Sunteți pe pagina 1din 83

INTERNATIONAL MASTER'S OF THE

THEORETICAL & PRACTICAL


APPLICATION
OF FINITE ELEMENT METHOD
AND CAE SIMULATION

AF.4 MATERIAL CONSTITUTIVE LAWS

ALBERTO FRAILE DE LERMA (*)


ARANCHA ALARCON-FLEMING
EDITION. - 9. 2018 ENRIQUE ALARCON ALVAREZ (*)

(*) E.T.S.I. INDUSTRIALES - U.P.M.


CHAPTER 4
PLASTICITY

4.1 Uniaxial behavior................................................................................................ 7


4.1.1 Yield stress. ........................................................................................... 7
4.1.2 Yield Hardening ..................................................................................... 7
4.1.3 Loading and unloading criteria ............................................................... 9
4.1.4 Idealized constitutive equations ............................................................. 9
4.1.4.1 Elastic-perfectly plastic solid ................................................. 9
4.1.4.2 Plastic solid with hardening ................................................. 16
4.1.5 Hardening rule ..................................................................................... 17
4.2 Plasticity theory. General formulation............................................................... 18
4.2.1 Yield surface ........................................................................................ 19
4.2.1.1 Von Mises yield surface ...................................................... 20
4.2.1.2 Tresca’s yield surface ......................................................... 22
4.2.2 Loading criterion .................................................................................. 23
4.2.3 Flow rule .............................................................................................. 24
4.2.4 Hardening ............................................................................................ 30
4.2.4.1 Isotropic hardening.............................................................. 31
4.2.4.2 Kinematic hardening ........................................................... 33
4.2.4.3 Mixed hardening.................................................................. 35
4.2.4.4 Experimental determination of the hardening parameters ... 36
4.3 Plasticity: Analysis with the finite element method ........................................... 38
4.3.1 Newton-Raphson method .................................................................... 38
4.3.2 Non-linear finite element method: Algorithm ........................................ 39
4.3.3 Evaluation of the residual vector: Integration of the elastoplastic
constitutive equations ....................................................................................... 40
4.3.3.1 Generalized trapezoidal rule ............................................... 42
4.3.4 Evaluation of the tangent operator. Consistent tangent modulus ......... 43
SELF EVALUATION EXERCISES ............................................................................. 46
SUMMARY OF MAIN IDEAS ..................................................................................... 77
REFERENCES .......................................................................................................... 79
CHAPTER 4

4
PLASTICITY

Plasticity theory mathematically describes the mechanical behavior of materials


working in the plastic regime. Its formulation is very useful to model the
behavior of ductile materials, such as steel, where the strain is mainly plastic.
Plasticity is also used for other materials such as concrete, soils and rocks.

From the early plasticity scientific work of Tresca in 1864, plastic theory has
developed greatly, especially with the use of high capacity computers.

Some aspects of modern plasticity theory were applied from the very beginning
of research on continuum mechanics. Take for instance, the well-known
Galileo’s experiment from his “Dialogues concerning two new sciences” (1638)
with a built-in cantilever beam (Figure 4.1.a) written almost 40 years before
Hooke´s linear constitutive equation.

Figure 4.1 Galileo’s experiment


He clearly understood that equilibrium is a basic requirement for structural
analysis and since he was interested in the collapse load of the beam, he

1
assumed a limit load R at the centroid of the fixed section AB (Figure 4.1.b) and
taking moments about B, he obtained the collapse load as

R = 2(L/h) P

This is tantamount as to write the equilibrium of horizontal forces, assuming a


compressive force of the same value concentrated at the lower corner B, which
means an infinite stress value there.

Had he assumed a couple of opposite forces, R and R´, acting symmetrically


around the centroid (Figure 4.1.c), he would have obtained the limit plastic
moment that leads to the cantilever collapse load.

Another precursor and influential study on the use of an assumed failure


mechanism along with an equilibrium condition, established in this case using
the principle of virtual work, is shown in Figure 4.2. These drawings belong to
the memoir written by Le Seur, Jacquier and Boscowich (1743) and they
studied, at the request of Pope Benedict XIV, the cracks that appeared in Saint
Peter’s dome in Rome as well as its analysis by Poleni.

b) Le Seur, Jacquier y Boscovich’s


a) Poleni’s sketch
sketch

Figure 4.2 St. Peter’s dome

2
It is possible to see (ref. 3) that the combination of a hypothetical collapse
mechanism along with the establishment of its global equilibrium, leads to a
bound to the collapse load. This is considered one of the bound theorems for
elastoplastic bodies accepted today for limit design.

It is also interesting that Coulomb choose this approach in his Memoir to the
French Academy (1773) when trying to obtain the design formula for collapse
loads of retaining walls, arches or columns. (For a lucid and instructive
description refer to J. Heyman (ref.4)).

Figure 4.3 Coulomb’s Memoir

3
Figure 4.3 is a copy of Plate I in Coulomb´s Memoir. It includes, in addition to
of the description of a testing device (Coulomb´s fig.1) and equilibrium
considerations (fig. 2, 3 and 4), the famous idea of a plane of rupture (fig.5) in
a column and the surface of rupture for retaining walls (figs. 7 and 8) where a
force of cohesion (stress concept was not yet discovered) acted simultaneously
with other due to friction. Equilibrium was written as a function of a parameter
(for instance the angle of the assumed failure plane), and a rationally based
minimum or maximum condition allowed the estimation of the collapse load. He
obtained the formula that is still taught and applied in engineering practice
today. In fact this was the first application of bounding plasticity ideas to soil
mechanics as shown in Heyman´s essay (ref. 4).

Interestingly enough, although out of our scope in this chapter, is Coulomb’s


fig. 6 where he solves Galileo’s problem from a purely elastic view point.

In the XIX century, with the advent of the elasticity theory, the research
emphasis was mainly on the linear elastic behaviour, but some substantial
progress occurred thanks to theoreticians like Saint Venant and Poncelet, or
bright experimentalists as Tresca and Bauschinger.

Figure 4 is taken from one of Henry Tresca´s communications to the French


Academy (1872) on the die extrusion of lead through a cylinder. This is the
starting point of the plasticity criteria based on the maximum absolute value of
shear stresses as opposed to the maximum strain value proposed by Poncelet
(1839) and Saint Venant (1870) or the maximum principal stress value
defended by Lamé (1852) and Rankine (1858).

4
Figure 4.4 Tresca’s apparatus used in his
flow experiments (1872)
At the same time Bauschinger (1875-1886) conducted and published many
results from his experiments on steel, iron, wood, etc. in which he discovered
the lowering of the yield stress in a second loading when the stress was of
opposite sign from that of the first loading, which is part of the experimental
justification of the today´s kinematic strain hardening concept.

In spite of those early attempts, when Hill (ref. 1) published his epoch-making
book in 1950, he wrote “significant advances did not follow (to the ones above
described) until after 1920. The theory of plasticity still remains a young science,
with few serious students”.

In those years of the first half of the twentieth century several points became
clear: the importance of the load trajectory in the stress space and the
impossibility of obtaining a direct stress-strain relationship. This lead to the
Saint Venant procedure of an incremental constitutive equation that later
received the name of “flow theory” as opposed to the “deformation theory”
proposed by Hencky in 1924 to simplify the resolution of practical problems.
The last one was applied intensively by URSS researchers (see Katchanov
ref.6). The development of “flow theory” applied to metals is mainly due to

5
German researches between the two World Wars. Some of them have been
already mentioned in Chapter 1.

Other important points of the theory are:

a. Under very general conditions, the solids can reach a “limit” or “collapse”
load whose estimation is a very important objective when we need to
calculate the ultimate strength of the structure to quantify its safety.
b. In ductile materials, the collapse load is reached after important
displacements. This capacity to make excursions on the plastic range
without losing strength is one of the most appreciated virtues of ductile
structures, especially in seismic engineering where energy dissipation
capacity guaranteed by structural ductility is as important as stiffness or
strength.
c. It is possible to obtain bounds of the collapse load using equilibrium
conditions. This is achieved either postulating a failure mechanism, as
Coulomb’s did, or searching for a statically admissible stress field along with
a limiting flow condition.

Those approaches constitute the basis for the spectacular development of


plasticity applied to soil mechanics after 1950 (see references 5, 8, 10, 12, 13),
structural steel structures (ref. 2, 7, 12), concrete structures (ref. 9, 10, 18, 20)
or even the reinterpretation of the behaviour of masonry structures (ref.11).
Some of the resulting brilliant ideas are the Cam-Clay model for soils, the
stress-fields approach in concrete or the capacity-based design in seismic
engineering. The development of all those applications was fuelled by the
advent of the computer and the possibility of solving complicated non-linear
problems iteratively using procedures like the Finite Element Method (ref. 15,
19, 21). Moreover, the interest on fast loading has induced the studies on
Viscoplasticity including the effects of time and temperature (see reference 16).

It is truly impossible to give a complete panorama of the current situation, so


our chapter will discuss only the most important and basic details of the theory.
We encourage the interested reader to study more advanced texts (see
references).

6
4.1 Uniaxial behavior

Recall some concepts from chapter 1.

The uniaxial behavior of any material is characterized by the stress-strain


(σ – ε) relationship obtained from a simple tensile test as shown in Figure 4.5.

Figure 4.5 Stress-strain (𝛔 – 𝛆) curve for a simple tensile test

4.1.1 Yield stress.

Figure 4.5, shows that the behavior is linear elastic up to point A. In this range
O − A, the behavior of the material is independent from the load path.

The stress at point A (σy ) is called yield stress or initial yield point.

4.1.2 Yield Hardening

From point A and beyond, the material is yielding, i.e., there is permanent strain
when the stress goes back to zero. We observe two clearly defined regions: a
flat region A − B where strain increases at constant stress (perfect plasticity)
and a region B − U where an increase in stress produces an increase in strain
(plasticity with hardening).

7
If unloading occurs in any point C of the range A − U, the only part of the strain
that is recovered is DE. The other part (OD) remains even without the load. This
irreversible strain which is associated to energy dissipation is the plastic strain.
Therefore, the strain at an arbitrary point C in the plastic region is the sum of
the plastic strain OD (εp ) and the elastic strain DE (εe ):

ε = εe + εp (1)
This separation of the strain into an elastic or reversible part and a plastic or
irreversible part is justified by the physics of the elastic and plastic phenomena
and is only valid for infinitesimal deformations.

