Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

International Journal of Plasticity

16 (2000) 1105±1129
www.elsevier.com/locate/ijplas

Prediction of localized thinning in sheet metal


using a general anisotropic yield criterion
Jian Cao a,*, Hong Yao a, Apostolos Kara®llis b,
Mary C. Boyce b
a
Department of Mechanical Engineering, Northwestern University, Evanston, IL 60208, USA
b
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

Received in final revised form 8 December 1999

Abstract
The forming limit diagram (FLD) is a useful concept for characterizing the formability of sheet
metal. The ability to accurately predict the FLD for a given material has been shown to depend
on the shape of the selected yield function. In addition, both experimental and numerical results
have shown that the level of the FLD is strongly strain path dependent. In this work, a combi-
nation of Marciniak±Kuczynski (M±K) analysis and a general anisotropic yield criterion devel-
oped by Kara®llis and Boyce (Kara®llis, A.P., Boyce, M.C., 1993. A general anisotropic yield
criterion using bounds and transformation weighting tensor. J. Mech. Phys. Solids 41, 1859) is
used to predict localized thinning of sheet metal alloys for linear and nonlinear strain paths. A
new method for determining the constants in the yield criterion is proposed. The optimal values
are obtained by ®tting the initial yield stresses and calculated FLD under linear strain paths with
the experimental measurement. Using this approach, accurate yield functions can be de®ned
for both Al2008-T4 and Al6111-T4. Comparisons of computed FLDs with the experimental
data of Graf and Hosford (Graf, A., Hosford, W.F., 1993b. E€ect of changing strain paths on
forming limit diagrams of Al 2008-T4. Metall. Trans. A. 24, 2503; Graf, A., Hosford, W.F.,
1994. The in¯uence of strain path changes on forming limit diagrams of Al 6111-T4. Int. J.
Mech. Sci. 36, 897) show good agreements. # 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Cutting and forming; Anisotropic material; Constitutive behavior; Stability and bifurcation;
Tearing prediction

1. Introduction

The usefulness of numerical simulation in designing sheet metal forming processes


requires good criteria to identify the modes of failure such as tearing, wrinkling and
* Corresponding author. Tel.: +1-847-467-1032; fax: +1-847-491-3915.
E-mail address: jcao@nwu.edu (J. Cao).

0749-6419/00/$ - see front matter # 2000 Elsevier Science Ltd. All rights reserved.
PII: S0749-6419(99)00091-1
1106 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

geometry deviation due to springback. Wrinkling and tearing are phenomena of


material and structure instability while springback is a result of elastic recovery
upon unloading. This work examines the failure caused by tearing due to excessive
thinning under various strain paths.
To determine the amount of deformation a material can withstand prior to necking
failure, Keeler and Backofen (1964) and Goodwin (1968) introduced the concept of the
forming limit diagram (FLD). The maximum strains that can be sustained by sheet
materials prior to the onset of localized necking are generally referred to as the forming
limit strains. A plot of the major and minor limit strains in the principal strain space
constitutes a FLD. A typical FLD has a forming limit curve (FLC) as shown in Fig. 1.
Material subjected to linear in-plane tension (plane stress state with a constant ratio of
principal strains) can be safely stretched until the major strain reaches the FLC. By
comparing the strains measured in the formed part to the FLC, the severity and
nature of the deformation can be assessed and process parameters such as lubricants
or draw beads can be designed accordingly to assist the forming operation.
Commonly, FLDs are based on linear strain paths where the ratio between the
major and minor principal strains is constant throughout the deformation process.
Experiments have shown that changes to the strain path could substantially increase
or decrease the levels of FLC. In practice, multi-step forming operations, geometry
of the tooling and interface friction will induce changes in the loading path, many of
which could be signi®cant. Hence, to assess the formability in industrial stamping
operations of complex panels, FLDs for nonlinear strain paths are a necessary tool.
Keeler (1965,1968) introduced the experimental method to obtain the FLD. How-
ever, the experimental determination of the forming limits in all sheet metal forming
processes is not only tedious and expensive but also nearly impossible, since the
strain paths of material points are quite non-linear and distinguished from each

Fig. 1. A typical forming limit diagram (FLD).


