ductile fracture experiments with locally proportional loading histories

27
Ductile fracture experiments with locally proportional loading histories Christian C. Roth a, c , Dirk Mohr a, b, c, * a Solid Mechanics Laboratory (CNRS-UMR 7649), Department of Mechanics, Ecole Polytechnique, Palaiseau, France b Impact and Crashworthiness Laboratory, Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA, USA c ETH Zurich, Department of Mechanical and Process Engineering, Zurich, Switzerland article info Article history: Received 19 July 2015 Available online 11 September 2015 Keywords: A. Ductility B. Finite strain C: Optimization Fracture experiments abstract Basic ductile fracture experiments for sheet metal (or at coupons extracted from bulk material) are presented to characterize the onset of fracture at different stress states. Special emphasis is placed on designing the experiments such that the stress triaxiality and the Lode angle parameter remain constant while the specimen is loaded all the way to fracture. A new in-plane specimen with two parallel gage sections is proposed to deter- mine the strain to fracture for approximately zero stress triaxiality. A FEA based meth- odology is shown to identify the optimal specimen geometry as a function of the material's ductility and strain hardening. A tension specimen with a central hole is investigated in detail with regard to determining the strain to fracture for uniaxial tension. It is found that the required hole-to-ligament width ratio decreases as a function of the material ductility and increases as a function of the strain hardening exponent. The bending of a wide strip is pursued to prevent the necking prior to fracture under plane strain tension conditions, while an Erichsen-type of punch test is used to characterize the material response for equi- biaxial tension. It is worth noting that the strain to fracture can be directly determined from surface strain measurements in the cases of shear, plane strain tension and equi- biaxial tension loading, thereby removing the need to perform nite element simula- tions for extracting the loading path to fracture. © 2015 Elsevier Ltd. All rights reserved. 1. Introduction There is a constant quest for reliable experimental data characterizing the effect of stress state on ductile fracture. Different stress states may be achieved through different initial specimen geometries or by applying different combinations of loading to the specimen boundaries. Examples for the rst approach are the works of Bao and Wierzbicki (2004), Brünig et al. (2008), Gao et al. (2010) or Driemeier et al. (2010). Example for the second approach are the tension-torsion experiments of Barsoum and Faleskog (2007a), Faleskog and Barsoum (2013), Haltom et al. (2013) and Papasidero et al. (2015), the internal pressure- tension testing of tubes (Kuwabara et al., 2005; Korkolis and Kyriakides, 2009), the tension-shear loading of buttery specimens (Wierzbicki et al., 2005; Mohr and Henn, 2007; Mae et al., 2007; Dunand and Mohr, 2011 , Abedini et al., 2015) and the biaxial loading of cruciform-like specimens (e.g. Abu-Farha et al., 2009; Brenner et al., 2014). * Corresponding author. ETH Zurich, Department of Mechanical and Process Engineering, Zurich, Switzerland. E-mail address: [email protected] (D. Mohr). Contents lists available at ScienceDirect International Journal of Plasticity journal homepage: www.elsevier.com/locate/ijplas http://dx.doi.org/10.1016/j.ijplas.2015.08.004 0749-6419/© 2015 Elsevier Ltd. All rights reserved. International Journal of Plasticity 79 (2016) 328e354

Upload: vonga

Post on 09-Dec-2016

224 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Ductile fracture experiments with locally proportional loading histories

International Journal of Plasticity 79 (2016) 328e354

Contents lists available at ScienceDirect

International Journal of Plasticity

journal homepage: www.elsevier .com/locate/ i jp las

Ductile fracture experiments with locally proportionalloading histories

Christian C. Roth a, c, Dirk Mohr a, b, c, *

a Solid Mechanics Laboratory (CNRS-UMR 7649), Department of Mechanics, �Ecole Polytechnique, Palaiseau, Franceb Impact and Crashworthiness Laboratory, Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge,MA, USAc ETH Zurich, Department of Mechanical and Process Engineering, Zurich, Switzerland

a r t i c l e i n f o

Article history:Received 19 July 2015Available online 11 September 2015

Keywords:A. DuctilityB. Finite strainC: OptimizationFracture experiments

* Corresponding author. ETH Zurich, DepartmentE-mail address: [email protected] (D. Mohr).

http://dx.doi.org/10.1016/j.ijplas.2015.08.0040749-6419/© 2015 Elsevier Ltd. All rights reserved.

a b s t r a c t

Basic ductile fracture experiments for sheet metal (or flat coupons extracted from bulkmaterial) are presented to characterize the onset of fracture at different stress states.Special emphasis is placed on designing the experiments such that the stress triaxialityand the Lode angle parameter remain constant while the specimen is loaded all the way tofracture. A new in-plane specimen with two parallel gage sections is proposed to deter-mine the strain to fracture for approximately zero stress triaxiality. A FEA based meth-odology is shown to identify the optimal specimen geometry as a function of the material'sductility and strain hardening. A tension specimen with a central hole is investigated indetail with regard to determining the strain to fracture for uniaxial tension. It is found thatthe required hole-to-ligament width ratio decreases as a function of the material ductilityand increases as a function of the strain hardening exponent. The bending of a wide strip ispursued to prevent the necking prior to fracture under plane strain tension conditions,while an Erichsen-type of punch test is used to characterize the material response for equi-biaxial tension. It is worth noting that the strain to fracture can be directly determinedfrom surface strain measurements in the cases of shear, plane strain tension and equi-biaxial tension loading, thereby removing the need to perform finite element simula-tions for extracting the loading path to fracture.

© 2015 Elsevier Ltd. All rights reserved.

1. Introduction

There is a constant quest for reliable experimental data characterizing the effect of stress state on ductile fracture. Differentstress states may be achieved through different initial specimen geometries or by applying different combinations of loadingto the specimen boundaries. Examples for the first approach are the works of Bao andWierzbicki (2004), Brünig et al. (2008),Gao et al. (2010) or Driemeier et al. (2010). Example for the second approach are the tension-torsion experiments of Barsoumand Faleskog (2007a), Faleskog and Barsoum (2013), Haltom et al. (2013) and Papasidero et al. (2015), the internal pressure-tension testing of tubes (Kuwabara et al., 2005; Korkolis and Kyriakides, 2009), the tension-shear loading of butterflyspecimens (Wierzbicki et al., 2005; Mohr and Henn, 2007; Mae et al., 2007; Dunand and Mohr, 2011, Abedini et al., 2015) andthe biaxial loading of cruciform-like specimens (e.g. Abu-Farha et al., 2009; Brenner et al., 2014).

of Mechanical and Process Engineering, Zurich, Switzerland.

Page 2: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 329

The main result of ductile fracture experiments on isotropic materials is the so-called loading path to fracture, i.e. theevolution of the equivalent plastic strain as a function of the stress triaxiality and the Lode parameter. The mechanical fieldswithin the specimen gage section are heterogeneous (at the macroscopic level) in most experiments and the local stress statecannot be calculated based on force measurements using analytical formulas. A hybrid experimental-numerical approach istherefore often required (e.g. Mohr and Henn, 2007; Bai and Wierzbicki, 2008; Dunand and Mohr, 2010; Gruben et al., 2011;Brünig and Gerke, 2011; Lou et al., 2012, 2014; Fourmeau et al., 2013; Malcher et al., 2012). This introduces additional un-certainty in the determined loading paths to fracture related to the employed finite strain plasticity model. A noteworthypredicament of ductile fracture experiments is that their outcome is often no longer a direct experimental observation, but acombined numerical-experimental result.

In micromechanical studies, ductile fracture has been thoroughly investigated for monotonic proportional loading, i.e.for loading histories during which the stress state remains constant up to the point of fracture initiation (e.g. Tvergaard,1981; Koplik and Needleman, 1988; Barsoum and Faleskog, 2007b, 2011; Scheyvaerts et al., 2011; Danas and PonteCastaneda, 2012; Tekoglu et al., 2012; Dunand and Mohr, 2014; Brünig et al., 2014). Moreover, knowledge of the strainto fracture as a function of the stress state for proportional loading is also a main ingredient of fracture initiation modelsthat fall into the category of damage indicator models (e.g. Wilkins et al., 1980; Bai and Wierzbicki, 2010; Lou et al., 2014;Mohr and Marcadet, 2015). Another important predicament of ductile fracture experiments is that proportional loadingconditions at a material point are extremely difficult to achieve. In most ductile fracture experiments, the local stress stateactually evolves throughout loading even if the ratios of the forces acting on the specimen boundaries are kept constant.This stress state evolution is due to changes in the specimen geometry that are almost inevitable in experiments involvinglarge deformations. For example, Ebnoether and Mohr (2013) have shown that in a conventional flat uniaxial tensionspecimen, the stress triaxiality can increase after the onset of necking from 0.33 to values as high as 0.8 at the instantof fracture initiation.

The above experimental predicaments partially prohibit the progress in the field since the direct validation offracture models for proportional loading through experimental results becomes almost impossible. Early works rep-resented the results from fracture experiments in terms of either the average stress triaxiality (e.g. Bao and Wierzbicki,2004) or the stress state at the instant of fracture initiation (Barsoum and Faleskog, 2007a). However, more recentconsiderations for non-proportional loading (Benzerga et al., 2012; Marcadet and Mohr, 2015; Papasidero et al., 2015)raise doubt about the meaningfulness of the representation of experimental data in terms of average or final stresstriaxialities.