Yield without hardening corresponds to the case when the stress remains
constant during yield (fixed yield stress). If there is hardening, yield only occurs
when the stress increases and consequently the yield stress also increases
(variable yield point). This is shown in Figure 4.5: after unloading, CD, if the
material is loaded again yielding would not occur until point C.

It is very important to point out a fundamental difference between linear


Elasticity and Plasticity. In Elasticity Theory there is a one to one relationship
between the total strain and the actual stress. In Plasticity there is no direct
relationship between the actual stress state and the permanent total strain.

Consider, for instance, that we reach point C (Figure 4.5) in a tensile test on a
thin-wall tube that, afterwards, is unloaded until point D. Now we reload again
but subjecting the tube to a torsional moment until point C is reached again with
a stress equivalent to the shear stress caused by the torsion moment (recall
Chapter 1). We shall have the same plastic strain, but the stress state will be
totally different!

Therefore, it is very important to know the loading trajectory and, if the yield limit
is reached, the only possibility we have is to establish a relation between the
infinitesimal increments of stress σ̇ and the infinitesimal increments of plastic
strain ε̇ p . This approach is called “flow theory” and was started by Barré de Saint
Venant in the XIX century.

8
4.1.3 Loading and unloading criteria

Since the constitutive relationships of a plastic material are different for the case
in which there is plastic loading, elastic loading or unloading, it is necessary to
identify if an increment of stress or strain causes plastic loading or elastic
unloading.

For uniaxial tension, we have the following expressions:

 σ = σy and σ dσ > 0 for plastic loading


 σ = σy and σ dσ < 0 for unloading

4.1.4 Idealized constitutive equations

4.1.4.1 Elastic-perfectly plastic solid

Yielding only occurs when stress reaches the yield stress σy :


σ
|σ| < σy ⇒ ε = εe =
E (2)
σ
|σ| = σy ⇒ ε = εe + εp = + εp
E
In this case, when σ = σy the plastic strain is undefined.

Truss structures (Figure 4.6.a) offer an interesting occasion to generalize the


tensile test and to explore the collapse of the truss when each of its members
follows a pure elastoplastic law such as in Figure 4.6.b. We can immediately
derive that the axial forces will follow a similar law (Figure 4.6.c). But, how can
the geometric configuration and the type of loading influence the shape of the
truss load-deflection diagram (Figure 4.6.d)? While all members remain in the
elastic domain, the global behavior will be linear, with the slope of the line (load,
displacements) reflecting the truss stiffness. But as soon as one bar reaches its
plastic-limit load Np = σy A, some load redistribution among the bars will occur.

On the other hand, the end of the structure serviceability, the so-called
“collapse load”, will occur when all the three members reach their plastic load
Np .

9
Let us analyze closely this particular example. Consider that the three members
are made of the same material: Young modulus E, yield stress σy and cross
section A.

The limit load will be reached when N1 = N2 = N3 = σy A

The collapse load Fp is obtained with the vertical equilibrium equation and
plugging in the values of the limit loads N3 and N1 previously obtained

Fp = N3 + 2 N1 cos α = σy A (1 + 2 cos α) (3)


So (what a wonderful discovery!) we are able to calculate the collapse value
of the truss right away!

Now let us see how we can follow the progression of plasticity from the elastic
situation.

Equilibrium (Figure 4.6.e):

N1 = N2
F = N3 + 2 N1 cos α (4)

Compatibility (Figure 4.6.f)

∆l1 = ∆l2 = v cos α


∆l3 = v (5)

Elastic solution

Constitutive equation
Nj Lj EA
∆lj = ; Nj = ∆l (6)
EA Lj j

Combining the equilibrium and compatibility conditions


EA EA EA
N3 = v ; N1 = N2 = v cos α = v cos 2 α
L L/ cos α L
EA
F=( v) [1 + 2 cos 3 α] = N3 (1 + 2 cos 3 α ); F − N3 = 2 N3 cos 3 α (7)
L
1
N1 = N2 = (F − N3 ) = cos 2 α N3 < N3
2 cos α

10
1 3 2

L
L/ α α
co

0 < α < π/2


v; F

(a)

σ N Fp
Fe ?
σy Np

ε Δl v

Np = σy A
σy ? Fe

(b) (c) (d)


Fp
N3 Fe
N1 N2
α EA
(1  2cos3 )
L
l2
v  l3 v
F l2  v cos 
Fe

(e) (f) (g)


Figure 4.6

11
Elastic limit load

When the stress in the middle bar is the yield stress σ3 = σy , bar 2 starts
yielding and, from there on, the linear behavior of the truss is characterized by
the constant slope (Figure 4.6.g)
F EA
= (1 + 2 cos 3 α) (8)
v L
stops to be valid. Then, since

N3 = A σ y
(9)
[Fe = N3 (1 + 2 cos 3 α) = A σy (1 + 2 cos 3 α)]

where Fe is the elastic limit load of the structure.

The corresponding displacement ve is obtained as


L L σy
[ve = N3 = A σy = L ] (10)
EA EA E
and the stresses in the inclined bars are

σ1e = σe2 = (cos2 α) σe3 = (cos 2 α) σy < σy (11)


For instance, if α = π/3
1
σ1e = σe2 = σy < σy (12)
4
See Figure 4.7.

Elastoplastic solution

From this moment on, the central member has lost its stiffness but the other two
maintain their loading capacity because only portion (cos α)σy has been used
and two symmetric members constitute a statically determinate (isostatic) truss
that can resist increments of load that will produce additional vertical
displacements. These displacements will be passively followed by the central
member that, after yielding, cannot share any more of the load.

12
Setting

∆N3 = 0 (13)
The equations now are

∆ F = 2(cos α) ∆N1
∆I3 = ∆v (14)
∆I1 = ∆I2 = ∆v cos α

N3 N1=N2

A σy
ΔN1 ΔN2

A σy/4
ve v ve v Δv; ΔF

(a) (b) (c)


Figure 4.7

EA EA
∆N1 = ∆N2 = ∆ l1 = (cos2 α) ∆v (15)
L/ cos α L
and
∆v
∆F = 2 ∆N1 cos α = 2(cos 3 α) EA (16)
L
The truss stiffness is
∆F EA EA
k ep = = 2 cos 3 α ( ) < (1 + 2 cos 3 α) = ke (17)
∆v L L
For instance, if α = π/3

1 3
k ep 2 ( ) 1/4 1
= 2 = = (18)
ke 1 3 1 + 1/4 5
1 + 2( )
2

13
and we can continue to plot the (F, v) diagram as a line, until member 1 and 2
start to yield and the structure reaches the so-called plastic limit collapse load
Fp .
F
Fp=Fe+ΔF
kep
Fe

A σy/4 ke

ve ve+Δve v
Figure 4.8 Force-displacement plot

The strength reserve in bars 1 and 2 is


1 3
∆N1 = σy A − σy A = σy A (19)
4 4
Then
3
AF = 2∆N1 cos α = (cos α)σy A
2
L 3 L
∆v = ∆N1 = σ (20)
EA cos 2 α 4 E cos 2 α y
3 Lσy
∆v = ( )
4 cos 2 α E
The collapse load is
3
Fp = Fe + ∆F = Aσy [(1 + cos 3 α) + cos α] (21)
2
and the displacement before the total yielding
σy 3 1
v = ve + ∆v = L (1 + ) (22)
E 4 cos 2 α
For instance, if α = π/3

14
Fe 1 5 ve
=1+ = σy = 1
Aσy 4 4 L( )
𝐸
∆F 3 ∆v
= σy = 3 (23)
Aσy 4 L( )
𝐸
Fp 8 v
= =2 σy = 4
Aσy 4 L( )
𝐸

F
pa
Ay ra
lle
l
2

5/4

v
1 4 σ 
L y 
 E 
 

Figure 4.9 Non-dimensional force displacement plot


Note that we have a strength reserve at the price of increasing displacements.

Here we can see the ductility of the structure, where it is capable of having
large excursions in the plastic domain without losing strength.

Moreover, for alternative loading and unloading cycles, an interesting problem


about the variation of plastic strains occurs. See Self Evaluation Exercise 1.

We can plot also the elastic domain and the yield surface of the structure by
representing the stress-state in a multidimensional space where each axis
corresponds to every member. For instance, in the truss model problem, we
only need two axes because due to the symmetry the inclined members present
the same behavior.

15
σ2=N2/A

σy

σ1=N1/A=σ3
-σy σy

-σy

Figure 4.10

It is clear that the elastic domain lies inside the square (+ σy ; + σy )

Something similar can be done with the strains (see Self Evaluation Exercise
1.).

4.1.4.2 Plastic solid with hardening

Depending on the hardening type, there are many constitutive models:

a) Elastic-linearly hardening model (Figure 4.10.a)

The stress-strain relationship is approximated by two straight lines:

σ
|σ| < σy ⇒ ε = εe =
E
σ (σ − σy ) (24)
|σ| > σy e p
⇒ ε=ε +ε = +
E Et

where the tangential modulus Et = is assumed to be constant.

16

B 
 B
A
F
B A
O
 O C
C  B 
D E

D
C

(a) elastic-linearly (b) isotropic hardening (c) kinematic hardening


hardening model rule rule
Figure 4.11 Hardening models

b) Ramberg-Osgood model

In this model, a non-linear expression is employed to represent the σ − ε


relationship
σ σ n
|σ| > σy ⇒ ε= + a( ) (25)
E b
where a, b and n are material constants to be determined by a curve-fitting with
the experimental data.

4.1.5 Hardening rule

The hardening rule is used to represent the evolution of the yield stress σy
throughout the loading process. That evolution is established through a
hardening parameter α which is related to the plastic deformation.

The most common hardening rules correspond to isotropic hardening and


kinematic hardening. Isotropic hardening (Figure 4.10.b) states that the
progressively increasing yield stresses under both tension and compression
loadings are always equal

|σ| = |σ(α)| (26)


Kinematic hardening states that the plastic strain accumulated in tension
loading affects not only the yield stress in tension but also its subsequent yield
stress in compression, as reflected by the Bauschinger effect (Figure 4.10.c).