J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1107

other. Therefore, theoretical and numerical methods of accurately predicting the


FLD become highly relevant.
Two basic approaches used in the past are the bifurcation analysis following the
work of Hill (1952) and the localization analysis introduced by Marciniak and
Kuczynski (1967). Using the ¯ow theory of plasticity, Hill (1952) ®rst developed a
general criterion for localized necking in thin sheets under plane stress conditions.
His theory predicts that localized necking occurs along a direction of zero extension
and is restricted to negative minor strain value. Therefore, it did not allow for loca-
lized necking of materials with smooth yield surface under biaxial stretching condi-
tions. StoÈren and Rice (1975) used a bifurcation analysis based on a simpli®ed
constitutive model of a pointed vertex on subsequent yield loci to predict the entire
FLD. They pointed out that the development of a vertex on the yield locus is
responsible for the onset of localized necking in thin sheets under biaxial tension.
The shape of both the left and right hand sides of the forming limit diagram can be
calculated by another procedure suggested by Marciniak and Kuczynski (1967)
(referred to as M±K analysis). In the M±K analysis, they introduced thickness
imperfections normal to the principal stress and strain direction as a groove simu-
lating preexisting defects in the material. Necking was considered to occur when the
ratio of the thickness in the groove to the nominal thickness is below a critical value.
More detailed review is presented in Section 2.
Research has shown that the calculated forming limit strains using M±K analysis
depend sensitively on the choice of yield surface. Localization occurs when plane
strain deformation takes place within the imperfection (i.e. the ratio of minor to
major strain increments becomes 0). Therefore, the curvature of the yield locus
determines how easy the change of strain ratio could be (Chan, 1989) and conse-
quently the level of the forming limit curve (FLC). Most of the early M±K calcula-
tions of FLDs were based on the quadratic isotropic Mises and/or anisotropic Hill
yield functions. Sowerby and Duncan (1971) used Hill's (1948) yield criterion to
calculate the FLD and showed that the right hand side FLD increases with an
increase in the strain hardening exponent n value and strongly depends on the nor-
mal anisotropic index R value. Comparison of the numerical and experimental FLD
showed that the Hill (1948) yield function over-predicted the FLC, especially for
aluminum material. A number of nonquadratic yield functions have been proposed
to describe material behavior. Those include the Hosford's (1972) nonquadratic
planar isotropic function, the Hill's (1979) high exponent yield function, the
Kara®llis and Boyce's (1993) general yield function (hereinafter referred to as K±B
yield criterion) and Barlat's general yield functions (Yld89, Yld91, Yld94 and Yld97)
(Barlat and Lian, 1989; Barlat et al., 1991,1997a,b). Graf and Hosford (1993a,1994)
calculated the right hand side of the FLD using M±K analysis with the non-quad-
ratic Hosford yield criterion. The stress exponent used in the yield criterion was 8 for
the aluminum alloys studied. The computed FLD was found to be raised by
increasing n and strain rate sensitivity index m, and was una€ected by the R value.
The predicted FLD, although still higher, was in better agreement with the experi-
mental FLD compared to the results obtained by using the Hill (1948) yield criter-
ion. Barlat et al. (1997c) computed the FLDs of one aluminum and two steel sheet
1108 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

materials using the Yld94 yield function. The predictions match the experimental
FLD reasonably well. Accurate forming limit diagrams have also been calculated
using textured polycrstalline yield surfaces computed by the Taylor or Bishop±Hill
approach. Barlat (1987,1989) and Chan (1989) computed the FLD using textured
polycrystalline yield surfaces and concluded that accurate predictions of the FLD
are possible at a relatively higher computation cost when microstructure of the
material is taken into account.
Some recent progresses made on the prediction of forming limits of metal sheets
are also worthy of attentions. Using M±K analysis, Mesrar et al. (1998) predicted
the necking failure of planar-isotropic viscoplastic materials subjecting to linear,
bilinear and curvilinear strain paths. A plane stress yield criterion proposed by Fer-
ron et al. (1994) is used in the numerical prediction of forming limits. Instead of
using the plot of limit principal strains, the author plotted the forming limit using
the e€ective strain as a function of the strain rate ratio. The limit curve obtained is
considered having good accuracy as an intrinsic curve that de®nes the forming limit
independent of the strain paths. An empirical expression ®tting this new limit curve
has been proposed for both positive and negative values of the strain ratios. In
another work, Hiwatashi et al. (1998) predicted FLDs using M±K analysis and an
anisotropic constitutive model based on texture and dislocation structure. The e€ect
of strain path was studied both for proportional and two stage straining. It has been
shown that forming limit for the case of two-stage straining depends on the amount
of the ®rst strain and the combination of the ®rst and second strain rate modes. The
proposed anisotropic hardening model provides forming limit predictions that
re¯ected the experimental tendencies which cannot be predicted by isotropic hard-
ening model. This is most pronounced for the case of strain paths changing from
equi-biaxial to uniaxial tension. The initial anisotropy and anisotropic hardening
including transient hardening are found to be the most important factors that a€ect
the accuracy of FLD prediction.
This paper examines failure detection using the M±K analysis together with the
K±B yield function. A brief review of M±K analysis, including its theoretical back-
ground and important contributions from previous works, is presented in Section 2.
A brief review of the anisotropic yield criterion developed by Kara®llis and Boyce
(1993) is given in Section 3. Section 4 proposes a new method for the determination
of the material constants in a yield function. The yield surface shape was determined
by matching numerical and experimental initial yield stresses and FLC for the case
when the major principal strain direction coincides with the transverse direction of
the material. Compared to the predicted values using the Hill and Hosford yield
criteria, the anisotropic model established in this work is found to predict the limit
strains on the right hand side of the FLD in much better agreement with the
experimental results for Al2008-T4. As demonstrated in Section 5, using the yield
surface that provides best results for the linear strain paths, improvement in the
predictions of FLC for nonlinear paths were obtained compared to the results given
in Graf and Hosford (1993a,b). Finally, the same approach is utilized for determin-
ing the yield function of Al6111-T4 and the comparisons of computed FLDs and the
experimental data of Graf and Hosford (1994) was presented in Appendix C.
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1109

