Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… ·...

16
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Transcript of Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… ·...

Page 1: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

This article appeared in a journal published by Elsevier. The attachedcopy is furnished to the author for internal non-commercial researchand education use, including for instruction at the authors institution

and sharing with colleagues.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Page 2: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

A study of intergranular fracture in an aluminium alloydue to hydrogen embrittlement

E. Pouillier a, A.-F. Gourgues a, D. Tanguy b, E.P. Busso a,⇑a Centre des Matériaux, École des Mines de Paris-ParisTech, CNRS UMR 7633, BP 87, 91003 Evry Cedex, Franceb Ecole des Mines de Saint-Etienne, 158 cours Fauriel, 42023 Saint-Etienne, France

a r t i c l e i n f o

Article history:Received 27 October 2011Received in final revised form 7 January2012Available online 1 February 2012

Keywords:Intergranular fractureAluminium alloysHydrogen embrittlementIn situ tensile testsCrystal plasticity modelling

a b s t r a c t

This work concerns a study of the effects of plasticity on the mechanism of intergranularcracking assisted by hydrogen induced embrittlement in an aluminium alloy. Here, tensilespecimens charged with hydrogen were used to investigate quantitatively the effect ofplastic deformation on the mechanism of intergranular crack initiation at the scale of theindividual grains. An experimental procedure was set up to monitor the evolution of sur-face strain fields on in situ tested SEM notched specimens using digital image correlationtechniques. In addition, measurements of the associated crystal orientation evolution atthe micron scale were carried out using electron backscatter diffraction (EBSD). These mea-surements were then compared with finite element predictions of the local strain fields onthe observed regions of the in situ specimen. The numerical predictions were obtainedusing a dislocation mechanics-based crystal plasticity model to describe the constitutivebehaviour of each individual grain. The crystallographic grain orientations of the regionof interest were discretised for the finite element analyses from EBSD maps. From thisstudy, it was found that intergranular cracking due to hydrogen embrittlement in the Alalloy is locally triggered by high tensile grain boundary tractions, here estimated to be170 ± 35 MPa. As importantly, the results also revealed that the conditions needed for grainboundary microcracks to initiate are greatly affected by the deformation of neighbouringgrains: i.e. it was established that boundaries between two ‘‘hard’’ grains, inside a neigh-bourhood of ‘‘softer’’ deformed grains, are the first to fail.

� 2012 Elsevier Ltd. All rights reserved.

1. Introduction

Aluminium alloys, strengthened by elements in solid solution, can be sensitive to intergranular stress corrosion cracking(ISCC) in some specific microstructural states (Hollinsworth and Hunsicker, 1987). Initially, the mechanism thought to leadto ISCC was understood to be stress assisted localised dissolution (Dix, 1940). It was supported by microstructural observa-tions showing that Mg-rich (Al3Mg2 b-phase) precipitates form at grain boundaries during ageing. Here, locally enhanced Mgconcentration was thought to create an electrochemical potential inhomogeneity that induced the anodic dissolution of theprecipitate and, in so-doing, creating a preferential crack path. This understanding prevailed till the early seventies (Gest andTroiano, 1974) when the crucial role of hydrogen in the fracture process was identified. Then, it was found that when a cer-tain amout of hydrogen is absorbed from the environment into materials such as steels, it can then diffuse along grainboundaries and eventually lead to their embrittlement.

0749-6419/$ - see front matter � 2012 Elsevier Ltd. All rights reserved.doi:10.1016/j.ijplas.2012.01.004

⇑ Corresponding author. Tel.: +33 1 60 76 30 01; fax: +33 1 60 76 31 50.E-mail address: [email protected] (E.P. Busso).

International Journal of Plasticity 34 (2012) 139–153

Contents lists available at SciVerse ScienceDirect

International Journal of Plasticity

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

Page 3: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

Initially, the effect of hydrogen contamination in Al alloys was ignored due to the vanishingly small solubility of hydrogenwhen the metal is in contact with the H2 gas. Nevertheless, it was shown that a hydrogen flux can develop through Al–Zn–Mg alloys (Gest and Troiano, 1974) even with a free corrosion potential. It was shown unambiguously later on (Gruhl, 1984)that hydrogen can diffuse over depths of some millimeters toward regions of high stress triaxialities (e.g. notch root region),be trapped at grain boundaries and induce intergranular fracture without any contribution of corrosion processes which maybe active at the crack tip. The precise nature of ‘‘brittle’’ fracture, i.e. decohesion vs intense localised plasticity, is still debatednowadays. In the specific case of Al–Zn–Mg, TEM studies by Scamans et al. (1976) and Scamans (1978)) show that it is a trueintergranular decohesion mechanism (see also Malis and Chaturvedi, 1982). No evidence of grain boundary voids, associatedwith a ductile-type failure mechanism, has been linked to hydrogen embrittlement in Al alloys.

The b intergranular precipitates play multiple roles in the stress corrosion cracking (SCC) of Al–Mg alloys. It was shown(Jones et al., 2001, 2004) that they dissolve actively during SCC and that enough hydrogen must have been produced to embrit-tle the grain boundary in between successive b particles (Jones and Danielson, 2003). Furthermore, pitting tests (Tanguy et al.,2002) followed by a tensile deformation in air showed that the anodic dissolution of b particles at triple junctions released en-ough hydrogen to embrittle precipitate-free grain boundary facets. Moreover, the observation of hydrogen bubbles in Al–8%Mg(Ben Ali et al., 2011) suggests that b-precipitates must indeed be preferential sites for hydrogen trapping. Such understandingand the results of the numerical simulation of the intergranular fracture process (Malis and Chaturvedi, 1982) revealed thatintergranular fracture is mostly dominated by the damage of the inter-precipitate regions. The role of plasticity in ISCC isknown to be important even though it has only been studied macroscopicaly. For instance, Tanguy et al. (2002) showed thatpre-straining of an Al–5%Mg alloy increases its sensitivity to ISCC under both tensile and fatigue loading conditions.

