European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

download European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

of 12

Transcript of European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    1/12

    Comparison of two models for anisotropic hardening and yield surface

    evolution in bcc sheet steels

    T. Clausmeyer a ,b, *, B. Svendsen c ,d

    a Institute of Mechanics, TU Dortmund University, Leonhard-Euler Strae 5, D-44227 Dortmund, Germanyb Institute of Forming Technology and Lightweight Construction, TU Dortmund University, Baroper Strae 303, D-44227 Dortmund, Germanyc Material Mechanics, RWTH Aachen University, Schinkelstrae 2, D-52062 Aachen, Germanyd Microstructure Physics and Alloy Design, Max-Planck-Institut fr Eisenforschung GmbH, Max-Planck Strae 1, D-40237 Dsseldorf, Germany

    a r t i c l e i n f o

    Article history:

    Received 23 June 2014

    Accepted 29 May 2015

    Available online 9 June 2015

    Keywords:

    Material modeling

    Cross hardening

    Yield surface

    a b s t r a c t

    The purpose of the current work is the investigation and comparison of aspects of the material behavior

    predicted by two models for anisotropic, and in particular cross, hardening in bcc sheet steels subject to

    non-proportional loading. The rst model is the modied form (Wang et al., 2008) of that due to Teo-

    dosiu and Hu (1995, 1998). In this (modied) Teodosiu-Hu model (THM), cross hardening is assumed to

    affect the yield stress and the saturation value of the back stress. The second model is due to Levkovitch

    and Svendsen (2007) and Noman et al. (2010). In the Levkovitch-Svendsen model (LSM), cross hardening

    is assumed to affect the ow anisotropy. As clearly demonstrated in a number of works applying the THM

    (e.g., Boers et al., 2010; Bouvier et al., 2005, 2003; Hiwatashi et al., 1997; Li et al., 2003; Thuillier et al.,

    2010; Wang et al., 2008) and the LSM (e.g., Clausmeyer et al., 2014, 2011b; Noman et al., 2010), both of

    these are capable of predicting the effect of cross hardening on the stress-deformation behavior observed

    experimentally in sheet steels. As shown in the current work, however, these two models differ

    signicantly in other aspects, in particular with respect to the development of the yield stress, the back

    stress, and the yield surface. For example, the THM predicts no change in the shape of the yield surface

    upon change of loading path, in contrast to the LSM and crystal plasticity modeling of bcc sheet steels(Peeters et al., 2002). On the other hand, the LSM predicts no hardening stagnation after cross hardening

    as observed in experiments, in contrast to the THM. Examples are given.

    2015 Elsevier Masson SAS. All rights reserved.

    1. Introduction

    Finite-element-based modeling and simulation of the material

    and structural behavior of sheet metal parts in various stages of

    design and manufacture is today standard. In general, one aim of

    this is to benet from the predictive capability of such simulations

    (Zienkiewicz et al., 2010). In this regard, Wagoner et al. (2013)emphasize the importance of improving material models to ac-

    count for the loading path-dependent behavior of metals during

    sheet metal forming. When subject to complex non-proportional

    loading processes such as those found in many technological ap-

    plications, a number of metals exhibit hardening behavior which is

    more complex than isotropic and kinematic hardening alone.

    Observed effects in this regard include cross hardening and hard-

    ening stagnation during orthogonal loading (e.g., tension to shear).

    Cross hardening is observed to occur for example in a number of

    steels such as austenitic fcc tube steels (e.g., SUS304: Ishikawa,

    1997; Wu, 2003), ferritic bcc tube steels (e.g., S355: Kowalewski

    and Sliwowski, 1997), multi-phase tube steels (e.g., X100:

    Shinohara et al., 2010), or ferritic bcc sheet steels (e.g., LH800:Ghosh and Backofen, 1973; Noman et al., 2010). Systematic studies

    (Bouvier et al., 2005, 2006a, 2003) of interstitial free (IF), high-

    strength low-alloyed (HSLA), transformation-induced plasticity

    (TRIP), and dual-phase (DP), sheet steels, found signicant kine-

    matic hardening, hardening stagnation, as well as cross hardening,

    the latter especially in IF sheet steels. In these investigations, the

    material was subjected to monotonic shear, reverse shear, as well as

    orthogonal tension-shear, loading. Clausmeyer et al. (2012); van

    Riel and van den Boogaard (2007); Wang et al. (2008)have docu-

    mented these effects in the IF sheet steel DC06 with the help of

    monotonic tension, reverse shear, and orthogonal tension-shear,

    * Corresponding author. Institute of Forming Technology and Lightweight

    Construction, TU Dortmund University, Baroper Strae 303, D-44227 Dortmund,

    Germany. Tel.:49 231 755 8429; fax:49 231 755 2489.E-mail address: [email protected](T. Clausmeyer).