2. Review of Marciniak and Kuczynski analysis

The original Marciniak and Kuczynski analysis, as discussed in Section 1, is based


on a simpli®ed model with assumed pre-existing thickness imperfection in the form
of a groove perpendicular to the principal strain directions as shown in Fig. 2 for the
case of  ˆ 0: The sheet is composed of the nominal area and weak groove area,
which are denoted by `a' and `b', respectively. The initial imperfection factor of the
groove, f0 , is de®ned as the thickness ratio f0 ˆ tb0 =ta0 ; where `t' denotes the thickness
and subscript `0' denotes the initial state. A biaxial stress state is imposed on the
nominal area and causes the development of strain increments in both the nominal
(a) and the weak area (b).
Assuming that compatibility and equilibrium are satis®ed across the imperfection,
the strain increment inside the groove, d"b ; is then greater than that outside the
groove, d"a : Eventually, d"b is so large that plane strain deformation occurs inside
the groove leading to localization failure.
Further research has shown that assuming that the imperfection is always oriented
perpendicular to the principal direction of loading is a severe restriction and leads to
inaccurate prediction of the FLD in the negative minor strain region for the general
case of anisotropic materials. Hutchinson and Neale (1978a,b) examined the case of
an imperfection in the form of a groove inclined at an angle  with respect to the
principal strain directions as shown in Fig. 2. They found that for each proportional
strain path at the left hand side of FLD, there exists a de®nite imperfection orien-
tation angle which gives the minimum limiting strain. This ®nding was further con-
®rmed by many researchers including Chan (1989) and Chan et al. (1984) and Lian
and Baudelet (1987). In the general case of anisotropic materials, Barata Da Rocha
(1989) pointed out that the hypothesis of normality between the groove and the
major principal stress axis is not valid because the limiting strains are sensitive to the
orientation of the groove, even on the right hand side of the FLD.
In the modi®ed M±K analysis (see Fig. 2), the angle  between the imperfection
and one of the principal directions is updated from an initial value 0 at each incre-
ment of the plastic deformation as

Fig. 2. Model of imperfection in the M±K analysis.


1110 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

1 ‡ d"a1
tan… ‡ d† ˆ tan…† …1†
1 ‡ d"a2

where d"a1 and d"a2 are the major and minor principal strain increments in the plane
of the sheet of the nominal area, respectively. The basic equations for the M±K
analysis are the geometric compatibility equations expressed as:

d"att ˆ d"btt …2†

and the force equilibrium equations across the imperfection groove, i.e.

Fann ˆ Fbnn …3†

Fant ˆ Fbnt …4†

where subscripts n and t denote the normal and tangential directions of the groove,
respectively, and F is the force per unit width in the t direction, i.e.
 a a "a3 a
Fnn ˆ nn e t0
b b "b3 b …5†
Fnn ˆ nn e t0
 a
a "3 a
Fant ˆ nt e t0
b b "b3 b …6†
Fnt ˆ nt e t0

The physical basis for the M±K analysis was well presented in McCarron et al.,
(1988). In their study, imperfections in the form of grooves were planted in samples
used in equibiaxial stretching. It was found that no reductions in the forming limit
strain were obtained with shallow grooves for which the imperfection indices were
greater than 0.990 and 0.992 for two di€erent steels. These f0 values represent the
pre-existing micro-structural defects in the two steels. In the M±K analysis, f0 is used
as an adjustable parameter to match the calculated FLD0, the limiting strain at
plane strain condition, with the experimental data.
The failure criterion in the M±K analysis can be de®ned in many ways. In the work of
Marciniak and Kuczynski (1967), necking was considered to occur when the weakness
factor …tb =ta † dropped below a critical value. Barata Da Rocha et al. (1984) associated
the necking in the M±K analysis to the ratio of the e€ective strain increment in the
groove to that in the nominal area. In their calculation, necking occurred when this
ratio was greater than 10. Similarly, Narasimhan and Wagoner (1991) suggested
necking occurred when the principal strain in the groove area was 10 times that in
the nominal area. In this work, the deformation severity indices are de®ned as

fnn ˆ d"bnn =d"ann
…7†
fnt ˆ d"bnt =d"ant

and necking occurs when either fnn or fnt is greater than 10.
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1111

Experimental and numerical analyses have shown that changes of strain path
could substantially increase or decrease the levels of FLC depending on a particular
strain path. Barata Da Rocha (1989) and Barata Da Rocha et al. (1984) numerically
investigated the FLD of anisotropic sheets in linear and non-linear loading. They have
shown that uniaxial prestrain followed by biaxial stretching considerably increases the
limit strains at the onset of necking. Conversely, premature instabilities were observed
for strain paths consisting of prior biaxial prestrain followed by uniaxial tension. In
addition, a change in strain path towards plane strain resulted in a signi®cant loss of
plastic stability. In their work, the Hill (1948) yield criterion was used.
The trend was found valid in Al2008-T4 and Al6111-T4 in experiments conducted
by Graf and Hosford (1993b,1994). Furthermore, they found that prestraining in
plane strain increases forming limits in both subsequent biaxial and uniaxial tension.
In addition, they tested the e€ect of changing the direction of the maximum principal
strains and concluded that rotating the directions of the principal strain lowers the
forming limits for the case of prestraining in uniaxial or plane strain tension. Graf and
Hosford (1993a) calculated the forming limit diagrams of Al2008-T4 for changing
strain paths using the M±K analysis with a high exponent Hosford yield criterion.

3. Review of Kara®llis and Boyce yield criterion

In this work, the yield surface of the anisotropic material with orthotropic sym-
metry is described by a non-quadratic yield criterion developed by Kara®llis and
Boyce (1993). The K±B yield criterion was constructed by mixing two yield func-
tions, 1 and 2 . As shown in Eq. (8), 1 represents a yield locus located between the
Tresca yield locus and the von Mises yield locus and 2 varies from the von Mises to
a theoretical upper bound as m changes from 2 to 1.

where

 ˆ …1 ÿ c†1 ‡ c2 ˆ 2Ym


1 ˆ jS1 ÿ S2 jm ‡jS2 ÿ S3 jm ‡jS3 ÿ S1 jm …8†
2 ˆ 3m =…2mÿ1 ‡ 1†…jS1 jm ‡jS2 jm ‡jS3 jm †

and Si is the principal values of the isotropic plasticity equivalent (IPE) stress tensor
as further de®ned below and Y is the average yield stress in uniaxial tension
obtained experimentally.
A fourth order tensorial operator, L, introduces the material anisotropy, i.e.

