Вы находитесь на странице: 1из 125

PLASTICITY

Professor Khanh Chau Le


Lehrstuhl f ur Allgemeine Mechanik
Ruhr-Universit at Bochum
Universit atsstr. 150, D-44780 Bochum
Lecture Notes
2
Contents
1 Fundamentals 7
1.1 Phenomenon of plastic deformation . . . . . . . . . . . . . . . 7
1.2 Mechanical framework . . . . . . . . . . . . . . . . . . . . . . 10
1.3 Thermodynamical framework . . . . . . . . . . . . . . . . . . 15
1.4 Constitutive law . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5 Closed system of equations . . . . . . . . . . . . . . . . . . . . 26
2 Elementary theory 29
2.1 Bending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2 Torsion of a cylinder . . . . . . . . . . . . . . . . . . . . . . . 39
2.3 Cylindrical shell under combined load . . . . . . . . . . . . . . 41
2.4 Simple metal forming processes . . . . . . . . . . . . . . . . . 45
3 Theory of plastic ow 51
3.1 Governing equations . . . . . . . . . . . . . . . . . . . . . . . 51
3.2 Torsion of prismatic bars . . . . . . . . . . . . . . . . . . . . . 53
3.3 Plane strain problems . . . . . . . . . . . . . . . . . . . . . . . 56
3.4 Plane stress problems . . . . . . . . . . . . . . . . . . . . . . . 70
4 Crystal plasticity 73
4.1 Physical background . . . . . . . . . . . . . . . . . . . . . . . 73
4.2 Continuum dislocation theory . . . . . . . . . . . . . . . . . . 79
4.3 Anti-plane constrained shear . . . . . . . . . . . . . . . . . . . 84
4.4 Plane constrained shear . . . . . . . . . . . . . . . . . . . . . 97
4.5 Single crystals deforming in double slip . . . . . . . . . . . . . 110
3
4 CONTENTS
Preliminary remark
In this course we shall restrict ourselves to the deformable solids. In solid
mechanics we distinguish
1. material independent universal relations such as
kinematic relations,
mechanical balance equations, as well as
thermodynamical balance equations
2. from the constitutive relation, which expresses the stress tensor of
a material point in terms of the local strain tensor and the local
temperature 1:
(. 1).
Let us rst classify the form of this constitutive relation.
As you know from the theory of elasticity, elastic materials are character-
ized by a single-valued scalar function, called free energy density (per unit
volume) and denoted by o(. 1), such that
=

o(. 1).
This is the so-called state equation for thermoelastic solids. The measures
of strain and stress tensors can still be chosen dierently for small and nite
deformations.
For inelastic materials this one-to-one relation is no longer valid. The
stress-strain relation depends now on the history of loading and deforma-
tion. We can roughly classify the inelastic material behavior according to
the following features
rate-independent phenomena. The material behavior does not depend
on the loading rate. Example: plastic deformation.
5
6 CONTENTS
rate-dependent phenomena. The material behavior depends on the
loading rate. Examples: visco-plastic deformation, creep, relaxation.
In this course we shall study isothermal deformation processes of elasto-
plastic bodies, where we sometimes even neglect the contribution of the elas-
tic strains as small compared with its plastic counterpart. After a short
discussion about the phenomenon of plastic deformation on the example of
the uniaxial tension or compression test we shall propose thermodynami-
cally consistent constitutive equations for elastoplastic materials under the
condition of small strains. Within the framework of
i) the so-called elementary theory, as well as
ii) the theory of plastic ow
we shall solve some simple problems to show how the elastoplastic defor-
mation of solids can be determined. Finally, we give a short introduction
to the modern crystal plasticity incorporating the continuously distributed
dislocations.
Chapter 1
Fundamentals
1.1 Phenomenon of plastic deformation
Simple tension or compression test
Let us consider rst the uniaxial tension test with the subsequent unloading
for two materials: i) pure cooper, and ii) soft-annealed carbon steel. The
corresponding stress-strain curves are shown below in Fig. 1.1,
e e
s
s
s
Y
s
Y
O O
A
A
B
B
i) ii)
C
C
e
p
e e
p e
+
e
p
e
e
Figure 1.1: Stress-strain curve
where the strain and stress are dened as follows
- =
|
|
0
. o =
1

0
. - < 10
3
,:.
One remark should be made concerning the denition of stress. Since the
deformed cross-section at tension shrinks, the true stress should actually be
dened as 1,, where is the current cross-section area. However, at small
strains of the order - < 1% the error is not so grave.
7
8 CHAPTER 1. FUNDAMENTALS
Looking at the stress-strain curve one can recognize two dierent types
of material response in the elastic and elasto-plastic regions. In the purely
elastic region (within the line OA) no residual strain is observed: the speci-
men assume its original length after the load is removed. For most of metals
the stress is proportional to the strain so that the Hooke law is valid. The
purely elastic region ends at point A corresponding to the yield stress o
Y
.
Beyond this purely elastic region we observe for cooper i) a mild transition
to the elasto-plastic region, while for steel ii) a sharp yield stress marked
by a nearly horizontal segment. If the specimen is loaded beyond this yield
stress, it begins to deform plastically. The specimen shows a residual strain
after unloading. The total strain is additively decomposed into the elastic
and plastic parts
- = -
c
+ -
j
=
o
1
+ -
j
.
After the unloading the plastic strain remains. In Fig. 1.1 this loading and
unloading processes are shown by the stress-strain curve (marked with ar-
rows) from O through A to B and nally from B back to C.
Determination of the yield stress
It is not always easy to determine in praxis the yield stress from the stress-
strain curve. Normally, one measures the elastic modulus, at which a xed
amount of residual strain occurs (say, -
o
= 0.2%). With this modulus of
elasticity 1
0
one can determine the yield stress o
Y
(see Fig. 1.2).
e e
off
s
s
Y
Figure 1.2: Fixing the yield stress
Hardening
In the elasto-plastic region, when the specimen is reloaded again after the
unloading, one observes approximately the same stress-strain curve (only
in the opposite direction), apart from a small hysteresis loop and a rather
milder transition to the elasto-plastic region. This means that the material
behaves elastically up to the point B (see Fig. 1.1), and the plastic deforma-
tion begins to increase again when the stress achieves its value o

at point B
1.1. PHENOMENON OF PLASTIC DEFORMATION 9
corresponding to the stress at the end of the previous loading process. This
stress o

can be regarded as the new yield stress o


Y
. Since o
Y
is higher
than the initial yield stress o
Y 0
, one speaks of the hardening behavior. The
following power law, which is phenomenological, can be used to describe the
hardening behavior
-
j
=
(
o
Y
o
Y 0
H
)
a
. : 1.
and inversely
o
Y
= H
u

-
j
+ o
Y 0
.
e
s
s
Y0
1
1
E
H/(1+H/E)
Figure 1.3: Linear hardening
For : = 1 we have a linear hardening (see Fig. 1.3)
-
j
=
1
H
(o
Y
o
Y 0
).
e
s
s
Y
Figure 1.4: Elastic-ideal-plastic material behavior
If we replace the increasing hardening curve by a horizontal line (see
Fig. 1.4), then the material behavior of this idealized material is called elastic-
ideal-plastic.
10 CHAPTER 1. FUNDAMENTALS
The same can be said about the hardening behavior for the compression
test. One needs just to inverse the signs of o. o
Y 0
as well as -. -
Y 0
. The
corresponding inequalities must also be modied appropriately.
Bauschinger eect
We consider now a loading process, for which the specimen is rst loaded
in tension to attain a certain amount of plastic strain, then is unloaded and
immediately loaded in compression. The corresponding stress-strain curve is
shown in Fig. 1.5.
e
s
s
Y0
s
Y+
s
Y-
Figure 1.5: Process with loading in compression
We observe that
o
Y +
o
Y
.
This phenomenon is called Bauschinger eect.
1.2 Mechanical framework
To keep the presentation as simple as possible let us use cartesian coordinates
to describe deformations of solids.
Kinematics
At the beginning of the process at time t = 0 the body occupies the region
of the three-dimensional Euclidean point space. The position vector of an
arbitrary material point is denoted by x, and its components by r
i
. i = 1. 2. 3.
The displacement vector of this material point is denoted by u(x. t), with
n
i
(x. t) being its components. The deformation gradient is given by
F = grad(x +u) = I + gradu
or, in components
1
i)
= o
i)
+ n
i,)
.
1.2. MECHANICAL FRAMEWORK 11
B
x
u
Figure 1.6: Motion of a solid in the Euclidean space
where the comma before an index denotes the partial derivative with respect
to the corresponding co-ordinate. The Green strain tensor is dened as E =
1
2
(F
T
F I), or in components
1
i)
=
1
2
[(o
iI
+ n
i,I
)(o
)I
+ n
),I
) o
i)
].
Unless otherwise specied we always use the Einstein summation convention:
summation on repeated indices from 1 to 3 is understood. For small displace-
ment gradients the quadratic term n
i,I
n
),I
in this formula can be neglected
giving
-
i)
=
1
2
(n
i,)
+ n
),i
).
Like the Green strain tensor, this small strain tensor -
i)
is also symmetric.
The trace of the strain tensor -
i)
-
11
+-
22
+-
33
= -
ii
= n
i,i
means nothing else but the volume change \,\
0
. To show this we use the
well-known Euler formula
\,\
0
= det F =

1 + n
1,1
n
1,2
n
1,3
n
2,1
1 +n
2,2
n
2,3
n
3,1
n
3,2
1 + n
3,3

.
Since the displacement gradients n
i,)
are small compared with 1, we can
neglect the quadratic and cubic terms of n
i,)
in this determinant to obtain
\,\
0
= 1 + n
i,i
\,\
0
= n
i,i
.
We recall the following preliminary result from linear algebra: any sym-
metric tensor of second rank A can be brought into the diagonal form by
12 CHAPTER 1. FUNDAMENTALS
some orthogonal transformation of coordinate system. For this purpose one
needs to nd all eigenvectors n and the corresponding eigenvalues ` of A
from the equation
(
i)
`o
i)
):
)
= 0
This homogeneous equation for the eigenvectors has nontrivial solutions if
and only if its determinant vanishes
det(
i)
`o
i)
) = 0.
This is a cubic equation for the eigenvalues which looks in the expanded form
as follows
`
3
1

`
2
+11

` 111

= 0.
The three coecients of this cubic equation 1

, 11

, 111

are called principal


invariants of the tensor A. The computations give
1

=
11
+
22
+
33
=
ii
.
11

11

12

21

22

11

13

31

33

22

23

32

33

.
111

= det A.
Denoting the eigenvalues of A by `
1
. `
2
. `
3
, we can simply express these
principal invariants as
1

= `
1
+ `
2
+`
3
.
11

= `
1
`
2
+ `
1
`
3
+ `
2
`
3
.
111

= `
1
`
2
`
3
.
According to the above result, we may diagonalize the strain tensor -
i)
too. Its eigenvalues, called principal strains, will be denoted by -
1
. -
2
. -
3
.
Balance equations
Let j be the mass density, o
i)
the Cauchy stress tensor, /
i
the body force.
We formulate the balance of momentum and moment of momentum in the
form

j n
i
dr =

j/
i
dr +

o
i)
:
i
dc. (1.1)

c
i)I
r
)
j n
I
dr =

c
i)I
r
)
j/
I
dr +

c
i)I
r
)
o
I|
:
|
dc
1.2. MECHANICAL FRAMEWORK 13
for an arbitrary regular volume of the body. Here dr = dr
1
dr
2
dr
3
denotes the volume element, dc the area element,
c
i)I
=

0 when at least two indices coincide


1 when i,/ is an even permutation
1 otherwise
is called the permutation symbol, and n
i
corresponds to the time derivative
of n
i
. The balance of momentum generalize Newtons second law of the
classical mechanics to continua; together with the balance of moment of
momentum they present the most general laws of mechanics. The surface
integrals in (1.1) can be transformed into the volume integrals in accordance
with Gauss formula. Since is arbitrary and since the integrand is assumed
to be continuous, we may derive from (1.1) the balance equations in local
form
j n
i
= j/
i
+o
i),)
. (1.2)
o
)i
= o
i)
.
Thus, the balance of moment of momentum implies the symmetry of the
stress tensor o
i)
.
In case of equilibrium the displacement vector n
i
does not depend on time
t so that the inertial term j n
i
vanishes. The equation of motion reduces then
to the equilibrium condition
o
i),)
+ j/
i
= 0. (1.3)
In plasticity we often have a very slow loading process. Therefore the defor-
mation process runs quite slowly, and the acceleration and the corresponding
inertial term turns out to be small compared with other terms. Such pro-
cesses are called quasi-static, and for them the equilibrium equation (1.3)
presents a good approximation.
Stress tensor
Since the stress tensor o
i)
is symmetric, it can also be diagonalized. The
eigenvalues of this tensor, o
1
. o
2
. o
3
, are called principal stresses. The prin-
cipal invariants of the stress tensor are
1
o
= o
1
+ o
2
+ o
3
.
11
o
= o
1
o
2
+ o
1
o
3
+ o
2
o
3
.
111
o
= o
1
o
2
o
3
.
14 CHAPTER 1. FUNDAMENTALS
The hydrostatic stress is dened as o
n
= 1
o
,3.
The stress deviator is dened as follows
:
i)
= o
i)
o
n
o
i)
.
The following invariants of the stress deviator are often used in the plasticity
theory
J
1
= :
ii
= 0.
J
2
=
1
2
:
i)
:
i)
=
1
6
[(o
1
o
2
)
2
+ (o
2
o
3
)
2
+ (o
3
o
1
)
2
]. (1.4)
J
3
= dets =
1
27
[(o
1
o
2
)
2
(o
1
o
3
+ o
2
o
3
) + (o
2
o
3
)
2
(o
2
o
1
+ o
3
o
1
) + (o
3
o
1
)
2
(o
3
o
2
+ o
1
o
2
)].
One can see that the invariants J
2
. J
3
are symmetric functions of o
i
o
)
.
Figure 1.7: Mohrs stress circles
The geometric interpretation of J
2
can be given in terms of the octahedral
shear stress. Consider the normal vector
n =
1

3
(1. 1. 1)
in the principal coordinates of the stress tensor and an area element per-
pendicular to it which lies on the side of the octahedron. The stress vector
acting on this area element is given by
t = n =
1

3
(o
1
. o
2
. o
3
).
The normal stress equals
o
oct
= n t =
1
3
(o
1
+ o
2
+ o
3
) = o
n
.
1.3. THERMODYNAMICAL FRAMEWORK 15
The shear stress acting on the side of the octahedron is obtained from the
formula
t
2
oct
=
1
3
(o
2
1
+ o
2
2
+o
2
3
)
1
9
(o
1
+o
2
+o
3
)
2
=
2
3
J
2
.
It is interesting to mention that the octahedral shear stress is the average
shear stress over all planes passing through a material point.
With the help of Mohrs stress circles we can also determine the maximum
shear stress (see Fig. 1.7)
t
max
=
1
2
max{o
1
o
2
. o
2
o
3
. o
3
o
1
}.
1.3 Thermodynamical framework
It is well known from the classical experiment by Taylor and Quinney that
about 90% of the work done to deform metals plastically will be dissipated
into heat. This heat suply leads in general to the change of temperature.
Thus, if the plastic deformations occur, the process we are dealing with
becomes thermo-mechanically coupled. The consequence is that, in plasticity,
thermodynamic balance equations should be taken into account.
Energy balance
We assume that the energy of an arbitrary sub-body is a sum of the kinetic
and internal energies
=

j(c +
1
2

i
) dr.
with
i
= n
i
being the material velocity. Here c corresponds to the internal
energy density. The balance of energy states

= +.
where is the power of the external forces, and is the rate at which heat
is supplied to the body. The power of the body and contact forces is given
in the form
=

j/
i

i
dr +

o
i)
:
)

i
dc.
The heat supply comes from two sources: the body heat supply and the heat
ow across the boundary; its rate is equal to
=

j: dr

i
:
i
dc.
16 CHAPTER 1. FUNDAMENTALS
Here :(x. t) is the body heat supply per unit mass and unit time,
i
(x. t) is
the heat ux vector across the surface dc per unit time. The heat ow is
positive if q and n are opposite; therefore the minus sign in the last equation
agrees with our common sense.
Substituting the above formulas for the power and the heat supply in the
right-hand side of the balance of energy and transforming the surface integral
into the volume integral, we get

[j( c +
i

i
/
i

i
:) (o
i)

i
)
,)
+
i,i
] dr = 0.
Since this equation holds true for an arbitrary regular sub-body , and since
the integrand is assumed to be continuous, we obtain the balance of energy
in the local form
j( c +
i

i
/
i

i
:) (o
i)

i
)
,)
+
i,i
= 0.
Taking into account the balance of momentum (1.2) we obtain nally
j c +
i,i
= o
i)
-
i)
+ j:. (1.5)
Second law of thermodynamics
In order to formulate the second law of thermodynamics we need two new
quantities. The rst one is the absolute temperature, referred to as an inten-
sive quantity and denoted by 1(x. t). The second one is the entropy, referred
to as an extensive quantity, whose density is denoted by :(x. t). The entropy
of the sub-body is given by

j: dr.
The second law of thermodynamics states that
d
dt

j: dr

j:
1
dr

i
:
i
1
dc. (1.6)
When the heat supply and the heat ow are absent (adiabatic process with
: = 0 and
i
= 0), the following inequality holds true
d
dt

j: dr 0
1.4. CONSTITUTIVE LAW 17
which means that the entropy of the closed system cannot decrease.
With the help of Gauss theorem we obtain

j : dr

[
j:
1
(
i
,1)
,i
] dr.
Since is arbitrary, this inequality leads to
j : j:,1 (
i
,1)
,i
= j:,1
i,i
,1 +
i
1
,i
,1
2
. (1.7)
We call = j :j:,1 +(
i
,1)
,i
the entropy production rate. The inequality
(1.7) says that 0.
There is an alternative form of the entropy production inequality often
used in plasticity. We introduce the free energy density
= c 1:.
Provided all other balance equations hold true, then the entropy production
inequality is equivalent to
j(:

1 +

) o
i)
-
i)
+
i
1
,i
,1 0. (1.8)
To prove (1.8) we use the denition of

= c

1: 1 : 1 : = c :

1

.
Substitute this into (1.7) and multiply by 1
j( c :

1

) j:
i,i
+
i
1
,i
,1.
Combining this equation with the energy balance equation (1.5), we arrive
at (1.8).
For isothermal processes with 1 = const the inequality (1.8) reduces to
j

o
i)
-
i)
0. (1.9)
1.4 Constitutive law
The formulation of the constitutive laws begins always with the specication
of all quantities characterizing the current state of the material element. Such
quantities are called state variables. Besides, one needs to specify all internal
variables which may inuence the dissipation and the irreversible behavior
of the material element. The constitutive laws for elastoplastic materials
include:
18 CHAPTER 1. FUNDAMENTALS
Specication of the free energy as function of all state variables. By
this the reversible behavior of the material is xed.
Evolution law for the internal variables (plastic strains + hardening
parameters)
A law for the heat ux (if the process under consideration is thermo-
mechanically coupled)
Dierent models of elasto-plastic materials can be proposed. Below we con-
sider some of them.
Elastic-ideal-plastic materials
We restrict ourselves to isothermal processes with 1 = const. For elastic-
ideal-plastic materials we include in the list of variables the following quan-
tities
-
c
i)
. -
j
i)
. (1.10)
We assume that the elastic strains -
c
i)
characterize completely the current
state of the deformed material element. This means that the stress tensor
depends only on -
c
i)
o
i)
= o
i)
(-
c
i)
).
The plastic strains -
j
i)
depend on the history of loading and therefore are
not the state variables. They present the internal variables which character-
ize irreversible behavior of the material element. The total strain tensor is
additively decomposed into the elastic and plastic strain tensors
-
i)
= -
c
i)
+ -
j
i)
. (1.11)
The free energy density assumes the form
= (-
c
i)
).
i. e., it depends only on the state variables. Let us dierentiate the free
energy with respect to time t

=

-
c
i)
-
c
i)
.
We substitute this formula into the dissipation inequality (1.9)
(j

-
c
i)
o
i)
) -
c
i)
o
i)
-
j
i)
0. (1.12)
1.4. CONSTITUTIVE LAW 19
We rst consider processes with -
j
i)
= 0, i. e. reversible processes. For these
processes the second term in (1.12) vanishes, so that
(j

-
c
i)
o
i)
) -
c
i)
0.
Since -
c
i)
can be arbitrary, and since the expression in the brackets does not
depend on -
c
i)
, it must be identically equal to zero and thus
o
i)
= j

-
c
i)
. (1.13)
If the free energy density per unit volume o = j is a quadratic form of -
c
i)
j =
1
2
C
i)I|
-
c
i)
-
c
I|
.
then (1.13) yield Hookes law
o
i)
= C
i)I|
-
c
I|
.
For isotropic elastic material we have
o
i)
= `-
c
II
o
i)
+ 2j-
c
i)
.
This equation of state can be decomposed into the volumetric and deviatoric
parts
o
II
= 31-
c
II
.
:
i)
= 2jc
c
i)
.
where c
c
i)
= -
c
i)

