12
Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts * , L.E. Govaert, F.P.T. Baaijens Materials Technology, Eindhoven University of Technology, PO Box 513, 5600 MB Eindhoven, The Netherlands Received 1 October 1999; received in revised form 7 November 2000; accepted 27 November 2000 Abstract This study is focussed on the evaluation of ductile fracture methodologies, which are needed to predict product shapes in the blanking process. In an earlier publication [Goijaerts et al., J. Manuf. Sci. Eng., Trans. ASME 122 (2000) 476], two approaches were elaborated using local ductile fracture models. The first strategy incorporates the characterisation of a ductile fracture model in a blanking experiment. The second methodology is more favourable for industry. In this approach, instead of a complex and elaborate blanking experiment, a tensile test is used to characterise a newly proposed criterion, which was shown to predict accurately the ductile fracture for different loading conditions. In this paper, finite element simulations and experiments are performed on both tensile testing and blanking to evaluate the validity of both approaches with corresponding criteria for five different metals. In the blanking process, different clearances as well as different cutting radii of the tools are considered. In conclusion, it can be stated that the first approach gives very good results close to, or within the experimental error for all five materials. The second approach, the more favourable one for industry, yields good results that deviate slightly more over the range of metals. # 2001 Published by Elsevier Science B.V. Keywords: Blanking; Ductile fracture; FEM; Metals; Tensile testing 1. Introduction Blanking is a common technique in high volume produc- tion. Since the beginning of this century, researchers have been analysing the blanking process. Blanking experiments on either planar [2,3] or axisymmetric [4–7] configurations have led to empirical guidelines for process variables such as punch and die radius, speed and clearance. Nevertheless, the blanking process is not yet fully understood. Nowadays, it can be observed that product specifications are getting more severe. This may lead to lengthy trial and error procedures in the development of industrial blanking applications, which is based on empirical knowledge. There- fore, a proper model of the blanking process is desired. Because of the constantly changing loading situations in the material, the process is too complex for an analytical approach [8–10]. Instead, the finite element method has been used to simulate the blanking process [11–13]. One major difficulty in the numerical analysis is the accurate description of ductile fracture initiation, which greatly determines the product shape (Fig. 1). The physical background for ductile fracture in metals is known to be the initiation, growth and coalescence of voids [14–16]. Voids can initiate at inclusions, secondary phase particles or at dislocation pile-ups. Brokken [17] showed with numerical simulations that even under large hydrostatic pressures, voids can initiate at inclusions in shear deformation. Growth and coalescence of voids are driven by plastic deformation. Consequently, in the modelling of ductile fracture initiation, the deformation history is required. In this paper, the class of local ductile fracture criteria that incorporate the stress and strain history are applied for the prediction of ductile fracture initiation (a short overview of these criteria is given in [18]). They can be written in a general form as an integral over equivalent plastic strain e p up to fracture of a certain function of the actual stress state reaching a threshold value C: Z e p f s d e p C (1) Journal of Materials Processing Technology 110 (2001) 312–323 * Corresponding author. Address: Philips Lighting B.V., Central Development Lamps, Mechanics, Design and Heat Technology, PO Box 80020, Building EEA-220, 5600 JM Eindhoven, The Netherlands. Tel.: 31-40-27-56368; fax: 31-40-27-56564. E-mail addresses: [email protected], [email protected] (A.M. Goijaerts). 0924-0136/01/$ – see front matter # 2001 Published by Elsevier Science B.V. PII:S0924-0136(00)00892-X

Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

Evaluation of ductile fracture models fordifferent metals in blanking

A.M. Goijaerts*, L.E. Govaert, F.P.T. BaaijensMaterials Technology, Eindhoven University of Technology, PO Box 513, 5600 MB Eindhoven, The Netherlands

Received 1 October 1999; received in revised form 7 November 2000; accepted 27 November 2000

Abstract

This study is focussed on the evaluation of ductile fracture methodologies, which are needed to predict product shapes in the blanking

process. In an earlier publication [Goijaerts et al., J. Manuf. Sci. Eng., Trans. ASME 122 (2000) 476], two approaches were elaborated

using local ductile fracture models. The ®rst strategy incorporates the characterisation of a ductile fracture model in a blanking experiment.

The second methodology is more favourable for industry. In this approach, instead of a complex and elaborate blanking experiment, a

tensile test is used to characterise a newly proposed criterion, which was shown to predict accurately the ductile fracture for different

loading conditions. In this paper, ®nite element simulations and experiments are performed on both tensile testing and blanking to evaluate

the validity of both approaches with corresponding criteria for ®ve different metals. In the blanking process, different clearances as well as

different cutting radii of the tools are considered. In conclusion, it can be stated that the ®rst approach gives very good results close to, or

within the experimental error for all ®ve materials. The second approach, the more favourable one for industry, yields good results that

deviate slightly more over the range of metals. # 2001 Published by Elsevier Science B.V.

Keywords: Blanking; Ductile fracture; FEM; Metals; Tensile testing

1. Introduction

Blanking is a common technique in high volume produc-

tion. Since the beginning of this century, researchers have

been analysing the blanking process. Blanking experiments

on either planar [2,3] or axisymmetric [4±7] con®gurations

have led to empirical guidelines for process variables such as

punch and die radius, speed and clearance. Nevertheless, the

blanking process is not yet fully understood.

