.9 32.7 22.8 11.4

The tensile yield stress may be obtained from Tc by multiplying by the appropriate
Taylor factor, which we will take as 3, for the case of an fcc crystal structure. The
third row in the table is thus obtained and the data shown graphically in Fig. 3.2.
The yield stress of pure aluminium is about 3 MPa, and it is apparent from the
graph that significant Orowan hardening is only achieved when sub-micron
inter-particle spacings are present.

EXAMPLE 3.6 CALCULATION OF CRITICAL INTERPARTICLE SPACING

Discuss the effect of particle spacing on the type of dislocation-particle interaction.


Calculate the precipitate spacing (L) necessary to give maximum hardening for
precipitates with a misfit (B) of 0.2 and with a volume fraction (f) of 0.02 in copper
(lattice parameter 0.354 nm, shear modulus 45 GPa).

36
Dislocation Interactions in Two-Phase Alloys
Mott and Nabarro ' realised that the obstacles become impenetrable when the back
stress arising from the energy contained within the particle becomes larger than
the applied stress, 'T. The energy generated per unit volume by the particles is
given by

E= Gef

This stress is equivalent to an energy per unit volume and is equal to the applied
stress needed to cause particle cutting rather than Orowan looping. Thus

'T = Gef= aGb/R

where R is the radius of curvature of the dislocation.


Therefore

R = aGblGsf = able]

Remembering that the looping stress is at a maximum when R = A/2 and that
a = 0.5, we obtain that the critical spacing, ACrit' for the maximum applied stress Gust
before the back stress is insufficient to resist the applied stress, and thus
particle-cutting starts to take place) is:

Acrit = 2bl2sf = bl sf

Substituting the values given, with b = atV2, we find:


Acrit = (0.354/1.414 X 4) 10-6 m = 63 nm,

which is the required answer.


N ate that 'T c will decrease with particle size and spacing for spacings below ACrit.

This corresponds to the situation that <f> > 0 and is described in Section 3.3.

COMBINATION OF STRENGTHENING PROCESSES

EXAMPLE 3.7 COMBININGGRAIN-SIZE, SOLUTE AND


PARTICLESTRENGTHENING

A specimen of a-brass (Cu-30 at. % Zn) contains a volume fraction of 0.028 of


spherical oxide particles of diameter 0.1 urn. The grain size is 20 urn. Estimate the
tensile yield stress given the following data:
Shear modulus of brass = 48.3 GPa
Lattice parameter, a = 0.361 nm
Slope of Hall-Petch plot, k, = 0.398 MPa m'"

37
Dislocation Theory for Engineers: Worked Examples
Lattice friction stress for pure copper = 14 MPa
Solute hardening combined mismatch parameter = 0.5
Value of A in equation (3.3) = 3.97 X 10-3

We will carry out this estimation in several stages:


(i) Calculate the increase in lattice friction stress due to the zinc in solid solution,
using the coherency strengthening mechanism described in Example 3.3.
(ii) Calculate the yield stress of the polycrystalline single-phase brass.
(iii) Calculate the increment in tensile yield stress arising from the dispersed oxide
phase.
(i) The presence of Zn in solid solution will raise the flow stress. Let us assume
that the increase in shear stress is given by (Example 3.3):

(3.3)

Substitution of the given parameters gives:

T = 0.548 X 4.83 X 109 X 0.354 X 3.97 X 10-3 Pa = 3.72 MPa

Taking a Taylor factor of 3, the corresponding increase in tensile yield stress,


~O"ss = 11.2 MPa. The lattice friction stress for brass 0"0 is thus = 14 + 11.2 =
25.1 MPa.

(ii) The yield stress of the polycrystalline brass will be given by the Hall-Petch
equation:

0"
Y
= 0"0 + kY d -112
d = 2.10-5 m, so d -1/2 = 223.6 m -1/2
O"y = 25.1 + (0.398 X 223.6) = 114 MPa

(iii) The increment in yield stress arising from the oxide dispersion can
be assumed to arise from dislocations having to loop between the particles
rather than shear them. It can thus be estimated from the Orowan equation
(Tc = N~/2Gb = Gb/L, see Example 3.5), where N, is the number of particles per unit
area, and L the square lattice spacing of the particles.
The Burger's vector b = atV2 = 0.361/1.414 nm = 0.255 nm. The number of
particles per unit volume, N; = volume fraction of particles, f, divided by the volume
of one particle, so

and

38
Dislocation Interactions in Two-Phase Alloys
200~----------------~

Dispersion strengthening
increment (Orowan)
150

co
o,
~
Cf)
Cf)
~ 100
Ui
"0
Q)
~ Grain size
strengthening
(Hall-Petch)
50

Solid solution strengthening by Zn


Friction stress of pure copper
0
0 100 200 300 400 500
d-1f2(m-1f2)

Fig.3.3 The grain-size dependence of the yield stress in a dispersion-strengthened


Cu-Zn alloy.

