You are on page 1of 25

ARTICLE IN PRESS

International Journal of Mechanical Sciences 47 (2005) 719–743


www.elsevier.com/locate/ijmecsci

Calibration and evaluation of seven fracture models


Tomasz Wierzbicki, Yingbin Bao, Young-Woong Lee, Yuanli Bai
Impact and Crashworthiness Laboratory, Massachusetts Institute of Technology, Room 5-218, Cambridge,
MA 02139, USA

Received 2 July 2004; received in revised form 24 February 2005; accepted 7 March 2005

Abstract

Over the past 5 years, there has been increasing interest of the automotive, aerospace, aluminum, and
steel industries in numerical simulation of the fracture process of typical structural materials. Accordingly,
there is a pressure on the developers of leading commercial codes, such as ABAQUS, LS-DYNA, and
PAM-CRASH to implement reliable fracture criteria into those codes. Even though there are several
options to address fracture in these and other commercial codes, no guidelines are given for the users as to
which fracture criterion is suitable for a particular application and how to calibrate a given material for
fracture. The objective of the present paper is to address the above issues and present a thorough
comparative study of seven fracture criteria that are included in libraries of material models of non-linear
finite element codes. A set of 15 tests recently conducted by the authors on 2024-T351 aluminum alloy is
taken as a reference for the present study. The plane stress prevails in all these tests. These experiments are
compared with the constant equivalent strain criterion, the Xue–Wierzbicki (X–W) fracture criterion, the
Wilkins (W), the Johnson–Cook (J–C) and the CrachFEM fracture models. Additionally, the maximum
shear (MS) stress model, and the fracture forming limit diagram (FFLD) are included in the present
evaluation. All criteria are formulated in the general 3-D case for the power law hardening materials and
then are specified for the plane stress condition. The advantage of working with plane stress is that there is
one-to-one mapping from the stress to the strain space. Therefore, the fracture criteria formulated in the
stress space can be compared with those expressed in the strain space and vice versa. Fracture loci for all
seven cases were constructed in the space of the equivalent fracture strain and the stress triaxiality.
Interesting observations were made regarding the range of applicability and expected errors of some of the
most common fracture criteria. Besides evaluating the applicability of several fracture criteria, a detailed
calibration procedure for each criterion is presented in the present paper. It was found rather unexpectedly

Corresponding author. Tel.: +1 617 253 2104; fax: +1 617 253 8125.
E-mail address: wierz@mit.edu (T. Wierzbicki).

0020-7403/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijmecsci.2005.03.003
ARTICLE IN PRESS

720 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

that the MS stress fracture model closely follows the trend of all tests except the round bar tensile tests. The
X–W criterion and the CrachFEM models predict correctly fracture in all types of experiments. The W
criterion is working well in certain ranges of the stress triaxiality.
r 2005 Elsevier Ltd. All rights reserved.

Keywords: Fracture criterion; Calibration; Plane stress

1. Introduction

The present paper is concerned with a class of problems involving ductile fracture of crack-free
bodies. Usually, material separation is a result of a complex physical process which occurs at the
micromechanical scale. On a macro scale the only variables that control fracture are current
values of components of the stress and strain tensors and their histories. These quantities are
readily available as output in all commercial nonlinear FE codes. Fracture theories that make use
of these macroscopic quantities are criteria of choice for finite element implementation. Several
leading commercial codes are indeed offering to their customers a number of different, ever more
sophisticated fracture options. At the same time no guidance is given for the users as to which
fracture criterion is suitable for a particular application and how to determine fracture parameters
from a minimum number of tests. The present paper aims at giving some insights into these
problems by comparing the relative performance of seven fracture criteria which are included in
the material libraries of several commercial codes. In addition, predictions from these criteria are
confronted with the results of a dozen of different types of fracture tests on 2024-T351 aluminum
alloy recently reported in the literature [1–3]. It is not our intention to present a comprehensive
literature review on fracture nor to overview capabilities of commercial codes in this regard.
Interested readers are referred to excellent review articles by Atkins [4] and McClintock [5].
Instead, we chose seven representative criteria that are relatively easy to calibrate and are
applicable not to one but many loading situations. A list of these criteria is given in Table 1. It is
difficult to compare fracture criteria which are strain based, stress based or mixed stress/strain
based in most general terms. However, in the special case of plane stress, such a comparison can
be made because one can uniquely transform the fracture locus from one space to the other. The
results are presented on a plane, with the equivalent plastic strain as an ordinate and the stress

Table 1
Examples of ductile fracture criteria available in FE codes

# Name Formulated in the space of Example of a code

1 Constant equivalent strain Strain Almost all leading codes


2 FFLR Strain Alcoa in-house codes [6]
3 MS stress Stress None
4 J–C Mixed strain–stress ABAQUS, LS-DYNA, AUTODYN
5 X–W Mixed strain–stress ABAQUS (UMAT)
6 Wilkins (or Dc ; Rc ) Mixed strain–stress LS-DYNA, PAM-CRASH
7 CrachFEM Mixed strain–stress PAM-CRASH
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 721

triaxiality as an abscissa. Each of the seven fracture criteria is represented by one fracture locus in
this space. Many interesting features of various fracture criteria can be revealed by working with
those types of representation.
In order to test a real merit of various theories, an effort was made to perform calibration and
validation on different types of tests. Three tensile tests on unnotched and notched round bars and
one shear test will be available for calibration. Then, all fracture criteria will be specified to the
plane stress condition and the theoretical prediction will be compared with the result of 12
different plane stress tests.

2. Presentation of seven fracture criteria

2.1. Criterion I: constant equivalent strains

Fracture is assumed to occur in a material element when the equivalent plastic strain ¯ reaches a
critical value ¯f
¯ ¼ ¯f . (1)
For an incompressible plastic material ¯ is defined by
rffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2
¯ ¼  þ 22 þ 23 , (2)
3 1
where 1 ; 2 ; 3 are the principal strains. There is an understanding that Eq. (1) is valid for all
possible stress states. Even though the origin of this criterion goes back to the beginning of 20th
century, [7], it has found its way into almost all nonlinear commercial codes in 21st century. While
this criterion lacks generality, practitioners like it because location of possible fracture site can be
determined simply by constructing color-coded plots of the equivalent strain and data can be
easily found in handbook.