Nowadays, it can be observed that product speci®cations

are getting more severe. This may lead to lengthy trial and

error procedures in the development of industrial blanking

applications, which is based on empirical knowledge. There-

fore, a proper model of the blanking process is desired.

Because of the constantly changing loading situations in the

material, the process is too complex for an analytical

approach [8±10]. Instead, the ®nite element method has

been used to simulate the blanking process [11±13]. One

major dif®culty in the numerical analysis is the accurate

description of ductile fracture initiation, which greatly

determines the product shape (Fig. 1).

The physical background for ductile fracture in metals

is known to be the initiation, growth and coalescence of

voids [14±16]. Voids can initiate at inclusions, secondary

phase particles or at dislocation pile-ups. Brokken [17]

showed with numerical simulations that even under large

hydrostatic pressures, voids can initiate at inclusions in shear

deformation. Growth and coalescence of voids are driven by

plastic deformation. Consequently, in the modelling of

ductile fracture initiation, the deformation history is

required.

In this paper, the class of local ductile fracture criteria that

incorporate the stress and strain history are applied for the

prediction of ductile fracture initiation (a short overview of

these criteria is given in [18]). They can be written in a

general form as an integral over equivalent plastic strain ��ep�up to fracture of a certain function of the actual stress state

reaching a threshold value C:

Z�ep

f �s� d�ep � C (1)

Journal of Materials Processing Technology 110 (2001) 312±323

* Corresponding author. Address: Philips Lighting B.V., Central

Development Lamps, Mechanics, Design and Heat Technology, PO Box

80020, Building EEA-220, 5600 JM Eindhoven, The Netherlands.

Tel.: �31-40-27-56368; fax: �31-40-27-56564.

E-mail addresses: [email protected], [email protected]

(A.M. Goijaerts).

0924-0136/01/$ ± see front matter # 2001 Published by Elsevier Science B.V.

PII: S 0 9 2 4 - 0 1 3 6 ( 0 0 ) 0 0 8 9 2 - X

Page 2: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

where f(s) can be interpreted as a function of the invariants

of the Cauchy stress tensor s: f(J1, J2, J3). If the integral

on the left-hand side reaches the critical value C during

the process, ductile fracture is supposed to initiate. In the

blanking process, this initiation determines the height of

the shear zone (Fig. 1) and thus the shape of the blanked

edge. In the formulations for the different criteria, some

parameters that in¯uence ductile fracture are expected to

appear: plastic strain and triaxiality (triaxiality is de®ned as

hydrostatic stress divided by equivalent Von Mises stress:

sh=�s). It is well known that hydrostatic pressure postpones

ductile fracture initiation which is generally rationalised by

the effect of this stress state on void initiation and

void growth [14±17]. Therefore, triaxiality is often repre-

sented in f(s). Large plastic strains permit voids to

grow and coalesce. This justi®es the integration over plastic

strain.

In the formulation of Eq. (1), C is generally regarded as a

material constant. One has to perform an experiment to

determine C and then it should be possible to use the

characterised criterion in any desired application, the criter-

ion ought to be valid for both the characterisation experi-

ment and the application. Unfortunately, there were no

examples found in the literature where the critical value

C is determined under a loading condition other than that of

the actual application. In other words, existing ductile

fracture criteria are only successful when they are both

characterised and applied under similar loading conditions.

This suggests that some information of the loading path is

represented by the parameter C.

The FEM model of the blanking process that is used in

this study [13] (incorporating a material model with iso-

tropic hardening) is extensively veri®ed on both deformation

®elds and process forces [19,20]. In a previous publication

[1], using the aforementioned FEM model, two approaches

were elaborated to predict fracture initiation in the blanking

process, using speci®c formulations of Eq. (1), for a ferritic

stainless steel. In one approach, C is determined in the

blanking process. No existing ductile fracture criteria were

found to yield satisfactory results. However, very good

results were achieved within the experimental error over

a wide range of clearances [1] when the Rice and Tracey [21]

criterion was adapted. The second approach involved char-

acterisation of C in a tensile test, which represents in a

methodology that is more favourable for industrial applica-

tions. For this approach, a new criterion was proposed that

showed good results within 6% of the blanking experiments.

Moreover, this new model could also predict ductile fracture

initiation for tensile tests under different superposed hydro-

static pressures (different triaxialities). This is important,

because it shows that the criterion is valid for different

loading conditions.

The two aforementioned approaches showed good results

for a ferritic stainless steel. The remaining question was [1]

whether the utilised criteria contain more material-depen-

dent parameters than only the critical C. Therefore, in this

paper, these two approaches were evaluated for four differ-

ent materials using exactly the same methodology, including

the material characterisation.

In Section 2, we explain the methodologies to characterise

ductile fracture initiation in both blanking and tensile test-

ing. In Section 3, we discuss the experimental and numerical

methods needed. In Section 4, the results are presented for

all ®ve materials. Finally, we discuss the results and con-

clude in Section 5.

2. Methodologies to characterise ductile fracture

In the following, the experimental determination of the

critical parameter C is referred as the characterisation of

local ductile fracture criteria with the general formulation of

Eq. (1). Two approaches will be utilised (as in [1]): deter-

mination of C in a blanking experiment and characterisation

of the criterion in a tensile test. These methodologies are

explained in Sections 2.1 and 2.2.

Fig. 1. The blanking process with an indication of the different zones determining the product shape.