1
3
-
c
II
o
i)
is the strain deviator, and 1 = ` + 2j,3. In rate
form we have
o
II
= 31 -
c
II
. (1.14)
:
i)
= 2j c
c
i)
.
With (1.13) we reduce the inequality (1.12) to the following dissipation
inequality
o
i)
-
j
i)
0.
The left-hand side of this equation is called plastic dissipation. The yield
condition as well as the associate ow rule should satisfy this inequality.
One speaks then of thermodynamically consistent constitutive equations. We
20 CHAPTER 1. FUNDAMENTALS
formulate the yield condition in the stress space: the stress tensor must
always satisfy the condition
)(o
i)
) 0
As long as )(o
i)
) < 0, no plastic strain occurs. The surface given by
)(o
i)
) = 0
is called the yield surface, and function )(o
i)
) the yield function. The elastic
region is found inside the yield surface. If the stress tensor lies on the yield
surface, then the associate ow rule states that -
j
i)
is either zero or shows in
the direction of the gradient of the yield function
-
j
i)
= `
)
o
i)
(1.15)
with
` = 0 for ) < 0 or ) = 0 and

) < 0 (unloading).
` 0 for ) =

) = 0 (loading).
If the yield surface has an edge, the above ow rule can still be applied if we
replace the gradient by the sub-gradient of ). An alternative procedure has
been proposed by Koiter: Instead of the product of ` and the gradient of the
f=0
f=0
f
f
1
2
2
1

s
s
e
.
p
Figure 1.8: Yield surface with an edge
yield function we take now the linear combination of : products
-
j
i)
=
a

c=1
`
c
)
c
o
i)
(1.16)
with
`
c
= 0 for )
c
< 0 or )
c
= 0 and

)
c
< 0 (unloading)
`
c
0 for )
c
=

)
c
= 0 (loading).
1.4. CONSTITUTIVE LAW 21
The validity of (1.15) or (1.16) follows from the so-called principle of
maximum of plastic dissipation, which claims that
(o
i)
o

i)
) -
j
i)
0 (1.17)
for an arbitrary stress state o

i)
within the yield surface.
s s
s
s
e
e
f=0
f=0
p
p
*
*
.
.
Figure 1.9: Convexity of the yield surface and the normality rule
This principle is equivalent to the requirement of convexity of the yield
surface, because in the case of non-convexity one can always nd the stress
state o

i)
which violate the inequality (1.17). Thus, the principle of maximum
of plastic dissipation implies the convexity of the yield surface as well as the
associate ow rule. One can show that the plastic dissipation 1 = o
i)
-
j
i)
is
a function of the plastic strain rate -
j
i)
only. When the stress tensor o
i)
is
found inside the yield surface, then -
j
i)
= 0 and the dissipation vanishes. If
the stress tensor during the loading is found on the yield surface, then the
dissipation must be a homogeneous function of rst order with respect to
-
j
i)
. To be consistent with the second law of thermodynamics we require that
1 0.
Examples of the yield surface
For isotropic materials the yield function must be a symmetric function of
three principal stresses
) = )(o
1
. o
2
. o
3
).
Since the principal stresses can be expressed in terms of the principal invari-
ants 1
o
. 11
o
. 111
o
, the yield function can also be presented in the following
form
) = )
1
(1
o
. 11
o
. 111
o
).
Various observations and experiments show that the hydrostatic stress does
not inuence the plastic yielding. This means that the yield function depends
only on the principal invariants 11
o
. 111
o
of the stress tensor, or alternatively,
on the invariants J
2
. J
3
of the stress deviator
) = )
2
(J
2
. J
3
).
22 CHAPTER 1. FUNDAMENTALS
Consequently, the yield function must be a symmetric function of o
i
o
)
) = )
3
(o
1
o
2
. o
2
o
3
. o
3
o
1
)
and any parallel translation in the direction (1. 1. 1),

3 in the 3-D space of


principal stresses does not change the yield surface.
The criterion of maximum shear stress (Trescas yield condition) states
that the plastic ow occurs when the maximum shear stress achieves some
critical value. According to this criterion the yield function must have the
form
) =
1
2
max{o
1
o
2
. o
2
o
3
. o
3
o
1
} t
Y
.
or, the equivalent form
) =
1
4
(o
1
o
2
+o
2
o
3
+o
3
o
1
) t
Y
.
In this equation t
Y
denotes the yield shear stress. For an uniaxial tension
Figure 1.10: Mohrs stress circle for uniaxial tension
test the plastic ow occurs when (see Fig. 1.10)
o
1
= o
Y
= 2t
Y
.
Thus, t
Y
= o
Y
,2. The projection of the Tresca yield surface onto the octa-
hedron plane (the so-called -plane) in 3-D space of principal stresses is a
hexagon (Fig. 1.11).
The normality rule then implies that
-
j
1
=
1
4
`[sign(o
1
o
2
) + sign(o
1
o
3
)].
where
signr =
d
dr
r =

+1 r 0.
. (1. 1) r = 0.
1 r < 0.
1.4. CONSTITUTIVE LAW 23
Figure 1.11: Projection of Trescas and Mises yield surfaces
Similar equations hold true for -
j
2
and -
j
3
. When all principal stresses are
dierent and are ordered so that o
1
o
2
o
3
, then -
j
1
= `,2, -
j
2
= 0,
-
j
3
= `,2. When o
1
= o
2
o
3
, then -
j
1
= `(1 + ),4, -
j
2
= `(1 ),4,
-
j
3
= `,2, and so on. For each combination of the principal stresses we
always have -
j
1
+ -
j
2
+ -
j
3
= `. Therefore the dissipation is equal to
1 = o
i
-
j
i
= `t
Y
= t
Y
( -
j
1
+ -
j
2
+ -
j
3
).
Von Mises proposed another yield function, which, in terms of the stress
invariants, takes the form
) =

J
2
/
With formula (1.4) this yield function can be written in the form
) =

1
6
[(o
1
o
2
)
2
+ (o
2
o
3
)
2
+ (o
3
o
1
)
2
] /.
Mises criterion states that the plastic yielding occurs when the octahedral
shear stress achieves a critical value. An alternative form of the Mises yield
function reads
) = J
2
/
2
=
1
6
[(o
1
o
2
)
2
+ (o
2
o
3
)
2
+ (o
3
o
1
)
2
] /
2
.
For the uniaxial tension test the plastic yielding occurs when
) =
1
3
o
2
Y
/
2
= 0.
Thus, / = o
Y
,

3. The projection of Mises yield surface onto the octahedron


plane in the 3-D space of principal stresses is a circle of radius

2/ (Fig. 1.11).
Using the normality rule we nd that
-
j
i)
= `:
i)
. (1.18)
24 CHAPTER 1. FUNDAMENTALS
Therefore the plastic dissipation for Mises yield condition is given by
1 = o
i)
-
j
i)
= :
i)
`:
i)
= /

2 -
j
i)
-
j
i)
.
Combining equation (1.18) with Hookes law in rate form for a linear elastic
isotropic material (see equation (1.14)) we obtain the rate of the total strain
in the form
-
II
=
1
31
o
II
.
c
i)
=
1
2j
:
i)
+`:
i)
.
This equation has been obtained by Prandtl and Reuss.
Models with hardening
We again restrict ourselves to isothermal processes with 1 = const. In order
to describe the hardening behavior one needs to include into the list of vari-
ables in (1.10) additional internal variables. Dierent types of hardening can
be described by introducing scalar variable i or tensor variable , where
is a traceless tensor of second rank. The hardening parameter i is dened in
a standard way (Odqvist)
i =

2
3
-
j
i)
-
j
i)
dt. (1.19)
If we interpret as the coordinates of the middle point of the yield surface
and

2/ as its radius, then various types of hardening can be displayed in


the -plane as shown in Fig. 1.12. Choosing the yield function in the form
Figure 1.12: Hardening: i) purely isotropic, ii) purely kinematic iii) combined
) = )
1
( ) /(i).
1.4. CONSTITUTIVE LAW 25
we can describe both isotropic and kinematic hardening.
For the plastic strain rate the normality rule in the form (1.15) remains
valid. Its validity for the hardening material behavior follows directly from
Druckers postulate, which requires that
1. additional stresses produce non-negative work during a loading process
(material stability)
do
i)
d-
i)
0.
2. for a complete cycle of additional loading and unloading, additional
stresses do positive work if plastic strains occur, and zero when the
strains are purely elastic

(o
i)
o

i)
)d-
i)
0.
Here o

i)
is an arbitrary starting point in the stress space. The elastic part
in the last inequality can be removed because it does not contribute to the
work done. Consider a cycle ABCA in the stress space. Since the plastic
strains occur only along the path BC (see Fig. 1.13), this inequality can also
be written in the form (1.17).
A
B
C
s
s
*
Figure 1.13: Normality rule and material stability
Druckers postulate is an additional requirement, that cannot be derived
from the second law of thermodynamics.
From the denition (1.19) follows that the hardening parameter i satises
the following equation
i =

2
3
-
j
i)
-
j
i)
.
For the evolution of the internal variables we propose a very simple equa-
tion in the form
c
i)
= c -
j
i)
.
Then the unknown factor ` in the ow rule for materials with hardening can
be determined. We consider for example the yield function in the form
) =
1
2
(:
i)
c
i)
)(:
i)
c
i)
) /
2
(i). (1.20)
26 CHAPTER 1. FUNDAMENTALS
which has been proposed by Melan, Prager, Ziegler and Shield. In addition
to it the consistence condition

) = 0 must be fullled for the loading process.
Thus
(:
i)
c
i)
)( :
i)
c
i)
) 2//

i = 0.
With the above evolution equations for the internal variables
i =

2
3
-
j
i)
-
j
i)
= `

2
3
(:
i)
c
i)
)(:
i)
c
i)
) = 2`/,

3.
(:
i)
c
i)
) c
i)
= c`(:
i)
c
i)
)(:
i)
c
i)
) = 2c`/
2
.
we can transform the consistence condition to
(:
i)
c
i)
) :
i)
`(2c/
2
+ 4/

/
2
,

3) = 0.
Therefore
` =
(:
i)
c
i)
) :
i)
2/
2
(c + 2/

3)
.
The ow rule becomes nally
-
j
i)
=
(:
I|
c
I|
) :
I|
2/
2
(c + 2/

3)
(:
i)
c
i)
). (1.21)
1.5 Closed system of equations
Restricting ourselves to the isothermal processes only, we have altogether the
following system of equations
3 balance equations of momentum (1.3)
6 stress-strain relations
1 yield condition
: evolution equation for the internal variables.
These 10 + : equations contain the following unknown functions
3 components of displacements (or 3 components of velocity)
6 components of the stress tensor o
i)
1 scalar factor
: internal variables.
1.5. CLOSED SYSTEM OF EQUATIONS 27
Thus, the system of equations is closed. To solve this system we may develop,
depending on the particular problems, dierent methods and approaches:
i) elementary theory of elasto-plastic deformation. This approach is char-
acterized by hypotheses which strongly simplify the boundary-value
problems. However, the ow rule is limited to simple loading situation
like uniaxial strain or pure shear.
ii) theory of plastic ow. This approach is based on the simplied mate-
rial models (elastic-ideal-plastic or rigid-ideal-plastic materials). Apart
from that no further simplications are made, and the boundary-value
problems will be solved exactly. Due to the mathematical complexity,
analytical solutions may be obtained only in exceptional cases.
iii) general theory of elasto-plastic deformation. This approach is free from
any simplifying assumption. Due to the mathematical complexity, only
numerical solutions of boundary-value problems based on the nite
element method are available.
Note that, in some special cases solutions based only on the equilibrium
equations and on the yield condition can be found without referring to the
ow rule. We call such problems statically determinate.
28 CHAPTER 1. FUNDAMENTALS
Chapter 2
Elementary theory
The elementary theory uses various simplifying assumptions concerning the
kinematics and the stress state. Justication of these assumptions cannot be
given in general, but for particular problems.
2.1 Bending
Pure bending of a beam
As the rst example let us consider the pure bending of a beam having a
constant rectangular cross-section
`
:
= `
a
= 0. `

= `.
Q

= Q
:
= ` = 0.
The chosen coordinate system and the sizes of the beam can be seen in
Fig. 2.1.
Figure 2.1: Straight beam with constant rectangular cross-section
According to the elementary theory we assume that the cross-sections
during bending remain plane and perpendicular to the beam axis, and
o

= o
::
= o
:
= o
a:
= 0.
29
30 CHAPTER 2. ELEMENTARY THEORY
The rst assumption is related to the kinematics of bending, the second to
the stress state. Both coincide with the commonly accepted assumptions of
the beam theory. It follows then from the rst assumption
-
a
= -
a:
= 0.
-
aa
= -(.) = -
0
+
.
1
.
where 1 is the radius of curvature of the beam axis. Consequently, we have
along the bers parallel to the beam axis an uniaxial stress state
o
aa
= o(.).
The remaining unknown quantities -
0
. 1 can be found from the equations
` =

o(.) dc = 0.
` =

o(.). dc
in which o(.) and -(.) are related to each other by a constitutive law.
e
s
s
Y
Figure 2.2: Stress-strain curve
For elastic-ideal-plastic materials (at loading) we have
o =
{
1- - <
o
Y
1
.
o
Y
-
o
Y
1
.
The stress distribution over the thickness is shown in Fig. 2.3.
Let us consider rst the purely elastic case
` =

o(.) dc = 1

(-
0
+
.
1
) dc = 0 -
0
= 0.
` = 1

-(.). dc =
1
1

.
2
dc =
1
1
J
:

1
1
=
`
1J
:
.
2.1. BENDING 31
-s
Y
s
Y
z z
z
o
z
u
i)
ii)
el.zone
Figure 2.3: Stress distribution: i) elastic, ii) elastic-plastic
Thus, we obtain the well-known formula
- =
`
1J
:
. o =
`
J
:
.. o
max
=
6`
/
2
.
The elastic stress distribution is valid until
o
max
= o
Y
`
c
= o
Y
/
2
6
.
1
1
c
=
2

o
Y
1
.
The plastic deformation occurs when ` `
c
. At the boundaries be-
tween the elastic and the plastic zone .
c
. .
&
the yield conditions hold true
1(-
0
+
1
1
.
c,&
) = o
Y
.
Together with two integral equations for the force and the bending moment
there are four equations to determine four unknowns -
0
. 1. .
c
. .
&
. With the
above stress distribution the force equation is simplied to
` =

:
o
:
u
1(-
0
+
.
1
)/ d. o
Y

:
u
2
/ d. +o
Y

2
:
o
/ d. = 0.
1/-
0

:
o
:
u
d. +
1/
1

:
o
:
u
. d. o
Y

:
u
2
/ d. + o
Y

2
:
o
/ d. = 0.
1-
0
(.
c
.
&
) +
1
21
(.
2
c
.
2
&
) o
Y
(.
&
+

2
) +o
Y
(

2
.
c
) = 0.
1-
0
(.
c
.
&
) +
1
21
(.
2
c
.
2
&
) o
Y
(.
c
+ .
&
) = 0.
The yield conditions at the boundaries .
c
. .
&
imply
.
c,&
= 1
o
Y
1
1-
0
.
c
+ .
&
= 21-
0
.
.
c
.
&
= 21
o
Y
1
. .
2
c
.
2
&
= 41
2
-
0
o
Y
1
.
32 CHAPTER 2. ELEMENTARY THEORY
and therefore
0 = 1-
0
21
o
Y
1

1
21
41
2
-
0
o
Y
1
+ o
Y
21-
0
-
0
= 0.
.
c,&
= 1
o
Y
1
.
c,&
=

2
1
1
c
.
We compute now the bending moment
` =
1
1

:
o
:
u
/.
2
d. o
Y

:
u
2
/. d. + o
Y

2
:
o
/. d..
` =
1
1
/
3
(.
3
c
.
3
&
) +
o
Y
/
2
(

2
4
.
2
c
.
2
&
+

2
4
).
` =
2
3
/1
2
o
3
Y
1
2
+
1
4
o
Y
/
2
/1
2
o
3
Y
1
2
.
` = o
Y
/(

2
4

1
2
3
o
2
Y
1
2
) = o
Y
/(

2
4

1
2
3

2
4
1
1
2
c
).
1 2 3 4
0.25
0.5
0.75
1
1.25
1.5
R /R
e
M/M
e
M/M
e
2z /h
o
2z /h
o
Figure 2.4: Plots of `,`
c
and 2.
c
,
Thus,
`
`
c
=
3
2
[1
1
3
(
1
1
c
)
2
].
The ultimate moment is achieved when .
c
= .
&
= 0 or, equivalently, when
1 = 0, and is equal to `
&
,`
c
= 3,2. The plots of `,`
c
and 2.
c
, versus
1,1
c
are shown in Fig. 2.4.
Mention that the above formula for the moment is valid only for small
strains, while the ultimate moment is achieved rst when 1 = 0, for which
the strains are innitely large. It must be emphasized, however, that even
2.1. BENDING 33
for relatively small plastic strains the value of the bending moment is close
to its limit. For example, 98% of this ultimate value is already achieved at
1
c
1
= 4.
If we are only interested in the limit state of the plastic bending, we may let
the elastic zone disappear completely so that
`
&
=

2
0
o
Y
/. d.

0
2
o
Y
/. d. = o
Y
/
2
4
.
The angle of bending c takes the value c = |
0
,1.
Spring-back, unloading
We consider a loading program for which the bending moment is rst in-
creased up to the value `

, where
`
c
< `

< `
&
.
Now, if we unload the beam by decreasing the moment to zero, its curvature
will decrease from
1
1

to
1
1
j
.
The dierence 1,1

1,1
j
is the elastic spring-back (elastic recovery) of the
beam. 1
j
is the residual radius of curvature.
The decrease of the moment from `

to zero is equivalent to the super-


position of the solution found above with the elastic solution corresponding
to the moment `

, provided, the beam behaves elastically during the un-


loading, what we may assume. Thus, for small curvatures we can express the
elastic spring-back as
1
1

1
1
j
=
`

1J
:
and, accordingly c

c
j
=
`

1J
:
|
0
The stress distribution at the end of the loading (corresponding to the mo-
ment `

) is
o

(.) =

1
1

. . 1

o
Y
1
.
o
Y
. 1

o
Y
1
.
o
Y
. < 1

o
Y
1
.
We have to superimpose this stress distribution with
o(.) =
`

J
..

2
.

2
.
34 CHAPTER 2. ELEMENTARY THEORY
-s
Y
s
Y
z
M
*
Figure 2.5: Stress distribution after the loading
Figure 2.6: Stress distribution due to `

We have thus after the unloading the eigenstress in the cross section of the
beam
o(.) =

(
1
1

J
:
). . 1

o
Y
1
.
o
Y

A

J
:
. . 1

o
Y
1
.
o
Y

A

J
:
. . < 1

o
Y
1
.
where 1

is determined in accordance with


z
Figure 2.7: Eigenstress after the unloading
`

`
c
=
3
2
[1
1
3
(
1

1
c
)
2
].
This eigenstress will aect the plastic yielding if we reload the beam in the
opposite direction: the absolute value of the moment at which the plastic
strain changes is lower than that of the rst loading. This is the Bauschinger
eect due to the eigenstrain. Consider for example the case when the beam
is bent up to
1
1
c
= 4
`

`
c
=
47
32
. .
c,&
=

8
.
2.1. BENDING 35
The elastic spring-back at the unloading is equal to
1
c
(
1
1

1
1
j
) =
`

`
c
`
c
1J
:
=
`

`
c
=
47
32
.
The stress distributions after loading and unloading are shown in Fig. 2.8. At
z
-s
Y
s
Y
z
M
*
-81 128 s /
Y
15 32 s /
Y
i)
ii)
Figure 2.8: Stress distributions after loading and unloading
the subsequent reloading in the opposite direction the beam deform elastically
until

`
`
c
=
17
32
.
so
`
`
c
=
`

`
c

`
`
c
= 2.0.
At the subsequent reloading in the same direction the beam behaves elasti-
cally for increasing moment until
`
`
c
=
`

`
c
=
47
32
.
The shakedown of the beam occurs for `,`
c
2.0.
z
-s
Y
s
Y
z
M
*
- 2 s /
Y
i)
ii)
s
Y
Figure 2.9: Reloading: i) in the same direction, ii) in the opposite direction
36 CHAPTER 2. ELEMENTARY THEORY
Figure 2.10: Hardening and Bauschinger eect at reloading in the same and
in the opposite direction
The previous solution can easily be generalized for materials with linear
or non-linear hardening. One needs just to replace the constant yield stress
in the plastic zone by a function o
Y
(-).
For linear hardening we have
o =

1- -
o
Y
1
.
1
1+1
(o
Y
+ H-) -
o
Y
1
.
1
1+1
(o
Y
+H-) - <
o
Y
1
.
If the hardening is symmetric with respect to tension and compression, then
-
0
= 0. - =
.
1
.
c,&
= 1
o
Y
1
.
As before, the ultimate moment is achieved when - and 1 0.
These deliberations can further be applied to study:
i) a general case of unsymmetric bending of the beam having rectangular
cross section. One has to assume that
-(. .) = -
0
+
.
1
:
+

1

.
o

= o
::
= o
:
= 0.
ii) the combination of tension and bending of the beam having rectangular
cross section.
iii) similar problems for beams with doubly-symmetric cross sections (in-
cluding thin-walled cross section).
2.1. BENDING 37
iv) the bending of a plate.
Plate bending
For the bending of a plate we assume that
o
::
= o
:
= o
:a
= o
a
= 0,
but instead of o

= 0 now -

= 0 (plane strain state).


