Você está na página 1de 14

Numerical analysis of damage for prediction of fracture initiation in

deep drawing
Ravindra K. Saxena, P.M. Dixit

Department of Mechanical Engineering, Indian Institute of Technology, Kanpur 208016, India


a r t i c l e i n f o
Article history:
Received 29 October 2010
Received in revised form
7 April 2011
Accepted 7 April 2011
Keywords:
Deep drawing
Ductile fracture
Continuum damage mechanics
Elasto-plastic analysis
Finite element method
a b s t r a c t
The sheet metal may have inherent voids/imperfections present because of preprocessing. These voids/
imperfections grow under the applied load resulting into nal fracture. The occurrence of ductile
fracture is often a limiting factor in metal forming processes. Prediction of the initiation of ductile
fracture allows a prior modication of the process which can result in a defect-free nal product with
nancial savings. The intensity of voids is often represented by introducing a variable called damage.
This paper deals with the prediction of fracture initiation in deep drawn cup using Lemaitres
continuum damage mechanics model. The damage is incorporated in the constitutive equation through
the principle of strain equivalence. The damage is evaluated using the damage growth law proposed by
Lemaitre. An in-house 3D nite element formulation is employed for the analysis. The maximum cup
height at which the fracture initiates is determined using the critical damage criterion. A parametric
study of the maximum cup height is carried out to investigate the inuence of material, geometric and
other process parameters.
& 2011 Elsevier B.V. All rights reserved.
1. Introduction
Manufacturing of sheet metal parts by means of press working
is a cost effective process since it eliminates expensive machining
and welding operations giving a better quality nished product.
Further, it enables production of components at a very high rate.
Among the manufacturing of these sheet metal parts, deep
drawing is an extensively used press working process [1]. Among
the deep drawing processes, circular cup drawing is the most
basic process (Fig. 1). In circular cup drawing, the ange region of
the cup is compressed in the tangential direction and stretched in
the radial direction. The stress condition in the wall and bottom
regions, however, is different. The stress elds exhibited in
non-circular deep drawing processes are more complicated as
the mechanics of deformation differs very appreciably from that
observed in the circular cup drawing process. Among the
non-circular deep drawing processes, box drawing is the most basic
process (Fig. 2). In box drawing process, the metal ow rates are
different in the straight walls than in the box corner region. This
results in an uneven material distribution around the box wall.
Traditionally, the formability of a fracture-free sheet metal
product is predicted using the forming limit diagram. The forming
limit of a sheet metal is dened to be a state at which the
localized necking initiates (i.e., when the existing state of strain
becomes unstable) as the sheet is formed into the product. This
limit is conventionally represented as a curve in the 2D strain
space of major and minor strains. This curve is called the forming
limit diagram (FLD). Initial attempt to measure the principal
strains at necking was made by Keeler and Backhofen [2] by
stretching sheets over hemispherical punches and recording the
strains at the discontinuous increase by a grid marking technique.
From this experiment, one can construct only the right side of
FLD. This technique of strain measurement was extended to other
experiments by Goodwin [3] so as to obtain the left side of the
FLD. Early attempts to construct theoretical FLDs were based on
the necking criteria, in terms of the major and minor strains,
proposed by Swift [4], Hill [5] and Marciniak and Kuczyn ski [6] for
the plane stress conditions. Hills [5] criterion assumed that
localized necking takes place along the direction of zero rate of
extension. Swifts [4] criterion, which assumed that necking takes
place whenever the yield stress increment caused by a change in
the strain state is less than the induced increment in the
equivalent stress, corresponds to diffuse necking. Whereas Swift
[4] and Hill [5] assumed that the sheet metal is homogeneous in
nature, Marciniak and Kuczyn ski [6] assumed a pre-existing
groove, perpendicular to the major stress axis, to obtain a necking
criterion when both the principal strains are positive. (This is
normally called the MK model.) Hutchinson and Neale [7]
extended the MK model by assuming the groove at an angle to
the major stress axis to obtain a necking criterion when the
principal strains are either tensile or compressive.
In practice, the FLD of a material is employed in predicting the
formability of a product by estimating the critical strains in the
product experimentally using the grid marking method and then
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/nel
Finite Elements in Analysis and Design
0168-874X/$ - see front matter & 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.nel.2011.04.003

Corresponding author. Tel.: 91 512 2597094; fax: 91 512 2597408.