Substituting the given values of particle size and volume fraction gives
N, = 5.35 X 1012 m-2•
This corresponds to a square lattice spacing of particles, 0.432 urn.
The increase in shear stress = Gb/ L
= 28.5 MPa.
Again assuming a Taylor factor of 3, this predicts an increase in tensile yield stress
~86MPa.

Assuming that the total yield stress may be obtained by the linear addition of
the Orowan contribution to the Hall-Petch increment, we estimate the yield stress
to be:

O"y = 114 + 86 = 200 MPa

The individual contributions to this stress are indicated in Fig. 3.3.

39
4
Effects at Elevated Temperatures

RECOVERY AND RECRYSTALLISATION

SUBGRAIN GROWTH DURING RECOVERY

Energy of symmetrical tilt boundary (misorientation 0)

This is treated as an array of parallel edge dislocations, all sharing the same Burger's
vector, b. If D is the dislocation spacing, then as illustrated in Fig. 4.1:

sin(S/2) = b/2D
for small S we have:

e = biD
In unit area the dislocations are of unit length, and the number per unit
area = 1/D = S/b. The energy per unit area of the boundary is simply the sum of the

Fig.4.1 A symmetrical tilt boundary in a simple cubic lattice.

41
Dislocation Theory for Engineers: Worked Examples
energies of the individual edge dislocations (Ed)' where:

where E; is the core energy per unit length of dislocation line, and ro the inner cut-off
radius from the dislocation. Therefore the boundary energy is:

(4.1)

where A and B are constants:

Gb Gb
A = Efb + 411"(1 _ v) In( bI2ro); B = 411"(1 - v)

Equation (4.1) is the Read-Shockley formula and its functional form applies to all
small angle boundaries, not just symmetrical tilt boundaries.

EXAMPLE 4.1 DRIVING FORCE FOR SUB GRAIN GROWTH

Calculate the reduction in energy (.dE) resulting from the coalescence at a 'Y' junction
of two small-angle boundaries of misorientations eJ and e2 into a single small-angle
boundary of misorientation (el + (2) as illustrated in Fig. 4.2.

8
1
..l ..l
1- ..l
..l 1-
1- ..l
1-
..l
1-
1-
..l
81 + 82

Fig.4.2 A 'Y' junction between three tilt boundaries of misorientations OI, O2, and
01 + 020

42
Effects at Elevated Temperatures
Egb1 = 81[A - B In 8d
Egb2 = 82[A - B In 82]
and E1+2 = (81 + (2) [A - B In(81 + (2)]
~E = Egb1 + Egb2 - E1+2

This is the driving force for sub grain growth during recovery annealling.

RECRYSTALLISATION

Stored energy of deformation

Heavily cold-worked metals recrystallise when annealed, the process being


driven by the stored energy of deformation. The principal source of this
energy lies in the form of dislocations, and calorimetric measurements during
a recrystallisation anneal provides a means of estimating the density of
dislocations in the cold-worked metal.

EXAMPLE 4.2 ENERGY RELEASED DURING RECRYSTALLISATION

The energy released when a cold-worked copper specimen is recrystallised is 1Mlm-3•


Taking the modulus as 48 GPa and a lattice parameter of 0.36 nm, estimate the
dislocation density in the cold-worked state.

For a dislocation density p the stored energy is

E = pEdis

where Edis is the energy per unit length of dislocation line.


Edis may be written:

e.; = Gb f(v) In(R)


2

4'lT ro

43
Dislocation Theory for Engineers: Worked Examples
where R is the upper cut-off radius (usually taken to be the separation of
dislocations (p -112).
ro is the inner cut-off radius (usually taken as between band 5b)
f( v) is a function of Poisson's ratio (v), which, for an average population of edge and
screw dislocations is ---(1 - v/2)/(1 - v).

The energy of a dislocation depends on its environment - being highest in a pile-up


and lowest when in a cell or subgrain wall. In the present case, only a very
approximate energy is needed, so the above equation can be simplified to

where c is a constant of the order 0.5.


So we may write:

E
p = 0.5Gb2

In the present case, b = atV2 = 0.255 nm

p = 0.5 x 48 X 109 X 0.065 X 10-18

i.e. p = 6.4 X 1014 m-2

The driving pressure for recrystallisation

During the recrystallisation of cold-worked metals, therefore, the pressure


acting on a migrating grain boundary arises from the line tension of the
dislocations being annihilated. We have seen in Example 4.2 that this
corresponds to an energy per m3 of 1 MJ, and so a pressure of 1 MPa will
be exerted per m2 of grain boundary, which can thus be regarded as the
driving force for primary recrystallisation.