17
4.2 Plasticity theory. General formulation

In the uniaxial formulation, the general stress-strain relationship should be


expressed in incremental form

ε̇ = G(σ̇ ) (27)

Plasticity is a phenomenon that is independent of time. Consider a function G


which is a first order homogeneous function of the stress rate1. To see that
consider two time references t and t′ such as:

t′ = k t
∂ ∂ (28)
=k
∂t ∂t′
Then:

ε̇ t = k ε̇ t′ σ̇ t = k σ̇ t′ (29)
Using the stress strain relationship

ε̇ t = G(σ̇ t )
(30)
k ε̇ t′ = G(k σ̇ t′ )
If G is homogeneous and first order function of the stress rate

G(k σ̇ t′ ) = k G(σ̇ t′ ) (31)


Therefore:

ε̇ t′ = G(σ̇ t′ ) (32)
which shows there is no dependence on the time reference as mentioned
previously. The dot over the symbols that is reserved for time derivate in
viscoelasticity and dynamics, here it only has meaning of increment or rate. An
incremental relationship linear with respect to stress rate is then proposed:

ε̇ = C e σ̇ (33)

1 A function f(x) is homogeneous of order r if f(k · x) = k r · f(x)

18
where C e is the material elastic (or compliance) tensor, or expressed in terms
of the stiffness matrix De

σ̇ = De ε̇ (34)
According to the constitutive equation, by extending the concepts from the one-
dimensional case, the plasticity theory is based on four fundamental aspects:

a) The existence of a yield surface in the stress space of Haig Westergaard


(equivalent to the yield point for one-dimensional yielding) that
determines if the material is in an elastic or a plastic state.

b) Formulation of hardening and softening rules that define the evolution


of the yield surface (equivalent to the yield point evolution for one-
dimensional yielding).

c) A flow rule that allows the establishment of the constitutive rules for
plastic strain increments.

d) Establishment of loading criteria to identify if a stress increment


produces a plastic strain (loading) or an elastic strain (unloading).

These four aspects establish a general theory of plasticity and will be developed
in the following sections.

4.2.1 Yield surface

In uniaxial stress states, the yield stress of a material is obtained by a well-


defined point in the stress-strain curve. In a general stress state, the yield
stress is defined with a surface called yield surface in the stress space.

F(σij ) = 0 (35)

For points inside that surface, the material behaves elastically. For points
located on that surface, the material begins to yield.

The stress space is a space with six stress components which is hard to
visualize. If the body is isotropic, the yield does not depend on the orientation
of axis but only on the values of principal stresses:

19
F(σ1 , σ2 , σ3 ) = 0 (36)
or the stress invariants I1 , J2 , J3

F(I1 , J2 , J3 ) = 0 (37)
which makes representing it easier. (Review Appendix Chapter 1)

4.2.1.1 Von Mises yield surface

In this criterion, yield occurs when the elastic distortion energy reaches a critical
value (Recall Appendix Chapter 1). The yield surface is defined:

F(σ) = J2 − k 2 = 0 (38)

Where J2 is the second invariant of the deviatoric stress tensor.

In this case, F is a cylinder of axis σ1 = σ2 = σ3 (Figure 4.12); if σ1 ; σ2 ; σ3 is


on the yield surface, then σ1 + α, σ2 + α, σ3 + α is also on the yield surface.
Why?

R 2k
σ3

σ2

σ1

Figure 4.12 Von Mises yield surface

20
Figure 4.13 Von Mises section with the deviatoric 𝛑 plane
(isometric projection)

The cross-section of the yield surface with a deviatoric plane is a circle of radius
R = √2 k (Figure 4.13).

If one of the principal stresses is zero (σ3 = 0), then the yield surface is an
ellipse in the σ1 − σ2 plane (Figure 4.14)

σ12 + σ22 − σ1 σ2 = σ2y = √3 k (39)

where σy is the yield stress in uniaxial tension.

Figure 4.14 Two-dimensional von Mises “surface”

21
4.2.1.2 Tresca’s yield surface

In this criterion, yield is not related to energy but to the maximum shear stress.
The yield surface is defined in terms of principal stresses by the six planes:
1
|σ − σ2 | = k
2 1
1
|σ − σ3 | = k (40)
2 2
1
|σ − σ1 | = k
2 3
Or more compactly written
1
F(σ) = Sup(|σi − σj |) − k = 0 (41)
2 i≠j
which represents a prism that has the same axes as the von Mises yield surface
and it is contained within it. In fact, von Mises’ idea came when he tried to
smooth the corners of the Tresca yield surface.

In the deviatoric plane, the yield surface is a regular hexagon with six corners
(Figure 4.15).

σ3

2 3 1
1   2  k
2 1
1
2   3  k
2 2
1 1
1
3   1  k
2 3
σ1 3 2 σ2

Figure 4.15 Tresca Yield surface in the deviatoric plane

22
If one of the stresses is zero (σ3 = 0), the yield surface becomes an irregular
hexagon (Figure 4.16). For example, in the quarter σ1 ≥ 0, σ2 ≥ 0 , then:

σ1 − σ2 = 2 k (42)

Figure 4.16 Two-dimensional Tresca yield surface

4.2.2 Loading criterion

For a perfectly plastic material, the yield surface F(σij ) = 0 is a fixed surface
in the stress space. Plastic strain occurs if the current stress point is on the
surface. In a loading case, this point stays there after an increment of load dσij .
On the other hand, when the load is removed, the stress point would move
inside the surface.

F(σij ) < 0 Elastic deformation


∂F
F(σij ) = 0 dF = dσ = 0 Loading
∂σij ij
∂F
F(σij ) = 0 dF = dσ
∂σij ij
< 0 Unloading

Figure 4.17 Loading criteria without hardening

23
For a hardening material, the yield surface changes when the stress point
reaches the yield surface2.

F(σij ) < 0 Elastic deformation


∂F
F(σij ) = 0 dσ > 0 Loading
∂σij ij
∂F
F(σij ) = 0 dσ = 0 Neutral load.
∂σij ij
∂F
F(σij ) = 0 dσ < 0 Unloading
∂σij ij

Figure 4.18 Loading criterion with hardening

4.2.3 Flow rule

The existence of the yield surface F that borders the elastic domain E (Figure
4.19) allows the definition of a relationship between the plastic strain increment
and the stress increment compatible with F.

Taking advantage of working in the space of principal stresses, the stresses


and strains will be six components vectors (σ1 ; σ2 ; σ3 )T ; (ε1 ; ε2 ; ε3 )T and the
material properties fourth order tensor can be simplified to be a second order
tensor.

Figure 4.19 Yield surface

2 Recall that the components of gradient of surface F(σ) is defined by vector 𝜕F/ ∂σ that points in the
direction of the normal n to the surface which can be, then, defined as n = 𝜕F/ ∂σ

24
Consider point A (Figure 4.20) which represents the stress state on the
boundary ∂E of the yield surface F . At that point, the normal vector is n .
Consider three different stress increments σ̇ 1 , σ̇ 2 , σ̇ 3 . The first increment points
towards the inside of F and according to the previous section, this implies
elastic behavior:

ε̇ = C e σ̇ (43)
where C e is the material’s elastic tensor or elastic compliance tensor.

n 𝜕𝐹
𝜕σ1
n1
𝜕𝐹
n = [n2 ] =
n3 𝜕σ2
𝜕𝐹
[𝜕σ3 ]
σ1
σ = [σ2 ]
σ3
ε1
ε = [ε2 ]
ε3
Figure 4.20 Plastic load
On the other hand σ̇ 3 produces and increment of load and therefore

ε̇ = C ep σ̇ (44)
where C ep is the elastoplastic tensor for the incremental behavior.

For stress increments with directions comprised between σ̇ 1 and σ̇ 3 (Figure


4.20), a logical rule should not include sudden changes in plastic response
when switching from loading to unloading, i.e., realize that to guarantee
continuity, a stress rate σ̇ 2 in the direction of the tangent plane to the elastic
border does not differentiate between elastic or elastoplastic behavior.

This implies:

25
a) Continuity through the yield surface

An arbitrary loading increment σ̇ can be expressed in its tangential and normal


components (Figure 4.21)

Figure 4.21 Stress rate decomposition


the previous continuity conjecture means that for each σ̇

C e σ̇ t = C ep σ̇ t (45)
b) Additive decomposition in the elastic and plastic components

According to the incremental law:

ε̇ = C ep σ̇ = C ep (σ̇ n + σ̇ t ) = C ep σ̇ n + Cep σ̇ t (46)


and using the information from the previous bullet point (45):

ε̇ = C ep σ̇ n + Ce σ̇ t (47)
Adding and subtracting C e σ̇ n to the right-hand-side leads to:

Ε̇ = (C ep − C e ) σ̇ n + C e (σ̇ n + σ̇ t ) (48)

the last term represents the elastic strain ε̇ e :

C e (σ̇ n + σ̇ t ) = C e σ̇ = ε̇ e (49)
Therefore, it follows the additive decomposition

ε̇ = ε̇ e + ε̇ p (50)
where:

26
ε̇ p = C p σ̇ n
(51)
C p = C ep − C e
c) Flow rule

According to step b), the plastic strain increment is:

ε̇ p = C p σ̇ n = C p n (nt σ̇ ) (52)
In general it is possible to write:
1
Cp n = m (53)
h
where h is a parameter that depends on the material and m is the orientation
vector of the strain increment.

Therefore:
1 t
ε̇ p = (n σ̇ ) m (54)
h
A particular case, called Associative Plasticity occurs when m = n. In this
case:
1 t
ε̇ p = (n σ̇ ) n (55)
h
Since:
∂F
∂σ
n=
(56)
∂F t ∂F
√( ) ·
∂σ ∂σ

∂F
Understanding that the vector expressed in components is
∂σ

∂F ∂F ∂F ∂F T
=[ ] (57)
∂σ ∂σ1 ∂σ2 ∂σ3
Then the plastic strain rate can be written as

27
∂F t
1 (∂σ) σ̇ ∂F
ε̇ p = t (58)
h ∂F ∂F ∂σ
( ) ·
[ ∂σ ∂σ]
The definition of h is found by multiplying ε̇ p by itself:

t 1 2
(ε̇ P ) ε̇ P = 2
(nt σ̇ )
h
(59)
P 2
1 2
|ε̇ | = 2 |σ̇ n |
h
And since

|ε̇ P | = |dεp |
(60)
|σ̇ n | = |dσn |
We obtain,

|dσn |
h= (61)
|dεp |

Parameter h is then a modulus that can be identified in a tensile test, after


selecting a yield function F. (See Self Evaluation Exercise 5).

For the case of Associative Plasticity, the elastoplastic stiffness matrix De is


obtained inverting C e and relates stresses and strains.