E-mail address: pmd@iitk.ac.in (P.M. Dixit).
Finite Elements in Analysis and Design 47 (2011) 11041117
comparing these critical strains with the FLD. However, there
have been some attempts to use the FLD of a material to predict
localized necking in a deep drawn product using the nite
element method (FEM). Evangelista et al. [8] proposed a mod-
ication in the MK model [6] (in the form of an inhomogeneity
factor) to construct the FLD of SAE 1010 steel using FEM. Then,
this FLD was employed to predict the maximum cup height at
localized necking in circular and square cups using FEM with shell
elements. However, there is no experimental validation of the
predicted maximum cup height at localized necking. Chen et al.
[9] determined the FLD of a magnesium alloy AZ31 using the
experimental method of Keeler and Backhofen [2]. Since the
magnesium alloy has a poor formability at room temperature,
the FLD is constructed at elevated temperatures. Then, they used
this FLD and the PAM-STAMP nite element software (with four-
noded shell elements) to predict the localized necking in square
cups of AZ31 material. They obtained the maximum cup height at
which the simulated values of the major and minor strains lie
below the FLD. They observed a good agreement between the
predicted maximum cup height with the experimental results up
to moderate temperatures. They also conducted the parametric
studies of the maximum cup height with respect to the three
parameters: the punch prole radius, the die corner radius and
the forming temperature.
The FLD of a material depends on the strain path. As a result,
for a single material, there are different FLDs for different paths.
To overcome the drawback of the strain path dependence of the
FLD, Arrieux et al. [10] and Stoughton [11] proposed an alternate
stress-based FLD which is path-independent in the stress space.
However, it is difcult to construct the stress-based FLDs experi-
mentally due to practical difculties involved in measuring the
principal stresses on a deformed sheet. As a result, normally the
stress-based FLDs are constructed theoretically. Arrieux et al. [12]
used the MK model [6,13] to construct the FLD of mild steel
in the stress space, instead of in the strain space, by employing
the fourth order RungeKutta method. They employed this FLD
(which they called as the forming limit stress curve or FLSC in
short) to predict the occurrence of localized necking in deep
drawing of a square cup using FEM with triangular at elements.
They found that the necking region is in the cup corner above the
punch radius region. Further, the location of this region matches
with the location of the strain localization region determined by
them experimentally.
Instead of using the FLD in the strain or stress space
(constructed experimentally or numerically) for predicting the
formability of a product, an alternative approach is to use an
appropriate ductile fracture criterion. Quite a few semi-empirical
criteria have been employed for predicting the fracture initiation
in metal forming processes. Takuda et al. [14] used four different
semi-empirical ductile fracture criteria (Cockroft and Latham [15],
Brozzo et al. [16], Clift et al. [17], Oyane et al. [18]) in nite
element simulation of circular cup drawing process for two
aluminum alloys (A1100 and A2024) to predict the maximum
cup height without fracture. They also conducted experiments
and observed that the fracture location predicted by these four
criteria matches with the experimental location of fracture. They
further observed that, whereas the fracture occurs in the cup wall
region near the cup bottom in A1100 material, in A2024 material,
the fracture location is either near the cup bottom or mid-way
between the ange and the cup bottom. About the latter fracture
location, Takuda et al. [14] observed that the fracture is not
preceded by localized necking and, hence, may not be predictable
by the FLD. The main limitation of using any of the semi-empirical
ductile fracture criteria for the prediction of fracture initiation is
that they are based on an inadequate knowledge of the physics of
the ductile fracture process in metals.
It is well-known that ductile fracture in metals occurs mainly
due to void nucleation, growth and nally coalescence into a
micro-crack. Three broad approaches have emerged that try to
r
pp
t
0
D
d
Punch
Z
r
dp
D
p
X
Die
Blankholder
D
b
Fig. 1. Schematic diagram of circular cup drawing process (D
b
blank diameter,
t
0
blank thickness, D
p
punch diameter, D
d
die opening diameter, r
pp
punch
prole radius, r
dp
die prole radius).
r
dc
r
pc
r
pp
S
d
S
p
S
b
r
dp
t
0
Punch
X
Y
Z
Blankholder
Die
Fig. 2. Schematic diagram of square cup drawing process (S
b
blank size, t
0
blank thickness, S
p
punch size, S
d
die size, r
pp
punch prole radius, r
dp
die prole radius,
r
pc
punch corner radius, r
dc
die corner radius).
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1105
predict ductile fracture initiation on the basis of the phenomena
of void nucleation, growth and coalescence:
porous plasticity model of Gurson [19,20],
combination of void nucleation, growth and coalescence mod-
els [2123],
continuum damage mechanics model of Lemaitre [2426].
In porous plasticity model, the material with voids is idealized
as a porous plastic material. Thus, the constitutive equation is
derived from the plastic potential of porous material. Based on
Bergs [19] model of dilatational plasticity, Gurson [20] proposed
a plastic potential for porous plastic materials. Later Tvergaard
[27,28] modied this plastic potential to account for the void
interaction. In this model, the rate of change of void volume
fraction
_
f is considered as the sum of the void nucleation rate
_
f
nucleation
and the void growth rate
_
f
growth
. The void nucleation rate
is assumed to depend on the equivalent plastic strain rate, which
is called as the cracking model. The void growth rate is related to
the hydrostatic part of the plastic strain rate tensor. In this
approach, fracture initiation is characterized by the critical value
of void volume fraction (f
c
) which should be determined from a
suitable void coalescence criterion, but is usually obtained
experimentally [29]. Doege et al. [30] used the Gursons model
[20] and the ABAQUS code (with eight-noded brick elements) to
study the void volume distribution in circular cup drawing of X5
CrNi 1810 stainless steel.
Thomason [23] combined the results of de-cohesion model of
Goods and Brown [21] on void nucleation, those of Rice and Tracy
[22] on void growth (a single spherical void in uniform stress and
strain rate elds) and his own on void coalescence (i.e., on plastic
instability of the inter-void matrix) to arrive at a fracture criterion
in the form of a graph of fracture strain versus the hydrostatic
part of stress. Thus, in this model, the effects of void nucleation
and growth are incorporated not in the constitutive equation but
in the fracture criterion itself. A limitation of this approach is that
while integrating the void growth equations of Rice and Tracy
[22], Thomason [23] assumed that the principal directions of the
strain rate tensor remain xed in the direction throughout the
deformation path. This, in general, is not valid for the nite
deformation and/or rotation.
Kachanov [31] described the deterioration of the material
strength due to voids by introducing a scalar variable D called
damage, which is dened as the surface density of void traces
in any plane of volume element. Based on this idea, and using
the theory of continuum thermodynamics, Lemaitre [2426]
proposed a continuum damage mechanics (CDM) model. In this
model, the plastic potential of a damaged material consists of two
parts: (i) rst part corresponding to the yielding and hardening
and (ii) second part corresponding to the damage. The rst part is
usually obtained from an appropriate yield function using the
principal of strain equivalence. This principal states that the
deformation behavior of a damaged material can be described
by the same constitutive equation as that of the virgin material if
the Cauchy stress is replaced by the effective stress, which is
the Cauchy stress divided by the factor (1D). The derivative of
the rst part with respect to the stress gives the constitutive
equation (i.e., the ow rule). The derivative of the second part
with respect to the thermodynamic force (corresponding to the
damage) gives the damage growth law. The second part of the
plastic potential is not well-established in the literature. Lemaitre
[2426] proposed a simple expression for the second part which
leads to the linear damage growth law (i.e., the damage depend-
ing linearly on the equivalent plastic strain) that does not account
for the void nucleation. Dhar et al. [32] proposed a damage
growth law based on the experimental results of Leroy et al.
[33] on the measurements of area void fractions at different strain
levels in tension test. This accounts for both the void nucleation as
well as growth. In continuum damage mechanics approach,
fracture initiation is characterized by the critical value of damage
(D
c
) which is determined either by a void coalescence criterion or
experiments.
Quite a few researchers have used the CDM model of Lemaitre
[2426] with or without his damage growth law to predict ductile
fracture in deep drawing processes. Elgueta [34] used the damage
growth law of Lemaitre [24,25] and the in-house nite element
code (with degenerate shell elements) to simulate the drawing of
the central sector of a steel kerosene stove. The critical value of
damage (D
c
) was determined experimentally using the micro-
hardness measurements in tension test. He evaluated the damage
distribution at the end of the drawing process both in single as well
as two-stage drawing operations. He observed that the maximum
damage in the two-stage drawing was less than D
c
but in the single-
stage, it reaches the level of D
c
. The maximum damage occurs in the
wall region close to the bottom. His experimental results on the
single-stage drawing showthat the experimental location of fracture
matches with the predicted location of the region where D reaches
the critical value D
c
. Khelifa et al. [35] used a modied version of the
Lemaitres damage growth law [24] and the ABAQUS/explicit code
(with eight-noded brick elements) to simulate the drawing of a
anged circular cup of aluminum alloy AL5754. They used Hills [36]
anisotropic yield function to obtain the constitutive relation of the
material. However, the damage is assumed to be isotropic as in the
Lemaitres model. They obtained the damage distributions at various
levels of the punch displacement. They did not determine the critical
value of the damage. Instead, the damage is allowed to reach the
maximum value of 1. The maximum value occurs in the wall region
close to the bottom. Their experimental results show that the
experimental location of fracture matches with the predicted
location of the region where D reaches the value of 1. Fan et al.
[37] also used the Lemaitres damage growth law [25] and the
ABAQUS/explicit code (with eight-noded brick elements) to simulate
the drawing of a square cup of mild steel. They obtained the damage
distributions at various levels of the punch displacement. They
found that the maximum damage occurs in the cup corner region
closer to the ange. They also conducted experiments and observed
that the experimental fracture location matches with the predicted
location of the region where D reaches the critical value. They also
conducted the parametric study of damage growth (during the
punch travel) with respect to the blank-holder force and friction
coefcient. They observed that the damage increases with the
blank-holder force and friction. The values of the blank-holder force
used by them are quite small. As a result, the wrinkling also occurs.
The above literature shows that even though the ductile fracture
prediction in circular and square cup drawing processes has been
attempted using various methods, there is a need to study the effects
of various geometric and material parameters on the maximum cup
height that can be drawn without fracture. This is the objective of the
present work. Since the use of FLD or semi-empirical ductile fracture
criteria has certain limitations as mentioned earlier, and the pre-
dicted fracture location using the Lemaitres CDM model (along with
his damage growth law) matches with the experimental results, the
latter is used for this purpose. Updated Lagrangian formulation is
employed to develop the incremental nite element equations using
eight-noded brick elements. Incremental logarithmic strain measure
is used. The incremental stress is made objective by evaluating it in a
frame rotating with the material particle. The material is assumed to
be elasto-plastic strain hardening yielding according to the von Mises
yield criterion. The strain hardening behavior is modeled by a power
law. The damage is incorporated in the constitutive equation through
the principal of strain equivalence. Further, the damage growth law
proposed by Lemaitre [24,25] is used. The critical value of damage
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1106
(available in the literature) is used for predicting the fracture
initiation. Body forces are neglected and due to small accelerations
inertial forces are not included. Modied NewtonRaphson iterative
technique is used to solve the non-linear incremental nite element
equations. The code is rst validated by comparing the predicted
punch variation and thickness strain distribution with the experi-
mental results available in the literature.
2. Formulation
2.1. Continuum damage mechanics
As stated in the introduction, a ductile fracture occurs mainly
due to micro-void nucleation, growth and nally coalescence into a
micro-crack. The void growth also affects the constitutive relation of
the material. A realistic model for the prediction of ductile fracture
must include the following three things:
the effect of micro-voids on the constitutive relations of the
material,
an evolution law for micro-voids which incorporates both the
void nucleation as well as the growth and
a condition for micro-crack initiation based either on a
coalescence model or experimental observations.
As stated earlier, in this work, the continuum damage mechanics
(CDM) model proposed by Lemaitre [2426] is employed for the
prediction of fracture initiation. In this model, the behavior of
a material containing micro-voids is described by introducing
an additional variable called the damage variable. If the material
behavior is isotropic, then this variable is scalar [2426]. For isotropic
material, the damage at a point (at a time) is dened as the area void
fraction in a plane at that point at that time:
D lim
DA-0
DA
v
DA
1
where DA is the innitesimal area around that point and DA
v
is the
area of void traces contained in DA. Since, this is an incremental
formulation, the damage at time t is denoted by
t
D and the
incremental damage at time t by
t
DD.
2.1.1. Incremental stressstrain relation for damaged material
When the temperature change is not signicant, the plastic
potential of a damaged material is given by [2426]
jf
t
s
ij
,
t
R;
t
D f
d