In this work, an attempt is made to present fracture experiments that feature (i) a constant stress state up to the onset offracture, and (ii) that allow for the direct determination of the strain to fracture from experimental measurements withoutany numerical simulations. In addition, we focus on stress states that are particularly useful for identifying the plane stressfracture envelope for proportional loading: (1) pure shear, (2) uniaxial tension, (3) plane strain tension, and (4) equi-biaxialtension. New specimen geometries are identified for the first two stress states through constrained shape optimization, whileV-bending and punch experiments are considered for the latter two stress states. To facilitate the fracture characterization inan industrial environment, all experiments are designed such that they can be performed in a universal testing machine.

It is emphasized that we take a 3D continuum mechanical point of view on failure. We are interested in determining theintrinsic failure response at a material point which is assumed to depend on the history of the local mechanical field variablesonly. For mechanical reasons, all specimens are flat and are thus extracted from steel sheets. However, we are not concernedwith sheet metal mechanics where it is common practice to distinguish between in-plane and bending properties. To avoidany confusion, it might beworth thinking of all proposed fracture specimens as thin specimens that have been extracted froma bulk material.

2. Plasticity model

Numerical simulations play a central role in developing, analyzing and validating ductile fracture experiments. Wetherefore begin our presentation with a brief summary of the plasticity and fracture initiation models that are employed inthe sequel. We calibrated both models based on experiments on specimens extracted from a 1 mm thick dual phase steelDP780 provided by US Steel. Thematerial serves as model material in the present work; it is composed of a ferrite matrix withmartensite inclusions and features an average grain size of about 8 mm.

2.1. Plasticity model formulation

A rate-independent non-associated quadratic plasticity model (Mohr et al., 2010) is used to model the material response.This particular model had been proposed based on the results from combined tension-shear experiments on DP590 andTRIP780 steels. It also provided a remarkably accurate description of multi-axial experiments on DP780 steel specimens(Mohr and Marcadet, 2015). Its isotropic yield function is written in terms of the von Mises equivalent stress, s, and adeformation resistance k,

f ½s; k� ¼ s� k ¼ 0: (1)

Page 3: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354330

The non-associated plastic flow is determined through the stress derivative of a flow potential g½s�,

dεp ¼ dlvg½s�vs

: (2)

dεp denotes the increment in the plastic strain vector (in the material coordinate system)

εp ¼ �εp11εp22εp332εp122εp232εp13�T ; (3)

with the 1-, 2-, and 3-directions corresponding to the rolling, transverse and thickness directions, respectively. s denotes the

Cauchy stress vector,

s ¼ ½s11s22s33s12s23s13�T : (4)

and dl � 0 is a scalar plastic multiplier. The flow potential is defined through the quadratic form

g½s� ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðGsÞ$s

p(5)

with

G ¼

26666664

1 G12 �ð1þ G12Þ 0 0 0G12 G22 �ðG22 þ G12Þ 0 0 0�ð1þ G12Þ �ðG22 þ G12Þ 1þ 2G12 þ G22 0 0 00 0 0 G33 0 00 0 0 0 3 00 0 0 0 0 3

37777775: (6)

Note that g½s� corresponds to a special case of the Hill'48 yield function which accounts for the direction dependence ofthe Lankford ratios through the anisotropy coefficients G12, G22 and G33. The case of associated plastic flow is recovered forG12 ¼ �0.5, G22 ¼ 1 and G33 ¼ 3.

The equivalent plastic strain increment dεp is defined as work-conjugate to the von Mises equivalent stress. The material'sstrain hardening behavior is modeled through a linear combination of a power and an exponential law,

kε�εp� ¼ akS

�εp�þ ð1� aÞkV

�εp�

(7)

with the weighting factor a 2 [0,1], the power law (Swift, 1952)

kS�εp� ¼ A

�εp þ ε0

�n; (8)

as defined through the Swift parameters {A,ε0,n}, and the exponential law (Voce, 1948),

kV�εp� ¼ k0 þ Q

�1� e�bεp

�: (9)

as defined through the Voce parameters {k0,Q,b}.

2.2. Plasticity model parameter identification

Uniaxial tension experiments on flat dogbone specimens (Fig. 1a) are performed along three different in-plane directions(0�, 45� and 90� with respect to the rolling directions). All specimens are tested on a hydraulic testing machine at a crossheaddisplacement of 2 mm/min which results in a pre-necking strain rate of about _εp ¼ 10�3=s. The strains along the axial andwidth directions are measured using virtual extensometers with an initial gage length of 7 mm and 5 mm, respectively.Evaluation of the slopes of the logarithmic plastic width strain versus the logarithmic plastic thickness strain (as calculatedassuming plastic incompressibility) along the thickness direction (Fig. 2b) yields the Lankford coefficients r0¼ 0.77, r45 ¼ 0.84and r90 ¼ 0.90. Using the analytical relationships

G12 ¼ � r01þ r0

; G22 ¼ r0r90

1þ r901þ r0

and G33 ¼ 1þ 2r45r90

r0 þ r901þ r0

; (10)

the anisotropy coefficients G12 ¼ �0.44, G22 ¼ 0.92 and G33 ¼ 2.8 are determined.The true stress-strain curves for the three different specimen orientations all lie on top of each other (Fig. 2a) which

supports the assumption of an isotropic yield function (Eq. (1)). In a first step, the Swift parameters {A,ε0,n} and the Voceparameters {k0,Q,b} are determined from fitting expressions (8) and (9) to the true stress versus logarithmic plastic axialstrain curves (Fig. 2c).

Page 4: Ductile fracture experiments with locally proportional loading histories

Fig. 1. Specimens used in this study: (a) uniaxial tension (UT), (b) miniature punch, (c) bending strip, (d) notched tension with radius r ¼ 20 mm (NT20), (e)tension with a central hole (CH), (f) in-plane shear. Blue solid dots highlight the position of the virtual extensometer for relative displacement measurements.

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 331

The difference between the Voce and Swift approximations becomes significant at large strains. The determination ofthe weighting factor a therefore requires the finite element analysis of the specimen response in the post-necking rangewhere very large strains are reached. To allow for the post-necking analysis of the specimen response through an eighth-model of the specimen, a notched tension experiment (Fig. 1d) is selected. The measured forceedisplacement curve for a10 mmwide specimen with two R ¼ 20 mm notches is shown by solid dots in Fig. 2d. The corresponding simulation resultusing the finite element model of Dunand and Mohr (2010) and a weighting factor of a ¼ 0.7 (obtained throughcomputational minimization) is shown as solid line. Note that equivalent plastic strains as high as ε

p ¼ 0:71 prevailwithin the gage section of that specimen. A summary of all plasticity model parameters for the DP780 steel is providedin Table 1.

3. Fracture initiation model

3.1. Fracture initiation model formulations

AHosfordeCoulomb fracture initiationmodel is used to predict the onset of fracture. Its backbone is a localization criterionfor proportional loading in stress-space of the form,

Page 5: Ductile fracture experiments with locally proportional loading histories

Fig. 2. Plasticity of DP780 steel: (a) True stress vs. logarithmic strain curves, and (b) plastic logarithmic width strain vs. thickness strain; (c) inter- andextrapolation of the isotropic hardening function, (d) forceedisplacement curve of notched tension experiment used for inverse identification of the mixedSwift-Voce law.

Table 1Plasticity model parameters for the DP780 steel examined.

A [MPa] ε0 [e] n [e] k0 [MPa] Q [MPa] b [e]

1315.40 0.28E-4 0.146 349.54 536.36 93.07

a [e] E [GPa] n [e] r [kg/m3]

0.70 194 0.33 7850

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354332

sHF þ cðsI þ sIIIÞ ¼ b* (11)

with the ordered principal stresses sI � sII � sIII, and the Hosford equivalent stress defined as

sHF ¼12�ðsI � sIIÞa þ ðsII � sIIIÞa þ ðsI � sIIIÞa

�1a

: (12)

This criterion is transformed into the mixed stress-strain space (Mohr and Marcadet, 2015) leading to an expression of theequivalent plastic strain at the onset of fracture for proportional loading as a function of the stress triaxiality h and the Lodeangle parameter q,

Page 6: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 333

εprf

�h; q� ¼ bð1þ cÞ1n

0@1

2�ðf1 � f2Þa þ ðf2 � f3Þa þ ðf1 � f3Þa

�1a

þ cð2hþ f1 þ f3Þ1A

�1n

(13)

with the model parameters {a,b,c} and the transformation constant n ¼ 0.1.1 The functions fi are Lode angle parameter

dependent trigonometric functions that are associated with the transformation from principal stresses to Haigh-Westergaardcoordinates,

f1�q� ¼ 2

3coshp6�1� q

�i; (14)

f2�q� ¼ 2

3coshp6�3þ q

�i; (15)

f3�q� ¼ �2

3coshp6�1þ q

�i: (16)

The Hosford exponent controls the deviatoric stressmeasure (12) with the limiting cases of Tresca (a¼ 1 or a¼∞) and vonMises (a ¼ 2 or a ¼ 4); the friction coefficient c � 0 controls the influence of the normal stress, while the parameter b � 0controls the overall level of the fracture strains; it has been defined based on b* such that it corresponds to the strain tofracture for uniaxial tension (and equi-biaxial tension).