2.2. Criterion II: fracture forming limit diagram (FFLD)

The concept of FFLD has been developed in the metal forming industry to characterize
transition from plane stress necking to transverse plane strain fracture. Lee [8], collected
experimental fracture data on sheets made from 11 different materials and ploted them in the
space of two principal strains, Fig. 1. Note that the plot was normalized with respect to the plane
strain fracture strains.
A rather large spread of experimental points can be explained by difficulties in measuring two
principal fracture strains inside a plane strain neck. Another parameter might also be responsible
for the spread of test points. In the first approximation, the FFLD forms a straight line in the
space of principal strains
1f þ 2f ¼ 3f ¼ C, (3)
where the subscript ‘‘f ’’ denotes the strain magnitude at the point of fracture. The range of
applicability of Eq. (3) is between the uniaxial and equi-biaxial tension.
ARTICLE IN PRESS

722 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

~
Uni-axial tension ε1

2.5 plane strain


5154 Al-Mg alloy: Equi-biaxial tension
Embury and LeRoy (1977) 2.0
AKsteel: Ghosh (1975)
HS3 steel: Ghosh (1975)
70-30 brass: Ghosh (1978) 1.5 Normalized FFLD
Steel (n=0.22): Marciniak et al. (1973)
Al (n=0.24): Marciniak et al. (1973)
Copper (n=0.36): Marciniak et al. (1973) 1.0
A1100: Takuda et al. (2000)
A5882: Takuda et al. (2000)
Mild steel (n = 0.22): Lee (2004) 0.5
Mild steel (n = 0.2): Lee (2004)

-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0


ε~2

Fig. 1. Normalized FFLDs for 11 different materials in classical punch indentation problems. Note that all curves are
made to pass through the same point ð~2 ¼ 0; ~1 ¼ 1:0Þ on the ~1 axis [8].

Fig. 2. Shear fracture in an upsetting test of 2024-T351 A1.

2.3. Criterion III: maximum shear (MS) stress

There is a bulk of evidence that ductile fracture may occur on a plane where the shear stress is
maximum. For example, in upsetting test on short aluminum cylinders a spiral fracture occurs in
the equatorial area of barreled specimens, as shown in Fig. 2.
It is then reasonable to postulate that fracture is governed by the condition
tmax ¼ ðtmax Þf , (4)
where
ns  s s  s s  s o
1 2 2 3 3 1
tmax ¼ max ; ; (5)
2 2 2
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 723

and s1 ; s2 ; s3 are the principal stresses. Eq. (4) is similar in form to the Tresca yield condition but
in general ðtmax Þf is different (larger) than the yield stress in shear k: The MS failure criterion has
been successfully applied in soil and rock mechanics and other geomaterials for decades [9,10].

2.4. Criterion IV: Johnson–Cook fracture model (J–C)

Johnson and Cook [11], postulated that the critical equivalent fracture strain (for constant
strain rate and temperature) is a monotonic function of the stress triaxiality
¯f ¼ C 1 þ C 2 expðC 3 ZÞ, (6)
where the stress triaxiality parameter Z is a ratio of the mean stress to the equivalent stress
sm
Z¼ . (7)

Similar forms to Eq. (6) were derived analytically by McClintock [12] and Rice and Tracey [13]
by studying expansion of cylindrical or spherical cavities under hydrostatic tension. Interestingly,
a numerical analysis of the process of linkage of a cluster of three voids gave a similar
expression, [14].
A real success of the J–C model in commercial codes is that Johnson and Holmquist [15]
provided a table of material fracture data for a number of structural materials. The constants C 1 ;
C 2 ; C 3 were determined from tensile tests with high triaxiality and in some cases from a shear test.
An extension of Eq. (6) to the range of small or even negative value of the stress triaxiality, where
no data was collected, may be quite risky. In fact Johnson and Cook [11] reported that shear
fracture strain in one steel was much lower from that predicted by Eq. (6). They admitted that the
‘‘insubordinate’’ data was not included in the curve fitting procedure. One difficulty of calibrating
the J–C model is that the triaxiality parameter Z changes in the loading process and the authors of
Ref. [11] do not indicate how to overcome this problem. In view of a mounting experimental
evidence that material ductility is not a monotonic function of Z; the J–C criterion is gradually
losing its primacy in fracture literature.

2.5. Criterion V: Xue–Wierzbicki model (X–W)

This model is of most recent origin [16,17]. It explains well most if not all experimental
observations and is relatively easy to calibrate. Fracture is postulated to occur when the
accumulated equivalent plastic strain, modified by the function of the stress triaxiality Z and the
deviatoric state parameter x; reaches a limiting value equal to one.
Z ¯f
d¯
¼ 1, (8)
0 FðZ; xÞ
where the deviatoric state parameter, which is related to the Lode parameter, is defined by
27 J 3
x¼ (9)
2 s̄3
in which J 3 is the third invariant of the stress deviators. In terms of principal components s1 ; s2 ; s3
the stress deviator is defined by J 3 ¼ s1 s2 s3 : It should be mentioned that because both parameters
ARTICLE IN PRESS

724 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

fZ; xg are varying in the loading process, additional assumptions are needed to calibrate the X–W
model. For this purpose, average values of these parameters are defined
Z
1 ¯f
Zav ¼ Zð¯Þ d¯, (10)
¯f 0
Z ¯f
1
xav ¼ xð¯Þ d¯ (11)
¯f 0

and used in constructing the fracture locus. Now, replacing current values of ðZ; xÞ by average
values, the integration in Eq. (8) can be performed to give
¯f ¼ F ðZav ; xav Þ. (12)
From now on the subscript ‘‘av’’ will be dropped and ðZ; xÞ will be understood as average values,
unless otherwise stated.
The next step is to specify the form of the function FðZ; xÞ: Xue [17] showed that the fracture
strain is always bounded by two lines corresponding to the axisymmetric stress state and the plane
strain state. The most favorable condition for fracture, meaning the highest ductility is achieved in
tension of round bars where the deviatoric state variable x ¼ 1: There is an abundance of
experiential results showing a decrease of ¯f with the triaxiality variable. This dependence can be
conveniently described by the exponential function
¯axi
f ¼ C1e
C 2 Z
; x ¼ 1, (13)
where the superscript ‘‘axi’’ means axisymmetric. The least favorable condition for fracture is the
plane strain state. The dependence of the equivalent fracture strain on the triaxiality is assumed as
another exponential function
¯ps
f ¼ C3e
C 4 Z
; x ¼ 0, (14)
where the deviatoric state variable x ¼ 0 and superscript ‘‘ps’’ means plane strain. A conceptual
sketch of the above two functions is shown in Fig. 3. In order to determine the intermediate values
of the material ductility for any combination of ðZ; xÞ; consider a plane Z ¼ constant:

_
εf

Upper bound, axial symmetry, ξ = 1

Lower bound, plane strain, ξ = 0

_
σm / σ
0.0

Fig. 3. The axisymmetric stress state provides an upper bound on the material ductility while the lower bound
corresponds to the plane strain condition.
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 725

