Sei sulla pagina 1di 25

International Journal of Plasticity 71 (2015) 62e86

Contents lists available at ScienceDirect

International Journal of Plasticity


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

A bifailure specimen for accessing failure criteria performance


Larissa Driemeier*, Rafael T. Moura, Izabel F. Machado, Marcilio Alves
~o Paulo, Polytechnic School, Av.Prof. Mello de Moraes, 2231 Sa
University of Sa ~o Paulo, SP, Brazil

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

Article history: Many engineering ductile materials, such as metals, show inelastic behavior with corre-
Received 18 June 2014 sponding large deformations. For these materials, the prediction of failure, defined as local
Received in revised form 20 February 2015 material separation, is still nowadays a scientific challenge. Failure initiation, its locus and
Available online 19 March 2015
evolution, has been discussed extensively in recent literature. This work presents a bifai-
lure specimen, especially developed to evaluate failure criteria at high and low triaxiality.
Keywords:
In order to explore further the failure phenomenon, uncoupled models are numerically
C. Finite elements
implemented in an explicit finite element code. Simulations were performed and do
A. Fracture
Triaxiality
indicate that, for the studied examples, uncoupled models are not capable of correctly
C. Mechanical testing predicting failure.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction

Nowadays, the engineering international scenario is characterized by rapid technological development. It is evident the
need for accurate and realistic virtual models, in order to minimize the number of experimental tests. Moreover, as pointed
out by Matsumoto et al. (2012), post-processing tools allow a more detailed and deep analysis of the behaviour and sensibility
of a component or structure.
The quality of virtual analyses is basically dependent on the knowledge of the material behavior, geometry, loading and
boundary conditions. The large number of constitutive models presented in the literature can describe different material
behaviors fairly well under different conditions of loading, strain rate and temperature. However, for a plasticity model to be
more general, the number of material constants should increase. Here it is opted for a simple isotropic plasticity model based
on a quadratic yield criterion (Hill, 1950), where temperature and strain rate effects are neglected. Such a model is readily
available in many codes, such that the user material subroutine here implemented could be validated with a small number of
physically-based parameters. More sophisticated yield surface definitions (Barlat, July 1987; Cazacu and Barlat, 2004; Barlat
et al., 2007; Brünig, 1999; Brünig and Driemeier, 2007) could enhance the numerical results, particularly for metal sheets, but
are not considered here.
On the other hand, damage and failure phenomena offer so many challenges that so far it remains a key issue in design and
they are explored in many experimental investigations in literature. Historically, Bridgman (1952) tested the effect of hy-
drostatic pressure on the fracture of different types of steel. The author performed uniaxial tensile tests under high pressure
conditions and concluded that deformation to fracture increases with increasing hydrostatic pressure. Rice and Tracey (1969)
concluded that, for moderate to high levels of stress triaxiality, defined as a ratio between hydrostatic and equivalent stress,
the voids growth rate increases exponentially with the hydrostatic stress.

* Corresponding author.
E-mail address: driemeie@usp.br (L. Driemeier).

http://dx.doi.org/10.1016/j.ijplas.2015.02.013
0749-6419/© 2015 Elsevier Ltd. All rights reserved.
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 63

Currently, notched specimens are used to develop different levels of localized hydrostatic stress, leading, consequently, to
different stress triaxialities - see, for example, Hancock and Mackenzie (1976), Alves and Jones (1999), Wilson (2002), Bao and
Wierzbicki (2004), Driemeier et al. (2010). Hancock and Mackenzie (1976) investigated the relationship between ductility and
stress triaxiality for three types of steel, testing axisymmetric specimens with different notches. Accordingly, the authors
concluded that the ductility decreased with increasing level of stress triaxiality for all materials studied. As a consequence, the
onset of failure is at highest stress triaxiality locus, as shown in Fig. 1 in a typical tensile test in steel.
Bao and Wierzbicki (2004) conducted a pioneering extensive experimental program in 2024-T351 aluminum with eleven
distinct specimen geometry. They covered a wide range of stress triaxialities, from negative levels (0.3) to values close to
unity (0.95). The results suggested that the equivalent fracture strain versus stress triaxiality is not a monotonic curve. Ac-
cording to the authors, there are different fracture mechanisms at different stress triaxiality ranges, as illustrated in Fig. 2. For
negative and low positive stress triaxialities, shear mechanisms govern damage evolution and failure, while for large positive
triaxialities, growth and coalescence of voids is dominant. At stress triaxialities between these two regimes, damage and
failure occur due to a combination of shear and void growth mechanisms. Later, Bao and Wierzbicki (2005) proposed a cut-off
value of 1/3 in the stress triaxiality for fracture. In contrast, Khan and Li (2012) found fracture surfaces along maximum
shear direction in a biaxial non proportional compression test at a stress triaxiality level of 0.495. Wierzbicki et al. (2005b)
present a complete review of seven fracture models, including the experimental tests necessary for the material parameter
calibration. Several authors developed experimental techniques and specimen geometries to investigate the onset of fracture
at large ranges of stress triaxialities. Mohr and Henn (2007), Mohr and Oswald (2008) and Driemeier et al. (2010) analyzed
positive low to intermediate stress triaxiality, while Haltom et al. (2013) studied specimens subjected to shear-dominant load.
The failure fractures at negative stress triaxiality matches the idea that hydrostatic pressure contributes to the closing of
voids, while high positive stress triaxiality cooperates with nucleation and growth of voids and material weakness. However,
in the range of low positive stress triaxiality, the literature agrees that there is another parameter that plays an important role
in controlling ductile fracture.
It can be shown that a given stress field can be uniquely expressed by the stress triaxiality, hydrostatic stress and the Lode
parameter, defined via the third invariant of the deviatoric stress state. Wilkins et al. (1980) were the first to introduce the
effect of the Lode parameter in a ductile failure model. According to Argon and Im (1975) and Wilkins et al. (1980) two factors
are responsible for the damage evolution: hydrostatic tensile stresses and asymmetric deformation. The first is responsible for
void growth and the second factor takes into account the experimental observations that final elongation before failure
decreases with increasing shear forces (Wilkins et al., 1980). In fact, Kim et al. (2007) and Gao and Kim (2006) remarked that
different stress states with the same stress triaxiality ratio level lead to different void growth and coalescence behavior.
Barsoum et al. (August 2007b) proposed a new specimen, a double notched tube, loaded in combined tensile and torsional
loading at fixed ratio, controlling stress triaxiality. The authors concluded that stress triaxiality is not a sufficient parameter to
characterize ductility, especially at low levels.
Nowadays, the Lode angle and the stress triaxiality are investigated by many authors, as it is considered that they control
the behavior of ductile failure (Cazacu and Barlat, 2004; Bai and Wierzbicki, 2008; Brünig et al., 2008; Gao et al., 2009; Mirone
and Corallo, 2010; Barsoum and Faleskog, 2011; Malcher et al., 2012). Gao et al. (2009) studied, numerically and experi-
mentally, the influence of stress triaxiality and Lode angle on the hardening evolution and failure process of flat and axis-
symmetric specimens of the aluminum alloy 5083. The authors suggested that the Lode angle plays a minor role on the
damage process, but significantly affects the plastic flow. The stress triaxiality has the opposite effect, it has negligible in-
fluence on the hardening evolution, but it is a significant parameter on the failure strain of the studied material. Mirone and
Corallo (2010) tested different ductile materials and arrived to similar conclusions. It is well established in the recent