    Contents lists available at ScienceDirect

    European Journal of Mechanics A/Solids

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c om / l o c a t e / e j m s o l

    http://dx.doi.org/10.1016/j.euromechsol.2015.05.016

    0997-7538/

    2015 Elsevier Masson SAS. All rights reserved.

    European Journal of Mechanics A/Solids 54 (2015) 120e131

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    2/12

    tests, all under plane-strain conditions. In particular, cross hard-

    ening occurs during discontinuous (e.g., tension-shear: Bouvier

    et al., 2005, 2006a, 2003) and continuous (e.g., tension-shear:

    Noman et al., 2010; van Riel and van den Boogaard, 2007; Wang

    et al., 2008) orthogonal tension-shear tests. Similar results were

    obtained by Verma et al. (2011) in a series of tension and

    compression tests on ultra-low carbon IF sheet steel in which the

    tension or compression direction changed from rolling to trans-

    verse. In these tests, cross hardening was correlated with a change

    of the tension axis. As attested to in particular by the continuous

    orthogonal tension-shear test results (Noman et al., 2010; van Riel

    and van den Boogaard, 2007; Wang et al., 2008), cross hardening is

    transient and strongly depends on the rate of transition. Its

    occurrence and strength are strongly inuenced by the particular

    path taken in stressspace in changing from one loading direction to

    another.

    Generally speaking, anisotropic hardening in sheet steels may

    be inuenced by the grain and dislocation (micro)structures. In

    particular, the former is related to the grain orientation distribution

    (texture). The inuence of texture on the hardening behavior of IF

    sheet steel was investigated by Bacroix and Hu (1995) and

    Nesterova et al. (2001a,b)using two-stage loading tests (e.g., shear

    to reverse shear, tension to shear). In particular, Bacroix and Hu(1995)concluded that, at least up to moderatestrains, the inu-

    ence of texture evolution on hardening in the specimens investi-

    gated was small compared to that of dislocation structure

    evolution. This conclusion was substantiated by later crystal plas-

    ticity modeling (e.g.,Peeters et al., 2002). Related to this are more

    recent EBSD investigations on DC06 (Boers et al., 2010; Clausmeyer

    et al., 2012), which imply that the rolling-induced texture in this

    steel does not change considerably for strains lower than 35% in

    simple tension. This may be the case in other ferritic steels (e.g.,

    LH800:Clausmeyer et al., 2012; Noman et al., 2010) as well. These

    results imply that it is sufcient to account for the effect of the

    initial (e.g., rolling) texture on the anisotropic hardening and ow

    behavior in the material model.

    Although texture evolution in this sense may be secondary,grain orientation (i.e., glide-system orientation) in relation to

    loading direction certainly inuences dislocation structure devel-

    opment. The development of certain characteristic dislocation

    structures related to cross hardening have been observed during

    quasi-static loading of mild steels such as DC06 at room tempera-

    ture (e.g., Rauch and Schmitt, 1989; Rauch and Thuillier, 1993;

    Thuillier and Rauch, 1994). These include for example dense

    dislocation wall structures. The morphology and orientation of such

    walls depends forexample on grain orientation, the type of loading,

    and the loading direction in relation to the grain orientation (e.g.,

    Clausmeyer et al., 2012; Nesterova et al., 2001a,b; Thuillier and

    Rauch, 1994). A change in loading direction or type activates new

    glide systems for which existing walls act initially as obstacles,

    resulting in cross hardening.One of the rst phenomenological models accounting in

