2009 Thuillier

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

International Journal of Plasticity 25 (2009) 733–751

Contents lists available at ScienceDirect

International Journal of Plasticity


journal homepage: www.elsevier.com/locate/ijplas

Comparison of the work-hardening of metallic sheets


using tensile and shear strain paths
S. Thuillier, P.Y. Manach *
Université de Bretagne-Sud, Laboratoire d’Ingénierie des MATériaux de Bretagne, Rue de Saint Maudé, BP 92116, 56321 Lorient, France

a r t i c l e i n f o a b s t r a c t

Article history: This work deals with the characterization of the kinematic work-
Received 11 January 2008 hardening of a bake-hardening steel. A shear test device has been
Received in final revised form 9 July 2008 designed and its use for the characterization of the work-hardening
Available online 22 July 2008
of sheet metals is described. Two main results are presented.
Firstly, a local strain measurement, based on the following of three
Keywords: dots drawn on the gauge area, gives the evolution of the strain ten-
B. Anisotropic material
sor eigenvalues during the test. It is shown, by comparing the the-
B. Elasto-viscoplastic material
B. Finite strain
oretical kinematics of simple shear with a slightly perturbated one,
B. Metallic materials that the strain state is close to the ideal one in the center of the
C. Mechanical testing gauge area. Secondly, reversal of the shear direction is performed
after several prestrain and the evolution of the kinematic work-
hardening with the equivalent plastic strain has been identified
using an anisotropic elasto-viscoplastic model of Hill 1948 type.
Isotropic and kinematic contributions of the work-hardening are
also calculated from loading–unloading tensile tests and are com-
pared to those obtained from the simple shear tests. The results
show a discrepancy between both identification for the isotropic
and the kinematic hardening. However, they are in agreement con-
cerning the evolution of the global work-hardening.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction

In the context of economic requirements for the lightening of structures, particularly in the automo-
tive industry, the mechanical properties currently required for steel sheets used for metal forming stim-
ulate the development of new materials whose mechanical resistance is increasingly high. However,
their performances and their reliability require the knowledge of their mechanical properties for

* Corresponding author.
E-mail address: pierre-Yves.Manach@univ-ubs.fr (P.Y. Manach).

0749-6419/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2008.07.002
734 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

complex loadings and for large plastic deformations, as well as the prediction of the springback effect,
phenomenon which is crucial for the dimensional accuracy of parts and tools. The simulation of spring-
back can be seen as a trivial problem compared to the simulation of the forming phases, since the equi-
librium state can be reached rather easily by removing the contact stresses. However, many researchers
have been studying this phenomenon for the last decades as it can be seen in the proceedings of the
Numisheet (1996), Numisheet (1999), Numisheet (2002), Numisheet (2005) conferences. The final
shape of the part depends mainly on the amount of elastic energy stored in the part during the drawing
process (Narasimhan and Lovell, 1999). The springback prediction strongly depends on the accurate
simulation and realistic modeling of the forming operation in order to correctly predict the final state
variables, such as stress and strain tensors (Oliveira et al., 2007; Lee et al., 2008). To obtain accurate
numerical solutions, mechanical models should use reliable descriptions of the material behavior, i.e.
anisotropy and work-hardening during loading as well as during unloading (Li et al., 2002).
At the moment, the most widely used constitutive model is the one based on the Hill (1948) yield
criterion to characterize the initial anisotropy and an isotropic work-hardening, with the material
parameters being identified from experimental uniaxial tensile tests. In that respect, hardening char-
acterization is central for the formulation of the constitutive equations. The importance of considering
the Bauschinger effect by using a kinematic hardening law instead of an isotropic hardening has been
demonstrated by many authors (Tang, 1990; Gau and Kinzel, 2001; Geng and Wagoner, 2002; Hahm
and Kim, 2008). Thus, more sophisticated constitutive models like the one suggested by Armstrong
and Frederick (1966) (Choi et al., 2006; Choi et al., 2006) or more complex ones like the Teodosiu and
Hu model (Teodosiu and Hu, 1998; Haddadi et al., 2006; Haddag et al., 2007), which take into account
non-linear kinematic hardening and more complex internal state variables, are expected to give an
improvement in the accuracy of the sheet metal forming simulations.
Unfortunately, in the three-dimensional case, the identification of such tensorial constitutive laws
is difficult to achieve using only tensile tests, and other deformation modes need to be investigated for
the identification of material parameters (Suprun, 2006). These parameters can be determined from
mechanical tests which need to be adapted to thin sheet metals. Conventional tension–compression
tests are generally very difficult to set up due to the buckling of the sheet specimen (Kuwabara et
al., 1995; Yoshida et al., 2002; Lee et al., 2005). For this reason, Zhao and Lee (2002) suggested an iden-
tification of the hardening characteristics by means of a cyclic three-point bending test; Yoshida et al.
(1998) and Brunet et al. (2001) developed a pure bending device and performed cyclic loadings of the
sheets. It should be emphasized that for thin metallic sheets of around 1 mm thick, the maximum
strain reached within each cycle is of the order of 0.05. Hu (2005) and Tong (2006) used standard uni-
axial-tension and equi-biaxial tension tests for the characterization of their models. A last method for
flat sheet metals consists in using the planar simple shear test (Miyauchi, 1992).
Among all mechanical tests that have been developed to characterize the material behavior of
sheets, the simple shear test has received an increasing attention during the past decade (Wack and
Tourabi, 1993; Manach and Favier, 1997; Rauch, 1998) and in recent years (Flores et al., 2005; Yoon
et al., 2005; Bouvier et al., 2006; Flores et al., 2007). The main reason is that this test is rather simple
and that it can be performed easily on a classical tensile machine by using a dedicated device. More-
over, shear tests have several advantages: firstly, both shear and tensile samples can be machined
from the same sheet. Secondly, the shear test enables cyclic deformations to be carried out and the
sheet shape is particularly suitable for investigating the influence of different thermomechanical treat-
ments such as cold rolling and/or annealing on the resulting properties (Fjeldly et al., 1998). Thirdly,
there are no artifacts due to thermal dilatation, which is an important point for many materials for
which the temperature plays an essential role in their behavior (Manach and Favier, 1997). Finally,
the deformation may be considered as homogeneous throughout the shear gauge section; actually
near the free ends of the sample, boundary effects disturb the stress distribution up to 5% of the aver-
age stress, but it has been shown that these effects may be decreased by using a long and thin shear
zone (G’Sell et al., 1983; Rauch and G’Sell, 1989; Bouvier et al., 2006; Duchêne et al., 2008). Another
advantage is that both homogeneous tensile and shear tests are complementary: the strain state in-
volves only diagonal components of the strain tensor in the case of tension while non diagonal com-
ponents of the strain tensor mainly appear in shear. Moreover, both tests can be used for the testing of
planar anisotropy which is essential for deep drawing process: this leads to a different and non
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 735