Fig. 3a shows a 3D plot of the HosfordeCoulomb fracture initiation model for proportional loading in the spacefh; q; εprf g. It describes a monotonic dependency of the strain to fracture on the stress triaxiality and a convex shaped

dependency on the Lode angle parameter with a minimum for generalized shear (q ¼ 0, i.e. stress states for which theintermediate principal stress is equal to the average of the minimum and maximum principal stresses). The above cri-terion for proportional loading is micromechanically-motivated, i.e. Dunand and Mohr (2014) have shown throughcomputational localization analysis that a HosfordeCoulomb type of criterion is suitable for predicting the onset ofcoalescence in a porous solid after proportional loading. To predict fracture initiation after non-proportional loading, Eq.(13) is used in a damage indicator model framework. Let D denote a scalar damage indicator of initial value D ¼ 0 and amaximum value of D ¼ 1 at the instant of fracture initiation, the evolution of the damage indicator is described throughthe differential equation

dD ¼ dεpεprf

�h; q� : (17)

For proportional loading, the current value of the damage indicator can be interpreted as the fraction of the fracture strainthat has been applied to the material.

3.2. Fracture model parameter identification

In view of calibrating the model from experiments on sheet metal, the plane stress representation is most relevant asgeneral 3D stress states cannot be easily generated in experiments. For plane stress conditions, the Lode angle parameter is afunction of the stress triaxiality (Fig. 3b),

q ¼ 1� 2parccos

�� 27

2h

�h2 � 1

3

�for �2=3 � h � 2=3 (18)

and the fracture initiation model can be conveniently represented in the strain to fracture versus stress triaxialityplane (Fig. 3c). Note that the apparent non-monotonic relationship between the strain to fracture and the stresstriaxiality is deceiving as it is actually due to the underlying Lode angle dependency. Also note that irrespective of thechoice of the model parameters, the strain to fracture for uniaxial and equi-biaxial tension are always identical ac-cording to the HosfordeCoulomb model. This is due to the fact that the Hosford equivalent stress is equal to the

1 In the so-called “consistent version” of the HosfordeCoulomb model, the transformation from stress to strain space would need to be performed usingthe material's isotropic hardening law (Mohr and Marcadet, 2015). However, in engineering practice, we recommend using a simple power law with theexponent n ¼ 0.1 to perform this transformation. Small adjustments of the parameters a and b can usually attenuate the effect of this approximation on thestrain to fracture.

Page 7: Ductile fracture experiments with locally proportional loading histories

Fig. 3. (a) Equivalent plastic strain to fracture as a function of the stress state according to the HosfordeCoulomb model for proportional loading; (b) Lode angleparameter as a function of the stress triaxiality for plane stress conditions (blue, black and red curves), (c) HosfordeCoulomb criterion in the fracture strain versusstress triaxiality plane for states of plane stress. Selected stress states are highlighted through solid dots: shear (SH), uniaxial tension (UT), plane strain tension(PST) and equi-biaxial tension (EBT). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354334

maximum principal stress for these two stress states. Furthermore, with the minimum principal stress being zero inboth cases, the HosfordeCoulomb model (Eq. (11)) reduces to maximum principal stress criterion for uniaxial andequi-biaxial tension.

In the following, we will present experimental techniques to characterize the strain to fracture for

� pure shear (SH):

εSHf ¼ b

ffiffiffi3

p 1þ c1

!1n

(19)

ð1þ 2a�1Þa

� uniaxial tension (UT):

εUTf ¼ b (20)

Page 8: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 335

� Plane strain tension (PST):

εPSTf ¼ b

ffiffiffi3

p 1þ c

ð1þ 2a�1Þ1a þ 2c

!1n

(21)

� Equi-biaxial tension (EBT):

εEBTf ¼ b (22)

Note that the strain to fracture for uniaxial tension as well as that for equi-biaxial tension are independent of the modelparameters a and c. Moreover, the strains to fracture for uniaxial tension and equi-biaxial tension are the same since theHosford equivalent stress and the sum of theminimum andmaximumprincipal stress (as normalized by the vonMises stress)are the same for these two stress states. Based on the proportional fracture strain triple fεSHf ; εPSTf ; εEBTf g or fεSHf ; εPSTf ; εUTf g, it isthen straightforward to identify the HosfordeCoulomb model parameters:

1. Determine b from UT or EBT experiment.2. Determine c from PST and SH experiments:

c ¼1�

εPSTf

εSHf

!n

2ffiffiffi3

p

εPSTf

b

!n

þ

εPSTf

εSHf

!n

� 1

(23)

3. Determine the real exponent a from solving the simple implicit equation:

�1þ 2a�1

�1a ¼

ffiffiffi3

pð1þ cÞ

bεSHf

!n

(24)

Note that we recommend limiting the Hosford exponent to the interval 1 � a � 2 to guarantee uniqueness of the solutionof Eq. (24).

4. Equi-biaxial tension

Fracture testing of sheet materials for equi-biaxial tension is probably the “most standard” among the four experimentsdiscussed. The punch test configuration is preferred over hydraulic bulge testing to avoid the evacuation of excess fluid afterfracture. This advantage of the punch test comes at the expense of the possible effect of tool friction on the experimentalresults. Here, we present a punch testing procedure which is similar to the so-called Erichsen cupping test.

4.1. Mini-punch testing device

Fig. 4 shows a photograph and a schematic of the proposed axisymmetric mini-punch testing device. It is designed as anindependent testing fixture which can be inserted into any universal testing machine or press. The main distinctive featureover conventional punch test set-ups is that the punch remains stationary throughout the experiment while the die andclamping platemove downward. This reduces the required focal depth of the DIC camera system, thereby allowing for shorterobject distances and ultimately increasing the spatial resolution of the acquired surface strain fields.

The main components (Fig. 4a and b) of the mini-punch device are:

(a) Stationary part composed of a 18 mm thick aluminum base plate (part①) and a 12.7 mm diameter cylindrical stainlesssteel punch (part ②) with a hemispherical head and an Ra-0.4 surface finish;

Page 9: Ductile fracture experiments with locally proportional loading histories

Fig. 4. (a) Photographs of the mini-punch testing device, and (b) schematic of the axisymmetric cross-section: ① base plate, ② punch, ③ bottom plate, ④ linearroller bearing, ⑤ specimen, ⑥ clamping plate,⑦ steel rods, (8) top plate; (c) Evolution of the equivalent plastic strain and the stress triaxiality for three differentpunch dimensions.

Page 10: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 337

(b) Movable part composed of a 35 mm thick bottom plate (part ③) with an embedded linear roller bearing (part ④), a60 mm diameter disc specimen (part ⑤ and Fig. 1b), a floating clamping plate (part ⑥) applying a clamping force ofabout 140 kN through eightM6 screws, three vertical 10mmdiameter steel rods (items⑦), and a 20mm thick top plate(part ⑧) with a recess to receive a 10 mm diameter steel ball for axial load application.

4.2. On the choice of the punch diameter (to specimen thickness ratio)

Finite element simulations are performed to demonstrate the effect of the punch size on the strain distribution inthe specimen. The FEA model describes a quarter of the punch system with first-order solid elements of a length ofle ¼ 100 mm (corresponding to 10 elements along the thickness direction). Frictionless contact is assumed between thepunch and specimen for the simulations presented in this subsection. Fig. 4c shows the stress triaxiality and equivalentplastic strain distribution at the outer specimen surface for three different punch diameters (D ¼ 12.7 mm, 25.4 mmand 50.8 mm). The simulations have been stopped at a maximum equivalent plastic strain of 1. The plots of theequivalent plastic strain as a function of the normalized radius lie almost on top of each other. Similarly, the stresstriaxiality versus normalized radius curves are all identical within the apex regions of a radius of about r/RPunch < 0.8.It is thus concluded from these simulation results that the choice of the small diameter punch has no noticeabledisadvantage as compared to commonly used large punch sizes as far as the stress and strain distribution on thespecimen surface is concerned (for zero friction). For the 12.7 mm diameter punch, the radial gradients in the surfacestrain field are of the order of 0.1 mm�1 which equates to macroscopic strain variations of less than 1% within a singlegrain.

The simulation results also show that the stress triaxiality decreases from its theoretical value of 2/3 at the apex to aslightly smaller value of 0.65 at a distance of about r=RPunchy0:5. For plane stress conditions, this apparently negligiblechange in stress triaxiality implies a significant change of the Lode angle parameter from �1 (for perfectly equi-biaxialtension) to �0.572. According to the calibrated2 HosfordeCoulomb fracture initiation model for the DP780 steel (Fig. 3c),the material ductility decreases from 0.72 to 0.59 due to this change in stress state. Despite the equivalent plastic strainmaximum at the specimen center, it is thus possible from a theoretical point of view that fracture initiates away from the apexat a strain that is smaller than the maximum strain measured within the specimen.

4.3. Experimental procedure

The miniature punch device is positioned in a hydraulic universal testing machine equipped with a 250 kN axial load cell.The disc specimen (Fig. 1b) with eight thru-holes for the clamping screws is extracted from the sheet metal using water-jetcutting. For optimal results, in particular if the flat specimens are extracted from bulk material stock, we recommendmachining the specimen surface with a sharp end mill to minimize the effect of machining and rolling on the measuredfracture properties (see also discussion in subsection 6.2 on machining artifacts). The experiment is carried out at a constantcross-head velocity of 2 mm/min until the onset of fracture. The surface displacement and logarithmic strain field aredetermined using stereo digital image correlation (see Table 2 for details).

4.4. Experimental results and validation

The measured forceedisplacement curve is shown in Fig. 5a. It increases monotonically with a change in curvature fromconvex to concave at a punch displacement of about 5 mm. Fracture initiates in a stable manner at a radial distance of about0.3mm from the specimen center. With the appearance of visible cracks on the specimen surface (Fig. 5e), the force begins todrop sharply. The repetition of the experiments resulted in nearly identical results, i.e. the measured forceedisplacementcurves lie on top of each other and fracture initiated at almost the same displacement of 7.97 mm.