ε̇ e = ε̇ − ε̇ p
C e σ̇ = ε̇ − ε̇ p (62)

σ̇ = De (ε̇ − ε̇ p )
Using the definition of the plastic strain rate in terms of h leads to
1 t
σ̇ = De [ε̇ − (n σ̇ ) n] (63)
h
Multiplying both terms by n and rearranging leads to:

28
t
nt De n
(n σ̇ ) [1 + ] = nt De ε̇
h
(64)
1 t nt De ε̇
(n σ̇ ) =
h h + nt De n
Therefore the tress rate is

e
De nnt De
σ̇ = [D − ] ε̇ (65)
h + nt De n
The elastoplastic stiffness matrix is:

ep
De nnt De
e
D =D − (66)
h + nt De n

Sometimes, the plasticity problem is expressed with the strain vector dεPij
(corresponding to a certain stress σij ) being normal to an assumed Plastic
Potential Function G(σij ):

p ∂G
dεij = dλ (67)
∂σij

In the case of Associative Plasticity, the plastic potential and the yield surface
coincide (G = F), while if G ≠ F the flow rule would be called Non-Associative.
The plastic multiplier dλ is a positive scalar than is only non-zero when there is
plastic strain. If dλ is negative there is unloading. Since elastic unloading is the
only case possible, a negative dλ should be replaced by a zero value.

To obtain yielding, the stresses have to remain on the yield surface, i.e.: at any
instant the yield surface passes through the stress point and

F(σij ) = 0 F(σij + dσij ) = 0 dF(σij ) = 0 (68)

Then, for a case of perfect plasticity with the associative rule the consistency
condition is:
∂F ∂F ∂F
dF = σ̇ = 0 ⇒ (Dijkl dεkl − dλ Dijkl )=0 (69)
∂σij ij ∂σij ∂σkl

Where we have used eq. (62) and (67).

29
And solving for the plastic consistency parameter:
∂F
D
∂σij ijkl
dλ = dεkl (70)
∂F ∂F
D
∂σpq pqrs ∂σrs

Substituting this value into the constitutive law leads to:


p
σ̇ ij = Deijkl (ε̇ kl − ε̇ kl ) (71)
ep
where Dijkl is the elastoplastic stiffness matrix:

∂F ∂F 𝑒
D𝑒ijkl D
ep e
∂σkl ∂σij ijkl
Dijkl = [Dijkl − ] dεkl (72)
∂F 𝑒 ∂F
D
∂σpq pqrs ∂σrs

This is the same matrix seen previously in eq. (66) but with h = 0 (perfectly
plastic case).

It can be demonstrated that the normality condition due to an associative flow


rule leads to a unique solution of the problem.

According to Drucker’s stability postulate, it can be demonstrated also that for


stable (work-hardening) materials there are two consequences: the normality
rule and the convexity of the yield surface.

4.2.4 Hardening

When a material is loading, there are changes in the limit values, either as an
increase (hardening materials) or a decrease (softening materials). In these
cases, as the plastic strain appears, the yield surface is transformed into a new
loading surface. Since the changes depend on the plastic deformation, the
constitutive laws depend on the loading path.

The changes that occur in the yield surface are the main difference between
plasticity with hardening and perfect plasticity. The changes are controlled
through two hardening parameters: αij and χ

30
F(σij , αij , χ) = 0 (73)

For elastoplastic materials, the loading criteria was seen in section 4.2.2 and
the flow rule does not change. Here we are going to describe three possible
hardening rules, depending on the nature of the parameter αij : isotropic
hardening, kinematic hardening and mixed hardening.

4.2.4.1 Isotropic hardening

The isotropic hardening rule assumes that the yield surface expands uniformly.
The size of the subsequent loading surfaces is controlled by a hardening
parameter χ that affects k 2 :

F(σij , αij ) = f(σij ) − k 2 (χ) = 0 (74)

In the von Mises criterion (Figure 4.22) the yield surface is

σ3 σ3 2 k(1 )

σ1=σ2=σ3
2 k(2 )
2 k(2 )

2 k(1 )

σ2

σ1 σ2 σ1

Figure 4.22 Von Mises criterion

F(σij , αij ) = J2 − k 2 (χ) = 0 (75)

In the one-dimensional case with stress σe


2 2 1
σe 0 σe 0 − σe 0
−J2 = |3 |+|3 |+| 3 | (76)
1 1 1
0 − σe 0 − σe 0 − σe
3 3 3

31
which simplifies to:
1
J2 = σ2e (77)
3
To verify the condition in the yield surface:
σe = σy
| 1 (78)
J2 = k 2 (χ) = σ2y
3
Therefore, for the general case, σe can be interpreted as an effective stress.

σe = √3 J2 (79)

Normally, the isotropic hardening parameter, χ is associated with the effective


p
plastic strain rate or equivalent plastic strain, ε̇ e . This parameter is the dual
amount to the equivalent stress σe from an energy stand point, i.e. the rate of
the plastic work.
p
σ ε̇ p = σe ε̇ e (80)
Due to the associative character of plasticity:

σ ε̇ p = |s| |ε̇ p | (81)


But:

|s| = √2 J2
t (82)
|ε̇ p | = √(ε̇ p ) ε̇ p

So, substituting (79) on the left-hand-side of (80) and eq. (81) & (82) on the
right-hand-side

p t
√3 J2 ε̇ e = √2 J2 √(ε̇ p ) ε̇ p (83)

and then, the effective plastic strain is:

p 2 t
ε̇ e = √ (ε̇ p ) ε̇ p (84)
3

32
The effective plastic strain represents a measure of the amount of plastic strain
p p
accumulated during the process. In simple tension σe = σy and ε̇ e = ε̇ y so the
relationship between the effective stress and the effective plastic strain can be
obtained from the one-dimensional plastic stress-strain relationship.

Two currently used alternatives for χ are based on the total amount of effective
plastic strain eq. (84).
t
p
χ1 = ∫ ε̇ e dt (85)
0

Or the accumulated plastic work


t
χ2 = ∫ √σ ε̇ p dt (86)
0

4.2.4.2 Kinematic hardening

The kinematic hardening represents a translation of the yield surface in the


stress space, without changing the shape, size or orientation of the initial yield
surface. This hardening rule provides a way to account for the Bauschinger
effect.

The hardening parameter is a tensor α that represents the center coordinates


of the yield surface in the stress state:

F(σ, α) = f(σ, α) − k 2 = 0 (87)


For the von Mises criterion (Figure 4.23):
1 T
F(σ, α) = (s − α) (s − α) − k 2 = 0 (88)
2
where k 2 is constant.

Several rules have been proposed to describe this type of hardening. One of
them is Prager’s rule which postulates describe that the yield surface does not
change its size or its shape but translates in stress space in the direction of its
normal

33
p
dα = c dε (89)

where c > 0 and is a material constant.

Due to some inconsistencies in Prager’s rule for a general stress space, Ziegler
proposed that the translations occurs in a radial direction determined by the
vector σ − α:

dα = (σ − α)dμ (90)

where dμ is a scalar determined by the yield criteria.

σ1=σ2=σ3

Figure 4.23 Kinematic hardening for von Mises criterion

Figure 4.24 represents both hardening rules for the two-dimensional von Mises
criterion.

34
P

P: Praguer
Z: Ziegler

Figure 4.24 Kinematic hardening rules


4.2.4.3 Mixed hardening

The mixed hardening rule combines the behaviors of isotropic and kinematic
hardenings so the yield surface can both translate and expand uniformly without
a change in orientation.

In this case, there are two hardening parameters, a “kinematic vector” α that
represents the translation, and an isotropic parameter,  that represents the
expansion:

F(σ, α, χ) = f(σ − α) − k 2 (χ) = 0 (91)


The mixed hardening von Mises surface (Figure 4.25) can be expressed as:
1 T
F(σ, α, χ) = (s − α) (s − α) − k 2 (χ) = 0 (92)
2

Figure 4.25 Mixed hardening von Mises surface

35
4.2.4.4 Experimental determination of the hardening parameters

Von Mises criterion can be expressed as eq. (38).


1
J2 = k 2 = s T s (93)
2
On the other hand, an equivalent unidimensional stress comparable with the
results of a tensile test is (eq. (79))

σe = √3 J2 (94)

Then, successively, we found

√J 2 = k
σe σ2e (95)
2
σe = √3 k k= k =
√3 3
Or, using eq. (90) and generalizing the concept of equivalent unidimensional
stress for mixed von Mises Hardening
1
3 T 2
(96)
(s − α)e = [ (s − α) (s − α)]
2
and, the yield criteria can be expressed as

F(σ, α, χ) = (s − α)e − R(χ) − σy (97)

that can be identified in a (σ, ε) diagram of the tensile test (Figure 4.26), where
R(χ) is variable.

From those tests some researches (see ref. 17) infer that:

a) The kinematic stress α increases, in matrix form, with the plastic strain and
tends to saturate at a limit value α∞ .

The evolution of α can be expressed as:


p
α = α∞ [1 − e− γ ε ] (98)

which is an equation that can be justified with micromechanics. Parameters α∞


and γ depend on the material and the temperature. Plotting α as a function of

36
εp (Figure 4.26) makes it possible to identify α∞ and calculate γ on a semi-log
plot or with least squares.

Eq. (98) Eq. (99)

Figure 4.26 Experimental determination of the hardening parameters

b) The isotropic hardening stress R = σ − σy − α , increases non-linearly with


the plastic strain and tends to saturate at a value R ∞ . Similar to the expression
for α, the equation for R is:
p
R = R ∞ [1 − e− b ε ] (99)

where R ∞ and b are parameters that depend on the material and temperature
and can be compared to α∞ and γ, respectively.

37
4.3 Plasticity: Analysis with the finite element method

In a classical step by step analysis by the finite element method, the equilibrium
equation is weakly satisfied, resulting in a system of equations:

(Fint )n+1 = (Fext )n+1 (100)

Where:

(Fint )n+1 = ∫ B t σn+1 dΩ


Ω
(101)
(Fext )n+1 = ∫ N t Xn+1 dΩ + ∫ N t Tn+1 dΩt
Ω Ωt

and the sub index n + 1 represents the values in the step t n+1 . Since:

σn+1 = σ (εn+1 ) = σ(un+1 ) (102)


then

(Fint )n+1 = Fint (un+1 ) (103)

The equilibrium equation written in residual form is

r = r(un+1 ) = (Fext )n+1 − (Fint )n+1 (104)

In a nonlinear analysis, such as plastic analysis, the internal forces vector Fint
is a non-linear function in u, so the calculation for un+1 can only be performed
iteratively with a numerical method such as Newton-Raphson.