Wierzbicki and Xue, [16], assumed that the drop of material ductility D¯f due to the deviatoric
state parameter can be described by a family of elliptic functions
!1=m
D¯f
pl
þ x1=m ¼ 1 (15)
axi
¯f  ¯f
in which m is the even integer closest to 1=n where n is the hardening exponent and the drop of
ductility, D¯f ; due to x is defined in Fig. 4.
Resolving Eq. (15) for D¯f and noting that ¯f ¼ ¯axi f ; the final expression for the function
f  D¯
F ðZ; xÞ becomes

¯f ¼ F ðZ; xÞ ¼ C 1 eC 2 Z  ðC 1 eC 2 Z  C 3 eC 4 Z Þð1  x1=n Þn . (16)

The above expression reduces to Eqs. (13) and (14) by taking respectively x ¼ 1 and x ¼ 0: There
are four free parameters in this models, C 1 ; C 2 ; C 3 ; C 4 : However, there is an indication that the
parameters C 2 and C 4 are related to the hardening exponent n; [14]. All calibration constants will
be determined in Section 4. An example of the 3-D fracture locus constructed for the 2024-T351
aluminum is shown in Fig. 5.
Three solid black lines in Fig. 5 denote the loci of points corresponding to plane stress, plane
strain, and axial symmetry.

2.6. Criterion VI: Wilkins model (W)

Even though the model was developed a quarter of a century ago, [18], it was through the work
of the developers of the PAM-CRASH FE code that interest in it has been renewed [19–21]. The
Wilkins model has been implemented in the 2003 release of PAM-CRASH and is now also
available in LS-DYNA. The authors are aware that the extended Wilkins Kamoulakos (EWK)
fracture model is under development at the company Engineering System International, [22].
However, it is the original W model which is evaluated here because the extent of modification of
the EWK model relative to the W model has not been released to public.

εf η=const

∆ε f

εfaxi

εf
ε fplane

-1 0 +1 ξ

Fig. 4. Dependence of the fracture strain in the deviatoric state variable x:


ARTICLE IN PRESS

726 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

Fig. 5. Representation of the fracture locus in the space of stress triaxiality and the deviatoric state variable. Note the
position of lines corresponding to axisymmetric state, plane strain, and plane stress state.

Wilkins postulated that fracture occurs when the following integral exceeds a critical value Dc
over a critical dimension Rc :
Z¯f
1
Dc ¼ ð2  AÞm d¯ in Rc , (17)
ð1  asm Þl
0

where

s2 s2
A ¼ max ; (18)
s1 s3
s1 4s2 4s3 and a; l and m are material constants. These constants as well as the critical parameters
Dc and Rc should be determined from suitably designed calibration tests. Note that the
dependence of fracture strain on the hydrostatic and deviatoric states is assumed in Eq. (17) in a
separable form. The W model includes one more parameter, which is the critical dimension of the
fracture zone Rc : It was shown that for the present material (2024-T351 aluminum alloy)
the output quantities of numerical simulations (stresses and strains) are practically independent of
the mesh size as far as fracture initiation is concerned. Therefore, the critical fracture volume and
Rc can be dropped as an additional parameter of the process.

2.7. Criterion VII: CrachFEM

A group of researches of the BMW R&D Center and MATFEM Co. in Munich proposed
recently a comprehensive fracture model that is now available in PAM-CRASH (Mat #52, 128).
The model is applicable for thin sheets and extrusions (plane stress) and it distinguishes between
two mechanisms responsible for ductile fracture. One is the void growth and coalescence and the
other one is the shear failure model. Accordingly, two different expression for the equivalent
fracture strain are introduced: ¯ductile
f and ¯shear
f : In the case of a ductile void growth model, the
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 727

equivalent strain to fracture is assumed to depend only on the stress triaxiality


¯ductile
f ¼ d 0 eð3cZÞ þ d 1 eð3cZÞ , (19)
where c; d 0 and d 1 are three material constants to be found from tests. In the shear dominated
zone, the fracture strain ¯shear
f is postulated to depend both on the hydrostatic and deviatoric state
according to
¯shear
f ¼ d 2 eðf yÞ þ d 3 eðf yÞ , (20)
where the new ‘‘shear fracture’’ parameter, y; is defined by

y¼ ð1  3ks ZÞ. (21)
tmax
Again, there are four material constants in the model, ks ; f ; d 2 and d 3 : According to the theory,
there are two competing fracture modes and the one for which the calculated fracture strain
(¯ductile
f or ¯shear
f ) is the lowest, determines the fracture limit. The model was clearly developed with
a view of industrial application but involves at total of seven free parameters. The shear fracture
model acknowledges the joint effect of stress triaxiality and the deviatoric state through the
definition of the parameter y: Thus, conceptually, it is similar to the Wilkins and X–W model.

2.8. Criterion VIII: Cockcroft–Latham

In addition to the above seven fracture models which are applicable to the wide range of the
stress triaxiality, the well known Cockcroft–Latham, [28] fracture criterion is also studied here. It
was developed for the bulk forming operations and therefore is applicable only to the range
of small and negative stress triaxilaity. According to this criterion, fracture occurs when
the accumulated equivalent strain modified by maximum principal tensile stress reaches a
critical value
Z ¯f
s1 d¯ ¼ C, (22)
0

where C is the calibration constant.

3. Summary of the experimental program

Bao [1] performed fifteen different tests on 2024-T351 aluminum alloy, covering a wide range of
stress triaxialities 0:3oZo0:9: Tests #1–4 are used for calibrations of the fracture models while
Tests 5–15 are used for validation and relative evaluation (see Table 2). All specimens shown in
Figs. 6 and 7 were cut from the same block of aluminum in the rolling direction. The tests were
performed on a universal testing machine and the following data were recorded: the
force–displacement relation, location of the first crack, and the displacement to fracture. In
some specimens (Tests #1–4, 15), measured was also the area reduction.
The true stress–strain curve was found from the data on the unnotched round specimen. Up to
the point of necking, the area reduction is uniform over the gauge section and the true strain could
ARTICLE IN PRESS

728 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

Table 2
Summary of the experimental program

Test number Specimen description ¯f Zav xav ¯f

1 Round, smooth 0.46 0.40 1.0 0.45


2 Round, large notch 0.28 0.63 1.0 0.3
3 Round, small notch 0.17 0.93 1.0 0.21
4 Flat-grooved 0.21 0.61 0.097 0.21
5 Cylinder ðd 0 =h0 ¼ 0:5Þ 0.45 0.278 0.91
6 Cylinder ðd 0 =h0 ¼ 0:8Þ 0.38 0.234 0.81
7 Cylinder ðd 0 =h0 ¼ 1:0Þ 0.356 -0.233 0.82
8 Cylinder ðd 0 =h0 ¼ 1:5Þ 0.341 0.224 0.80
9 Round notched (compression) 0.62 0.248 0.84
10 Flat dog-bone tensile 0.21 0.0124 0.055
11 Flat 0.26 0.117 0.50
12 Plate with a circular hole 0.31 0.343 1.0
13 Dog-bone 0.48 0.357 0.979
14 Pipe 0.33 0.356 0.984
15 Solid bar 0.36 0.369 1.0 0.29

Note that ¯f is the average cross-section strain evaluated from the measured area reduction according to Eq. (25).