    particular for cross hardening is the Teodosiu-Hu model (THM: e.g.,

    Hu et al., 1992; Teodosiu and Hu, 1995, 1998). In the THM, cross

    hardening is assumed to affect the yield stress sY in the yield

    function fY. The THM has been employed in a number of works

    (e.g., Bouvier et al., 2005, 2003; Haddadi et al., 2006; Hiwatashi

    et al., 1997; Li et al., 2003; Thuillier et al., 2010) to model aniso-

    tropic ow and hardening behavior in sheet metals. This has

    motivated similar work on models for anisotropic hardening in the

    continuum (Barlat et al., 2013; Butuc et al., 2011; Carvalho Resende

    et al., 2013; Clausmeyer et al., 2014; Pietryga et al., 2012; Shi and

    Mosler, 2012; Tarigopula et al., 2008, 2009) and crystal plastic

    (Peeters et al., 2002; Viatkina et al., 2007) contexts. More recently,

    the THM has been modi

    ed, extended and generalized to deal with

    arbitrary changes of loading path by Wang et al. (2008). This

    modied version of the THM is that considered in the current work.

    A second model for cross hardening was introduced by

    Levkovitch and Svendsen (2007) and Noman et al. (2010). This

    model has been referred to byShi and Mosler (2012)as the Lev-

    kovitcheSvendsen model (LSM), whodiscussed related models for

    distortional hardening and the strength differential effect in mag-

    nesium alloys. In the LSM, cross hardening is assumed to inuence

    the ow anisotropy through the corresponding tensor A deter-

    mining fY. Common to both the THM and the LSM is the consti-

    tutive form

    fYffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiM X$AM X

    p sY (1)

    forfYwith respect to the intermediate (local) conguration. Here,

    Mis the Mandel stress (e.g., Mandel, 1971, 1974), and Xis the back

    stress. To be more precise, in the THM,sYis assumed to depend on

    both isotropic and cross hardening, and A is assumed constant. On

    the other hand, in the LSM, cross hardening is assumed to inuence

    the evolution ofA, andsYis assumed to depend only on isotropic

    hardening. As shown in the previous works discussed above, both

    models are capable of quantitatively predicting experimentally

    observed cross hardening. The question arises as to how the THMand the LSM compare in other respects. To this end, in the current

    work, a direct comparison of these two has been carried out. Tothis

    end, both models have been identied from the same data set for

    the ferritic sheet steel DC06 (for comparison, the ferritic-pearlitic

    steel LH800 is also briey discussed). As the current comparison

    of the THM and the LSM shows, the two models are not equivalent

    in other respects. Among these, the prediction of yield surface

    evolution is perhaps the most prominent.

    The current work begins with a brief review of the formulation

    of the two models in Section 2. This is carried out within the

    framework of the multiplicative decomposition of the deformation

    gradient and the assumption of small elastic strain relevant to

    metal inelasticity. Again, for a meaningful comparison, the two

    models are identied in Section3using the same test data sets forthe ferritic bcc sheet steel DC06. The identied THM and LSM are

    then compared on the basis of their respective predictions for yield

    and back stress evolution in Section 4 as well as yield surface

    development in Section 5. The latter results are also compared

    qualitatively with analogous results from the crystal plasticity

    model ofPeeters et al. (2002)for IF steel. Lastly, these two models

    are compared in the context of their application to the modeling of

    sheet metal forming during non-proportional loading in Section6.

    The work ends with a summary and discussion in Section 7.

    2. Model formulation

    2.1. Notation

    In this work, Euclidean vectors (i.e., rst-order Euclidean ten-

    sors) are represented by lower-case bold italic characters a,b,; in

    particular, let i1,i2,i3represent the Cartesian basis vectors. Likewise,

    upper-case bold italic characters A,B, represent second-order

    Euclidean tensors; in particular, let I represent the second-order

    identity tensor. Such tensors are dened in this work as linear

    mappings between (three-dimensional) Euclidean vectors. In other

    words, Ab is a vector for allA and all b. Let I,A and devAA(I,A)I/3 represent the trace and deviatoric part, respectively, of any A.

    Likewise, let symA:(AAT)/2 and skwA:(AAT)/2 representthe symmetric and skew-symmetric parts, respectively, of any A.

    Fourth-order tensors are represented by upper-case calligraphic

    characters A; B;, in this work. Interpreting these as linear

    mappings between second order tensors,AB

    represents a second

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131 121

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    3/12

    order tensor for all A and all B. The dyadic product C A5BofanyA andanyB is a fourth-order tensor with Cartesian components

    Cijkl AijBkl. The scalar product of two tensors A and B of any orderis represented by A,BAijBij via the summation convention.The (Euclidean) norm of any A is dened by jAj:(A,A)1/2.