44
Effects at Elevated Temperatures

DISLOCATION CLIMB - CLIMB RATE CALCULATION

EXAMPLE 4.3 MEASUREMENT OF STACKING- FAULT ENERGY FROM THE RATE


OF ANNIHILATION OF DISLOCATION Loops

A prismatic edge dislocation loop of vacancy character in an fcc metal is located near
the centre of an equiaxed grain with grain size much larger than the loop diameter.
Estimate the time taken for the loop to anneal out at 200°C if the initial radius of the
loop is 50 nm. How would the annealing time differ if the loop were a Frank loop on
{J11} enclosing an intrinsic stacking fault?
[Assume shear modulus = 25 GPa, Burger's vector of the prismatic disloca-
tion = 290 pm, stacking fault energy 200 ml m:', self-diffusion coefficient at
200°C = 3.5 X 10-20 m2 S-l.j

When an edge dislocation encounters a high supersaturation of vacancies, it will act


as a vacancy sink and the dislocation will climb. The dislocation may be thought of
as being influenced by an effective stress somewhat analogous to osmotic pressure.
The magnitude of this chemical stress can be derived from Example 2.1 as
rr = (kT/b3)[ln(c/co)]. The force per unit length on the dislocation, F, will be given by
ob, i.e.

F = (kT/b2)[ln(c/co)] (4.2)

which may be rearranged to express the concentration c of vacancies in the vicinity


of an edge dislocation which is SUbjected to a force F per unit length:

c= Co exp(Fb2/kT)

But from the exponential series we may write for small x: e' ~ (1 + x).
Therefore

(4.3)

Turning now to a consideration of the shrinking dislocation loop of radius r in a


spherical crystal of radius R2(R ~ r), the problem is one of spherically symmetrical
diffusion. The dislocation experiences a force per unit length due to its curvature,
giving rise to a local increase in vacancy concentration. (Fig. 4.3).

45
Dislocation Theory for Engineers: Worked Examples
n

Fig. 4.3 Showing the number of vacancies per unit volume as a function of radius (p)
from the centre of the dislocation loop (radius r) in a crystal of radius R.

There will thus be a flow of vacancies which can be expressed by Fick's Law.

-41Tp2D .dnldp = N = constant (4.4)

Where N is the total number of vacancies per second flowing and D the coefficient
of diffusion.
If n = number of vacancies per unit volume at a radius p = n(p)

n varies as lip, say n = no + (X/p (4.5)

Where (X is a constant, and no is the equilibrium number per unit volume (remote
from the dislocation loop).
If n is the volume of one vacancy, we can rewrite equation (4.3):

n = no + (co/fl)(Fb2IkT), which gives the value of n where p = r

Therefore from equation (4.5),


rcoFb2
:. (X = OkT

So we can write the equation to the curve of Fig. 4.3:

rcoFb2
n(p) = no + pflkT

So we can substitute into equation (4.4), giving

46
Effects at Elevated Temperatures
But DCo = Dsd, the coefficient of self-diffusion, since a given vacancy jumps Nln
times as often as a given atom.
If .(27fr/b) vacancies arrive at the loop, its radius will shrink by b, so
dr b2 N 2Fb4 ti;
- dt = 27fr = f!kT (4.6)

The force per unit length exerted on the dislocation arising from its line tension, I',
when the dislocation is in the form of a loop of radius r, is given by fir. Taking
F = !Gb2, and assuming the atomic volume n = b3, equation (4.6) becomes
Gb3 Dsd
dr
- r dt = kT

The time taken for the loop to shrink may thus be obtained by integrating this
equation, giving

[r]r2 o
= _ Gb3 Dsd [ ]0
kT t t

= 375 seconds, which is the required answer.

If the dislocation was in the form of a Frank loop, it would shrink more rapidly,
because the stacking fault of energy 0.2 J m-2 provides an extra (constant) force per
unit length of magnitude 0.2 Nm-1 (independent of the radius of curvature). If this
is substituted into equation (4.6) we find:

-drldt = (2Fb . Dsd)/kT

0.4 X 290 X 10-12 X 3.5 X 10-20


1.38 X 10-23 X 473
= 0.623 X 10-9 m S-1

So t = r10.623 X 10-9 = 5010.623 = 80 seconds.


This indicates a four-fold reduction in the shrinkage time of the loop when it
contains a stacking-fault.

DISLOCATION CREEP

The significant' engineering parameter during high temperature deformation of


polycrystals is the secondary (steady state) creep rate, which is the slope of the curve
during secondary creep, indicated in Fig. 4.4.

47
Dislocation Theory for Engineers: Worked Examples
Tertiary
creep

\ Primary creep
Initial {
elastic
strain L-- ~

Time, t

Fig. 4.4 Typical creep curve.