Fig. 1. Failure initiation, in the center of the necking.


64 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 2. Failure mechanisms according to stress triaxiality level. Experimental results extracted from Bao and Wierzbicki (2004).

literature, however, that failure can occur by two very different mechanisms at high and low stress triaxiality; the effect of the
Lode parameter plays a significant role for moderate or negative triaxialities and it is marginal in the high stress triaxiality
regime. The subject is deeply studied in Brünig et al. (2013), for example. The authors numerically studied void growth and
covered a wide range of stress triaxialities and Lode parameters.
Bai and Wierzbicki (2008) postulated a general metal plasticity model, considering both the pressure sensitivity and the
Lode dependence. The authors proposed a three-dimensional fracture surface for aluminum, stress triaxiality vs Lode vs
deformation fracture, extrapolating the bidimensional curve by Bao and Wierzbicki (2004). In fact, the requirement of a three-
dimensional view of fracture behaviour can explain the apparently different behavior found by Beese et al. (2010) and Haltom
et al. (2013) for the same material, aluminum alloy 6061-T6. In tests conducted by Haltom et al. (2013) the bidimensional
curve stress triaxiality vs equivalent strain to fracture decreases monotonically with increasing stress triaxiality, while Beese
et al. (2010) found the behaviour stated by Bao and Wierzbicki (2004) of three different branches. According to Haltom et al.
(2013), these differences may be due to differences in the materials analyzed, including the effect of processing, but possibly
also due to differences in test types and measurements of stresses and strains. However, the Lode parameter is not discussed
by the authors. The results should be shown in a fracture surface, not a two-dimensional curve, so that overlapping points in
the curves were, in fact, at different Lode parameter coordinates. Recent work from Erice et al. (2014) shows the three
dimensional surface equivalent plastic strain to failure-stress triaxiality-Lode angle for Inconel 718 nickel-base superalloy.
Also Lou et al. (2014) present a star-shape Lode dependence in the proposed ductile fracture criterion. However, the loss of
convexity is not studied in terms of maximum-dissipation postulate and normality rule.
Indeed, attention has been given to experimental tests to characterize the material. The Arcan specimen (Arcan et al., 1978)
was designed for fracture in plane stress condition, of mode I, mode II and mixed mode. Different modes are obtained with
different loading angles, using a tensile testing machine. Along the years, many researchers have used Arcan's original or
modified specimen (Voloshin and Arcan, 1980; Bao and Wierzbicki, 2004; Wierzbicki et al., 2005a; Mohr and Henn, 2007;
Cognard et al., 2011; Dunand and Mohr, 2010). Driemeier et al. (2010) proposed a modified Arcan specimen that can be
tested at high or low triaxialities, depending the direction the specimen is loaded. Brünig et al. (January 2014a) tested the
specimen simultaneous loading in the two perpendicular directions. The authors argue that by changing the relationship
between the two applied forces they can cover, in a more continuous way, the range of stress triaxiality shown in Fig. 2.
Barsoum (2007a) and later Haltom et al. (2013) used notched tubes for combined tensile and shear tests. The major disad-
vantage of these tests in comparison with those proposed by Brünig et al. (January 2014a) is the tube-shaped specimen, since
industry deals mostly with material in thin plates.
A more detailed study on influence of the Lode angle and stress triaxiality is hampered both by the mandatory use of a
numerical-experimental hybrid approach and the fact that it is not possible to keep neither the stress triaxiality nor the Lode
angle constant during analysis. As already discussed, the localization is the precursor for the ultimate failure and, therefore,
the stress state becomes three-dimensional and stressestrain curve can no longer be estimated using analytical uniaxial
formulation. Additionally, local pressure values and Lode parameter are important for one to obtain conclusive results in the
study of failure. In other words, the finite element analysis complements the experimental test for the identification of stress
and strain state at failure. The real problem with this approach, described by Dunand and Mohr (2010), is the risk of adding
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 65