A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323 313

Page 3: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

2.1. Characterisation in blanking

We consider ductile fracture initiation criteria in the

formulation of Eq. (1). The right-hand side of this formula-

tion is assumed to be a material constant. The characterisa-

tion of a ductile fracture model in the blanking process can

be performed with the four following steps:

1. For one clearance, the punch displacement at ductile

fracture initiation is experimentally determined.

2. That particular blanking process is simulated with a

validated FEM model of the blanking process [1] and

the left-hand side of Eq. (1) is updated and stored as a

®eld variable.

3. When the experimental punch displacement at fracture

is reached in the simulation, C is determined to be the

maximum value ofR

f �s� d�ep over the entire FEM mesh.

4. At the end of the simulation (the moment of fracture

initiation), we check if the location of the maximum

value ofR

f �s� d�ep is in agreement with the experi-

mental location of ductile fracture initiation.

If this ®nal check is satis®ed, we declare the ductile

fracture initiation criterion to be characterised.

If a ductile fracture initiation model is characterised, we

can evaluate the validity of it for the blanking process over

the entire range of clearances. This evaluation is performed

using FEM simulations of the blanking process for the other

clearances. During the simulationsR

f �s� d�ep is again

updated and stored as a ®eld variable and as soon as this

®eld variable reaches the critical C, the punch displacement

at fracture is predicted. Goijaerts et al. [1] showed that the

adapted Rice and Tracey criterion of Eq. (2) performed very

well, within the experimental error, for X30Cr13, using this

methodology. The adaptation consists of the change of a

constant a � 32

(theoretically obtained considering one void

in an in®nite matrix [21]) into a � 2:9:Z�ep

exp 2:9sh

�s

� �d�ep � CRT (2)

2.2. Characterisation in tensile testing

For industrial applications it would be a great advantage if

a fracture criterion could be characterised by performing a

more common and easy test, instead of a complicated and

dif®cult, well-conditioned, blanking experiment. The char-

acterisation of a ductile fracture model in tensile testing can

be explained with the steps presented below:

1. A tensile test is performed and the minimum thickness

of the neck after fracture is measured. This minimum

thickness of the neck determines the moment of ductile

fracture initiation.

2. Then, the tensile test is simulated, using an FEM model

and the left-hand side of Eq. (1) is updated and stored as

a ®eld variable.

3. When the experimentally determined minimum thick-

ness of the neck is reached in the simulation, C is

determined to be the maximum value ofR

f �s� d�ep over

the entire FEM mesh.

4. At the end of the simulation (the moment of fracture

initiation), we check if the location of the maximum

value ofR

f �s� d�ep is in agreement with the experi-

mental location of ductile fracture initiation, which is

the overall centre of the specimen.

If this ®nal check is satis®ed, we declare the ductile

fracture initiation criterion to be characterised.

If a ductile fracture initiation model is characterised, we

can evaluate the validity of it for the blanking process in the

same way as was explained in Section 2.1 over the range of

clearances. Goijaerts et al. [1] showed that for X30Cr13 no

existing criterion could predict ductile fracture initiation in

the blanking process satisfying well, following this strategy.

Moreover, the application of the criterion de®ned in Eq. (2)

exhibited deviations in the order of 30%, which was

regarded not to be acceptable. Therefore, a new criterion

was proposed:Z�ep

1� 3:9sh

�s�e0:63

p

� �d�ep � CG (3)

where the brackets b� � �c, de®ned as

bxc � x; x > 0;0; x � 0

�(4)

are used to ensure that the criterion cannot decrease.

This newly posed criterion performed well in the blanking

process with maximum deviations of only 7% for the

predicted punch displacement at fracture over a range of

clearances for X30Cr13. Moreover, it was capable of pre-

dicting ductile fracture initiation for tensile tests under

superposed hydrostatic pressures up to 500 MPa. This indi-

cates that the criterion of Eq. (3) captures the in¯uence of

triaxiality on ductile fracture initiation correctly, at least for

X30Cr13.

3. Methods

In this section all methods, needed for the methodologies

of Section 2, are explained. In Section 3.1, ®ve materials

were presented for which the characterisations were per-

formed. For one of these materials, X30Cr13, the strategies

of Section 2 were elaborated in an earlier publication [1].

The experimental results for X30Cr13, presented in that

publication, are used to explain the experimental techniques

in Section 3.2. In Section 3.3, the FEM models of the

blanking process and the tensile test are presented.

3.1. Materials

The two aforementioned methodologies were elaborated

for the blanking process to characterise ductile fracture

314 A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323

Page 4: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

initiation for X30Cr13 [1]. To verify these strategies, exactly

the same procedures were applied to four other metals which

are presented, along with X30Cr13 in Table 1. X30Cr13 is a

ferritic stainless steel and 316Ti is an austenitic stainless

steel. Ms64 is a brass that is often used in cutting and

forming processes. DC04 is a high quality steel with a large

formability that is applicable in deepdrawing processes.

Al51ST is employed often in cutting and has a low form-

ability (is relatively brittle).

3.2. Experimental

In order to obtain a satisfying material description, also

for large plastic strains, a material characterisation techni-

que is used, which is brie¯y explained in Section 3.2.1 with

the help of the results for the aforementioned ferritic stain-

less steel (X30Cr13). To characterise and verify ductile

fracture initiation criteria of the form of Eq. (1), experiments

were needed. As was done previously [1] an axisymmetric

blanking set-up is used with different geometries (Section

3.2.2). We chose to vary the clearance because the effect on

the product shape of a change in clearance is known to be

large [4,5,7].