t
Y;
t
D 2
where f and f
d
are the plastic potentials associated with yielding
and damage, respectively, and
t
s
ij
is the Cauchy stress tensor at
time t. The quantities
t
Y and
t
R are the dissipative parts of the
thermodynamic forces at time t corresponding, respectively, to
the rates of the damage variable
t
D and the hardening variable
t
p.
The expression for
t
Y is given by [2426]

t
Y
t
s
2
eq
t
R
v
2E1
t
D
2
3
t
R
v

2
3
1n 312n
t
s
m
t
s
eq
!
2
4
where E is the Youngs modulus and n is the Poissons ratio,
t
s
m

1
3
t
s
ii
5
is the hydrostatic part of
t
s
ij
(called the mean stress) and
t
s
eq

3
2
t
s
0
ij
t
s
0
ij
r
6
is the equivalent stress at time t. Here,
t
s
0
ij
is the deviatoric part of
t
s
ij
. The quantity
t
s
m
=
t
s
eq
is called the triaxiality. For the case of
strain hardening,
t
p is identied as the equivalent plastic strain
t
e
pL
eq
and is given by
t
p
t
e
pL
eq

X
t
De
pL
eq
,
t
De
pL
eq

2
3
t
De
pL
ij t
De
pL
ij
r
7
Here, the sum is to be carried over all the increments up to time t
and
t
De
pL
ij
is the plastic part of the incremental logarithmic strain
(dened in Section 2.3) in an earlier increment. Note that,
t
e
pL
eq
is
the work conjugate variable corresponding to
t
R. Therefore, for f,
often
t
e
pL
eq
is chosen as one of the independent variables rather
than
t
R. Thus, f is considered as a function of
t
s
ij
,
t
e
pL
eq
and
t
D.
The principle of strain equivalence states that the deformation
behavior of a damaged material can be represented by the same
constitutive relation as that of the virgin material (i.e., material
free from defects) if the Cauchy stress
t
s
ij
is replaced by the
effective Cauchy stress
t
s

ij
[2426]:
t
s

ij

t
s
ij
1
t
D
8
Using this principle, the plastic potential f for a material yielding
according to the von Mises criterion is given by
f
t
s
ij
,
t
e
pL
eq
;
t
D
t
s

eq

t
s
y
9
where
t
s

eq
is the equivalent stress corresponding to
t
s

ij
dened
by a relation similar to Eq. (6) and s
y
is the variable yield stress of
the material. The dependence of s
y
on
t
e
pL
eq
(i.e., the hardening
curve of the material) is assumed to be given by a power law:
t
s
y
H
t
e
pL
eq
s
y

0
K
t
e
pL
eq

n
10
where s
y

0
is the initial yield stress and K and n are the harden-
ing parameters. Using the associated ow rule where the plastic
potential is given by Eq. (9), employing the additivity of elastic
and plastic parts of the incremental logarithmic strain and assum-
ing the elastic behavior to be linear, the incremental elasto-plastic
stressstrain relationship for isotropic material becomes [38,39]
t
Ds
ij

Z
t Dt
t
t
C
EP
ijkl
d
t
De
pL
kl
11
where the fourth order elasto-plastic constitutive tensor
t
C
EP
ijkl
is
given by
t
C
EP
ijkl
2m d
ik
d
jl

n
12n
d
ij
d
kl

9m
t
s
0
ij
t
s
0
kl
23m
t
H
0

t
s
2
eq
8
<
:
9
=
;
1
t
D 12
Here, d
ij
is the Kroneckers delta, m is the shear modulus and
t
H
0
is
the slope of the hardening curve, modeled by Eq. (10). The tensor
t
Ds
ij
on the left side of Eq. (11) must be an objective stress tensor.
To make it objective, the stress updating procedure discussed in
Ref. [40] is followed.
2.2. Damage growth law
As stated in the introduction, the damage growth law is
obtained as the derivative of f
D
with respect to (
t
Y). Lemaitre
[2426] has proposed the following form for f
D
:
f
D

t
Y;
t
D
1
1
t
D
1
s1

t
Y
S
s 1
13
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1107
where s and S are the material parameters. When the temperature
change is not signicant, the value of s is taken as unity [25]. To
express the other material parameter S in terms of the measurable
quantities, we assume that, at large values of the equivalent plastic
strain, the material behaves like a perfectly plastic material. Then,
using Eq. (13), the incremental damage growth lawcan be expressed
as [25]
t
DD
D
c
t
R
v
e
C
e
D

t
De
pL
eq
14
where D
c
is the value of damage at fracture initiation called the
critical damage value, e
C
is the value of the equivalent plastic
strain at fracture initiation, i.e., when the damage reaches the
critical value D
c
and e
D
is the value of the equivalent plastic strain
at which the damage initiates (i.e., the threshold value). The
material constants D
c
, e
C
and e
D
are determined from experiments
[41,42]. Alves et al. [43], Bennani and Lauro [44] and Bonora et al.
[42] have proposed numerical techniques to determine these
material constants from the experimental data.
2.3. Governing equations
In updated Lagrangian formulation, the governing equations
consist of (a) incremental straindisplacement relations, (b) incre-
mental stressstrain relations, (c) incremental equations of motion.
The incremental logarithmic strain measure, used in the present
formulation, is dened by [45]
t
De
L
ij
ln
t

i
d
ij
no sum over i 15
where
t

i
are the principal values of the incremental right stretch
tensor
t
U
ij
. The square of this tensor is dened by
t
U
2
ij

t
F
T
ik
t
F
kj
,
t
F
ij
d
ij

t
Du
i,j
16
where
t
Du
i,j
denotes the derivative of the incremental displacement
vector
t
Du
i
at time t with respect to the position vector
t
x
j
at time t.
Note that, in Eq. (15), the components are with respect to the
principal axes of
t
U
ij
. (This is used as the material frame in the
present work.) The incremental elasto-plastic stressstrain relation-
ship is described in Section 2.1.1.
The incremental form of the equilibrium equation is not found
to be convenient for the nite element formulation. Instead, an
integral form of the equilibrium equation at time t Dt is used. It
is given by the following virtual work expression [45]:
Z
t Dt
V
t Dt
s
ij
d
t Dt
e
ij
d
t Dt
V
t Dt
R 17
Here,
t Dt
V is the domain,
t Dt
R is the virtual work of the external
forces and
t Dt
s
ij
is the Cauchy stress tensor, all at time t Dt.
Further, d
t Dt
e
ij
represents the virtual linear strain tensor
corresponding to the virtual displacement vector d
t Dt
u
i
at time
t Dt. Since, the conguration at time t Dt is unknown, this
expression is transformed to an integral over the known domain
at time t (i.e.,
t
V):
Z
t
V
t Dt
t
P
ij
d
t Dt
t
e
ij
d
t
V
t Dt
R 18
Here,
t Dt
t
P
ij
is the 2nd PiolaKirchoff stress tensor and d
t Dt
t
e
ij