the errors of both the experiment and numerical simulation. These errors can be minimized by using modern experimental
techniques, such as digital image correlation (DIC), as presented, for example, in Mohr and Henn (2007), Luo et al. (2012),
Daiyan et al. (November 2012) and in the present work.
From the numerical point of view, efficient constitutive models of damage and plasticity or poroplasticity are required for
accurate modeling of material behavior, since it is the damage variable or porosity that progressively affects the strength and
stiffness of the material until failure. These models are said to be coupled. Classical examples are damage model from Lemaitre
(Lemaitre, 1985a; Lemaitre, 1985b) and Gurson-Tvergaard-Needleman (GTN) poroplasticity model originaly proposed by
Gurson (January 1977) and modified by Tvergaard (1982), Tvergaard and Needleman (1984). Both models are continuously
discussed and enhanced by researchers. Models for the evolution of the damage variable have been proposed in the literature,
considering the influence of parameters such as stress triaxiality and, more recently, Lode parameter (Xue and Wierzbicki,
2008; Basaran, 2011; Brünig et al., 2014b). It is already stated that the damage variable becomes anisotropic after a certain
level of deformation. Second order damage tensors are proposed by (Chaboche, 1990; Voyiadjis and Kattan, 1992; Brünig,
2003; Bonora et al., 2005). Particularly, several other authors have, over the past decades, including to the original Gurson
or GTN model, effects such as nucleation of voids (Chu and Needleman, 1980), consideration of non-spherical voids (Gologanu
et al., 1994), shear (Nahshon and Hutchinson, 2008; Xue, 2008; Malcher et al., 2014), void size (Monchiet and Bonnet, 2013),
among others.
Coupled models are complex and computationally expensive. They require the identification of more material constants,
whose physical significance and independence are not always proved and, furthermore, aspects such as convergence and
localization due to softening should be considered. For these reasons, despite the importance of the process of damage, many
models ignore this step, and failure modelling occurs without loss of strength and stiffness. In this case, the damage variable is
uncoupled, do not necessarily have representative elementary volume and defines the rupture without modifying the ma-
terial constitutive law. The term sudden failure is used to emphasize that failure is controlled by an external variable that,
when it reaches certain critical value, represents an abrupt decrease in the local resistance of the material. Many uncoupled
models are available in the literature since the pioneer one proposed by Freudenthal (1950), who postulated that failure
occurs when the total work reaches a critical value. Sudden failure criteria are widely applied in industry and the performance
of the most known will be evaluated here.
The objective of the current work is to address all these issues by an indepth analysis of the performance of some well
known uncoupled or sudden failure criteria. The analyses will be made using an ad hoc specimen, designed to generate two
distinct failure modes, hence the name bifailure. The first failure occurs at high stress triaxiality, and the second occurs in a
state of almost pure shear, ie, at low stress triaxiality. These failure modes occur in a simple tensile loading and they generate a
complex loading history. The ability of the here studied failure criteria of predicting the loadedisplacement pattern is seen as
a limitation or as a strength of the given studied models.

2. Experimental aspects

2.1. Material

The base material used here is the aluminum alloy 2024-T351 delivered in a 9.72 mm thick plate. There are extensive
previous experimental and numerical studies on failure models using the same material (Wierzbicki et al., 2005b; Xue, 2007;
Malcher et al., 2012; Morales, 2013). Fig. 3a shows the load-engineering strain curve as measured in the tensile test performed
in an Instron machine model 3369 with load capacity of 50 kN. The flat tensile specimen, as illustrated in Fig. 3b, was

Fig. 3. Experimental tests for material characterization. Specimen thickness of 9.72 mm.
66 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

submitted to imposed displacement rate of 0.005 mm/s. Load and displacements were measured, respectively, by the cali-
brated machine load cell and by clip gauges with length of 50 mm, fixed symmetrically to the center of the specimens.
An exponential law will govern the strain hardening evolution in the numerical analyses. It is defined as,
 p n
sy ðεp Þ ¼ sy0 þ H εn (1)

where sy ðεp Þ and εp are, respectively, the updated yield stress and equivalent plastic strain; sy0, H, n are positive material
constants. Table 1 resumes the elastoplastic material constants for aluminum alloy 2024-T351.

2.2. Bifailure specimen

2.2.1. Geometry
The Bifailure geometry is depicted in Fig. 4. It can be noticed that the specimen shows three critical regions of failure that
arise when it is pulled: two of high triaxiality around the side notches (n. 1 in Fig. 4a), and one of low triaxiality, around the
central notch (n. 2 in Fig. 4a). On both main planes of the specimen, i.e. front and back, small notches are machined as
indicated in Fig. 4b. This concentrates further the stresses, increasing the triaxiality in the side notches and cooperating with a
more pure shear stress state in the middle.
Although not the case here, it is possible to have the side notches machined with different radii, so leading to a non-
symmetric state and to a different fracture sequence as the one here obtained.

2.2.2. Bifailure test


The bifailure specimen was tested using standard tensile 50 kN Instron machine, with a crosshead velocity of 0.1 mm/min.
The specimen was painted in white color and sprinkled with toner powder to create the random pattern for DIC analysis.
The specimen is connected to the tensile machine via rods passing through holes located at the extremes of the specimens.
Any lateral motion of the specimen is accommodated by this type of fixture. Such a motion may take place if one of the
notched sides breaks first than the other, if different notch radii were machined.
Three computers received the images from two synchronized cameras and the load cell. A Visual Instrumentation Cor-
poration illumination set, model 900445, was used with 2 led arrays (4 rows by 6 columns) for general illumination of the
specimen (Fig. 5a). The whole illumination set-up provides 12000 lumens without generating heat, allowing a good quality
for the pictures for subsequent digital image correlation analyses.
Both cameras are of the same model, Nikon D90, with 12.3 megapixels and capable of ISOs from 100 to 3200 and shutter
speeds between 1/4000s and 30s. In order to obtain as many details as possible around the high triaxiality notch (Fig. 5b), it
was used a macro lens, model vivitar series 1L 105 mm f.2.5 macro (Nikon f-mount), to capture the side view images. For
capturing images of the front view (Fig. 5c), it was used an AF-S DX NIKKOR lens (18e105 mm f/3.5-5.6G ED VR). Fig. 6 shows
the images from the camera positioned to capture the frontal test images.
Although the geometry of the Bifailure specimen is three-dimensional, failure occurs in a plane perpendicular to the
camera. It is clear that failure proceeds in a local fashion with little, if some, out of plane deformation. Therefore, it is not
necessary the use of a three-dimensional DIC.
Fig. 7 shows the force vs displacement curve obtained in two tests performed with the Bifailure specimen. It can be easily
noticed the moment of failure at the side notches, u1 z 1.38 mm, and at the central notch u2 z 5.56 mm. The tests also put in
evidence the repeatability of the response and the simultaneous failure of side notches. Fig. 8 shows the specimen after test.
Due to the asymmetry in the geometry, the specimen had undergone side plastic deformation while being subjected to near
pure shear state at the central notch.