Hydrogen induced embrittlement (HIE) processes can be studied experimentally using local strain measurements (Sacht-leber et al., 2002). However, even though as more accurate local measurement techniques become available, it still remainsdifficult to measure intragranular strains in fine grained materials. For such cases, large focal lenses and microdiffractiontomography (Ludwig et al., 2009) techniques are not suitable. Fortunately, plastic strain fields can nowadays be more easilymeasured at the micrometrer scale thanks to techniques which rely on fine grids or speckles spaced out at distances consid-erably smaller than the grain size. Examples are those based on local pitting corrosion, photolithography (Schroeter andMcDowell, 2003; Liu and Fischer, 1997), micro-lithography (Allais et al., 1994; Biery et al., 2001), and digital image correla-tion techniques (DIC) (Allais et al., 1994; Tatschl and Kolednik, 2003; Hild et al., 2002; Sachtleber et al., 2002).

Even when the measurement of local strains is possible in small grained materials, numerical techniques are needed toidentify the local stress fields. Intergranular stresses are known to play a significant role in intergranular fracture and resultfrom complex phenomena (e.g. strain incompatibilities between adjacent grains) in a variety of materials such as Zn (Parisotet al., 2000), TiAl (Simkin et al., 1999) and duplex stainless steels (Marrow, 1996). The mechanical response of a polycrys-talline aggregate can in principle be predicted by using either mean-field homogenisation or finite element (FE) techniques.However, when the spatial local distributions of strain and stress heterogeneities at the level of each individual grain areneeded, then full-field computations are required. The finite element method has been used to model the intragranular re-sponse of polycrystalline materials under various macroscopic loading conditions, such as homogeneous (Becker and Pan-chanadeeswaran, 1995) and fretting (Goh et al., 2006) loading, and to study crack initiation and growth using methodssuch as those based on cohesive zone models (e.g. Clayton and McDowell, 2004). Crystal plasticity constitutive formulationshave been based on semi-phenomenological (Asaro, 1983; Barbe et al., 2001a), and dislocations mechanics based local(Busso and McClintock, 1996; Harder, 1999) and non-local (Busso et al., 2000; Cheong et al., 2004) approaches. They are alsoable to account in some cases for, e.g., statistical effects of grain boundaries and lattice misorientations on local stress andlattice energy density distributions (Bieler et al., 2009; Clayton and McDowell, 2003), and twinning as a deformationmechanism (Staroselsky and Anand, 1998).

The main objective of this work is to understand the effect of local grain plasticity on the mechanism of intergranularcrack initiation due to hydrogen induced embrittlement in an Al–5Mg alloy. To that purpose, notched specimens were locallycharged with hydrogen and then tested under tension in situ within a scanning electron microscope. A combination of in situ2D strain field and EBSD measurements, and finite element simulations of the observed specimen regions using a dislocationmechanics-based crystal plasticity model was relied upon to understand the local conditions responsible for intergranularcrack initiation.

2. Experimental investigations

2.1. Material

The Al–Mg alloy studied, known as AA 5083, was thermomechanically treated to obtain equiaxed grains. Its chemicalcomposition is given in Table 1. Charging Al alloys with hydrogen is problematic because of the barrier effect of the oxide

Table 1Chemical composition of the AA 5083 alloy.

Element Si Fe Cu Mn Mg Zn Ti Cr Al

wt.% 0.40 0.40 0.10 0.70 4.50 0.25 0.15 0.25 Balance

140 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 4: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

layer. For a discussion of these issues see, e.g., Jones and Danielson (2003), Buckley and Birnbaum (2002), Ben Ali et al.(2011). The presence of the b-phase at the grain boundary was necessary for the charging procedure to induce embrittlementover a significant depth. The procedure consists of a cathodic polarisation at a current density of �1 mA/cm2, during 12 h in adeaerated solution containing AlCl3 with a concentration of 120 g/l (with addition of H2SO4 and saturated with boric acid)and a pH 2. Embrittlement was over at least one grain boundary facet when b precipitates were present and less than 1 lmwhen they were not present. Note that longer charging times may damage the surface and not lead to an embrittlement. Inorder to prevent hydrogen desorption, specimens were stored in liquid nitrogen until the mechanical tests were performed.Preliminary uniaxial tensile tests, conducted up to a 10% elongation, confirmed that HIE was effective up to a depth of 5 lmand that intergranular cracking was observed on grain boundaries perpendicular to the tensile direction (see Fig. 1) after afew percent of plastic strain. Embrittlement was fully reversible by heat treatment.

2.2. Room temperature tensile tests and strain field measurements

Experimental and numerical investigations were performed using two different types of 1 mm-thick specimens, see Fig. 2.The smooth specimen was relied upon to calibrate the crystal plasticity model, and the specimen with two symmetricalnotches to localise a priori the deformation and intergranular fracture events in a region small enough to be suitable forthe local strain and grain orientation measurements. Note that in Fig. 2, the plotted stress is the axial component on the spec-imens, calculated as the applied axial load divided by the smallest specimen cross-section, rather than the Mises equivalentstress. Consequently, one can see the strengthening effect of the notch due to the higher local triaxiality in the notch regionwith respect to that in the smooth specimen.

After the surface preparation, the contour of the specimen zone to be studied was defined by microindentation marks, seeFig. 3(a). These marks were also used as references for optical extensometry in order to confirm the average measured strainobtained with the digital image correlation (DIC) technique.

The use of both EBSD analysis and DIC techniques required a specific surface finishing. After machining, specimens wererecrystallised at 460 �C for one hour and water-quenched to the keep Mg in solid solution. Then a mechanical polish wasperformed down to 1 lm in order to remove any surface damage. The specimens were then tempered at 150 �C for 7 daysin order to precipitate the Al–Mg b-phase and relieve any residual stresses which may have been introduced due to themechanical polishing. A final polishing step from 1 lm down to 1/4 lm was carried out before polishing chemically the spec-imen with colloidal silica to optimise the surface condition for the EBSD scanning (see Fig. 4(a)). After the EBSD mapping, theregion of interest was precharged with hydrogen following the procedure described previously. To be able to measure thestrain heterogeneities inside the grains, the region of interest on the specimen surface had to be patterned with a specificspeckle. However, to obtain accurate results by DIC at this scale, surface heterogeneities should be less than 200 nm in size.Slight corrosion pits, i.e. the smallest white features in Fig. 4(b), observed after hydrogen charging, were kept and employedas surface heterogeneities for contrast enhancement purposes.