classical point for the identification of yield criteria. Finally, this test enables large equivalent plastic
strains as well as the characterization of kinematic work hardening, which is generally difficult to
reach in experiments on thin products.
However, the shear test has also several disadvantages that have already been pointed out (Bouvier
et al., 2006). Firstly, the measurement of the local strain is difficult because of the narrowness of the
shear zone. Secondly, the ratio between the length, the width and the thickness of the shear zone
should be chosen in order to obtain a homogeneous strain state and to avoid buckling (G’Sell et al.,
1983).
The aim of this work is twofold: to check the kinematics of the experimental shear test and to com-
pare the evolution of the kinematic hardening part identified in simple shear with the one identified
during loading–unloading tensile tests. In that respect a shear test device and its use for the charac-
terization of the work-hardening of sheet metals is presented. Several aspects are discussed, such as
the kinematics and the stress state of the test. A particular attention is paid to the measurement of the
shear strain by the use of an optical method that gives all components of the planar strain tensor. This
measurement leads to a better knowledge of the perturbating effects that affect the strain state, par-
ticularly during strain path reversal. Reversal of the shear direction is performed after several pre-
strain and the evolution of the kinematic work-hardening with the equivalent plastic strain is
discussed. The Bauschinger effect has been explained in terms of long-range stresses acting on mobile
dislocations, e.g. (Feaugas, 1999); these internal stresses, or back stresses, can be measured from ten-
sion–compression tests (Choteau et al., 2005) or during unloading (Feaugas, 1999; Cleveland and
Ghosh, 2002). In this work, both simple shear tests and loading–unloading sequences in tension were
performed, to compare the magnitude of these internal stresses.

2. Experimental device

The shear test specimens, presented Fig. 1, have a rectangular shape of L  l ¼ 50  18 mm2 with a
shear gauge width h of 4.5 mm. The shear width can be adjusted from 2 to 5 mm; the sheet thickness
can be within 0:2 6 e 6 1:2 mm and the shear direction is along the length of the specimen. The width

Fig. 1. Shear test samples: (a) undeformed, (b) after clamping the grips, (c) after deformation and (d) using image correlation
painting. Three points of different sizes are drawn in the center zone of the samples (b) and (c) with a serigraphic painting to
measure the deformation. ~ I1 is aligned to the shear direction, ~
I2 to the transverse direction and ~
I3 is normal to the sample.
736 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

of the gauge area can be changed by moving the grips perpendicularly to the shear direction. This
shear test device (Fig. 2) has been especially designed for the study of metallic sheets. The sample
(1) is clamped between two grips (2) and (3), immovably attached to the fixed part (4) and the moving
part (5) of the apparatus, respectively. Conversely to the earlier experimental devices (Miyauchi, 1977;
G’Sell et al., 1983; Rauch and G’Sell, 1989) for which the guides are not symmetric, the relative motion
between (4) and (5) is obtained by pairs of linear guides symmetrically positioned with respect to the
sample (Wack and Tourabi, 1993; Bouvier et al., 2006). The device, presented in Fig. 3, is directly con-
nected to a tensile test machine (not seen in this figure). The clamping of the sample under the grips
(Fig. 3) is obtained by the tightening of six screws with a torque wrench; the torque is dependent on
the tested material. The optimal value is obtained with the lowest torque that minimizes the sliding
between the sample and the grips. For quasi-static tests, the imposed strain rate is usually c_ ¼ 103 s1
but higher strain rates have already been performed (Manach and Couty, 2002). Conversely to most of
the existing shear apparatus that enables the adjustement of the width of the shear zone, the grips can
move symmetrically from the shear axis, which means that the shear zone remains centered, what-
ever the shear width.
Many numerical simulations have been carried out to study the effect of the shape of the sample on
stress and strain fields, in particular near the free edges of the gauge area (Manach, 1989; Yoon et al.,
2005; Bouvier et al., 2006). These simulations show that the stress state is nearly homogeneous in the
shear zone and that the tensile components are weak compared to the shear stress components. Due

4 6

1
3
2 10

7
9

Fig. 2. Schematic overview of the simple shear test apparatus: 1. sample; 2. movable grip; 3. fixed grip; 4. upper frame part; 5.
lower frame part; 6. load cell; 7. hydraulic actuator; 8. actuator displacement sensor; 9. frame and 10. computer.
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 737

Fig. 3. Pictures of the shear device and zoom of the grip system.

to the edge effects, the Cauchy stress tensor involves also tensile components. In the frame (~ I1 ;~
I2 ;~
I3 ),
where ~I1 is aligned to the shear direction, ~
I2 to the transverse direction and ~
I3 to the normal direction,
it can be written:
0 1
r11 r12 0
B C
r ¼ @ r12 r22 0 A ð1Þ
0 0 0
The components r11 and r22 can reach high values near the free ends of the sample but can be ne-
glected by using a length-over-width ratio L=h greater than 10 for metallic materials (Bouvier et al.,
2006). However, it should be emphasized that the prediction of these normal stresses depend on
the large strain formalism used in the numerical model.

3. Shear strain measurement

The strain measurement in shear is not as direct as in tension. Within the large deformation frame-
work, different strain measures have been introduced: e.g., the non-diagonal component of the Euler-
Almansi or Green-Lagrange strain tensor is equal to c=2 in the test frame, where c is the non-diagonal
and non-zero component of the transformation gradient tensor F. However, the integration of the
strain rate tensor in the corotational frame leads to a non-diagonal strain component equals to
sin c=2 (Gilormini, 1994). For the sake of simplicity, experimental data of shear tests are most of
the time displayed by using the parameter c, which describes the kinematics of the simple shear test.
The measurement of c in shear is quite complex (Wack and Tourabi, 1993). During a simple shear
test the principal axes of stress and strain tensors rotate and, in most of the shear test apparatus in
which only the shear strain is measured it is not possible to access the strain state completely. Most
often, the shear strain is evaluated from the relative grip displacement (Miyauchi, 1977; Rauch and
G’Sell, 1989), this shear strain value being sometimes corrected in order to take into account the slid-
ing of the sample under the grips (Rauch, 1998; Yoon et al., 2005). Another measure consists in draw-
ing a line on the sample normal to the shear direction and to follow the rotation of this line during
the test with a video camera (Mohanraj et al., 2006; Bouvier et al., 2006). Both methods lead to a
738 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

measurement of c ¼ x=h where x is the maximum displacement of the line in the shear direction. In
this study, an original measurement of the shear strain also based on an optical method has been
developed. Three points of different sizes are drawn in the center zone of the sample with a serigraph-
ic painting (see Figs. 1 and 4). During the test, a video camera records the actual position of these
points; the definition of two vectors joining these points enables the calculation of all components
of the planar strain tensor. Thanks to the quality of the high resolution camera, this method enables
an accuracy on the c measurement of the order of 104 . This measurement uses the notion of con-
vected material frame which is detailed in the following part.