2.3. Image analyses

2.3.1. Digital image correlation


The 7D software, used for the DIC analysis, calculates de Green Strain. A Matlab software script was created to calculate the
equivalent strain for each mesh element. Fig. 9 shows the experimental curve and equivalent strain results obtained by
processing the images. The values are an average from the elements highlighted in the illustrations in the graph. It is noticed
that no significant deformation takes place in the central part during the initial loading phase. The deformations therein only
grow after lateral failure.
Figs. 10e11 show the profile of the equivalent strain. Fig. 10 shows the side failure, where the deformation grows from the
side to the center, despite failure starts in the center. Fig. 11 shows central failure. In this case, due to the movement of the

Table 1
Material properties for aluminum alloy 2024-T351.

E (MPa) n r (ton/mm3) syo(MPa) H (MPa) n


74660 0.33 2.7E9 352 440 0.42
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 67

Fig. 4. Bifailure geometry, in mm.

Fig. 5. Pictures show test setup and the angle view of the speckle-painted specimens from the two cameras positioned to capture images for the DIC analyses.
68 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 6. Test evolution pictures obtained with the frontal camera.

toner powder after the first failure, the analysis is troublesome. However, the deformation profile is very similar to that
obtained numerically, as it will be discussed later.

2.3.2. Failure surface


In addition to the phenomenological analyses outlined in last section, a study of the fracture surfaces was performed. The
images were obtained using traditional techniques and amplification magnifier (Figs. 12e13) and by scanning electron mi-
croscopy or SEM (Fig. 14).
With higher magnifications compared to Figs. 14 and 15a,b indicate the presence of different ductile failure mechanisms in
the material at the side and central regions, respectively. Fig. 15a illustrates the failure at high stress triaxiality, characterized
by the nucleation, growth and coalescence of voids by internal necking (presence of dimples). On the other hand, Fig. 15b
shows a completely different aspect, without the presence of voids observed in the previous case. Here, the failure occurs by
internal shear mechanism at low levels of stress triaxiality. Barsoum (2007a) present similar experimental results.

3. Numerical aspects

3.1. Failure models

Coupled models are not only more complex and computationally expensive, but require the identification of more material
constants. This is why, in many practical applications, uncoupled models are still used. In Børvik et al. (2009), for example, the
authors use CockcroftLatham and JohnsoneCook fracture criteria to model the complex phenomenon of steel plates perfo-
ration at high strain rates. Five uncoupled models were chosen in order to study their performance based on the Bifailure
response. They are the maximum plastic deformation, Cockroft-Latham and JohnsoneCook traditional models used in the
industry; Wilkins is the pioneering model in considering the influence of stress field asymmetry, reformulated later and
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 69

Fig. 7. Force vs displacement curve for the Bifailure test.

Fig. 8. Specimen after test.

Fig. 9. Curves force and equivalent strain vs displacement. Displacement is imposed on the upper end.
70 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 10. Profile of the equivalent strain by digital image processing superimposed on the experimental test, for failure at side notches. Figs. 1e6 were obtained in
instants related to the following imposed displacement levels 0; 0.276; 0.552; 0.828; 1.104; 1.38 mm, respectively.

incorporated in the model of Xue-Wierzbicki. The models were implemented uncoupled to von Mises model with isotropic
hardening, as detailed in Section 2.1.
Table 2 shows model and material parameters of the sudden failure criteria adopted. Regarding the parameters, those for
equivalent plastic strain, Cockroft-Latham, Wilkins and Xue-Wierzbicki criteria were obtained from Wierzbicki et al. (2005b),
and for JohnsoneCook were taken from Morales (2013). Here, simplified JohnsoneCook model is adopted since strain rate
and temperature effect were disregarded.

3.1.1. Maximum accumulated equivalent plastic strain (MaxEPS)


The simplest criteria assumes that fracture occurs when the equivalent plastic strain εp reaches a critical value εpf , where
the equivalent plastic strain ε_ p is defined as the integral of the equivalent plastic strain rate,

Zt
p
p
ε ¼ ε_ dt; (2)
0

with the equivalent plastic strain rate defined in terms of the plastic strain rate tensor ε_ p,
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 71

Fig. 11. Profile of the equivalent strain by digital image processing superimposed on the experimental test, for failure at central notch. Figs. 1e4 were obtained in
instants related to the following imposed displacement levels 3.0; 4.25; 5.5 mm, respectively, and Fig. 5 shows the end of the test.

Fig. 12. Fracture surface of the specimen in the region of high triaxiality at different magnifications.

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p 2 p p
ε_ ¼ ε_ : ε_ (3)
3

Although most important commercial codes present this option as failure criteria, it is an old criterion, and lacks the
generality necessary for nonlinear analyses.
72 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 13. Fracture surfaces of the specimen bifailure in the region of low triaxiality at different magnifications.

3.1.2. CockcrofteLatham (CeL)


Cockcroft and Latham (1968) suggest that fracture takes place when the plastic work associated with the maximum
principal stress s1,

εp
Zf
p
w¼ s1 dεp  wf for s1 > 0 (4)
0
reaches a threshold value, wpf . Thus, the authors essentially assumes that the defect growth is controlled by the maximum
p
principal stress. The critical value wf can be easily determined from a tensile test. However, it should be noted that different
tests, such as torsion or even tension test with different notches in the specimen, result in different wpf . In a study by Dey et al.
(Dey et al., 2007) the authors concluded that the C-L criterion of a single parameter gives as satisfactory results as the
JohnsoneCook criterion with its five parameters in the numerical simulations of steel plates perforation. Gruben et al.
(Gruben et al., 2012) extended the model to consider the effect of the Lode parameter and obtained good results in the
analysis of steel failure at different triaxial stress states.

3.1.3. Wilkins
Wilkins et al. (Wilkins et al., 1980) proposed a failure criterion based on the accumulation of plastic deformation weighted by
two independent functions, u1 and u2, accounting for hydrostatic and deviatoric load components, respectively. According to the
criterion proposed by the authors, failure occurs when the cumulative damage variable D reaches the critical value Dc. Thus,

εp
Zf
D¼ u1 u2 dεp  Dc (5)
0

where dεp is the equivalent plastic strain, u1 and u2 are respectively,

1
u1 ¼ ;
ð1  apÞa (6)
u2 ¼ ð2  AÞb

where
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 73

Fig. 14. Sections of failure analyzed with SEM: high (top) and low (bottom) triaxiality.

Fig. 15. Failure surface.

Table 2
Sudden failure models tested with Bifailure specimen.