S ˆ L …9†

where  is the Cauchy stress in the anisotropic material, S is the IPE stress tensor,
and L is a fully symmetric and traceless fourth order tensor.
The constitutive equations using the above yield function and the associated ¯ow
rule are expressed in the principal axes of orthotropic symmetry as:
1112 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

@
dp ˆ l …10†
@S
p
where l is a scalar related to the plastic work rate, d is the plastic work conjugate of
the IPE deviatoric stress S. The plastic work rate is:
: :
W ˆ   Dp ˆ S  dp ˆ Y" …11†
:
where W is the plastic work rate, Dp is the symmetric
: part of the plastic velocity
gradient tensor, Y is the equivalent stress and " is the equivalent strain rate. Then by
@
using the homogeneity of ; (S  @S ˆ m), combined with Eq. (8), we obtain:
: @
W ˆ lS  ˆ 2lm m …12†
@S
Therefore,
:
W
lˆ …13†
2mYm
and
: :
W @ W @
dp ˆ ˆ QT Q …14†
2mYm @S 2mYm @
where Q is the rotation matrix whose rows are the eigenvectors of S and K is the
principal tensor of S obtained as:

X
3
K ˆ Si vi
vi …15†
iˆ1

where i are the unit vectors.


Also, from the ¯ow normality rule and by using the symmetry of L, we obtained:
: :
p W @ W @
D ˆ m
ˆ m
L ˆ Ldp …16†
2mY @ 2mY @S

The constitutive relation for the deformation can be expressed as:


:
 ˆ C…D ÿ Dp † …17†

where C is the elastic constant for the anisotropic material, D is the symmetric part
of the velocity gradient.
Another important feature of K±B yield function is that it allows the yield stresses
in tension and compression to be di€erent by o€setting the center of the yield sur-
face with backstresses. Expending Eq. (9) , we have

S ˆ L… ÿ B† …18†
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1113

where B is an irreducible symmetric traceless tensorial state variable of the second


order. The ¯ow rule in this case remains as given by Eq. (16). Detailed discussion
was given by Kara®llis and Boyce (1993).

4. Yield surface determination

To describe the anisotropic behavior of a given material, we can adjust material


constants in the K±B yield criterion, which are the mixing factor c, the stress exponent
m, the fourth order tensor L and backstress B. After c; m and B are known, components
in the tensor L can be determined by ®tting the experimental R ratios, setting the value
of equivalent yield stress Y to be the average yield stress, and ®tting the predicted initial
yield stresses in 0, 45 and 90 directions closest to the experimental measurements.
Consequently, the K±B yield criterion predicts the yield locus.1 In the work of Kara®llis
and Boyce (1993), two methods were suggested for obtaining c and m based on experi-
mental material data. In the ®rst method, the stress exponent m is ®xed to a high value
(m ˆ 30) to provide a nearly lower and upper representation of 1 and 2 in Eq. (8).
The mixing coecient c is modi®ed until the desired values of the initial yield stresses
are best reached. In the second method, both m and c are modi®ed and determined to
provide a best ®t to the Lode's (1926) diagram as derived from the precise description of
Bishop±Hill yield surface. As discussed earlier, the FLD depends sensitively on the
shape of the yield locus. Although the di€erence in the yield surface obtained by each of
the above two methods is subtle, the corresponding FLD predictions could be di€erent.
Here, a new method is proposed where the combination of m and c in Eq. (8) can be
selected by comparing the right hand side of the calculated FLD with the experimental
FLD from linear loading paths.
The calculation procedure established for the selection of material constants is shown
as a ¯ow chart in Appendix A. In this procedure, various combinations of c and m are
tested. For each combination, the linear transformation tensor L is calculated in terms of
the R ratios obtained from the uniaxial tensile tests in 0, 45 and 90 to the rolling direc-
tion. The imperfection factor f used in the M±K analysis is obtained by matching the
calculated FLD0 to the experimental value. The complete FLD corresponding to that
particular c and m combination can then be calculated by M±K analysis with rotations of
the groove. The calculation procedure and the basic equations involved are described in
Section 3. The ¯ow chart for the calculation of FLD with rotation of groove is given in
Appendix B. The best combination, which FLD is closest to the experimental FLD, can
then be selected. We believe that the yield surface corresponding to that combination of c
and m best describes the actual material behavior. In case such a combination cannot
make the predicted yield stress distribution match the experimental yielding stress at 0, 45
and 90 directions, backstress B can be added to the yield function. Once the backstress is
determined, procedures in Appendix A need to be repeated with the backstress to calcu-
late the combination of c and m that enables the match of FLD.