Based on the DIC determined principal logarithmic surface strains εI and εII (Hencky strains) and the assumptionof incompressibility, we compute the effective strain

ε ¼ 2ffiffiffi3

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiε2I þ εIεII þ ε

2II

q: (25)

The effective strain field is shown in Fig. 5c at the instant of fracture initiation. At the top apex of the punched specimen,we have approximately proportional loading conditions for which the effective strain provides a good approximation of theequivalent plastic strain, εpyε: Not knowing the Lode angle and stress triaxiality histories at the locations of fracture initi-ation, the maximum strain at the apex ðε ¼ 0:72Þ is retained as a lower bound for the strain to fracture for equi-biaxial tension(h ¼ 2/3, q ¼ �1).

2 The calibration has been performed for the DP780 material using the formulas provided in Section 3 after completing all experiments described inSections 5e7.

Page 11: Ductile fracture experiments with locally proportional loading histories

Table 2Details on the digital image correlation. All images are acquired using AVT-Pike 505B/C cameras with a type 2/3 CCD sensor (2452 � 2054 pixels). A 150Wcold light source is used in all experiments to illuminate the area of interest. An eight-tap filter is used to achieve sub-pixel accuracy through spatial grayscaleinterpolation. The strain fields are computed using a Gaussian filter (as built into VIC software).

Exper-iment DIC type camera(s) Lenses Specklesize [mm]

Resol-ution[mm/pixel]

Freq-uency[Hz]

Soft-ware Subset[pixel]

Step[pixel]

Punch stereo 2x AVT Pike(17� azimut, 50� polar)

Nikon 90 mm1:1 macro

100 10 1 VIC-3D 35 10

V-bending stereo 2x AVT Pike(17� azimut, 50� polar)

Nikon 90 mm1:1 macro

100 16 1 VIC-3D 25 6

CH tension planar AVT Pike Nikon 90 mm1:1 macro

100 21 0.5 VIC-2D 15 4

In-plane shear (global) planar AVT Pike Nikon 90 mm1:1 macro

100 21 0.5 VIC-2D 15 4

In-plane shear (local) planar AVT Pike Mitutoyo 1xtelecentric

<50 3.4 0.5 VIC-2D 27 7

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354338

The numerical simulation of the experiment has been performed using the same FE model as in subsection 4.2 with afriction coefficient of m¼ 0.05 for the contact between the punch and the specimen. The computed forceedisplacement curve(solid line in Fig. 5a) matches the experimental curve (solid dots) after correcting its stiffness through an assumed machinestiffness of 19 kN/mm. The strain at the apex obtained from the simulation (Fig. 5d) is εfp ¼ 0:708, which is remarkably close tothat measured experimentally (Fig. 5c).

5. V-bending

Bending of a small rectangular plate specimen is a standardized experiment for evaluating the edge bending (andhemming) capacity of sheet materials (VDA 238-100, DIN 50111, ASTM E290, ISO 7438, and JIS Z2248). Here, V-bendingexperiments are performed to determine the strain to fracture for plane strain conditions. In a typical V-bending experiment arectangular sheet material coupon (Fig. 1c) is placed on top of two parallel rollers featuring a diameter D that is much largerthan the sheet thickness t. The sheet specimen is then loaded through a thin knife-like tool. Provided that the gap between thetwo rollers is not much larger than twice the sheet thickness, a sharp V-bend is formed.

The main modification proposed is keeping the knife stationary, while moving the rollers downwards. As for the punchtest, this configuration has the advantage that the surface strain field can be measured by means of stereo DIC without anyspatial resolution limitations imposed by the limited depth of focus of the optical system.

5.1. Experimental set-up

Fig. 6a shows a 3D sketch of the loading device. It features a base plate (part ①) with four vertical guidance rods (items②) and the central knife-like support point (item ③). The latter is made from a high strength tool steel and features a0.8 mm wide initially flat tip (Fig. 4b). The large roller support points (items ④) are rigidly connected with the movablecross-head (item ⑤) of the loading device. The roller diameters and center-to-center distance are 30 mm and 33.8 mm,respectively.

A 60mm long and 20mmwide specimen is tested (Fig. 1c). Before inserting the specimen (item⑥) into the testing device,a random speckle pattern is applied to the specimen top surface within a 5 mmwide band. The specimen is initially held inplace by the surface friction and dead weight pressure of the cross-head. Two cameras are positioned at an object distance ofabout 450 mm and an angle of 17� with respect to each other. Note that the rollers feature a chamfer to facilitate theobservation of the bent specimen surface. We inserted the loading device into our universal testing machine equipped with a10 kN load cell, which provides the force and cross-head displacement histories in addition to the local surface strainmeasurements. However, note that the determination of the strain to fracture per se only requires a simple press, i.e. the loadmeasurement is not needed.

5.2. Experimental results

Experiments are performed at a constant cross-head velocity of 2 mm/min until cracks are observed by eye on thespecimen outer surface. As detailed in Table 2, stereo digital image correlation is used to obtain the surface strain fields. Fig. 6dshows the recorded forceedisplacement curves for V-bending. The measured surface strain field (Fig. 6c) shows the uni-formity of the axial strain εx along the bending axis (y-axis). The maximum width strain measured along that axis does notexceed

��εy��<0:003 which confirms the validity of the plane strain assumption on the specimen surface. It is assumed that

fracture initiates when a first crack becomes visible on the specimen surface. For the DP780 steel, this instant coincides withthe maximum in the forceedisplacement curve. Based on the measurement of the axial logarithmic strain to fracture, εfx, i.e.

Page 12: Ductile fracture experiments with locally proportional loading histories

Fig. 5. Punch testing of disk specimens: (a) forceedisplacement curve of the experiments (solid lines) and the FE simulation (solid dots); (b) loading path tofracture for the element closest to the center, (c) DIC-based effective strain field for the last picture before fracture initiation. (d) computed equivalent plasticstrain field at force maximum, (e) fractured specimen right after the force maximum, (f) damage indicator field reaching unity (onset of fracture) along the blackline at a radial distance of 1.89 mm from the center.

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 339

the surface strain at the specimen center at the instant of fracture initiation, the corresponding equivalent plastic strain(“fracture strain”) for plane strain conditions is estimated as

εpfy

2ffiffiffi3

p εxf : (26)

Page 13: Ductile fracture experiments with locally proportional loading histories

Fig. 6. (a) Drawing of the V-bending device with ① base plate, ② guidance rods, ③ knife-like punch, ④ rollers, ⑤ upper cross-head, ⑥ specimen (red); thechamfers in the rollers and the upper plate are introduced to free the view for two cameras. (b) basic design principle, (c) DIC contour of the logarithmic strain inx-direction. Besides the visual examination of the image, fracture can also be determined from the loss of correlation (see gray default pixel near center ofzoomed area). (d) Recorded forceedisplacement curves for two experiments and the evolution of the log. strain in x-direction. The first visible crack is denoted bythe arrow. The four black dots in both force and strain plot denote the time points of the four images in Fig. 8. (For interpretation of the references to color in thisfigure legend, the reader is referred to the web version of this article.)

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354340

According to the non-associated anisotropic flow rule, the stress triaxiality for plane strain tension loading along therolling direction is

hPST ¼ G22 � G12

3ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiG222 þ G2

12 þ G12G22

q : (27)

The corresponding Lode angle parameter follows from application of Eq. (18). For the presentmaterial, we have hPST¼ 0.57and qPST ¼ 0:06. In the case of an isotropic flow potential function (i.e. G22 ¼ 1 and G12 ¼ �0.5), the stress triaxiality is hPST ¼1=

ffiffiffi3

py0:58 and qPST ¼ 0.

The repeatability of the experimental procedure is good with a variation of less than 2% in the strain to fracture measuredin two experiments εfp ¼ f0:518; 0:509g. At the same time, the correspondingmeasured forceedisplacement curves lie on topof each other (Fig. 6d).

5.3. Finite element analysis

A finite element simulation is performed to gain more insight into the strain and stress state distribution in a V-bendingexperiment. The results shown in Fig. 7 are obtained using a quarter model of the experimental set-up with eight first-ordersolid elements along the sheet thickness direction, 200 elements along the bending axis (y-direction) and 60 elements alongthe x-direction, resulting in an element size of approximately le ¼ 100 mm in the bent zone; the rollers and the knife aremodeled as rigid bodies, and a roller-to-sheet friction coefficient of 0.1 is assumed.

Page 14: Ductile fracture experiments with locally proportional loading histories

Fig. 7. Results from the numerical simulation of a V-bending experiment: (a) plot of the equivalent plastic strain field, (b) distribution of the equivalent plasticstrain (black), the stress triaxiality (red) and the Lode angle parameter (blue) on the specimen surface along the y-direction (bending axis), (c) evolution of theequivalent plastic strain against the stress triaxiality for an element in the middle of the top surface.

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 341

The simulation results show that the mechanical fields are uniform at a distance greater than 7 mm from the specimenedges. The stress triaxiality is close to 1/3 (uniaxial tension) at the specimen boundary and about 0.55 (close to plane straintension) at the specimen center. The latter deviates slightly from its theoretical valuewhich is attributed to elastic strains. Theplane strain constraint needs to be satisfied for the total strains; due to the positive elastic strains, the flow rule must thus besatisfied for a slightly negative (instead of a zero) plastic strain increment. As a result, the flow rule drives the plane stressstate from plane strain tension towards uniaxial tension, i.e. the stress triaxiality for zero total strain is lower than that of zeroplastic strain along the y-direction. The equivalent plastic strain is lower at the free edge than at the specimen center whichcan be explained by the equivalent plastic strain definition. If the latter is evaluated for UT and PST for the same axial strain,we obtain a factor of εPSTp =εUTp ¼ 2=

ffiffiffi3

p¼ 1:15. Fig. 7b also shows the variation of the Lode angle parameter which is very

significant, i.e. it increases from 0 at the specimen center to 1 at the free edges. However, different from the stress state

Page 15: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354342

gradients encountered in a punch experiment, the location of fracture initiation is not expected to shift according to theHosfordeCoulomb model due to two benign effects: firstly, the ductility is expected to increase, and secondly, the equivalentplastic strain actually decreases towards the free edge.