Model Equation Parameters


Maximum Equivalent Plastic Strain Z εp p
εf ¼ 0:15  0:60
f
p
(dated from 1966) εp ¼ dεp  εf
0
Cockcroft and Latham (1968) Z εpf
p
wf ¼ 0:058  0:485
p
w¼ s1 dεp  wf
0
Wilkins et al. (1980) Z εp a ¼ 1.2E3 MPa1 a ¼ 2.15
f
D¼ u1 u2 dεp  Dc
0
1
u1 ¼ ð1þapÞa u2 ¼ ð2  AÞb b¼2.18 Dc¼0.93
Johnson and Cook (1985) ¼ C1 þ C2 exp ðC3 hÞ  εpf
εp C1 ¼ 0.13 C2 ¼ 0.13 C3¼1.5
Wierzbicki and Xue (February 2005) Z εp p D1 ¼ 0.87 D2 ¼ 1.77
f dε
D¼ <1
0 εf
εf ¼ D1 eD2 h  ðD1 eD2 h  D3 eD4 h Þð1  x1=n Þn D3 ¼ 0.21 D4 ¼ 0.01 n ¼ 0.153
74 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 16. Wilkins model.

 
s2 s2
A ¼ max ; (7)
s1 s3

s1 > s2 > s3 are the principal deviatoric stresses, p is the hydrostatic pressure. The material constants a, a and b are obtained
from experimental tests that ideally should cover a wide range, from pure shear to purely hydrostatic stress states.

Fig. 17. Influence of parameter b in the shape of deviatopric part of the Wilkins model.
L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 75

Fig. 18. 3D and 2D views of Wilkins model, in the plane of principal stresses.

Fig. 19. FE mesh for bifailure virtual analyses.


76 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 16a,b show, respectively, the influence of Lode angle and of hydrostatic pressure on fracture. For high pressures, the
limit of fracture tends to infinity, while for high hydrostatic tension the limit tends to zero, precluding the occurrence of
plastification. Fig. 17 shows the influence of the parameter b in bloom geometry shape in the Pi plane. For b ¼ 0, the circular
profile in deviatoric plane is similar to the von Mises criterion, whose radius is defined by u2. Fig. 18 shows the envelope of
fracture in tri-and bi-dimensional plane. Particularly, the two-dimensional shape is presented for values of b ¼ 1.8 and
b ¼ 0.

3.1.4. JohnsoneCook (JeC)


The criterion proposed by Johnson and Cook (1985) postulates that fracture occurs when the accumulated plastic strain,
calculated as a function of three independent variables - triaxiality, temperature and rate of plastic deformation,
p
εp ¼ up ðhÞuε_ ð_εp Þuq ðTÞ  εf (8)

reaches its critical value εpf . In Equation (8),

up ¼ C1 þ C2 exp ðC3 hÞ
ε_ p
uε_ ¼ 1 þ C4 log
ε_ 0 (9)
T  T0
uq ¼ 1 þ C5
Tm  T0

where h is the triaxiality; Ci are material parameters, ε_ p is the plastic strain rate and ε_ 0 is the reference rate; T is the tem-
perature; T0 and Tm are, respectively, the reference and melting temperature of the material. As the model has independent
contribution of each variable, the parameters C1,C2,C3 can be obtained from tests at constant temperature and deformation
rate and the parameters C4 and C5 can be obtained with tensile tests varying strain rate and temperature, respectively.
Johnson and Holmquist (Johnson and Holmquist, 1989) present the characterization of the model for different materials,
which initially helped to leverage the use of the model.

3.1.5. XueeWierzbicki (XeW)


According to Wierzbicki and Xue (February 2005), the general equation for equivalent plastic strain at fracture is given by,

  n
εf ¼ D1 eD2 h  D1 eD2 h  D3 eD4 h 1  x1=n
p
(10)

where h is the triaxiality; n is the hardening material parameter and x is the Lode parameter, defined as,

Fig. 20. Force vs displacement for MaxEPS criterion.


L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 77

Fig. 21. Equivalent plastic strain evolution for side notch.


78 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

pffiffiffi
4J3 3 3 J3 27 J3
x¼ ¼ ¼
r3 2 J 3=2
2
2 q3 (11)
pffiffiffiffiffiffiffi
q ¼ 3J2

where J2 and J3 are the invariants of deviatoric stress tensor, q is the von Mises equivalent stress.
Expression (10) defines a fracture surface (Wierzbicki and Xue, February 2005; Wierzbicki et al., 2005b). As the failure
criterion, the authors assume an incremental relationship for the damage indicator D

Fig. 22. Equivalent plastic strain evolution for central notch.


L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 79

Fig. 23. Triaxiality evolution in side and central notches.

εp
Zf
dεp
D¼ 1 (12)
εpf
0

p
and equivalent plastic strain, εf .
A subroutine was implemented in LS-Dyna® code with the von Mises elastoplastic model with exponential hardening law
and options for the various sudden failure criteria here used.

3.2. Numerical results

The virtual model of the Bifailure has 167989 nodes and it is discretized into 152228 solid elements, with 30 elements in
the thickness of the more refined region, Fig. 19. Different meshes were tested in order to guarantee mesh independence
(Morales, 2013).
p
The overall result obtained with the MaxEPS criterion is shown in Fig. 20, for εf ¼ 0:21. This limit value, suggested in
Wierzbicki et al. (2005b), refers to pure shear test. Traditionally, this value is taken from a uniaxial tensile test with unnotched

Fig. 24. Force vs displacement, for C-L criterion.


80 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 25. Force vs displacement, for Wilkins criterion.

p
specimen (corresponding, to present material, the εf ¼ 0:45). Nevertheless, the error in the response at low triaxiality is
considerable. In the figure, it is noticed instability after the first failure. The same behavior is observed in all simulations. This
instability occurs because, in the numerical simulation with uncoupled models, fracture is represented by eliminating the
finite element which is under high stress state and that achieved, at any Gauss point, the limit value imposed by the sudden

Fig. 26. Critical value for failure according to Wilkins Model.