1
The alternative method to determine tensor L is by ®tting initial yield stresses in the three directions
for a given combination of c and m. In this case, the K±B criterion will predict R ratios.
1114 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

In this work, the yield surfaces of Al2008-T4 and Al6111-T4 are determined by using
the above method. Results for Al2008-T4 are presented here and results for Al6111-T4
are given in Appendix C. The Al2008-T4 tensile properties are obtained from the
experimental data published in Graf and Hosford (1993a,b) and are listed in Table 1.
Fig. 3a shows the calculated yield locus of Al 2008-T4 using Hill (1948), Hosford
(1972) and K±B yield criterion. Fig. 3b and c show the predicted R values and nor-
malized yield stresses versus the angle to the rolling direction for Al2008-T4 using two
combinations of c and m values in K±B yield criterion along with the predictions from
the Hill (1948) and the Hosford yield criteria. Fig. 4 shows the corresponding right
hand side of the FLDs. Notice that the two combinations of c and m with backstresses
for the K±B yield function provide identical R values in Fig. 3b and almost identical
yield stresses in Fig. 3c but generate di€erent FLDs (Fig. 4). Table 2 lists the non-
zero components of the linear operator L and backstress components B11=0.018Y
and B22=ÿ0.024Y when c=0.63 and m=26. For the case when c=0.5 and m=16,
B11=0.02Y and B22=ÿ0.027Y. Therefore, to best describe material behavior, it is
recommended that the FLD be used to determine the material constants.
The calculated FLDs are compared with the numerical results given by Graf and
Hosford (1993a) and by using the Hill (1948) yield criterion. The results demonstrated
that the forming limit at the right hand side drops with the increase of the stress expo-
nent and the FLD0 (the plane strain value) is independent of the yield surface shape.
Very good agreement is obtained between numerical results and experimental data
when c=0.63 and m=26 used with non-zero backstress in the K±B yield criterion. The
imperfection factor of 0.9909 is taken to match the FLD0 at the transverse direction.
From Fig. 42, we can conclude that K±B yield criterion with the right material
constants can accurately predict the FLC with major strains in the transverse direc-
tion using M±K analysis. The predicted FLC with major strains in the rolling
direction are 3±5% lower than the experimental value. Graf and Hosford
(1993b,1994) attributed the di€erence in FLD0s in rolling and transverse directions
to the ridging oriented to the rolling direction on the surface. When loading is

Table 1
Tensile properties of Al2008-T4

Angular o€set from rolling direction Yield stress (MPa) R values

0 160 0.58
45 150 0.485
90 146 0.78
Young's modulus 70.0 (GPa)
Poisson's ratio 0.33
Strain hardening  ˆ 530:0…0:008 ‡ p †0:25 (MPa)

2
The meaning of abbreviations used in the legends of the curves in Figs. 4±8 are as follows: KB,
numerical analysis using K±B function; Hosford, numerical analysis using Hosford yield function; Exp,
experimental results; bia, prestraining in equal biaxial tension; uni, prestraining in uniaxial tension; pln,
prestraining in plane stain tension; mix, prestraining and test are conducted in two di€erent directions;
Exp-Al2008, linear strain path (reference curve).
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1115

Fig. 3. (a) Calculated yield locus of Al2008-T4 using Hill (1948), Hosford (1972) and K±B yield criterion.
(b) Comparison of the R values with respect to the angle from the rolling direction for Al2008-T4 using
various yield criteria [experimental data was published by Graf and Hosford (1993b) and the reported R
value distribution in Hosford m=8 was obtained from Graf and Hosford (1993a)]. (c) Comparison of the
normalized yield stress with respect to the angle from the rolling direction for Al2008-T4 for di€erent
combinations of c and m values (Y is the averaged yield stress).
1116 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

Fig. 4. Comparisons of the experimental and numerical results of FLDs calculated by using various yield
criteria for Al2008-T4 [experimental data was published by Graf and Hosford (1993a,b)].

Table 2
Constants used in the K±B theory for Al2008-T4

Variable Value Variable Value Variable value

L1111 0.667 L2222 0.680 L3333 0.698


L1122 ÿ0.325 L2233 ÿ0.355 L1133 ÿ0.343
L1212 0.476 c 0.63 m 26
B11 0.0184Y B22 ÿ0.0245Y B33 0

applied at the transverse direction, the ridges and valleys parallel to the rolling
direction develop and help to lower the forming limits in the experiments. In our
simulation, the surface condition is considered isotropic and therefore cannot
account for the FLD0 di€erence other than that caused by the material planar
anisotropy.

5. M±K analysis for changing strain paths

As discussed in the introductory section and demonstrated in Fig. 4, the calculated


forming limit strains depend sensitively on the choice of yield surface. In this section,
the e€ectiveness of our method to determine the yield surface and the accuracy of
K±B yield criterion are further examined by calculating FLDs of Al2008-T4 with
nonlinear strain paths. The FLDs for various levels of prestraining in uniaxial,
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1117

biaxial and plane strain tension are calculated by using M±K analysis with the
rotation of groove and the K±B yield criterion with the constants listed in Table 2.
The results are compared with the experimental and numerical results of Graf
and Hosford (1993a,b). The general tendency discovered by Graf and Hosford as
discussed in Section 2 of this work is con®rmed. Detailed discussions for each case
of nonlinear path FLD prediction are given in the following.