    2.2. Relations common to both models

    The formulation of both models for cross hardening and yield

    surface evolution under consideration in this work is carried out in

    the framework of the standard multiplicative decomposition

    F FEFP (2)

    of the deformation gradient F (e.g., Lee, 1969) into elastic FE and

    inelastic FP parts, respectively. Note that such a decomposition

    arises naturally in the context of the modeling ofFPas a change of

    local reference conguration (Svendsen, 2001). More precisely, the

    current formulation is based on evolution relations derived from

    (2) and the right polar decompositionFEREUEofFE in the contextof small elastic strain as follows. To begin, note that the time de-

    rivative ofF

    E FF1

    P REUE results in the relation_REUE RE _UE REUELP LREUE (3)

    between the velocity gradient L: _FF1 and its inelastic counter-partLP : _FPF1P . Multiplying (3) ontherightby U1E and on the leftby RTE R1E , we obtain

    RTE _RE _UEU1E UELPU1E RTELRE: (4)Restrict the formulation now to metals and small elastic strain

    as gauged or measured by (the magnitude of) the right logarithmic

    stretch lnUE, i.e., jlnUEj1. In this case,

    U1E

    Iln UE

    O

    ln UE

    2; (5)

    and so

    UELPU1E LP Ojln UEj; _UEU1E _ln UE Ojln UEj;

    (6)

    then hold. Here, O(jAjn) represents all terms depending explicitlyon all jAjm form n. Substituting (6) into (4), rearrangement yields

    RTE _RE _ln UE LP RTELRE Ojln UEj: (7)Lastly, since RTERE I implies

    _R

    TERE 2 sym RTE _RE 0, the

    symmetric and skew-symmetric parts of (7) then imply the (small-

    elastic-strain-based) evolution relations

    _ln UE RTEDRE DP; (8a)

    _REWRE REWP; (8b)

    to O(jlnUEj) for lnUEand the elastic rotation RE, respectively. Here,D symL is the continuum rate of deformation, DP symLP itsinelastic counterpart, W skwL the continuum spin, andWP skwLP the plastic spin. In particular, the usual associated owrule

    DP _aPvfY

    vM (9)

    is assumed in both models for DP in terms of the yield functionfYfrom(1), withaP the equivalent inelastic deformation. Since various

    investigations (e.g., Bacroix and Hu,1995; Boers et al., 2010) suggest

    that the inuence of grain orientation evolution on hardening in

    ferritic bcc sheet steels such as DC06 is not signicant in the range

    of deformation relevant to the forming processes considered in this

    work, WP is neglected here for simplicity. Consequently, the evo-

    lution relation (8b) for REdepends only on the continuum spin W,

    and as such is purely kinematic. In addition, attention is restricted

    here to forming below the forming limit, and damage or any other

    process resulting in inelastic volume changes are assumed negli-

    gible. In this case, plastic incompressibility I,DP 0 pertains, andDP devDPis purely deviatoric as usual.

    Since the elastic range and elastic strain are small, the effect of

    any elastic anisotropy on the material behavior is assumed negli-

    gible here. In this case, the isotropic form

    M kI$EEI 2m dev EE (10)

    holds for the Mandel stress to lowest order in the small elastic

    strain measure EE:lnUEfrom (5), withk the bulk modulus, andmthe shear modulus. Likewise in the context of small elastic strain,

    K REMRTE (11)

    holds between the KirchhoffKand Mandel Mstresses. In this case,

    note that the trace I,KI,M is independent of RE. Lastly, bothmodels are based on common saturation-based forms

    _r crsr r _aP; (12a)

    _X cxsxNP X _aP; (12b)

    for the evolution of isotropic and kinematic hardening, respectively.

    Here, r represent the contribution of isotropic hardening to sY. In

    addition,crand cxrepresent the respective growth rates, srand sx

    are the respective saturation values, andN

    P:DP/jDPj is the direc-tion of inelastic ow.Since the THM(e.g., Hu et al.,1992; Teodosiu and Hu,1995,1998;

    Wang et al., 2008) and the original version (e.g. Levkovitch and

    Svendsen, 2007; Noman et al., 2010) of the LSM1 are limited to

    rate-independence, the current model comparison is carried out on

    this basis. To this end, attention is restricted in this work to quasi-

    static loading rates at which the rate-dependent behavior of the

    steels in question is small. As is well-known, the rate-independent

    formulation is based in particular on the concepts of an elastic

    range, yield surface and yield conditions, resulting in the so-called

    KaresheKuhneTucker conditions: _fY 0,fY 0, fY _aP 0. In thiscontext, aP is determined as usual by the consistencycondition:_fY 0 whenfY 0. For the case of rate-dependence, of course, theyield condition generalizes to an activation condition, and the

    consistency condition no longer applies.