At low stresses and high temperatures, the creep process is controlled by


stress-directed atomic diffusion. This involves the migration of vacancies along a
gradient from grain boundaries experiencing tensile stresses to boundaries in
compression. Atoms move in the opposite direction, leading to elongation of the
grains parallel to the axis of principal tensile stress. The concentration gradient, and
thus the creep rate, is grain-size dependent and this diffusional creep does not
therefore involve the movement of dislocations. It is characterised by the creep
strain rate being directly proportional to the applied stress.
At higher stress levels, dislocation creep mechanisms operate, the main ones
responsible for assisting creep deformation being glide and climb, Climb is a
thermally activated process because it needs vacancies and these are only available
in large quantities at significant proportions of the melting point, i.e. at temperatures
above 0.4 T m» where T m is the melting point in K. The grain boundaries assist
diffusion and grain boundary sliding, and so a fine grain size provides more
opportunity for these processes to take place and thus to raise the strain rate during
secondary creep. Naturally, as we have seen earlier, the applied stress will accelerate
plastic deformation, so there are three main components in any description of the
secondary creep rate:
(1) Temperature initiated vacancies, reflected in a term involving the self-
diffusion coefficient, D, for the matrix material.
(2) the grain size, d.
(3) the applied stress, a.
A universal equation, the so-called Monkman-Grant equation, is used to predict the
strain rate during secondary creep

(4.7)

48
Effects at Elevated Temperatures
where Ao is a constant - 1010
D is the self-diffusion coefficient
b is the Burger's vector
k is Boltzmann's constant
T is the absolute temperature
d is the grain size
a is the applied stress
ao is the back -stress caused by already-existing dislocation obstacles
G is the shear modulus.

The parameters nand p refer to the creep mechanisms in operation, which depend
on the stress and the temperature as already observed. For grain boundary
dominated, or Coble creep, D is the grain boundary diffusion coefficient, p is 3 and
n is unity. For lattice diffusion assisted, or Nabarro-Herring creep, D is the lattice
diffusion coefficient,. p is 2 and n is again unity.
At higher stresses, D is the lattice diffusion coefficient, p is zero, and n usually lies
between 3 and 8. This regime of creep is known as power law creep. The regimes of
stress and temperature where these mechanisms operate are material-dependent,
and maps defining them for different materials are to be found in Frost and
Ashby,"
Nomograph methods are also available for predicting creep failure (see Fig. 4.5).
A number of sophisticated models have been developed to describe the
phenomenon, but we will indicate how a simple power law creep relationship may
be derived from first principles.

EXAMPLE 4.4 RECOVERY CREEP MODEL

Assuming steady-state creep (deldi) to arise from a balance between the recovery rate
(-da/dt) and the hardening rate (dolde), derive an expression for the stress and
temperature dependence of the creep rate.

We assume that the dislocations exist as three-dimensional network of average link


length I. Recovery processes will cause this network to coarsen. The applied stress
will cause the links in the network to bow out and act as dislocation sources. Work
hardening during dislocation multiplication refines the network size, so that in a
steady state the hardening and recovery processes balance to give a constant
dislocation density, p, and a constant creep rate.

Considering first the hardening rate, h:


The magnitude of the applied stress will determine the average value of I, and hence
the dislocation density, p:

49
Dislocation Theory for Engineers: Worked Examples
We can therefore write:

h = drr/ds ex: 1/e1l2 ex: 1/cr

N ow consider the recovery rate, r:


The rate of growth of the network, dlldt = mobility (M) X force, i.e.

dUdt = M X Til

where T is the line tension of the dislocation, i.e.

l dl = M. T dt; so l2 ex: t; i.e. p -1 ex: t; and cr-2 ex: t, i.e. rr ex: t-1I2

We can now write

Using these expressions for rand h we obtain

Steady state creep rate, i; = rlh ex: cr4

The dislocation mobility, M, will depend on the climb rates. The temperature
dependence of M and therefore e
will be characterised at high temperatures by an
activation energy equal to that for lattice self diffusion (QSD), so that we obtain:

e = Acr 4
exp( - QSDIRT)

where A is a constant.
This power-law creep relationship is seldom obeyed in practice, and it is usually
expressed in the general form:

(4.7)

where the values of the constants A, nand Q vary from material to material and
have to be found experimentally.

EXAMPLE 4.5 POWER-LAW CREEP CALCULATION

A nickel alloy creeps in steady state according to the relation (4.7) above. The alloy
is used to make a tensile strut inside a furnace. The strut is in the form of a rod 2 mm
in diameter, and supports a load of 20 kg. Each day the furnace is run at 800°C for
14 h and 850°C for 10 h. How long will it be before the strut breaks? [For this alloy,

50
Effects at Elevated Temperatures
n = 4.6, Q = 270 kJ mot:', A = 8.5 X 10-4 with a in MPa. Strain to rupture at 800°C
is 8% and is 12% at 850°C.]

The stress in the strut, = 20 X 9.81hr MPa = 62.45 MPa


(J"

Neglecting any tertiary creep, the time to fracture by creep tf will be given by:

where sf is the strain to fracture at temperature T.


At 800°C:

tf = 0.08/ 8.5 x 10-4 X 62.454.6 exp( -270000/8.31 X 1073)


= 7.31 X 10 s 6

= 2030.6 h

tf = 0.12/8.5 X 10-4 X 62.454.6 exp( -270000/8.31 X 1123)


= 2.84 X 10 s
6

= 788.9 h

Each day in service, the strut spends a fractional life of 14/2027.6 at 800°C and
10/788.9 at 850°C. If n is the total lifetime in days, and assuming that the duration
of tertiary creep is negligible and that the mechanism of creep failure is identical in
the two situations, we may write:

n[(14/2030.6) + (10/788.9)] =1

whence n = 51 days.