4.3.1 Newton-Raphson method

The Newton-Raphson method is used to solve a system of non-linear equations


such as the system defined with the equilibrium equations r(un+1 ) = 0.

The well-known relation that defines the Newton-Raphson method to obtain the
solution for the equation r(un+1 ) = 0 is:

r(ukn+1 )
uk+1
n+1 = ukn+1 − (105)
r ′ (ukn+1 )

38
where k represents each iteration, uk+1 k
n+1 is the value we are looking for, r(un+1 )
is the residual at the previous iteration, and r ′ (ukn+1 ) is the derivative of the
residual.

Figure 4.27 shows this method graphically for a general function f. (See Self
Evaluation Exercise 13).

Figure 4.27 Newton-Raphson’s method

4.3.2 Non-linear finite element method: Algorithm

When the Newton-Raphson method is used for the equation:

(Fext )n+1 = Fint (un+1 ) (106)

the algorithm in Figure 4.28 is obtained.

In this algorithm, it can be noted that it is necessary to solve a system of


equations A ∆u = b for each iteration. The solution cost for a non-linear problem
can be very high for large problems.

The residual vector contains both the external force and the internal force
evaluated at each iteration. Matrix A is called tangent operator and changes for
each iteration. In a modified Newton-Raphson method, matrix A is fixed for a
certain number of iterations, reducing the solution cost, but also convergence.

In the following section, it will be discussed how to evaluate the residual vector,
b, and the tangent operator, A, in an elastoplastic problem.

39
4.3.3 Evaluation of the residual vector: Integration of the elastoplastic
constitutive equations

The residual vector, b, is evaluated as:

b = r(ukn+1 ) = (Fext )n+1 − ∫ B t σkn+1 dΩ =


Ω
(107)
= (Fext )n+1 − ∫ B t (σn + ∆σk ) dΩ
Ω

where ∆σk represents the stress increment at iteration k.

For each step, the external force vector is constant. The problem, however,
consists of calculating the stress at each iteration σkn+1 so that the plastic
consistency condition is satisfied. The way to do this is by using robust
algorithms for integrating the plastic constitutive rules.

40
MAIN PROGRAM

Data input

For each step

k  0, unk 1  un
Start……………….

For each element N = 1, NUMEL


For each integration point L = 1, NINT

MATERIAL SUBROUTINE
given n and  kn 1 obtain…..
kn 1


b=r unk 1 
A=F'INT unk 1
Assembly…….....……y……………....  
YES Check Convergence
Print results
r k   r 0 ???

NO

A u=b
Solve………….

Check Convergence
unk 11  unk 1  u
k  k 1

Figure 4.28 Flowchart: Finite element algorithm for the resolution


of non-linear problems

41
4.3.3.1 Generalized trapezoidal rule
p
Starting from the known values of strain (εn , εn ) and stress (σn ), in a step t n , it
is necessary to obtain those values in a next step t n+1 .

Integrating the plastic constitutive equations between t n , and t n+1 leads to

σn+1 = σn + D ∶ ∆ε − D ∶ ∆εp = σtest


n+1 − D ∶ ∆ε
p
(108)
The increment of strain, ∆ε, is calculated from the incremental displacement
field, ∆u, obtained in the finite element method. To obtain the plastic strain
increment, ∆εp , a trapezoidal integration scheme is used.

p p ∂G ∂G
∆εp = εn+1 − εn = ∆λ [(1 − α) +α ] (109)
∂σn ∂σn+1

where the integration parameter,α varies from 0 to 1 and the plastic multiplier
is obtained by consistency, as shown in 4.2.3.

The stresses σn+1 are calculated and therefore the internal force vector.
Separating the stress, σn+1 , into an elastic part, σtest
n+1 , and a plastic part,
(D ∶ ∆εp ), it can be shown that the stress value at t n+1 is obtained in two
phases. The first phase, which corresponds to the elastic stress-strain
relationships, transforms the initial stress state σn into an elastic predictive
stress state, σtest
n+1 . In the second phase, this predictive state returns to the yield
surface, F(σn+1 ) = 0, thereby restoring the plastic condition. The plastic return
can also be divided into two steps: in the first step the stress is projected onto
the direction of the initial yield ∂G/ ∂σn; the stresses obtained are then projected
onto the yield surface in the direction of the final yield ∂G/ ∂σn+1 (Figure 4.29).

42
ntest
1

F(n )  0
G
(1  )
n
n1
G

n

F(n1 )  0

Figure 4.29 Generalized trapezoidal rule

According to this interpretation, the variable parameter α would give the relative
weight assigned to the directions of the initial and final yield and the integration
would only be exact in the case of radial loading where the normal does not
vary from t n to t n+1 . If α = 0, the algorithm is termed explicit and called forward
Euler algorithm. For α > 0 the algorithm is implicit, and in particular for α = 1
and associative plasticity is called implicit backward Euler method. This choice
of α was initially proposed as a generalization of the radial return algorithm for
plasticity models other than von Mises.

4.3.4 Evaluation of the tangent operator. Consistent tangent modulus

Tangent operator A is:

̂ k B dΩ
A = Fint (ukn+1 ) = ∫ B t D (110)
Ω

where

∂σ(εkn+1 )
̂k =
D (111)
∂εkn+1
represents the consistent tangent modulus, which is strongly influenced by the
constitutive laws and the algorithms used for integration.

43
Let us calculate the expression of the tangent modulus consistent with the von
Mises plastic perfect model with the radial return algorithm.

According to the Self Evaluation Exercise 14, using the radial return, the
following expression is obtained:
test
R sn+1
σn+1 = +pI
test test
(112)
√sn+1 : sn+1

where
1
p = tr(σtest
n+1 )
test
sn+1 = σtest
n+1 − p I σtest
n+1 = σn + D: ∆ε (113)
3
In initial form and without writing the sub index n + 1:

R sijtest
σij = + p δij
test s test
(114)
√spq ij

The consistent tangent modulus is:


∂σij
̂ ijkl =
D =
∂εkl
R ∂sijtest test
1 test test −12 test
test
∂spq (115)
test
= test test [√spq spqtest − sij (s s ) (2 spq )]
spq spq ∂εkl 2 pq pq ∂εkl
∂p
+ δ
∂εkl ij
Since, by definition,

σtest
ij = σn,ij + Dijkl (εkl − εn,kl ) (116)

where σn,ij and εn,kl represent the values of stress and stress for t n .

Taking the derivative of this expression:

∂σtest
ij 2
= Dijkl = G(δik δjl + δil δjk ) + (K − G) δij δkl (117)
∂εkl 3

44
∂p 1 ∂σtest
ii
= = K δkl
∂εkl 3 ∂εkl
∂sijtest ∂σtest
ij ∂p 2
= − δij = G (δik δjl + δil δjk − δij δkl )
∂εkl ∂εkl ∂εkl 3
̂ ijkl leads to:
Substituting these derivatives in D

2GR 1 sijtest test


skl
̂ ijkl
D = [I − δij δkl − ] + K δij δkl (118)
test s test ijkl 3 test s test √s test s test
√spq pq √s rs rs tu tu

Where:
1
Iijkl = (δik δjl + δil δjk ) (119)
2

45
SELF EVALUATION EXERCISES

EXERCISE 1

The three-bar truss shown below is loaded by a vertical force P. All bars have
the same cross-sectional area, A. The bars are made of an elastic-perfectly
plastic material with elastic modulus E and yield stress σ0 . The load P is first
increased to the plastic limit Pp and then is unloaded to zero. Afterward, P is
increased in the reversed direction until all three bars yield again in
compression and then is unloaded again to zero.

a) Determine the elastic limit load Pe+ and the plastic limit load Pp+ in the
initial loading.
b) Determine the elastic limit load Pe− and the plastic limit load Pp− in the
reversed loading.
c) Determine the residual stresses and strains of the bars and the residual
horizontal and vertical displacements at load point C at the end of the
loading history.

Figure 4.30

Solution:

The basic equilibrium and compatibility equations are:

46
P
σ1 = σ2 σ1 + σ3 =
A
ε1 = ε2 4 · ε1 = ε3

Using the elastic constitutive laws, we obtain the elastic solution of the problem.

In this case, the results are in the plane (σ1 , σ2 ). The elastic domain has a
square shape bounded by the lines σ1 = σ0 , σ1 = −σ0 , σ2 = σ0 , σ2 = −σ0
(Figure 4.31).

2
0

0

1
0

0

Figure 4.31 Elastic domain

a) Determine the elastic limit load and the plastic limit load

Since the stress in bar 3 is the largest, this bar yields first. The elastic limit load
is obtained by equating σ3 to σ0 .

Pe+ 5
= σ0
A 4
at P = Pe+ the stress, strain and displacements are
1 1
σ1 = σ2 = σ0 σ3 = σ0 ε1 = ε2 = ε0 ε3 = ε0
4 4
u=0 v = L ε0

47
where u and v are the horizontal and vertical displacements respectively at the
point of application of the load.

Once bar 3 yields, the structure loses a degree of indetermination and it is


transformed into a statically determined structure.