    2.3. THM for cross hardening

    Besides the common ones just discussed, the THM (Teodosiu

    and Hu, 1995, 1998; Wang et al., 2008) is based in particular on

    the constitutive relations

    sY sY0 rfjSj; (13a)

    1

    This model was generalized to rate-dependence recently inBarthel et al. (2013).

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131122

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    4/12

    A A 0; (13b)

    for the yield stress sYand ow anisotropy tensor A, respectively,

    appearing in (1). In (13b), A 0 is the initial value2 ofA determined

    by the initial grain structure (i.e., rolling texture). As discussed in

    the introduction, for the current case of ferritic bcc sheet metals,

    the grain (orientation) structure is not observed to vary signi-

    cantly in the loading range of interest, resulting in (13b). In (13a),sY0is the initial (i.e.,aP 0) yield stress, andfis a material constantdetermining the fraction of oriented dislocation structures

    contributing to isotropic hardening. In addition, S is a (symmetric

    traceless) fourth-order structure tensor, the magnitude of which

    represents the effective strength of (persistent) oriented disloca-

    tion structures and their effecton the level of yield. The evolution of

    S as a standard internal variable3 is modeled by the relation

    _S cd

    hpssN P

    hp hxS d

    _aP cljS l=ssjnlS l_aP (14)

    in the back-rotated frame with N P NP5NP. Here, S drepresentsthe strength of the dislocation structures (e.g., walls) associated

    with the currently active slip systems, and S

    lS

    S d is thatassociated with the latent (i.e., currently inactive) slip systems.

    Note that cd (hp hx) determines the effective saturation rate, andhp ss/(hp hx) the effective saturation magnitude, associated withS d. By analogy, cljS l=ssjnl is the effective saturation rate. Thesequantities contain the material constants cd, cl, ss, and nl. hp rep-

    resents a further material function depending on S d, the direction

    NP, a further material constant np, and the symmetric traceless

    second-order tensor-valued internal variable P with constitutive

    relation

    _P cpNP P _aP (15)

    which accounts for the effect of dislocation polarization on cross

    hardening (Teodosiu and Hu, 1998; Wang et al., 2008). cp is thecorresponding saturation rate. The effect of hardening stagnation is

    accounted for via material function hx depending in particular on

    NP and X (Teodosiu and Hu, 1998; Wang et al., 2008). Lastly, the

    saturation magnitudesxofXis given by the constitutive relation

    sx x0 1 fffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffijS dj2 mjS lj2

    q (16)

    depending on the directional strength of dislocation structures.

    Here,x0represents the initial (i.e., jS dj 0 jS lj) value ofsx. Theparameter m determines the inuence of latent as opposed to

    current dislocation structures on the saturation magnitude sx of the

    back stress X.

    This model contains 21 material parameters. These include theelastic bulkk and shearm moduli, the initial yield stresssY0, the six

    parameters F, G, H, L, M, and N determining the initial ow

    anisotropy tensorA 0, the ve saturation rates cr,cx, cd,cl and cp, the

    three saturation valuessr,x0, andss, the two exponentsnlandnp, as

    well asfand m.

    2.4. LSM for cross hardening

    As just discussed, the effect of cross hardening on the material

    behavior is accounted for in the THM and (13) by the dependence of

    theyield stresssYonS . Instead of this, the effectof cross hardening

    on the material behavior is accounted for in the LSM by a depen-

    dence of the ow anisotropy tensor A on a structure tensor H

    formally analogous to S . Instead of (13), then, the LSM for cross

    hardening is based on the constitutive relations

    sY sY0 r; (17a)

    A A 0 H; (17b)

    for the yield stress and ow anisotropy tensor, respectively, in (1).