is the virtual GreenLagrange strain tensor. This virtual work


expression is further simplied by decomposing
t Dt
t
P
ij
as the
sum of
t
s
ij
and
t
DP
ij
, decomposing d
t Dt
t
e
ij
into the linear and
non-linear parts and neglecting the higher order terms. Then, it
becomes
Z
t
V
t
DP
ij
d
t
De
ij
d
t
V
Z
t
V
t
s
ij
d
t
DZ
ij
d
t
V
Z
t
V
t
s
ij
d
t
De
ij
d
t
V
t Dt
R
19
Approximating
t
DP
ij
as the product of the elastic-plastic con-
stitutive tensor and the incremental linear strain tensor, it is
written as
Z
t
V
t
C
EP
ijkl t
De
kl
d
t
De
ij
d
t
V
Z
t
V
t
s
ij
d
t
DZ
ij
d
t
V
Z
t
V
t
s
ij
d
t
De
ij
d
t
V

t Dt
R 20
where the tensors
t
De
ij
and
t
DZ
ij
are given by
t
De
ij

1
2

t
Du
i,j

t
Du
j,i
,
t
DZ
ij

1
2

t
Du
k,it
Du
k,j
21
2.4. Finite element formulation
When the solid nite elements are used, the sheet is treated as a
3D domain, which is the more realistic way to model the process.
The simultaneous contact with the die on the bottom side of sheet
and with the punch and blank-holder on the top side of the sheet is
naturally taken care of without any particular strategy. However, the
aspect ratio of the elements should be small to avoid deterioration
of the solution [46]. The domain is discretized into a number of
eight-noded brick elements and the incremental displacement eld
is approximated over each element by
t
fDug
t
Du
x
t
Du
y
t
Du
z
8
>
<
>
:
9
>
=
>
;

t
F
t
fDug
e
22
Here, the vector
t
fDug
e
contains the nodal values of the incremental
displacement components at the elemental nodes and the matrix
t
F contains the shape functions that are known functions of the
coordinates. Substitution of Eq. (22) into the virtual work expression
(20) and assembly over all the elements leads to the following
algebraic equation:
t
K
t
fDug
t
ff g
t Dt
fFg 23
Here,
t
fDug is called the global (incremental) displacement vector
and
t
K,
t
ff g and
t Dt
fFg, respectively, denote the global coefcient
matrix (at time t), global internal force vector (at time t) and global
external force vector (at time t Dt), the expressions for which are
given in any standard text on non-linear FEM [45]. Since
t Dt
fFg
t
fFg
t
fDFg and
t
fFg
t
ff g 24
Eq. (23) can be written as
t
K
t
fDug
t
fDFg 25
The solution of Eq. (25) represents only an approximate
solution to the governing equations, because of the linearization
and approximation involved in arriving at expression (20). To
minimize the error of the approximate solution, the modied
NewtonRaphson algorithm [45] is used. Here, the equation
t
K
t
fDug
i

t Dt
fRg
i1
for i 1,2, . . . 26
where
t Dt
fRg
i1

t Dt
fFg
t Dt
ff g
i1
for i 2,3, . . .
t Dt
fRg
0

t
fDFg 27
is solved till the (global) unbalanced force vector
t Dt
fRg
i
becomes sufciently small. After solving Eq. (26), the net incre-
mental displacement vector
t
fDug
X
i
t
fDug
i
28
is used to compute rst the incremental logarithmic strain tensor
using Eqs. (15) and (16) and then the incremental stress tensor
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1108
using Eq. (11). The integration in Eq. (11) is carried out using
the Euler forward integration scheme. Finally, the Cauchy stress
tensor at time t Dt is updated using the procedure described in
Ref. [40].
2.4.1. Evaluation of damage
At each convergent equilibrium state (i.e., at time t), the
Cauchy stress tensor
t
s
ij
is decomposed into the hydrostatic
(or mean) part
t
s
m
and the deviatoric part
t
s
ij
0
. From the
deviatoric part, rst the equivalent stress
t
s
eq
is evaluated using
Eq. (6) and then the triaxiality (
t
s
m
=
t
s
eq
) is calculated. From the
triaxiality, the quantity
t
R
n
is evaluated using Eq. (4). Then, at the
end of increment, the incremental logarithmic strain is calculated.
As its elastic part is small, the whole of the incremental logarith-
mic strain is considered as its plastic part. From the plastic part,
the equivalent plastic strain increment
t
De
pL
eq
is calculated using
Eq. (7). From
t
R
n
and
t
De
pL
eq
, the incremental damage
t
DD is
evaluated using the damage growth law given by Eq. (14).
Finally, the damage is updated as
t Dt
D
t
D
t
DD 29
The fracture initiation is checked using the condition
t Dt
D D
c
.
If this condition is satised for at least one nite element, the
analysis is stopped.
3. Results and discussion
An in-house 3D FE code based on the formulation of Section 2
is used. The code is rst validated for the simulation of square and
circular cup drawings by comparing the predicted punch load
variation and thickness strain distribution with the experimental
results available in the literature. Since the damage data is not
available for these materials, the validation study is carried out
without incorporating the damage. Then, the code is used for the
prediction of fracture initiation in square and circular cup drawing.
The geometry of the problem for circular cup drawing from
circular blank is shown in Fig. 1 and for square cup drawing from
square blank in Fig. 2. The punch, die and blank-holder are assumed
to be rigid. Due to the symmetry in geometry and loading, only
one-quarter of the blank is considered for the analysis. Sliding friction
is assumed at sheet-die interface. The coefcient of friction is
assumed to be constant. Sticking friction is assumed at sheet-punch
interface, which is a fair assumption considering the large compres-
sive force generated at the sheet-punch interface. As the deformation
progresses in the sheet, the position of the nodes and the related
boundary conditions get affected. Therefore, the boundary conditions
are updated in each increment. Flags are assigned to the nodes
according to their position, to designate the change in the boundary
conditions.
The algorithm for the contact problem assumes that the
tooling geometry is represented by a set of straight line segments
and circular/elliptical arcs. Such a representation of tooling
geometry makes it possible to employ simple formulae to control
whether a node crosses the tooling surface (straight line segment
or arc). The nodes, if at any particular increment, penetrate the
punch, die or blank-holder proles, are checked for penetration
into these proles and are given specied displacement to project
them onto the punch/die/blank-holder prole. The whole incre-
ment is repeated by considering these nodes at the specied/
projected points on the prole.
At the beginning of each incremental step, the contact state and
the displacements eld obtained from the previous step are used to
initiate the described iterative procedure. The converged results of
the iterative procedure at a given contact state are checked to
determine whether the contact state needs to be changed. If it
appears that for ith contact node, the sign of contact force has
changed, the node is declared to be free and if jth non-contact node
penetrates the punch/die/blank-holder prole, the node is projected
on the prole. Taking into account such a change of contact
conditions, the incremental step is repeated. The transition to the
next increment is accomplished only if there is no change of contact
state after such a recalculation. The blank-holder force (F
b
) is
assumed to be uniformly distributed over the sheet. The total
blank-holder force is applied incrementally for a rst fewincrements.
3.1. Validation
As stated earlier, the FE code is rst validated for the simulation of
square cup drawing. First, the predicted punch load variation and
thickness strain distributions are compared with the experimental
results of Ref. [47]. The geometric dimensions used in the square
cup drawing experiment are blank size110.0 mm, blank thick-
ness0.86 mm, punch size40.0 mm, die size42.5 mm, die pro-
le radius5 mm, punch prole radius5 mm, punch corner
radius3.2 mm, die corner radius4.45 mm [47]. The material used
is Aluminum-killed steel [47] with Youngs modulus211.744 GPa,
Poissons ratio0.291, stressstrain curve: s
y
172574:5e
pL
eq