5.4. Remarks

� Effect of edge condition: It is worth noting that the long edges of the bending specimen need to be machined with a sharptool to avoid premature fracture from the edges. It is known that pre-damage induced during water-jet and shear cuttingresults in a significant reduction of ductility (e.g.Wang, 2015), thereby increasing the fracture initiation from the edges andpossibly overwriting the above benign effects. For example, we had performed some preliminary experiments on spec-imens with shear-cut edges where fracture initiated from one of the free edges (instead of the specimen center).

� Maximum achievable strain: Assuming that the neutral bending axis coincides with the sheet mid-plane, the equivalentplastic strain in a bent sheet of thickness t and inner radius of curvature R is

ε ¼ 2ffiffiffi3

p ln�

Rþ tRþ t=2

: (28)

In the limit of R/ 0, themaximum achievable equivalent strain is about 0.8. In practice, the neutral axis shifts towards thecompression side, thereby generating strains greater than predicted by (28). For our 2R ¼ 0.8 mm wide tool and a sheetthickness of t ¼ 1.06mm, the maximum achievable strain according to (28) is 0.52. Even though slightly larger strains can beproduced in practice (as the neutral axis will be slightly below the center of the sheet), we have not been able to achievefracture in a V-bending experiment on a highly ductile 1 mm thick 22MnB5 steel (before hot-stamping).

� Detection of fracture initiation: The detection of the instant of fracture initiation by eye based on the acquired photographsrequires careful inspection. The photographs shown in Fig. 8 demonstrate this difficulty. We found that the correlationcoefficient maps computed by the DIC software are useful in detecting fracture initiation provided that a sufficiently smallstep and subset size are chosen. It is also worth noting that the difficulty in detecting the instant of fracture initiation in aV-bending experiment is material dependent. For example, a recent V-bending test series on complex phase steels showedthe successive emergence of microcracks and the progressive loss of specimen load carrying capacity. For such materials,both the experimental technique and the fracture modeling framework may not be applicable.

6. Tension with a central hole

As discussed in the introduction, the stress state in dogbone specimens changes throughout loading and the experimentalresults are not suitable for evaluating the fracture response after uniaxial tension. Dunand andMohr (2010) advocated using atensile specimenwith a central hole to characterize the material ductility under uniaxial tension. Experiments on specimenswith a central hole (CH) therefore potentially provide insight into the fracture response for uniaxial tension provided that (a)fracture initiates near the hole boundary, and (b) that the effect of hole machining defects on the material's ductility isnegligible.

6.1. Evolution of the stress state in a CH specimen

A finite element simulation of a tensile experiment on a 20 mm wide CH specimen with a 2.25 mm radius hole has beenperformed to gain some insight into the evolution of themechanical fields. One eighth of the specimen ismodeled using a fine

Fig. 8. Detection of the instant of fracture initiation in a V-bending experiment; emerging crack on the specimen surface. The numbers below the images denotethe acquisition time in seconds.

Page 16: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 343

solid element mesh with eight first-order solid elements along the thickness direction. The evolution of the specimen ge-ometry and equivalent plastic strain fields are shown in Fig. 9 next to the forceedisplacement curve. In addition, theequivalent plastic strain is plotted as a function of the stress triaxiality for the integration point at which the highestequivalent plastic strain is observed in the simulation (i.e. the point where the onset of fracture will be assumed).

In the purely elastic regime, the stress triaxiality is about 0.35. With the onset of plasticity, it decreases to the theoreticalvalue of 0.33 for uniaxial tension (point①). Between points① and②, the equivalent plastic strain distribution is still more orless uniform along the thickness direction, and the stress-triaxiality therefore remains constant up to a strain of 0.4. Note that

Fig. 9. (a) Forceedisplacement curves of the experiment (solid dots) and the simulation (solid line) for a central hole specimen with R ¼ 2.25 mm. (b) Computedevolution of the equivalent plastic strain as a function of the stress triaxiality; (c) effect of central hole size on the evolution of the equivalent plastic strain field:CH-specimen with R ¼ 2.25 mm (left column), and CH-specimen with R ¼ 4 mm (right column). A small arrow highlights the location of the highest strainedelement.

Page 17: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354344

the in-plane gradients in the plastic strain field also induce a non-uniform thickness distribution. As the thickness changesbecome more pronounced (at point ②, the thickness at the hole boundary is 0.81 mm as compared to 1.0 mm at the outeredge), gradients along the thickness-direction become apparent in the mechanical fields. In particular, an out-of-plane stressdevelops due to the curvature of the specimen surface. At point③, the strain at the specimen top surface is 0.52 while a strainof 0.6 prevails at the specimen center. This effect becomes more pronounced as the specimen is loaded further. At the instantof specimen failure (point ④), a maximum equivalent plastic strain of 0.7 is reached at a stress triaxiality of 0.35. The spatialgradients in the mechanical fields around the highest strained material point are dh=dxy0:61=mm and dεp=dxy� 0:15=mmat the instant of fracture initiation.

A second simulation for a CH specimenwith a larger hole (4mm radius) showed different results (right column of Fig. 9). Inparticular, the highest strained material point moves away from the edge towards the hole ligament center. In very simpleterms, for large holes, the two hole ligaments may be seen as two parallel uniaxial tension specimens. The stress stateevolution in large diameter CH specimens therefore resembles that of dogbone specimens, i.e. the stress triaxiality is onlyconstant and close to 1/3 up to the point of the onset of through-thickness necking; thereafter, it constantly increases at a rateof about dh=dεpy0:13 (Fig. 9b).

The results from a simulation with a very small 1 mm radius hole are very similar to those obtained above. The maindifference is that the gradients in the mechanical fields increase. Practical arguments regarding the machining of smalldiameter holes certainly impose a minimum hole size. Furthermore, the size lu of the uniformly loaded (e.g. stress triaxialityand plastic strain variations of less than 5%, which yields lu ¼ 100 mm for the present example) vicinity of the point of stressinitiation should be large as compared to the size of the machining affected zone, lMAZ, and the grain size, lg,

lu > >max�lMAZ ; lg

�: (29)

Given that the gradients in the mechanical fields decrease with the size of the hole, the determination of the “optimal”hole diameter corresponds to finding the largest hole diameter for which the highest strained point is still sufficiently close tothe hole boundary, such that the stress state remains approximately uniaxial throughout the entire loading history.

6.2. Size of the machining affected zone (MAZ)

The pre-damaged zone around the hole boundary is referred to as “Machining Affected Zone (MAZ)”. The following fourprocedures for introducing a 5 mm diameter hole into a 1 mm thick DP780 steel sheet have been considered:

� Water-jet cutting: The edge imperfections introduced by water-jet cutting are visible to the naked eye. In particular, thewater-jet produces a slanted surface with a difference of approximately 0.12 mm between the radius on the top and thebottom surface of the sheet.

� Drilling: Microscopic analysis of the drilled holes reveals that the material is plastically deformed with a zone of 10 mmaround the edge (Fig. 10a).

� Drilling followed by reaming: In the case of reaming after drilling, the width of the MAZ is reduced to about 5 mm (Fig. 10b).� CNC milling: Using a sharp 4 mm diameter four-flute end mill, the MAZ size could be reduced to less than lMAZ ~ 2 mm(Fig. 10c).

As shown by Dunand and Mohr (2010), the fracture strains obtained from experiments on water jet cut specimens can bemore than 50% lower than that obtained from CNCmilled specimens. EDMmachining has been disregarded due to the knowndeep (>10 mm) penetration of wire atoms into the substrate material. Similarly, laser cutting has not been considered due tothe unavoidable significant material property changes within the solidified molten material near the hole boundary.

Based on the above considerations, it is thus recommended to introduce the central hole through CNCmilling tominimizethe effect of MAZ on the fracture strains determined from experiments on CH specimens.

6.3. Parametric study on optimal hole diameter

As stated above, the largest hole diameter is considered as “optimal” for a CH specimen of givenwidth as long as the stressstate at the location of fracture initiation remains close to uniaxial tension all theway to fracture. The optimal hole diameter isexpected to be a function of the material hardening response and the material ductility due to the important effect ofthrough-thickness necking on the mechanical fields in a CH specimen. A parametric study is thus performed to shed somelight on the relationship between the “optional” radius and the material properties. Using the same boundary conditions andmesh densities as in the above examples, we prepared FE models for:

� 33 different hole radii, varying in 0.25 mm increments from r ¼ 1 mm to r ¼ 9 mm� 42 different hardening behaviors: using an elasto-plastic J2-plasticity material model, the following grid of Swift hard-ening parameters was considered: A¼ 500, 700,…, 1500MPa and n¼ 0.01, 0.05, 0.10,…, 0.30; the same reference strain ofε0 ¼ 0.002 has been used in all simulations.