In the elastic zone we have
-

= 0 =
1
1
(o

io
aa
) o

= io
aa
= io.
-
aa
= - =
1
1
(o
aa
io

) =
1 i
2
1
o.
Figure 2.11: Plate bending
In the plastic zone we assume the ideal-plastic material behavior with
Mises yield condition, so
-

= 0 =
1
1
( o

i o
aa
) +
2
3
`(o

1
2
o
aa
).
-
aa
= - =
1
1
( o
aa
i o

) +
2
3
`(o
aa

1
2
o

).
) = o
2
aa
+ o
2

o
aa
o

o
2
Y
= 0.
We eliminate ` from the rst two equations
- =
1
1
( o
aa
i o

)
1
1
( o

i o
aa
)
o
aa
o

,2
o

o
aa
,2
.
38 CHAPTER 2. ELEMENTARY THEORY
From the yield condition we derive
o

1
2
o
aa
=

o
2
Y

3
4
o
2
aa

- =
o
1
[1 i(
1
2

3
2
o
2

...
) + (
1
2

3
2
o
2

... i
)(
1
2

3
2
o
2

...
)]
=
o
1
[1 + (
1
2

3
2
o
2

...
)
2
2i(
1
2

3
2
o
2

...
)]
=
o
1
1
4o
2
Y
3o
2
[o
2
Y
(5 4i) 3o(1 2i)(

... +
1
2
o)].
Therefore the following dierential equation holds true
- = op(o).
In the special case i = 1,2 we have
- =
o
1
3o
2
Y
4o
2
Y
3o
2
=
o
1
3
4
1
3
4
o
2
.
where o = o,o
Y
. Introducing a new variable n, with
o =
2

3
cos n.
we obtain
- =

3
2
o
Y
1
n
sin n
.
The integration gives
- =

3
2
o
Y
1
ln

3
2
o
1 +

3
2
o
+ c.
The stress at the boundary of the elastic zone equals
o
c
=
o
Y

1 i +i
2
and for i = 1,2 o
c
=
2

3
o
Y
.
The Ansatz for the solution remains as before
- = -
0
+
.
1
.
2.2. TORSION OF A CYLINDER 39
together with the integral equations for the force and bending moment
: =

o d. = 0. : =

o. d..
Beam with symmetric cross section about one axis
If the cross section of the beam is symmetric only with respect to the .-
axis, then, due to the redistribution of stress during the plastic yielding, the
position of the neutral ber (with - = 0. o = 0) will change. From the
condition
` =

o dc = 0
follows in the case of ultimate moment
` =

o dc = o
Y

&
o
Y

c
= 0
c
=
&
= ,2.
The ultimate neutral ber will therefore be the line dividing the cross section
into equal areas. Take for example the triangle, we have
=
1
2

2

c
=
&
=
1
4

2
. =

2
2
.
Figure 2.12: Neutral and ultimately neutral ber
2.2 Torsion of a cylinder
We consider a pure torsion of a cylinder shown in Fig. 2.13, with
`
|
= const.
40 CHAPTER 2. ELEMENTARY THEORY
Figure 2.13: Torsion of a cylinder
We assume that cross sections remain plane and perpendicular to the axis
of cylinder, and that the straight lines in radial directions remain straight.
Besides
o
aa
= o

= o
::
= o
:
= 0.
From the rst assumption follows
u
,
= r:0.
with 0 being the twist. Therefore the strain is equal to
-
a,
(:) =
1
2
:0.
For the elastic torsion we have
o
a,
= t(:) = 2G-
a,
= G0:.
`
|
=

1
0
2:t(:): d: = 2G0

1
0
:
3
d:.
`
|
=

2
G01
4
bzw. `
|
= GJ
0
0. J
0
=

2
1
4
.
So, the twist 0 is given by
0 =
`
|
GJ
0
.
The maximum shear stress is achieved at : = 1
t
max
=
`
|
J
0
1 =
`
|
\
|
.
The threshold value for the plastic strain to occur reads
`
|c
= t
Y

2
1
3
.
where
t
Y
=
{
1

3
o
Y
Mises.
1
2
o
Y
Tresca.
2.3. CYLINDRICAL SHELL UNDER COMBINED LOAD 41
When `
|
`
|c
the plastic strain occurs, and for ideal-plastic material
behavior the yield condition at the boundary :
j
between the elastic and
plastic zone is
t = t
Y
= G0:
j
:
j
=
t
Y
G0
.
The torsion moment `
|
is computed as follows
`
|
=


g
0
2:G0:
2
d: +

g
2:t
Y
: d:
=

2
G0:
4
j
+
2
3
t
Y
(1
3
:
3
j
)
=

2
G0(
t
Y
G0
)
4
+
2
3
t
Y
[1
3
(
t
Y
G0
)
3
]
`
|
=

6
t
Y
[41
3
(
t
Y
G0
)
3
].
With 0
c
= `
|c
,GJ
0
= t
Y
,G1 we get for `
|
`
|
=

6
t
Y
[41
3
(
10
c
0
)
3
]
`
|
=
2
3
t
Y
1
3
[1
1
4
(
0
c
0
)
3
].
`
|
`
|c
=
4
3
[1
1
4
(
0
c
0
)
3
].
The ultimate moment is achieved at :
j
= 0, i. e. 0 and
`
|&
`
|c
=
4
3
.
This result can be obtained directly if we let the elastic zone disappear
`
|&
=

1
0
2t
Y
:
2
d: =
2
3
t
Y
1
3
=
4
3
`
|c
.
For cylindrical pipe with the internal radius d
i
and external radius d
o
the
ratios `
|&
,`
|c
decreases with the decreasing ratio d
i
,d
o
, and in the thin-wall
limit d
o
d
i
it tends to 1.
2.3 Cylindrical shell under combined load
We consider a thin-walled tank loaded by a longitudinal force `, and an
internal pressure j (see Fig. 2.14). We assume that the cross-section remains
42 CHAPTER 2. ELEMENTARY THEORY
Figure 2.14: A closed tank under combined loading
plane and perpendicular to the .-axis. Besides,
o
::
= o
:
=
jc
2t
+
`
2tc
.
o
,,
= o
,
=
jc
t
.
o
,:
= o

= o
:
= o
,
= 0.
In addition to this we shall neglect the elastic strains.
From these assumptions follows immediately
-
:
=

|
|
. -
,
=
c
c
. -

t
t
.
where |, c, and t are the length, the radius of the cross section, and the
thickness of the cylindrical shell, respectively. Besides,
o
i)
=

0 0 0
0 o
,
0
0 0 o
:

.
i. e. we have a homogeneous plane stress state. For materials with Mises
yield function and isotropic hardening
) =
1
2
:
i)
:
i)
/
2
(i).
The computation of p =
1
2
:
i)
:
i)
gives
p =
1
2
:
i)
:
i)
=
1
6
[(o
,
o
:
)
2
+ o
2
,
+o
2
:
]
=
1
3
(o
2
,
o
,
o
:
+o
2
:
).
The ow rule (1.21) take the form
-
j
i)
=

3:
I|
:
I|
4/
2
/

:
i)
.
2.3. CYLINDRICAL SHELL UNDER COMBINED LOAD 43
The loading condition requires that
:
I|
:
I|
= p 0.
Besides, for linear hardening

3
12/
2
/

=
1
/p
Therefore
-
:
=
p
/p
(2o
:
o
,
) =

|
|
.
-
,
=
p
/p
(2o
,
o
:
) =
c
c
.
-

=
p
/p
(o
:
+ o
,
) =

t
t
.
These are three governing dierential equations for three unknowns c. t. |, for
which the loading paths of o
,
and o
:
are given.
We can for example load the tank in two steps:
Step 1: Increase the internal pressure up to the yield stress at j = j
Y
.
Step 2: Keep j = j
Y
= const and increase the longitudinal force `.
p p
Y
N
s
j
s
z
s
jY
Figure 2.15: Loading in two steps
In the rst step
o
:
=
jc
2t
=
1
2
o
,
. o
,
=
jc
t
o
,Y
= j
Y
c
Y
t
Y
.
The yield condition leads to
p
Y
=
1
3
o
2
Y
=
1
3
(o
2
,
o
,
o
:
+ o
2
:
)
=
1
3
o
2
,Y
(1
1
2
+
1
4
) =
1
4
o
2
,Y
.
44 CHAPTER 2. ELEMENTARY THEORY
Consequently
j
Y
=
2

3
t
Y
:
Y
o
Y
.
In the second step (j
Y
= const, o
:
0)
-
,
=
c
c
=
p
/p
(2o
,
o
:
).
-
:
=

|
|
=
p
/p
(2o
:
o
,
)
with
p =
1
3
(o
2
,
o
,
o
:
+ o
2
:
).
p =
1
3
[(2o
,
o
:
) o
,
+ (2o
:
o
,
) o
:
].
Now
o
,
= j
Y
c
t
o
,
= j
Y

(c,t) = j
Y
(
c
t

c

t
t
2
).
Due to the incompressibility

t
t
+
c
c
+

|
|
= 0
we obtain
o
,
= j
Y
c
t
(

|
|
+ 2
c
c
) = j
Y
c
t
( -
:
+ 2 -
,
).
We can therefore express o
,
as follows
o
,
= o
2
,
3
/
p
p
.
The solution of this equation gives
1
o
,
+
3
/
ln p + c = 0.
At o
,
= o
,Y
we have p = p
Y
= o
2
Y
,3, so
o
,
o
,Y
=
1
1
3o
,Y
o
ln
3j
o
2
Y
.
This is a transcendental equation for o
,
. There are still two remaining equa-
tions to determine : and | at a given o
:
and o
,
, where
p
/p
=
2o
:
o
,
p
3
2o
o
2
,
(2o
,
o
:
)
o
:
2/
.
2.4. SIMPLE METAL FORMING PROCESSES 45
2.4 Simple metal forming processes
The elementary theory of metal forming processes (cold work) is quite sim-
ilar to the elementary theory of elasto-plastic deformation. We also begin
here with the simplifying assumptions about the kinematics of the process
under consideration. On the other side the stress states appearing in most
metal forming processes are much more complicated compared with those of
say, simple bending or torsion. Drastic simplications may lead to erroneous
results. That is why we normally made in this type of problems only as-
sumptions about some integral characteristics (for example the plastic work)
rather than assumptions about the stress state.
Since the elastic strains are normally quite small compared with its plastic
counterpart which may be of the order 1, we shall neglect the former in the
cold working problems.
Most metal forming processes such as forging, rolling, drawing and so
on can be explained by quite simple models. We restrict ourselves to the
plane strain problems. Since plastic deformations in cold-worked materials
are normally produced under pressure, we shall regard the latter as positive
despite the traditional convention in mechanics. We denote the pressure by
j. .
We consider for instance:
i) rolling (Fig. 2.16),
ii) drawing, extrusion (Fig. 2.17),
iii) pressing, forging (Fig. 2.18).
Figure 2.16: Rolling: Boundary conditions are time-independent
The strip model
We consider now a strip obtained by cutting (mentally) the piece of metal
at cold rolling along the cuts perpendicular to the plane of motion. For
simplicity we assume the symmetry with respect to the (r. .)-plane. Further
46 CHAPTER 2. ELEMENTARY THEORY
Figure 2.17: Drawing: Boundary conditions are time-independent
Figure 2.18: Forging: Boundary conditions are time-dependent
we assume i) the plane strain state, i. e. the planes perpendicular to r-axis
remain plane, ii) the work done by external forces per unit time is the same
as that of the uni-axial compression (the shear strain are neglected), iii) the
process is quasi-static, the body forces are negligibly small. We take for
Figure 2.19: Force equilibrium of a strip
granted that
a
0. In the planes r = const the shear stress is absent, while
the normal stress (pressure) is uniformly distributed over the cross section.
Thus, the resultant forces in r-direction acting on the strip is at r and
+(),rdr at r +dr. The normal force exerted from the roller on the
strip, j
a
d:, causes the friction
td: = jj
a
d:
2.4. SIMPLE METAL FORMING PROCESSES 47
Thus, the resultant force from the roller on the strip is decomposed into the
vertical component
j

dr = jdr = j
a
d:(cos c + jsin c)
and the horizontal component
j
a
dr = j
a
d:(sin c jcos c).
With
dr = cos cd:.
we obtain
j

dr = jdr = j
a
dr(1 +jtan c).
j
a
dr = j
a
dr(tan c j) = jdr
tan c j
1 + jtan c
.
Denoting the coecient of friction by
j = tan i.
we obtain for the horizontal force
j
a
dr = jdr
tan c tan i
1 + tan i tan c
= jdrtan(c i).
From the assumption ii) we set the power of external forces equal to that
obtained under the uniaxial compression
u = o
Y
dr

= jdr

+ 2j
a
dr
a
+
a
[
a
+

r
(
a
)dr]
= jdr

+ 2j
a
dr
a


r
(
a
)dr.
Dividing this equation by dr and substituting the formula for j
a
into it, we
obtain
o
Y

= j

+ 2
j

tan(c i)
a

r
()

a
r
.
The incompressibility condition requires that

a
= const

a
r
+
a

r
= 0.
48 CHAPTER 2. ELEMENTARY THEORY
Besides

=

r

a
.
Therefore

a
r
=

.
In the elementary theory of forming processes we always use Trescas yield
condition, for which
j = o
Y
.
It follows from here

r
+

r
2j tan(c i) = 0.
Replacing j = + o
Y
and ,r = 2 tan c, we obtain nally

r
+ 2

[tan c tan(c i)] 2


o
Y

tan(c i) = 0.
This ordinary dierential equation of rst order can be written in the form
d
dr
+ )(r) + p(r) = 0.
so, its general solution reads
(r) = c

i
i
0
)oa
[
0

a
a
0
(pc

i
i
0
)oa
)dr].
with =
0
being the value of at r = r
0
. This solution is useful if the
integrals can be computed analytically. Otherwise the numerical solution
turns out to be more eective. After the stress is known, the stress j can
easily be determined.
The tube model
If we deal with the axisymmetric forging process, we may imagine to have
a tube cut from the cold worked metal at forging which is symmetric about
the .-axis as shown in Fig. 2.20.
We assume i) axisymmetric deformation, i. e. cylindrical surfaces about
the .-axis remain cylindrical, ii) the work done by the external forces is the
same as that in the uni-axial compression, iii) the body forces are negligibly
small. Except that we take for granted that

0.
As in the previous strip model we obtain for the force acting on the die
in the .-direction
j
:
d: = jd: = j
a
d:(cos c jsin c) = j
a
d:(1 tan i tan c).
2.4. SIMPLE METAL FORMING PROCESSES 49
Figure 2.20: The tube model
and in the :-direction
j

d: = j
a
d:(sin c tan i cos c) = j
a
d:(tan c tan i) = jd: tan(c i).
From the assumption ii) we set the power of external forces equal to that of
the pressure acting on the tube
u = o
Y
2:d:

= j2:d:

+ 2j

2:d:

2: 2[:

+

:
(:

)d:]
= 2:d:[j

+ 2j tan(c i)

] 2d:[

:
():

+

:
(:

)].
The incompressibility requires

:
+

:

= 0.
or
:

=

:
(:

).
We obtain (after dividing by d\ = 2:d:)
(o
Y
+j )

= [2j tan(c i)

:
()]

.
Using Trescas yield condition we have
j = o
Y
.
It follows that

:
+

:
2j tan(c i) = 0.
50 CHAPTER 2. ELEMENTARY THEORY
and respectively

:
+ 2

[tan c tan(c i)] 2


o
Y

tan(c i) = 0.
provided

< 0.

0. Thus, we obtain the same equation as that of the


strip model, with r replaced by :.
As an example let us consider the compression of a cylinder shown in
Fig. 2.21). In this case c = 0, so the equation reduces to
Figure 2.21: Compression of a cylinder

+ 2j

+ 2j
o
Y

= 0.
The boundary condition is
= 0 at : = 1.
This equation yield the following general solution
(:) = c

r
1
2jo
[(1)
2j

o
Y


1
c

r
1
2jo
d:]
= c

(1)
[
2j

o
Y


1
c
2

(1)
d:]
= o
Y
[c
2

(1)
1].
For j(:) we have
j(:) = o
Y
+ = o
Y
c
2

(1)
The solution found remains valid as long as the stick zone where t = jj
o
Y
,2 does not occurs. This means that the shear stress distribution
t = jj
must be controlled and checked, whether its maximum exceed the value o
Y
,2
or not. With this solution we can compute the total compressive force acting
on the cold forged material
1 = 2o
Y

1
0
c
2

(1)
:d: = 2o
Y
[

2
4j
2
(c
21

1)
1
2j
].
We can easily generalize this model to the so-called plate model.
Chapter 3
Theory of plastic ow
3.1 Governing equations
According to the classication given at the end of Chapter 1 we shall assume
in the theory of plastic ow the ideal plastic material behavior. Thus,
o
Y
= const.
i. e. any hardening behavior is excluded from consideration.
We further admit that the plastic materials obey Mises yield condition
) = J
2
/
2
0
= 0. /
2
0
=
1
3
o
2
0
.
In most of cases we neglect the elastic strains as small compared with the
plastic strains. Thus, as a rule, the material considered in the theory of plastic
ow is rigid-ideal-plastic and obey Mises yield condition. The associate ow
rule reads
d
i)
= -
i)
= -
j
i)
= `
)
o
i)
= `:
i)
where ` remains still an unknown parameter. It can be determined only with
the help of the kinematic boundary condition. From this ow rule we derive
d
i)
d
i)
= `
2
:
i)
:
i)
= 2`
2
/
2
0
.
Therefore
` =
1

2/
0

d
i)
d
i)
.
51
52 CHAPTER 3. THEORY OF PLASTIC FLOW
If the material behaves as elastic-plastic, then the Prandtl-Reuss equa-
tions hold true (see Section 1.4)
-
II
=
1
31
o
II
.
c
i)
=
1
2j
:
i)
+`:
i)
.
The closed system of governing equations must include the equilibrium
conditions, which, in the absence of body forces, read
o
i),)
= 0.
Thus, the whole system of governing equations consists of
3 equilibrium conditions
6 stress-strain relations
1 yield condition
These 10 equations contain 10 unknowns, among which
3 components of velocity
i
6 components of the stress tensor o
i)
1 scalar factor `.
The theory of plastic ow is primarily applied for three types of problems:
onset of plastic ow of a rigid-plastic or elastic-plastic material,
non-stationary plastic ow, provided the boundary conditions of the
states under consideration are suciently known (the so-called pseudo-
stationary plastic ow),
stationary plastic ow.
Analytical solutions are available only if some additional assumptions con-
cerning the kinematics of the plastic ow can be made, for examples in
the torsion of prismatic bars, or
the plane strain problems.
3.2. TORSION OF PRISMATIC BARS 53
3.2 Torsion of prismatic bars
Stress distribution
For the torsion of prismatic bars with an arbitrary cross-section we use the
St. Venant Ansatz
n

= 0.r. n
:
= 0r. n
a
= n(. .).
Here 0 is the twist, and n the warping. The following components of the
stress tensor are zero
o
aa
= o

= o
::
= o
:
= 0.
The equilibrium conditions reduce to
o
a,
+ o
a:,:
= 0.
Figure 3.1: Cross section of a bar
This equation is automatically fullled if there exists a stress function
(. .) such that
o
a
=
,:
. o
a:
=
,
.
The dierential of
d =
,
d +
,:
d. = o
a:
d +o
a
d.
vanishes for
d.
d
=
o
a:
o
a
i. e., when the lines of = const has the same direction as the resultant of
t (Fig. 3.2). The lines = const will therefore be called stress trajectories.
54 CHAPTER 3. THEORY OF PLASTIC FLOW
Figure 3.2: Lines = const
Since the normal stresses vanishes at the boundary of the cross section,
= const at for simply connected cross section.
The plastic zone in the cross section is determined from the yield condition
o
2
a
+ o
2
a:
= t
2
Y
.
where
t
Y
=
{
1
2
o
Y
Tresca.
1

3
o
Y
Mises.
Thus,
(
,
)
2
+ (
,:
)
2
= t
2
Y
= t
Y
.
Nadai has found from the obtained condition (that the gradient or the max-
imal slope of remains constant) the sand roof analogy which is con-
structed by pouring sand on a horizontal sheet of cardboard set out in the
shape of a cross section. Due to the constant internal friction of sand, the
constructed roof satises the above equation.
Figure 3.3: Sand roof
The torque is calculated according to
`
|
=

(o
a:
o
a
.) dc =

(
,
+
,:
.) dc = 2

dc.
so it is equal to twice of the volume under this roof.
3.2. TORSION OF PRISMATIC BARS 55
The determination of the stress distribution o
i)
within the cross section
does not requires the stress-strain relation. Thus, this is the statically deter-
minate problem.
Determination of the warping
According to the ow rule we have in the plastic zone
-
a:
= `o
a:
. -
a
= `o
a
.
Eliminating ` we derive
d-
a:
d-
a
=
o
a:
o
a
.
Since the stress components in the plastic zone are time-independent, we can
integrate this equation to get
-
a:
-
a
=
o
a:
o
a
+ const.
The constant of integration must be equal to zero because it vanishes iden-
tically at the boundary between the elastic and plastic zones. Based on the
St. Venants Ansatz we have
-
a:
=
1
2
(n
,:
+0).
-
a
=
1
2
(n
,
0.).
Substitution in the above equation gives
o
a:
o
a
=
n
,:
+ 0
n
,
0.
= tan ,.
what is completely analogous to the purely elastic torsion, where , denotes
as before the angle between the -Achse and the direction of t. A simple
transformation leads to
n
,:
cos , n
,
sin , = 0(. sin , + cos ,).
or
n
,a
= 0d.
where : is the direction perpendicular to t. With the known stress dis-
tribution and the known direction of t we can nd the warping from this
equation.
56 CHAPTER 3. THEORY OF PLASTIC FLOW
Figure 3.4: Determination of the warping
3.3 Plane strain problems
Governing equations
We are dealing with a 2-D plastic ow, if the components of the velocity are
given by

a
=
a
(r. ).