0:269
(stress in MPa). The other process parameters are (i) blank-holder
force F
b
20.0 kN and (ii) friction coefcient m 0:04 [47]. Fig. 3
shows the comparison of the predicted punch load variation (with
punch displacement) with the experimental data [47]. There seems
to be a reasonably good agreement between the two.
Figs. 4 and 5 show the thickness strain distributions in Alumi-
num-killed steel sheet along the X- (or Y-) direction and the diagonal
direction of the square cup. For comparison purpose, the experi-
mental results [47] are also shown in the gures. Figs. 4 and 5 show
that the thinning of the sheet is most severe in the clearance space.
Further, the magnitude of the thickness strain is maximum along
the diagonal direction compared to the transverse direction. There-
fore, the potential fracture initiation site in square cup is along the
diagonal direction near the die throat region where the thinning is
maximum. The trends of the thickness strain distributions are
consistent with the experimental results of Ref. [47].
Next, the FE code is validated for the simulation of circular cup
drawing. The predicted punch load variation is compared with the
experimental result of Ref. [48]. The geometric dimensions used
in the circular cup drawing experiment are blank diameter
110 mm, blank thickness0.74 mm, punch diameter50 mm,
Fig. 3. The variation of punch load with punch displacement in square cup
drawing.
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1109
die opening diameter52.5 mm, die prole radius8 mm, punch
prole radius8 mm [48]. The material used is Aluminum-killed
steel [48] with Youngs modulus200 GPa, Poissons ratio0.3,
yield strength (s
y