3.1. Convected material frame and strain tensor

When studying the deformation of a solid body, its position relative to a general frame is not
important since only the relative position of its material points is necessary. In that respect, a material
frame can be chosen and the description of the sample configuration is performed using convected
material coordinates.
A general 2D kinematics involving convected material frames is studied according to Green and
Zerna (1968). Let us consider a continuous solid X in motion and M a point of this solid. The position
of M is defined in a fixed cartesian frame (e.g., the laboratory frame) ~ Ia , a ¼ 1; 2 by its coordinates X að0Þ
at time t ¼ 0 and xaðtÞ at time t > 0. To follow the point M in its motion, it is marked using material
coordinates. This marking is constant throughout the deformation and the simplest one corresponds
to the coordinates in the fixed frame at t ¼ 0. A corresponding natural, or covariant, frame (M; ~ g^i ) is
associated with point M:
~ ðtÞ oxaðtÞ 
oM
g^iðtÞ ¼
~ ¼ ~
Ia ð2Þ
oX i oX i
where M~ ðtÞ ¼ xa ~ ~^
ðtÞ Ia . The g iðtÞ vectors are calculated at time t and they evolve during the deformation.
The covariant components of the metric tensor:

2
2

1
1

Fig. 4. Strain measurement and kinematics of the shear test.


S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 739

g^ðtÞ
g^ ¼ g^ ijðtÞ~ i
g^ðtÞ
~ i
ð3Þ

are derived from these vectors:

g^iðtÞ :~
g^ijðtÞ ¼ ~ g^jðtÞ ð4Þ
g^ðtÞ
where the contravariant or dual vectors ~ i
are defined as follows:

g^ðtÞ
~ i ~
:g^jðtÞ ¼ dij ð5Þ

where d is the Kronecker symbol. Finally, the Almansi strain tensor , which is expressed in the final
configuration, is calculated from:
1 
¼ g^ijðtÞ  g ijð0Þ ~g^ðtÞ
i
g^ðtÞ
~ j
¼ ij~ g^j ¼ ij g ik~
g^i  ~ g^j ¼ kj~
g^k  ~ g^j
g^k  ~ ð6Þ
2
In the last equation, the subscripts t and 0 are omitted and the convected material vectors at t are ref-
erenced using a hat. In the 2D case, at the beginning of the test, g ijð0Þ is stored and for each picture ob-
tained by the camera, the components of the metric tensor given by Eq. (4) are calculated from the
coordinates of the center of mass of each dot drawn on the sample. The covariant components of
the Euler-Almansi strain tensor are then obtained with Eq. (6) and its eigenvalues (ki ) are calculated
using the mixed tensor. This calculation gives quantities that are frame indifferent and the knowledge
of the initial frame is not necessary. During the test, the ki values are recorded as a function of time.
This method can be applied for tensile tests where it gives access to both longitudinal and transverse
strains (Gallée et al., 2007; Thuillier et al., 2008); it is then compared satisfactorily with the measure-
ment of a classical extensometer (see Fig. 5) and for the shear test, where c is then calculated from the
eigenvalues as it is explained in the next section.

3.2. Shear test kinematics

To check whether or not a simple shear kinematics is actually imposed during the test, the eigen-
values measured experimentally are compared with the analytical expressions derived from a shear
test kinematics. This kinematics takes into account a small displacement b of the grips perpendicularly

0.2
logarithmic strain

0.1
longitudinal strain - camera
longitudinal strain - extensometer
transverse strain - camera
0

-0.1

-0.2
0 50 100 150 200 250
time (s)
Fig. 5. Validation of the optical strain measurement on a tensile test. The optical measurement is made in the center of the
sample while for practical reasons, the extensometer is located outside the center area. Both measurements start to diverge
when necking occurs. The irregular behavior observed for t ¼ 10 s is due to the Lüders plateau presented by this material.
740 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

to the shear direction (in the transverse direction ~


I2 ): this is called the perturbated shear kinematics.
Let us consider the kinematics:
(
~¼ x1 ¼ X 1 þ c X 2
M 2
ð7Þ
x2 ¼ X 2 þ bX
Let us choose the natural frame in the initial state as:
~
oM
gi ¼
~ ¼~
Ia ð8Þ
oX i
According to this kinematics, the natural and dual frames in the deformed state can be calculated as
follows:
    ! !
1 c 1 0
g^1 ¼
~ g^2 ¼
~ g^1 ¼
~ c g^2 ¼
~ 1 ð9Þ
0 1þb  1þb 1þb

the covariant metric tensor components can then be defined as:


!
1 c
g^ij~ g^j ¼
g^i  ~ ð10Þ
c ð1 þ bÞ2 þ c2
and
0 1
c2 c
B 1 þ ð1 þ bÞ2 
ð1 þ bÞ 2 C
^ ^
ij~ ^
~ B C
g gi  gj ¼ B C ð11Þ
@ c 1 A

ð1 þ bÞ2 ð1 þ bÞ 2

which leads to the components of the Almansi strain tensor in the natural frame:
0 1
c2 c
  
1 B ð1 þ bÞ2 ð1 þ bÞ CC 1 A C
2
kj~g^j  ~g^k ¼ BB C¼ ð12Þ
2@ c bð2 þ bÞ A 2 C B
ð1 þ bÞ2 ð1 þ bÞ2
In order to calculate the eigenvalues of this tensor, the strain tensor has to be written in mixed coor-
dinates (contravariant and covariant) in the deformed frame. It is then possible to calculate the eigen-
values of this tensor:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 1 1
kpi ¼ ðA þ BÞ  ðA  BÞ2 þ C 2 ð13Þ
4 2 4
In the case of a theoretical shear test kinematics (b ¼ 0), Eq. (13) leads to the simplified expression ki
of the eigenvalues kpi :
 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
ki ¼ c2  c2 ðc2 þ 4Þ ð14Þ
4
It should be emphasized that the eigenvalues are independent of the sign of c. In order to obtain this
information, it can be shown that the eigenproduct of the two eigenvectors ~ V i of the Almansi strain
tensor is equal to the opposite of the sign of c. Therefore, c is calculated from:
k1  k2
jcj ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi and signðcÞ ¼ signð~
V1 ^ ~
V 2Þ ð15Þ
1  k1 þk 2
2