(r. ).
:
= 0.
and thus,
-
::
= -
:
= -
a:
= 0.
Based on the rigid-ideal-plastic material behavior and the Mises yield condi-
tion, the normality rule reads
-
i)
= `:
i)
. ` 0.
This implies
:
::
= o
:
= o
a:
= 0.
and
o
::
= o =
1
2
(o
aa
+ o

).
and accordingly
o
i)
=

o
aa
o
a
0
o
a
o

0
0 0 o

.
So o is one of the three principal stresses. The remaining two are obtained
as follows
o
1,2
= o
1
2

(o
aa
o

)
2
+ 4o
2
a
3.3. PLANE STRAIN PROBLEMS 57
in the coordinate system obtained by an anticlockwise rotation of the original
axes through an angle , (see Fig. 3.5)
, =
1
2
arctan
2o
a
o
aa
o

.
Thus, the maximum shear stress
t
max
=
1
2
(o
1
o
2
) =
1
2

(o
aa
o

)
2
+ 4o
2
a
= t
is achieved in the direction inclined at the angle ,4 with respect to the
principal axes. The stress state
o
1
= o + t. o
2
= o t. o
3
= o
is characterized by superposition of the pure shear stress in the (r. )-plane
on the hydrostatic pressure.
Figure 3.5: Principal axes and c- and -lines
We denote the directions corresponding to o
1
and o
2
as rst and sec-
ond principal direction, respectively, while the directions obtained by the
clockwise rotations of the principal axes through the angle 45
0
, in which the
maximum and minimum of the shear stresses occur as rst and second slip
direction.
The curve, whose tangent coincides with the slip direction in each point,
is called slip line. It is obvious that there are two orthogonal families of slip
lines. We denote them as c- and -lines, respectively.
Let 0 be the angle between the c-line and the r-axis. Then we obtain
d
dr
= tan 0. for c-lines.
d
dr
= cot 0. for -lines
58 CHAPTER 3. THEORY OF PLASTIC FLOW
as dierential equations for these two families of slip lines.
The stress state must satisfy during the plastic ow the yield condition
1
6
[(o
1
o
2
)
2
+ (o
2
o
3
)
2
+ (o
3
o
1
)
2
] /
2
= 0
For the plane strain problems this condition becomes
t = /.
with / = o
Y
,

3.
The stress state in a particular point is given through o
aa
. o

. o
a
. Due to
the yield condition these components of the stress tensor are not independent.
Let us introduce two dimensionless quantities, . and 0 as follows
. =
o
2/
. 0 =
1
2
arctan
2o
a
o
aa
o


4
= ,

4
.
Thus, 0 is the angle, through which the r-axis is rotated in anti-clockwise
direction to the rst slip line. Then
o
aa
= 2/. / sin 20.
o

= 2/. + / sin 20. (3.1)


o
a
= / cos 20.
We have in addition the equilibrium equations which reduce to
o
aa,a
+ o
a,
= 0.
o
a,a
+ o
,
= 0.
There are altogether three equations to determine three unknown functions.
The plane strain problem is therefore static determined.
Slip lines and their properties
The stress components can be expressed in the plastic zone in terms of vari-
ables . and 0 according to (3.1). Substituting formulas (3.1) into the equi-
librium conditions, we obtain
.
,a
0
,a
cos 20 0
,
sin 20 = 0.
.
,
0
,a
sin 20 + 0
,
cos 20 = 0.
This is the system of homogeneous quasi-linear partial dierential equations
of rst order. Its solution as well as the suitable method of solution depend
3.3. PLANE STRAIN PROBLEMS 59
strongly on the type of these equations. To recognize the latter, we consider
the dierential equations of characteristics
d
dr
=
1
c
(/

/
2
cc)
with
c = sin 20.
2/ = 2 cos 20.
c = sin 20.
The discriminant
/
2
cc = cos
2
20 + sin
2
20 = 1
is positive, so the system under consideration is hyperbolic. For the charac-
teristic directions we have
d
dr
=
1
sin 20
(cos 20 1).
or
d
dr
=
{
tan 0.
cot 0.
Thus, the characteristics of this problem coincide with the slip lines found
above.
Choosing in an arbitrary point 1(r. ) the directions of the r--axes such
that they coincide with the directions of the slip lines (:
c
. :
o
), and taking
into account that
sin 20 = 0. cos 20 = 1
we transform the above equations to
(. 0)
,-
o
= 0. (. + 0)
,-

= 0.
These equations imply that . 0 is constant along the rst slip line, while
. + 0 is constant along the second slip line
d
dr
= tan 0. . 0 = const. c-line.
d
dr
= cot 0. . + 0 = const. -line.
So, provided the slip lines as well as the values of constants on the c- and
-lines are known, then . and 0 and the stress state in each point of the
60 CHAPTER 3. THEORY OF PLASTIC FLOW
(r. )-plane can be determined. However, the directions of characteristics
depend on the solution of this problem.
The slip lines possess several important properties to be discussed below.
Proposition 1: The pressure o changes along a slip line in proportion with
the angle 0.
This property follows directly from the above formulas, according to which
. 0 = const along the c. -lines.
Proposition 2: The change in the angle 0 and the pressure o is the same for
a transition from one slip line of the -family to another along any slip line
of the c-family (Henkys rst theorem).
Figure 3.6: Henckys rst theorem
From the formula
. 0 =
{
/
1
along c
1
.
/
2
along c
2
.
and respectively,
. + 0 =
{

1
along
1
.

2
along
2
.
it follows
0
o
11
=
1
2
(
1
/
1
). 0
o
12
=
1
2
(
2
/
1
).
0
o
21
=
1
2
(
1
/
2
). 0
o
22
=
1
2
(
2
/
2
).
Therefore
,
1
= 0
o
21
0
o
11
=
1
2
(/
1
/
2
).
,
2
= 0
o
22
0
o
12
=
1
2
(/
1
/
2
) = ,
1
.
3.3. PLANE STRAIN PROBLEMS 61
a
b
A
B
C
Figure 3.7: Determination of the pressure
Proposition 3: If the value of o is known at some point of a given slip grid,
then it can be found everywhere in the eld.
Since 0 is known everywhere, we have
o
1
= 2/(
1
0
1
).
and
o
C
= 2/(/
1
+ 0
C
).
Proposition 4: If some segment of a - (or c)-slip line is straight, then all
the corresponding segments of - (or c)-lines are also straight and have the
same length.
E
A
B
A
B

a
b
Figure 3.8: Straight slip lines
This proposition follows from the second proposition, since the angle between
the tangents to any two -slip lines remains constant as we move along
the prescribed -line. The evolute (locus of the center of curvature) of an
arbitrary curve is the envelope of the normals to the curve (Fig. 3.8). It is
evident that the slip line AA

and BB

have the same evolute E. From the


denition of the evolute follows that AB=A

.
Proposition 5: Suppose that we move along some slip line; then the radii of
curvature of the slip lines of the other family at the points of intersection
change by the distance traveled (Henckys second theorem)
Consider neighboring slip lines of the c and families, bounding a slip
62 CHAPTER 3. THEORY OF PLASTIC FLOW
b
a
A
B
C
D
E
E
a
a
a
b
b
b
ds
ds
d
d
d
g
g
g

R
R
Figure 3.9: Henckys second theorem
element d:
c
d:
o
shown in Fig. 3.9. It is apparent that
1
c
d

= d:
c
= 1.
1
o
d = d:
o
= C.
For the arc-length AC we have
C = 1
o
d.
On the other side, from proposition 2 follows
C = (1
o
+ d1
o
+ d:
c
)d.
so
d1
o
= d:
c
. d1
c
= d:
o
.
Proposition 5a: The center of curvature of the -lines at point of intersection
with c-lines generate the involute PT of the c-line (Prandtls theorem).
T
S
R
Q
P
A
A
B
B
C
C
D
D

dg
a
b
Figure 3.10: Prandtls theorem
Since AP=ABQ, PQ must be the involute of the c-line.
3.3. PLANE STRAIN PROBLEMS 63
Proposition 5b: The envelope of the slip lines of one family is the geometric
locus of the cusps of the slip lines of the other family.
It can be seen from Fig. 3.10 that in point T the distance between the c-lines
as well as the radius of curvature of the -lines tend to zero. Therefore: i)
T belongs to the envelope of the family of slip lines, ii) the -line passing
through this point has a cusp at T.
Since they have a cusp at T the -lines cannot intersect the envelope. In
other words, the envelope is the boundary of an analytic solution.
According to Caratheodory and Schmidt an orthogonal grid of slip lines
possessing the above properties is called Hencky-Prandtls grid.
Boundary conditions
The properties of the slip lines discussed in the previous subsection can be
used to construct the Hencky-Prandtls grid, provided one slip line is known
from each family, and they intersect each other.
In order to construct the grid and the solution of a given boundary-value
problem, the boundary conditions should be taken into account.
Figure 3.11: Static boundary condition
For a static boundary condition we must specify the tractions o
a
and t
a
along the boundary .
We compute now o
a
, o
|
and t
a
o
a
=
1
2
(o
aa
+ o

) +
1
2
(o
aa
o

) cos 2 +o
a
sin 2.
o
|
=
1
2
(o
aa
+ o

)
1
2
(o
aa
o

) cos 2 o
a
sin 2.
t
a
=
1
2
(o
aa
o

) sin 2 + o
a
cos 2.
Substitution of o
aa
. o
a
and o

from (3.1) yields


o
a
= 2/. / sin 2(0 ).
o
|
= 2/. + / sin 2(0 ).
t
a
= / cos 2(0 ).
64 CHAPTER 3. THEORY OF PLASTIC FLOW
With the given values of o
a
. t
a
and we get
0 =
1
2
arccos
t
a
/
+ :.
. =
o
a
2/
+
1
2
sin 2(0 ).
The presence of two solutions agrees with the quadratic character of the
yield condition. To choose the sign we consider the normal stress o
|
at the
boundary
o
|
= 4/. o
a
.
The sign of o
|
can sometimes be predetermined, and this enables the correct
choice of the solution.
An important special case occurs when there is no tangential stress at
the boundary (t
a
= 0). Then
0 =
1
2
arccos 0 + : =

4
+:.
. =
o
a
2/
+
1
2
sin(2:

2
) =
o
a
2/

1
2
.
Further
o
|
= 4/. o
a
= o
a
2/.
Consider for example a traction free straight boundary shown in Fig.3.12,
where = 90
c
. o
a
= t
a
= 0. On this boundary
Figure 3.12: Traction free straight boundary
0 =

2


4
+ :.
o = 2/. = /. o
|
= 2/.
Depending on whether the ber in tangential direction is in tension or com-
pression, one chooses the sign + or .
In order to study particular plastic ows, we must nd the solution of the
above quasilinear hyperbolic system of equations satisfying certain boundary
conditions. As a rule, the corresponding boundary value problems can only
3.3. PLANE STRAIN PROBLEMS 65
A
Q
C
B
P
P
P
x
y

Figure 3.13: Cauchys problem


be solved numerically. We now give a brief account of the most important
boundary-value problems.
Boundary-value problems.
1. Cauchys problem. Let AB be a smooth arc (described by an arc-length
:), which nowhere coincides with characteristics and intersects each charac-
teristic once only. Let the functions .(:) and 0(:) be continuous together
with their derivatives on AB. Then the solution exists and is unique in a
triangular region APB, bounded by the arc AB and the c- and -slip lines
originating at A and B. The solution at poit P depends only on the data
along AB, therefore the region APB is called the domain of dependence for
point P.
Note that if derivatives of the initial data are discontinuous at some point
C on the curve AB, then the jumps propagate along the characteristics CP

and CP

.
If the shear stress t
a
is zero at the boundary, the normal to this boundary
is one of the principal directions and the slip lines approach the contour
at an angle of 45
c
. Consequently, the contour coincides nowhere with a
characteristics, and we have Cauchys problem whose solution is unique.
A
B
x
y
(i,j)
a
b
(i+1,j)
(i,j-1)
Figure 3.14: Determination of . and 0
If the boundary is traction-free, then the stress eld as well as the grid of
66 CHAPTER 3. THEORY OF PLASTIC FLOW
slip lines in the inuence domain depend only on the shape of the boundary.
Provided the values of o
a
and t
a
are given at the boundary 1, then
we can compute . and 0 in each point of this curve. The proper choice of
solution is based on the predetermined sign of o
|
.
Let us denote the nodal points of the grid of characteristics such that
on c-lines the rst index i is constant and on -lines the second index , is
constant. Then according to proposition 2 the quantities . and 0 at some
point (i. ,) are given by
.(i. ,) =
1
2
[.(i. , 1) + .(i + 1. ,) + 0(i + 1. ,) 0(i. , 1)].
0(i. ,) =
1
2
[.(i + 1. ,) .(i. , 1) + 0(i + 1. ,) + 0(i. , 1)].
Thus, the grid as well as the stress eld can be determined by these equa-
tions in nite dierence. Note that such the grid of slip lines satises the
orthogonality condition only approximately, since the small arcs connecting
nodal points are replaced by the chords.
A
B
x
y
a
b
Figure 3.15: Riemanns problem
2. Riemanns problem. Suppose the initial data are known on the segment
AB of the slip line. In this case the domain of inuence of AB covers the whole
strip between the slip lines of the other family passing through A and B. In
this case the unique solution must be found by the following requirements:
i) the second slip lines are orthogonal to AB, and ii) . + 0 = const along
these slip lines.
3. Mixed boundary value problem. Let us consider now the combination of
the above problems. Namely, assume that . and 0 are known on a segment
of the slip line OB. This segment joints a chracteristic curve OA, along which
the angle 0 is prescribed. The solution to the mixed boundary value problem
is determined in the triangle C1. The approximate solutions to 2. and 3.
are similar to 1.
3.3. PLANE STRAIN PROBLEMS 67
A
B
x
y
a
b
O
Figure 3.16: Mixed boundary value problem
Velocity eld.
If the stress eld is known, then the velocity eld can also be determined.
According to the ow rule we have
-
aa
= `:
aa
.
-

= `:

.
-
a
= `o
a
.
From the rst two equations
-
aa
-

= `(o
aa
o

).
Combine this with the third equation
( -
aa
-

)
o
a
o
aa
o

= -
a
.
Taking into account the relation
o
a
o
aa
o

=
1
2
tan 2, =
1
2
tan(20

2
) =
1
2
cot 20
and the incompressibility, we obtain

a,a
+
,
= 0.
(
a,
+
,a
) tan 20 +
a,a

,
= 0.
This system of homogeneous quasilinear partial dierential equations of rst
order is of hyperbolic type, and its characteristics coincide with the slip lines.
With the transformation

a
=
c
cos 0
o
sin 0.

=
c
sin 0 +
o
cos 0.
68 CHAPTER 3. THEORY OF PLASTIC FLOW
we obtain for the derivatives

a,a
=
c,a
cos 0
c
sin 00
,a

o,a
sin 0
o
cos 00
,a
.

,
=
c,
sin 0 +
c
cos 00
,
+
o,
cos 0
o
sin 00
,
.
Choosing in an arbitrary point 1(r. ) the directions of the r--axes such
that they coincide with the directions of the slip lines (:
c
. :
o
), and recalling
that
sin 0 = 0. cos 0 = 1
we obtain

a,a
=
c,a

o
0
,a
= 0 (c).

,
=
c
0
,
+
o,
= 0 ().
These equations are called Geiringers equations for the velocity eld.
If the grid of the slip lines is known from the solution of the stress problem,
then the velocity eld can also be determined from these equations for all
three types of the boundary value problems.
Note that the line separating plastic and rigid regions is a slip line or
the envelope of slip lines. To show this we rst assume that the velocities
are continuous on the line of separation. Then
a
=

= 0 on this line.
If the boundary has nowhere a characteristic direction, then the solution of
Cauchys problem gives
a
=

= 0 everywhere in the plastic region, which


contradicts the original assumption. Consider now the case when the velocity
is discontinuous on the line of separation. The discontinuity can only be in

|
, otherwise a crack appears. One can then show that o
|
/. Thus, the
boundary will be a slip line or an envelope of slip lines.
Using Geiringers equation and considering the velocity at some nodal
point (i. ,) of the grid of slip line, we can easily obtain the following approx-
imate nite dierence formulas

c
(i. ,)
c
(i. , 1) =
o
(i. ,)[0(i. ,) 0(i. , 1)].

o
(i + 1. ,)
o
(i. ,) =
c
(i. ,)[0(i + 1. ,) 0(i. ,)].
For statically determinate problems, i. e. when the static boundary condi-
tions are specied at the boundary , the grid of slip lines as well as the
stress eld can be determined numerically, so the velocity eld can be found
from these formulas.
For statically undeterminate problems, which have not been considered
up to now, the coupling cannot be removed. To construct the grid of slip
lines all four dierence equations need to be solve simultaneously.
3.3. PLANE STRAIN PROBLEMS 69
For problems with mixed boundary conditions (where the tractions are
specied on one part of the boundary and the velocity on the remaining part)
we nd the grid of the slip lines as well as the stress and velocity elds in
each of the zone of inuence as above. The initial data for the velocity can be
obtained from the transition conditions at the boundary between the zones
of inuence.
Family of straight slip lines.
Many of the above statements become quite simple for the family of straight
slip lines.
Due to the Proposition 4 there are only two possible grids with straight
slip lines
grid with two orthogonal families of straight slip lines,
grid with one family of straight slip lines and the curved slip lines
orthogonal to them.
In any case
. 0 = const. c line.
. + 0 = const. line.
Since 0 remains constant along a straight line, . and o must also be constant
there.
For the rst case we have a uniform stress state in the whole domain and
the velocity eld of the form

a
= )().

= p(r).
A domain in which one family of slip lines contains only straight line is
called a fan. We distinguish between central and non-central fans.
The central fan has the -lines as concentric circles. Here the envelope
degenerate to a point O which is the middle point of all circles. In this case
0 = const . = const on c-line.
and
. + 0 = const on -line.
Consequently
. + 0 = const in the whole domain.
70 CHAPTER 3. THEORY OF PLASTIC FLOW
Because . = c 0 is independent of :, we obtain for the stresses
o

= o
,
= 2/. = 2/(c 0).
o
,
= /.
With the help of Geiringers equation we compute the velocity

= )(0)
,u
.

,
= )(0) + p(:).
For the non-central fan the similar results can also be obtained.
3.4 Plane stress problems
Governing equations
We are dealing with a plane stress state when, in some cartesian coordinate
system,
o
aa
= o
aa
(r. ).
o

= o

(r. ).
o
a
= o
a
(r. ).
o
::
= o
a:
= o
:
= 0.
and, accordingly
o
i)
=

o
aa
o
a
0
o
a
o

0
0 0 0

. o =
1
3
(o
aa
+ o

).
For the principal stresses in the (r. )-plane we have as before
o
1,2
=
1
2
(o
aa
+ o

)
1
2

(o
aa
o

)
2
+ 4o
2
a
.
These principal stresses occur in a coordinate system which is rotated through
the angle
, =
1
2
arctan
2o
a
o
aa
o

with respect to the original coordinate system.


This plane stress state must fulll in the plastic zone the yield condition,
which is either i) Mises
1
6
[(o
1
o
2
)
2
+ (o
2
o
3
)
2
+ (o
3
o
1
)
2
] /
2
= 0.
3.4. PLANE STRESS PROBLEMS 71
with the consequence
o
2
1
+ o
2
2
o
1
o
2
3/
2
= 0.
or ii) Tresca
1
2
max{o
1
o
2
. o
2
o
3
. o
3
o
1
} t
Y
= 0.
with the consequence
o
1
o
2
2t
Y
= 0. o
1
o
2
0.
o
1
2t
Y
= 0. o
1
o
2
0. o
1
o
2
.
o
2
2t
Y
= 0. o
1
o
2
0. o
2
o
1
.
Figure 3.17: Yield surface for the plane stress state
The local stress state at an arbitrary point is xed by three components
o
aa
. o

. o
a
. For them the yield condition as well as the two remaining
equilibrium conditions
o
aa,a
+ o
a,
= 0.
o
a,a
+ o
,
= 0.
hold true. The problem is therefore statically determinate. However, in
contrast to the plane strain problems the above two yield conditions lead to
dierent results.
Solution based on Mises yield condition
In addition to the rotation angle , let us introduce a new angle .
cos . =

3
o
2/
.
72 CHAPTER 3. THEORY OF PLASTIC FLOW
so that, with the help of
o
1,2
= 2/ cos(.