0
162 MPa, stressstrain curve: s
y
562
0:00941e
pL
eq

0:266
(stress in MPa). The other process parameters
are (i) blank-holder force F
b
20.0 kN and (ii) friction coefcient
m 0:17 [48]. Fig. 6 shows the comparison of the predicted punch
load variation (with punch displacement) with the experimental
data [48]. There seems to be a reasonably good agreement between
the two.
3.2. Typical results
First, the FE code is used for the analysis of damage up to
fracture initiation in square cup drawing. The geometric dimen-
sions used for the analysis are given in Table 1. The material used
for the analysis is forming steel [34], whose properties are given
in Table 2. The analysis is performed with a total blank-holder
force of F
b
15.0 kN to avoid wrinkling. The coefcient of fric-
tion used is m 0:05. The number of nite elements used for
the analysis is 1390. The size of the element is 1.5 mm along the
symmetry direction of the blank. There are two elements along
the thickness direction.
The growth of damage with punch displacement is shown in
Figs. 710. In these gures, the values of D/D
c
are plotted on
the deformed congurations, where unity represents the critical
damage i.e., DD
c
. It is observed that, rst, the damage initiates at
the cup corner near the cup bottom because the initial plastic
deformation takes place in this region (Fig. 7). With a further
punch displacement, the damage zone grows in size at the cup
bottom and it also initiates at the region in contact with the die
throat. The zone of maximum damage shifts from the corners at
the cup bottom to the corners at the die throat with a further
increase in the punch displacement (Figs. 8 and 9). Ultimately, the
two damage zones merge in the region between the cup bottom
and the contact with the die throat. The damage reaches the
Fig. 5. The variation of thickness strain along diagonal direction in square cup
drawing.
Fig. 6. The variation of punch load with punch displacement in circular cup
drawing.
Table 1
Geometric dimensions.
Circular cup Square cup
Blank diameter 140 mm Blank size 124 mm
Blank thickness 1 mm Blank thickness 1 mm
Punch diameter 85 mm Punch size 70 mm
Die opening diameter 88 mm Die size 74 mm
Die prole radius 8 mm Die prole radius 5 mm
Punch prole radius 7 mm Punch prole radius 8 mm
Punch corner radius 10 mm
Die corner radius 12 mm
Table 2
Material properties of forming steel and AA6060-T5.
Forming steel AA6060-T5
Youngs modulus (E) 210 GPa 70.8 GPa
Poissons ratio (n) 0.3 0.3
Initial yield stress (sy
0
) 276 MPa 108 MPa
Hardening coefcient (K) 421 MPa 231 MPa
Hardening exponent (n) 0.8 0.1807
Damage parameters
e
D
0.04 0.06
e
C
0.56 0.11
D
c
0.32 0.8
Fig. 4. The variation of thickness strain along transverse direction in square cup
drawing.
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1110
critical value at a punch stroke of 17.01 mm at the cup corners
near the die throat, for the element number 157, at a depth of
9.5 mm from the die surface (Fig. 10). The predicted location of
fracture initiation is in good agreement with the experimental
results on steel about the fracture location [37] (Fig. 11) as well as
the zone of strain localization [12]. Further, the predicted zone of
maximum damage is consistent with the zone of maximum
damage of Ref. [37] and the zone of localized necking obtained
using the stress-based FLD [12]. It is observed that, at the at
bottom of the cup and at the ange portion of the cup corners, the
total damage value is very small at the initiation of fracture.
Fig. 10 shows that the maximum cup height that can be achieved
without fracture is 17.01 mm.
Fig. 12 shows the damage growth with punch displacement for
a representative element (element number 157) in which the
damage reaches the critical value (Fig. 10). It shows that the
damage initiates at the punch displacement of 8.4 mm and
increases till the punch displacement of 11.01 mm. After this,
the damage growth is small up to a punch displacement of
12.7 mm. Again, the damage starts growing up to a punch
displacement of 14.41 mm except for a small region of constant
value at 13.1 mm. The damage growth is again small from a
punch displacement of 14.4116.31 mm. Finally, it reaches the
critical value at the punch displacement of 17.01 mm.
Since the damage depends on the triaxiality (i.e., the ratio of
the mean stress to the equivalent stress) and the equivalent
plastic strain (through Eqs. (4) and (14)), to understand the
behavior of Fig. 12, the graphs of the variations of triaxiality
and equivalent plastic strain with punch displacement are
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Element No. 157
Fig. 7. Damage distribution at a punch displacement of 2.6 mm.
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Element No. 157
Fig. 8. Damage distribution at a punch displacement of 6.01 mm.
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Element No. 157
Fig. 9. Damage distribution at a punch displacement of 13.6 mm.
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Element No. 157
Fig. 10. Damage distribution at a punch displacement of 17.01 mm.
Fig. 11. Experimental result for sheet fracture in square cup.
Fig. 12. The variation of damage with punch displacement for element number
157.
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1111
constructed for the representative element (element number
157). They are shown in Figs. 13 and 14, respectively. Initially,
the element is in the ange region and therefore, the tensile stress
in the radial direction is more than the sum of the compressive
stresses in the circumferential and thickness directions. Thus, the
triaxiality is positive. However, it starts decreasing and continues
to decrease up to the punch displacement of 8.31 mm, except
for a small increase at the punch displacement of 5.81 mm.
It decreases because the circumferential compressive stress
increases as the element gets drawn in the ange region and
moves toward the die throat. It starts increasing and becomes
positive at a punch displacement of 9.81 mm. It continues to
increase up to the punch displacement of 14.41 mm except for a
small decrease at the punch displacement of 10.21 mm. This
increase in the triaxiality of the element is due to the stretching
of the sheet between the die and punch. From this point onward,
it again decreases up to the fracture initiation. It becomes
negative at the punch displacement of 16.31 mm. It decreases
again because the circumferential compressive stress increases as
the element moves into the die cavity away from the die throat.
On the other hand, the equivalent plastic strain in the represen-
tative element continuously increases with punch displacement up
to the fracture initiation (Fig. 14). This is because the element gets
plastically deformed continuously as it rst gets drawn in the ange
region up to the die throat and then gets stretched after it enters the
die cavity. The equivalent plastic strain reaches the threshold value
e
D
0:04 (i.e., the value at which the damage initiates) at the punch
displacement of 8.4 mm. This explains why the damage is zero
till the punch displacement of 8.4 mm in Fig. 12. In spite of the
continuous increase in the equivalent plastic strain, the damage
growth is small between the punch displacement of 11.01 and
12.7 mm because the triaxiality is very small. From the punch
displacement of 14.41 to 16.31 mm, the damage growth is again
small because of the decrease in the triaxiality. Beyond the punch
displacement of 16.31 mm, even though the triaxiality continues to
decrease, its magnitude increases. Since, the damage growth
depends on the square of the triaxiality, it starts increasing beyond
the punch displacement of 16.31 mm.
Next, the code is used for the analysis of damage up to fracture
initiation in circular cup drawing. The geometric dimensions used
for the analysis are given in Table 1. The material used for the
analysis is forming steel [34] whose material properties are given
in Table 2. The analysis is performed with a total blank-holder
force of F
b
18.0 kN to avoid wrinkling. The coefcient of friction
used is m 0:05. The number of nite elements used for the
analysis is 1390. The growth of damage with punch displacement
is shown in Figs. 1518. Fig. 15 shows that the damage in circular
cup drawing initiates at the cup bottom radius region and the
ange region (near the die throat) because the initial plastic
deformation takes place in these regions. With a further punch
displacement, the damage initiates in the wall region and grows
at the cup bottom radius region. After the punch displacement of
Fig. 13. The variation of triaxiality (ratio of the mean stress to the equivalent
stress) with punch displacement for element number 157.
Fig. 14. The variation of equivalent plastic strain with punch displacement for
element number 157.
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
Fig. 15. Damage distribution at a punch displacement of 3.41 mm.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
Fig. 16. Damage distribution at a punch displacement of 13.01 mm.
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1112
13.0 mm, all the nodes under the punch prole radius region
come into contact with the punch and there is no further plastic
deformation in the cup bottom radius region. Therefore, the zone
of maximum damage shifts from the cup bottom radius region to
the wall region with a further increase in the punch displacement
(Figs. 16 and 17). Finally, the damage reaches the critical value at
the punch stroke of 20.01 mm in the cup wall (Fig. 18). The
predicted location of fracture initiation is in agreement with
the experimental result of material A2024 from Ref. [14]. As
mentioned in Ref. [14], this fracture is not preceded by localized
necking and hence may not be predictable by FLD.
It is further observed that at the at bottom of the cup and at
the ange region, the total damage value is very small at the
initiation of fracture. Fig. 18 shows that the maximum cup height
that can be achieved without fracture is 20.01 mm.
For circular cup also, one can analyze the damage growth in a
representative element (in which the damage reaches the critical
value) using the procedure similar to that used for the square cup.
In circular cup, the value of equivalent plastic strain at fracture
initiation is observed to be 0.1805 which is much less than the
corresponding value of 0.3063 in square cup (Fig. 14). This shows
that the fracture initiation in square cup is inuenced mostly by
the severe plastic deformation (occurring in the corner regions)
than by the triaxiality.
3.3. Parametric study
The modied FE code is used for carrying out the parametric
study of the maximum cup height that can be achieved without
fracture for various sets of process parameters. The sheet size/
diameter and the coefcient of friction are kept constant in the
parametric study. The materials used for the parametric study are
forming steel [34] and aluminum alloy (AA6060-T5) [49], whose
material properties are given in Table 2. The results of the
parametric study are expressed in terms of the normalized
maximum cup height d/b where d is the maximum cup height
without fracture (i.e., the cup height when DD
c
) and b is the
sheet-diameter for circular cup and sheet-size for square cup.
3.3.1. Effect of sheet thickness
The variation of maximum cup height (that can be achieved
without fracture) is studied for the following sheet thicknesses in
this section: t0.86, 0.90, 1.0, and 1.1 mm. The analysis is carried
out with the total blank-holder force of F
b
15.0 kN for square cup
and F
b
18.0 kN for circular cup drawing. The tooling geometry
remains the same. The material is forming steel. Figs. 19 and 20
show the variations of the normalized maximum cup height d/b
(without fracture) with respect to the normalized sheet thickness
t/b for square and circular cup drawing processes, respectively. It
is observed that the maximum cup height (without fracture)
increases with the sheet thickness for both the cups. This is
because a thicker sheet is expected to fracture at a higher level of
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
Fig. 17. Damage distribution at a punch displacement of 17.01 mm.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Fig. 18. Damage distribution at a punch displacement of 20.01 mm.
Fig. 19. Effect of sheet thickness on the maximum cup height for square cup
(material: forming steel).
Fig. 20. Effect of sheet thickness on the maximum cup height for circular cup
(material: forming steel).
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1113
deformation. However, the increase in the maximum cup height
is more signicant for circular cup. It is observed that, if only one
element is used in the thickness direction, the expected trend is
not observed (i.e., the maximum cup height decreases with the
sheet thickness).
3.3.2. Effect of die prole radius
The die prole radius affects the maximum cup height that can
be achieved without fracture. For square cup, this effect is studied
for the following values of the die prole radii: r
dp
3.0, 4.0, 5.0,
and 7.0 mm. The analysis is carried out with the total blank-
holder force of F
b
15.0 kN. The analysis for circular cup is carried
out for the following values of the die prole radii; r
dp
6.0, 8.0,
9.0, and 10.0 mm. Further, the total blank-holder force is
F
b
18.0 kN. For both the cases, the material is forming steel.
The sheet thickness remains the same (1.0 mm). The other
dimensions of the tooling geometry also remain the same.
Figs. 21 and 22 show the variations of the normalized maximum
cup height d/b (without fracture) with respect to the normalized
die prole radius r
dp
/b for square and circular cup drawing
processes, respectively. For both the cups, the maximum cup
height (without fracture) increases with an increase in the die
prole radius. The material ow from the die contact point to the
punch contact point becomes more smooth with an increase in
the die prole radius. Therefore, the stretching in the cup wall
reduces and a larger punch displacement is required for the
damage to reach the critical value when the die prole radius is
increased. For square cup, however, the increase in the maximum
cup height is insignicant at higher values of the die prole
radius. The material ow in the corners of square cup is con-
strained. Because of this, the improvement in the smoothness of
the material ow is less signicant at higher values of the die
prole radius.
3.3.3. Effect of punch prole radius
The punch prole radius also has an important inuence on
the fracture initiation in deep drawing. The analysis for square
cup is performed for the following values of the punch prole
radii: r
pp
6.0, 8.0, 9.0, and 10.0, and with the total blank-holder
force of F
b
15.0 kN. The analysis for circular cup is carried out for
the following values of the punch prole radii: r
pp
6.0, 8.0, 9.0,
and 10.0 mm. Further, the total blank-holder force is F
b
18.0 kN.
For both the cases, the material is forming steel. The sheet
thickness remains the same (1.0 mm). The other dimensions of
the tooling geometry also remain the same. Figs. 23 and 24 show
the variations of the normalized maximum cup height d/b (with-
out fracture) with respect to the normalized punch prole radius
r
pp
/b for square and circular cup drawing processes, respectively.
For both the cups, the maximum cup height (without fracture)
increases with an increase in the punch prole radius. When the
punch prole radius is increased, then also the material ow from
the die contact point to the punch contact point becomes more
smooth thereby reducing the stretching in the cup wall. Because
of this, a larger punch displacement is needed for the damage to
reach the critical value. This result is consistent with the observa-
tion of Ref. [9] that the higher punch prole radius enhances the
formability of square cup drawing. The observation of Ref. [9] is
based on the use of the FLD in the strain space as the formability
criterion.
3.3.4. Effect of blank-holder force
The change in the maximum cup height (without fracture) is
studied for different values of the blank-holder force: F
b
16.0,
18.0, 20.0, and 22.0 kN. The sheet thickness (1.0 mm) and the
Fig. 21. Effect of die prole radius on the maximum cup height for square cup
(material: forming steel).
Fig. 22. Effect of die prole radius on the maximum cup height for circular cup
(material: forming steel).
Fig. 23. Effect of punch prole radius on the maximum cup height for square cup
(material: forming steel).
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1114
tooling geometry remain the same. The material used is forming
steel. It is found that the blank-holder force has no effect on the
maximum cup height (without fracture). The blank-holder force
does affect the triaxiality in the ange region. However, when the
critical element is in the ange region, the equivalent plastic
strain is less than the threshold value e
D
(Figs. 7 and 14). Thus,
the damage in the critical element does not initiate till it enters
the die cavity. Therefore, the blank-holder force does not affect
the maximum cup height (without fracture).
3.3.5. Effect of material properties
The material properties also affect the maximum cup height
that can be achieved without fracture. This effect is studied in two
stages. First, for the forming steel, this effect is analyzed by
varying the initial yield stress s
y