Fig. 6. Tests on unnotched round bar #1; two notched bars #2 and 3, and flat grooved specimens used for calibration
of seven fracture models.

be calculated from the measured relative displacement. After necking, the equivalent stress and
strain could be determined from a continuous measurement of the diameter of the neck. Such a
technique was not available in the lab. Instead, the true stress–strain curve was determined
through an iterative procedure based on a detail numerical simulation of the necking process. This
technique was described in a number of previous publications, see for example [1–3]. The true
stress–strain obtained in this way is shown in Fig. 8. Note that the solid line represents the portion
of the curve before necking while the dotted line was determined by means of above mentioned
iterative procedure. The open circle in Fig. 8 corresponds to the final fracture point for which the
area reduction was measured. This point seems to lie very close to the experimentally/numerically
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 729

Fig. 7. Upsetting #5–9, shear #10; 11, and tensile #12–15 tests used for validation.

Fig. 8. A power law fit for Al 2024-T351 up to the necking point (solid line), extrapolated beyond necking (dashed line).

obtained stress–strain curve. A good power law fit s ¼ Kn of the entire stress–strain curve gives
the amplitude K ¼ 744 MPa and the hardening exponent n ¼ 0:153:
The exact s   curve (not the power law fit) was then input into the material model of
ABAQUS/Explicit, and numerical simulation of all 15 specimens was performed. The history of
the equivalent plastic strain ¯ðuÞ was calculated as a function of the cross-head displacement u at
the potential fracture location. The equivalent strain at the expected fracture location
corresponding to the measured displacement to fracture uf was then taken as the fracture strain
¯ðu ¼ uf Þ ¼ ¯f : In addition, histories of stress triaxiality parameter Zð¯Þ and the deviatoric stress
parameter xð¯Þ were also calculated.
ARTICLE IN PRESS

730 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

Fig. 9. Representation of all 15 experimental points in the space of Z and x: Note that all points lie almost exactly on
three lines corresponding, respectively to plane stress, plane strain, and axial symmetry.

The corresponding average values of parameters Z and x were then calculated from Eqs. (10)
and (11). A summary of the experimental results obtained with the help of numerical simulation is
given in Table 2. Various types of specimens generate a different stress state. It is interesting to
show a plot the results of Table 2 on the plane fZ; xg which uniquely characterize the stress state.
The solid line in Fig. 9 represents the locus of points corresponding to the plane stress case. It was
shown in [16] that the condition s3 ¼ 0 in the space of principal stresses uniquely maps into the
third order parabola in the space of stress invariants.

27 2 1
x¼ Z Z  . (23)
2 3
Experimental points #1–3 stay on the line corresponding to axial symmetry, x ¼ 1: The flat
grooved specimen (experimental point #4) is in the state of transversal plane strain. Indeed, it lies
at the intersection of the plane stress curve, Eq. (16) and the plane strain line, x ¼ 0: These four
tests will be used for calibration of seven fracture criteria. The point with the coordinates ð0; 0Þ
represents pure shear. In some cases, it will also be used for calibration instead of the Test #4.
Points corresponding to Tests #5–15 are seen to congregate around the solid line and thus must be
in or close to the plane stress state. These additional 11 tests will serve as a check of accuracy of
various fracture models. Because in the plane stress state the variables Z and x are related through
Eq. (23), the 3-D plot of the fracture locus, shown in Fig. 5 can now be projected into a plane
f¯f ; Zg:
Additional comment is needed regarding the upsetting test. The diameter to height ratio in
upsetting Tests #5–8 was in the range 0.5 to 1.5. Cylinders with the smaller ratio (tall) were also
tested but most cases short column shear instability developed and the test had to be interrupted.
However, other authors reported on successful tests covering a wide range of diameter to height
ratio. Collection of experimental results from five different sources (including our own five points)
is shown in Fig. 10. It was pointed out by Kudo and Aoi [26] that the trend of test points could be
fitted by a straight line
1f þ 12 2f ¼ C 8 , (24)
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 731

Experiments from literature


Al 2024: Bao and Wierzbicki (2004a)
Steel: Kudo and Aoi (1967)
Al 2014: Erturk and Kazaoglu (1982)
ε~1
Pure shear 1045 steel: Kuhn (1976)
UNSL52905 Lead alloy:
Gouveia et al. (2000)
branch (III) 2.5
Normalized fracture locus: Eq. (24)
2.0

Uni-axial 1.5
compression

1.0

0.5

-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0
ε~2

Fig. 10. Normalized fracture loci for five different materials in upsetting tests. Note that all curves are made to pass
through the same point ð~2 ¼ 1:0; ~1 ¼ 1:0Þ [8].

where C 8 is the magnitude of plane fracture strain. A transformation of this equation to the space
of equivalent strain and stress triaxiality is given in Section 4.8.
Questions might be raised regarding the accuracy of our experimental/numerical method for the
determination of the equivalent fracture strain ¯f : However, some of the results were double
checked by measuring the area reduction in axi-symmetric tensile specimens. From this data the
true logarithmic strain, averaged over the cross-section, can be calculated
A0
¯f ¼ ln . (25)
Af
An additional column was created in Table 2 with fracture strain ¯f measured directly from the
experiment. There are five entries in this column. In three cases, the calculated and measured
fracture strains were quite close. There were however large difference in the remaining cases. The
mesh size effect was also studied. Results for three different meshes were compared and they were
found to be close (within 4%). It can be concluded that in the present experimental problem
dealing with fracture initiation, the gradients of stresses and strains are not large, leading to a
practical insensitivity of the results to the mesh size. For a thorough study of the mesh sensitivity
on initiation and propagation of cracks, the reader is referred to the work by Lee, [8].