5.1. FLDs of sheets prestrained in equal biaxial tension

Fig. 5 shows the comparisons of the experimental and numerical FLDs pre-
strained in equal biaxial tension and tested with major strains perpendicular to the
rolling direction. The FLDs shifted to the right and down with respect to the original
FLD. Results of the experiment and our calculation matched very well qualitatively.
Quantitatively, the results are much better than that obtained by using Hosford
yield criterion, which over-predicted the FLDs compared to their experimental
results. The quantitative di€erence between the numerical and experimental results
for FLDs of this and other plots to be shown for changing strain paths could be
caused by two reasons. First, as mentioned in the last section, the surface ridging
development that occurred in the test is not accounted for in the calculation. Sec-
ondly, in the experiments conducted by Graf and Hosford (1994), there was a time
delay between the prestrain and the following tests. For the aluminum material, this
idle time can cause property changes in the material, so called time recovery e€ect,
and consequently cause the rise of forming limits. In Fig. 5, for example, the fact
that the failure strain for the case of the equal biaxial prestraining of 0.17 followed

Fig. 5. Comparisons of the experimental and numerical FLDs of Al2008-T4 prestrained in equal biaxial
tensions of 0.04 and 0.17 and tested perpendicular to the rolling direction [Hosford's numerical results and
experimental data was published by Graf and Hosford (1993a,b)].
1118 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

Fig. 6. Comparisons of the experimental and numerical FLDs of Al2008-T4 prestrained in plane strain
tensions of 0.08 and 0.13 perpendicular to the rolling direction and tested perpendicular to the rolling
direction [Hosford's numerical results and experimental data was published by Graf and Hosford
(1993a,b)].

by plane strain stretch is higher than numerical results could be caused by the time
recovery e€ect in the experiment. Our numerical calculation cannot take this e€ect
into account unless additional testing of material behavior after the time delay is
available.

5.2. FLDs of sheets prestrained in plane strain tension

Fig. 6 shows the comparisons of the experimental and numerical FLDs pre-
strained in plane strain tension in the transverse direction and tested with major
strains in the transverse direction. Both experimental and numerical results show
that prestraining in plane strain increases forming limits in the subsequent uniaxial
and biaxial tension. Once again the FLDs predicted by using K±B yield criterion is
lower than the results of Graf and Hosford (1993a) and closer to the experimental
results.

5.3. FLDs of sheets prestrained in uniaxial tension

Fig. 7 shows the comparisons of the experimental and numerical FLDs pre-
strained in uniaxial tension perpendicular to the rolling direction and tested in the
same direction. In both the experiments and calculations the FLDs shifted to the left
and up in the biaxial stretching.
Fig. 8 shows the case of the experimental and numerical FLDs prestrained in
uniaxial tension parallel to the rolling direction and tested perpendicular to the
rolling direction. In this case the rotation of principal strain directions causes reverse
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1119

Fig. 7. Comparisons of the experimental and numerical FLDs of Al2008-T4 prestrained in uniaxial ten-
sions of 0.05 and 0.12 perpendicular to the rolling direction and tested perpendicular to the rolling direc-
tion [Hosford's numercal results and experimental data was published by Graf and Hosford (1993a,b)].

Fig. 8. Comparisons of the experimental and numerical FLDs of Al2008-T4 prestrained in uniaxial ten-
sions of 0.04 and 0.18 parallel to the rolling direction and tested perpendicular to the rolling direction
[experimental data was published by Graf and Hosford (1993b)].

of loadings on the sheet. In contrast to the case in Fig. 7 the FLDs shift to the right
and lower with the increase in the level of prestraining. The FLC predicted by using
K±B yield function is much closer to the experimental results compared to the FLC
calculated by using Hosford yield function for the case of prestraining of 0.04. The
1120 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

FLCs of calculations and experiments matched very well except the raise of the
experimental FLC potentially caused by time recovery e€ect in the case of large
prestraining of 0.18.
In spite of some errors in the FLD0 prediction in di€erent directions, the forming
limit in the range of biaxial stretching following the prestraining have been accurately
predicted. Similar predictions using the K±B yield criterion for Al6111-T4 are shown
in Appendix C.

6. Conclusion

Brief reviews are given with respect to the development of the Marciniak
and Kuczynski analysis and the general anisotropic yield criterion developed
by Kara®llis and Boyce (1993). Using the combination of M±K analysis and K±B
yield criterion, accurate predictions of the FLD can be obtained. In addition, this
work has provided a reliable numerical tool, which can not only provide the
conventional FLDs following the linear strain paths but also predict localization
failure under nonlinear strain paths. The calculated FLDs exhibit good agreement
with the experimental results (Graf and Hosford, 1993b) on Al2008-T4 (Figs. 4±8).
The small discrepancy is mainly due to the surface anisotropy in the experiments,
pre-existence and development of the surface ridging, which is not considered in the
calculation. The high stress exponents in the K±B yield criterion have the e€ect of
lowering the forming limit prediction to the desired level and therefore improve the
accuracy of the prediction of the FLD.
A new method of using the experimental FLD of linear strain paths as a
calibration tool to determine the yield locus is proposed. It has been demonstrated
that certain combinations of the mixing factor c and stress exponent m may provide
same predictions of R values and yield stresses in 0, 45 and 90 to the rolling direc-
tion (Fig. 3b) but result in di€erent FLDs (Fig. 4). Because of this sensitivity, using
the conventional FLD to determine material constants in the K±B yield criterion
will have a better description of material deformation in 3-D deformation. This
method may be generalized to determine material constants in the other yield
functions.
With the accurate yield surface description of the anisotropic material, the
proposed combination of M±K analysis and K±B yield criterion is proved to be a
powerful tool in predicting the tearing failure caused by linear and nonlinear load-
ing. Those conclusions are furthered illustrated by the satisfactory predictions using
the proposed approach for Al6111-T4 sheet as shown in Appendix C.

Acknowledgements

The authors would like to thank Dr. Alexander Graf for providing the material
tensile test data of Al6111-T4 and J.C. wishes to acknowledge the support from
NSF grant No. DMI-9703249.
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1121

Appendix A. The general ¯ow chart for the determination of constants in K±B
yield criterion
1122 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

Appendix B. Flow chart for the M±K analysis with rotations of groove

(continued on next page)


J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1123
1124 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

Appendix C. FLDs of Al6111-T4

To further demonstrate the e€ectiveness of our method to determine the yield


surface and the accuracy of K±B yield criterion, FLDs with linear and nonlinear

Fig. C1. Comparisons of the experimental results of Hosford to FLD calculated by using K±B yield cri-
terion with c=0.5 and m=16 for Al6111-T4 [experimental data was published by Graf and Hosford (1994)].