For P > Pe+ , ∆σ3 = 0 since the material behaves as a perfectly plastic material.
From the equilibrium equations:
P
σ1 + σ0 = σ1 = σ2
A
The plastic limit load is reached when:

σ1 = σ2 = σ0

so

Pp+
= 2 σ0
A
In Pp+ the stress, strain and displacements are:

σ1 = σ2 = σ3 = σ0 ε1 = ε2 = ε0 ε3 = 4 ε0

u=0 v = 4 L ε0

The accumulated plastic strain is:


p p p
ε1 = ε2 = 0 ε3 = 4 ε0 − ε0 = 3 ε0

Unloading

When the load P decreases from Pp+ to zero, the material is unloaded
elastically. The incremental stresses are related to the load increments ∆P
through the elastic solution:
1 ∆P 4 ∆P
∆σ1 = ∆σ2 = ∆σ3 =
5 A 5 A
This leads to:

σ1 = σ2 = σ0 + ∆σ1 σ3 = σ0 + ∆σ3

48
∆P = −Pp+ = − 2 σ0 A the residual stresses are obtained as:

3 3
σ1 = σ2 = σ0 σ3 = − σ0
5 5
The residual strains are:
∆σ1 2 3
ε1 = ε2 = ε0 + = ε0 − ε0 = ε0
E 5 5
∆σ3 8 12
ε3 = 4 ε0 + = 4 ε0 − ε0 = ε
E 5 5 0
12
u=0 v= L ε0
5
b) Determine the compression elastic load and plastic limit loads

Applying an increment of load ∆P = +Pe− the additional stresses that have to be


added to the residual stresses are obtained:
3 3 1 Pe−
σ1 = σ2 = σ0 + ∆σ1 = σ0 +
5 5 5 A
3 3 4 Pe−
σ3 = − σ0 + ∆σ3 = − σ0 +
5 5 5 A
Since σ3 is closer to −σ0 , bar 3 yields first. With σ3 = −σ0 :
Pe− 1
= − σ0
A 2
at P = Pp− :

Pp−
= −2σ0
A
And

σ1 = σ2 = σ3 = −σ0 ε1 = ε2 = −ε0 ε3 = −4 ε0

u=0 v = −4 L ε0

49
c) Residual values

When unloading occurs ∆P = − Pp− , the residual stresses and strains are:

3 3 3 12
σ1 = σ2 = − σ0 σ3 = σ0 ε1 = ε2 = − ε0 ε3 = − ε
5 5 5 5 0
12
u=0 v=− L ε0
5
Figure 4.32 represents a plot in the load-vertical displacement P − v plane with
the load path including loading, unloading and residual values.

Figure 4.32 Plot P-v

50
EXERCISE 2

Plot the load path in the stress-strain plane from the load in Example 1.

Figure 4.33 Load path in the stress space

Solution:

Using the elastic domain in stresses (Figure 4.33), during the elastic path the
load follows the straight line OP. At P, σ3 reaches the yield stress σ0 and due to
the perfect plasticity the stress in bar 3 is constant.

During the elastoplastic phase, the load follows path PQ until Q, where yielding
of all the structure is reached.

From point Q there is unloading parallel to OP . Point R is obtained at the


intersection of this unloading line with line AA’ (the line that represents
equilibrium without loads, i.e.: σ1 + σ3 = 0. This intersection point corresponds
to the residual stresses previously calculated (σ1 = 3/5 σ0 ; σ3 = −3/5 σ0 ).

If there is reloading in compression from point R, bar 3 yields first (point S)


followed by elastoplastic behavior until point T.

Unloading occurs along path TU and point U represents the residual final
stresses (σ1 = −3/5 σ0 ; σ3 = 3/5 σ0 )..

The strain path is represented in the same way in Figure 4.34.

51
Figure 4.34 Loading path in the strain space
During the elastic phase the path is along line OA. Past this point, the strains
are plastic and continue along the same line until point B:

ε1 = ε0 ε3 = 4 ε0

Upon unloading, the path reaches point C which corresponds to the residual
strains (ε1 = 3/5 ε0 ; ε3 = 12/5 ε0 ).

The following final reload and unload steps follow path AD and DE respectively.

It can be observed that, as expected, the strain path follows along a straight line
corresponding to the compatibility equation of the structure 4 ε1 = ε3 .

52
EXERCISE 3

Solve the structure from Exercise 1 using the finite element method.

Solution:

The stiffness matrix in the global coordinate system of a truss structure is:

c2 sc −c 2 −sc
EA s2 −cs −s 2 )
K= (
L c2 sc
sim. s2
where c = cos α and s = sin α.

For each element, the stiffness matrix is:

3 √3 3 √3
− −
4 4 4 4
1 √3 1 0 0 0 0
EA − EA 1 0 −1
K1 = 4 4 4 K2 = ( )
2L 3 √3 L 0 0
− sim. 1
4 4
1
(sim. 4 )
3 √3 3 √3
− −
4 4 4 4
1 √3 1
EA − −
K3 = 4 4 4
2L 3 √3
4 4
1
(sim.
4 )
Assembling the overall system of equations and applying the boundary
conditions, the equilibrium equation at node C (load node) is obtained.

53
3
EA 4 0 uC
0
( )= ( ) (v )
−P L 1 C
0 1+
4
Assuming unit values for E and A, the solution is:
uC = 0
4
vC = − PL
5
Starting from these global values, the stresses and strains in each bar can be
calculated in their local coordinates as
1 uj − ui
εij = (cos α sin α) ( v − v )
L j i

and then:
2
ε1 = ε2 = P = σ1 = σ2
5
4
ε3 = P = σ3
5
In the numerical solution process, the load is applied incrementally and the
elastic stresses are calculated until the stress in any bar reaches over the yield
stress σ0 . From that point, iterations are made until equilibrium is achieved.

Therefore, if an initial load P = σ0 is applied, the stresses are:

σ1 = 0.4 σ0 σ3 = 0.8 σ0

In this case, the yield stress has not been reached. Increasing the load by ∆P =
0.5 σ0, the total applied load is P = 1.5 σ0 and

σ1 = 0.6 σ0 σ3 = 1.2 σ0

The stress in bar 3 now exceeds the yield stress. Since the material behavior is
perfectly plastic, if the force σ3 = σ0, equilibrium is not achieved:

3
0.6 σ0 + σ0 ≠ σ0
2

54
To achieve equilibrium, subtract the excess stress value in bar 3, 0.2 σ0 . This
leads to P = 1.3 σ0 which produces the following stresses:

σ1 = 0.52 σ0 σ3 = 1.04 σ0

Verify that the excess stress in bar 3 has decreased.

Repeating the same iterative process with the excess values (0.04 σ0 ), the
solution is reached when the difference between the calculated stress σ3 and
σ0 is less than a specified residual value.

EXERCISE 4

A material yields with the following stress state

σ1 = −σ0 ; σ2 = −σ0 ; σ3 = 0

Using the von Mises and Tresca criteria, find the stresses in tension and pure
shear as well as the yield load Py , for the following load path σ1 = P, σ2 =
2P, σ3 = 3P.

Solution:

 von Mises criterion

F(σ) = J2 − k 2 = 0

Knowing that the material yield criteria is σ1 = σ0 , σ2 = σ0 , σ3 = 0 the value of


k can be calculated recalling the definition of J2 from Chapter 1
1
J2 = (I12 − 3I2 )
3
I1 = σ1 + σ2 + σ3
I2 = σ1 σ2 + σ1 σ3 + σ2 σ3

Plugging in the values

55
1
J2 = σ20 = k 2
3
σ0
k=
√3
In a simple tensile test there is only one component of stress, σ1

σ1 = σ1 σ2 = 0 σ3 = 0

so the tensile strength can be calculated similarly as before


1 1
J2 = σ12 = k 2 = σ20
3 3
Therefore

σ1 = σ0

In pure shear the stresses are:

σ1 = τ1 σ2 = 0 σ3 = −τ1

so the shear strength is:

1 2
σ20 σ0
(1 + 1)σ1 = ⇒ σ1 = τ1 =
2 3 √3
For the given load path, plugging in the yield criterion leads to a yield load Py

1 σ20 σ0
(1 + 1)Py2 = ⇒ Py =
2 3 √3
 Tresca criterion
1
max ( |σi − σj |) = k
i≠j 2
- k calculation
1
σ =k
2 0
-Tensile strength

56
σ1 = σ1 , σ2 = 0, σ3 = 0
1 1
σ1 = σ0 ⇒ σ1 = σ0
2 2
- Shear strength

σ1 = τ1 , σ2 = 0, σ3 = −τ1
1 1 1
2τ1 = σ0 ⇒ τ1 = σ0
2 2 2
-Yield load Py

1 σ0 σ0
(3 Py − Py ) = → Py =
2 2 2

EXERCISE 5

Determine h for a von Mises yield surface F and the stress-strain (σ − ε) rule
from Figure 4.35

dσn
H=
dεp

(a) (b)
Figure 4.35

Solution:

57
In the uniaxial tensile test
σ̇ 2
σ̇
3 3
σ̇ σ̇ σ̇
σ̇ = ( 0 )= + − = σ̇ oct + ṡ
3 3
0 σ̇ σ̇
( ( − )
3) 3
The projection of σ̇ on n gives the deviatoric stress tensor ṡ since σ̇ oct is
perpendicular to n. (See Figure 1.18 Appendix in Chapter 1).

4 2 1 2 2
|σ̇ n | = nt σ̇ = √ |σ̇ | + 2 |σ̇ | = |σ̇ |√
9 9 3

According to Figure 4.35.b:

|dσn | 2 dσ 2
h= =√ = √ H
|dεp | 3 dεp 3

where H is measured in the one-dimensional diagram after subtracting the


elastic strain (Figure 4.35.b).

58
EXERCISE 6

Determine the flow rule associated with von Mises yielding.

Solution:

As shown in section 4.2.1.1 the von Mises yield surface can be written as:

F(σ) = J2 − k 2 = 0

Using an associative flow rule leads to:


∂F
dεp = dλ = dλ s
∂σ

i.e., the direction of the plastic strain increment is parallel to the deviatoric stress
tensor (Figure 4.36).

Expressing the above equation in component form leads to the Prandtl-Reuss


equation
p p p p p p
dεx dεy dεz dγyz dγzx dγxy
= = = = = = dλ
sx sy sz 2 syz 2 szx 2 sxy

Figure 4.36 Flow rule for von Mises criterion

59
EXERCISE 7

Determine the flow rule associated to Tresca’s criterion.

Solution:

As shown in 4.2.1.2., this criterion is expressed as:

F(σ) = max(|σi − σj |) − 2 k = 0
i≠j

Similar to the von Mises criterion, the plastic strain increment is purely deviatoric
since the walls of the yield surface are parallel to the hydrostatic axis (σ1 = σ2 =
σ3 ).

In the deviatoric plane, the directions of dεp have to be determined for the six
planes and six corners of the Tresca regular hexagon (Figure 4.37).

For example, for the case σ1 > σ2 > σ3 (plane AB).

Figure 4.37 Flow rule for Tresca’s criterion

F(σ) = σ1 − σ3 − 2 k = 0
p p p
dεp = (dε1 dε2 dε3 ) = dλ (1 0 −1)

For the corners, the direction of the plastic flow is undefined and lies between
the cone formed between the normal vectors of neighboring faces.

60
EXERCISE 8

Obtain the expression for the plastic multiplier dλ for a perfectly plastic material
using the associative flow rule for the von Mises criterion. Assume that in the
elastic domain the behavior is linear and isotropic.