EXAMPLE 4.6 USE OF THE MONKMAN-GRANT EQUATION (4.7)

Superheater tubes in a power station boiler are made from Type 316 stainless steel with
a 100 J.Lmgrain size and operate at 550°C at a maximum stress of 50 MPa.
Calculate the design lifetime of the superheater assuming that accumulated strains
of less than 1% are specified. Compare your predictions with the empirical predictions
for the same alloy using the Larson-Miller plot provided (Fig. 4.5).

51
Dislocation Theory for Engineers: Worked Examples

Larson-Miller parameter, T (20 + log t) = 10-3; T in oK, t in hours


5 10 15 20 25 30 35

,
C')
c
C')

,cen :9 (\J

E
c 0
E
0
Ol
~
0
~ ~
300 200
10o 150
90 200
80
70 150
60 100
90
50 80
100 70
40 90 60
80
30 70 50
1"-. .••..
r-, 25 60 40
50
r- 20
15
40 30

\ 30 220
\ 10

°C
\.
, 9
8
7
20
15
15

1\ 6 10
100 20~ \ 9

,
5 8
200 100~
1\ 4
10
9
7
6
300 ~ 3
8
7 5
400
500 -
200 ~
~~ \ 1\ 2
6
5
4

3
300 ~~~~~~ 4
600 1.5
~~~~~~ 3 2
700
400 ~~~~ 1 1.5
800 ~~~~~~~ 2
900
500 ~~~~~~ 1.5 1
1000 ~~~~~~'"
1100 600 ~~~~~~~

-, -, -, -,-.
1200 ~~~~~~~
700
~~~~~~'"
1300
~"",
1400
800 ~~~~~~~
1500
~~~~~
-, -, -, -, -. -. -.
1600 ~~~~~~'"
900
1700
1800
1000 ~~~~~~~
1900 ~~~~~~~
""'~~""~~'"
2000 1100
2100 ~~~~~~"",
10-1 100'101 "'10"1
2 03 "104

2200 1200
Time - hours
Failure times at 1% strain

52
Effects at Elevated Temperatures
Data:
In the Monkman-Grant equation:
Ao = 1010
G = 80GPa
b = 2.58 X 10-10 m
p=O
n=5
lattice friction stress = 45 MPa

Pre-exponential constant in the diffusion equation = 10-4 rrr' S-l

Activation energy for self-diffusion = 240 kJ mor '

3
. _ -4 (_ 240 X 10 ) _ X -20 2-1
.. D - 10 exp 8.31 X 823 - 5.7 10 m s .

Under power law creep, equation (4.7) becomes:

. _ A (DGb)
E-
(a - ao)5 s -1
0-- ---
kT G

.. = 10 (5.7 X 10-20 X 8 X 1010 X 2.58 X 10-10) ( 5 X 106 )5


•• E 10 1.38 X 10-23 X 823 8 X 1010

i.e.

If the maximum strain is 10/0, then the time to reach this strain, t, is given by

8 0.01
t = 8 = 9.9 X 10-10 s

i.e. t = 2811 hours.


The stress is 50 MPa (5 kg mm "), reading this from the Larson-Miller plot (Fig.
4.5) gives a lifetime of around 1000 hours, which is adequate confirmation that the
mechanism-based calculation complies well with the empirical data.

53
5
Dislocations in Fracture Processes

DUCTILE FRACTURE

Ductile fracture occurs by a linking of small voids set up around second phase
particles in the alloy, as shown in Fig. 5.1. Assuming that the void linkage process
does not contribute significantly to the failure strain (which will be the case when
the spacing of the second phase particles is small), the critical step in the ductile
fracture process is the opening of a void around the particle. Ashby'? has analysed
this problem and the mechanism turns out to be the accumulation of dislocation loops
on secondary slip planes intersecting the particles. Enough dislocation loops at the
interface eventually introduce sufficient half-planes or stress that a void is opened
up beneath the slip plane. It is useful in the design of particle-strengthened systems
to identify the ideal requirements of the particle size distribution for the maximum
resistance to ductile failure, and the Ashby analysis enables us to do this.

APPliej stress

Voids

IH
Particles

1
Fig. 5.1 Ductile fracture.

55
Dislocation Theory for Engineers: Worked Examples

/aXiS Tensile

/
/

Cavity
Primary
Shear

Fig. 5.2 Showing how tensile stress in the particle-matrix interface leads to the
opening up of a cavity.'

EXAMPLE 5.1 CALCULATION OF DUCTILE FRACTURE STRAIN IN A STEEL

Calculate the fracture strain for a steel with a volume fraction of 1% carbides with a
mean radius of 10 nm. The properties of the steel are as follows:
Carbide-matrix interfacial energy = 0.6 J m ?
Poisson's ratio = 0.28
Shear modulus = 81 GPa

The required equation is derived by assuming that self-interstitial loops


accumulating on secondary slip systems during plastic deformation will exert a stress
on the particle-matrix interface (see Fig. 5.2). The loop closest to the interface
carries the majority of the stress to the interface, and its distance from the interface
is the length of the loop stack divided by the number of loops in the stack. This stress
varies as the reciprocal of the distance from the loop to the particle-matrix interface.
The stress at the interface is thus proportional to the number of loops in the stack
divided by the length of the dislocation loop stack. Unless the particle spacing is very
large, the loop stack size can be approximated to the interparticle spacing, 'A. The