Table C1
Tensile properties of Al6111-T4

Angular o€set from rolling direction Yield stress (MPa) R values

0 185 0.665
45 169 0.63
90 165 0.785
Young's modulus 70.0 (GPa)
Poisson's ratio 0.33
Strain hardening  ˆ 560:66…0:012 ‡ p †0:2588 (MPa)

Table C2
Constants used in the K±B theory for Al6111-T4

Variable Value Variable Value Variable Value

L1111 0.680 L2222 0.663 L3333 0.702


L1122 ÿ0.320 L2233 ÿ0.342 L1133 ÿ0.360
L1212 0.4845 c 0.5 m 16
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1125

Fig. C2. Calculated yield locus of Al6111-T4 using K±B yield criterion.

Fig. C3. Comparisons of the experimental and numerical FLDs of Al6111-T4 prestrained in equal biaxial
tensions of 0.03 and 0.085 and tested perpendicular to the rolling direction [experimental data was pub-
lished by Graf and Hosford (1994)].
1126 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

Fig. C4. Comparisons of the experimental and numerical FLDs of Al6111-T4 prestrained in plane strain
tensions of 0.05 and 0.11 perpendicular to the rolling direction and tested perpendicular to the rolling
direction [experimental data was published by Graf and Hosford (1994)].

Fig. C5. Comparisons of the experimental and numerical FLDs of Al6111-T4 prestrained in plane strain
tensions of 0.05 and 0.12 parallel to the rolling direction and tested perpendicular to the rolling direction
[experimental data was published by Graf and Hosford (1994)].
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1127

Fig. C6. Comparisons of the experimental and numerical FLDs of Al6111-T4 prestrained in uniaxial
tensions of 0.05, 0.095 and 0.14 perpendicular to the rolling direction and tested perpendicular to the
rolling direction experimental data was published by Graf and Hosford [1994].

Fig. C7. Comparisons of the experimental and numerical FLDs of Al6111-T4 prestrained in uniaxial
tensions of 0.05, 0.095 and 0.14 parallel to the rolling direction and tested perpendicular to the rolling
direction [experimental data was published by Graf and Hosford (1994).

strain paths of aluminum Al6111-T4 are calculated here. By choosing a particular


combination of constants c and m in the yield function [Eq. (8)], the corresponding
FLC of linear strain paths can be calculated following the procedures listed in
Appendix B. As proposed in Section 4, the best combination was determined by
matching numerical and experimental FLCs of linear strain paths for the case when
the major principal strain direction coincides with the transverse direction of the
material. Fig. C1 shows the comparison of the calculated FLD on both the left and
right hand sides to the experimental one at the combination of c=0.5 and m=16.
1128 J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129

Good agreement is obtained. In this exercise, the tensile properties of Al6111-T4 are
obtained from the experimental data tested by Graf and listed in Table C1. The
imperfection factor used is 0.9872 for matching FLD0 in the transverse direction and
the non-zero components of the linear operator L are given in Table C2. The
resulting yield locus for Al6111-T4 in the plane of 1ÿ2 is shown in Fig. C2.
The FLDs for various levels of prestraining in uniaxial, biaxial and plane strain
tension are then calculated by using M±K analysis with the rotation of groove and
the K±B yield criterion with the obtained constants in Table C2. The results are
compared with the experimental results of Graf and Hosford (1994) and are illu-
strated in Figs. C3±C7. In spite of some discrepancies in the FLD0 prediction in
di€erent directions, which mainly due to the development of the surface ridging after
the prestraining and the time recovery e€ect as discussed in Sections 4 and 5, the
computed FLDs are generally in good agreement with the experimental results. In
Fig. C5 the FLC predicted by using K±B yield function is otherwise more close to
the experimental results should the FLC not be raised by the time recovery e€ect,
which is evident for the case of prestraining of 0.12 and the second stage straining is
plane strain. Notice that the predicted FLC slopes from K±B yield criterion are very
close to those of the experimental curves. Once again, we have demonstrated our
claims.