3.2.1. Material characterisation

As was done earlier for X30Cr13, all materials considered

were assumed to be plastically deformed according to the

Von Mises yield condition with isotropic hardening. In

formulating this plastic deformation, the yield stress

increases with increasing equivalent plastic strain. The

relationship between the yield stress and the equivalent

plastic strain is dif®cult to be obtained experimentally for

large strains, using conventional tests such as tensile or shear

experiments. This relationship is determined by performing

approximately 10 tensile tests with each tensile specimen

being subjected to a different amount of rolling to obtain

different initial plastic deformations. The assumption of

isotropic hardening allows the addition of the rolling and

tensile equivalent plastic strains. We quanti®ed the relation-

ship between the yield stress and the equivalent plastic

strain, by ®tting a mastercurve through the maxima of the

stress±strain curves of these tensile tests. For X30Cr13, the

results are presented in Fig. 2. At the left-hand side, the true

stress±true strain results of 17 tensile tests are presented with

their maxima (triangles), together with a number of points

(circles) of the tensile test of as-received material. At the

right, a mastercurve is ®tted through the markers. In this

®tting procedure, the circles are essential to correctly cap-

ture the hardening behaviour for small strains. For this ®tting

procedure, the power law of the Nadai [22] hardening model

does not suf®ce because it only incorporates two constants to

be determined and cannot be accurately ®tted along the

markers. The hardening model of the Voce type [23]

involves three material parameters but grows to a constant

yield stress level for large strains. Therefore, this model is

phenomenologically expanded with a square root and a

linear term in strain so that different hardening behaviour

for large strains can be captured:

sy � sy;0 �M1�1ÿ eÿ�ep=M2� �M3

�����ep

p �M4�ep (5)

where sy is the Von Mises yield stress, sy,0 the initial yield

stress, M1±M4 are material constants and

�ep �������������������2

3�ep : ep

qthe equivalent plastic strain, with ep:ep the double inner

product of the logarithmic plastic strain tensor with itself.

Table 1

Five different metals that are employed in the total veri®cation procedurea

Name used Material code Norm table

X30Cr13 X30Cr13 DIN 17006

316Ti X10CrNiMoTi 18 10 DIN 17440

Ms64 CuZn36 DIN 17660

DC04 FeP 04 Euronorm 32±66

Al51ST Al MgSi 1 Mg DIN 1746/1747

a The names used in this paper are presented along with the material

codes and their corresponding norm tables.

Fig. 2. Fitting procedure for the hardening curve of X30Cr13 [1]. Left: the separate tensile curves are presented and their maxima (triangles), together with a

number of points on the as-received tensile curve (circles). Right: the mastercurve is ®tted through all markers.

A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323 315

Page 5: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

3.2.2. Blanking experiments

An axisymmetric blanking set-up with a die-hole dia-

meter of 10 mm, including a blankholder with constant

pressure, was built [1] with ®ve different punches (dia-

meters: 9.98, 9.94, 9.88, 9.80 and 9.70 mm) resulting in

®ve different clearances, covering the industrially used

range of clearances (1, 3, 6, 10 and 15% of the sheet

thickness of 1 mm). To avoid exorbitant simulation times,

the cutting radii of the punches and die are enlarged to

approximately 0.1 mm. The punch radius is somewhat

smaller (�0.08 mm) and the die radius is a bit larger

(�0.14 mm), to make sure that fracture will initiate at the

punch and grow to the die radius. We want to determine the

punch displacements at fracture experimentally, to have

reference points in the numerical simulations for the initia-

tion of ductile fracture. In our blanking set-up, six experi-

ments were performed for every clearance. The shear zone

(b) and the burr (c) are measured afterwards at eight posi-

tions over the circumference of the blanked products, and

averaged to justify the misalignment of the punch. Then, the

values are averaged over the six experiments and the stan-

dard deviation is calculated (as an example, results are

presented for X30Cr13 at the left-hand side of Fig. 3).

It was shown that the roll-over could be accurately

predicted for X30Cr13 with a validated model [19,24].

Because it is dif®cult to measure the roll-over in the

axisymmetric set-up, due to the spring-back of the specimen,

the roll-over (a) is taken from the numerical simulations.

The element size near the transition of roll-over and shear

zone is taken as the corresponding standard deviation.

Although, the mentioned numerical model is not validated

for the other materials, it is assumed correctly to predict the

roll-over for those materials as well.

The roll-over height is very low for small clearances and

becomes larger for wider clearances because the broader

deformation zone allows more bending. The shear zone is

getting smaller for larger clearances and this is caused by the

hydrostatic stress state; for small clearances, the hydrostatic

pressure is larger and this postpones ductile fracture initia-

tion, despite of the fact that the deformation is more

localised and that the strains are larger. The burr height is

very small (in the order of 5 mm) and is largely determined

by the punch radius. The average punch displacements at

fracture a� b� c for X30Cr13 are plotted at the right-hand

side of Fig. 3, along with twice the standard deviations (95%

interval). The combination of the trend for roll-over height

and shear height (plus burr) explains the minimum in the

curve. There is a small experimental deviation for the

clearance of 10%. This is a result of the larger punch radius

for the speci®c 10% clearance punch. A larger punch radius

postpones ductile fracture initiation because the deformation

becomes less localised.