number of loops in the stack is related to the Burger's vector b, the particle diameter
derit, and the strain at the interface, e. Hence the interface stress, UT required to erit

generate the fracture strain e F is given by:

e deril
ITT
crit
= a--'Ab (5.1)

where a is a constant.

56
Dislocations in Fracture Processes
This shows that the shear fracture stress for the particle-matrix interface is
proportional to the particle size and inversely proportional to the inter-particle
spacing. Re-arranging, the fracture strain, e F, is proportional to the shear stress
multiplied by a constant involving 'Aid:

(5.2)

The strain-related term in equation (5.2) involving b, UT and o., is difficult to


erit,

estimate from first principles. So another method is used for Griffith planar cracking
with a crack length equivalent to the grain size.
Now the fracture stress in Griffith crack-like conditions for fracture is given by:

UF= ~ = GeF
~~
where db is grain size.
This shear fracture stress is related to the fracture strain at the interface with the
particles and the shear modulus as shown above. Since Young's modulus E is equal
to 2G(1 + v):

2(1+v)'Y
eF=
'ITd1 G

If we gather together the o, b and fracture stress UT terms in equation (5.2) to


erit

define the above strain effect, then equation (5.2) can be re-written for fracture
strain:

'A 2(1 + v)'Y


eF= -
d 'ITd1 G

In the present problem, we first need to calculate 'A.

The volume of one particle is given by ~1T(10 X 10-9) rrr',

Th e num b er 0 f partie. Ies per cu biIC metre IS


., b
given y
3 X 0.01
4'IT X10-24'

The elongation to failure is then given by:

74.8 2(1 + 0.28) X 0.6


eF = 20.0 'IT X 81 X 109 X 20 X 10-9'

giving OF = 6.5%.
57
Dislocation Theory for Engineers: Worked Examples

THE DUCTILE-BRITTLE TRANSITION IN nee METALS

Ashby and Embury!' have proposed a model for the influence of dislocation density
on the ductile-brittle transition in bee metals.

EXAMPLE 5.2 THE EFFECT OF DISLOCATIONDENSITY ON THE


DUCTILE-BRITTLE TRANSITION

Example: Derive an approximate expression for the critical dislocation density below
which a brittle crack will run. What is the effect of dislocation dynamics and work
hardening?

In many bee metals the ductility can be improved and the ductile to brittle transition
temperature can be lowered by prior plastic working. Figure 5.3 shows a crack in (a)
a specimen of low dislocation density and in (b) a specimen of higher dislocation
density. For crack velocities which are less than about half the Rayleigh wavespeed,
the elastic field can be estimated using a quasi-static result, so that for Mode I

Zone of
influence

(a)

Zone of
influence

(b)

Fig. 5.3 Showing how the crack tip field may be unable to capture dislocations and
cause them to multiply when the dislocation density is low (a), but may do so when
the dislocation density is higher (b).8

58
Dislocations in Fracture Processes
loading in plane strain, the shear stress at a distance r from the crack tip is
approximately

where KJ is the mode I stress intensity.


If the dislocation density is p, then their number per unit length is p1I2,so their
separation is p-1I2. If "» is shear stress opposing dislocation motion in the lattice,
then, as shown in Fig. S.3(a), the crack will run if the stress falls below "» in a distance
r which is less than the dislocation spacing, giving a low energy fracture.
If, as in Fig. 5.3(b), dislocations are always within reach of the crack tip, so that
they may move and multiply, this will give a higher energy fracture.
The transition to brittle behaviour occurs when

Thus the critical dislocation density, Pc, below which the crack runs without capturing
dislocations is

p,
= (61TT;)2
K2 (5.3)
Ie

For a typical bee transition metal at room temperature, KIe:::::::: 2 MPa m1l2 and
Tp:::::::: 400 MPa, suggesting a value of Pc:::::::: 6 X 1011m-2.

The effect of dislocation dynamics and work hardening


When the crack moves with velocity v, the dislocations must also move if they are
to contribute to crack-tip plasticity. The velocity, v, of a dislocation acted on by a
shear stress T may be written:

V=Vo
(
-T )n exp (-dF)
--
Tpo kT

where Vo is a constant with the dimensions of velocity, and dF


is the activation
energy for lattice resistance controlled glide, k is Boltzmann's constant, T is the
absolute temperature, and Tpo is the lattice resistance at 0 K. The exponent n is of the
order 40 at room temperature.
If "» is the resistance to dislocation motion at temperature T, we may rearrange
the above equation to give

T
~= ( V
-exp-
dF)lIn (5.4)
Tpo Vo kT

59