L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 81

Fig. 27. Profile for triaxiality (left) and Lode parameter (right).

failure criterion. In coupled models this instability is expected to be less important, since the material strength decreases with
increase of the damage and the eliminated finite element should have low rigidity.
Fig. 21 shows the evolution of equivalent plastic strain in the side notch. It can be verified, due to the dynamic caption in
the figure, that although plastic deformation begins in the outer surface, the values grow faster in the center. Since experi-
mental evidences have shown that fracture locus is in the center of the specimen, the model predicts the onset of failure in the
correct location. As discussed in Brünig et al. (2008) and Malcher et al. (2012), less smooth notches tend to concentrate the
plastic deformation at the lateral sides. Thus, it would be interesting to examine the locus transition from new side notch
dimensions for the specimen.
Fig. 22 shows the evolution of equivalent plastic strain in the central notch. The region is free of plastic deformation until
the occurrence of the first failure. The failure, according to the MaxEPS criterion, starts on the surface of the shear zone. As
shown in Fig. 11 (also in agreement with Bai and Wierzbicki (2008)), it is believed that the crack initiation occurs in this
region. Again, changes in the notch shape would dislocate the locus to the center of the shear zone (Malcher et al., 2012).
As observed in experiments and in the numerical simulations, there is a side plastic deformation due to rotation of the
specimen during the evolution of the test.
Fig. 23 shows the numerical result of triaxiality during the test. Accordingly, the first failure occurs under a state of high
triaxiality, above 0.4, while the second failure takes place at a value below 0.1. Looking at the experimental points in Fig. 2, one
see that these values are in different regions and governed by different failure mechanisms.
Fig. 24 shows the results obtained using limit values of the C-L criterion, extracted from literature (Wierzbicki et al.,
2005b). The results are unsatisfactory, suggesting the existence of an intermediate value that brings the results closer to
the experimental values. This suggests that the proposed specimen could be used in the characterization of criteria that
depend on a small number of parameters, such as C-L and MaxEPS.

Fig. 28. Force vs displacement, for J-C criterion.


82 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 29. Critical value for failure according to J-C criterion.

Fig. 30. Force vs displacement, for XeW and JeC criteria.


L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 83

Wilkins and J-C model results are indicated by, respectively, Figs. 25e26 and 28e29.
Particularly for the Wilkins model, side notch failure occurred simultaneously in two regions, separating the specimen into
three parts; and the second failure is displaced from the notch center. As the model considers the triaxiality and Lode angle, the
parameters are plotted in Fig. 27. It is noticed that there is equivalence between the maximum triaxiality and rupture of the side
notch section and between the region where the Lode parameter is higher and the rupture at low triaxiality in central notch.

Fig. 31. Evolution of damage, according to XeW criterion.


84 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Fig. 32. Evolution of equivalent plastic strain to failure, in the side and central notches, according to JeC and XeW criteria.

For J-C criterion, damage variable correctly represent failure loci, as illustrated in Fig. 29, but its evolution is slow in the
shear zone, and failure occurs in the central notch at higher displacement level when compared to experimental result.
Figs. 30e32 show the results obtained with XeW model. Fig. 30, where the graph also shows the JC model, points to an
interesting aspect of the models. Given that the same formulation for plasticity was adopted and the failure criterion is just a
post processing, the models led to very similar results.
Fig. 32 shows the evolution of equivalent plastic strain for XeW and JeC criteria. It is noticed that, in comparison with
Fig. 9, the evolution of deformation in the central notch is much slower in the numerical models, as a consequence of slow
damage progress at low triaxiality levels.

4. Conclusions

Failure criteria consisting of the posteriori evaluation of scalar functions of state variables, here called sudden failure
criteria, are simple, easy to implement, do not require a large number of parameters and therefore widely used in practice. On
the other hand, ductile materials may undergo large deformations before rupture, with significant changes in the material
microstructure and should not be ignored. The number of scientific articles that discuss the microstructure, the return to GTN
model, and many new micromechanical models, illustrate this need. Thus, constitutive models coupled with representative
volume element to consider the presence of voids, whether specific or diffuse, are the most consistent way to model ductile
materials. The confrontation between Bifailure experimental and numerical results obtained from an implemented consti-
tutive traditional von Mises plastic model uncoupled with different failure criteria - MaxDPE, CL, Wilkins, JC, XW - corrob-
orates these conclusions. Criteria for abrupt failure using few or even only one variable, such as the criteria for Cokroft-Latham
or MaxDPE, are widely used in industry. Simulations with these criteria must be made with care, noting that different
characterization tests lead to quite different parameters.
An efficient criterion has to have a good performance in a limited range of applicability using not so many parameters, wich
should have physical meaning. Models with a large amount of parameters can in the limit model any phenomenon! In short, a
model formulated to perform well in all loading conditions and for many different materials would be closer to the caos than to the
efficiency.
Digital Image Correlation and Scanning Electron Microscopy were explored. The results obtained through the correlation
of digital image are very important in failure studies, whith numerical and experimental results merging and complementing
each other. Parameters such as, for example, equivalent plastic strain, stress triaxiality and Lode are numerically obtained and
set in the experimental curve of the material, together with the equivalent strain to fracture. Scanning Electron Microscopy, on
the other side, revealead different aspects of failure surface at high and low stress triaxiality.
Bifailure specimen, whose geometry is detailed in this work, is a promising specimen to assist in the evaluation of failure
criteria performance. It can also be used, through inverse process, to characterize one- or multi-parameter criteria in an
efficient way.

Appendix A. Supplementary data

Supplementary data related to this article can be found at http://dx.doi.org/10.1016/j.ijplas.2015.02.013.