3.3. Numerical

In Section 3.3.1, the FEM model of the blanking process,

validated for X30Cr13 [19,24], is brie¯y discussed. In

Section 3.3.2, the three-dimensional FEM model of a tensile

test is presented [1].

3.3.1. Blanking

We simulated the blanking process using a two-dimen-

sional, axisymmetric FEM model, described by Brokken

et al. [25]. Quasi-static analyses were performed on the

model geometries that match the experimental set-up for the

®ve different clearances. We modelled the specimen with an

isotropic elasto-plastic material, using the material proper-

ties that are determined in Section 4.1. The plastic material

behaviour is described by the Von Mises yield condition,

with isotropic hardening and by the Prandtl±Reuss repre-

sentation of the ¯ow rule [26]. The mesh, used for the 15%

clearance, is shown in Fig. 4. The left boundary at the top

(specimen centre) is the axis of symmetry. The other

boundaries are either free surfaces or in interaction with a

contacting body (punch, die or blankholder).

Linear quadrilateral elements are used that become smal-

ler as they approach either the die radius or the punch radius.

Near those radii (�0.1 mm) the element proportions need to

be in the range 5 mm [17], resulting in up to 3000 elements in

the entire mesh. This element size is not necessary to predict

the process force correctly, but it will be vital accurately and

smoothly to describe the ®eld variables, needed to predict

Fig. 3. Experimental results for ductile fracture initiation for varying clearance for X30Cr13 [1].

316 A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323

Page 6: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

ductile fracture initiation. The punch moves down and

penetrates the specimen, resulting in constantly changing

boundary conditions. To deal with these dif®cult boundary

conditions and the localised large deformations, the FEM

application that we used, combines three numerical proce-

dures: the commercial implicit package MARC [26] (using

an updated Lagrange formulation), an Arbitrary Lagrange±

Euler approach [27,28] and an automatic adaptive remesh-

ing algorithm [17,25], to overcome severe mesh distortion

problems. This model was experimentally validated prior to

fracture on both deformation ®elds Ð using digital image

correlation Ð and process forces, using a planar blanking

set-up [19,20]. Therefore, the deformation history in the

blanking process can be calculated adequately, which is a

prerequisite for the local modelling of ductile fracture.

3.3.2. 3D FEM model of a tensile test

For the determination of C in a tensile test (the second

approach), specimens were used with sizes that are pre-

sented at the right-hand side of Fig. 5. In previous work a 3D

FEM model was presented [1]. The undeformed specimen is

shown at the left, of which 18

is modelled (shown in the centre

of Fig. 5). The three-dimensional calculation is needed to

correctly simulate the necking process, for which no imper-

fection was needed. At the location where the neck is

expected to occur the mesh is re®ned. Exactly, the same

material model will be used as in the blanking model of

Section 3.3.1.

4. Results for different materials

Two different approaches were elaborated for the blank-

ing process to characterise ductile fracture initiation for

X30Cr13 [1]. To verify these methodologies, exactly the

same procedures are applied to the four other materials of

Table 1.

In Section 4.1, the material parameters concerning elasto-

plastic deformation are determined assuming the same

constitutive model to be valid for all materials. Results

for tensile testing, in which the parameter C is determined

for the second approach, are presented in Section 4.2. For the

blanking process, the resulting maximum blanking forces

and punch displacements at fracture initiation are outlined in

Section 4.3. For two materials some special applications of

both criteria are outlined in Section 4.4.

4.1. Material characterisation

All materials are assumed to behave according to the

elasto-plastic model with Von Mises plasticity and isotropic

hardening. The elastic properties (Young's modulus E and

Poisson's ratio n) are taken from the material supplier. For

the determination of the plastic material parameters, the

procedure of Section 3.2.1 is applied, where a number of

pre-rolled specimens are tested in rolling direction to give

yield stress information for large strains. The results are

plotted in Fig. 2 for X30Cr13 and in Fig. 6 for the other

materials. The triangles represent the maxima of the separate

tensile tests (the maximum stress occurring in homogeneous

deformation, obtained just before necking). The circles are

extra points, taken from the tensile test of the as-received

material, which are needed to supply information for the

yield stress in the region of small strains. Through all

markers a mastercurve is ®tted using the formulation of

Eq. (5). The parameters resulting from the ®tting procedures

are all reported in Table 2 together with the elastic constants.

As can be seen in Table 2 some parameters are ®tted to be

zero. This is due to the fact that in the ®tting procedure,

employing Eq. (5), the parameters M3 and M4 are not

Fig. 4. Example of an FEM mesh for the 15% clearance experiment at the

punch displacement, where experimentally ductile fracture initiates for

X30Cr13.

Fig. 5. The undeformed tensile specimen together with the 3D FEM model of 18

of the specimen. The sizes are presented on the right (thickness is 1 mm).

A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323 317

Page 7: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

allowed to become negative. If one of these parameters is

®tted to be negative, it is made zero and the ®tting procedure

is repeated, keeping the speci®c parameter zero. M3 and M4

are not allowed to become negative because an extrapolation

of the mastercurve with negative parameters M3 or M4 may

introduce intrinsic softening for large strains which is not

realistic for metals.