4. Calibration of seven fracture models

Various fracture criteria require a different number of calibration tests and ranges from one to
four. Table 3 provides information on a minimum number of tests needed for each of the seven
fracture criteria.
ARTICLE IN PRESS

732 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

Table 3
Minimum number of fracture test to calibrate a given fracture criterion

# Name # of tests

1 Constant equivalent strain 1


2 FFLD 1
3 MS stress 1
4 J–C 3
5 X–W 4
6 Wilkins (or Dc ; Rc ) 4
7 CrachFEM 4

4.1. Calibration of the constant equivalent strain criterion

There is a great deal of arbitrariness in deciding which particular test should be used for
calibration. To a large extent, the decision would depend on the range of expected application.
For example, in problems of punch indentation or explosive loading of thin clamped plates, the
state of stress varies between the equi-biaxial tension at the center ðZ ¼ 23Þ to plane strain at the
clamped boundary ðZ ¼ p1ffiffi3Þ: From Table 2, one would then choose Test # 4, corresponding to
the transverse plane strain. In this paper, the magnitude of the critical equivalent fracture strain is
assumed to be ¯f ¼ 0:21: The line at constant ¯f comes thorough two experimental points
corresponding to transverse plane strain (Z ¼ p1ffiffi3; Test #4) and pure shear (Z ¼ 0; Test #10). It
should be noted that traditionally, the calibration of the constant strain fracture criterion is
performed on the basis of a tensile test on an unnotched specimen. The corresponding fracture
strain is much higher ð¯f ¼ 0:45Þ leading to a poor approximation of most of test points. Even
more often, ¯f is treated as an adjustable parameter to get the best correlation between component
tests and simulations. This procedure works, but it is fundamentally incorrect.

4.2. Calibration of the FFLD criterion

Because the FFLD is represented in the ð¯1 ; ¯2 Þ space by a straight line with 45
slope only, one
test is needed for calibration. The FFLD applies only for sheets, which are in the state of plane
stress. Lee [8] proved that the straight line in the space of principal strains can be transformed into
the hyperbola in the ð¯f ; ZÞ space, as shown in Fig. 11.
Furthermore, the above transformation gives a unique relationship between the calibration
constants (see boxes in Fig. 11). Theoretically, the position of the FFLD line could be based on
either uniaxial strain, plane strain or equi-biaxial strain. In the present paper, the data for the
transverse plane strain are used for calibration (see Fig. 12).

4.3. Calibration of the MS stress criterion

There is only one parameter in this model, which is the MS stress at fracture, ðtmax Þf : In the
typical cup-and-cone fracture mode, a crack initiates at the center as a result of the void growth
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 733

εf
Dc
εf =
η

0 1 1
η
2
3 3 3
ε1
Uni-axial

Plane strain

Equi Bi-axial

ε2
0
3Dc
ε1f + ε 2 f =
2

Fig. 11. Transformation of the FFLD fracture locus into the space of (¯f ; Z).

_
εf

experiment (plane stress)


experiment (axial symmetry)
0.8

9
FFLD
13
1
5
6 0.4
15
7 14 2
8 11
12
10 4
3

full point used for calibration

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
m /

Fig. 12. The range of applicability for FFLD is very limited in the present space and its applicability to plane stress
fracture problem is inconclusive.

and coalescence mechanism. It is only later that the shear fracture takes over. Calibration should
not be made on the basis of the round bar tensile tests. A more representative test to calibrate the
present criterion is the pure shear test (Test #10) or the transverse plane strain test (Test #4). The
fracture locus is composed of two branches corresponding to two sides of the Tresca fracture
hexagon in the space of principal stresses. Both Tests #4 and 10 are in the state of plane stress as
ARTICLE IN PRESS

734 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

well. The fracture locus is given in the parametric form by the following equation
(pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi)1=n
1 þ a þ a2 1 1 2
¯f ¼ C for  oao1 or oZo ,
2þa 2 3 3
(pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi)1=n
1 þ a þ a2 1 1 1
¯f ¼ C for  2oao  or  oZo , ð26Þ
1a 2 3 3
pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where Z ¼ ð1= 3Þðða þ 1Þ= 1 þ a þ a2 Þ and a ¼ d2 =d1 is the strain ratio parameter and C is the
calibration constant. By picking any pair f¯f ; Zg from Table 2, the calibration constant can be
found. Taking Test #10 in pure shear ðZ ¼ 0; a ¼ 1Þ; the calibration constant becomes C ¼ 0:21:
A plot of the function ¯f ðZÞ superimposed on all test points is shown in Fig. 13. Surprisingly, there
is an almost perfect correlation.

4.4. Calibration of the J–C criterion

Johnson and Holmquist [15] reported the following values of the material constants for 2024-
T351 aluminum C 1 ¼ 0:13; C 2 ¼ 0:13; C 3 ¼ 1:5: The ductility function, Eq. (6), corresponding
to these values is shown in Fig. 14 by a broken line.
This curve misses almost all experimental points except for pure shear and tension. It would
appear that Johnson and Holmquist provided a lower bound curve for material ductility. Besides,
three tests are needed to find three free parameters but the solid line curve in Fig. 14 passes
approximately only through two experimental points. By making Eq. (6) pass through three high
triaxiality points (#1, 2 and 3), a different set of calibration constants is obtained, C 1 ¼ 0:07;
C 2 ¼ 1:02; C 3 ¼ 1:62: The corresponding ductility curve, which is depicted in Fig. 14 by a solid
line, is the upper bound for the equivalent fracture strain.

_
εf

experiment (plane stress)


experiment (axial symmetry)
0.8

Maximum shear stress τmax = const.


9

13
5 1
6 0.4
15
7 14 2
8 11 12
10 4
3

full point used for calibration

-0.4 -0.2 0.0 0.2 _0.4 0.6 0.8 1.0


σm /σ

Fig. 13. Prediction of the MS stress criterion agrees quite well with all plane stress experimental points.
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 735

_
εf full points used for calibration

Johnson-Cook; Present
C1=-0.07, C2=1.02, C3=-1.62

0.8

Johnson-Cook; Ref. [15]


C1=0.13, C2=0.13, C3=-1.5

0.4

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
σm /σ

Fig. 14. Comparison of the ductility curve for 2024-T351 aluminum.

The present discussion raises a number of questions. Is the block of 2024-T351 aluminum that
Johnson and Cook tested in 1985 the same has the block form which the test samples were cut for
the present study? Can one rely on textbook values of material constants? Which experiments are
most representative for calibration. The last question will be ever present in calibrating the
remaining three fracture models.