A 450 � 300 lm area was scanned with EBSD, the result of which can be seen in Fig. 5. EBSD maps were colour codedaccording to the orientation of the tensile direction (x) of the specimen with respect to the crystal axes, as depicted in

Hydrogen charged No hydrogen charged

200 µm

y

x

Directionof loading

Fig. 1. Optical micrograph of intergranular fracture after 10% axial strain along the x-axis.

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 141

Page 5: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

Fig. 5. In the present study, a ‘grain’ was defined from the EBSD maps as a crystal of at least 6 lm in size, surrounded byboundaries with misorientation angles of at least 5�. In Fig. 5, the mean grain size was found to be 27.0 lm with a standarddeviation of 8.6 lm. The distribution of misorientations between grains was found to be in the range of 10–60�, without anyevidence of texture observed. The selected region in Fig. 5 is composed of 458 textureless grains with a total boundary lengthof �25 cm. This region was assumed to be representative and sufficiently large to conduct the study. Even though 10 spec-imens were tested, only the results of one of them will be reported as they were found to be representative of all the otherones obtained in this study.

Tensile specimens were mounted on a mini-tensile testing machine in a Leo SEM 1450VP vacuum chamber. The machinehas a dimension of 502 � 327 � 250 mm, a total screw travel of 150 mm, and a maximum load capacity of 5000 N. A com-puter integrated stepping motor control was used to carry out the tensile test at an imposed displacement rate of 0.36 mm/min. The microindented decorated specimen surface was imaged under both secondary electron (SE) and backscatter elec-tron (BSE) imaging modes and a set of 8-bit grey scale digital images was obtained. The specimen was then loaded up to a

Fig. 2. Measured and predicted tensile stress–strain curves and specimen geometries (dimensions in mm).

x

y

(a)

(b)

(c)

(d)

(e)

250 µm

Fig. 3. Double notched specimen region of interest: (a) region delimited by micro-indentation (black diamond symbols) marks used to calculate the averagestrains to be applied on the FE model, (b) region analysed by EBSD, and SEM image acquisition regions using (c) low, (d) medium, and (e) high resolutions,respectively.

142 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 6: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

total axial strain of 5%. A series of digital images within the microindentation decorated surface were acquired in situ at sev-eral applied strain levels in order to record the surface strain field evolution and to identify intergranular crack initiationevents.

Fig. 4. Optimum surface preparation for (a) EBSD (before hydrogen charging), and (b) DIC (after hydrogen charging) with a superimposed correlation grid of32 � 32 pixels.

Fig. 5. Inverse pole figure map of the region studied by EBSD. The y-axis represents the specimen tensile axis and the zone delineated in the black box wasthat modelled for the FE calculations.

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 143

Page 7: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

An optimised digital image correlation technique was used to compute surface strain fields from the recorded images usingthe software CorreliQ4GD (Hild et al., 1999, 2002). The field of view of the digital images was approximately 850 � 600,300 � 225, and 150 � 112 lm2 for the regions (c)(d)(e) identified in Fig. 3, respectively. These three scales of correlation pro-vided a range of fine-scale strain field measurements as well as the boundary conditions to be applied to the crystal aggregateregion used in the modelling part of the study. The spatial resolution of the system is 32 � 32 pixels, with an average displace-ment resolution of at least a pixel size. The pixel size, which determines the minimum spatial resolution for each magnifica-tion, is therefore 0.5, 0.12 and 0.07 lm for the large, medium and low magnification images, respectively. With this methodand with the level of applied macroscopic strains (0.5% up to 3%), the relative computed error in the measured strains is lessthan 3%. The use of three scales of correlation allowed a link between microscopic and macroscopic scales. The local strainfields were obtained from the DIC data by strain averaging over the smallest area investigated, viz. (e) in Fig. 3.

After the test, the deformed specimens were removed from the mini-tensile test stage and analysed again by EBSD toidentify the grain orientations after deformation. Due to the surface roughening caused by plastic deformation, the speci-mens were polished to improve the quality of EBSD patterns by removing a layer of approximately 0.2 lm.

3. Model of the polycrystal region

The dislocation-mechanics based rate-dependent crystal plasticity model for FCC materials proposed by Cheong and Bus-so (2006) was used in the FE simulations of the in situ test to compute the strain and stress fields. This formulation will beoutlined next.

3.1. Crystal plasticity model

The rate-dependent single crystal constitutive model is formulated following finite-strain kinematics. It relies in the clas-sical multiplicative decomposition of the total deformation gradient F,

F ¼ FeFp ð1Þ

into a plastic, Fp, associated with crystallographic slip, and an elastic component, Fe, which accounts for the elastic stretchingand rotation of the crystal lattice. The velocity gradient of plastic slip is given by

Lp ¼ _Fp _Fp�1 ¼XN

a¼1

_cama � na ð2Þ

Here, the sum extends over N active slip systems, _ca represents the crystallographic slip rate on a generic slip system (a), andma � na the dyadic product between the slip direction (ma) and associated slip plane normal (na).