6
)
the ow rule is satised identically.
Then
o
aa
= /(

3 cos . + sin . cos 2,).


o

= /(

3 cos . sin . cos 2,).


o
a
= / sin . sin 2,.
Substituting these formulas into the equilibrium conditions we obtain nally
(cos .

3 sin . cos 2,).


,a

3 sin . sin 2,.


,
+ 2 sin .,
,
= 0.

3 sin . sin 2,.


,a
(cos . +

3 sin . cos 2,).


,
+ 2 sin .,
,a
= 0.
With
c = 2 sin .(cos .

3 sin . cos 2,).


/ = 2 sin
2
.

3 sin 2,.
c = 2 sin .(cos . +

3 sin . cos 2,).


we derive for the discriminant
/
2
cc = 4 sin
2
.(3 4 cos
2
.) = 4 sin
2
.(.).
Depending on the sign of (.) we get the hyperbolic, parabolic or elliptic
system of partial dierential equations (see Fig. 3.18).
1 2 3 4
1
0.75
w
hyperbolic elliptic
Figure 3.18: Graph of function cos
2
.
For the characteristic directions we have
d
dr
=

3 sin . sin 2,

(.)
cos .

3 sin . cos 2,
.
In the hyperbolic case the problem can be solved with the help of char-
acteristics.
Chapter 4
Crystal plasticity
This Chapter discusses crystal plasticity. We aim at studying the dislocation
nucleation and accumulation, the resulting work hardening and the inuence
of the resistance to dislocation motion. Among various problems we select
antiplane shear and plane constrained shear of single crystals which admit
analytical solutions. The interesting features of these solutions are the ener-
getic and dissipative thresholds for dislocation nucleation, the Bauschinger
translational work hardening, and the size eects.
4.1 Physical background
One of the remarkable properties of crystals is their ability to glide easily
on certain crystallographic planes along certain crystallographic directions.
The simplest way to see this is to do a tension test for a bar made of a single
crystal whose slip planes are inclined at some angle, say, 45

to the bar axis


(see Fig. 4.1a). By measuring the load 1 and the corresponding elongation
| of the bar, we can plot the stress o = 1, versus strain - = |,| curve,
which is shown (schematically) in Fig. 4.1b.
At small stresses the bar deforms elastically, and the stress-strain curve
is nearly linear. If the force is removed, the bar length returns to its initial
value | so that | = 0. Beginning from a critical value o
Y
, called yield stress,
the bar deforms plastically and then, at larger stresses, easily stretches out
to a narrow ribbon (necking) which may be four or ve times longer than the
original length. If one examines the surface of the plastically deformed bar,
steps can be seen which run more or less continuously around the boundary
in the form of ellipses. These steps appear as witnesses of the plastic slips
on the active slip planes (see Fig. 4.2).
It is observed experimentally that plastic slips occur on the close packed
73
74 CHAPTER 4. CRYSTAL PLASTICITY
F
-F
s
s
Y
e
a)
b)
Figure 4.1: Tensile test and the stress-strain curve
slip
plane
a b
Figure 4.2: Magnied schematic view of plastic slips: a) front view, b) side
view
planes along the directions of shortest interatomic distances. To recognize
these planes and directions one needs to consider the periodic lattice structure
of the crystal. We see for example in Fig. 4.3 the unit cell of a body-centered
cubic (bcc) crystal. The unit cell is characterized by three lattice vectors a
1
,
a
2
, and a
3
which, for bcc crystals, are mutually orthogonal and have an equal
magnitude. The translations of the unit cell by :
1
a
1
+ :
2
a
2
+ :
3
a
3
for all
integers :
1
. :
2
. :
3
will develop the regular periodic lattice structure. In each
unit cell of the bcc crystal there are eight corner atoms and one atom located
at its center. Crystals may have also other lattice structures, for instance,
face-centered cubic (fcc) or hexagonal closed-packed (hcp) lattice structures.
When referring to crystallographic directions and planes, we need the
way to indicate them. We shall use Miller indices for this purpose. For
example, the Miller indices of the direction d in Fig. 4.3 are [111], while
those of the direction : in the same Figure are [101], where the negative sign
of the r
1
-component is indicated by a bar over the corresponding index. We
dene the Miller indices of a family of directions as a set of three integers
4.1. PHYSICAL BACKGROUND 75
:
1
. :
2
. :
3
, enclosed in square brackets, which indicates all directions parallel
to the vector :
1
a
1
+ :
2
a
2
+ :
3
a
3
. Since :
1
. :
2
. :
3
are dened uniquely up
to a common integer factor, we will choose the smallest multiples of them
for the Miller indices. For illustration let us apply this denition to nd the
Miller indices of the direction c in Fig. 4.3. Since the vector drawn from the
origin parallel to c has the smallest integers 1,-1,0 as its components, we get
for the direction c the Miller indices [110]. Crystallographic planes are also
indicated by sets of integers. For example, the Miller indices for the (r
1
. r
2
),
(r
2
. r
3
), and (r
3
. r
1
)-planes are (001), (100), and (010), respectively. Note
that the Miller indices for planes are enclosed in parentheses instead of square
brackets as for directions. We dene the Miller indices of a family of planes
as a set of three integers :
1
. :
2
. :
3
, enclosed in parentheses, which indicates
all planes perpendicular to the vector :
1
a
1
+ :
2
a
2
+ :
3
a
3
. Since the normal
vector :
1
a
1
+:
2
a
2
+:
3
a
3
is dened uniquely up to a common integer factor,
we choose again the smallest multiples of :
1
. :
2
. :
3
for the Miller indices. For
illustration let us apply this denition to nd the Miller indices of the plane
shown in Fig. 4.3. The normal vector to this plane is obviously a
1
+a
2
+a
3
,
so the Miller indices of this plane are (111).
x
1
x
2
x
3
s
d
e
a
1
a
2
a
3
Figure 4.3: Unit cell of bcc crystals
In terms of these Miller indices a precise description of the plastic slip
can be given. For example, the plastic slip may occur in bcc crystals on the
slip planes of type (110) along the directions [111]. For fcc crystals, the slip
planes are of type (111), while the slip directions of type [110]. Each slip
system is characterized by the family of slip planes and slip directions. For
fcc crystals there are 12 slip systems.
The critical shear stress of a perfect crystal at which the plastic slip occurs
can roughly be estimated as follows. Consider the glide of the upper half of
crystal relative to its lower half under a shear stress t. The two adjacent rows
of atoms in the initial state are shown in Fig. 4.4 in which the interplanar
spacing is c while the interatomic distance in the slip direction is /.
Denoting the relative displacement by n, the applied shear stress must be
76 CHAPTER 4. CRYSTAL PLASTICITY
a
b
u
E
t
Figure 4.4: Perfect crystal and energy vs. relative displacement
calculated in accordance with t = d1,dn, where 1 is the energy of crystal
in terms of n. Due to the periodicity of the lattice structure, 1 must be
a periodic function with the period /, having a minimum at n = 0 and a
maximum at n = /,2 (see Fig. 4.4). The simplest formula for 1 satisfying
these requirements can be proposed as follows
1(n) =
1/
2
(
1 cos
2n
/
)
.
The derivative of 1 yields the shear stress
t =
d1
dn
= 1 sin
2n
/
.
For small displacement n / we have approximately t = 12n,/. Com-
paring this with the Hookes law t = jn,c, with j being the shear modulus,
we obtain the value of 1 which coincides with the maximum shear stress
achieved at n = /,2
t
|
= 1 =
j/
2c
.
In typical situations we have / = c, so the theoretical shear strength is
estimated to be
t
|

j
2
.
For most of crystals this theoretically estimated shear strength is three or
four order of magnitude larger than the experimentally observed value of o
Y
.
In 1934 three scientists, Taylor, Orowan, and Polanyi, simultaneously
came to the idea that the plastic slip can so easily occur at very low shear
stress because of dislocations which are the line defects in crystals. One
should mention that, actually, the theory of dislocations in solids was origi-
nally developed by Volterra in 1905. But only the discovery by these pioneers
gave birth to the dislocation based plasticity. In Fig. 4.5 one can see the
so-called edge dislocation in a simple cubic lattice structure which can be
4.1. PHYSICAL BACKGROUND 77
A
D
C
B
Figure 4.5: Edge dislocation in primitive cubic crystal
regarded as the extra half-plane of atoms inside the crystal (the half-plane
ABCD). The distortion of the crystal is localized near the edge of this half-
plane which is called the dislocation line (the line AB). Far away from the
dislocation line the crystal has a nearly perfect lattice structure.
t t
t t
Figure 4.6: Migration of dislocation through crystal
Fig. 4.6 illustrates how the edge dislocation migrates through the crystal
from left to right under an applied shear stress t. This migration, from one
position to the next, involves only a small rearrangement of the atomic bonds
near the dislocation line. The nal results of these many small steps is the
plastic slip considered before. But the shear stress to break the atomic bond
and move dislocation in one interatomic distance is much lower that that
required for the simultaneous movement of the whole upper half of crystal.
Let us now consider the creation of a single dislocation in a simple cubic
crystal. Cut this crystal along any of the surfaces indicated in Fig. 4.7a,b,c.
Let the atoms on one side of the cut shift in a direction parallel to the cut
through a distance equal to one lattice spacing. Then rejoin the atoms on
both sides of the cut and let the crystal be elastically relaxed. The lattice
structure is again almost perfect except near the boundary AB of the cut
surface called dislocation line. If the atoms below the cut are shifted in the
direction parallel to the line AB, a screw dislocation is created; if the shift is
perpendicular to AB, an edge dislocation is produced; if the shift is neither
parallel nor perpendicular to AB, the created dislocation is of the mixed type.
So, each dislocation is characterized by the dislocation line and the vector
78 CHAPTER 4. CRYSTAL PLASTICITY
pointing in the direction of shift whose magnitude is equal to one lattice
spacing. This vector is called Burgers vector.
A
A
A
B
B
B
a)Screw
b)Edge
c)Mixed
C
D
C
D
C
Figure 4.7: Creation of dislocation in a simple cubic crystal
1 2 3 4 5
6
7
8
9 10 11 12
13
1 2 3 4 5
6
7
8
9 10 11
12
13
14
14
b
a)
b)
Figure 4.8: Burgers circuit around an edge dislocation
The denition of Burgers vector depends on the sense of a dislocation
line. We dene the sense of the dislocation line by assigning a unit vector
tangent to the dislocation line and taking the positive sense in the positive
direction of . Burgers vector can be determined by the following geometric
construction. Consider two circuits drawn in Fig. 4.8a,b. The left close circuit
is drawn in the reference dislocation-free crystal; the right is drawn around
an edge dislocation in real crystal. When moving along each circuit in the
positive direction agreeing with the sense of the dislocation line (anticlockwise
4.2. CONTINUUM DISLOCATION THEORY 79
if the positive sense of the dislocation line is taken to be out of the paper)
the same number of jumps from atom to atom are made to the right as to
the left, up as down. The starting and the end points correspond to the
same atom in the case of the circuit in the reference crystal. However the
starting and end points are not the same atom for the circuit that includes
the dislocation in the real crystal. Thus, there is a closure failure (or a mist)
in this circuit dened as Burgers vector.
4.2 Continuum dislocation theory
Macroscopically observable plastic deformation of single and polycrystals is
produced essentially by the motion of a large number of dislocations. On the
other side, these newly formed dislocations in crystals pile up near various
obstacles like grain or phase boundaries, or particulate inclusions, giving rise
to size dependent hardening of the material. Dislocations appear in the de-
formed crystal lattice to reduce its energy. Motion of dislocations generates
the dissipation of energy which, in turn, results in a resistance to disloca-
tion motion. The understanding of nucleation mechanism and the motion
of dislocations is therefore a cornerstone for describing plastic yielding, work
hardening, and hysteresis eects in crystal plasticity.
Furthermore, dislocations are not only a key microstructural defect for
plastic slip but also the core ingredient for forming microstructural patterns
and substructures. There are numerous examples illustrating this. The rst
one is the formation of lamellar twin patterns in manganese steels and other
twinning induced plasticity (TWIP)-alloys, which has signicant impact on
the macroscopic stress-strain response. The formation of twins provides
TWIP-alloys with excellent hardening behavior, allowing for higher stresses
and larger strains than in common fcc or bcc metals. The other example
is the recrystallization produced by severe plastic deformation during equal
channel angular extrusion which leads to almost dislocation-free grains of
an average diameter of a few hundred nanometers, yielding materials with
exceptional room-temperature strength.
From the above examples, one can clearly see the importance of being
able to predict the nucleation and motion of dislocations in crystals. The
high dislocation densities accompanying plastic deformation, in the range of
10
8
-10
15
m
2
, along with the complexity of the dislocation network necessitate
the continuum approach to dislocations, which is the subject of this Chapter.
Continuum dislocation theory deals with the ensembles of a huge number of
dislocations by the methods of continuum mechanics. The complexity of the
system makes the phenomenological approach unavoidable. However, one
80 CHAPTER 4. CRYSTAL PLASTICITY
needs some guiding principles to limit the feasible choices of the available
models. Such guiding principles are the laws of thermodynamics.
s
m
e
p
e
e
e e e = +
e p
Figure 4.9: Additive decomposition of the total strain
Consider rst a crystal deforming in single slip. The strains of an in-
nitesimal element of the crystal is additively decomposed into the plastic
strains and the elastic strains as shown schematically in Fig. 4.9. The plas-
tic strain tensor is the symmetric part of the plastic distortion which maps
the undeformed crystal to the reference stress-free crystal. The subsequent
elastic strains deform the crystal in accordance with the Hookes law. Both
plastic and elastic strains are incompatible so that the sum of them becomes
compatible and derivable from a displacement eld. For a large number of
loops lying on the parallel slip planes with the mean distance between them
being much smaller than the characteristic size of the specimen, we propose a
spatial average description of plastic distortion produced by this slip system
as follows

i)
= (x):
i
:
)
.
with s the unit vector pointing in the slip direction, and m the normal vector
to the slip plane. The essential dierence to the case of discrete dislocations
is that now function (x) is assumed to be continuously dierentiable. If the
crystal has : active slip systems, the plastic distortion is the sum of those
produced by these slip systems

i)
=
a

=1

(x):

i
:

)
.
4.2. CONTINUUM DISLOCATION THEORY 81
The small Gothic index indicating the slip systems runs from 1 to :. One
can see that, in general,
ii
= 0, consequently continuous plastic distortions
do not cause any volume change.
The plastic strains -
j
i)
and the plastic rotations .
i)
are the symmetric and
skew-symmetric parts of the plastic distortion
-
j
i)
=
1
2
(
i)
+
)i
). .
i)
=
1
2
(
i)

)i
).
The elastic strains are then the dierences between the total compatible
strains and the incompatible plastic strains
-
c
i)
= -
i)
-
j
i)
=
1
2
(n
i,)
+ n
),i
) -
j
i)
.
with n
i
being the components of the displacement vector.
Nye introduced an important characteristic of dislocations, the dislocation
density tensor
1
c
i)
= c
)I|

i|,I
. (4.1)
For a single dislocation loop the dislocation density tensor has the following
physical meaning: if we take an arbitrary innitesimal surface dc with the
unit normal which is the tangent vector to the dislocation line crossing
this surface, c
i)
:
)
dc gives the Burgers vector of this dislocation (see Section
3.4). Within the continuum dislocation theory, we interpret this quantity as
the resultant Burgers vector of all dislocations whose dislocation lines cross
the surface dc. For a crystal deforming in single slip c
i)
:
)
= :
i
c
)I|

,I
:
|
:
)
,
so the resultant Burgers vector turns out to be parallel to the slip direction.
The number of dislocations per unit area can then be computed as
j =
1
/
c
)I|

,I
:
|
:
)
. (4.2)
with / the magnitude of Burgers vector.
Let be any regular sub-region of the crystal in its initial state. The free
energy of the crystal conned in the region reads
=

o(-
i)
.
i)
. c
i)
. o) dr.
with o being the absolute temperature which is assumed to be constant.
Under this assumption the laws of thermodynamics state that the rate of
1
In fact, Nye introduced the dislocation density tensor produced by a plastic distortion
in form of rotation only. Formula (4.1) was proposed independently by Bilby and Kroner.
82 CHAPTER 4. CRYSTAL PLASTICITY
change of the free energy minus the power of the external forces must be
non-positive for arbitrary processes

1 =
d
dt

o(-
i)
.
i)
. c
i)
. o) dr 1 0. (4.3)
The structure of power must be controlled by the form of the energy. In our
case the power is given by
1 =

(o
i)
:
)
n
i
+ o
i)I
:
I

i)
) dc. (4.4)
where is the boundary of with :
i
being the unit outward normal to
. We see that some stresses of higher order enter the theory as a result
of the dependence of the free energy density on the gradient of the plastic
distortion.
Transforming the surface integral in (4.4) into a volume integral by Gauss
theorem and requiring that (4.3) is satised for an arbitrary , we obtain
the inequality
(
o
-
i)
o
i)
)
n
i,)
+
(
o

i)
o
i)I,I
)

i)
+
(
o
c
in
c
nI)
o
i)I
)

i),I
o
i),)
n
i
0. (4.5)
For rigid translations the energy does not change while n
i,)
,

i)
and

i),I
are zero. Since the inequality (4.5) must be fullled for an arbitrary transla-
tion, the stress must obey the equilibrium equation
o
i),)
= 0. (4.6)
Similarly, the inequality (4.5) can be satised for arbitrary rigid rotations
only if the stress tensor is symmetric
o
i)
= o
)i
. (4.7)
Let us introduce the following notation
t
i)
= o
i)

o
-
i)
. t
i)I
= o
i)I

o
c
in
c
nI)
. (4.8)

i)
=
o

i)
+ o
i)I,I
. (4.9)
4.2. CONTINUUM DISLOCATION THEORY 83
Then the combined rst and second laws of thermodynamics become
t
i)
n
i,)
+
i)

i)
+ t
i)I

i),I
0. (4.10)
Equation (4.10) shows that t
i)
and t
i)I
are those parts of the stresses and
the higher order stresses which cause energy dissipated in heating of the
crystal. Tensor t
i)
describes heating in a non-uniform ow, so it has the
meaning of viscous stresses. Tensors
i)
and t
i)I
describe heating caused by
homogeneous and inhomogeneous plastic deformation, respectively.
The widely used closure of non-equilibrium thermodynamics assumes that
there exists a dissipation potential
1 = 1( n
i,)
.

i)
.

i),I
) (4.11)
such that the tensors t
i)
,
i)
, and t
i)I
controlling the irreversible processes
are linked to n
i,)
,

i)
, and

i),I
by the relations
t
i)
=
1
n
i,)
.
i)
=
1

i)
. t
i)I
=
1

i),I
. (4.12)
The set of equations (4.6),(4.7),(4.8),(4.9), and (4.12) is closed with respect
to the unknown functions n
i
and
i)
.
Thus, each continuum model of dislocations is xed by two functions,
namely, the free energy density and the dissipation potential. For isothermal
processes we require the free energy density to depend only on the elastic
strain -
c
i)
and on the dislocation density c
i)
:
o(-
i)
.
i)
. c
i)
) =
1
2
C
i)I|
-
c
i)
-
c
I|
+ o
n
(c
i)
). (4.13)
where o
n
(c
i)
) corresponds to the energy of the dislocation network. In this
case the stresses are given by
o
i)
=
o
-
c
i)
= C
i)I|
-
c
I|
.
The crucial question is then how the energy of the dislocation network de-
pends on the dislocation density. For a single crystal deforming in single slip
we shall adopt the following formula
o
n
= /jln
1
1 j,j
-
. (4.14)
where j
-
is the saturated dislocation density and / a material constant. The
logarithmic energy stems from two facts: i) for small dislocation densities the
84 CHAPTER 4. CRYSTAL PLASTICITY
energy of the dislocation network must be proportional to the dislocation den-
sity, and ii) there exists a saturated dislocation density which characterizes
the closest packing of dislocations of equal signs admissible in the discrete
crystal lattice. The logarithmic term ensures a linear increase of the energy
for small dislocation density j and tends to innity as j approaches the sat-
urated dislocation density j
-
hence providing an energetic barrier against
over-saturation.
Concerning the dissipation potential several models can be considered.
The simplest model assumes that the dissipation is zero. In this case all
tensors t
i)
,
i)
, and t
i)I
vanish, and functions n
i
and
i)
should be found
from energy minimization. The next model, also quite simple, neglects the
viscous eect as well as the dissipation caused by

i),I
. In this model 1 is
assumed to depend only on

i)
so that t
i)
= 0 and t
i)I
= 0 and
o
i)
=
o
-
i)
. o
i)I
=
o
c
in
c
nI)
. (4.15)
If, furthermore, 1 is a homogeneous function of rst order with respect to

i)
, then the evolution equation for
i)
becomes

i)
=
1

i)
. (4.16)
Note that, for the free energy density in form (4.13), equation (4.16) can be
written in a variational form
1

i)
=
o
-
o
o
i)

o

i)