As shown in Figs. 2 and 6 the ®tted mastercurves describe

the yield stress data well, for all ®ve materials.

4.2. Tensile tests

For all ®ve materials, a 3D simulation was performed with

the FEM model of Section 3.3.2 using the material para-

meters of Table 2. The deformed mesh and the fractured

specimen are shown in Fig. 7 for X30Cr13 [1]. This ®gure

shows that the FEM model predicts the deformation of the

tensile specimen well. Also the wedge-like shape of the

specimen at fracture is predicted correctly.

Besides this veri®cation on deformation behaviour for

X30Cr13, the FEM simulation is also checked on the force±

displacement curve. The experimental and numerical force±

displacement curves for all ®ve materials are depicted in

Fig. 8. It can be seen that the shape of the predicted force±

displacement curves deviates only for Al51ST and DC04.

For these two materials only ®ve and six circles, respec-

tively, could be employed in the ®tting procedure (Fig. 6),

because otherwise deviations become too large for large

strains. Because less experimental data for small strains is

used during ®tting, the stress±strain relationship is not ®tted

Fig. 6. Fitting procedures for four different materials. The triangles represent the maxima of the separate tensile curves and the circles denote some points of

the as-received tensile curve. The curves are ®tted on Eq. (5).

Table 2

Material parameters of ®ve different materials for elasto-plastic model with Von Mises plasticity, with hardening according to Eq. (5)

E (MPa) n (ÿ) sy,0 (MPa) M1 (MPa) M2 (MPa) M3 (MPa) M4 (MPa)

X30Cr13 187�103 0.28 420 133 0.0567 406 70.7

316Ti 200�103 0.28 298 1003 0.414 0 0

Ms64 100�103 0.35 251 345 0.320 56.0 0

DC04 210�103 0.29 185 309 0.378 113 0

Al51ST 70�103 0.33 306 61.0 0.0476 73.0 0

318 A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323

Page 8: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

accurately in this strain region for these two materials. The

shape of the curves for the other materials is well predicted.

The only deviation seems to be the clamp displacement,

which might be under-predicted numerically.

4.3. Blanking

Axisymmetric blanking experiments were performed for

®ve different materials with ®ve different clearances using

the experimental set-up described in Section 3.2.2. The

maximum blanking forces are measured, as well as some

measures of the blanked edge, from which the punch dis-

placement at fracture initiation is determined. Using the

material parameters, characterised in Section 4.1, axisym-

metric FEM simulations are performed for all clearances for

the ®ve different materials.

The numerical and experimental maximum blanking

forces are presented in Fig. 9 at the left axes. For

X30Cr13, Al51ST and Ms64, deviations are observed in

the order of 5% or smaller. For 316Ti and DC04, the

predictions deviate approximately 10% from the experi-

ments. The stresses in the shear zone where strains occur

larger than 3, largely determine this maximum blanking

force. An error in the material description for large strains

will consequently result in a corresponding error in the

maximum blanking force. Therefore, the results and devia-

tions for the maximum blanking force provide information

Fig. 7. Simulation of a tensile test for X30Cr13 and experimental veri®cation on deformations taken from [1]. On top, the calculated deformations at fracture

initiation are shown with ®ve levels of the equivalent plastic strain, together with a fractured half. In the centre, the three orthogonal views of the deformed

specimen are shown with a zoomed plot of the re®ned mesh in the neck. At the bottom, the experimental fracture surface is compared with the calculated

cross-sectional area in the neck at fracture initiation (mind the wedge-like shape).

Fig. 8. Experimental and numerical comparison of the tensile force versus

clamp displacement for the tensile tests of all materials.

A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323 319

Page 9: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

about the performance of the material model, including the

assumption of Von Mises plasticity with isotropic hardening

and characterisation for large strains.

Two approaches were employed to predict the punch

displacement at fracture initiation for all materials. The ®rst

approach utilises the adapted criterion of Rice and Tracey

[21] of Eq. (2). CRT is determined in the blanking experiment

with the 6% clearance and with this CRT the punch dis-

placement at ductile fracture initiation is predicted for all

other clearances. Results for the quanti®ed CRT's are given

in Table 3 and the predictions are presented in Fig. 9. For the

approach of Section 2.2, the thickness at fracture initiation in

the tensile tests are measured and the critical CG's (for the

criterion proposed by Goijaerts et al. [1]) are determined in

the 3D FEM simulations of the tensile tests for all ®ve

materials (Table 3). Predictions for the punch displacements

Fig. 9. Comparison of blanking experiments (99% interval) and numerical predictions for ®ve different materials. Results are presented for the maximum

blanking force (left axes), and the prediction of punch displacement at fracture initiation (right axes), considering the adapted Rice and Tracey [21] criterion

(Eq. (2), CRT, characterised in blanking) and the criterion of Eq. (3) (CG, characterised in tensile testing) proposed by Goijaerts et al. [1].

Table 3

For ®ve different metals, the experimentally determined critical C's are

presented for both applied criteriaa

Material CRT (blanking)

(ÿ)

CG (tensile testing)

(ÿ)

Thickness at

fracture (mm)

X30Cr13 2.70 3.53 0.448

316Ti 7.42 6.53 0.329

Ms64 7.27 6.39 0.342

DC04 8.70 14.8 0.132

Al51ST 1.49 1.45 0.642

a For the Rice and Tracey [21] criterion, CRT is quanti®ed in the

blanking experiment with the 6% clearance. For the Goijaerts et al. [1]

criterion, CG is resolved from the presented experimental thickness at

fracture initiation using the 3D FEM simulations of the tensile tests.