References
Barata Da Rocha, A., 1989. Theoretical forming limit diagrams of anisotropic sheets. In: Wagoner, R.H.,
Chan, K.S., Keeler, S.P. (Eds.), Forming Limit Diagrams: Concepts, Methods, and Applications. TMS
Press.
Barata Da Rocha, A., Barlat, F., Jalinier, J.M., 1984. Prediction of forming limit diagrams of anisotropic
sheets in linear and nonlinear loading. Mat. Sci. Eng. 68, 151.
Barlat, F., 1987. Crystallographic texture, anisotropic yield surfaces and limits of sheet metals. Mat. Sci.
Eng. 91, 55.
Barlat, F., 1989. Forming limit diagrams Ð predictions based on some microstructural aspects of mate-
rials. In: Wagoner, R.H., Chan, K.S., Keeler, S.P. (Eds.), Forming Limit Diagrams: Concepts, Meth-
ods, and Applications. TMS press.
Barlat, F., Lian, J., 1989. Plastic behavior and stretch ability of sheet metals. Part I: a yield function for
orthotropic sheet under plane stress conditions. Int. J. Plasticity 5, 51.
Barlat, F., Lege, D.J., Berm, J.C., 1991. A six component yield function for anisotropic materials. Int. J.
Plasticity 7, 693.
Barlat, F., Becker, R.C., Hayashida, Y., Maeda, Y., Yanagawa, M., Chung, K., Brem, J.C., Lege, D.J.,
Matsui, K., Murtha, S.J., Hattori, S., 1997a. Yielding description for solution strengthened aluminum
alloys. Int. J. Plasticity 13, 385.
Barlat, F., Maeda, Y., Chung, K. et al., 1997b. Yield function development for aluminum alloys sheets. J.
Mech. Phys. Solids 11/12, 1727.
Barlat, F., Brem, J.C., Lege, D.J., Chung, K., 1997c. Characterization of the formability for aluminum
alloy and steel sheets. In: Predelaenu, M., Gilormini, P. (Eds.), Advanced Methods in Material Pro-
cessing Defects, Elsevier, p. 265.
Chan, K.S., 1989. Marciniak±Kuczynski approach to calculating forming limit diagrams. In: Wagoner,
R.H., Chan, K.S., Keeler, S.P. (Eds.), Forming Limit Diagrams: Concepts, Methods, and Applications.
TMS press.
Chan, K.S., Koss, D.A., Ghosh, A.K., 1984. Localized necking of sheet at negative minor strains. Metall.
Trans. A 15, 323.
J. Cao et al. / International Journal of Plasticity 16 (2000) 1105±1129 1129

Ferron, G., Makkouk, R., Morreale, J., 1994. A parametric description of orthotropic plasticity in metal
sheets. Int. J. Plasticity 10, 431.
Goodwin, G.M., 1968. Application of strain analysis to sheet metal forming problems in the press shop.
SAE paper no. 680093.
Graf, A., Hosford, W.F., 1993a. Calculations of forming limit diagrams for changing strain paths. Metall.
Trans. A 24, 2497.
Graf, A., Hosford, W.F., 1993b. E€ect of changing strain paths on forming limit diagrams of Al 2008-T4.
Metall. Trans. A 24, 2503.
Graf, A., Hosford, W.F., 1994. The in¯uence of strain path changes on forming limit diagrams of Al
6111-T4. Int. J. Mech. Sci. 36, 897.
Hill, R., 1948. A theory of the yielding and plastic ¯ow of anisotropic metals. Proc. R. Soc. London A193,
281.
Hill, R., 1952. On discontinuous plastic states, with special reference to localized necking in thin sheets. J.
Mech. Phys. Solids 1, 19.
Hill, R., 1979. Theoretical plasticity of textured aggregates. Math. Proc. Cambridge Philos. Soc. 85, 179.
Hiwatashi, S., Bael, A.V., Houtte, P.V., Teodosiu, C., 1998. Prediction of forming limit strains under
strain path changes: application of an anisotropic model based on texture and dislocation structure. Int.
J. Plasticity 14, 647.
Hosford, W.F., 1972. A generalized isotropic yield criterion. J. Appl. Mech. 39, 607.
Hutchinson, J.W., Neale, K.W., 1978a. Sheet necking Ð II: time-independent behavior. In: Koistinen,
D.P., Wang, N.M. (Eds.), Mechanism of Sheet Metal Forming. Plenum Press, New York, p. 127.
Hutchinson, J.W., Neale, K.W., 1978b. Sheet necking Ð III: strain-rate e€ects. In: Koistinen, D.P.,
Wang, N.M. (Eds.), Mechanism of Sheet Metal Forming. Plenum Press, New York, p. 269.
Kara®llis, A.P., Boyce, M.C., 1993. A general anisotropic yield criterion using bounds and transformation
weighting tensor. J. Mech. Phys. Solids 41, 1859.
Keeler, S.P., 1965. Determination of forming limits in automotive stampings. SAE paper no. 650535.
Keeler, S.P., 1968. Circular grid system Ð a valuable aid for evaluating sheet metal formability. SAE
paper no. 680092.
Keeler, S.P., Backofen, W.A., 1964. Plastic instability and fracture in sheets stretched over rigid punches.
ASM Trans. Quart 56, 25.
Lian, J., Baudelet, B., 1987. Forming limit diagram of sheet metal in the negative minor strain region.
Mat. Sci. Eng. 86, 137.
Lode, W., 1926. Versuche ueber den ein¯uss der mittleren hauptspannung auf das ¯iessen der metalle
eisen kupfer und nickle. Z. Physik. 36, 913.
Marciniak, Z., Kuczynski, K., 1967. Limit strains in the processes of stretch-forming sheet metal. Int. J.
Mech. Sci. 9, 609.
McCarron, T.J., Kain, K.E., Hahn, G.T., Flanagan, W.F., 1988. E€ect of geometrical defects in forming
sheet steel by biaxial stretching. Metall. Trans. A 19A, 2067.
Mesrar, R., Fromentin, S., Makkouk, R., Martiny, M., Ferron, G., 1998. Limits to the ductility of metal
sheets subjected to complex strain paths. Int. J. Plasticity 14, 391.
Narasimhan, K., Wagoner, R.H., 1991. Finite element modeling simulation of in plane forming limit
diagrams of sheets containing ®nite defects. Metall. Trans. A 22, 2655.
Sowerby, R., Duncan, D.L., 1971. Failure in sheet metal in biaxial tension. Int. J. Mech. Sci. 13, 137.
StoÈren, S., Rice, J.R., 1975. Localized necking in thin sheets. J. Mech. Phys. Solids 23, 421.

You might also like