This kinematics is applied to fit an experimental cyclic simple shear test on a E220BH steel sample.
Starting from a given signal cðtÞ, the eigenvalues ki are calculated with Eq. (14) and plotted in Fig.
6. The input signal cðtÞ corresponds to a monotonic shear test up to c ¼ 0:2 and then a reversal test
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 741

0.3
Lambda1
Lambda2
0.2 Gamma

0.1
Shear strain

-0.1

-0.2

-0.3
0 250 500 750 1000
Time (s)
Fig. 6. Evolution of c and of the eigenvalues ki of the Almansi strain tensor for a theoretical shear kinematics during a cyclic
shear test.

down to c ¼ 0:28. The results presented Fig. 6 show that the ki values are nul when reaching the
point c ¼ 0. This is not exactly the case for the experimental test, where k1 remains always negative
in Fig. 8. Due to the momentum applied to the device, both grips tends to nearer together which mod-
ifies the strain state. This is validated by the evolution of the components of the transformation gra-
dient tensor F recorded during a Bauschinger shear test, Fig. 7. These results are obtained in the

1.2
Transformation gradient tensor component

0.8 F11
F12
F21
0.6 F22

0.4

0.2

-0.2

-0.4
0 100 200 300 400 500 600 700
Time (s)
Fig. 7. Evolution of the components F 11 , F 12 , F 21 and F 22 of the transformation gradient tensor F during a reverse shear test
obtained by image correlation.
742 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

0.3 0.03
Lambda1
Lambda2
Gamma
Gamma mod.
delta eps22
0.15 0.015

Normal strain component


Shear strain

0 0

-0.15 -0.015

-0.3 -0.03
0 250 500 750 1000
Time (s)
Fig. 8. Optimization of c from the eigenvalues kexp
i of the Almansi strain tensor, for a perturbated shear kinematics (Gamma
mod) during a cyclic shear test (calculated values). Comparison with gamma calculated using a theoretical shear kinematics
(Gamma). On the right axis, evolution of d22 .

homogeneous central area, using an image correlation technique that follows the deformation of a
random painting on the sample (see Fig. 1d). It can be observed that the component F 22 is not perfectly
constant and equal to 1, as in a theoretical shear test, but it decreases slightly especially when the
shear direction is reversed.
This perturbation is now taken into account in the kinematics and the perturbated shear test kine-
matics is applied to fit the previous experimental cyclic shear test. The values measured during the
test kexp
i are compared with the calculated values kpi given by Eq. (13). c is then calculated by minimiz-
ing the following cost function:
X kp ðcÞ  kexp 2
i i
F¼ ð16Þ
i¼1;2
kexp
i

The results presented Fig. 8 show that this kinematics describes perfectly the kexp i values, even at the
c ¼ 0 point reached after the inversion of the shear direction. Due to this perturbation, the component
22 of the Almansi strain tensor is modified such as p22 ¼ th 22 þ d22 , where 22 is calculated with the
th

theoretical kinematics. The evolution of d22 is drawn in Fig. 8, on the right axis. It shows a slight com-
pression during the test, the magnitude of which is less than 1.7%. This contribution increases both
with the deformation and the number of cycles. Anyway, by taking into account this contribution, c
is very close to the one calculated from the theoretical kinematics (see Fig. 8).
This test is used as the main test for the characterization of the work-hardening of metallic sheets.
As it has been shown in the previous part, the measurement of the shear strain is not very sensitive to
the compression strain that is inevitably induced by the device, even if the eigenvalues of the strain
tensor are better described by the perturbated kinematics. In consequence, as the shear strain is the
main observable variable of this test, the treatment of monotonic and Bauschinger tests is performed
considering a theoretical shear strain kinematics. This method of characterization of the work-hard-
ening of metallic sheets is then compared to a different approach using loading–unloading tensile
tests.
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 743

4. Work-hardening characterization

The modeling of the anisotropic and plastic behavior is extended to include the concept of com-
bined isotropic and non-linear kinematic hardening, which describes the induced anisotropy. Since
we are concerned with sheet metal forming analysis, models need to be identified up to large strains.
Indeed, the isotropic hardening law over-estimates the hardening component by missing the Bausch-
inger effect and the work-hardening stagnation. The pure kinematic hardening rule under-estimates
the hardening component and over-estimates the Bauschinger effect. The combination of isotropic
and non-linear kinematic hardening is intended to predict accurately both effects. Several methods
exist to characterize the kinematic hardening of metallic materials. All methods need cyclic tests or
reversal tests, e.g., in bending-unbending (Brunet et al., 2001), in tension–compression (Yoshida
et al., 1998) or in tension with loading–unloading sequences (Feaugas, 1999; Feaugas and Gaudin,
2004).
The Bauschinger effect corresponds to an early yielding under reversed loading with a rounded
yield point sometimes called the micro Bauschinger process (Hu et al., 1992), followed for high pre-
strains, by a low work-hardening rate. The first step is related to the disappearing of unstable dislo-
cation structures such as pile-ups. The microplasticity recorded during unloading is also related to
dislocation motion with a small mean-free path and therefore this work is focused on the comparison
of the internal stresses recorded during unloading and strain path reversal. In the following para-
graphs, a constitutive law is presented in order to describe the mechanical behavior of a bake harden-
ing steel E220BH. The model is developed within a large deformation framework based on the use of
the corotational frame, i.e. constitutive equations are written in this local objective frame, the orien-
tation of which evolves with the deformation. Then an inverse identification technique is proposed
based both on shear tests and tensile tests. Reverse simple shear tests are used to analyze the Bausch-
inger effect and the evolution of the kinematic hardening for a wide range of equivalent strain (0 up to
0.15). The parameters of the mixed non-linear kinematic hardening model are optimized over the
whole experimental database. These parameters are compared to those measured using loading–
unloading tensile tests.