320 A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323

Page 10: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

at fracture initiation for all clearances are presented in Fig. 9

for all materials.

For the ®rst approach, the predictions fall within the

experimental errors or in a very close range, for all materials.

The maximum deviation from an experiment is 15% for

Ms64 for the blanking geometry with the 15% clearance. All

other predictions deviate in the range of 10% or smaller. For

the second approach, with characterisation in the tensile

tests, predictions for X30Cr13 and Al51ST maximally

deviate 6 and 10%, respectively, from the experiments.

For 316Ti and DC04, the maximum deviations are 12 and

15% and the maximum deviation for Ms64 is 24%. The ®rst

approach performs better for all materials.

4.4. Special applications

So far in this paper, only relatively large cutting radii of

the tools are considered in order to reduce the calculation

times in the large number of simulations, of which the results

were presented in the previous section. Now, for X30Cr13,

results are displayed for both small (read industrial) and

large cutting radii of the punch for two extreme clearances in

Fig. 10. The ®rst methodology, applying the adapted Rice

and Tracey criterion, and the second methodology, applying

the Goijaerts et al. criterion, show good results for different

geometries: a change in clearance which does not exhibit a

drastic change in punch displacement at fracture (Fig. 9), is

captured by both methodologies; moreover, a change in

cutting radii of the tools, which shows a drastic change in

punch displacement at fracture (Fig. 10), is also accurately

described by both approaches (CRT and CG are taken from

Table 3). These results strongly suggest that both methodol-

ogies with the corresponding criterion can be applied suc-

cessfully in the blanking process over a wide variety of

geometries.

Up to this point, for the ®rst approach Ð characterisation

in blanking Ð only the adapted Rice and Tracey criterion of

Eq. (2) is utilised, and for the second strategy Ð character-

isation in tensile testing Ð only the criterion of Eq. (3) is

used. To check their applicational area, both criteria are

applied in both methodologies for Al51ST. Results are

presented in Fig. 11. As can be seen, the criterion of

Eq. (3) performs well for both methodologies. From

Table 3, one can ®nd that for Al51ST, CG is determined

to be 1.45 in the tensile test. In blanking it is quanti®ed quite

close to CG � 1:29, which shows that the criterion is valid

for both tensile testing and blanking. Unfortunately, the

adapted Rice and Tracey criterion performs badly when

characterised in tensile testing (CRT � 3:18, determined in

tensile testing compared to CRT � 1:49 quanti®ed in blank-

Fig. 10. Experimental and numerical results of the punch displacement at fracture for both small and large cutting radii of the tools. At the left, results are

presented for the 1% clearance and at the right for the 15% clearance.

Fig. 11. Experimental and numerical results of the punch displacement at fracture for both methodologies utilising both criteria for Al51ST. Left: the ®rst

approach: characterisation in blanking. Right: the second strategy: characterisation in tensile testing.

A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323 321

Page 11: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

ing, Table 3). This result proves that this criterion is valid

only for blanking.

5. Discussion and conclusion

Complete strategies to predict ductile fracture initiation in

the blanking process, which resulted in a full blanking

model, were presented for one metal in an earlier publication

[1]. These methodologies, including the assumptions for the

material behaviour, are applied to four other metals. For all

®ve materials, blanking results are presented in this paper

concerning maximum process force and punch displacement

at ductile fracture initiation. The maximum process force,

occurring in blanking, is important for engineering aspects

when wear and duration of life of the tools is considered. The

punch displacement at fracture initiation is important

because it largely determines the shape of the blanked

edge.

Material characterisation is a signi®cant issue in model-

ling of the blanking process. Because very large strains

occur in the localised shear zone it is important to determine

accurately the stress±strain relation for large strains, which

is not straightforward. Tensile tests are performed on pre-

rolled specimens to evolve this relation up to large strains for

all metals. The calculated force±clamp displacement curves

for the tensile tests and the predicted maximum blanking

forces suggest that the plastic material behaviour is char-

acterised successfully over the entire strain range.

Predictions for maximum blanking forces are good for all

metals with maximum deviations of approximately 10%.

For the prediction of punch displacement at ductile fracture

initiation, two methodologies are evaluated. The strategy

where the fracture criterion is characterised in a blanking

experiment performs well for all materials. The second

methodology, which utilises a tensile test to characterise

ductile fracture with a newly proposed criterion [1], yields

satisfactory results. The advantage of the ®rst approach is the

higher accuracy. The second strategy has the advantage of a

much simpler experiment that is needed for the character-

isation of ductile fracture. Moreover, the ductile fracture

initiation criterion of the latter is not only valid for blanking

but also for tensile testing under different pressures.

Although, the criteria may look promising for other forming

processes, it should be noted that the results presented here

do not guaranty the applicability under the corresponding

loading conditions.

For both criteria it is concluded that the only material

parameter is the critical C.

Acknowledgements

This research was funded by the Dutch innovative

research projects (IOP-C.94.702.TU.WB). Jan Post of

Philips DAP/LTM kindly supplied the X30Cr13 material

and the experimental data for the corresponding material

model.