Solution:

Substituting the constitutive law into the consistency condition:


∂F ∂F ∂F ∂F
dF = σ =0 ⇒ Dijkl dεkl = dλ Dijkl
∂σij ij ∂σij ∂σij ∂σkl

which leads to:


∂F
D dε
∂σij ijkl kl
dλ =
∂F ∂F
Dpqrs
∂σpq ∂σrs

For a linear elastic and isotropic material the elastic tensor Dijkl is

2
Dijkl = (K − G) δij δkl + G(δik δjl + δil δjk )
3
Therefore:

∂F 2 ∂F ∂F
Dijkl = (K − G) ( δij ) δkl + 2 G
∂σij 3 ∂σij ∂σkl

∂F ∂F 2 ∂F ∂F ∂F ∂F
Dijkl = (K − G) ( δij ) ( δkl ) + 2 G =
∂σij ∂σkl 3 ∂σij ∂σkl ∂σkl ∂σkl
2
2 ∂F ∂F ∂F
= (K − G) ( δij ) + 2 G
3 ∂σij ∂σij ∂σij

∂F 2 ∂F ∂F
Dijkl dεkl = (K − G) ( δij ) dεkk + 2 G dε =
∂σij 3 ∂σij ∂σkl kl

61
2 ∂F ∂F
= (K − G) ( δij ) dεkk + 2 G dε
3 ∂σij ∂σij ij

From the von Mises criteria:


∂F
= sij
∂σij

Therefore
∂F
δ = sij δij = sii = 0
∂σij ij

(no volume changes in the deviatoric planes).

Substituting these results in the expression for the plastic multiplier:


2 G sij dεij
dλ =
2 G sij sij

EXERCISE 93

A long thin-walled circular tube with average diameter D and wall-thickness t is


subjected to an internal pressure p1 and an external pressure p2 , as shown in
Figure 4.38. The ends of the tube are closed. Let p2 = r p1 , r ≥ 0. The tube
yields when p1 = p0 , p2 = 0

a) Find the limit pressure p1 = py , at which the tube yields, in terms of p0


and r for the case r > 0 according to Tresca and von Mises criteria.
b) Determine the plastic strain increments when the tube yields using an
associative flow rule
p
ε̇ c
ε̇ p = [ε̇ pa ]
p
ε̇ r

3 Example from “Structural Plasticity”, W. F. Chen y H. Zhang, Springer-Verlag, 1991

62
Figure 4.38

Solution:

Assuming a membrane-type stress state, the circumferential stress using


Laplace’s equation is:
D D
σc = (p1 − p2 ) = (1 − r)p1
2t 2t
By equilibrium, the axial stress is:
D
σa = p
4t 1
For r = 0, the limit pressure is p0 and the corresponding stresses:
D D
σ1 = σc = p σ2 = σa = p σ3 = 0
2t 0 4t 0
Using the Tresca criterion, the limit value for k is:
1 D
k = (σ1 − σ3 ) = p0
2 4t
The two yield criteria for a biaxial state of stress can be expressed as
1
- Tresca: max ( |σi − σj |) = k
i≠j 2

- von Mises F(σ) = J2 − k 2 = 0

63
Limit pressure 𝐩𝐲 , and plastic strain increment analysis

- von Mises criterion

Substituting σa , and σc in the von Mises expression:

p1 D 2 p1 D 2 2
p1 D 2
( ) + 4( ) (1 − r) − 2 ( ) (1 − r) = 3 k 2 = σ2e
4t 4t 4t
and plugging in the value of k and p1 = py leads to:

4t 3
p1 = py = k√ 2
D 4r −6r+3

4t
If r = 0 ⇒ p0 = k
D

and then

3
p1 = py = √ 2 p
4r −6r+3 0

 Plastic strain increment: the ε̇ p direction follows that of s

1 2
− σa + σc 1 3
3 3 2σ − σ 2(1 − r) − − 2r
2 1 1 c a 1 p1 D 2 1 p1 D 2
s= σ − σ = [ 2σa − σc ] = 1−1+r = r
3 a 3 c 3 −σ − σ 3 2t 3 6 t 3
a c

1
σ −
1
σ [ −
2
+ r ] [ 2 + r]

[ 3 a 3 c]
3 − 4r
p ̇
ε̇ = λ [ 2r ]
−3 + 2r
- Tresca criterion

Tresca’s criterion varies with different values of r, so it is necessary to consider


each case separately:

Let us take this order for all cases

64
ε̇ p = [ε̇ pc p
ε̇ a
p
ε̇ r ]
1
(1) 0≤r≤
2

σ1 = σc σ2 = σa σ3 = 0

For this case, the criterion becomes:


1 D D
k = σc ⇒ (1 − r) p1 = p0
2 4t 4t
and the limit pressure can be calculated as
p0
p1 = py =
1−r
 Plastic strain increment:

σc − σr = σy
p
dεij = (dεpc dεa
p p
dεr ) = (1 0 −1) dλ

1
(2) ≤r≤1
2

σ1 = σa σ2 = σc σ3 = 0

 Tresca’s criterion
1 D D
k = σa ⇒ p1 = p0
2 8t 4t
 Limit pressure

p1 = py = 2 p0

 Plastic strain increment

σa − σr = σy
p
dεij = (dεpc dεa
p p
dεr ) = (0 1 −1) dλ

65
(3) r≥1

σ1 = σa σ2 = 0 σ3 = σc < 0

 Tresca’s criterion
1 D D D
k = (σa − σc ) ⇒ p1 − (1 − r) p1 = p0
2 8t 4t 4t
 Limit pressure
p0
p1 = py =
1
r−
2
 Plastic strain increment

σa − σc = σy
p
dεij = (dεpc dεa
p p
dεr ) = (−1 1 0) dλ

66
EXERCISE 10

Obtain the expression for the plastic strain increment for the von Mises loading
surface considering associative plasticity and an isotropic hardening rule with
the slope of the effective stress-effective plastic strain curve H
dσe
H= p
dεe

Solution:

The von Mises criterion with isotropic hardening can be written as

F(σij , χ) = J2 − k 2 (χ) = 0

where k is not a constant.

Using the effective stress concept, σe , from section 4.2.4.1, the criterion can be
written as:
1
F(σij , χ) = σ2e − k 2 (χ) = 0
3
which leads to:

σe
k(χ) =
√3
Therefore, the von Mises criterion in terms of the effective stress is:
1 1
F(σij , χ) = sij sij − σ2e (χ)
2 3
Applying the consistency condition leads to:
∂F 2
dF = 0 ⇒ dσij − σe dσe = 0
∂σij 3

or, expressed in other terms:

67
p 2 p
sij Dijkl (dεkl − dεkl ) = σe H ε̇ e
3
But:

p ∂F
dεkl = dλ = dλ skl
∂σij

2
sij Dijkl dεkl = sij [(K − G) δij δkl + G(δik δjl + δil δjk )] dεkl = 2 G sij dεij
3
p 2
sij Dijkl dεkl = sij Dijkl skl dλ = 2 G sij sij = 2 G σ2e
3

p 2 ∂F ∂F 2 2
ε̇ e = dλ√ = dλ√ sij sij = dλ σe
3 ∂σij ∂σij 3 3

Rearranging this equation leads to:


2 G sij deij
dλ =
4
(3G + H) σ2e
9
Therefore:

p ∂F 2G
dεij = dλ = sij skl dekl
∂σij 4
(3G + H) σ2e
9
This expression could have been obtained as a function of the stress increment
as:

∂F p 4 sij dσij
dσij = dλ H σ2e ⇒ dλ
∂σij 9 4
H σ2e
9
Then:

p ∂F 1 1 ∂J2 1
dεij = dλ = sij skl dσkl = sij dσkl = sij dJ2
∂σij 4 2 4 2 ∂σkl 4 2
H σe H σe H σe
9 9 9

68
EXERCISE 114

Start with the same vessel as Exercise 9, and load it with the same pressure.
The ends of the tube are closed. Assume a von Mises loading function with
isotropic hardening rule and the relationship εep = a σ3e , where a is a constant.
p p
Determine the plastic strains (εa , εc ) at the end of the following two loading
paths:

a) Proportional loading path

(p1, p2 ) = (0, 0) → (p1, R p1 )

b) Non-proportional loading path

(p1, p2 ) = (0, 0) → (0, R p1 ) → (p1, R p1 )


p p
where (εa , εc ) are the axial and circumferential plastic strains, respectively, and
p1 and R are constants, R = 3/2.

Solution:

As shown in Example 9, assuming a membrane stress state leads to:


D D
σa = p σc = (p − p2 )
4t 1 2t 1
Then:
1
J2 = (σ2a + σ2c − σa σc ) σ2e = 3 J2 = σ2a + σ2c − σa σc
3
1 1
dJ2 = (2σa − σc ) dσa + (2σc − σa ) dσc
3 3
From Exercise 10:

4 Example from “Structural Plasticity”, W. F. Chen y H. Zhang, Springer-Verlag, 1991

69
p 1
dεij = sij dJ2
4
H σ2e
9
and for this problem:
dσe 1 4 4
H= p = ⇒ H σ2e =
dεe 3 a σ2e 9 27a

Then:

p 9a
dεij = [(2σa − σc ) dσa + (2σc − σa ) dσc ]sij
4

a) Proportional loading path (0, 0) → (p1, R p1 )


3
p2 = R p1 = p1
2
D D
σa = p σc = − p = −σa
4t 1 4t 1
sa = σa sc = σc

Then:

p 9 27a
dεij = 6 a σa dσa sij = σ dσ s
4 2 a a ij
which in component form is:

p 27a 2 27a D 3 2
dεa = σ dσ = ( ) p1 dp1
2 a a 2 4t

p 27a 2 27a D 3 2
dεc =− σ dσ = − ( ) p1 dp1
2 a a 2 4t
Integrate the above expressions to obtain the plastic strains at the end of
the load path as:

p 27a D 3 p1 2 9a D p1 3
εa = ( ) ∫ p1 dp1 = ( )
2 4t 0 2 4t

70
p 27a D 3 p1 2 9a D p1 3
εc =− ( ) ∫ p1 dp1 = − ( )
2 4t 0 2 4t

b) Non-proportional loading path (0, 0) → (0, R p1 ) → (p1, R p1 )