-
.
+

r
I
o

i),I
. (4.17)
We will consider in what follows only these simplied dissipation potentials.
4.3 Anti-plane constrained shear
We start as always with the simplest problem of a beam made of a single
crystal undergoing an anti-plane shear deformation. Let be the cross sec-
tion of the beam by planes . = const. For simplicity, we consider to be
a rectangle of width c and height , 0 < r c, 0 < . We place the
crystal in a hard device with the prescribed displacement at the boundary
[0. 1] (see Fig. 4.10)
u = at [0. 1].
4.3. ANTI-PLANE CONSTRAINED SHEAR 85
x
y
a
z
h
L
Figure 4.10: Anti-plane constrained shear
where u(r. ) is the .-component of the displacement vector and corre-
sponds to the overall shear strain. The height of the cross section, , and
the length of the beam, 1, are assumed to be much larger than the width
c (c , c 1) to neglect the end eects and to have the stresses and
strains depending only on one variable r in the central part of the beam. If
the shear strain is suciently small, then the crystal deforms elastically and
u = everywhere in the beam. If exceeds some critical value, then the
screw dislocations may appear. We allow only the slip planes parallel to the
plane = 0 and the dislocation lines parallel to the .-axis. Our aim is to
determine the distribution of dislocations as function of within the frame-
work of continuum theory of dislocations proposed in Section 4.2. For screw
dislocations with the slip planes parallel to the plane = 0, the tensor of
plastic distortion,
i)
, has only one non-zero component
:
. We assume
that depends only on r-coordinate: = (r). Since the displacements are
prescribed at the boundary of the crystal, dislocations cannot penetrate the
boundaries r = 0 and r = c, therefore
(0) = (c) = 0. (4.18)
The plastic strains are given by
-
j
:
= -
j
:
=
1
2
(r).
The only non-zero component of Nyes tensor of dislocation density is
c
::
=
,a
.
86 CHAPTER 4. CRYSTAL PLASTICITY
The free energy density of the crystal with dislocations takes a simple form
o =
1
2
j( )
2
+ j/ ln
1
1
,a
,j
-
/
. (4.19)
If the resistance to the dislocation motion is negligible (and, hence, the
dissipation is zero), the true plastic distortion minimizes the total energy, ,
which is a functional of (r),
[(r)] = 1

o
0
[
1
2
j( )
2
+ j/ ln
1
1
,a
,j
-
/
]
dr (4.20)
among all admissible function (r) satisfying the boundary conditions (4.18).
The total strain, , is regarded as a given function of time, so one can study
the evolution of the dislocation network which accompanies the change of
the total strain.
If the resistance to the dislocation motion cannot be neglected, then the
energy minimization must be replaced by a ow rule. In case of the rate-
independent plasticity, when the dissipation potential is
1 = 1

.
with 1 denoting the critical resolved shear stress, the ow rule reads: for

= 0
1

=
o

o
o
. (4.21)
The right-hand side of (4.21) is the negative variational derivative of energy
with respect to at xed total strain

o

o
o
=
o

+

r
o

,a
.
With o from (4.19) we obtain
= j( ) + j/

,aa
(j
-
/
,a
)
2
. (4.22)
According to (4.22), if = const in some sub-interval of (0. c), coincides
with the shear stress o = j( ). For

= 0 the evolution equation (4.21)
needs not be satised: it is replaced by the equation

= 0.
So, for

= 0, the evolution equation for is
1sign

= j( ) + j/

,aa
(j
-
/
,a
)
2
(4.23)
4.3. ANTI-PLANE CONSTRAINED SHEAR 87
which must be subject to the boundary conditions (4.18).
According to (4.23) the plastic distortion may evolve only if the yield
condition
j( ) + j/

,aa
(j
-
/
,a
)
2
= 1 (4.24)
is fullled. If < 1 then is frozen:

= 0. If the yield condition (4.24)
holds, function (t. r) may evolve or may stay unchanged: this depends on
the time dependence of the control parameter .
We rst analyze the situation when the resistance to the dislocation mo-
tion is negligible (and, hence, the dissipation is zero). In this case the de-
termination of (r) reduces to the minimization problem (4.20). Note that,
since o is convex with respect to and
,a
, this variational problem has a
unique solution. It is convenient to introduce the following dimensionless
quantities
=
r/j
-

. n() =
(r)

. : =
/

2
. 1 =
/j
-
j1
3
. (4.25)
The dimensionless variable changes on the interval (0. c), where c = c/j
-
,.
The functional (4.20) reduces to
1[n()] =

c
0
[
1
2
(1 n)
2
+ :ln
1
1 n

]
d. (4.26)
where the prime denotes dierentiation with respect to . We minimize
functional (4.26) among functions n() satisfying the boundary conditions
n(0) = n(c) = 0.
It is instructive to analyze rst the variational problem in which the
logarithmic term is replaced by an asymptotic formula
ln
1
1 n

+
1
2
n
2
.
The functional to be minimized becomes
1[n()] =

c
0
[
1
2
(1 n)
2
+:(n

+
1
2
n
2
)
]
d. (4.27)
Due to the boundary conditions n

should change its sign on the interval


(0. c). One-dimensional theory of dislocation pile-ups suggests to seek the
minimizer in the form
n() =

n
1
() for (0. |).
n
n
for (|. c |).
n
1
(c ) for (c |. c).
(4.28)
88 CHAPTER 4. CRYSTAL PLASTICITY
where n
n
is a constant, | an unknown parameter, 0 | c,2, and n
1
(|) = n
n
at = |. We have to nd n
1
() and the constants, n
n
and |. Since n

0
for (0. |), the functional becomes
1 = 2

|
0
[
1
2
(1 n
1
)
2
+ :
(
n

1
+
1
2
n
2
1
)]
d +
1
2
(1 n
n
)
2
(c 2|). (4.29)
Function n
1
() is subject to the boundary conditions
n
1
(0) = 0. n
1
(|) = n
n
. (4.30)
Varying this energy functional with respect to n
1
() we obtain the Euler
equation for n
1
() on the interval (0. |)
1 n
1
+:n

1
= 0. (4.31)
The variation of (4.29) with respect to n
n
and | yields the two additional
boundary conditions at = |
n

1
(|) = 0. 2: = (1 n
n
)(c 2|). (4.32)
Condition (4.32)
1
means that the dislocation density must be continuous.
Equations (4.31), (4.30)
1
, and (4.32)
1
have the solution
n
1
() = 1 cosh

:
+ tanh
|

:
sinh

:
. 0 |. (4.33)
Equations (4.30)
2
and (4.32)
2
give the following transcendental equation to
determine | in terms of the constants : and c
)(|) 2| + 2:cosh
|

:
= c. (4.34)
According to (4.28) | must lie in the segment [0. c,2]. Since cosh(|,

:) 1,
2| c2:. Thus, equation (4.34) has no positive root if c < 2:. Returning
to the original variables according to (4.25) we see that inequality c < 2:
corresponds to the condition <
ca
, where

ca
=
2/
c/j
-
.
and for <
ca
no dislocations are nucleated. Note that the threshold
value,
ca
, is inversely proportional to the product of the size c of specimen
times the saturated dislocation density (a kind of Hall-Petch relation). For
c 2: equation (4.34) has only one root in the interval (0. c,2). Indeed,
4.3. ANTI-PLANE CONSTRAINED SHEAR 89
since )(0) = 2: < c and )(c,2) c, the curve = )(|) crosses the line
= c, so, there are roots of the equation (4.34) in (0. c,2). In fact, there is
only one root because )(|) is strictly increasing function of |.
Fig. 4.11 shows the evolution of ( r) (where r = r/j
-
) as increases. For
the numerical simulation we took / = 0.04, j
-
= 3 .10
14
m
2
, / = 4 .10
10
m,
c = 10
4
m, so that c = c/j
-
= 12.
2 4 6 8 10 12
0.01
0.02
0.03
0.04
0.05
0.06
x
b
a
b
c
Figure 4.11: Evolution of : a) = 0.01, b) = 0.05, c) = 0.07
It is interesting to plot the average shear stress
o =
1
c

o
0
j( (r)) dr (4.35)
as function of the shear strain. For <
ca
the plastic distortion = 0, so
o = j. For
ca
from (4.32)
2
and (4.33) we obtain
o
j
=
2/
c
+
2

/
c
tanh
|

:
. (4.36)
Fig. 4.12 shows the normalized average shear stress versus shear strain curve
OAB. There is a work hardening section AB for
ca
due to the second
term in (4.36) caused by the dislocation pile-up. Mention, however, that
there is no residual strain as we unload the crystal by decreasing : the
stress-strain curve follows the same path BAO, so the plastic deformation
is completely reversible, and no energy dissipation occurs. In the course
of unloading the dislocations nucleated annihilate, and as we approach the
point A they all disappear.
Consider now the functional (4.26) with the logarithmic energy of dislo-
cation network. Assuming n() as before in the form (4.28), we obtain for
n
1
() the nonlinear equation
1 n
1
+ :
n

1
(1 n

1
)
2
= 0. (4.37)
90 CHAPTER 4. CRYSTAL PLASTICITY
0.01 0.02 0.03 0.04 0.05
0.002
0.004
0.006
0.008
O
A
B
g
s/m
_
Figure 4.12: Normalized average stress versus shear strain curve
and the boundary conditions (4.30) and (4.32). Let j() be the function
j() =
:
1 n

1
.
Equation (4.37) and the boundary conditions can be written as the following
boundary-value problem
j

= n
1
1. n

1
= 1
:
j
. n
1
(0) = 0. n
1
(|) = n
n
. n

1
(|) = 0. (4.38)
Function j() is a decreasing function of . Since n

1
(|) = 0, j(|) = :.
The system of ordinary dierential equations (4.38) admits the rst inte-
gral
1
2
(1 n
1
)
2
+:ln j j = const.
The value of the constant can be found from the boundary conditions at the
point = |
1
2
(1 n
1
)
2
+ :ln j j =
1
2
(1 n
n
)
2
+:ln ::.
Hence
1 n
1
() =

(1 n
n
)
2
+ 2:(
j()
:
1 ln
j()
:
). (4.39)
The expression under the square root is positive because j() : and
r ln r+1 for r 1. For the function () = j(),:1 we have the initial
value problem

=
1
:

(1 n
n
)
2
+ 2:( ln(1 + )). (0) =
0
. (4.40)
4.3. ANTI-PLANE CONSTRAINED SHEAR 91
where, according to (4.39),
0
is related to n
n
by a transcendental equation
(1 n
n
)
2
+ 2:(
0
ln(1 +
0
)) = 1. (4.41)
Expressing n
n
in (4.41) by
0
and putting the result in (4.40), we obtain the
initial value problem

=
1
:

1 + 2:(
0
ln
1 +
1 +
0
). (0) =
0
. (4.42)
The length | is determined by the formula
| =

q
0
0
:d

1 + 2:(
0
ln((1 + ),(1 +
0
)))
.
Together with the condition (4.32)
2
, we get the following equation for
0
(
c 2

q
0
0
:d

1 + 2:(
0
ln((1 + ),(1 +
0
)))
)

1 + 2:(ln(
0
+ 1)
0
) = 2:.
(4.43)
Having found
0
from (4.43) we can solve the initial value problem (4.42) to
determine n
1
(). The solution of this problem shows that the minimizer of
(4.26) does not dier much from that of (4.27) for the values of the parameters
chosen. Fig. 4.13 shows the plot of n
1
() for (0. |) obtained by minimizing
(4.26) and (4.27) for = 0.01.
2.5 5 7.5 10 12.5 15 17.5
0.05
0.1
0.15
0.2
0.25
0.3
x
u
1
Figure 4.13: The graph of n
1
() for = 0.01: a) bold line: minimizer of
(4.26), b) dash line: minimizer of (4.27)
We turn now to the case when the resistance to the dislocation motion
(and hence the dissipation) cannot be neglected. In this case the plastic
92 CHAPTER 4. CRYSTAL PLASTICITY
distortion may evolve only if the yield condition (4.24) is satised. It is
convenient to divide both side of (4.24) by j to obtain
+ /

,aa
(j
-
/
,a
)
2
=
c
1,j. (4.44)
We regard again as a given function of time (the driving variable) and
try to determine (t. r). We consider the following close loading path: is
rst increased from zero to some value


c
, then decreased to
c
, and
nally increased to zero (Fig. 4.14). The rate of change of (t) does not aect
the results due to the rate independence of the dissipation. The problem is
to determine the evolution of as function of t and r, provided (0. r) = 0.
Since is initially zero, we see from (4.44) that = 0 as long as <
c
.
Thus, the dissipative threshold stress (the yield stress) o

= 1 in this case.
For
c
the yield condition
+ /

,aa
(j
-
/
,a
)
2
=
c
(4.45)
takes place everywhere in (0. c). Equation (4.45) follows from (4.44) and the
initial condition (0. r) = 0, since for small (t. r) and
c

+ /

,aa
(j
-
/
,a
)
2

= +/

,aa
(j
-|
/
,a
)
2
.
It is convenient to introduce the deviation of (t) from the critical shear,
c
,

=
c
0, and to dene the following dimensionless quantities
=
r/j
-

. n() =
(r)

. : =
/

. (4.46)
They are similar to those of (4.25), with

replacing . Equation (4.45)


takes the dimensionless form
1 n + :
n

(1 n

)
2
= 0. (4.47)
One-dimensional theory of dislocation pile-ups with non-zero resistance
suggests that the solution of (4.47) is symmetric
n() =
{
n
1
() for (0. c,2).
n
1
(c ) for (c,2. c).
(4.48)
where c = c/j
-
,

. Function n
1
() is determined from equation (4.37) and
the boundary conditions
n
1
(0) = 0. n

1
(c,2) = 0. (4.49)
4.3. ANTI-PLANE CONSTRAINED SHEAR 93
t
g
g
*
-g
c
Figure 4.14: A close loading path
The rst condition means that dislocations cannot reach the boundary of the
region because the boundary is clamped. The second condition follows from
the continuity of plastic distortion and the symmetry property (4.48).
The boundary-value problem (4.37), (4.49) can be solved in exactly the
same manner as for the case without dissipation. For the function () =
j(),: 1 we have the initial value problem (4.42). Integrating equation
(4.42) over from zero to c,2 and taking into account the condition (c,2) =
0 which is the consequence of (4.49)
2
, we get
p(
0
)

q
0
0
:d

1 + 2:(
0
ln((1 + ),(1 +
0
)))
= c,2. (4.50)
Thus,
0
is the root of the transcendental equation (4.50). Note that, as long
as
0
is smaller than the root of the equation

0
ln(1 +
0
) =
1
2:
=

2

2/
. (4.51)
the expression under the square root of integral (4.50) is positive and the
integral is well dened. This follows from the inequalities 1 2:(
0
ln(1 +

0
)) 0 for
0
smaller than the root of (4.51) (because 1 2:(r ln(1 +
r)) is strictly decreasing function of r) and ln(1 + ) 0 for 0.
Dierentiating p(
0
) with respect to
0
, it is easy to check that dp,d
0
0,
so p(
0
) is a strictly increasing function. Since p(0) = 0 and p(
0
) as

0
approaches the root of (4.51), equation (4.50) has only one root which is
smaller than the root of (4.51).
Numerical calculations show that the root of (4.50) is very close to that
of (4.51). Therefore we may nd n
1
() approximately by solving the initial
value problem on (0. c,2)

2
:
( ln(1 + )). (0) =
0
.
94 CHAPTER 4. CRYSTAL PLASTICITY
where
0
is the root of the equation (4.51). The length | of the boundary
layer may be dened by the formula
| =

:
2

q
0
c
d

ln(1 + )
.
where o is a small positive number. We take o = 10
5
.
Fig. 4.15 shows the evolution of ( r) (where r = r/j
-
) as

increases.
For the numerical simulation we took as before / = 0.04, c = c/j
-
= 12.
One can see that the dislocation density as well as the length of boundary
layers increase as

increases. Note that the expression


+ /

,aa
(j
-
/
,a
)
2
=

((t)
c
)
does not depend on r at any instant.
2 4 6 8 10 12
0.0005
0.001
0.0015
0.002
0.0025
0.003
x
b
a
b
c
Figure 4.15: Evolution of : a)

= 0.0001, b)

= 0.001, c)

= 0.003
After reaching some value of shear strain


c
we unload the crystal
by decreasing . Since becomes smaller than 1, does not change ( =

(r)) until

+ /

,aa
(j
-
/

,a
)
2
=
c
. (4.52)
where

(r) is the solution of equation (4.45) for (t) =

+/

,aa
(j
-
/

,a
)
2
=
c
. (4.53)
From (4.52) and (4.53) one can see that the ow begins when (

c
) =

c
, i.e. for =

2
c
. From that value of the yield condition
4.3. ANTI-PLANE CONSTRAINED SHEAR 95
= 1 takes place leading to the decrease of which should be now
determined by the equation
+ /

,aa
(j
-
/
,a
)
2
=
c
. (4.54)
Since for (
c
.

) the deviation
|
= +
c
is positive, equation (4.54)
can again be transformed to equation (4.47) and solved in exactly the same
manner if we replace

= (t)
c
in all formulas (4.46)(4.51) by
|
=
(t) +
c
. As approaches
c
, tends to zero because
|
0. The further
increase of from
c
to zero does not cause change in which remains
identically zero.
0.5 1 1.5
0.002
0.004
0.006
0.008
0.01
0.012
0.014
x
a
a
b
c
Figure 4.16: Evolution of the normalized dislocation density: a)

= 0.0001,
b)

= 0.001, c)

= 0.003
The dislocation density c =
,a
can be calculated from the solution (4.48).
In terms of the dimensionless variable (4.46) we have
c() =

/j
-
(),(1 + ()) for (0. |).
0 for (|. c |).
/j
-
(c ),(1 + (c )) for (c |. c).
Since () is decreasing, the maximum dislocation density is achieved at
= 0 giving c
max
= /j
-

0
,(1 +
0
). So, the parameter
0
is simply linked
to the maximum dislocation density. Fig. 4.16 shows the distributions of
the normalized dislocation density c( r),/j
-
within the interval (0.

|) (where

| = |

) as

changes.
As soon as the plastic deformation develops, the shear stress o = j()
becomes inhomogeneous. It is interesting to calculate the average shear stress
(4.35) which is a measurable quantity. During the loading, according to
96 CHAPTER 4. CRYSTAL PLASTICITY
(4.48), we have for the normalized average shear stress (or the average elastic
shear strain)
-
(c)
=
o
j
=
1
c
[

c 2

|
0
(1 n(r)) dr
]
=
c
+ 2

2

c/j
-

|
0
(1 n()) d.
The integral is evaluated using equation (4.38)

|
0
(1 n()) d =

|
0
j

() d = (j(|) j(0)) = :
0
.
Thus
-
(c)
=
c
+
2/
c/j
-

0
. (4.55)
where the second term causing hardening depends on the maximum dislo-
cation density and is inversely proportional to the product of the size c of
specimen times the saturated dislocation density. Formula (4.55) describes
the size eect in this model.
During the inverse loading, when the yield condition = 1 holds true,
formula (4.55) changes to
-
(c)
=
c
+
2/
c/j
-

0
.
where
0
is the root of the equation

0
ln(1 +
0
) =

2
|
2/
.
0.005 0.01 0.015 0.02
-0.004
-0.002
0.002
0.004
g
g
g
g
-g
s/m
-
O
A
B
C
D
c c
*
*
Figure 4.17: Normalized average stress versus shear strain curve
Fig. 4.17 shows the normalized average shear stress (or average elastic
shear strain) versus shear strain curve for the loading program of Fig. 4.14.
4.4. PLANE CONSTRAINED SHEAR 97
We took c/j
-
= 12, / = 0.04,
c
= 0.004,

= 0.02. The straight line


OA corresponds to the purely elastic loading with increasing from zero to

c
. The line AB corresponds to the plastic yielding with = 1. The yield
begins at the point A with the yield stress o

= 1. The work hardening


due to the dislocation pile-up is observed which is described by the second
term in (4.55). During the unloading as decreases from

to

2
c
(the line BC) the plastic distortion =

is frozen. As decreases further


from

to
c
, the plastic yielding occurs with = 1 (the line CD). The
yield stress o

= j

21 at the point C, at which the inverse plastic ow


sets on, is larger than 1 (because


c
1,j). Along the line CD,
as is decreased, the created dislocations annihilate, and at the point D all
dislocations disappear. Finally, as increases from
c
to zero, the crystal
behaves elastically with = 0. In this close cycle ABCD dissipation occurs
only on the lines AB and CD. It is interesting that the lines DA and BC
are parallel and have the same length. In phenomenological plasticity theory
this property is modelled as the translational shift of the yield surface in the
stress space, the so-called Bauschinger eect.
4.4 Plane constrained shear
0
y
x
h
gh
j
s
m
a
L
z
Figure 4.18: Plane constrained shear
As the next boundary-value problem we consider a strip made of a single
crystal undergoing a plane strain shear deformation (see Fig. 4.18). Let the
cross-section of the strip be a rectangle of width c and height , 0 r c,
0 . We realize the shear deformation by placing the strip in a hard
98 CHAPTER 4. CRYSTAL PLASTICITY
device with the prescribed displacements at its upper and lower sides
n(0) = 0. (0) = 0. n() = . () = 0. (4.56)
where n() and () are the longitudinal and transverse displacements, re-
spectively, with being the overall shear strain. We assume that the length
of the strip 1 is large, and the width c is much greater than the height
(1 c ) to neglect the end eects and to have the stresses and strains
depending only on one variable in the central part of the strip.
For the plane strain state the components of the strain tensor are
-
aa
= 0. -
a
= -
a
=
1
2
n
,
. -

=
,
. (4.57)
If the shear strain is suciently small, then the crystal deforms elasti-
cally and n = , = 0 everywhere in the strip. If exceeds some critical
threshold, then edge dislocations may appear. We admit only the slip direc-
tions (or the directions of the Burgers vectors) perpendicular to the .-axis
and inclined at an angle , with the r-axis and the dislocation lines paral-
lel to the .-axis. Since only one slip system is active, the plastic distortion
is given by
i)
= :
i
:
)
, with : = (cos ,. sin ,. 0) being the slip direction,
and : = (sin ,. cos ,. 0) the normal vector to the slip plane. We assume
that depends only on : = () (translational invariance). Because of
the prescribed boundary conditions (4.56), dislocations cannot penetrate the
boundaries = 0 and = , therefore
(0) = () = 0. (4.58)
The in-plane components of the plastic strain tensor are
-
j
aa
=
1
2
sin 2,. -
j
a
=
1
2
cos 2,. -
j

=
1
2
sin 2,.
With these total and plastic strain tensors we obtain the in-plane components
of the elastic strain tensor
-
c
aa
=
1
2
sin 2,. -
c
a
=
1
2
(n
,
cos 2,). -
c

=
,

1
2
sin 2,.
As depends only on , there are two non-zero components of Nyes dislo-
cation density tensor (4.1), namely, c
a:
=
,
sin ,cos , and c
:
=
,
sin
2
,.
Thus, the resultant Burgers vector of all dislocations whose dislocation lines
cut the area perpendicular to the .-axis is parallel to the slip direction : and
the scalar dislocation density equals
j =
1
/

c
2
a:
+c
2
:
=
1
/

,
sin ,.
4.4. PLANE CONSTRAINED SHEAR 99
Assuming for simplicity the isotropic elastic property of the crystal, we
write the free energy per unit volume of the crystal with dislocations as
o(-
c
i)
. c
i)
) =
1
2
`(-
c
ii
)
2
+ j-
c
i)
-
c
i)
+ j/ ln
1
1
o
,
sin ,
oj
s
. (4.59)
Thus, the total energy functional becomes
[n. . ] = c1


0
[
1
2
`
2
,
+
1
2
j(n
,
cos 2,)
2
+
1
4
j
2
sin
2
2, (4.60)
+j(
,

1
2
sin 2,)
2
+j/ ln
1
1
o
,
sin ,
oj
s
]
d.
Functional (4.60) can be reduced to a functional depending on () only.
Indeed, by rst xing () and taking the variation of (4.60) with respect to
n and we derive the equilibrium equations
j(n
,

,
cos 2,) = 0.
(` + 2j)
,
j
,
sin 2, = 0.
Integrating these equations and using the boundary conditions (4.56) we get
n
,
= + ( ) cos 2,.