4.1. Large deformation framework

The description of anisotropic elasto-viscoplastic behavior at large strain requires the use of local
objective frames. Particularly in shear, differences can appear between the different approaches
(Sidoroff and Dogui, 2001; Yeganeh, 2007; Duchêne et al., 2008). In this paper, the shear strain re-
mains lower than 1.0 and it is established that for such values, differences are weak. In this study,
the corotational frame is chosen. This frame is associated to W, the skew-symmetric part of the veloc-
ity gradient tensor L. Assuming small elastic strains, an additive decomposition of the strain rate ten-
sor in the corotational frame is chosen:

_ ¼ Q Tc DQ c ¼ _ e þ _ vp with Q_ c Q Tc ¼ W ð17Þ
where D ¼ L  W is the strain rate tensor and Q c the rotation between the current space frame and the
corotational frame. The constitutive law is written in the corotational frame, using the co-rotated Cau-
chy stress tensor s defined by:

s ¼ ðdet FÞQ Tc rQ c ð18Þ


where F is the transformation gradient tensor and r the Cauchy stress tensor. The associated deriva-
tive is the Jaumann derivative.

4.2. Constitutive law

The constitutive law is a modification of the elasto-plastic model of Hill (1948) in which kinematic
work-hardening and the viscous character of the material are taken into account. The plasticity crite-
rion takes into account the orthotropic symmetry of the sheet in its reference frame, the axis ~
1 being
744 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

aligned with the rolling direction (RD), the axis ~


2 with the transverse direction (TD) and ~3 with the
normal direction. The work-hardening combines isotropic and kinematic contributions, and the evo-
lution of isotropic work-hardening is related to the cumulated plastic strain. The evolution law of the
kinematic work-hardening is non-linear.
The yield function is expressed in the form:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3 d
f ðs; X; RÞ ¼ s  R ¼ ðs  X Þ : M : ðsd  X Þ  R ð19Þ
2
where sd represents the deviatoric part of s, s the equivalent stress and M the 4th order Hill’s tensor.
The 6 non-zero coefficients of M depend on the 6 coefficients F; G; H; L; M; N of the quadratic Hill’s cri-
terion. The condition on the initial elastic limit along the RD imposes the relation G þ H ¼ 2. The sec-
ond-order tensor X represents the back stress and R the term of isotropic work-hardening.
The viscoplastic component of the strain follows a flow rule derived from a viscoplastic potential X
which is a power function of the yield function (Lemaitre and Chaboche, 1996):
 þ nv þ1
Kv f
Xðf Þ ¼ ð20Þ
nv þ 1 K v
where nv is the strain rate sensitivity coefficient, K v a weighting coefficient of the viscous part of the
stress and f þ the positive part of f. The behavior is thus elastic if f 6 0 and if f > 0 the viscoplastic
strain rate is written as:
oX of
_ vp ¼ ¼ X0 ðf Þ ð21Þ
os os
The equivalent viscoplastic strain rate _ vp
eq is defined from the plastic work conservation principle:

ðsd  XÞ : _ vp
_ vp
eq ¼ s
ð22Þ

The evolution of the isotropic work-hardening follows a Swift law:


r 1=n
n 0
R ¼ Kðvp
eq þ 0 Þ with 0 ¼ ð23Þ
K
where K is a material parameter, n the hardening coefficient and r0 is the initial yield stress in tension
along the RD. The evolution law of the kinematic work-hardening is similar to that proposed by Arm-
strong and Frederick (1966) with a linear component of Prager type:
2 2
X¼ Q a þ Hp vp with a_ ¼ _ vp  bx _ vp
eq a ð24Þ
3 x 3
where a is the internal variable associated to X, Q x =bx determines the intensity of the non-linear kine-
matic work-hardening and Hp is the slope of the linear kinematic work-hardening.

4.3. Material

The studied material is a bake hardening mild steel E220BH provided by the Arcelor-Mittal
company. The chemical composition in weight is provided in Table 1 as well as the mechanical
characteristics given by the supplier. The material is supplied as cold rolled sheets of thickness
0.7 mm.

Table 1
Chemical composition (103 % weight) of the studied E220BH mild steel and supplier measured mechanical properties (Re and Rm
are expressed in MPa)

C Mn P N Al Re Rm A%
19 200 4 5 50 220 350 40
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 745

4.4. Experimental results

Tensile tests were carried out on rectangular samples of dimension 20  180  0:7 mm3 . The free
edges were machined in order to eliminate the hardened areas induced by the cutting and thus to in-
crease the range of homogeneous deformation. Monotonous tensile tests were carried out at 0°, 45°,
and 90° to RD in order to study the material anisotropy. For these tests, a cross-head speed of
20 mm=min is imposed which leads to _ ’ 2:4  103 s1 . The logarithmic strain as well as the Cauchy
stress are calculated from the raw data. Fig. 9 shows that the mechanical characteristics are higher in
the RD and lower at 45° to the RD. The plastic anisotropy coefficients r a ¼ dpyy =dpzz , where ~
x denotes
y the transverse and ~
the tensile, ~ z the normal directions respectively, are calculated from the trans-
verse strain yy and the assumption of volume conservation in the plastic area (see Table 2).
Both the average anisotropy coefficient r ¼ ðr0 þ r90 þ 2r45 Þ=4, which characterizes the normal
anisotropy and the planar anisotropy, measured by the coefficient Dr ¼ ðr0 þ r90  2r 45 Þ=2 are
significant.
Monotonous shear tests were performed on samples at 0°, 45°, and 90° to the RD, at a cross-head
speed of 0.5 mm=min, which corresponds to c_ ¼ 2:1  103 s1 . Moreover, cyclic tests were performed
in order to highlight the Bauschinger effect and to measure kinematic work-hardening parameters.
These tests are composed of a loading up to several values of c followed by a load in the opposite
direction until c ¼ 0:4. Each kind of test is performed three times to check the reproducibility. A rep-
resentative test is chosen for the database and experimental stress-strain curves are plotted in Fig. 9.
Tensile samples were also repeatedly loaded then unloaded with longitudinal strain steps of the
order of 0.005, up to a total strain of 0.12. A constant cross-head speed was used, leading to a strain
rate of 4  104 s1 . This gives a large number of recorded points during the unloading, to ensure a

450
RD Tension
TD
DD
300

Shear
Cauchy stress (MPa)

150

-150

-300
-0.4 -0.2 0 0.2 0.4
logarithmic strain or gamma
Fig. 9. Experimental tensile tests at 0° (RD), 45° (DD) and 90° (TD) to the RD and monotonic and Bauschinger shear tests in the
RD on the E220BH steel.