� 10 different fracture strains: εf ¼ 0:1; 0:2; ::: ; 1:0 (the effect of stress state is neglected here);

Page 18: Ductile fracture experiments with locally proportional loading histories

Fig. 10. SEM images of the Machining Affected Zone (MAZ) for different machining techniques: (a) drilling, (b) reaming, and (c) CNC milling.

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 345

In sum, 33 � 6 � 7 � 10 ¼ 13860 simulations are carried out. For a given set fA;n; εf gi of material parameters, finiteelement simulations are performed for 10 different radii rj to determine the average stress triaxiality <h(i,j)> at thehighest strained point. Subsequently, we report the optimal radius RðiÞopt as the maximum among all radii for which theaverage stress triaxiality is closest to 1/3. In all simulations, we observed that the stress triaxiality is always greater than1/3. In other words, we did not observe any loading path with an average close to 1/3 that featured stress triaxialitiesbelow and above 1/3. The deviation of the average stress triaxiality from the target 1/3 is thus also a meaningful measureof the proportionality of the loading paths, i.e. an average triaxiality of 1/3 can only be achieved if the stress state isindeed constant throughout loading. The influence of A on the optimal radius turned out to be small, changing itmaximally by 0.25 mm. Fig. 11 therefore shows the obtained function Ropt ¼ Ropt ½n; εf � for A ¼ 1100 MPa only. The mainobservations are that:

Page 19: Ductile fracture experiments with locally proportional loading histories

Fig. 11. Results of the parameter study for the size of the central hole. Surface plot of the optimal radius versus assumed equivalent plastic strain to fracture andhardening exponent.

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354346

� the optimal radius increases as a function of the strain hardening exponent: this may be explained by the fact that lig-ament necking is more likely to occur for low n values; consequently smaller radii are required to promote the localizationat the hole boundary;

� the optimal radius decreases as a function of the strain to fracture; the same argument as abovemay be applied: the higherthe strain to fracture, the more strain hardening capacity is consumed prior to fracture, and consequently, smaller radii arerequired to promote the localization at the hole boundary.

The strain hardening coefficient of most metallic engineering materials is seldom higher than 0.2, and we thereforerecommend a hole radius of about 2 mm to ensure that fracture initiates under uniaxial conditions near the hole boundary.

6.4. Experimental validation

Tensile experiments are performed on CH-specimens with two different hole radii: one with a radius of R¼ 2mm and onewith R¼ 4mm (of a different batch of DP780 steel). The specimens were extracted from the steel sheet initially usingwater jetcutting for the main geometry, followed by CNC milling to enlarge the holes to their final dimensions. Using custom highpressure clamps, the specimen is clamped into a universal testing machine and loaded at a cross-head speed of 0.4 mm/minuntil fracture. To perform planar DIC (see Table 2 for details), a random black and white speckle pattern is applied onto thespecimen surfaces.

The solid dots in Fig. 12a show the measured forceedisplacement curves next to those obtained from numerical simu-lations. The solid lines in Fig. 12b show the evolution of the equivalent plastic strain as a function of the stress triaxiality forthe highest strained elements. While the loading path of the specimen with R ¼ 2 mm stays close to 1/3 and fractures ath ¼ 0.41, the one for the specimen with R ¼ 4 mm deviates early and fails at h ¼ 0.53; the specimen with the smaller radiusfailed at a strain of εf ¼ 0:98, while the specimen with the larger radius could only achieve a maximum strain of εf ¼ 0:79.This result is consistent with the HosfordeCoulomb model which predicts a loss of ductility if the stress triaxiality increasesunder plane stress conditions between uniaxial and plane strain tension.

However, the difference between these two specimens becomes important when the strain at the hole boundary of theR ¼ 4mm specimen (see dashed loading path in Fig. 12b) is used to estimate the strain to fracture for uniaxial tension. In thatcase, the fracture strain estimate of the R¼ 4 mm specimenwould be almost 25% lower than that for the R ¼ 2 mm specimen(εUT ¼ 0:98 versus εUT ¼ 0:74). It is thus reemphasized here that an inappropriate choice of hole size and/or machiningtechnique may lead to significant errors (underestimation) in the determined strain to fracture for uniaxial tension.

7. In-plane shear

Themain challenge of in-plane shear experiments on flat specimens is premature fracture initiation near free gage sectionboundaries (Mohr and Henn, 2007; Ghahremaninezhad and Ravi-Chandar, 2013). By definition, the shear stress componentsalong the tangent direction tmust be zero at a free boundary. For plane stress specimens it is straightforward to show that thestress state at a free boundary is uniaxial tension, i.e. s ¼ sjjt5t. Unless the strain to fracture for pure shear is much lowerthan that for uniaxial tension, fracture is more likely to initiate under uniaxial tension at the free specimen boundary (orunder plane strain tension near the grips) than under pure shear at the specimen center. To address this issue, Mohr and Henn(2004) proposed reducing the gage section thickness such that the strains are significantly higher at the specimen center,thereby increasing the probability of fracture initiation under pure shear near the specimen center. Here, an in-plane shear

Page 20: Ductile fracture experiments with locally proportional loading histories

Fig. 12. Results of the hybrid experimental-numerical approach of two central hole specimens with radius R ¼ 2 mm and R ¼ 4 mm from a different batch ofDP780. (a) Force-Displacement curve of the experiment (dotted) and the simulation (line). (b) Loading paths as extracted from the element with the highestequivalent plastic strain at the onset of failure (solid lines) and from the element at the edge of the R ¼ 4 mm specimen (dashed line).

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 347

fracture specimen for sheet metal is presented that does not require any thickness reductions. This is achieved through a FEA-based optimization of the specimen shape.

7.1. Specimen design

Inspired by thework of Till and Hackl (2013) andMiyauchi (1984), a “smiley” shear specimen comprising two parallel gagesections is designed for fracture testing (Fig. 1f). Themain geometric feature of the specimen is a set of notches that define thecontours of the shear gage sections. To prevent shear buckling, the gage sectionwidth and height shall be of the same order asthe sheet thickness. A finite element simulation of a shear experiment3 using the basic specimen geometry (Fig. 14a) confirmsthe above concerns regarding pre-mature fracture initiation from the free boundaries (Fig. 14c).

We therefore create a shape optimization problem for the outer contour of the shear gage section (Fig. 13a). In particular,we consider notch contours of overall width h that.

(i) are positioned point-symmetrically at an off-set Dx with respect to the gage section center {xc,yc}(ii) feature a notch radius Rn around the notch centers fxc±ðDxþ DxnÞ ; yc±Dygwith Dy ¼ w/2 þ Rn and Dxn, an additional

offset between the centers of the radii;(iii) feature a smooth transition of filet radius Rf from the notch tangent to the parallel notch boundaries.

3 The HosfordeCoulomb parameters {a,b,c} ¼ {1.61, 0.71, 0.062} were chosen for the optimization.

Page 21: Ductile fracture experiments with locally proportional loading histories

Fig. 13. (a) Illustration of the parameters of the model. (b) Finite element model of the optimized shear specimen geometry. The boundary zone (red) and centralzone (blue) are highlighted. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354348

As illustrated in Fig. 13, the vector p ¼ fxc;Dx;h;w;Rn;Dxn;Rf g summarizes all parameters characterizing the specimengeometry. Using symmetry boundary conditions, only the right quarter of the shear specimen is modeled (Fig. 13b). Themodel is created in a manner such that a uniform mesh with an element length of approximately le ¼ 100 mm is achieved inthe gage section area. The finite element software Abaqus/Explicit (Abaqus, 2012) is used to perform the numerical simu-lations. The lower boundary of the specimen is clamped, while a constant vertical velocity is applied to the upper boundary.All simulations are stopped when the damage indicator reaches unity.

The scalar cost function F ¼ F½pk� is defined as

F½pk� ¼maxi2SN

fDigmaxi2SG

fDigþZT

jhjdt þZT

��q��dt; (30)

with Di denoting the damage indicator in the final time step for an element i, and SN and SG denoting two distinct element setscomprising all elements on the free gage section boundary and gage section center respectively (Fig. 13b). The first termtherefore ensures that fracture is more likely to initiate near the gage section center than at the free gage section boundary.The second and third terms are added tominimize variations in the stress triaxiality and Lode angle parameter histories at thepoint of maximum D. Recall that the target values for pure shear are h ¼ 0 and q ¼ 0.

The above cost function is minimized using a derivative-free simplex algorithm (Nelder and Mead, 1965). A solutionpredicting fracture initiation within the gage section center, with small variations in the stress state, is found after about 200iterations. The minimization has been repeated for ten different starting points p0 which all led to the specimen shown inFig. 13b. The overall notch width and gage section width are h ¼ 2.04 mm and w ¼ 2.36 mm, with a notch radius ofRN ¼ 0.62 mm and an off-set Dx ¼ 0.19 mm.

The distribution of the equivalent plastic strainwithin the gage section of the optimized specimen at the predicted instantof fracture initiation is shown in Fig. 14d. A comparison with the seed geometry simulation result shows that much higherstrains could be achieved in the sheared gage section (0.86 vs. 0.74). Another way of illustrating the advantage of the opti-mized over the seed geometry is to compare the critical strains at the free gage section edge for the same shear strains at thespecimen center. For example, for an equivalent plastic strain at the center of 0.74, the seed specimen experiences a strain of0.61 at the edge (i.e. it fractures) while the optimized specimen features an edge strain of 0.52 only.

Page 22: Ductile fracture experiments with locally proportional loading histories

Fig. 14. (a) Finite element model of the geometry proposed by (a) Till&Hackl and (b) the optimized geometry. (c) and (d) show the equivalent plastic strain at theonset of fracture (D ¼ 1) of the specimen proposed by Till&Hackl. Note how much more strain the optimized geometry can endure. (e) The evolution of theequivalent plastic strain as a function of the stress state in the element where fracture commences.