4.5. Calibration of the X–W fracture criterion

There are four free parameters in Eqs. (13)–(15), C 1 ; C 2 ; C 3 ; and C 4 : The pair C 1 and C 2 can be
uniquely found from two axisymmetric tests in which there is no loss of ductility due to deviatoric
state (Tests #2 and 3) and x ¼ 1: The other pair C 3 and C 4 are determined from two test points
corresponding to the pure shear (Test #10) and the transverse plane strain (Test #4) in which there
is a maximum loss of ductility due to the deviatoric state, and x ¼ 0: The hardening exponent for
the considered aluminum alloy is n ¼ 0:153 and 1=n ¼ 6:53: The closest even integer of 6.53 is 6.0.
Therefore, the magnitude of the power of the nonlinear ellipse in Eq. (15) is m ¼ 6:0:
The solution of two pairs of algebraic equations give the following values of the calibration
constants C 1 ¼ 0:87; C 2 ¼ 1:77; C 3 ¼ 0:21; C 4 ¼ 0:01: The correlation of theoretical prediction
with the test results, shown in Fig. 15, is very good in the range of both low and high stress
triaxialities. For comparison, curves corresponding to m ¼ 4 and m ¼ 2 are also given in Fig. 15.
It is seen that the curves corresponding to m ¼ 4 and m ¼ 6 are almost identical.
It is interesting to note that Eq. (16) predicts a high but finite value of the fracture strain in the
case of an uniaxial compression Z ¼ 1=3; x ¼ 1: In this case, from Eq. (13), one gets ¯f ¼
0:87e1:77=3 ¼ 1:57: The magnitude of the fracture strain is indeed very high, almost unattainable in
ARTICLE IN PRESS

736 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

_
εf

experiment (plane stress)


experiment (axial symmetry)
0.8 full points used for calibration

9 Xue-Wierzbicki, m = 2
Xue-Wierzbicki, m = 4
Xue-Wierzbicki, m = 6
13
5 1
6 0.4
15
7 14 2
8 11 12
10 4
3

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
σm /σ

Fig. 15. Prediction of the X–W fracture model for three values of the parameter m:

standard tests. This effect was captured by Bao–Wierzbicki’s [2] empirical fracture locus which
assumed a ‘‘cut-off’’ value of the stress triaxiality at Z ¼ 1=3 below which no fracture would
occur.

4.6. Calibration of the Wilkins criterion

There are four parameters in the W model, a; l; m and Dc ; see Eq. (10). Three tests on round
bars (#1, 2 and 3) and one Test #10 on pure shear are chosen for the present calibration. It can be
shown that in axisymmetric tests the variable A ¼ 1 due to symmetry srr ¼ syy at the center of the
round bar. Then Eq. (17) reduces to
f ¼ Dc ð1  as̄ZÞl . (27)
On the other hand, in a pure shear test sm ¼ 0; A ¼ 0 and in the limit Eq. (17) yields
f ¼ Dc =2m. (28)
Substituting the experimental values from Table 2 to Eqs. (27) and (28), a system of four
nonlinear algebraic equations is obtained for four unknowns. The system is partially uncoupled.
First, the three equations can be solved for a; m; and Dc : Then, from the fourth equation the
parameter m is calibrated. The following set of numbers satisfies approximately the above system
a ¼ 1:2½GPa1 ; l ¼ 2:15; m ¼ 2:18; Dc ¼ 0:93: The W model predicts reasonably well the trend of
experimental points in the range 1=3oZo1=3 but underpredicts the material ductility for
1=3oZo2=3; see Fig. 16. This is due to the assumption of a separable form of the dependence on
x and Z: Also, the double ‘‘U’’ shape is too narrow leaving test points #6–9 too fat to the left.
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 737

_
εf

experiment (plane stress)


experiment (axial symmetry)
0.8
full points used for calibration

Wilkins
13
1
5
6 0.4
15
7 14 2
8 11 12
10 4
3

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
σm /σ

Fig. 16. Prediction of the Wilkins model compared with a set of plane stress experimental points.

Table 4
Values of stress triaxialities in three calibration tests on ductile fracture used in Ref. [23] and the present paper

Hooptura et al. [23] Z Present Zav

Notched flat tensile specimen 0.333 Flat tensile specimen, #13 0.35
Plane strain 3-point bending 0.58 Flat grooved #4 specimen 0.61
Punch loading equi-biaxial 0.667 Round notched tensile specimen #3 0.93

4.7. Calibration of the CRACH FEM model

Two different branches of the fracture locus can be calibrated separately. The three free
parameters fc; d 0 ; d 1 g characterizing ductile fracture, Eq. (19), will be determined from the Tests
#1, 3 and 4. In making a choice of the most appropriate tests for calibration, an effort was made
to recreate the same type of test as in the original paper by Hooptura et al. [23] and Gese et al.
[24]. A comparison of the value of main parameters between Ref. [23] and the present paper is
given in Table 4.
Substituting the data from Tests #13, 3 and 4 into Eq. (19) and solving a system of nonlinear
algebraic equation, one gets c ¼ 1:33; d 0 ¼ 0:0029 and d 1 ¼ 2:016: With the above values, the
predicted fracture strain in the ductile fracture is shown in Fig. 17 by thick solid line. Clearly, the
curve passes through the coordinates of test points #13, 3, and 4. A different set of tests (#4, 10
and 13) were selected for the calibration of the shear fracture branch of the fracture locus. Again,
an effort was made to keep the parameter y as close as possible to the value reported in Ref. [23],
ARTICLE IN PRESS

738 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

_
εf experiment (plane stress)
experiment (axial symmetry)

full points used for calibration

0.8 CrachFEM (ductile), #13, #3, #4


c = -1.33; d0 = 0.003; d1 = 2.02
9
CrachFEM (shear), #4, #10, #13
ks = 0.01; d2 = 1560.42; d3 = 7.01e-06; f = 5.6
13
5 1
6 0.4
15
7 14 2
8 11 12
10 4
3

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
σm /σ

Fig. 17. Comparison of the prediction of the CrachFEM model (both ductile and shear fracture modes) with present
experimental points.