Table 2
Plastic anisotropy coefficients of the E220BH steel

Material r0 r 45 r 90 r Dr
E220BH 1.90 1.42 2.27 1.75 0.66
746 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

correct determination of the linear part and the deviation from the linearity. It was found that, after a
very fast stress drop near the inversion point, there is a linear decrease of the stress. For small plastic
strains (up to 0.01), the slope is close to the initial Young modulus. Then a decrease is observed as
mentionned in (Morestin and Boivin, 1996; Li et al., 2002; Yang et al., 2004). Two tests were
performed, post-treated, and compared to monotonic tensile test to check the validity of the
database.

4.5. Work-hardening measurement

To quantify the change of the flow stress upon stress reversal, different phenomenological param-
eters can be used (Choteau et al., 2005; Han et al., 2005). Such parameters describe the decrease of the
flow stress in the reversed direction as a function of the amount of forward strain. However the deter-
mination of these parameters depends on the strain path and requires the definition of an offset value
and an equivalent plastic strain.
In this paper, the identification of the parameters is carried out by inverse optimization using both
tensile and shear tests with the software SiDoLo (Pilvin, 1988). A cost function LðAÞ is defined in the
least square sense by Eq. (26), and starting from an initial guess of material parameters A0 , the cost
function is minimized with a Levenberg-Marquardt algorithm.
X
N
LðAÞ ¼ Ln ðAÞ ð25Þ
n¼1

1 X Mn
Ln ðAÞ ¼ ðZ s ðA; ti Þ  Z s ðti ÞÞT Dn ðZ s ðA; t i Þ  Z s ðt i ÞÞ ð26Þ
M n i¼1

with N the number of tests in the database, Z s ðA; ti Þ  Z s ðti Þ the gap between experimental and simu-
lated variables at time t i , and Dn a weighting matrix for the test n. The experimental database consists
of tests with two observable variables, namely stress and strain components. A different weighting
coefficient is affected for each of these observable variables, the value of which is chosen according
to the uncertainty on the experimental measures. For the shear stress, the value of the weighting coef-
ficient is Ds12 ¼ 3 MPa and for the tensile tests, Ds11 ¼ 5 MPa and D11 ¼ 0:005. The database is made
up of monotonic tensile tests at 0°, 45°, and 90° to RD, taking into account the transverse strain, of
three monotonic shear tests for the same orientations and of three cyclic shear test in the RD. Several
coefficients are fixed, i.e. L ¼ M ¼ 3, E ¼ 210 GPa, m ¼ 0:29. The viscous contribution is of the order of
20 MPa at a strain rate of 103 s1 and the strain rate sensitivity parameters where fixed to
K v ¼ 22:6 MPa s1=n and nv ¼ 50. The values of the other identified coefficients are given in Table 3.
The results obtained with this set of parameters are presented in Fig. 10 for tensile tests and for
monotonic and Baushinger shear tests in the RD. Good accordance in tension can be observed, as well
as the fact that if the kinematic hardening is taken into account both the monotonic and Bauschinger
tests can be described. It should be emphasized that the Hill 1948 yield criterion fails to reproduce
perfectly both tensile and shear data.
Another method consists in using loading–unloading tensile tests such as the test presented Fig. 11.
The unloading sequence offers a measure of the back stress X and the effective stress ref . In order to
provide this stress partition, the analysis proposed by Dickson et al. (1984) was used (Fig. 12). The re-
verse yield stress rr is obtained on the unloading sequence with a plastic strain offset equal to
5  105 . The concepts of back stress and effective stress have been extensively discussed in the past
(Dickson et al., 1984; Feaugas, 1999; Feaugas and Gaudin, 2004). The back stress which is the direc-
tional component of hardening is linked to the local straining process that introduces long-range

Table 3
Values of material coefficients for the elasto-viscoplastic model of Hill (1948) (r0 ; K; Q x and Hp are expressed in MPa)

Coefficient F G N r0 K n Qx bx Hp
Value 0.64 0.74 3.99 100 426.3 0.30 40841.6 331.3 162.7
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 747

450

Tension
300
Exp.
Sim.
Cauchy stress (MPa)

150 Shear

-150

-300
-0.3 -0.15 0 0.15 0.3
strain
Fig. 10. Identification of tensile and shear tests using the elasto-viscoplastic model of Hill (1948).

450
Cauchy stress (MPa)

300

150

0
0 0.05 0.1
logarithmic strain
Fig. 11. Experimental loading–unloading tensile test on the E220BH steel.

interactions with mobile dislocations. The effective stress ref is an isotropic component which can be
split into a thermal component r and an athermal component rl ; it represents the stress locally re-
quired for a dislocation to move (short-range interactions). In the stress space, the effective stress ref
and the back stress X correspond, respectively, to the radius and to the shift of the yield surface. Then
the partition can be expressed as follows:
r  rr r
ref ¼ þ ð27Þ
2 2
748 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

Fig. 12. Measurement of the isotropic R and kinematic X hardening from tensile unloading curves following the Feaugas (1999)
stress partition (upper left corner).

and
X ¼ r  ref ð28Þ
Fig. 13 presents the results of work-hardening measurement obtained using both previous identifica-
tion methods for a large range of plastic deformation (up to 0.12). Results referred as X and R are
obtained directly by the measurement of the loading–unloading tensile curves. It can be observed
that both methods lead to slightly different results, even if the total time-independent work-harden-
ing (R þ X) is satisfying. However, the back stress magnitude obtained from the simple shear test in-
creases rapidly for small strains and tends to a saturation value of 150 MPa. This saturation value is

450
X tension
X shear
R tension
R shear
R+X shear
R+X tension
Cauchy stress (MPa)

300

150

0
0 0.05 0.1 0.15
equivalent plastic strain
Fig. 13. Work-hardening evolutions obtained using the inverse optimization on tensile and shear tests (lines) or tensile
loading–unloading tests (linepoints).
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 749

not observed in tension. It should be emphasized that in tension, the magnitude of the back stress is
slightly higher than the isotropic contribution. It directly comes from the measurement method; in-
deed, the non-linearity enhanced during unloading is attributed to the kinematic hardening. This is
not the case in simple shear. The results show that even if both methods are strongly different and
comes from different strain paths, the obtained work-hardening evolutions are of the same order of
magnitude.