Dislocation Theory for Engineers: Worked Examples
As the dislocation density increases, work hardening ('Tn) superimposes on the lattice
resistance, given by

'Tn = exGbp1l2 (5.5)

where o r= O.l.
The critical dislocation density for brittle behaviour is given by substituting
('Tp + 'Tn) from equations (5.4) and (5.5) into equation (5.3). Rearrangement gives the
critical temperature T; for the ductile to brittle transition, as a function of the
dislocation density. For brittleness:

T< t; = Il:(In [~ (~~ (bpl12)I12(1- B(bp1l2))112 n)-I

(6'IT) 112 exGb 112


with B = :::::::::1

K1c

This shows that the critical temperature has a minimum at a particular value of the
dislocation density.

FATIGUE FRACTURE

Typical engineering fatigue design diagrams are of the S-N type where S is the
maximum applied stress and N is the number of cycles to failure at that stress S (see
Fig. 5.4). The approach allows designers to read off the maximum stress of the
application against the predicted lifetime for the material concerned. To understand

Cf)

a>
"'0

E
0. LCF
E
co
en
en
~
·05 HCF Endurance
limit

105 107
Cycles to failure, N,

Fig. 5.4 A typical fatigue S-N curve.

60
Dislocations in Fracture Processes
curves like those in Fig. 5.4 from a dislocation mechanism view point it is important
initially to realise that there are at least two distinctly different mechanisms
operating in different parts of the diagram.
At low stress the material is undergoing high cycle fatigue (HCF) and the failure
occurs after many millions of cycles. Commonly there is a fixed stress below which
fatigue failure is never expected, called the endurance limit. This behaviour does not
seem to occur in non-ferrous materials where the curve asymptotically approaches
zero. The majority of the failure damage is felt in the initiation phase, where cracks
are initiated at surfaces due to rachetting type loading. This damage is very slight
and many millions of cycles are required to form the necessary defects on the surface
(called intrusions and extrusions) to initiate a crack. The initiation can take several
million cycles to accomplish, and it is followed by the crack propagation phase.
At high stress, low cycle fatigue (LCF) occurs. Crack propagation is the important
factor: the initiation stage is relatively insignificant and nearly all of the cycles are
responsible for propagating the crack, a process that lasts for a few hundred cycles.
The stress demarcating HCF from LCF is usually about the yield point.

High Cycle Fatigue


A characteristic of fatigue microstructures on a fine scale is a cellular network of
thickly populated dislocation walls, often arranged orthogonally (Fig. 5.5). These
walls lie perpendicular to the persistent slip bands along which the intrusions and
extrusions are created at the intersection with the surface (Fig. 5.6).
It has been found using transmission electron microscopy that the walls consist of
edge dislocations, whereas those dislocations bowing out between the walls have
screw character. Plastic deformation within the persistent slip band is achieved by
moving the more mobile screw components back and forth as shown in Fig. 5.7.

Fig.5.5 Dislocation structure in 316L stainless steel after cycling to failure at a strain
range of ±O.25% at 550°C.12

61
Dislocation Theory for Engineers: Worked Examples

Ten~ile
axis
1 /
/

/
/
/
/
/
/

/
/

/
/

/
/
/
/

Fig. 5.6 Schematic diagram of a persistent slip band.

(a)

(b) HH r
Fig.5.7 Fatigue deformation showing the movement of screw dislocations back and
forth between the walls of edge dislocations

62
Dislocations in Fracture Processes
Since the pinning is effected by the walls, the wall spacing is a critical parameter in
determining the critical stress required to start plastic deformation. This stress is
similar to the Orowan Stress and represents the lower limit required to form
intrusions and extrusions. Above this limiting stress the screws will cross slip in the
inter-wall regions and annihilate each other. A typical equation describing this,
endurance limit, (TSAT, therefore is

2Gb
(TSAT = sr, + (1 - v)d

Where o, is the lattice friction stress, v is Poisson's ratio, b is the Burger's vector, G
is the shear modulus and d is the wall spacing. The dislocations in the walls are
effectively dipoles separated by a distance s, and are locked because they need
stresses of the order of

Gb
81T(1 - v)s

to allow them cross slip and annihilate each other. This stress must be equal to (TSAT
and so the spacing is very much less than the wall spacing.