,
= i( ) sin 2,.
(4.61)
where i =
j
A+2j
, and =
1

0
d. Substituting (4.61) into (4.60) and
collecting common terms we obtain the energy functional in terms of
[] =c1


0
j
[
1
2
(1 i)
2
sin
2
2, +
1
2
i
2
sin
2
2, (4.62)
+
1
2
( cos 2,)
2
+ / ln
1
1
o
,
sin ,
oj
s
]
d.
If the dissipation is negligible, then the plastic distortion minimizes (4.62)
under the constraint (4.58). The overall shear strain is regarded as given
function of time (control parameter), so one can study the evolution of the
dislocation network which accompanies the change of .
If the resistance to dislocation motion cannot be neglected, then the
energy minimization must be replaced by a ow rule. In case of rate-
independent plasticity, the ow rule for

= 0 reads
1

=
o

o
o
. (4.63)
100 CHAPTER 4. CRYSTAL PLASTICITY
where the dissipation potential for crystals deforming in single slip is
1 = 1

.
with 1 being a positive constant called critical resolved shear stress. The
right hand side of (4.63) is the negative variational derivative of the energy
with respect to at xed overall strain

o

o
o
=
o

,
.
For

= 0, the evolution equation (4.63) does not have to be satised: It is
replaced by the equation

= 0.
For small up to moderate dislocation densities the logarithmic term in
(4.62) may be approximated by the formula
ln
1
1
j
j
s

=
j
j
-
+
1
2
j
2
j
2
-
. (4.64)
so that
1() =c1


0
j
[
1
2
(1 i)
2
sin
2
2, +
1
2
i
2
sin
2
2, (4.65)
+
1
2
( cos 2,)
2
+ /
(

,
sin ,
/j
-
+
1
2

2
,
sin
2
,
(/j
-
)
2
)]
d.
For simplicity of the analysis, we shall further deal with this functional only.
In the case of zero resistance (and hence the energy dissipation is zero) the
determination of () reduces to the minimization of the total energy (4.65).
Since l is convex with respect to and
,
, the variational problem has a
unique solution. It is convenient to introduce the dimensionless quantities
1 =
/j
-
c1j
. = /j
-
.

= /j
-
. (4.66)
The dimensionless variable changes on the interval (0.

). Functional (4.65)
reduces to
1[] =


0
[
1
2
(1 i)
2
sin
2
2, +
1
2
i
2
sin
2
2, (4.67)
+
1
2
( cos 2,)
2
+/

sin , +
1
2
/
2
sin
2
,
]
d.
4.4. PLANE CONSTRAINED SHEAR 101
where the prime denotes dierentiation with respect to , and, for short,
the bars over and are dropped. We minimize functional (4.67) among
functions satisfying the boundary conditions
(0) = () = 0. (4.68)
For the variational problem of this type, there exists a threshold value
ca
such that when <
ca
no dislocations are nucleated and = 0. Near the
threshold value the dislocation density must be small so that the last term
in (4.67) can be neglected. Besides, the width of the boundary layer tends
to zero as
ca
. This gives us the idea of nding the threshold value by
employing a minimizing sequence of the form
=

o
r
c
. for (0. c).

n
. for (c. c).
o
r
c
( ). for ( c. ).
(4.69)
where
n
is an unknown constant, and c is a small unknown length which
tends to zero as
ca
. Substituting (4.69) into the energy functional
(4.67) (with the last term being removed) and neglecting all small terms of
order c and higher, we obtain
1(
n
) =
1
2
[(
n
cos 2,)
2
+
2
n
sin
2
2,] + 2/
n
sin , . (4.70)
A rather simple analysis shows that the minimum of (4.70) is achieved at

n
= 0 if and only if

ca
=
2/ sin ,
cos 2,
.
otherwise it is achieved at
n
= 0 (no dislocations are nucleated). Note that
the sign of
n
depends on the angle ,:
n
is positive if 0

< , < 45

and
negative if 45

< , < 90

. In terms of the original length the energetic


threshold value reads

ca
=
2/
/j
-
sin ,
cos 2,
. (4.71)
showing clearly the size eect. Equation (4.71) deviates from the well-known
Hall-Petch relation. The reason for this deviation can be explained by the
boundary conditions (4.58) which do not permit the penetration of disloca-
tions through the grain boundaries.
Due to the boundary conditions,

should change its sign on the interval


(0. ). The one-dimensional theory of dislocation pile-ups as well as the
102 CHAPTER 4. CRYSTAL PLASTICITY
solution of the previous problem suggest to seek the minimizer in the form
() =

1
(). for (0. |).

n
. for (|. |).

1
( ). for ( |. ).
(4.72)
where
n
is a constant, | an unknown length, 0 |

2
, and
1
(|) =
n
.
We have to nd
1
() and the constants,
n
and |. With from (4.72) the
total energy functional becomes
1 = 2

|
0
[
1
2
(1 i)
2
1
sin
2
2, + /

1
sin , +
1
2
/
2
1
sin
2
,
]
d
+
1
2
(1 i)
2
n
sin
2
2,( 2|) +
1
2
[i
2
sin
2
2, + ( cos 2,)
2
].
(4.73)
where
=
1

(
2

|
0

1
d + ( 2|)
n
)
. (4.74)
Varying the energy functional (4.73) with respect to
1
we obtain
/

1
sin
2
,+(1i)
1
sin
2
2,+(cos
2
2,+isin
2
2,) cos 2, = 0. (4.75)
where
1
() is subject to the boundary conditions

1
(0) = 0.
1
(|) =
n
. (4.76)
The variation of (4.73) with respect to | gives an additional boundary con-
dition at = |,

1
(|) = 0. (4.77)
which means that the dislocation density must be continuous. Varying the
energy functional with respect to
n
, we obtain a condition for
n
,
2/ sin ,(sign

1
) + [(cos
2
2, + isin
2
2,) cos 2,
+ (1 i)
n
sin
2
2,]( 2|) = 0. (4.78)
Equations (4.75), (4.76)
1
, and (4.77) yield the solution

1
=
1j
(1 cosh j + tanh j| sinh j) . 0 | (4.79)
with

1j
=
cos 2, (cos
2
2, + isin
2
2,)
(1 i) sin
2
2,
. (4.80)
4.4. PLANE CONSTRAINED SHEAR 103
and
j = 2

1 i
/
cos ,.
With (4.74), (4.76)
2
, (4.79), and (4.80) we obtain the average of
=
cos 2,
[
2
(
|
tanh j|
j
)
+
(
1
1
cosh j|
)
( 2|)
]
p(|)
. (4.81)
where
p(|) = (1 i) sin
2
2, + (cos
2
2, + isin
2
2,)

[
2
(
|
tanh j|
j
)
+
(
1
1
cosh j|
)
( 2|)
]
.
and

n
=
cos 2, (cos
2
2, +isin
2
2,)
(1 i) sin
2
2,
(
1
1
cosh j|
)
. (4.82)
Substitution of (4.82) into (4.78) gives the following equation to determine |:
)(|) 2/ sin ,(sign

1
)
cos 2, (cos
2
2, + isin
2
2,)
cosh j|
( 2|) = 0.
j=30
j=60
a
b
c
d
e
f
b
y
-
Figure 4.19: Evolution of for single-slip constrained shear at zero dissipa-
tion: a,d) = 0.0068, b,e) = 0.0118, c,f) = 0.0168
Fig. 4.19 shows the evolution of ( ) as increases, for , = 30

(contin-
uous lines) and , = 60

(dashed lines), where = /j


-
. For the numerical
simulation we took the material parameters from Table 4.1. All material
parameters used in this Chapter (except /, j
-
, and which is responsible
104 CHAPTER 4. CRYSTAL PLASTICITY
for the cross-slip interaction) are well-known for aluminum. We choose these
additional parameters to have good agreement of discrete dislocation simu-
lations and the continuum dislocation theory with respect to the yield stress
and the hardening rate for both single and double slip (see also Section 4.3).
In all numerical simulations we take = 10
6
m so that

= /j
-
= 0.349.
It is interesting to plot the shear stress t = j( cos 2,) as a function
of the shear strain. As we know, for <
ca
no dislocations are nucleated
and = 0, so the shear stress t = j. For
ca
we take from (4.81)
to compute the shear stress.
A
A
B
B
t/m
g
0
j=30
j=60
Figure 4.20: Normalized shear stress versus shear strain curve for single-slip
constrained shear at zero dissipation
Fig. 4.20 shows the normalized shear stress versus shear strain curve OAB
for , = 30

and OAB for , = 60

. There is a work hardening section


AB for
ca
caused by the dislocation pile-up. Note, however, that
there is no residual strain as we unload the crystal by decreasing : the
stress-strain curve follows the same path BAO, so the plastic deformation
is completely reversible and no energy dissipation occurs. In the course of
unloading the dislocations nucleated annihilate, and as we approach point A
they all disappear.
If the resistance to dislocation motion (and hence the dissipation) cannot
be neglected, the plastic distortion may evolve only if the yield condition
= 1 is fullled. If < 1, then is frozen, the dislocation density
remains unchanged and the crystal deforms elastically. Computing the vari-
Material j (GPa) i / (

A) j
-
(m
2
) /
Aluminum 26.3 0.33 2.5 1.396 10
15
0.000115 0.576
Table 4.1: Material characteristics
4.4. PLANE CONSTRAINED SHEAR 105
ational derivative of (4.65) we derive from (4.63) the yield condition
j

,
sin
2
,
/
2
j
2
-
(1 i) sin
2
2, (cos
2
2, + isin
2
2,) + cos 2,

= 1.
Consider rst the case , < 45

. We divide this equation by j and introduce


the dimensionless variable = /j
-
to transform the yield condition to

sin
2
, (1 i) sin
2
2, (cos
2
2, +isin
2
2,) + cos 2,

=
c
cos 2,. (4.83)
with
c
1,jcos 2, and the prime denoting the derivative with respect to
. We shall further omit the bar over for short.
We regard as a given function of time (the driving variable) and try
to determine (t. ). We consider the loading path similar to that shown in
Fig. 4.14. The problem is to determine the evolution of as a function of t
and , provided (0. ) = 0 and , < 45

.
Since the plastic distortion, , is initially zero, we see from (4.83) that
= 0 as long as <
c
. Thus, the dissipative threshold stress (the yield
stress) t

= 1, cos 2, in this case. For small (t. r) and


c
, the yield
condition becomes
/

sin
2
,(1i) sin
2
2,(cos
2
2,+isin
2
2,) + cos 2, =
c
cos 2,.
(4.84)
Let us introduce the deviation of (t) from the critical shear
c
,

=
c
,
and simplify (4.84) to obtain
/

sin
2
, (1 i) sin
2
2, (cos
2
2, + isin
2
2,) +

cos 2, = 0.
Since this equation is linear, is proportional to

such that =

1
,
where
1
is the solution of the equation
/

1
sin
2
,(1i)
1
sin
2
2,(cos
2
2,+isin
2
2,)
1
+cos 2, = 0. (4.85)
The analogous problem of anti-plane constrained shear at nonzero resis-
tance suggests that the solution of (4.85) is symmetric, i.e.

1
() =
1
( ) for (,2. ). (4.86)
Function
1
() is determined from equation (4.85) and the boundary condi-
tions

1
(0) = 0.

1
(,2) = 0. (4.87)
106 CHAPTER 4. CRYSTAL PLASTICITY
The rst condition means that dislocations cannot reach the boundary of the
region because of the prescribed displacement. The second condition follows
from the continuity of plastic distortion and the symmetry property (4.86).
Equations (4.85) and (4.87) admit the solution

1
=
1j
(
1 cosh j + tanh j

2
sinh j
)
. 0

2
. (4.88)
with

1j
=
cos 2, (cos
2
2, + isin
2
2,)
1

(1 i) sin
2
2,
. (4.89)
and
j = 2

1 i
/
cos ,. (4.90)
The average of
1
is obtained in the form

1
=
cos 2,
(
1
2 tanhj

2
j
)
(1 i) sin
2
2, + (cos
2
2, + isin
2
2,)
(
1
2 tanh j

2
j
). (4.91)
j=30
j=60
y
_
b
1
Figure 4.21: Graphs of
1
( ) for single-slip constrained shear at non-zero
dissipation
Fig. 4.21 shows the graphs of
1
( ) for , = 30

(continuous line) and


, = 60

(dashed line). For the numerical simulation we took the material


parameters from Table 4.1, and = 1jm, so that

= /j
-
= 0.349.
After reaching


c
, we unload the crystal by decreasing . Since
becomes smaller than 1, does not change ( =

()) until
/

sin
2
, (1 i)

sin
2
2, (cos
2
2, + isin
2
2,)

+ cos 2, =
c
cos 2,.
(4.92)
4.4. PLANE CONSTRAINED SHEAR 107
where

() is the solution of (4.84) for (t) =

. From (4.92) we can see that


the plastic ow begins when (

c
) =
c
, i.e. for =

2
c
.
From that value of , the yield condition = 1 holds leading to a decrease
of which should now be determined from
/

sin
2
,(1i) sin
2
2,(cos
2
2,+isin
2
2,)+ cos 2, =
c
cos 2,.
(4.93)
Since for (
c
.

), the deviation
|
= +
c
is positive, equation (4.93)
can again be transformed to equation (4.85) and solved in exactly the same
manner if we replace

=
c
in all formulas (4.88)-(4.91) by
|
= +
c
.
As approaches
c
, tends to zero because
|
0. The further increase
of from
c
to zero does not cause change in which remains zero.
It is not dicult to modify the construction given above to nd the solu-
tion for , 45

.
j=30
j=60
y
_
a
1
Figure 4.22: Graphs of c
1
( ) for single-slip constrained shear at non-zero
dissipation
The normalized dislocation density c =
,
sin , can be calculated from
the solution (4.88). Since () is proportional to

, c() is also proportional


to

such that c() =

c
1
(). For (0. ,2) we have
c
1
() =
1j
(
j sinh j + j cosh j tanh
j
2
)
sin ,.
with
1j
from (4.89) and j from (4.90). For (,2. ) we have c
1
() =
c
1
( ) due to symmetry. Fig. 4.22 shows the graphs of c
1
for , = 30

(continuous line) and , = 60

(dashed line) for (0.

,2).
It is interesting to calculate the shear stress t which is a measurable
quantity. During loading, we have for the normalized shear stress (or the
elastic shear strain)

c
=
t
j
=
c
+
(

(
1
2 tanh
j
2
j
)

1j
cos 2,
)
. (4.94)
108 CHAPTER 4. CRYSTAL PLASTICITY
t/t
0
g
h/d=1.15
h/d=2.3
h/d=80
withdissipation
energyminimization
Figure 4.23: The normalized shear stress versus shear strain curve
with
1j
from (4.89). The second term of (4.94) causes hardening due to the
dislocation pile-ups. Equation (4.94) describes the size eect in this model.
y
_
h
u,
y
h/d=2.3
h/d=80
energyminimization
withdissipation
g=0.0068
g=0.0118
g=0.0168
g=0.0218
Figure 4.24: Comparison of the total shear strain prole obtained from this
approach and from [6]
Fig. 4.23 shows a comparison of the stress-strain curves during loading
obtained from energy minimization, from (4.94), and from the discrete dis-
location simulations reported in [6]. In order to compare with the discrete
dislocation simulations we took , = 60

, t
0
= 1.9 10
3
j and let all other ma-
terial constants remain the same as in previous simulations. The stress-strain
curves from the discrete dislocation simulations are provided for three dier-
ent ratios ,d, where d is the spacing between the active slip planes. Both
4.4. PLANE CONSTRAINED SHEAR 109
curves obtained from energy minimization as well as from the ow rule nearly
coincide and show good agreement with the discrete dislocation simulations
for ,d = 80.
Fig. 4.24 shows a comparison of the total shear strain proles obtained
from energy minimization, from the ow rule, and from the discrete dislo-
cation simulations reported in [6]. Both proles obtained from energy mini-
mization and from the ow rule again show good agreement with the discrete
dislocation simulations for ,d = 80.
During inverse loading, when the yield condition = 1 holds true,
equation (4.94) changes to
t
j
=
c
+
(

(
1
2 tanh
j
2
j
)

1j
cos 2,
)
.
with the deviation
|
= +
c
used instead of

=
c
for the average
of .
A
B
C
D
0
g
*
g
t/m
-g
cr
g
cr
g
*
Figure 4.25: Normalized shear stress versus shear strain curve for single-slip
constrained shear at non-zero dissipation, for , = 60

Fig. 4.25 shows the normalized shear stress (or elastic shear strain) ver-
sus shear strain curve for the loading path of Fig. 4.14, with

= 0.006,
, = 60

, while all other parameters remain the same. The straight line OA
corresponds to purely elastic loading with increasing from zero to
c
. Line
AB corresponds to plastic yielding with = 1. Yielding begins at point A
with the yield stress t

= 1, cos 2,, and we can observe the work hard-


ening due to the dislocation pile-up, which is described by the second term
of (4.94). During unloading, as decreases from

to

2
c
(line
BC), the plastic distortion =

is frozen. As decreases further from

to
c
, plastic yielding occurs with = 1 (line CD). From Fig. 4.25
it becomes obvious that the yield stress t

= t

+ 21, cos 2, at point C,


at which the inverse plastic ow sets on, is larger than 1, cos 2, (because
110 CHAPTER 4. CRYSTAL PLASTICITY
t

1, cos 2,). Along line CD, as is decreased, the created dislocations


annihilate, and at point D all dislocations have disappeared. Finally, as
increases from
c
to zero, the crystal behaves elastically with = 0. In
this closed cycle OABCDO dissipation occurs only along lines AB and CD.
It is interesting that lines DA and BC are parallel and have the same length
exhibiting again the Bauschinger eect.
4.5 Single crystals deforming in double slip
j
r
j
l
y
h
0
z
x a
L
gh
s
l
m
l
m
r
s
r
Figure 4.26: Plane constrained shear of single crystals deforming in double
slip
Consider the same problem as in the previous Section, but now for crystals
deforming in double slip. We admit two slip directions (or the directions of
the Burgers vectors) perpendicular to the .-axis and inclined at an angle
,
|
(0 ,
|
,2) and ,

(,2 ,

) with the r-axis, respectively,


and the dislocation lines parallel to the .-axis. Since two slip systems are
active, the plastic distortion is given by
i)
=
|
:
|
i
:
|
)
+

i
:

)
, with :

i
=
(cos ,

. sin ,

. 0) being the slip directions, and :

)
= (sin ,

. cos ,

. 0) the
normal vectors to the slip planes ( = |. :).
We assume that
|
and

depend on only:
|
=
|
() and

()
(translational invariance). Because of the prescribed boundary conditions
dislocations cannot penetrate the boundaries = 0 and = , therefore

|
(0) =

(0) =
|
() =

() = 0. (4.95)
As previously, the boundaries = 0 and = serve as obstacles to disloca-
tion motion.
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 111
Under plane strain state conditions, the non-zero components of the plas-
tic strain tensor are
-
j
aa
=
1
2
(
|
sin 2,
|
+

sin 2,

).
-
j
a
=
1
2
(
|
cos 2,
|
+

cos 2,

).
-
j

=
1
2
(
|
sin 2,
|
+

sin 2,

).
(4.96)
With (4.57) and (4.96) we obtain the non-zero components of the elastic
strain tensor
-
c
aa
=
1
2
(
|
sin 2,
|
+

sin 2,

).
-
c
a
=
1
2
(n
,

|
cos 2,
|

cos 2,

).
-
c

=
,

1
2
(
|
sin 2,
|
+

sin 2,

).
(4.97)
As
|
and

depend only on , there are two non-zero components of


Nyes dislocation density tensor
c
a:
=
|,
sin ,
|
cos ,
|
+
,
sin ,

cos ,

.
c
:
=
|,
sin
2
,
|
+
,
sin
2
,

.
These are the components of the resultant Burgers vector of all edge disloca-
tions whose dislocation lines cut the area perpendicular to the .-axis. Thus,
the dislocations produced on two slip systems belong to two dierent groups:
the rst one with the resultant Burgers vector showing in the direction :
|
i
,
the second one with the resultant Burgers vector parallel to :

i
. Therefore
the scalar dislocation densities (or the numbers of dislocations per unit area)
equal
j
|
=
1
/

|,
sin ,
|
. j

=
1
/

,
sin ,

. (4.98)
The free energy per unit volume of the crystal with dislocations takes the
form
o(-
c
i)
. c
i)
) =
1
2
`(-
c
ii
)
2
+ j-
c
i)
-
c
i)
+ j/
(
ln
1
1
j
|
j
s
+ ln
1
1
j
r
j
s
+
j
|
j

j
2
-
)
.
(4.99)
The last term in parentheses corresponds to the energy of the dislocation
network which consists of energies of each group of dislocations plus the
112 CHAPTER 4. CRYSTAL PLASTICITY
energy of cross-slip interaction, with being the interaction factor. With
(4.97) and (4.99), the total energy functional becomes
[n. .
|
.