The slip rate, _ca, is assumed to be dominated by the thermally activated glide of dislocations over obstacles (i.e. mainlyforest dislocations as the heat treatment used in the alloy of interest did not lead to full intergranular precipitation of Mg(Tanguy et al., 2002)). The slip rate is related to the resolved shear stress, sa, through the exponential function proposedby Busso and McClintock (1996) and Busso et al. (2000),

_ca ¼ _c0 exp � F0

kh1� jsaj � Sa

c l=l0

s

� �p( )q" #

sgnðsaÞ ð3Þ

which accounts for the absolute temperature (h, K) and the stress dependence of the activation energy. In Eq. (3), F0 repre-sents the Helmholtz free energy of activation at 0 K, k the Boltzmann constant, _c0 a reference slip rate and, s, the maximumglide resistance at which dislocations can be mobilised without thermal activation. Furthermore, l and l0 are the shearmoduli at h and 0 K respectively. The exponents p and q describe the shape of the energy barrier vs. stress profile associatedwith interactions between dislocations and obstacles. The main contributions to the slip resistance to plastic flow, Sa

c , aredue to the friction stress induced by the Mg atoms in solid solution, Sa

ss, and to the athermal slip resistance SaT . As the slip

resistance contributions are assumed to be additive, then

Sac ¼ Sa

ss þ SaT ð4Þ

Since only a very small part of the Mg atoms precipitate during the heat treatment, the concentration of Mg in solid solu-tion is set to be equal to the average concentration in the material. Thus the friction stress due to Mg atoms in solid solutionis calculated based on the atomic size and concentration of Mg in the alloy (see Table 1), as proposed by Saada (1968). Theathermal slip resistance Sa

T is expressed as:

SaT ¼ klba

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiXN

b¼1

habqbT

!vuut ð5Þ

where the overall dislocation density for a given slip system b, qbT , is obtained from a discretisation of the dislocation struc-

ture into pure edge and pure screw types, of densities qbse

and qbss

, respectively. Thus,

144 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 8: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

qbT ¼ qb

seþ qb

ssð6Þ

In Eq. (5), k is a statistical constant and hab represents the dislocation interaction matrix defined as

hab ¼ x1 þ ð1�x2Þdab ð7Þ

where x1 and x2 are two cross-hardening constants and dab the Kronecker delta. The evolutionary equations of the individ-ual dislocation densities account for the competing dislocation storage-dynamic recovery processes in FCC metals and areexpressed as (Cheong et al., 2004; Cheong and Busso, 2006),

_qae ¼

Ce

ba Ke

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiXN

b¼1

qbT

vuut � 2deqae

24 35j _caj ð8Þ

_qas ¼

Cs

ba Ks

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiXN

b¼1

qbT

vuut � qas pd2

s Ks

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiXN

b¼1

qbT

vuut þ 2ds

0@ 1A24 35j _caj ð9Þ

Here, the parameters Ce and Cs describe the relative contributions to the overall slip from edge and screw dislocations, whileKe and Ks are mobility constants associated with their respective mean free paths. Recovery processes are associated with theparameters de and ds, which represent critical annihilation distances between dislocations of opposite Burgers vectors forboth edge and screw types. The theory is completed by considering the stress–strain relation,

T ¼ C : Ee ð10Þ

where T is the lattice-based second Piola–Kirchhoff stress tensor, C is the anisotropic elasticity tensor, while Ee = [FeTFe � I]/2is the Green–Lagrange strain tensor, with I, the second-order identity tensor. Furthermore, T can be expressed in terms of theCauchy stress tensor, r, as

T ¼ ðdet FeÞFe�1rFe�T ð11Þ

Details of the numerical implementation of the above constitutive theory into the finite element method can be seen inBusso et al. (2000) for an implicit implementation and in Abrivard (2009) for an explicit one.

The values of the model’s parameters are given in Table 2. They are inspired by those calibrated for pure aluminium inCheong and Busso (2006), except for the additional term which accounts for the Mg solid solution effect. A comparison be-tween the uniaxial stress–strain tensile curve obtained from the smooth specimen and that predicted using a 100 grainaggregate and the single crystal model is shown in Fig. 2.

3.2. FE mesh and boundary conditions of the region studied by EBSD

To mesh automatically the region of interest, namely Region (b) in Fig. 3, an arbitrary grain colour map was first plotted.To this aim, grains were defined as sets of at least 10 pixels in size with a misorientation between neighbouring pixels of atmost 5�. From this map, triple junctions were first identified using an in-house developed image processing tool. Then, grainboundaries were revealed using both triple junctions and colour gradient detection. The grain boundary discretisation wascarried out by dividing them into small straight segments while taking their local curvatures into account (Fig. 6(b)). Finally,grains were meshed using 6-node generalised plane strain triangular elements to allow a proper geometric description ofgrain boundaries. Note that generalised plane strain boundary conditions imply that the out-of-plane strain (�z) is uniformon the (x,y) plane. For the free out-of-plane (z) surfaces of interest in the specimen, it implies that �z – 0 and that, on average,rz = 0. Such assumption constitutes the best 2D approximation to a full 3D model of the polycrystal. Other approaches rely-ing on square meshing based on EBSD scan step or on correlation microgrids were not deemed to be appropriate as regularmeshing can result on an over estimate of the strains and stresses at grain boundaries (e.g., see Barbe et al. (2001a)). Here, amesh composed of 15,230 elements was found to be suitable to account for the spatial resolution of the EBSD measurements,the description of grain boundary morphologies, and to yield acceptable computational times.

Table 2Single crystal model parameters.

Elastic constants (GPa) Flow rule (Eq. (3)) Evolutionary laws (Eqs. (5), (8) and (9))

C11 = 112.9 F0 = 3 � 10�19 J b = 2.57 nmC22 = 66.4 bs ¼ 8:0 MPa de = 7.0nm; dsw = 35 nmC33 = 27.9 Sa

ss ¼ 42:7 MPa x1 = 1.5; x2 = 1.2l0 = 49 _c0 ¼ 1:73� 106 s�1 Ce = Csw = 0.5

l = 45 p = 0.141 Ksw = 2Ke; Ke = 15.7 � 10�3

q = 1.10 k = 0.3

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 145

Page 9: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

When comparing numerical and experimental results, the influence of the neighbouring grains on the simulated micro-structure must be taken into account, since they can affect the mechanical behaviour from as far away as 5–10 times theaverage grain size (Doumalin et al., 2003). In order to take neighbouring grains into account, the measured displacementsat the fringes of the meshed region, whose locations are defined by the microindentations shown in Fig. 3, were interpolatedand applied as boundary conditions.