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 349

Page 23: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354350

7.2. Experimental procedure

To validate the proposed shear experiment, the optimized specimen is extracted from the sheet metal using wire electric-dischargemachine (EDM). It is inserted into our hydraulic tensile testing machine and loaded at constant cross-head speed of0.2 mm/min. Throughout loading, the specimen is observed with two cameras: a first camera equipped with a 90 mm 1:1macro lens monitors the relative displacement of the specimen shoulders, while a second camera equipped with a Mitutoyo1x telecentric lens takes high resolution pictures (3.4 mm/pixel) of the two gage sections only (see Table 2 for details on DIC).

7.3. Results

The recorded forceedisplacement curves are shown in Fig. 15a. Note that the force drops after the instant of fractureinitiation. The curves for the two experiments lie on top of each other, which confirms the repeatability of the experimentalprocedure. In addition, Fig. 15b shows the measured effective strain field, which provides a good approximation of theequivalent plastic strain field for nearly proportional loading histories.

At the instant of fracture initiation, the strain has reached amaximumvalue of 0.86. Despite the above efforts in optimizingthe specimen geometry, this measurement is still interpreted as a lower bound estimate for the strain to fracture under pureshear loading due to the absence of a clear experimental proof of fracture initiation at the specimen center. The crackpropagation after fracture initiation is unstable, i.e. wewere unable to stop the experimentwith a partial crack only. Assuminga crack velocity of the order of 103m/s, a high resolution high speed camerawith an acquisition frequency above 1MHzwouldbe required to monitor the crack propagation. However, it can be concluded from Fig. 15c that the observed final crack path isconsistent with the assumption of crack initiation under pure shear away from the gage section boundaries.

Fig. 15. Results for the in-plane shear specimen: (a) Comparison of the experimental to the simulated forceedisplacement curve. (b) Surface plot of the effectivestrain field just before the onset of fracture; (c) fractured gage section.

Page 24: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 351

The finite element simulation results agree well with the experimental observations. The predicted forceedisplacementcurve lies on top of those measured experimentally, while the contour plot of the equivalent plastic strain distribution(Fig. 14d) is also very similar to that obtained from DIC (Fig. 15b). In particular, it predicts nearly the same value of themaximum equivalent plastic strain (0.85 vs 0.86). The simulation estimates of the evolution of the Lode angle parameter andthe stress triaxiality are shown in Fig. 14e. The average values are <h> ¼ 0.027 and <q> ¼ 0.069, while fracture initiates athf ¼ 0.12 and qf ¼ 0:32. With regards to the fracture initiation model for the DP780 steel, these variations are “small”, i.e. the

Fig. 16. Results for the shape optimization for three engineering materials: A 1.4 mm thick DP590 dual phase steel (black), a 2 mm thick AA2198-T8R (red) and a1.5 mm thick Ti6Al4V (blue). (a) Comparison of the hardening behavior of the three materials. (b) Comparison of the fracture loci for the materials. The solid partof the curves represents the calibrated fracture locus, while the dashed part corresponds to the assumed fracture locus. Optimized geometries for the (c) DP590,the (d) AA2198-T8R and the (e) Ti6Al4V. (For interpretation of the references to color in this figure legend, the reader is referred to the web version ofthis article.)

Page 25: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354352

strain to fracture for proportional loading changes by less than 5%within the range of stress states experienced at the locationof maximum equivalent plastic strain.

7.4. Influence of the material properties on the shear specimen geometry

It is emphasized that the above specimen shape “optimization” has been performed for one specificmaterial (DP780 steel).As for the central hole tension specimen, it is expected that the optimal specimen shape depends on the elasto-plastic andfracture properties of a material. If the identified specimen geometry is used for other materials, it may be expected that theobtained fracture strain is only a lower bound for the material ductility under pure shear. Given the high computational costsassociated with the optimization, a comprehensive parametric study is omitted here. However, using the abovemethodology,we have identified the “optimal” shear specimen geometry for three other engineering materials: (i) a 1.4 mm thick DP590dual phase steel, (ii) a 2 mm thick AA2198-T8R aluminum alloy, and a 1.5 mm thick Ti6Al4V titanium alloy (Tancogne-Dejeanet al., 2016). The corresponding stress-strain curves and assumed fracture envelopes for plane stress conditions are shown inFig. 16a and b respectively. The obtained “optimal” specimen geometries are defined through the parameters given in Table 3and are shown in Fig. 16cee. The main geometrical features changing are the notch radii, the offset between the two notchesand the notch width. Even though the DP590 and the Ti6Al4V exhibit a similar strain hardening exponent, their fracture locidiffer significantly which leads to a different shear specimen geometries. Different geometries are also obtained for AA2198and the DP590 which feature similar fracture loci, but differing hardening behavior.

8. Concluding remarks

Four basic fracture experiments with approximately proportional loading condition all the way until fracture initiation areanalyzed in detail. They provide the strains to fracture for pure shear, uniaxial tension, plane strain tension and equi-biaxialtension. Two original testing devices are presented to perform reliable punch and V-bending experiments. The distinctivefeature of both devices is that the location of fracture initiation remains stationary throughout the experiment, therebyincreasing the spatial resolution of the stereo DIC surface strain measurements. The in-depth analysis of experiments ontensile specimenswith a central hole demonstrated the importance of the size of the chosen hole. If the hole size is not chosenappropriately, significant errors in the estimated strains to fracture for uniaxial tension are expected. A comprehensiveparametric study revealed that a 4 mm diameter hole in a 20 mm wide tensile gage section is expected to yield satisfactoryresults for most engineering materials.

Table 3Geometry parameters for the smiley shear specimens for the DP780 examined, a DP590, an AA2198 and a Ti6Al4V. All dimensions are given in mm.

[mm] xc Dx h w Rn, Dxn Rf

DP780 2.85 0.05 2.04 2.36 0.62 0.14 1.03DP590 3.18 0.11 1.86 2.70 0.55 0.23 1.81AA2198 2.99 0.14 1.20 1.50 0.23 0.00 1.61Ti6Al4V 2.87 0.10 1.48 1.66 0.35 0.00 1.35

Fig. 17. Evolution of the principal strain paths for the punch (black), bending (red) and shear (blue) experiments as extracted from digital image correlation at thepoint where fracture occurred. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Page 26: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354 353

In addition, a new flat shear fracture specimen is proposed. It features two parallel gage sections without any thicknessreduction. The FEA based shape optimization elucidated the known issues of in-plane shear specimens: if the geometry is notchosen appropriately, fracturewill initiate prematurely from the free gage section boundaries under a tensile stress state. As aresult, a poor lower bound estimate of the strain to fracture for pure shear loading is obtained. All proposed experiments havebeen performed on specimens extracted from 1 mm thick DP780 steel sheets. A simple analytical procedure is shown toidentify the three material parameters of the HosfordeCoulomb fracture initiation model based on measured strains tofracture.

Compared to other multi-axial experimental techniques such as notched tension, tension-torsion, or Nakazima experi-ments, the proposed four experiments have the important qualitative advantage of approximately constant stress statesthroughout the entire loading history up to the instant of fracture initiation (Fig. 17). As a result, important points on the so-called fracture surface (fracture strain as a function of the stress triaxiality and Lode angle parameter) for proportional loadingcan finally be determined from experiments thereby removing uncertainty related to non-proportional loading effects. Themain limitation of the present work is that the fracture strain can be identified for only four distinct stress states. However,based on our current understanding of the ductile fracture process of modern engineering materials, three of the four pointsare local extrema for plane stress conditions: the V-bending experiment features the most critical stress state (ductilityminimum for plane strain tension), while local ductility maxima are expected to prevail for uniaxial tension (CH experiment)and equi-biaxial tension (mini-punch experiment). The result from the pure shear experiment (when compared to that foruniaxial tension) provides insight into the competition of the stress triaxiality and Lode parameter effects which is crucial forextrapolating the fracture surface into the range of negative stress triaxialities. The entire fracture surface cannot be con-structed from four experimental points only. However, given that the proposed experimental program includes importantextreme points, it is expected that more reliable predictions may be achieved when existing fracture theories are calibratedbased on the obtained material data.

If ductile fracture is indeed due to the loss of ellipticity at a material point (e.g. Rice, 1976), it can probably only be detectedin experiments when the emerging shear/normal localization band reaches a critical length. This process of finite bandformation is expected to depend on spatial gradients in the mechanical fields at the instant of fracture initiation which is aremaining experimental uncertainty in our search for the “intrinsic” fracture envelope for proportional loading.

Acknowledgments

The partial financial support through theMIT Industrial Fracture Consortium is gratefully acknowledged. Thanks are due toDr. B. Hackl fromVoestAlpine for the suggestion of the initial shear specimen geometry and discussion. Thanks are also due toDr. B. Erice (Ecole Polytechnique) for valuable discussions.