    Since the effect of cross hardening on the material behavior is then

    accounted for in the modeling (17b) of A here, that (17a) ofsYaccounts only for dislocation processes resulting in isotropic

    hardening. As discussed in detail elsewhere (Levkovitch and

    Svendsen, 2007; Noman et al., 2010), the form

    _H

    cd

    hdN P

    Hd

    _aP

    cl

    fhl

    Idev

    N P

    Hl

    g_aP (18)

    of the evolution relation for H (and so for A since A 0 is constant)

    in the LSM ismotivatedby that (14) for S in the THM. Here, IdevI 13 I5I is the fourth-order deviatoric identity. Further,Hd N P$HN P is the part ofH parallel, and Hl H Hd thatperpendicular, to N P. The idea here is that currently active dislo-

    cation structures oriented parallel to NPpersist after an orthogonal

    loading-path in inactive form (e.g., as cell-block boundaries). This

    strengthens existing obstacles to glide-system activation in the

    new loading direction.

    Comparison of (13) and (17) shows clearly the qualitative dif-

    ference between the two models in regards to the assumed effect of

    cross hardening on the material behavior. In addition, note that

    isotropic and kinematic hardening are coupled to cross hardening

    in the THM, but basically decoupled from each other in the LSM. Asit turns out, the LSM is an example of the class of models distin-

    guished by a yield function of the general form introduced in Baltov

    and Sawczuk (1965). As will be shown in detail inwhat follows, one

    result of this is that cross hardening results in a change in shape of

    the yield surface in the LSM. This is completely analogous to the

    association of isotropic hardening with a change in size, and ki-

    nematic hardening with a change in center, of this surface. Before

    delving into this, however, we rst summarize briey the identi-

    cation of the above models upon which their subsequent com-

    parison is based.

    3. Parameter identication

    For completeness, the results of the model parameter identi-cation are briey summarized and discussed in this section. For this

    study, DC06 steel sheets (DIN EN 10130) of 1.0 mm thickness sup-

    plied by ThyssenKrupp Steel were tested in uniaxial tension, plane-

    strain tension, monotonic shear, shear to reverse-shear, and plane-

    strain tension to shear. The advantage of shear tests is that larger

    strains can be achieved with common testing facilities (Bouvier

    et al., 2006b; Klepaczko et al., 1999; Miyauchi, 1987 ) than gener-

    ally possible in uniaxial tension tests. The steel DC06 is often used

    in auto-body parts4 due to its high ductility and formability. The

    mechanical tests involving shear and plane-strain deformation

    2A0 is often assumed to be of the type due to Hill (1948). Other choices (such as

    in Cazacu and Barlat, 2004; Shi and Mosler, 2012) for the representation of the

    initial ow anisotropy in non-bcc metals can be combined with the presented

    approaches for anisotropic hardening.3 In the original form of the THM due to Teodosiu and Hu (1995, 1998), S is not

    modeled as such a variable. To this end, and for the purpose of modeling continuous

    loading path changes, the model was reformulated byWang et al. (2008), resulting

    in ( 14) for _S

    .

    4 For example, see http://incar.thyssenkrupp.com/download/Broschueren/

    Tiefziehstaehle.pdf.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131 123

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    5/12

    were carried out using the biaxial testing facility at the Faculty of

    Engineering Technology, University Twente, The Netherlands. De-

    tails of the mechanical testing procedure are given elsewhere (e.g.,

    Clausmeyer et al., 2011b; Noman et al., 2010; van Riel and van den

    Boogaard, 2007). The region of the specimen subject to deforma-

    tion is initially 45.0 mm wide and 3.0 mm high. The strain eld is

    homogeneous in the range of deformation investigated. As an un-

    ambiguous and transparent measure of the state of deformation,

    the deformation gradient Fis used here; let Fij ii,Fijrepresent itsCartesian components. For the tests, the tension/rolling direction is

    oriented in the i2 direction, and the shear/transverse direction in

    the i1 direction. Consequently, in the case of plane-strain testing,

    F22 is the deformation component for tension, and F12 the defor-

    mation component for shear. Given plastic incompressibility and

    small elastic strain, the Cauchy stress TK/detF is well-approximated by the Kirchhoff stress K; let Kij ii,Kij representits Cartesian components in what follows.