First increasing p2 from 0 to R p1 :
D
σa = 0 σc = − p dσa = 0 dσc ≠ 0
2t 2
1 2
sa = − σc sc = σc
3 3
and

p 9a 9a
dεij = 2 σc dσc sij = σc dσc sij
4 2
In component form, this is:

p 9a 1 3a D 3 2
dεa = σc dσc (− σc ) = ( ) p2 dp2
2 3 2 2t

p 9a 2 D 3 2
dεc = σc dσc ( σc ) = −3a ( ) p2 dp2
2 3 2t
At the end of path p1 = 0, p2 = R p1 , the equations become:

p 3a D 3 Rp1 2 a 3D p1 3
εa = ( ) ∫ p2 dp2 = ( )
2 2t 0 2 4t

p D 3 Rp1 2 3D p1 3
εc = −3a ( ) ∫ p2 dp2 = −a ( )
2t 0 4t

and

1 2 3 D 2 2
J2 = σc = ( ) p1
3 4 2t
In the second part, p1 increases from 0 to p1 and p2 is constant and equal to
R p1 . The stresses are:

71
D D D D
σa = p σc = p1 − R p1 = 2 σa − R p1
4t 1 2t 2t 2t
When the value of σc decreases, it is necessary to check the loading criteria. If
J2 is equal at the beginning of the step than the current J2 , we can plug in the
values of stresses σa and σc from above into the expression for J2 from the
beginning of the example:
2
1 2 D D 3 D 2 2
[σ + (2σa − Rp1 ) − σa (2σa − Rp1 )] = ( ) p1
3 a 2t 2t 4 2t

Or:
D
σ2a − σa R p1 = 0
2t
Solving for σa , leads to:
D 3D D
σa = 0 or σa = R p1 = p1 > p1
2t 4t 4t
This result implies that the tube is either in an unloading state or in an elastic
loading state and there is no new plastic strain. Therefore, at the end of this
path, the strains are:

p 27a D p1 3 p D p1 3
εa = ( ) εc = −27a ( )
2 4t 4t

72
EXERCISE 12

Write an algorithm for the Newton-Raphson method.

Solution:

1. k = 0 ukn+1 = u0

2. Calculate r(ukn+1 )

3. If |r(ukn+1 )| < TOL → END

4. Calculate r′(ukn+1 )

k
r(ukn+1 )
5. Obtain ∆u =
r′(ukn+1 )

6. Evaluate uk+1 k
n+1 = un+1 − ∆u
k

7. Next k k = k+1

8. Go to 2

73
EXERCISE 13

Solve the equation f(u) = sin u = 0 by Newton-Raphson’s method starting at


u0 = 2.

Solution:

Using the algorithm shown in the previous example:

k uk f(uk ) f ′(uk ) −f(uk )/f ′(uk )'

0 2 9.09E-01 -4.16E-01 2.19E+00

1 4.1850398633 -8.64E-01 -5.03E-01 -1.72E+00

2 2.4678936745 6.24E-01 -7.82E-01 7.98E-01

3 3.2661862776 -1.24E-01 -9.92E-01 -1.25E-01

4 3.1409439123 6.49E-04 -1.00E+00 6.49E-04

5 3.1415926537 -9.10E-11 -1.00E+00 -9.10E-11

6 3.1415926536 1.23E-16 -1.00E+00 1.23E-16

Notice that, as expected, we are approaching the solution u = π and that


function f here is playing the role of a residual.

74
EXERCISE 14

Develop the algorithm to calculate the stresses at time step t n+1 , by applying
the minimum distance point integration scheme for a von Mises yield surface
without hardening with an associative flow rule.

Solution:

For the algorithm with α = 1, and an associative flow rule


∂F
σn+1 = σtest
n+1 − D ∆λ
∂σn+1

Since von Mises criterion is purely deviatoric, the direction of the plastic return
according to an associative flow rule in σn+1 , is normal to the hydrostatic axis
and therefore radial in the deviatoric plane. (Figure 4.39)

radial return

Figure 4.39 Algorithm of the radial return

The volumetric stress is not affected by yielding and remains unchanged.

75
This algorithm, known as the radial return algorithm, follows these steps:

1. Calculate σtest
n+1 = σn + D: ∆ε

2. Obtain of the volumetric 1


p = tr(σtest
n+1 )
component of σtest
n+1 3

3. Obtain the deviatoric test


sn+1 = σtest
n+1 − p I
component of σtest
n+1

test test
4. Calculate A = sn+1 : sn+1

5. Check for plasticity by comparing with von Mises circle radius R

 If A ≤ R2 → Elastic process σn+1 = σtest


n+1 → END

R
 If A > R2 → Plastic process σn+1 = s test +pI → END
√A n+1

76
SUMMARY OF MAIN IDEAS

1. Our study is limited to isotropic materials. That means that mechanical


properties around a point in the domain do not depend on the local axis
orientation.

2. Another limitation is the assumption of isothermal plasticity that allows us


to avoid some advanced problems, where temperature influence is
determinant, in spite of their practical importance.
3. We also assume that plastic deformation is “instantaneous” without any
time dependence. This has its origin in the beginning of plastic theory
when researchers were interested mainly in metals at room temperature.
If we need to incorporate time in our model, it can be added to viscoelastic
models (Recall Bingham body in chapter 2) leading to a branch called
“viscoplasticity” that is also out of this chapter scope.
4. In most examples, following the experiments with metals, we shall assume
isochoric behaviour, that is, no change of volume, which means that
plastic deformations are due to shear stresses.
5. This chapter presents only the fundamentals of the so-called “Flow theory”,
using incremental computations because the final state depends on the
load trajectory. Here it is worthwhile to emphasize that, although the
notation used for increments is the dot over the letter representing stress
or strain (reserved in viscoelasticity and dynamics to represent time
derivative), this has to be interpreted as a symbolic time. In fact, we
choose a first order homogeneous function to model the functional
relationship between stress and strain increments to avoid the influence of
any change in time scale. Valanis and other authors developed in the 70’s
a theory called “Endochronic plasticity” motivated by the goal of giving
some physical meaning to the time used in computations that is applied
occasionally.
6. The importance that has to follow the stress or strain trajectory adds value
to the representation in the corresponding spaces. Due to the assumed
isotropic stress space, for instance, it can be reduced to Haig-Westergard
representation (see Appendix in Chapter 1) in the principal stresses,
where stress tensor can be reduced to a (3x1) vector. Moreover, if the
volume changes are not produced by plastic strains, the octahedral stress
plays no role and we can limit ourselves to the representation in isometric
projection in a deviatoric plane. This will not be possible in concrete or soil
mechanics as we are going to discuss below.

77
7. All possible stress combinations that maintain the elastic response of the
body define an “Elastic domain” E in the stress space where the material
behaviour follows the Linear Elasticity equations.
8. The elastic domain is bounded by a “Yield Surface” 𝐅(σ) that separates
the elastic from the elastoplastic stress states. Since the XIX century, the
definition of that surface has been one of the main research objectives in
plasticity theory. Von Mises criterion is a well-established alternative for
metals but in concrete or soil modelling, where dilatancy is coupled with
shear strains, more appropriated Yield Locus are being used.
9. When the point representing the stress-state has a trajectory that reaches
or crosses the yield surface, plastic strains appear. Under very reasonable
hypothesis of continuity of response and linear relation between stress and
strain rates, it is possible to establish a “Flow Rule” that depends on the
existence of a new surface 𝐆(𝛔), called “Plastic Potential”, whose normal
vector defines the direction of the plastic strain increment.

10. Sometimes it is possible to assume G(σ) = F(σ), in which case we shall


speak of “Associated flow rule” or Associative plasticity. Then, the
direction of the plastic flow increment ε̇ will be along the normal vector n to
the yield surface F and we can draw the increments of strain in the same
plot that the stress trajectory.

11. To reproduce Bauschinger effect, the yield surface depends on a


parameter that allows its isotropic growth, its translation in the stress space
following the evolution of the loading path or a combination of both. The
first alternative was favoured in pre-computer times to simplify calculations
while the second, a combination or even a multisurface approach as
defended by Zenon Mroz, has been made possible by modern
computationally based numerical methods.

12. The key to numerical models is the development of criteria to follow the
nonlinear path and the definition of a stiffness matrix that depends on all
the concepts defined above.

78
REFERENCES

1. R. Hill “The mathematical theory of plasticity” Oxford 1950

2. J. Baker & Heyman “Plastic design of frames” 2 vol. Cambridge UP 1969

3. C.R. Calladine “Engineering plasticity” Pergam Press 1969

4. J. Heyman. “Coulomb’s memoir on Statics”. Cambridge U.P. 1972

5. W.F. Chen “Limit analysis and soil plasticity” Elsevier 1975

6. L. Katchanov “Elements de la théorie de la plasticité” MIR 1975

7. B.G. Neal “The plastic methods of structural analysis” Science


paperback Chapman Hall 1977

8. J.H. Atkinson; P.L. Bransby “The mechanics of soil. An introduction to


critical state soil mechanics” Mc Graw 1978

9. B.P. Hughes “Limit state theory for reinforced concrete structures”


Pitman 1980

10. W.F. Chen & A.F. Saleeb “Constitutive equations for engineering
Materials vol1.” Wiley 1982

11. J. Heyman “The masonry arch”. Ellis Horwood 1982

12. W. Szczepinski; J. Szlagowski. “Plastic design of complex shape


structures” Ellis Horwood 1985

13. W.F. Chen, X.L. Liu “Limit analysis in soil mechanics” Elsevier 1990

14. D. Muir Wood “Soil behaviour and critical state soil mechanics”
Cambridge UP 1990

15. W.F. Chen & .Zhong “Structural Plasticity Theory, problems and CAE
software” Springer 1991

16. G.A. Maugin “The thermomechanics of plasticity and fracture”


Cambridge U.P. 1992

17. J. Lemaitre “A course on damage mechanics” 2nd ed. Springer. 1996

79
18. A. Muttoni, J. Schwartz, B. Thürlimann “Design of Concrete Structures
with Stress Fields” Birkhäuser 1997

19. J.C. Simo, T.J. Hughes “Computational inelasticity” Springer 1998

20. P. Nielsen “Limit Analysis and concrete plasticity” 2nd ed. CRC press
1999

21. T. Belytschko; W.K. Lin; B. Moran. “Nonlinear Finite Elements for


Continua & Structures” Wiley 2000

22. M. Kojic, K. J. Bathe “Inelastic Analysis of solids & structures” Springer


2005

80

S-ar putea să vă placă și