L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86 85

References

Alves, M., Jones, N., 1999. Influence of hydrostatic stress on failure of axisymmetric notched specimens. J. Mech. Phys. Solids 47, 643e667.
Arcan, M., Hashin, A., Voloshin, A., 1978. A method to produce uniform plane stress states with application to fiber - reinforced materials. Exp. Mech. 18 (4),
141e146.
Argon, A.S., Im, J., 1975. Separation of second phase particles in spheroidized 1045 steel, Cu - 0.6 % Cr alloy, and maraging steel in plastic straining. Metall.
Trans. 6A, 839e851.
Bai, Y., Wierzbicki, T., 2008. A new model of metal plasticity and fracture with pressure and lode dependence. Int. J. Plast. 24 (6), 1071e1096.
Bao, Y., Wierzbicki, T., 2004. On the fracture locus in the equivalent strain and stress triaxiality space. Int. J. Mech. Sci. 46, 81e98.
Bao, Y., Wierzbicki, T., 2005. On the cut-off value of negative triaxiality for fracture. Eng. Fract. Mech. 72, 1049e1069.
Barlat, F., July 1987. Crystallographic texture, anisotropic yield surfaces and forming limits of sheet metals. Mater. Sci. Eng. 91, 55e72.
Barlat, F., Yoon, J., Cazacu, O., 2007. On linear transformations of stress tensors for the description of plastic anisotropy. Int. J. Plast. 23, 876e896.
Barsoum, Faleskog, 2007a. Rupture mechanisms in combined tension and shear - experiments. Int. J. Solids Struct. 44, 1768e1786.
Barsoum, I., Faleskog, J., August 2007b. Rupture mechanisms in combined tension and shear - micromechanics. Int. J. Solids Struct. 44 (17), 5481e5498.
Barsoum, I., Faleskog, J., 2011. Micromechanical analysis on the influence of the lode parameter on void growth and coalescence. Int. J. Solids Struct. 48 (6),
925e938.
Basaran, M., 2011. Stress State Dependent Damage Modeling with a Focus on the Lode Angle Influence. Ph.D. thesis. Rheinisch-Westflischen Technischen
Hochschule Aachen.
Beese, A.M., Luo, M., Li, Y., Bai, Y., Wierzbicki, T., 2010. Partially coupled anisotropic fracture model for aluminum sheets. Eng. Fract. Mech. 77 (7), 1128e1152.
Bonora, N., Gentile, D., Pirondi, A., Newaz, G., 2005. Ductile damage evolution under triaxial state of stress: theory and experiments. Int. J. Plast. 21,
981e1007.
Børvik, T., Dey, S., Clausen, A., 2009. Perforation resistance of five different high-strength steel plates subjected to small-arms projectiles. Int. J. Impact Eng.
36 (7), 948e964.
Bridgman, P.W., 1952. Studies in Large Flow and Fracture. McGraw Hill, New York.
Brünig, M., 1999. Numerical simulation of the large elastic-plastic deformation behavior of hydrostatic stress-sensitive solids. Int. J. Plast. 15, 1237e1264.
Brünig, M., 2003. An anisotropic ductile damage model based on irreversible thermodynamics. Int. J. Plast. 19, 1679e1713.
Brünig, M., Chyra, O., Albrecht, D., Driemeier, L., Alves, M., 2008. A ductile damage criterion at various stress triaxialities. Int. J. Plast. 24, 1731e1755.
Brünig, M., Driemeier, L., 2007. Numerical simulation of taylor impact tests. Int. J. Plast. 23, 1979e2003.
Brünig, M., Gerke, S., Brenner, D., January 2014a. Stress-state-dependence of damage behavior: experiments and numerical simulations. In: International
Symposium on Plasticity and its Current Applications. Freeport, Bahamas.
Brünig, M., Gerke, S., Hagenbrock, V., 2013. Micro-mechanical studies on the effect of the stress triaxiality and the lode parameter on ductile damage. Int. J.
Plast. 50, 49e65.
Brünig, M., Gerke, S., Hagenbrock, V., 2014b. Stress-state-dependence of damage strain rate tensors caused by growth and coalescence of micro-defects. Int.
J. Plast. 63, 49e63.
Cazacu, O., Barlat, F., 2004. A criterion for description of anisotropy and yield differential effects in pressure-insensitive metals. Int. J. Plast. 20, 2027e2045.
Chaboche, J.L., 1990. On the description of damage induced anisotropy and active passive damage effect. In: Damage Mechanics in Engineering Materials,
pp. 153e166.
Chu, C.C., Needleman, A., 1980. Void nucleation effects in biaxially streched sheets. J. Eng. Mater. Technol. 102, 249e256.
Cockcroft, M.G., Latham, D.J., 1968. Ductility and the workability of metals. J. Inst. Metals 96, 33e39.
Cognard, J., Sohier, L., Davies, P., 2011. A modified arcan test to analyze the behavior of composites and their assemblies under out-of-plane loadings.
Compos. Part A: Appl. Sci. Manuf. 42 (1), 111e121.
Daiyan, H., Andreassen, E., Grytten, F., Osnes, H., Gaarder, R., November 2012. Shear testing of polypropylene materials analysed by digital image correlation
and numerical simulations. Exp. Mech. 52 (9), 1355e1369.
Dey, S., Brvik, T., Hopperstad, O., Langseth, M., 2007. On the influence of constitutive relation in projectile impact of steel plates. Int. J. Impact Eng. 34 (3),
464e486.
Driemeier, L., Brünig, M., Micheli, G.B., Alves, M., 2010. Experiments on stress-triaxiality dependence of aluminum alloys. Mech. Mater. 42 (2), 207e217.
Dunand, M., Mohr, D., 2010. Hybrid experimental-numerical analysis of basic ductile fracture experiments for sheet metals. Int. J. Solids Struct. 47 (9),
1130e1143.
Erice, B., Perez-Martín, M.J., G alvez, F., 2014. An experimental and numerical study of ductile failure under quasi-static and impact loadings of inconel 718
nickel-base superalloy. Int. J. Impact Eng. 69, 11e24.
Freudenthal, A.M., 1950. The Inelastic Behaviour of Engineering Materials and Structures. Jonhn Wiley & Sons, New York.
Gao, X., Kim, J., 2006. Modeling of ductile fracture: significance of void coalescence. Int. J. Solids Struct. 43, 6277e6293.
Gao, X., Zhang, T., Hayden, M., Roe, C., 2009. Effects of the stress state on plasticity and ductile failure of an aluminum 5083 alloy. Int. J. Plast. 25, 2366e2382.
Gologanu, M., Leblond, J., Devaux, J., 1994. Approximate models for ductile metals containing nonspherical voids - case of axisymmetric oblate ellipsoidal
cavities. J. Eng. Mater. Tech. 116, 290e297.
Gruben, G., Hopperstad, O.S., Brvik, T., 2012. Evaluation of uncoupled ductile fracture criteria for the dual-phase steel docol 600dl. Int. J. Mech. Sci. 62,
133e146.
Gurson, A.L., January 1977. Continuum theory of ductile rupture by void nucleation and growth: part I- yield criteria and flow rules for porous ductile media.
J. Eng. Mater. Technol. 99, 2e15.
Haltom, S., Kyriakides, S., Ravi-Chandar, K., 2013. Ductile failure under combined shear and tension. Int. J. Solids Struct. 50, 1507e1522.
Hancock, J.W., Mackenzie, A.C., 1976. On the mechanics of ductile failure in high-strength steels subjected to multi-axial stress-states. J. Mech. Phys. Solids
24, 147e169.
Hill, R., 1950. The Mathematical Theory of Plasticity. Oxford.
Johnson, G.R., Cook, W.H., 1985. Fracture characteristics of three metals subjected to various strains, strain rates, temperatures and pressures. Eng. Fract.
Mech. 21, 31e48.
Johnson, G.R., Holmquist, T.J., 1989. Test Data and Computational Strength and Fracture Model Constants for 23 Materials Subjected to Large Strain, High
Strain Rates, and High Temperature. Tech. rep., Los Alamos National Laboratory Technical Report LA-11463-MS.
Khan, A.S., Li, H., 2012. A new approach for ductile fracture prediction on Al 2024-T351 alloy. Int. J. Plast. 35, 1e12.
Kim, J., Zhang, G., Gao, X., 2007. Modeling of ductile fracture: application of the mechanism-based concepts. Int. J. Solids Struct. 44, 1844e1862.
Lemaitre, J., 1985a. A continuous damage mechanics model for ductile fracture. J. Eng. Mater. Technol. 107, 83e89.
Lemaitre, J., 1985b. Coupled elasto-plasticity and damage constitutive equations. Comput. Methods Appl. Mech. Eng. 51, 31e49.
Lou, Y., Yoon, J.W., Huh, H., 2014. Modeling of shear ductile fracture considering a changeable cut-off value for stress triaxiality. Int. J. Plast. 54, 56e80.
Luo, M., Dunand, M., Mohr, D., 2012. Experiments and modeling of anisotropic aluminum extrusions under multi-axial loading - part ii: ductile fracture. Int.
J. Plast. 32 (33), 36e58.
Malcher, L., Pires, F.A., de Sa, J.C., 2012. An assessment of isotropic constitutive models for ductile fracture under high and low stress triaxiality. Int. J. Plast.
30, 81e115.
Malcher, L., Pires, F.M.A., de Sa , J.M.A.C., 2014. An extended GTN model for ductile fracture under high and low stress triaxiality. Int. J. Plast. 54, 193e228.
Matsumoto, A.T., Driemeier, L., Alves, M., 2012. Performance of polymeric reinforcements in vehicle structures submitted to frontal impact. Int. J. Crash-
worthiness 17, 479e496.
86 L. Driemeier et al. / International Journal of Plasticity 71 (2015) 62e86