4. Results and discussion

4.1. Strain field measurements and intergranular cracking observations

Fig. 7 shows regions oriented 90� and 45� to the loading direction where deformation is localised, as also reported else-where at similar scales (Barbe et al., 2001a,b). The approximate width of the more highly deformed regions is 80 lm (i.e. 3times the average grain size). Microcracks were found within these deformed regions, which suggests that crack initiationrequires a certain minimum average level of deformation. Cracks were only seen to nucleate at grain boundaries oriented

Fig. 6. (a) Grain colour map of the zone studied by EBSD, and (b) corresponding FE mesh using 15,230 quadratic triangular generalised plane strainelements.

Fig. 7. Measured DIC axial strain field after 0.0045 macroscopic strain along the x-axis. Black arrows indicate the grain boundaries where crack initiationwas observed.

146 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 10: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

normal to the tensile direction. Thus, intergranular crack initiation seems to be related to both average strain level and grainboundary orientation.

The strain field was then studied at a finer scale (see upper half of Region (d) in Fig. 3) to investigate the heterogeneity ofthe deformation within the more highly deformed regions. The measured strain distributions at medium magnifications areshown in Fig. 8 for macroscopic applied strains of (a) 0.5%, (b) 1.6%, (c) 2.6%. In Fig. 8(d), a SEM micrograph taken after 4.5%strain is shown for comparison. Here, white arrows point out strongly deformed regions and the black ones cracked grainboundaries. These results reveal that microcracks appear only after a small amount of macroscopic strain, between grainswhich are relatively less deformed compared to their neighbours (less than 1% in Fig. 8(a)). To confirm these results, the levelof lattice rotation due to plastic deformation was investigated within each individual grain. Fig. 9 shows the crystal orien-tation evolution of Grains 1, 2 and 3 identified in Fig. 8 before and after deformation. It can be seen that Grains 2 and 3 under-go a very small amount of rotation, whereas Grain 1 rotates considerably more towards the [101] orientation. These EBSDresults are in good agreement with the fact that Grains 2 and 3 are less deformed than Grain 1, as revealed by Fig. 8.

These observations also reveal that grain boundaries between undeformed or only slightly deformed grains do not nec-essarily crack. This implies that, in order to describe the cracking phenomena correctly, grains neighbouring the failed grainboundaries need to be taken into account. The DIC results also showed that grains surrounding the failed grain boundariesdeformed more than those where cracks were found. As it will be discussed below, the numerically predicted strain fields atsmaller scales confirm this experimental observation.

In the region where detailed EBSD analyses were conducted (see Fig. 5), 12 grain boundaries were identified as havingfailed during the test. Furthermore, 11 out of the 12 were found to be partly located in the vicinity of highly deformed re-gions, and between grains which had not undergone a great deal of deformation. This is illustrated in Fig. 10 which gives anschematics of (a) the corresponding grain configuration and (b) an idealised configuration to be used in the model presentedbelow. Thus, the following preliminary conclusions can be made: micro-cracks seem to initiate at grain boundaries (i) con-tained within a strongly deformed region, (ii) oriented normal to the loading direction, and (iii) located at the boundary be-tween two ‘‘hard’’ grains which are, in turn, contained within a larger ‘‘softer’’ region.

0.00 0.01 0.02 0.03 0.04 0.05Exx

0.00 0.0025 0.005 0.0075 0.01Exx

0.00 0.01 0.02 0.03 0.04 0.05Exx

10 µm

x

y

(a)

(c) (d)

(b)

C

A

B C

A

B

C

A

B C

A

B

Fig. 8. Measured strain distributions at medium magnification along the x-direction for macroscopic applied strains of (a) 0.5%, (b) 1.6% and (c) 2.6%, and (d)SEM image taken after 4.5% applied strain. White arrows indicate strongly deformed regions and the black ones cracked grain boundaries.

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 147

Page 11: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

A failure criterion for intergranular cracking can also be proposed by considering that the grains at either side of the failedgrain boundary develop high normal tractions. Typically, the critical value of these tractions were seen to be attained wherethe local inelastic strain reaches approximately 0.33%. Such low value of plastic deformation indicate that those grain bound-aries fail in a brittle manner. It is assumed here that crack initiation is controlled by a critical normal grain boundary traction,henceforth denoted rd.

4.2. Crack initiation criteria

4.2.1. Effect of strain incompatibility across grain boundariesAs discussed above, grain boundary cracking seems to be driven by local grain boundary tractions arising from strain

incompatibilities between ‘‘harder’’ and ‘‘softer’’ grains. Furthermore, cracking events could not be related to either a stronggradient in the Schmid factor (assuming grains with uniform lattice orientations and stresses, as in a Sachs assumption) or toa strong strain gradient across boundaries (as can seen in Fig. 8). In addition, a local study of misorientation line profiles be-fore and after deformation, across both ‘‘sound’’ and ‘‘cracked’’ grain boundaries, did not reveal any significant role of localmisorientations either near (i.e. local lattice rotations due to plasticity) or across grain boundaries as it is sometimes ob-served in intergranular fracture (Gourgues, 2002). In can be said that our findings indicate that strain incompatibilities atgrain boundaries are not sufficient to explain intergranular fracture in the present case. To address this issue and to obtaina detailed local description of both strain and stress fields, FE calculations were carried out and the corresponding results areshown next.

4.2.2. Numerical predictions of local stress and strain fieldsFig. 11(a) and (b) shows the measured and predicted strain fields in Region (e) of Fig. 3 after 0.45% applied macroscopic

strain (which corresponds to Point A in Fig. 2), respectively. Fig. 11(c) is a SEM micrograph of Fig. 11(a)’s region and Fig. 11(d)gives the predicted Mises stresses at the time when the first grain boundary failure was observed (also at Fig. 2’s Point A).