Table 5
Parameters characterizing three calibration tests on shear fracture used in Ref. [23] and the present paper
tmax
Hooptura et al. [23] y Present s̄ y Z ¯f
pffiffiffi
Pure shear 3 Pure shear # 10 p1ffiffi 1.732 0.0124 0.21
3
Shear/tension 1.47 Transverse plane strain, # 4 p1ffiffi 1.4 0.611 0.21
3
Biaxial tension 1.6 Uniaxial tension # 13 0.5 1.75 0.357 0.48

see Table 5 below. The parameter y is defined by Eq. (21). By solving a set of nonlinear algebraic
equations, the following values of the free parameters were obtained: d 2 ¼ 1560; d 3 ¼ 7:01e  06;
f ¼ 5:60:
It should be noted that the parameter y was calculated from Eq. (21) taking a pre-defined value
ks ¼ 0:01: Two of the tests that were used for ductile fracture calibration were used again for the
calibration of shear fracture. By doing this, a number of free parameters in this model was
reduced from six to four (see solid black points in Fig. 17). Curves corresponding to the ductile
and shear fracture modes are shown in Fig. 17 with a corresponding legend.
The branch corresponding to the shear fracture is always lower than the ductile fracture branch.
It is believed that a different set of tests would significantly improve the accuracy of the
CrachFEM model.
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 739

4.8. Calibration of the Cockcroft–Latham model

In the case of plane stress, it is possible to prove that the empirical fracture locus from the
upsetting test, given by Eq. (24) is equivalent to the Cockcroft–Latham fracture criteria presented
in a normalized form
Z ¯f
syy 4
d ¯ ¼ C 8 , (29)
0 s̄ 3

where syy is the hoop stress in the equatorial region of the barreled cylinder. Using the plane stress
yield condition and the associate flow rule, Eq. (24) can be further transformed to the space of the
equivalent plastic strain and stress triaxiality to give
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
C 8 3Z þ 12  27Z2
¯f ¼ pffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi . (30)
3 2ð1 þ Z 12  27Z2 Þ

There is one calibration constant characterizing the Cockcroft–Latham fracture model.


Therefore, only one test is needed to find unknown coefficient C 8 : We chose for that purpose
Test #7. Taking the corresponding values of ¯f and Z from Table 2, one obtains C 8 ¼ 0:12: A plot
of Eq. (30) with superposed experimental points is shown in Fig. 18. It is seen that the
corresponding curve predicts extremely well fracture in all upsetting tests but fails to describe
fracture all other tests including pure shear. For this reason, the Cockcroft–Latham fracture
criterion has not found its way to the material libraries of commercial codes and will not be
pursed any further.

_
εf

experiment (plane stress)


experiment (axial symmetry)
0.8
full points used for calibration
9

13
5 1
Cockcroft-Latham
6 0.4
15
7 14 2
8 11 12
10 4
3

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
σm /σ

Fig. 18. Comparison of the prediction of the Cockcroft-Latham model with present experimental points.
ARTICLE IN PRESS

740 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

5. Discussion and conclusions

A summary of the prediction of seven fracture models is shown in Fig. 19. There are clear
‘‘winners’’ and ‘‘losers’’. In the industrial environment, the quality of a product (which in this case
is a fracture predictive technology) is a weighted function of performance vs. cost. Performance is
measured by the accuracy of numerical simulation as compared to tests. The cost is related here to
a number of calibration tests and to the ease in performing standard experiments.
An absolute winner of the present ‘‘contest’’ is the old and faithful MS stress failure condition.
This criterion not only follows the trend of experimental points with the ‘‘engineering’’ type accuracy
but require only one test for calibration. Any test from the collection of plane stress or plane strain
category can be taken for that purpose. A partial success of this formulation should not be surprising
because the MS stress carries information on the second and third stress invariants, [16]. It does not
carry information on the hydrostatic state. A drawback of the method is that it cannot predict
fracture in axisymmetric loading situations. In general, good results are obtained when fracture is
preceded by shear localization (as in plane strain) or occurs in the shear decohesive mode (Fig. 2).
This criterion holds promise to predict slant fracture which so far defies conventional methods.
The X–W fracture criterion was constructed in the three-dimensional space of invariants. It can
adequately deal with all stress states which can then be projected on the space of the equivalent
fracture strain and stress triaxiality. The model can predict fracture in all types of experiment with
superior accuracy. It requires however four independent tests for calibration. Bandstra and Koss
[14] suggested that the rate of growth of ductility with diminishing pressure depends on the degree

_
εf experiment (plane stress)
experiment (axial symmetry)

Xue-Wierzbicki

CrachFEM (ductile)
0.8 CrachFEM (shear)

9 τ max = const.
FFLD
_
13 ε f = const.
5 1
Wilkins
6
0.4
7 15 Johnson-Cook:
8 14 2 C1=-0.07, C2=1.02, C3=-1.62
11 12
10 4 Johnson-Cook:
3 C1=0.13, C2=0.13, C3=-1.5

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0


_
σm /σ

Fig. 19. Comparison of prediction of all seven fracture criteria relatively to the set of 12 test points (plane stress) on
2024-T351 aluminum specimens.
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 741

of hardening. Had such a relationship been established, the number of needed tests for calibration
would have been reduced to three or even two.
The Wilkins criterion recognizes the importance of both the hydrostatic and deviatoric states on
fracture. However, the joint effect on these two variables on the material ductility is assumed to be
of a separable form. As a result, good prediction is obtained either in the range of large
ð1=3oZo2=3Þ or small stress triaxiality ð1=3oZo1=3Þ; but not in both.
The CrachFEM criterion holds a lot of promise because it is physically based and distinguishes
between the ductile fracture driven by void growth and shear fracture. In fact, there are
advantages and disadvantages of a separation of these two effects. For example, the
monotonically varying dependence of the fracture strain with the hydrostatic pressure (first term
in Eq. (19)) cannot capture the difference in ductility between the equi-biaxial tension and the
transverse plane strain. Hooptura et al. [23] proposed to fix this discrepancy by introducing
another term depending on the stress triaxiality. The present authors believe that this effect is due
to the loss of ductility caused by the deviatoric state rather than by the hydrostatic state. On the
positive side, it might be easier to calculate the accumulated damage separately in the ductile
fracture and in the shear fracture for nonlinear strain path, [25].
It should be noted that the parameter ks in Eq. (21) is very small. Thus, practically the shear
fracture parameter is mainly controlled by the deviatoric state parameter tmax =s̄ with very weak effect
of the hydrostatic term 3ks Z: In this regard, the CrachFEM shear fracture model is similar to the MS
stress criterion. It remains to be seen if such a decoupling will produce good results for other materials.
The remaining three fracture criteria i.e. the J–C model, the FFLD approach and the constant
equivalent strain method can only be used in situation when the stress triaxiality and/or the
deviatoric state parameter vary in very narrow ranges which are known ahead of time.
Four fracture models predict sharp peak in the fracture strain around the stress triaxiality
corresponding to uniaxial tension. One would then expect a large scatter of the test data where a
small difference in test condition (stress triaxiality) could lead to a large difference in ductility. By
comparing the positions of the points #12–15, such a scatter can clearly be distinguished in Fig. 19.
The present assessment of the accuracy of various models relative to experimental data was
made for one material only. Care should be taken in extending some of the conclusions to other
materials. At the same time, comments made on the general formulation in terms of
interdependence on hydrostatic and deviatoric state are valid for any material. There is one
important difficulty in the validation of the present macroscopic approach to fracture. In almost
all types of experiments (perhaps with the exception of pure shear), the independent variables
fZ; xg change in the loading process. Therefore, average values of these variables fZav ; xav g have to
be used in constructing a fracture locus. The above limitation calls for a new type of test, for
example a shear test with a constant hydrostatic pressure or a modified Bridgman test with a
suitably controlled and varying pressure to compensate for the varying triaxiality inside the
axisymmetric neck.