5. Conclusion

This paper presents a characterization of the kinematic work-hardening on a E220BH steel. A shear
test device has been developed and its use for the characterization of the work-hardening of sheet
metals is presented. It is shown that the device provides a test which is close to a theoretical simple
shear test, even if a small compression strain is measured thanks to the optical method developed on
this purpose that gives all components of the planar strain tensor. Reversal of the shear direction is
performed after several prestrains and the evolution of the kinematic work-hardening with plastic
strain has been identified using a classical anisotropic elastoviscoplastic model of (Hill, 1948). This ap-
proach is compared to another way to identify kinematic hardening using loading–unloading tensile
tests (Feaugas, 1999). The results show a discrepancy between both identification for the isotropic and
the kinematic hardening. However, the results are in agreement concerning the evolution of the global
work hardening.

References

Armstrong, P.J., Frederick, C.O., 1966. A mathematical representation of the multiaxial Bauschinger effect. GEGB Report RD/B/
N731.
Bouvier, S., Haddadi, H., Levée, P., Teodosiu, C., 2006. Simple shear tests: experimental techniques and characterization of the
plastic anisotropy of rolled sheets at large strains. Journal of Materials Processing Technology 172, 96–103.
Bouvier, S., Gardey, B., Haddadi, H., Teodosiu, C., 2006. Characterization of the strain-induced plastic anisotropy of rolled sheets
by using sequences of simple shear and uniaxial tensile tests. Journal of Materials Processing Technology 174, 115–126.
Brunet, M., Morestin, F., Godereaux, S., 2001. Non linear kinematic hardening identification for anisotropic sheet metals with
bending-unbending tests. Journal of Engineering Materials and Technology 123, 378–383.
Choi, Y., Han, C.S., Lee, J.K., Wagoner, R.H., 2006. Modeling multi-axial deformation of planar anisotropic elasto-plastic materials,
part I: theory. International Journal of Plasticity 22, 1745–1764.
Choi, Y., Han, C.S., Lee, J.K., Wagoner, R.H., 2006. Modeling multi-axial deformation of planar anisotropic elasto-plastic materials,
part II: applications. International Journal of Plasticity 22, 1765–1783.
Choteau, M., Quaegebeur, P., Degallaix, S., 2005. Modelling of Bauschinger effect by various constitutive relations derived from
thermodynamical formulation. Mechanics of Materials 37, 1143–1152.
Cleveland, R.M., Ghosh, A.K., 2002. Inelastic effects on springback in metals. International Journal of Plasticity 18, 769–785.
Dickson, J.I., Boutin, J., Handfield, L., 1984. A comparison of two simple methods for measuring cyclic internal and effective
stress. Materials Science and Engineering A 64, L7–L11.
Duchêne, L., Lelotte, T., Flores, P., Bouvier, S., Habraken, A.M., 2008. Rotation of axes for anisotropic metal in FEM simulations.
International Journal of Plasticity 24 (3), 397–427.
Feaugas, X., 1999. On the origin of the tensile flow stress in the stainless steel AISI 316L at 300 K: back stress and effective stress.
Acta Materialia 47, 3617–3632.
Feaugas, X., Gaudin, C., 2004. Ratchetting process in the stainless steel AISI 316L at 300 K: an experimental investigation.
International Journal of Plasticity 20, 643–662.
Fjeldly, A., Roven, H.J., Rauch, E., 1998. Shear deformation properties of extruded AlZnMg alloys. Scripta Materialia 38, 709–714.
Flores, P., Rondia, E., Habraken, A.M., 2005. Development of experimental equipment for the identification of constitutive laws.
International Journal of Forming Processes 8, 117–137.
Flores, P., Duchêne, L., Bouffioux, C., Lelotte, T., Henrard, C., Pernin, N., Van Bael, A., He, S., Duflou, J., Habraken, A.M., 2007. Model
identification and FE simulations: effect of different yield loci and hardening laws in sheet forming. International Journal of
Plasticity 23, 420–449.
Gallée, S., Manach, P.Y., Thuillier, S., 2007. Mechanical behavior of a metastable austenitic stainless steel under simple and
complex loading paths. Materials Science and Engineering A 466, 47–55.
Gau, J.T., Kinzel, G.L., 2001. A new model for springback prediction in which the Bauschinger effect is considered. International
Journal of Mechanical Sciences 43, 1813–1832.
Geng, L., Wagoner, R.H., 2002. Role of plastic anisotropy and its evolution on springback. International Journal of Mechanical
Sciences 44, 123–148.
Gilormini, P., 1994. Sur les référentiels locaux objectifs en mécanique des milieux continus. Comptes Rendus de l’Académie des
Sciences de Paris 318, 1153–1159.
Green, A.E., Zerna, W., 1968. Theoretical Elasticity, 2nd ed. Oxford University Press.
G’Sell, C., Boni, S., Shrivastava, S., 1983. Application of the plane simple shear test for determination of the plastic behavior of
solid polymers at large strains. Journal of Materials Science 18, 903–918.
750 S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751