It should be noted that at the locations 1, 2 and 3 indicated in Fig. 11, the high levels of axial strain measured by the DICsystem are to a great extent different from the real values due to the distortion introduced by the opening of the cracks dur-ing the digital image correlation measurements. Note also that the spatial resolution of the scanning electron microscopyimage was not high enough to detect cracks for applied strains lower than 0.1%.

Fig. 9. Comparison between the inverse pole figures for (a) and (d) Grains A, (b) and (e) Grain B, and (c) and (f) Grain C identified in Fig. 8(c) beforedeformation (a–c) and after (d–f) 4.5% applied strain.

148 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 12: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

Fig. 10. Schematics of the typical grain configuration observed around a cracked grain boundary (indicated by the dashed line): (a) polycrystalline regionand (b) idealised configuration to be used in the simplified model. The dashed line delimits the boundary that is assumed to fail.

Fig. 11. (a) Measured and (b) predicted axial strain fields after 0.45% applied macroscopic strain, (c) micrograph of the same region as (a) and (b) at the endof the test (4.5% applied strain), and (d) predicted Mises stresses after 0.45% applied strain. (Loading direction is parallel to the x-axis).

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 149

Page 13: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

Fig. 12 shows both the predicted and measured axial strains along the two dashed lines in Fig. 11(a) and (b). Note that thevery high value of the measured axial strain at the crack location (distance = 18 lm in Fig. 12) is due to crack opening effectsrather than to strain localisation. It can be seen that the model predictions along the vertical line are in satisfactory agree-ment generally with the measurements. However, the model over estimates the local axial strains along the horizontal lineat a distance greater than 60 lm (that is, greater than 60 lm from Fig. 11(a) and (b)’s left border), as it is the case over a 4 lmlong region close to the crack at distances of between 10 and 14 lm. The latter is understood to be due to the stress redis-tribution which follows the crack initiation event.

At d = 0 (both vertical and horizontal lines) and d = 60lm (vertical line), the model predictions and experimental mea-surements are close as expected since the experimental measurements were prescribed as the finite element model’s bound-ary conditions. However, the predictions and measurements do not coincide for the horizontal line at d = 68lm due to thefact that the size of the region studied by digital image correlation was smaller than that discretised with finite elements.Thus, here the boundary conditions were prescribed slightly further away in the numerical model. Nevertheless, the effectintroduced by this extrapolation of the boundary conditions was found to be only local and it can be seen that the axial strainfield is reasonably well predicted close to the boundary which is the first to fail. The predicted local stress fields around grainboundaries just before crack initiation will be discussed next.

4.2.3. Simplified stress based grain boundary failure criterionA first approximation to the normal tractions needed to fracture a given grain boundary, Rd, can be obtained by solving an

Eshelby problem consisting of a ‘‘hard’’ spherical elastic inclusion embedded in a ‘‘softer’’ infinite elasto-viscoplastic matrix(as illustrated in Fig. 10(b)). It will be assumed that the grain boundary belongs to two ‘‘hard’’ grains which behave elasti-cally, very much as the grain boundary in Fig. 10(b)’s elastic inclusion. Under this assumption, it can be shown that the solu-tion to this problem yields uniform stresses within the inclusion, and that the maximum principal stress is given by(Beremin, 1981),

rxx ¼ Rxx þ kðReq � ryÞ ð12Þ

where rxx is the average axial stress in the polycrystal at the moment the first grain boundary fails. This value can be approx-imated from Fig. 2’s Point A, which yields Rxx = 120 MPa. Also, ry is the material yield stress (=85 MPa), and Req the averageMises stress in the inclusion. From the FE calculations, the average von Mises stress in Grains B and C (see Fig. 11(d)) wasfound to be Req = 140 MPa. Finally, k is a function of the inclusion shape – equal to 1 for the spherical particle assumed here(Pineau and François, 2001). With these values, Eq. (12) yields rxx = 175 MPa. This value will henceforth be considered to be afirst estimate of the critical grain boundary traction, rd, required to fracture a grain boundary.

4.3. Critical grain boundary traction predicted by a full FE analysis of the crystal aggregate

The predictions obtained from the previous section’s analytical model were compared to the local normal tractions ob-tained from the numerical simulations of the crystal aggregate. Here, the normal grain boundary tractions were determinednumerically by resolving the local stress tensor into a direction normal to the grain boundary.

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0 10 20 30 40 50 60 70

Exx

Distance (µm)

Crack

Data for vertical linePrediction for vertical line

Data for horizontal linePrediction for horizontal line

Fig. 12. Comparison between the measured and predicted axial strain profiles along the dashed lines of Fig. 11, which correspond to 0.45% macroscopicapplied strain.

150 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 14: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

The predicted normal tractions along some of the most highly stressed grain boundaries within the region studied by DIC,labelled 1–10 in Fig. 13(a), are shown in Fig. 13(b). Here, the tractions are expressed in terms of the distance along the grainboundary of interest from a triple point, l, normalised by the overall grain boundaries length, L. It can be seen that the normalgrain boundary tractions tend to be higher in the vicinity of triple junctions, as has also been observed elsewhere (Masayukiet al., 2007). This is to a great extent due to the complex deformation constraints imposed by the material surrounding thetriple junctions.

It is worth noting from Fig. 13(a) and (b) that, out of the 10 grain boundary regions along which their normal tractionswere calculated, only those numbered 1–3 failed during this particular test. It should also be recalled that these three failedgrain boundaries were already identified in Fig. 11(a). Thus the predicted tractions shown in Fig. 13(b) reveal that, at theonset of cracking, the tractions in Grain Boundaries 1–3 vary between an upper bound of 205 MPa to a lower bound of135 MPa. Thus, writing the critical tractions in these grain boundaries in terms of their mean value, one finds170 ± 35 MPa. This numerical prediction is close to the one obtained analytically in the previous section using a simplifiedmodel based on an extension of Eshelby’s theory, namely 175 MPa. This good correlation is encouraging as far as the suit-ability of the configuration considered in Fig. 10 to describe intergranular crack initiation due to hydrogen embrittlement isconcerned.