    Among the material parameters, note that k,m,sY0,F,G,H,L,M,

    N,cr,sr,cx, andx0are common to both models. The Hill parameters

    F,G,H,L,M,Nare computed from average r-values (seeClausmeyer

    et al., 2011b). Ther-values are obtained from uniaxial tension tests

    performedat 0, 45and 90with respect to the rollingdirection. Inparticular, for DC06 at room temperature, we have k151 GPa,m69.6 GPa, sY0 132 MPa, F 0.252, G 0.302, H 0.698,N 1.36, L 1.5, and M 1.5. The remaining parameters valuesidentied for all models are shown inTable 1. The hardening pa-

    rameters cr,sr,cx, andx0are identied from uniaxial tension, plane-

    strain tension, shear-reverse shear, and plane-strain tension to

    simple shear, tests (Haddadi et al., 2006; Noman et al., 2010; Wang

    et al., 2008). Since for example the modeling of kinematic hard-

    ening is different in the two models, note that the corresponding

    material parameter values in Table 1 are different. In particular,sx is

    constant in the LSM and variable in the THM.

    The identied models for DC06 are compared with experi-

    mental data for the case of plane-strain tension to simple shear

    loading in Fig.1 (left). To document the applicabilityof the THMand

    the LSM to steels other than DC06, and for comparison, analogousresults for the ferritic-pearlitic steel LH800 (thickness 0.7 mm)

    from Noman et al. (2010) and Noman (2010) are also shown in Fig.1

    (right). Clearly, for both DC06 and LH800, cross hardening is

    captured well by both models; this hardening is particularly pro-

    nounced in the former material (i.e., more than 50 MPa;Fig. 1, left).

    Also more pronounced in DC06 than in LH800 is hardening stag-

    nation after cross hardening and change of loading direction (e.g.,

    from tension to shear). Such stagnation is accounted for in the THM

    (via (15) and (16)) but not in the LSM. This is the reason why, after

    loading path change and cross hardening, the THM-based result

    (red dashed curve, in the web version) inFig. 1(left) lies below the

    LSM-based one (green dashed curve, in the web version) and closer

    to the data (crosses). In any case, this difference between the two

    models is hardly surprising and certainly to be expected. As

    investigated in more detail in the next section, another difference

    between the two models regards the yield surface evolution pre-

    dicted by each, to which we turn after a more detailed analysis of

    the contribution of isotropic as well as kinematic hardening.

    4. Comparison of isotropic and kinematic hardening

    modeling

    As discussed in Section2, the relations (13a) and (17a) forsYin

    the THM and the LSM, respectively, differ in the assumed inuence

    of cross hardening via S on the former. Note that sY0 and the

    evolution relation (12a) forrare the same in both models. As well,

    the evolution relation (12b) forXis of exactly the same form in both

    the THM and the LSM. On the other hand, the saturation magnitude

    sxin (12b) is treated differently in the two models, i.e., dependent

    on S and variable in the THM via (16), and constant in the LSM.

    As an example of the consequences of these model differences,

    consider the orthogonal loading of DC06 in discontinuous tension-

    shear. In particular, this loading path consists of (i) loading in ten-

    sion to F22 10.1, (ii) unloading, and (iii) reloading in shear toF22

    1

    F12

    0.5. Consider rst the results for aP and sYfor this

    loading case displayed inFig. 2. Although the relation (12a) for the

    evolution of r is the same in both models, the coupled nature of

    model identication results in different values for the samema-

    terial coefcients in Table 1. For example, note thatcris a factor of 3

    smaller, andsris an order of magnitude larger, in the LSM than in

    the THM. This is the reason why, as shown inFig. 2 (right), isotropic

    hardening saturates much more quickly in the THM than in the

    LSM. The much lower level ofrin the THM is due of course to the

    additional contribution fromS to isotropic hardening not assumed

    in the LSM. Given the coupled natureof model identication, from a

    quantitative point of view, the contribution ofXto fYin (1) will also

    inuence the level of isotropic hardening in both models.

    Consider next the development of the components of the back

    stress

    bX :

    REXR

    TE in the current conguration during tension-

    shear loading displayed in Fig. 3. During the tension stage, theevolution of the normal componentsbX11,bX22 andbX33 is qualita-tively similar for both models. The reduction in these components

    predicted by both models during shear (re)loading after tension

    (pre)loading evident in Fig. 3 is due to the fact that the corre-

    sponding components ofNPvanish in the shear stage, resulting in a

    zero saturation level for the evolution of these components as

    determined by (12b). As is the case for DC06 in Fig. 3, due to the

    increase ofjSj (i.e., jS dj; jS lj) and its contribution to sx via (16), thelevel of kinematic hardening predicted by the THM will generally

    be higher than that of the LSM, in which sx is assumed constant.