Haddadi, H., Bouvier, S., Banu, M., Maier, C., Teodosiu, C., 2006. Towards an accurate description of the anisotropic behavior of
sheet metals under large plastic deformations: Modelling, numerical analysis and identification. International Journal of
Plasticity 22, 2226–2271.
Haddag, B., Balan, T., Abed-Meraim, F., 2007. Investigation of advanced strain-path dependent material models for sheet metal
forming simulations. International Journal of Plasticity 23, 951–979.
Hahm, J.H., Kim, K.H., 2008. Anisotropic work hardening of steel sheets under plane stress. International Journal of Plasticity 24,
1097–1127.
Han, K., Van Tyne, C.J., Levy, B.S., 2005. Effect of strain and strain rate on the Bauschinger effect response of three different steels.
Metallurgical and Materials Transactions A 36, 2379–2384.
Hill, R., 1948. A theory of the yielding and plastic flow of anisotropic metals. Proceedings of the Royal Society of London A 193,
281–297.
Hu, Z., Rauch, E.F., Teodosiu, C., 1992. Work-hardening behavior of mild steel under stress reversal at large strains. International
Journal of Plasticity 8, 839–856.
Hu, W., 2005. An orthotropic yield criterion in a 3D general stress state. International Journal of Plasticity 21, 1771–1796.
Kuwabara, T., Morita, Y., Miyashita, Y., Takahashi, S., 1995. Elastic-plastic behavior of sheet metal subjected to in-plane
reverse loading. Proceedings of Plasticity95 Dynamic Plasticity and Structural Behaviors. Gordon and Breach, London. pp.
841–844.
Lee, M.G., Kim, D., Kim, C., Wenner, M.L., Wagoner, R.H., Chung, K., 2005. Springback evaluation of automotive sheets based on
anisotropic-kinematic hardening laws and non-quadratic anisotropic yield functions. Part II: characterization of material
properties. International Journal of Plasticity 21, 883–914.
Lee, M.G., Kim, S.J., Wagoner, R.H., Chung, K., Kim, H.Y., 2008. Constitutive modeling for anisotropic/asymmetric hardening
behavior of magnesium alloy sheets. International Journal of Plasticity 24, 545–582.
Lemaitre, J., Chaboche, J.L., 1996. Mécanique des matériaux solides. Editions Dunod, Paris France.
Li, X., Yang, Y., Wang, Y., Bao, J., Li, S., 2002. Effect of the material-hardening mode on springback simulation accuracy of V-free
bending. Journal of Materials Processing and Technology 123, 209–211.
Manach, P.Y., 1989. Analyse par simulations numériques des non-homogénéités au cours d’essais mécaniques. Master’s Thesis,
Grenoble France.
Manach, P.Y., Favier, D., 1997. Shear and tensile thermomechanical behavior of near equiatomic NiTi alloy. Materials Science
and Engineering A 222, 45–57.
Manach, P.Y., Couty, N., 2002. Elastoviscohysteresis constitutive law in convected coordinate frames: application to finite
deformation shear tests. Computational Mechanics 28, 17–25.
Miyauchi, K., 1977. A proposal of a planar simple shear. Journal of Materials Science 18, 903–918.
Miyauchi, K., 1992. Deformation path effect on stress–strain relation in sheet metals. Journal of Materials Processing Technology
34, 195–200.
Mohanraj, J., Barton, D.C., Ward, I.M., Dahoun, A., Hiver, J.M., G’Sell, C., 2006. Plastic deformation and damage of
polyoxymethylene in the large strain range at elevated temperatures. Polymer 47, 5852–5861.
Morestin, F., Boivin, M., 1996. On the necessity of taking into account the variation in the Young modulus with plastic strain in
elastic-plastic software. Nuclear Engineering and Design 162, 107–116.
Narasimhan, N., Lovell, M., 1999. Predicting springback in sheet metal forming: an explicit to implicit sequential solution
procedure. Finite Elements in Analysis and Design 33, 29–42.
Numisheet, 1996. In: Lee, J.K., Kinzel, G.L., Wagoner, R.H. (Eds.), Proceedings of the 3rd International Conference on Numerical
Simulation of 3D Sheet Metal Forming Processes. Dearborn, Michigan, USA.
Numisheet, 1999. In: Gelin, J.C., Picart, P. (Eds.), Proceedings of the 4th International Conference and Workshop on Numerical
Simulation of 3D Sheet Forming Processes. Besançon, France.
Numisheet, 2002. In: Yang, D.Y., Oh, S.I., Huh, H., Kim, Y.H. (Eds.), Proceedings of the 5th International Conference and Workshop
on Numerical Simulation of 3D Sheet Forming Processes. Jeju Island, Korea.
Numisheet, 2005. In: Smith, L.M., Pourboghrat, F., Yoon, J.W. (Eds.), Proceedings of the 6th International Conference and
Workshop on Numerical Simulation of 3D Sheet Forming Processes. Detroit, Michigan, USA.
Oliveira, M.C., Alves, J.L., Chaparro, B.M., Menezes, L.F., 2007. Study on the influence of work-hardening modeling in springback
prediction. International Journal of Plasticity 23, 516–543.
Pilvin, P., 1988. Identification des paramètres de modèles de comportement. International Seminar on the inelastic behavior of
solids Mécamat, Besançon, France, 155–161.
Rauch, E.F., G’Sell, C., 1989. Flow localization induced by a change in strain path in mild steel. Materials Science and Engineering
A 111, 71–80.
Rauch, E.F., 1998. Plastic anisotropy of sheet metals determined by simple shear tests. Materials Science and Engineering A 241,
179–183.
Sidoroff, F., Dogui, A., 2001. Some issues about anisotropic elastic–plastic models at finite strain. International Journal of Solids
and Structures 38, 9569–9578.
Suprun, A.N., 2006. A constitutive model with three plastic constants: the description of anisotropic work hardening.
International Journal of Plasticity 22, 1217–1233.
Tang, S., 1990. An anisotropic hardening rule for the analysis of sheet metal forming operations. Advanced Technology of
Plasticity 3, 1149–1154.
Teodosiu, C., Hu, H.C., 1998. Microstructure in the continuum modelling of plastic anisotropy. In: 19th Riso International
Symposium on Materials Science, pp. 445–455.
Thuillier, S., Le Maoût, N., Manach, P.Y., Debois, D., 2008. Numerical simulation of the roll hemming process. Journal of Materials
Processing Technology 198, 226–233.
Tong, W., 2006. A plane stress anisotropic plastic flow theory for orthotropic sheet metals. International Journal of Plasticity 22,
497–535.
Wack, B., Tourabi, A., 1993. Cyclic simple shear of metallic sheets: application to aluminum–lithium alloy. Journal of Materials
Science 28, 4735–4743.
S. Thuillier, P.Y. Manach / International Journal of Plasticity 25 (2009) 733–751 751

Yang, M., Akiyama, Y., Sasaki, T., 2004. Evaluation of change in material properties due to plastic deformation. Journal of
Materials Processing Technology 151, 232–236.
Yeganeh, M., 2007. Incorporation of yield surface distortion in finite deformation constitutive modeling of rigid-plastic
hardening materials based on the Hencky logarithmic strain. International Journal of Plasticity 23, 2029–2057.
Yoon, J.W., Barlat, F., Gracio, J.J., Rauch, E., 2005. Anisotropic strain hardening behavior in simple shear for cube textured
aluminum alloy sheets. International Journal of Plasticity 21, 2426–2447.
Yoshida, F., Urabe, M., Toropov, V.V., 1998. Identification of material parameters in constitutive model for sheet metals from
cyclic bending tests. International Journal of Mechanical Science 2-3, 237–249.
Yoshida, F., Uemori, T., Fujiwara, K., 2002. Elastic-plastic behavior of steel sheets under in-plane cyclic tension–compression at
large strain. International Journal of Plasticity 18, 633–659.
Zhao, K.M., Lee, J.K., 2002. Finite element analysis of the three-point bending of sheet metals. Journal of Materials Processing
Technology 122, 6–11.

You might also like