5. Conclusions

An in situ tensile testing procedure was set up to investigate intergranular fracture due to hydrogen embrittlement at thefree surface of an aluminium–magnesium alloy. Results obtained from a combination of the in situ tests, local EBSD analyses,

(a)

(b)Fig. 13. (a) Identification of the grain boundaries analysed, with the arrows indicating the triple junctions or origin of the grain boundary coordinate l usedin (b), which shows the computed normal grain boundary tractions along the grain boundaries labelled from 1 to 10 in (a). Grain Boundaries 1, 2 and 3 werefound to have failed.

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 151

Page 15: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

strain field measurements by digital image correlation techniques and finite element modelling have led to the followingconclusions:

� Grain boundary cracking occurs on boundaries normally oriented to the applied tensile stress when the average localaxial strain is as low as 0.45%. In addition, failed grain boundaries were found to be between grains that undergo verylimited plastic deformation despite being embedded in large localised deformation regions.� The critical local grain boundary tractions obtained from an analytical Eshelby-type approach, which was based on a

‘‘hard’’ inclusion embedded in a ‘‘softer’’ matrix, of 175 MPa is consistent with the numerical prediction of170 ± 35 MPa obtained from a finite element model of the polycrystal using crystal plasticity concepts.

The aluminium alloy investigated in the present study exhibits a high stacking fault energy, which prevents it fromannealing twinning. This behaviour strongly limits the possibility of increasing the intrinsic grain boundary resistance toenvironmentally assisted cracking through the use of grain boundary engineering. The results of this study suggests that,in such cases, grain boundary cracking sensitivity can be limited or reduced provided that in-service mechanical loading con-ditions are well known. This approach, which could be referred to as ‘‘computer-assisted texture engineering’’, should enablethe tailoring of the crystallographic texture of a material so as to lower the probability of finding clusters of ‘‘harder grains’’embedded in a softer surrounding.

To provide guidelines for an ‘‘optimal’’ targeted texture, a number of numerical simulations of polycrystal aggregates(obeying a given average texture) could be carried out in parallel and the probability of finding the above mentioned grainconfiguration could be estimated on a statistical basis. Moreover, in cases where clusters of ‘‘harder’’ grains were ‘‘fre-quently’’ found, the orientation of grain boundary planes within these clusters should also be optimised (by proper designof grain morphologies obtained after thermo-mechanical processing) to limit the normal tractions on such grain boundaries.In such an approach, crystal plasticity models appear as a key tool to help designing new metallic alloys resistant to grainboundary cracking.

Acknowledgements

This research was supported by the French National Research Agency (ANR) under contract H-Inter (06-BLAN 0231). Thecontribution of Mrs. Maria Betbeder (Centre des Matériaux) to the EBSD and SEM work is gratefully acknowledged.

References

Abrivard, G., 2009. A coupled crystal plasticity – phase field formulation to describe microstructural evolution in polycrystalline aggregates duringrecrystallisation, Doctoralt Thesis, Ecole des Mines de Paris, France.

Allais, L., Bornert, M., Bretheau, T., Caldemaison, D., 1994. Experimental characterization of the local strain field in a heterogeneous elastoplastic material.Acta Metallurgica et Materialia 42 (11), 3865–3880.

Asaro, R.J., 1983. Crystal plasticity. Journal of Applied Mechanics 50, 921–934.Barbe, F., Decker, L., Jeulin, D., Cailletaud, G., 2001a. Intergranular and intragranular behavior of polycrystalline aggregates. Part I: FE model. International

Journal of Plasticity 17 (4), 513–536.Barbe, F., Forest, S., Cailletaud, G., 2001b. Intergranular and intragranular behavior of polycrystalline aggregates. Part II: Results. International Journal of

Plasticity 17 (4), 537–566.Becker, R., Panchanadeeswaran, S., 1995. Effects of grain interactions on deformation and local texture in polycrystals. Acta Metallurgica et Materialia 43,

2701–2719.Ben Ali, N., Tanguy, D., Estevez, R., 2011. Effects of microstructure on hydrogen-induced cracking in aluminum alloys. Scripta Materialia 65, 210–213.Beremin, F.M., 1981. Cavity formation from inclusions in ductile fracture of A508 steel. Metallurgical and Materials Transactions A 12 (5), 723–731.Bieler, T., Eisenlohr, P., Roters, F., Kumar, D., Mason, D., Crimp, M., Raabe, D., 2009. The role of heterogeneous deformation on damage nucleation at grain

boundaries in single phase metals. International Journal of Plasticity 25, 1655–1683.Biery, N., De Graef, M., Pollock, T., 2001. Influence of microstructure and strain distribution on failure properties in intermetallic Ti Al-based alloys. Materials

Science and Engineering A 319, 613–617.Buckley, C., Birnbaum, H., 2002. Characterization of the charging techniques used to introduce hydrogen in aluminum. Journal of Alloys and Compounds,

649–653.Busso, E.P., McClintock, F.A., 1996. A dislocation mechanics-based crystallographic model of a B2-type intermetallic alloy. International Journal of Plasticity

12 (1), 1–28.Busso, E.P., Meissonnier, F.T., O’Dowd, N.P., 2000. Gradient-dependent deformation of two-phase single crystals. Journal of the Mechanics and Physics of

Solids 48 (11), 2333–2361.Cheong, K., Busso, E., Arsenlis, A., 2004. A study of microstructural length scale effects on the behavior of FCC polycrystals using strain gradient concepts.