0
and the hardening parameters
K and n. In the second stage, this effect is studied for two different
materials (the forming steel and the aluminum alloy AA6060-T5)
for which the D
c
(the critical damage value) is also different
besides s
y

0
, K and n.
In the rst stage, the effect of material properties is studied by
rst keeping the hardening parameters K and n constant and
varying only the initial yield stress: s
y

0
200, 276, 350, and
450 MPa. The analysis is carried out with the total blank-holder
force of F
b
18.0 kN for circular cup and F
b
15.0 kN for square
cup drawing. The sheet thickness (1.0 mm) and the tooling
geometry remain the same. Figs. 25 and 26 show the variations
of the normalized maximum cup height d/b (without fracture)
with respect to the normalized yield stress s
y

0
=276 (276 MPa is
the initial yield stress of the forming steel as mentioned in
Table 2) for square and circular cup drawing processes, respec-
tively. It is observed that the maximum cup height (without
fracture) increases with an increase in the initial yield stress. An
increase in the initial yield stress delays the occurrence of plastic
deformation, thereby increasing the maximum cup height (with-
out fracture). However, the effect of variation in s
y

0
is not very
signicant in square cup drawing.
Next, the hardening parameters K and n are varied one at a
time, by keeping s
y

0
and the other hardening parameter xed. It
is observed that an increase in the hardening coefcient (K)
increases the maximum cup height (without fracture) in circular
cup drawing, whereas, in square cup drawing it has no inuence
on the maximum cup height (without fracture). Further, it is
observed that an increase in the hardening exponent (n) reduces
the maximum cup height (without fracture) in both the cups. This
is because the triaxiality at fracture initiation increases with an
increase in n. However, the effect of n is not very signicant in
square cup drawing. In general, the plastic deformation is inu-
enced mostly by the geometric parameters whereas the triaxiality
is affected largely by the material properties. As stated earlier,
compared to circular cup, the fracture initiation in square cup is
inuenced mostly by the severe plastic deformation (occurring in
the corner regions) than by the triaxiality. This explains why the
material properties do not have much effect on the maximum cup
height (without fracture) in square cup.
The parameters used in the second stage of the parametric
study on material properties are as follows. The sheet thicknesses
for the forming steel are t0.86, 0.90, 1.0, and 1.1 mm and for the
aluminum alloy are t0.9, 1.0, 1.1, and 1.2 mm. The total blank-
holder force is F
b
15.0 kN for square cup and F
b
18.0 kN for
circular cup for the forming steel and F
b
22.0 kN for square cup
and F
b
28.0 kN for circular cup for the aluminum alloy. This
has been done to avoid wrinkling. Since, the blank-holder force
does not affect the maximum cup height, this difference in the
Fig. 25. Effect of initial yield stress on the maximum cup height for square cup
(material: forming steel).
Fig. 26. Effect of initial yield stress on the maximum cup height for circular cup
(material: forming steel).
Fig. 24. Effect of punch prole radius on the maximum cup height for circular cup
(material: forming steel).
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1115
blank-holder force does not affect the objective of the parametric
study, i.e., to study the effect of only the material properties. The
tooling geometry remains the same. Figs. 27 and 28 show the
variations of the normalized maximum cup height d=b (without
fracture) with respect to the normalized sheet thickness t/b for
square and circular cup drawing, respectively. It is observed that
for square cup, the maximum cup height (without fracture) is
more for the aluminum alloy whereas for circular cup, it is more
for the forming steel.
The reasons for these trends can be explained as follows. As
per Table 2, all the parameters s
y

0
, K and n are higher for the
forming steel than for the aluminum alloy AA6060-T5. The
maximum cup height (without fracture) increases with s
y

0
and K and decreases with n. The net effect seems to be that the
maximum cup height (without fracture) is more for the forming
steel in circular cup (Fig. 28).
However, the trend in square cup is different (Fig. 27). Note
that the above observations about the dependence of the
maximum cup height (without fracture) on s
y

0
, K and n are
based on the parametric study in which the critical damage value
D
c
is kept constant. However, the maximum cup height (without
fracture) is the cup height at which the damage reaches D
c
.
Therefore, if the D
c
is more for one material, the maximum cup
height (without fracture) is also more for that material if the
remaining things are equal. The value of D
c
is more (0.8) for the
aluminum alloy than for the forming steel (0.32). (This is because,
the density of the secondary particles is much less in the
aluminum alloy than in the forming steel thereby requiring
higher levels of deformation and triaxiality for the voids to
coalesce.) Further, for square cup, K has no effect and s
y

0
and
n have only small effect on the maximum cup height (without
fracture). Therefore, in square cup, the maximum cup height
(without fracture) is more for the aluminum alloy in-spite of
having lower values of s
y