Acknowledgements

The authors wish to thank Dr. Heinz Werner of BMW R&D Center in Munich for the helpful
discussion on the manuscript. Thanks also to Dr. Helmut Gese for his help in the calibration of
ARTICLE IN PRESS

742 T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743

the CrachFEM shear fracture model. The assistance of Li Zheng and Xiaoqing Teng in the
technical preparation of the manuscript is greatly appreciated. Support of this work came in part
from an ONR/MURI award to MIT, an NNR initiative through the Office of Naval Research,
and a GE Global Research grant.

References

[1] Bao Y. Prediction of ductile crack formation in uncracked bodies. Ph.D. thesis. Cambridge, MA: Department of
Ocean Engineering, Massachusetts Institute of Technology; 2003.
[2] Bao Y, Wierzbicki T. On fracture locus in the equivalent strain and stress triaxiality Space. International Journal
of Mechanical Sciences 2004;46:81–98.
[3] Bao Y, Wierzbicki T. A comparative study on various ductile crack formation criteria. Journal of Engineering
Materials and Technology 2004;126(3):314–24.
[4] Atkins AG. Fracture mechanics and metal forming: damage mechanics and the local approach of yesterday and
today. Fracture Research in Retrospect—An Anniversary Volume in Honor of George R. Irwin’s 90th Birthday.
Rotterdam, Brookfield: A.A. Balkema; 1997.
[5] McClintock FA. Slip line fracture mechanics: A new regime of fracture mechanics. Fatigue and Fracture
Mechanics, vol. 33, ASTM STP-1417, West Conshohocken, PA; 2002.
[6] Yeh JR, Summe TL, Seksaria DC. The development of an aluminum failure model for crashworthiness design.
Crashworthiness, Occupant protection and biomechanics in transportation systems, AMD-vol. 237/BED-vol. 45.
ASME; 1999. p. 97–105.
[7] Huber MT. Contribution to the foundation of the strength of the material (in Polish, translated to English by
Professor. M. Zyczkowski in connection with the M.T. Huber Century Symposium, Krakow, August, 2004).
Czasopismo Techniczne, Lwow 1904; 22: 81.
[8] Lee YW. Fracture prediction in metal sheets. Ph.D. thesis, Cambridge, MA: Department of Ocean Engineering,
Massachusetts Institute of Technology, 2004.
[9] Bardet JP. Lode dependences for isotropic pressure-sensitive elastoplastic materials. Transactions of the ASME
1990;57:498–506.
[10] Bigoni D, Piccolroza A. A new yield function for geomaterials. In: Viggiani C, editor. Constitutive modelling and
analysis of boundary value problems in geotechnical engineering, Napoli; 22–24 April 2003, p. 266–81.
[11] Johnson GR, Cook WH. Fracture characteristics of three metals subjected to various strains, strain rates,
temperatures and pressures. Engineering Fracture Mechanics 1985;21(1):31–48.
[12] McClintock FA. A criterion for ductile fracture by growth of holes. Transactions of the ASME Journal of Applied
Mechanics 1968;35:363–71.
[13] Rice JR, Tracey DM. On the ductile enlargement of voids in triaxial stress fields. Journal of the Mechanics and
Physics of Solids 1969;17:201–17.
[14] Bandstra JP, Koss DA. A simulation of growth and coalescence of voids during ductile fracture. Materials Science
and Engineering, A 2004;387–389:399–403.
[15] Johnson GR, Holmquist TJ. Test data and computational strength and fracture model constants for 23 materials
subjected to large strain, high strain rates, and high temperature. Technical Report LA-11463-MS, Los Alamos
National Laboratory; 1989.
[16] Wierzbicki T, Xue L. On the effect of the third invariant of the stress deviator on ductile fracture. Impact and
Crashworthiness Lab Report #136; 2005, International Journal of Fracture, submitted for publication.
[17] Xue L. Damage accumulation and fracture initiation in uncracked ductile solids under triaxial loading—Part I:
Pressure sensitivity and Lode dependence. Impact and Crashworthiness Lab Report #138; 2005, submitted for
publication.
[18] Wilkins ML, Streit RD, Reaugh JE. Cumulative-strain-damage model of ductile fracture: simulation and
prediction of engineering fracture tests. Technical Report UCRL-53058, Lawrence Livermore National
Laboratory; October 1980.
ARTICLE IN PRESS

T. Wierzbicki et al. / International Journal of Mechanical Sciences 47 (2005) 719–743 743

[19] Kamoulakos A, Culiere P, Araki T. Prediction of ductile metal rupture with the E-W Model in PAM-CRASH. In:
IBEC 2003 Chiba, Japan; 2003.
[20] Lemoins X, Kamoulakos A. Calibrating and using the E-W rupture model: Application on Arcelor high strength
steel. In: EuroPAM 2003, Mainz, Germany; 2003.
[21] Kamoulakos A. The ESI-Wilkins–Kamoulakos (EWK) rupture model. In: Raabe D, Chen LQ, Barlat F, Roters F,
editors. Continuum Scale Simulation of Engineering Materials. Berlin, Germany: Wiley-VCH, Verlag, Berlin
GmbH; 2004.
[22] Kamoulakos A. Private communication; 2004.
[23] Hooputra H, Gese H, Dell H, Werner H. A comprehensive failure model for crashworthiness simulation of
aluminum extrusions. International Journal of Crashworthiness 2004;9(5):449–63.
[24] H. Gese, H. Werner, H. Hooputra, H. Dell, A. Health, CrachFEM—a comprehensive failure model for metallic
structures in sheet metal forming and crash simulation. EuroPAM 2004; 12 October 2004.
[25] H. Gese, Private communication; 2005.
[26] Kudo H, Aoi K. Effect of compression test conditions upon fracturing of medium carbon steel. Journal of the
Japan Society of Technical Plasticity 1967;18:17–27.
[28] Cockcroft MG, Latham DJ. Ductility and the workability of metals. Journal of the Institute of Metals
1968;96:33–9.

Further Reading

[27] Atkins AG, Mai YW. Crack and craze nucleation. In: Elastic and plastic fracture. Ellis Horwood: Chichester;
1985. p. 369–431.

You might also like