Mirone, G., Corallo, D., 2010. A local viewpoint for evaluating the influence of stress triaxiality and lode angle on ductile failure and hardening. Int. J. Plast. 26
(3), 348e371.
Mohr, D., Henn, S., 2007. Calibration of stress-triaxiality dependent crack formation criteria: a new hybrid experimental-numerical method. Exp. Mech. 47,
805e820.
Mohr, D., Oswald, M., 2008. A new experimental technique for the multi-axial testing of advanced high strength steels. Exp. Mech. 48, 65e77.
Monchiet, G., Bonnet, G., 2013. A Gurson-type model accounting for void size effects. Int. J. Solids Struct. 50 (2), 320e327.
Morales, E.D., 2013. Analysis of Failure Criteria for Ductile Materials: a Numerical and Experimental Study (in Portuguese). Master’s thesis. Sa ~o Paulo
University.
Nahshon, K., Hutchinson, J., 2008. Modification of the Gurson model for shear failure. Eur. J. Mech. A/Solids 27, 1e17.
Rice, J.R., Tracey, D.M., 1969. On the ductile enlargement of voids in triaxial stress fields. J. Mech. Phys. Solids 17, 201e217.
Tvergaard, V., 1982. Material failure by void coalescence in localized shear bands. Int. J. Solids Struct. 18 (8), 659e672.
Tvergaard, V., Needleman, A., 1984. Analysis of the cup-cone fracture in a round tensile bar. Acta Metall. 32 (1), 157e169.
Voloshin, A., Arcan, M., 1980. Pure shear moduli of unidirectional fibre-reinforced materials (FRM). Fibre Sci. Technol. 13, 125e134.
Voyiadjis, G.Z., Kattan, P.I., 1992. A plasticity-damage theory for large deformation of solids - I. theoretical formulation. Int. J. Eng. Sci. 9, 1089e1108.
Wierzbicki, T., Bao, Y., Bai, Y., 2005a. A new experimental technique for constructing a fracture envelope of metals under multi-axial loading. In: Proceedings
of SEM Annual Conference & Exposition on Experimental and Applied Mechanics.
Wierzbicki, T., Bao, Y., Lee, Y.-W., Bali, Y., 2005b. Calibration and evaluation of seven fracture models. Int. J. Mech. Sci. 43, 719e743.
Wierzbicki, T., Xue, L., February 2005. On the Effect of the Third Invaraint of the Stress Deviator on Ductile Fracture. Tech. Rep. 136. MIT Impact and
Crashworthiness.
Wilkins, M.L., Streit, R.D., Reaugh, J.E., 1980. Cumulative-strain-damage Model of Ductile Fracture: Simulation and Prediction of Engineering Fracture Test.
Tech. rep., UCRL- 53058. Lawrence Livermore National Laboratory.
Wilson, C.D., 2002. A critical reexamination of classical metal plasticity. J. Appl. Mech. Trans. ASME 69, 63e68.
Xue, L., 2007. Ductile Fracture Modeling - Theory, Experimental Investigation and Numerical Verification. Ph.D. thesis. Massachusetts Institute of
Technology.
Xue, L., 2008. Constitutive modeling of void shearing effect in ductile fracture of porous materials. Eng. Fract. Mech. 75, 3343e3366.
Xue, L., Wierzbicki, T., 2008. Ductile fracture initiation and propagation modeling using damage plasticity theory. Eng. Fract. Mech. 75, 3276e3293.

Potrebbero piacerti anche