] = c1


0
[
1
2
`
2
,
+
1
4
j(
|
sin 2,
|
+

sin 2,

)
2
+
1
2
j(n
,

|
cos 2,
|

cos 2,

)
2
+ j(
,

1
2

|
sin 2,
|

1
2

sin 2,

)
2
+ j/
(
ln
1
1
j
|
j
s
+ ln
1
1
j
r
j
s
+
j
|
j

j
2
-
)
]
d. (4.100)
Functional (4.100) can be reduced, in exactly the same manner as in the
previous case, to a functional depending on
|
() and

() only. The result


is
[
|
.

] = c1


0
j
[
1
2
(1 i)(
|
sin 2,
|
+

sin 2,

)
2
+
1
2
i(
|
sin 2,
|
+

sin 2,

)
2
+
1
2
(
|
cos 2,
|

cos 2,

)
2
+ /
(
ln
1
1
j
|
j
s
+ ln
1
1
j
r
j
s
+
j
|
j

j
2
-
)
]
d. (4.101)
Using (4.64) for small up to moderate dislocation densities we reduce (4.101)
to
[
|
.

] = c1


0
j
[
1
2
(1 i)(
|
sin 2,
|
+

sin 2,

)
2
+
1
2
i(
|
sin 2,
|
+

sin 2,

)
2
+
1
2
(
|
cos 2,
|

cos 2,

)
2
+ /
(
j
|
j
-
+
1
2
(
j

j
-
)
2
+
j

j
-
+
1
2
(
j

j
-
)
2
+
j
|
j

j
2
-
)
]
d. (4.102)
If the dissipation can be neglected, then
|
and

minimize (4.102) under


the constraints (4.95).
If the resistance to dislocation motion cannot be ignored, then the energy
minimization must be replaced by the ow rules. For

|
= 0 and

= 0 we
have
1

|
=
o

o
o
|
.
1

=
o

o
o

. (4.103)
where the dissipation potential is
1 = 1
|

|
+ 1

.
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 113
with 1
|
and 1

being positive constants called critical resolved shear stresses


of the corresponding slip systems. The right hand sides of (4.103) are the
negative variational derivatives of the energy with respect to
|
and

, re-
spectively, at xed overall shear strain

o
o

=
o

,
. = |. :.
The corresponding evolution equation (4.103) does not have to be fullled
for

|
= 0 or

= 0. It is replaced by the equation


|
= 0 or

= 0.
If the resistance to dislocation motion can be neglected (and hence the
energy dissipation is zero) the determination of
|
() and

() reduces to
the minimization of the total energy (4.102). We analyze this variational
problem rst in the special case ,

= ,
|
= , which corresponds
to symmetric double slip. It is convenient to introduce the dimensionless
variables (4.66), in terms of which the functional (4.102) reads
1[
|
.

] =


0
[
1
2
(1 i)(
|

)
2
sin
2
2, +
1
2
i(
|

)
2
sin
2
2,
+
1
2
( (
|
+

) cos 2,)
2
+ / sin ,(

|
+

)
+
1
2
/ sin
2
,(
2
|
+
2

+ 2

)
]
d. (4.104)
where the prime denotes dierentiation with respect to , and, for short,
the bars over and are dropped. We minimize functional (4.104) among

satisfying the boundary conditions (4.95). To guarantee existence and


uniqueness of the minimizer we must ensure convexity of the free energy
density o with respect to

. For this purpose let us consider the matrix


(
o
,o

|
o

|
o
,o

|
o

r
o
,o

r
o

|
o
,o

r
o

r
)
= / sin
2
,
(
1 sign

|
sign

sign

|
sign

1
)
It is easy to see that this matrix is positive denite (and, consequently,
the energy is convex with respect to

) if , = 0 and < 1. For =


1 the determinant of the matrix vanishes and there exists an eigenvector
corresponding to a zero eigenvalue. Thus, for = 1 the energy is no longer
strictly convex and one may expect non-uniqueness of the minimizer as well
as some numerical instability. In order to avoid this deciency we will assume
that < 1.
Similarly to the single slip problem, there exists a threshold value
ca
such that when <
ca
no dislocations are nucleated and
|
=

= 0. In
114 CHAPTER 4. CRYSTAL PLASTICITY
order to nd the threshold value, we employ minimizing sequences of the
form

|
=

o
|r
c
. for (0. c).

|n
. for (c. c).
o
|r
c
( ). for ( c. ).
(4.105)
and

o
rr
c
. for (0. c).

n
. for (c. c).
o
rr
c
( ). for ( c. ).
(4.106)
where
|n
and
n
are unknown constants, and c is a small unknown length
which tends to zero as
ca
.
Substituting (4.105) and (4.106) into the energy functional (4.104) (with
the last term being removed) and neglecting all small terms of order c and
higher, we obtain the function of two variables
1(
|n
.
n
) =
1
2

(
(
|n

n
)
2
sin
2
2, + ( (
|n
+
n
) cos 2,)
2
)
(4.107)
+ 2/ sin ,(
|n
+
n
).
The partial derivatives of (4.107) with respect to
|n
and
n
must vanish
giving
1

|n
= sin
2
2,(
|n

n
) cos 2,( (
|n
+
n
) cos 2,)
+ 2/ sin ,sign
|n
= 0.
and
1

n
= sin
2
2,(
|n

n
) cos 2,( (
|n
+
n
) cos 2,)
+ 2/ sin ,sign
n
= 0.
It is easy to see that these equations imply
|n
=
n
=
n
. Then, a rather
simple analysis shows that the minimum of (4.107) is achieved at
n
= 0 if
and only if (in terms of the original length )

ca
=
2/
/j
-
sin ,
cos 2,
.
otherwise it is achieved at
n
= 0 (no dislocations are nucleated). Note that
the sign of
n
depends on the angle ,:
n
is positive if 0

< , < 45

and is
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 115
negative if 45

< , < 90

. The energetic threshold value for the symmetric


double slip turns out to be exactly the same as that of the single slip problem,
see equation (4.71).
Based on the previous analysis, we now assume that
|
() =

() = ()
for (0. ). Due to the boundary conditions,

should change its sign on


the interval (0. ). The one-dimensional theory of dislocation pile-ups as well
as the solution of the previous problem suggest to seek the minimizer in the
form
() =

1
(). for (0. |).

n
. for (|. |).

1
( ). for ( |. ).
(4.108)
where
n
is a constant, | an unknown length, 0 |

2
, and
1
(|) =
n
.
The total energy functional becomes
1[] =

|
0
(4/ sin ,

1
+4/ sin
2
,
2
1
) d+2
(
1
2
cos 2,
)
2
. (4.109)
where
=
1

(
2

|
0

1
d + ( 2|)
n
)
. (4.110)
We have to nd
1
() and the constants,
n
and |.
This variational problem can be solved in exactly the same manner as in
the previous case. Here, we present the solution:

1
=
cos 2,( 2 cos 2,)
/(1 + ) sin
2
,
(
|
1
2

2
)
. 0 |. (4.111)
where the average of is given by
=
(3 2|)|
2
cos 2,
6/(1 + ) sin
2
, + 2(3 2|)|
2
cos
2
2,
. (4.112)
with

n
=
cos 2,( 2 cos 2,)
2/(1 + ) sin
2
,
|
2
. (4.113)
The equation to determine | reads
(

2 cos 2,

/ sin ,(sign

1
)
( 2|) cos
2
2,
)

(3 2|)|
2
cos 2,
6/(1 + ) sin
2
, + 2(3 2|)|
2
cos
2
2,
= 0
Fig. 4.27 shows the evolution of ( ) as increases, for , = 30

(contin-
uous lines) and , = 60

(dashed lines), where = /j


-
. For the numerical
116 CHAPTER 4. CRYSTAL PLASTICITY
j=30
j=60
a
b
c
d
e
f
y
_
b
Figure 4.27: Evolution of for double-slip constrained shear at zero dissipa-
tion: a,d) = 0.02, b,e) = 0.04, c,f) = 0.06
simulation we took the material parameters from Table 4.1, and = 1jm,
so that

= /j
-
= 0.349.
It is interesting to plot the shear stress t = j(2 cos 2,) as a function
of the shear strain. As we know, for <
ca
no dislocations are nucleated
and = 0, so the shear stress t = j. For
ca
we take from (4.112)
to compute the shear stress.
A
A
B
B
t/m
g
0
j=30
j=60
Figure 4.28: Normalized shear stress versus shear strain curve for double-slip
constrained shear at zero dissipation
Fig. 4.28 shows the normalized shear stress versus shear strain curve OAB
for , = 30

and OAB for , = 60

. The behavior is quite similar to that of


a crystal deforming in single slip.
If the resistance to dislocation motion (and hence the dissipation) can-
not be neglected, the plastic distortion may evolve only if one of the yield
conditions
|
= 1
|
and

= 1

is fullled. If

< 1

, then

is
frozen, the dislocation density remains unchanged, and the crystal deforms
elastically. Again, it is useful to analyze the evolution of

in the special
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 117
case of symmetric double slip for which we assume that
|
() =

() = ()
for (0. ) and 1
|
= 1

= 1. Introducing the dimensionless variable


= /j
-
, we write the total energy functional as
1() =


0
j
[
2
(
1
2
cos 2,
)
2
+ 2/ sin ,

+ /(1 + ) sin
2
,
2
]
d.
(4.114)
Computing the variational derivative of (4.114), we derive from (4.103) the
yield condition
j

2 cos 2,( 2 cos 2,) + 2/(1 + ) sin


2
,

= 1.
Consider rst the case , < 45

. We divide this equation by j to transform


the yield condition to

2 cos 2,( 2 cos 2,) + 2/(1 + ) sin


2
,

= 2
c
cos 2,. (4.115)
with
c
1,2jcos 2, and the prime denoting dierentiation with respect
to . We shall further omit the bar over for short.
We regard as a given function of time (the driving variable) and try
to determine (t. ). We consider the loading path shown in Fig. 4.14. The
problem is to determine the evolution of as a function of t and , provided
(0. ) = 0 and , < 45

.
Since the plastic distortion, , is initially zero, we see from (4.115) that
= 0 as long as <
c
. Thus, the dissipative threshold stress (the yield
stress) t

= 1,2 cos 2, in this case. For small (t. r) and


c
, the yield
condition becomes
2 cos 2,( 2 cos 2,) + 4/ sin
2
,

= 2
c
cos 2,. (4.116)
Let us introduce the deviation of (t) from the critical shear
c
,

=
c
and simplify (4.116) to obtain
2 cos 2,(

2 cos 2,) + 4/ sin


2
,

= 0. (4.117)
Since this equation is linear, is proportional to

such that =

1
,
where
1
is

1
=
cos 2,(1 2
1
cos 2,)
2/(1 + ) sin
2
,
( ). (4.118)
The average of
1
is obtained in the form

1
=

2
cos 2,
12/(1 + ) sin
2
, + 2
2
cos
2
2,
. (4.119)
118 CHAPTER 4. CRYSTAL PLASTICITY
j=30
j=60
y
_
b
1
Figure 4.29: Graphs of
1
( ) for double-slip constrained shear at non-zero
dissipation
Fig. 4.29 shows the graphs of
1
( ) for , = 30

(continuous line) and


, = 60

(dashed line). For the numerical simulation we took the material


parameters from Table 4.1, and = 1jm, so that

= /j
-
= 0.349.
After reaching


c
, we unload the crystal by decreasing . Since
becomes smaller than 1, does not change ( =

()) until
2 cos 2,( 2

cos 2,) + 2/(1 + ) sin


2
,

= 2
c
cos 2,. (4.120)
where

() is the solution of (4.116) for (t) =

. From (4.120) we can see


that the plastic ow begins when (


c
) =
c
, i.e. for =

2
c
. From that value of , the yield condition = 1 holds leading
to a decrease of which should be now determined from the equation
2 cos 2,( 2 cos 2,) + 2/(1 + ) sin
2
,

= 2
c
cos 2,. (4.121)
Since for (
c
.

), the deviation
|
= +
c
is positive, equation
(4.121) can again be transformed to equation (4.117) and solved in exactly
the same manner if we replace

=
c
in all formulas (4.118) and (4.119)
by
|
= +
c
. As approaches
c
, tends to zero because
|
0. The
further increase of from
c
to zero does not cause change in which
remains zero.
It is not dicult to modify the construction given above to nd the solu-
tion for , 45

.
For symmetric double slip the dislocation densities are the same, c
|
=
c

sin ,. The resultant Burgers vector of all dislocations has only one
non-zero component in the -direction: c
:
= 2
,
sin
2
,. Thus, couples of
dislocations near the boundaries form super dislocations with the Burgers
vector in the -direction. The normalized dislocation density c =

sin ,
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 119
j=30
j=60
y
_
a
1
Figure 4.30: Graphs of c
1
( ) for double-slip constrained shear at non-zero
dissipation
can be calculated from the solution (4.118). Since () is proportional to

,
c() is also proportional to

such that c() =

c
1
() with
c
1
() =
cos 2,(1 2
1
cos 2,)
2/(1 + ) sin ,
( 2).
and with
1
from (4.119). Fig. 4.30 shows the graphs of c
1
for , = 30

(continuous line) and , = 60

(dashed line) for (0.

).
h/d=40
h/d=160
h/d=80
h/d=240
withdissipation
energyminimization
t/t
0
g
Figure 4.31: Comparison of the normalized shear stress versus shear strain
curves obtained from this approach and in [6]
It is interesting to calculate the shear stress t which is a measurable
quantity. During loading, we have for the normalized shear stress (or the
elastic shear strain)
c
(c)
=
t
j
=
c
+ (

2
1
cos 2,) . (4.122)
120 CHAPTER 4. CRYSTAL PLASTICITY
with
1
from (4.119). The second term of (4.122) causes hardening due to
the dislocations piling up. Equation (4.122) describes the size eect in this
model.
Fig. 4.31 shows the stress-strain curves during loading obtained from
energy minimization and from the ow rule as compared with the average
stress-strain curves from discrete dislocation simulations in [6]. In order
to compare with the discrete dislocation simulations we took 1 such that

c
=
ca
= 0.00118, , = 60

, t
0
= 1.9 10
3
j and let all other material
constants remain the same as in the previous Section. The average stress-
strain curves in the discrete dislocation simulations are provided for four
dierent ratios ,d, where d is the spacing between the active slip planes. It
is seen that reasonably good agreement of the discrete and the continuum
approaches is observed at ,d = 240.
g=0.0068
g=0.0118
g=0.0168
g=0.0218
h/d=80
energyminimization
withdissipation
y
_
h
u
,y
Figure 4.32: Comparison of the total shear strain proles obtained from the
present approach and in [6]
Fig. 4.32 shows the total shear strain proles obtained from energy min-
imization and from the ow rule as compared with the average proles pro-
duced by the discrete dislocation simulations in [6]. A somewhat larger dis-
crepancy between the shear strain proles obtained by the discrete and con-
tinuum approaches is perhaps due to the rather low ratio ,d = 80 taken to
simulate these curves in [6].
During inverse loading, when the yield condition = 1 holds true,
equation (4.122) changes to
t
j
=
c
+ (
|
2
1
cos 2,) .
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 121
where

1
=

2
cos 2,
12/(1 + ) sin
2
, + 2
2
cos
2
2,
.
with the deviation
|
= +
c
used instead of

=
c
.
A
B
C
D
0
-g
cr
g
cr
g
*
g
t/m
g
*
Figure 4.33: Normalized shear stress versus shear strain curve for double-slip
constrained shear at non-zero dissipation for , = 60

Fig. 4.33 shows the normalized shear stress (or elastic shear strain) versus
shear strain curve for the loading path of Fig. 4.14, with

= 0.025, , = 60

,
= 1jm, while all other parameters remain the same. The straight line OA
corresponds to purely elastic loading with increasing from zero to
c
. Line
AB corresponds to plastic yielding with = 1. Yielding begins at point
A with the yield stress t

= 1,2 cos 2,, and we can observe the work


hardening due to the dislocation pile-up which is described by the second
term of (4.122). During unloading, as decreases from

to

2
c
(line BC), the plastic distortion =

is frozen. As decreases further


from

to
c
, plastic yielding occurs with = 1 (line CD). From
Fig. 4.33 it is seen that the yield stress t

= t

+21,2 cos 2, at point C, at


which the inverse plastic ow sets on, is larger than 1,2 cos 2, (because
t

1,2 cos 2,). Along line CD, as is decreased, the created dislocations
annihilate, and at point D all dislocations have disappeared. Finally, as
increases from
c
to zero, the crystal behaves elastically with = 0. In
this closed cycle OABCDO dissipation occurs only along lines AB and CD.
It is interesting that lines DA and BC are parallel and have the same length.
This corresponds again to the Bauschinger eect as in the case of single slip.
Exercises
4.1 The only non-zero component of the plastic distortion is

:
= r
2
(1 +
4
).
122 CHAPTER 4. CRYSTAL PLASTICITY
Find the dislocation density.
4.2 Prove that Nyes dislocation density tensor satises the following equa-
tion
c
i),)
= 0.
which means that dislocations cannot end in the lattice.
4.3 Prove the following identity
c
ioo
c
)co
-
j
oo,oc
+
1
2
(c
ioo
c
o),o
+ c
)oo
c
oi,o
) = 0.
4.4

Given the plastic strain tensor -


j
i)
and the dislocation density tensor
c
i)
satisfying the equations in Exercises 4.2 and 4.3. Find the plastic
distortion tensor.
4.5 The plastic distortion tensor is given by

co
= (r. ):
c
:
o
.
where s = (cos ,. sin ,) and m = (sin ,. cos ,). Find the dislocation
density tensor.
x
y
M
Figure 4.34: Bending of a beam
F
Figure 4.35: Indentation
4.6

For the plane strain bending of a beam with the slip planes parallel to
the plane = 0 and the dislocation lines parallel to the .-axis shown
in Fig. 4.34 write the energy functional of the crystal beam.
4.7 Develop the numerical procedure for the plane constrained shear of a
single crystal deforming in non-symmetric double slip.
4.8 The similar problem for a bicrystal having non-symmetric slip system.
4.5. SINGLE CRYSTALS DEFORMING IN DOUBLE SLIP 123
4.9

A rigid wedge is pressed into a crystal by a force 1 as shown in Fig. 4.35.


Develop a numerical procedure to nd the dislocation density beneath
the wedge and the force versus displacement curve.
4.10

Find the dislocation density near a rigid spherical inclusion embedded


in an innite crystal under tension.
124 CHAPTER 4. CRYSTAL PLASTICITY
Bibliography
[1] Calladine, C. R., Engineering Plasticity, Pergamon 1969.
[2] Hill, R., The Mathematical Theory of Plasticity, Clarendon Press 1983.
[3] Hull, D., Introduction to Dislocations, Pergamon Press 1975.
[4] Le, K. C., Introduction to Micromechanics, Nova Science 2010.
[5] Lubliner, J., Plasticity theory, Macmillan Publishing Company 1990.
[6] J. Y. Shu, N. A. Fleck, E. Van der Giessen, and A. Needleman. Boundary
layers in constrained plastic ow: comparison of nonlocal and discrete
dislocation plasticity. J. Mech. Phys. Solids, 49:13611395, 2001.
125

Вам также может понравиться