International Journal of Plasticity 21, 1797–1814.Cheong, K.S., Busso, E.P., 2006. Effects of lattice misorientations on strain heterogeneities in FCC polycrystals. Journal of the Mechanics and Physics of Solids

54 (4), 671–689.Clayton, J., McDowell, D., 2003. A multiscale multiplicative decomposition for elastoplasticity of polycrystals. International Journal of Plasticity 19,

1401–1444.Clayton, J.D., McDowell, D.L., 2004. Homogenized finite elastoplasticity and damage: theory and computations. Mechanics of Materials 36 (9), 799–824.Dix, E., 1940. Transactions American Institute of Mining and Metallurgical Engineers 137, 11.Doumalin, P., Bornert, M., Crépin, J., 2003. Caractérisation de la répartition de la déformation dans les matériaux hétérogènes. Mécanique & Industries 4 (6),

607–617.Gest, R.J., Troiano, A.R., 1974. Stress corrosion and hydrogen embrittlement in an aluminum alloy. Corrosion – NACE 30, 274–279.Goh, C.H., McDowell, D.L., Neu, R.W., 2006. Plasticity in polycrystalline fretting fatigue contacts. Journal of the Mechanics and Physics of Solids 54 (2),

340–367.Gourgues, A.F., 2002. Electron backscatter diffraction and cracking. Materials Science and Technology 18 (2), 119–133.Gruhl, W., 1984. Stress corrosion cracking of high strength aluminum alloys. Zeitschrift fuer Metallkunde 75, 819.

152 E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153

Page 16: Author's personal copymatperso.mines-paristech.fr/People/esteban.busso/Papers/IJP-H-cra… · Author's personal copy A study of intergranular fracture in an aluminium alloy due to

Author's personal copy

Harder, J., 1999. A crystallographic model for the study of local deformation processes in polycrystals. International Journal of Plasticity 15 (6), 605–624.Hild, F., Perie, J.-N., Coret, M., 1999. Mesure de champs de deplacements 2D par intercorrélation d’images: CORRELI2D, Internal report 230, LMT-Cachan.

Internal report 230 LMT-Cachan.Hild, F., Raka, B., Baudequin, M., Roux, S., Cantelaube, F., 2002. Multiscale displacement field measurements of compressed mineral-wool samples by digital

image correlation. Applied Optics 41 (32), 6815–6828.Hollinsworth, E.H., Hunsicker, H., 1987. Metals Handbook, vol. 13. ASM, Metals Park, OH.Jones, R., Baer, D., Danielson, M., Vetrano, J., 2001. Role of mg in the stress corrosion cracking of an Al–Mg alloy. Metallurgical and Materials Transactions A

32, 1699–1711.Jones, R., Danielson, M., 2003. Role of hydrogen in stress corrosion cracking of low strength Al–Mg alloys. Corrosion/2003, NACE, Houston, TX, Paper 03513.Jones, R.H., Gertsman, V.Y., Vetrano, J.S., Windisch, C.F., 2004. Crack-particle interactions during intergranular stress corrosion of AA5083 as observed by

cross-section transmission electron microscopy. Scripta Materialia 50, 1355–1359.Liu, Y., Fischer, G., 1997. Local strain fields in particulate metal matrix composites characterised by object grating technique. In: Proceedings of the Metal

Matrix Composites and Physical Properties, vol. 1.Ludwig, W., King, A., Reischig, P., Herbig, M., Lauridsen, E., Schmidt, S., Proudhon, H., Forest, S., Cloetens, P., Roscoat, S., et al, 2009. New opportunities for 3D

materials science of polycrystalline materials at the micrometer lengthscale by combined use of X-ray diffraction and X-ray imaging. Materials Scienceand Engineering: A 524 (1–2), 69–76.

Malis, T., Chaturvedi, M., 1982. Grain–boundary segregation in an Al–8 wt% Mg alloy. Journal of Materials Science 17, 1479–1486.Marrow, T.J., 1996. The fracture mechanism in 475 degrC embrittled ferritic stainless steels. Fatigue & Fracture of Engineering Materials & Structures 19,

919–933.Masayuki, K., Yoshihiro, K., Takayuki, K., 2007. Three-dimensional local stress analysis on grain boundaries in polycrystalline material. International Journal

of Solids and Structures 44 (10), 3267–3277.Parisot, R., Forest, S., Gourgues, A.-F., Pineau, A., Mareuse, D., 2000. Modeling the mechanical behavior of a multicrystalline zinc coating on a hot-dip

galvanized steel sheet. Computational Materials Science 19 (1–4), 189–204.Pineau, A., François, D., 2001. Fracture of metals. Part II: Ductile fracture. In: Bouchaud, E., Jeulin, D., Prioul, C., Roux, S. (Eds.), Physical Aspects of Fracture.

Kluwer, Dordrecht.Saada, G., 1968. Limite élastique et durcissement dessolutions solides. Pont à Mousson 16, 255–269.Sachtleber, M., Zhao, Z., Raabe, D., 2002. Experimental investigation of plastic grain interaction. Materials Science and Engineering A 336 (1–2), 81–87.Scamans, G.M., 1978. Hydrogen bubbles in embrittled Al–Zn–Mg alloys. Journal of Materials Science 13, 27–36.Scamans, G.M., Alani, R., Swann, P.R., 1976. Pre-exposure embrittlement and stress corrosion failure in Al–Zn–Mg Alloys. Corrosion Science 16, 443–459.Schroeter, B.M., McDowell, D.L., 2003. Measurement of deformation fields in polycrystalline OFHC copper. International Journal of Plasticity 19 (9),

1355–1376.Simkin, B., Bieler, T.R., Crimp, M.A., 1999. In: Proceeding of the 12th International Conference on Textures of Materials. NRC Research Press, Ottawa, Canada,

pp. 529–534.Staroselsky, A., Anand, L., 1998. Inelastic deformation of polycrystalline face centered cubic materials by slip and twinning. Journal of the Mechanics and

Physics of Solids 46 (4), 671–673.Tanguy, D., Bayle, B., Dif, R., Magnin, T., 2002. Hydrogen effects during IGSCC of pure Al–5Mg alloy in NaCl media. Corrosion Science 44, 1163–1175.Tatschl, A., Kolednik, O., 2003. A new tool for the experimental characterization of micro-plasticity. Materials Science and Engineering A 339 (1–2), 265–280.

E. Pouillier et al. / International Journal of Plasticity 34 (2012) 139–153 153