References

[1] A.M. Goijaerts, L.E. Govaert, F.P.T. Baaijens, Prediction of ductile

fracture in metal blanking, J. Manuf. Sci. Eng., Trans. ASME 122

(2000) 476±483.

[2] T.M. Chang, H.W. Swift, Shearing of metal bars, J. Inst. Met. 78

(1950) 119±146.

[3] A.G. Atkins, Surfaces produced by guillotining, Phil. Mag. 43 (1981)

627±641.

[4] T.M. Chang, Shearing of metal blanks, J. Inst. Met. 78 (1951) 393±

414.

[5] W. Johnson, R.A.C. Slater, A survey of the slow and fast blanking of

metals at ambient and high temperatures, in: Proceedings of the

International Conference of Manufacturing Technology, Michigan,

1967, pp. 773±851.

[6] R. Balendra, F.W. Travis, Static and dynamic blanking of steel of

varying hardness, Int. J. Mach. Tool Des. Res. 10 (1970) 249±271.

[7] C.M. Choy, R. Balendra, Experimental analysis of parameters

in¯uencing sheared-edge pro®les, in: Proceedings of the Fourth

International Conference on Sheet Metal, Twente, 1996, pp. 101±

110.

[8] C.F. Noble, P.L.B. Oxley, Crack formation in blanking and piercing,

Int. J. Prod. Res. 2 (1963) 265±274.

[9] S. Fukui, K. Konda, K. Maeda, Smooth shearing by stepped pro®le

tool, Ann. CIRP 20 (1971) 57±58.

[10] Q. Zhou, T. Wierzbicki, A tension zone model of blanking and

tearing of ductile metal plates, Int. J. Mech. Sci. 38 (1996) 303±324.

[11] J. Post, R. Voncken, FEM analysis of the punching process, in:

Proceedings of the Fourth International Conference on Sheet Metal,

Twente, 1996, pp. 159±169.

[12] E. Taupin, J. Breitling, W.-T. Wu, T. Altan, Materials fracture and

burr formation in blanking: results of FEM simulations and

comparison with experiments, J. Mater. Process. Technol. 59

(1996) 68±78.

[13] D. Brokken, W.A.M. Brekelmans, F.P.T. Baaijens, Numerical

modelling of the metal blanking process, J. Mater. Process. Technol.

83 (1998) 192±199.

[14] D. Broek, A study on ductile failure, Ph.D. Thesis, Delft University

of Technology, the Netherlands, 1971.

[15] B. Dodd, Y. Bai, Ductile fracture and ductility Ð with applications to

metalworking, Academic Press, London, 1987.

[16] P.J. Bolt, Prediction of ductile failure, Ph.D. Thesis, Eindhoven

University of Technology, the Netherlands, 1989.

[17] D. Brokken, Numerical modelling of ductile fracture in blanking,

Ph.D. Thesis, Eindhoven University of Technology, the Netherlands,

1999.

[18] S.E. Clift, P. Hartley, C.E.N. Sturgess, G.W. Rowe, Fracture

prediction in plastic deformation processes, Int. J. Mech. Sci. 32

(1990) 1±17.

[19] Y.W. Stegeman, A.M. Goijaerts, D. Brokken, W.A.M. Brekelmans,

L.E. Govaert, F.P.T. Baaijens, An experimental and numerical study

of a planar blanking process, J. Mater. Process. Technol. 87 (1999)

266±276.

[20] A.M. Goijaerts, Y.W. Stegeman, L.E. Govaert, D. Brokken, W.A.M.

Brekelmans, F.P.T. Baaijens, Can a new experimental and numerical

study improve metal blanking? J. Mater. Process. Technol. 103

(2000) 44±50.

[21] J.R. Rice, D.M. Tracey, On the ductile enlargement of voids in

triaxial stress ®elds, J. Mech. Phys. Solids 17 (1969) 201±217.

[22] A. Nadai, Theory of ¯ow and fracture of solids, McGraw-Hill, New

York, 1950.

[23] E. Voce, The relationship between stress and strain, J. Inst. Met. 74

(1948) 537.

322 A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323

Page 12: Evaluation of ductile fracture models for different …Evaluation of ductile fracture models for different metals in blanking A.M. Goijaerts*, L.E. Govaert, F.P.T. Baaijens Materials

[24] A.M. Goijaerts, Prediction of ductile fracture in metal blanking, Ph.D.

Thesis, Eindhoven University of Technology, the Netherlands, 1999.

[25] D. Brokken, A.M. Goijaerts, W.A.M. Brekelmans, C.W.J. Oomens,

F.P.T. Baaijens, Modelling of the blanking process, in: Computational

Plasticity: Fundamentals and Applications, CIMNE, Barcelona,

Spain, Vol. 2, Barcelona, 1997, pp. 1417±1424.

[26] MARC Manuals: Vol. A±F, Version K7, MARC, Palo Alto, CA 1997.

[27] P.J.G. Schreurs, F.E. Veldpaus, W.A.M. Brekelmans, Simulation of

forming processes using the arbitrary Eulerian±Lagrangian formula-

tion, Comput. Meth. Appl. Mech. Eng. 58 (1986) 19±36.

[28] F.P.T. Baaijens, An U-ALE formulation of 3D unsteady viscoelastic

¯ow, Int. J. Numer. Meth. Eng. 36 (1993) 1115±1143.

A.M. Goijaerts et al. / Journal of Materials Processing Technology 110 (2001) 312±323 323