The plastic strain, sP' accumulated at the endurance limit can also be calculated
and is given by

21T(TSAT
S =---
p G

Low Cycle Fatigue


This mechanism sets in when the stresses are higher and the whole crack
propagation stage is usually finished in a few thousand cycles. Figure 5.4 indicates
that the value of S is number of cycles-dependent in the LCF regime. Therefore any
dislocation based mechanism used to describe the curve must contain some
dependence of N. Mughrabi':' has considered the damage created by dislocations
cycling back and forth between two pinning planes spaced s apart. The total relative
displacement between the planes is given by

V4dsbspNm

Where de is the local plastic strain maximum, b is the Burger's vector, p is the
fraction of slip that is irreversible (usually taken to be unity), and m is equal to 1 - n
where n is the work hardening exponent.
A dislocation pinned on each of these planes will therefore experience a stress
required to move it which gradually decreases according to the Orowan formula

63
Dislocation Theory for Engineers: Worked Examples
because the pinning distance will increase as indicated above. Therefore the failure
stress, (J"p, during LCF becomes

Cb
(J" - + (J"
F - V4b.sbspNm SAT

A comparative engineering approach for LCF behaviour using constitutive equa-


tions has been adopted by Peterson."

-lOIC

N = 1000 ( :: )

Where (J"y is the yield stress and C is the stress concentration factor for the notch
initiating the crack.

EXAMPLE 5.3 S-N CURVE PREDICTION

An aluminium alloy is to be used in an aircraft wing application where fatigue


properties are likely to be an important issue. It is desired to produce a design S-N
engineering curve for the alloy. Assuming dislocation theories (HCF and LCF) for
fatigue crack initiation and propagation, make a prediction of the S-N curve for the
alloy, assuming the following properties. Compare your curve with that predicted from
the constitutive Peterson equation.
Wall spacing = 0.5 JLm
Probability of irreversible slip = 1
Work hardening exponent = 0.3
Shear modulus = 26 CPa
Al is face centred cubic with a lattice parameter = 0.404 nm
Lattice friction stress = 50 MPa
Assumed stress concentration factor = 5
Yield stress = 300 MPa
Poisson's ratio = 0.33
Assume the local plastic strain maximum during crack propagation is 0.25%, and that
the transition from HCF to LCF occurs after 10,000 cycles.

Consider N, at 1000, 10,000 and 107 cycles. First the endurance limit must be
calculated. This is equivalent to the stress for a life of 107 cycles.

0.404 2 X 26 X 109 X 10-9


(J"SAT = 50 + v2 (1 - 0.33) X 0.5 X 10-6

(J"SAT = 50 + 44.3 = 94.3 MPa

64
Dislocations in Fracture Processes
At 1000 cycles:
26 X 109 X 0.404 X 10-9
UF= Vz + 94.3MPa
0.404
4 X 0.0025 X 0.5 X 10- 9
X 10- 6
X 1000 X 0.7 X v'2

:. Up = 234 + 94.3 = 328.3 MPa


At 10,000 cycles:
Up = 74 + 94.3 = 168.3 MPa
Using the Peterson formula:
328.3)-10/5
N = 1000 ( 300 = 836 cycles

168.3)-10/5
N = 1000 ( 300 = 3177 cycles

Thus the S-N curve appears as in Fig. 5.8.15


400

350

300

8:.. 250
~
en
en
Q)
~ 200
(ij ,,
.~ ,,
·c
::> 150 ,,
,
x-_ .....................

--- --- ---


----~
100

50
o Experimental data
X Theory (dislocation stress model)

o~----~----~------~----~----~
o 200000 400000 600000 800000 1000000
Number of cycles, N

Fig. 5.8 S-N curves for AI-Mg-Si alloy.IS

65
References

1. M. 1. WHELAN,Worked Examples in Dislocations, Institute of Materials, 1990.


2. R. E. SMALLMAN,Modern Physical Metallurgy, 4th Edition, Butterworth, 1985.
3. A. SEEGER,1. DIEHL, S. MEDER and K. REBSTOCK,Phil. Mag., 1957, 2, 323.
4. G. R. ODETTEand D. FREY, J Nuclear Mat., 1979, 85-86, 818.
5. 1. D. ESHELBY,F. C. FRANKand F. R. N. NABARRO,Phil. Mag., 1951, 42, 351.
6. R. L. FLEISCHER,Acta Metall., 1963, 11, 203.
7. N. F. MOTT and F. R. N. NABARRO, Report on the Strength of Solids, Physical Society,
1948, 1.
8. V. GEROLDand K. HARTMANN,Trans. Japan Inst. Metals, 1968, 9, 509.
9. H. 1. FROST and M. F. ASHBY, Deformation Mechanism Maps - The plasticity and creep of
metals and ceramics, Pergamon Press, 1982.
10. M. F. ASHBY,Phil Mag., 1966, 14, 1157.
11. M. F. ASHBYand 1. D. EMBURY,Scripta Met., 1985, 19, 557.
12. S. N. GHAFOURI,R. G. FAULKNERand T. E. CHUNG,Materials Science and Technology, 1986,2,
1223.
13. H. MUGHRABI,Proc. Conf. on Dislocations and properties of real materials, Institute of Metals,
1985, 244-262 ..
14. R. E. PETERSON,Stress concentration design factors, John Wiley, 1966.
15. T. E. CHUNGand R. G. FAULKNER,Materials Science and Technology, 1990,6,1187.

67

You might also like