References

Abaqus, 2012. Reference Manuals v6. Abaqus Inc, pp. 12e13.Abedini, A., Butcher, C., Anderson, D., Worswick, M., Skszek, T., 2015. Fracture characterization of automotive alloys in shear loading. In: SAE 2015 World

Congress & Exhibition, Paper #2015-01-0528.Abu-Farha, F., Hector, L.G., Khraisheh, M., 2009. Cruciform-shaped specimens for elevated temperature biaxial testing of lightweight materials. JOM 61 (8),

48e56.Bai, Y., Wierzbicki, T., 2008. A new model of metal plasticity and fracture with pressure and lode dependence. Int. J. Plast. 24, 1071e1096.Bai, Y., Wierzbicki, T., 2010. Application of the extended Coulomb-Mohr model to ductile fracture. Int. J. Fract. 161, 1e20.Bao, Y., Wierzbicki, T., 2004. On fracture locus in the equivalent strain and stress triaxiality space. Int. J. Mech. Sci. 46 (1), 81e98.Barsoum, I., Faleskog, J., 2007a. Rupture mechanisms in combined tension and shear e experiments. Int. J. Solids Struct. 44, 1768e1786.Barsoum, I., Faleskog, J., 2007b. Rupture mechanisms in combined tension and shear e micromechanics. Int. J. Solids Struct. 44, 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,

925e938.Benzerga, A.A., Surovik, D., Keralavarma, S.M., 2012. On the path-dependence of the fracture locus in ductile materials e analysis. Int. J. Plast. 37, 157e170.Brenner, D., Gerke, S., Brünig, M., 2014. Numerical simulation of damage and failure behavior of biaxially loaded specimens. In: Proceedings of 11th World

Congress on Computational Mechanics, pp. 1211e1219.Brünig, M., Chyra, O., Albrecht, D., Driemeier, L., Alves, M., 2008. A ductile damage criterion at various stress triaxialities. Int. J. Plast. 24 (10), 1731e1755.Brünig, M., Gerke, S., 2011. Simulation of damage evolution in ductile metals undergoing dynamic loading conditions. Int. J. Plast. 27 (10), 1598e1617.Brünig, M., Gerke, S., Hagenbrock, V., 2014. Stress-state-dependence of damage strain rate tensors caused by growth and coalescence of micro-defect. Int. J.

Plast. 63, 49e63.Danas, K., Ponte Casta~neda, P., 2012. Influence of the lode parameter and the stress triaxiality on the failure of elasto-plastic porous materials. Int. J. Solids

Struct. 49 (11e12), 1325e1342.Driemeier, L., Brünig, M., Michelia, G., Alves, M., 2010. Experiments on stress-triaxiality dependence of material behavior of aluminum alloys. Mech. Mater.

42 (2), 207e217.Dunand, M., Mohr, D., 2010. Hybrid experimentalenumerical analysis of basic ductile fracture experiments for sheet metals. Int. J. Solids Struct. 47 (9),

1130e1143.Dunand, Matthieu, Mohr, Dirk, 2011. Optimized butterfly specimen for the fracture testing of sheet materials under combined normal and shear loading.

Eng. Fract. Mech. ISSN: 0013-7944 78 (17), 2919e2934. http://dx.doi.org/10.1016/j.engfracmech.2011.08.008.Dunand, M., Mohr, D., 2014. Effect of lode parameter on plastic flow localization after proportional loading at low stress triaxialities. J. Mech. Phys. Solids 66,

133e153.Ebnoether, F., Mohr, D., 2013. Predicting ductile fracture of low carbon steel sheets: stress-based versus mixed stress/strain-based Mohr-Coulomb model.

Int. J. Solids Struct. 50 (7e8), 1055e1066.Faleskog, J., Barsoum, I., 2013. Tension-torsion experiments e part I: experiments and a procedure to evaluate the equivalent plastic strain. Int. J. Solids

Struct. 50, 4241e4257.

Page 27: Ductile fracture experiments with locally proportional loading histories

C.C. Roth, D. Mohr / International Journal of Plasticity 79 (2016) 328e354354

Fourmeau, M., Børvik, T., Benallal, A., Hopperstad, O.S., 2013. Anisotropic failure modes of high-strength aluminium alloy under various stress states. Int. J.Plast. 48, 34e53.

Gao, X., Zhang, G., Roe, C., 2010. A study on the effect of the stress state on ductile fracture. Int. J. Damage Mech. 19, 75e93.Ghahremaninezhad, A., Ravi-Chandar, K., 2013 March. Ductile failure behavior of polycrystalline Al 6061-T6 under shear dominant loading. Int. J. Fract. 180

(1), 23e39.Gruben, G., Fagerholt, E., Hopperstad, O.S., Børvik, T., 2011. Fracture characteristics of a cold-rolled dual-phase steel. Eur. J. Mech. A/Solids 30 (3), 204e218.Haltom, S.S., Kyriakides, S., Ravi-Chandar, K., 2013. Ductile failure under combined shear and tension. Int. J. Solids Struct. 50, 1507e1522.Koplik, J., Needleman, A., 1988. Void growth and coalescence in porous plastic solids. Int. J. Solids Struct. 24 (8), 835e853.Korkolis, Y.P., Kyriakides, S., 2009. Path-dependent failure of inflated aluminum tubes. Int. J. Plast. 25 (11), 2059e2080.Kuwabara, T., Yoshida, K., Narihara, K., Takahashi, S., 2005. Anisotropic plastic deformation of extruded aluminum alloy tube under axial forces and internal

pressure. Int. J. Plast. 21 (1), 101e117.Lou, Y.S., Huh, H., Lim, S., Pack, K., 2012. New ductile fracture criterion for prediction of fracture forming limit diagrams of sheet metals. Int. J. Solids Struct.

49, 3605e3615.Lou, Yanshan, Yoon, Jeong Whan, Huh, Hoon, 2014. Modeling of shear ductile fracture considering a changeable cut-off value for stress triaxiality. Int. J.

Plasticity. ISSN: 0749-6419 54, 56e80. http://dx.doi.org/10.1016/j.ijplas.2013.08.006.Mae, H., Teng, X., Bai, Y., Wierzbicki, T., 2007. Calibration of ductile fracture properties of a cast aluminum alloy. Mater. Sci. Eng. A 459, 156e166.Malcher, L., Andrade Pires, F.M., C�esar de S�a, J.M.A., 2012. An assessment of isotropic constitutive models for ductile fracture under high and low stress

triaxiality. Int. J. Plast. 30e31, 81e115.Marcadet, St�ephane J., Mohr, Dirk, 2015. Effect of compressionetension loading reversal on the strain to fracture of dual phase steel sheets. Int. J. Plasticity.

ISSN: 0749-6419 72, 21e43. http://dx.doi.org/10.1016/j.ijplas.2015.05.002.Miyauchi, K., 1984. A proposal for a planar simple shear test in sheet metals. Sci. Pap. RIKEN 81, 27e42.Mohr, D., Henn, S., 2004. A New Method for Calibrating Phenomenological Crack Formation Criteria. Technical Report 113. Impact and Crashworthiness

Laboratory, Massachusetts Institute of Technology.Mohr, D., Henn, S., 2007. Calibration of stress-triaxiality dependent crack formation criteria: a new hybrid experimentalenumerical method. Exp. Mech. 47,

805e820.Mohr, D., Marcadet, S.J., 2015. Micromechanically-motivated phenomenological Hosford-Coulomb model for predicting ductile fracture initiation at low

stress triaxialites. Int. J. Solids Struct. ISSN: 0020-7683. Available online 23 February 2015. http://dx.doi.org/10.1016/j.ijsolstr.2015.02.024.Mohr, D., Dunand, M., Kim, K.H., 2010. Evaluation of associated and non-associated quadratic plasticity models for advanced high strength steel sheets

under multi-axial loading. Int. J. Plast. 26 (7), 939e956.Nelder, A., Mead, R., 1965. A simplex method for function minimization. Comput. J. 7, 308e313.Papasidero, J., Doquet, V., Mohr, D., 2015. Ductile fracture of aluminum 2024-T351 under proportional and non-proportional multi-axial loading: Bao-

Wierzbicki results revisited. Int. J. Solids Struct. 69e70, 459e474.Rice, J.R., 1976. The localization of plastic deformation. In: Proceedings of the 14th International Congress on Theoretical and Applied Mechanics, Delft, 1976,

ed. W.T. Koiter), vol. 1, pp. 207e220.Scheyvaerts, F., Onck, P.R., Tekoglu, C., Pardoen, T., 2011. The growth and coalescence of ellipsoidal voids in plane strain under combined shear and tension.

J. Mech. Phys. Solids 59, 373e397.Swift, H.W., 1952. Plastic instability under plane stress. J. Mech. Phys. Solids 1, 1e18.Tancogne-Dejean, T., Roth, C.C., Woy, U., Mohr, D., 2016. Probabilistic fracture of Ti-6Al-4V made through additive layer manufacturing. Int. J. Plasticity 78,

145e172.Tekoglu, C., Leblond, J.B., Pardoen, T., 2012. A criterion for the onset of void coalescence under combined tension and shear. J. Mech. Phys. Solids 60,

1363e1381.Till, E., Hackl, B., 2013. Calibration of plasticity and failure models for AHSS sheets. In: Proceedings of the International Deep Drawing Research Conference

IDDRG 2013.Tvergaard, V., 1981. Influence of voids on shear band instabilities under plane strain conditions. Int. J. Fract. 17 (4), 389e407.Voce, E., 1948. The relationship between stress and strain for homogeneous deformation. J. Inst. Metals 74, 537e562.Wang, K., 2015. MIT Department of Mechanical Engineering (PhD thesis).Wierzbicki, T., Bao, Y., et al., 2005. Calibration and evaluation of seven fracture models. Int. J. Mech. Sci. 47 (4e5), 719e743.Wilkins, M.L., Streit, R.D., Reaugh, J.E., 1980. Cumulative-strain-damage Model of Ductile Fracture: Simulation and Prediction of Engineering Fracture Tests.

Lawrence Livermore National Laboratory Report, UCRL-53058.