0
, K and n than for the forming steel.
4. Conclusions
At microscopic level, ductile fracture occurs rst due to the
nucleation of micro-voids, then their growth and nally their
coalescence into micro-cracks. At the continuum level, the ductile
fracture depends on two continuum parameters: (i) the equiva-
lent plastic strain and (ii) the stress triaxiality. In the continuum
damage mechanics model, these two continuum parameters are
combined into a single parameter called damage (which repre-
sents the intensity of micro-voids) and its critical value is used for
the prediction of fracture initiation. In the present work, Lemai-
tres continuum damage mechanics (CDM) model is used for the
analysis of damage in square and circular cup drawing processes.
From this study, the following conclusions can be drawn:
The fracture location predicted by the present formulation is in
agreement with the results reported in the literature. Thus,
Lemaitres CDM model seems to be capable of predicting the
fracture in deep drawing processes.
The maximum cup height (i.e., the cup height at which the
fracture initiates) increases with the sheet thickness, the die
prole radius and the punch prole radius.
In circular cup, the fracture initiation is inuenced by the plastic
(material) properties (like the initial yield stress and the hard-
ening parameters) as well as the critical damage value.
In square cup, because of the presence of the corner regions,
the fracture initiation is inuenced mostly by the plastic
deformation in the corner regions and less by the triaxiality.
As a result, the plastic (material) properties like the initial
yield stress and the hardening parameters, which normally do
not inuence the plastic deformation much, do not have much
effect on the fracture initiation. However, the critical damage
value certainly affects the fracture initiation.
Acknowledgment
This work is a part of the research work under QIP scheme
sanctioned by AICTE, INDIA.
References
[1] K. Lange, Handbook of Metal forming, McGraw-Hill Book Company, New
York, 1985.
[2] S.P. Keeler, W.A. Backhofen, Plastic instability and fracture in sheet stretched
over rigid punches, ASM Trans. Q. 56 (1964) 2548.
[3] G.M. Goodwin, Application of strain analysis to sheet metal forming in the
press shop, SAE paper no. 680093, 1968.
[4] H.W. Swift, Plastic instability under plane stress, J. Mech. Phys. Solids 1 (1)
(1952) 118.
[5] R. Hill, On discontinuous plastic states, with special reference to localized
necking in thin sheets, J. Mech. Phys. Solids 1 (1) (1952) 1930.
[6] Z. Marciniak, K. Kuczyn ski, Limit strains in the processes of stretch-forming
sheet metal, Int. J. Mech. Sci. 9 (1967) 609620.
Fig. 27. Effect of material properties on the maximum cup height for square cup.
Fig. 28. Effect of material properties on the maximum cup height for circular cup.
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1116
[7] J.W. Hutchinson, K.W. Neale, Sheet necking-II: time-independent behaviour,
in: D.P. Koistinen, N.M. Wang (Eds.), Mechanics of Sheet Metal Forming,
Plenum Press, New York, 1978, pp. 127153.
[8] S.H. Evangelista, J. Lirani, H.A. Al-Qureshi, Implementing a modied
MarciniakKuczyn ski model using the nite element method for the simula-
tion of sheet metal deep drawing, J. Mater. Process. Technol. 130131 (2002)
135144.
[9] F. Chen, T. Huang, C. Chang, Deep drawing of square cups with magnesium
alloy AZ31 sheets, Int. J. Mach. Tool Manuf. 43 (15) (2003) 15531559.
[10] R. Arrieux, Determination and use of the forming limit stress diagrams in
sheet metal forming, J. Mater. Process. Technol. 53 (1995) 4756.
[11] T.B. Stoughton, A general forming limit criterion for sheet metal forming, Int.
J. Mech. Sci. 42 (2000) 127.
[12] R. Arrieux, M. Brunet, P. Vacher, T.N. Nhat, A method to predict the onset of
necking in numerical simulation of deep drawing operations, Ann. CIRP 45
(1) (1996) 255258.
[13] Z. Marciniak, K. Kuczyn ski, T. Pokora, Inuence of the plastic properties of a
material on the forming limit diagram for sheet metal in tension, Int. J. Mech.
Sci. 15 (1973) 789805.
[14] H. Takuda, K. Mori, N. Hatta, The application of some criteria for ductile
fracture to the prediction of the forming limit of sheet metals, J. Mater.
Process. Technol. 95 (13) (1999) 116121.
[15] M.G. Cockroft, D.J. Latham, Ductility and workability of metals, J. Inst. Metals
96 (1968) 3339.
[16] P. Brozzo, B. DeLuca, R. Rendina, A new method for the prediction of
formability in metal sheets, in: Proceedings of the Seventh Biennial Con-
ference of the International Deep Drawing Research Group (IDDRG), Amster-
dam, Netherlands, 1972.
[17] S.E. Clift, P. Hartley, C.E.N. Sturgess, C.W. Rowe, Fracture prediction in plastic
deformation process, Int. J. Mech. Sci. 32 (1990) 117.
[18] M. Oyane, T. Sato, K. Okimoto, S. Shima, Criteria for ductile fracture and their
applications, J. Mech. Work. Technol. 4 (1980) 6581.
[19] C.A. Berg, Plastic dilation and void interaction, in: M.F. Kannien, W.F. Adler,
A.R. Rosenfeld, R.I. Jaffe (Eds.), Inelastic Behavior of Solids, McGraw-Hill, New
York, 1969, pp. 171210.
[20] A.L. Gurson, Continuum theory of ductile rupture by void nucleation and
growth: Part Iyield criteria and ow rules for porous ductile media, J. Eng.
Mater. Technol. 99 (1977) 215.
[21] S.H. Goods, L.M. Brown, The nucleation of cavities by plastic deformation,
Acta Metall. 27 (1979) 115.
[22] J.R. Rice, D.M. Tracey, On the ductile enlargement of voids in triaxial stress
eld, J. Mech. Phys. Solids 17 (1965) 201217.
[23] P.F. Thomason, Ductile Fracture, Pergamon Press, UK, 1990.
[24] J. Lemaitre, A continuous damage mechanics model for ductile fracture,
J. Eng. Mater. Technol. 107 (1985) 8389.
[25] J. Lemaitre, Coupled elasto-plasticity and damage constitutive equations,
Comput. Methods Appl. Mech. Eng. 51 (1985) 3149.
[26] J. Lemaitre, R. Desmorat, Engineering Damage Mechanics, Springer-Verlag,
Berlin, Heidelberg, 2005.
[27] V. Tvergaard, Inuence of voids on shear band instabilities under plane strain
condition, Int. J. Fract. 17 (1981) 389406.
[28] V. Tvergaard, On localization in ductile materials containing spherical voids,
Int. J. Fract. 18 (1982) 237252.
[29] L. Xia, C.F. Shih, J.W. Hutchinson, Computational approach to ductile crack
growth under large scale yielding conditions, J. Mech. Phys. Solids 43 (1995)
389413.
[30] E. Doege, T. El Dsoki, D. Seibert, Prediction of necking and wrinkling in sheet
metal forming, J. Mater. Process. Technol. 50 (1995) 197206.
[31] L.M. Kachanov, Introduction to Continuum Damage Mechanics, Martinus
Nijhoff Dortrecht, The Netherlands, 1986.
[32] S. Dhar, R. Sethuraman, P.M. Dixit, A continuum damage mechanics model for
void growth and micro-crack initiation, Eng. Fract. Mech. 53 (1996) 917928.
[33] G. Le Roy, J.D. Embury, G. Edwards, M.F. Ashby, A model of ductile fracture
based on the nucleation and growth of voids, Acta Metall. 29 (1981)
15091522.
[34] M. Elgueta, Ductile damage analysis of sheet metal forming, J. Mater. Process.
Technol. 121 (2002) 148156.
[35] M. Khelifa, M. Oudjene, A. Khennane, Fracture in sheet metal forming: effect
of ductile damage evolution, Comput. Struct. 85 (34) (2007) 205212.
[36] R. Hill, A theory of the yielding and plastic ow of anisotropic metals, Proc. R.
Soc. London 193 (1948) 281297.
[37] J.P. Fan, C.Y. Tang, C.P. Tsui, L.C. Chan, T.C. Lee, 3D nite element simulation of
deep drawing with damage development, Int. J. Mach. Tool Manuf. 46 (2006)
10351044.
[38] S. Dhar, A continuum damage mechanics model for ductile fracture, Ph.D.
Thesis, Department of Mechanical Engineering, IIT Kanpur, September 1995.
[39] D.R.J. Owen, E. Hinton, Finite Elements in Plasticity: Theory and Practice,
Pineridge Press, Swansea, 1980.
[40] S.N. Varadhan, Dynamic large deformation elasto-plastic analysis of continua,
Masters Thesis, Mechanical Engineering Department, IIT Kanpur, 1997.
[41] J. Lemaitre, J. Dufailly, Damage measurements, Eng. Fract. Mech. 28 (1987)
643661.
[42] N. Bonora, A. Ruggiero, D. Gentile, S. De Meo, Practical applicability and
limitations of the elastic modulus degradation technique for damage mea-
surements in ductile metals, Strain 47 (2011) 241254.
[43] M. Alves, J. Yu, N. Jones, On the elastic modulus degradation in continuum
damage mechanics, Comput. Struct. 76 (2000) 703712.
[44] B. Bennani, F. Lauro, Damage models and identication procedures for
crashworthiness of automotive light materials, Latin Am. J. Solid Struct. 3
(2006) 7587.
[45] K.J. Bathe, Finite Element Procedures, Prentice Hall of India, New Delhi, 1996.
[46] K. Lange, M. Herrmann, P. Keck, M. Wilhelm, Application of an elasto-plastic
nite element code to the simulation of metal forming processes, J. Mater.
Process. Technol. 27 (1991) 239261.
[47] S. Kobayashi, S. Oh, T. Altan, Metal Forming and The Finite Element Method,
Oxford University Press, Oxford, 1989.
[48] T.H. Choi, H. Huh, Sheet metal forming analysis of planar anisotropic
materials by a modied membrane nite element method with bending
effect, J. Mater. Process. Technol. 8990 (1999) 5864.
[49] D. Al Galib, A. Limam, A. Combescure, Inuence of damage on the prediction
of axial crushing behavior of thin-walled aluminum extruded tubes, Int. J.
Crash. 11 (1) (2006) 112.
R.K. Saxena, P.M. Dixit / Finite Elements in Analysis and Design 47 (2011) 11041117 1117

Você também pode gostar