    This together with the effect of hardening stagnation on S is also

    the reason for qualitative differences in the results forbX12 in Fig. 3in the shear stage of discontinuous tension-shear loading. Indeed,

    Table 1

    Material parameter values for DC06 determined for the THM (above) and the LSM (below) from room-temperature uniaxial tension, monotonic shear, cyclic shear, and

    orthogonal tension-shear, test data (Clausmeyer et al., 2011b). See text for details.

    THM

    parameter cr sr cx x0 cd ss cl nl cp np f m

    units MPa MPa MPa

    value 20.1 20.8 499.0 8.1 8.4 221.0 4.1 1.0 4.0 90.0 0.2 1.6

    LSM

    parameter cr sr cx sx cd hd cl hl

    units MPa MPa

    value 6.64 192.0 33.1 56.0 23.9 0.0 87.3 -0.447

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131124

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    6/12

    in the LSM (Fig. 3, right), sx is constant (e.g., 56 MPa for DC06 in

    Table 1). Consequently, the result forbX12 from the LSM has thetypical Voce-based growth-saturation form as expected from (12b).

    In contrast, as shown inFig. 3 (left), the coupling of the evolution of

    X to S via sx in the THM results in qualitative deviations of the

    result forbX12 from ideal Voce form. In particular, the effect ofhardening stagnation and the increase of sx are evident in Fig. 3

    (left). As well, in the case of DC06 at least, the much larger value

    ofcxfor the THM in Table 1results in much faster growth ofbX12inthe THM (Fig. 3, left) than in the LSM (Fig. 3, right). Note the sim-

    ilarity ofbX12 in Fig. 3(left) andK12in Fig. 1(left).5. Comparison of yield surface modeling

    In this section, the THM and the LSM identied for DC06 via the

    model parameter values inTable 1are used to model yield surface

    Fig. 1. Comparison of model and experimental results for shear stress as a function of deformation in DC06 (left) and LH800 (right: Noman et al., 2010; Noman, 2010). Both

    materials were subject to plane-strain tension in the rolling direction (up to 10% in DC06 and 13% in LH800) followed by an orthogonal loading path change to simple shear in the

    transverse direction. See text for discussion and web version for color.

    Fig. 2. Development of aP (left) and isotropic hardening (right) predicted by the THM and the LSM during discontinuous tension-shear loading. This loading path consists of (i)

    loading in tension to F2210.1, (ii) unloading, and (iii) reloading in shear to F221 F12 0.5. See text for details.

    Fig. 3. Back stress evolution predicted by the THM (left) and the LSM (right) during discontinuous tension-shear loading. Here, the material is (i) loaded in tension to F221 0.1,

    (ii) unloaded, and (iii) reloaded in shear toF221 F12 0.5. See text for details and web version for color.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131 125

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    7/12

    development in DC06. In the current model context and relation (1)

    for the yield functionfY, the yield surface represents the boundary

    fY 0 in stress space of the so-called elastic range of the material,i.e., those stress states for which fY

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    8/12

    proles (consisting of mild steel and DP steel: Haddag et al., 2007),split rings from a drawncup (consisting of DC06: Wang et al., 2008),

    or even more complex parts (e.g., channel die and S-rail consisting

    of DC06:Clausmeyer et al., 2014). Depending on factors such as the

    geometry of the workpiece, different types of loading path changes

    occur locally in the structure.

    For example, in the case of deep drawing with a cross-shaped

    die, nite-element (FE) simulations (e.g., Clausmeyer et al., 2011a)

    show that, during this process, certain regions in the sides of the

    workpiece structure experience a loading path change from plane-

    strain tension to uniaxial tension. Stress-deformation results for

    DC06 subject to such loading are shown in Fig. 6. As shown, both

    models predict a moderate increase of cross hardening with

    increasing (pre)tension. In particular, after 10% (pre)tension (Fig. 6,

    right), the level of cross hardening predicted by the THM is largerthan predicted by the LSM. This is in contrast to the tension-shear

    case inFig. 1. Although less pronounced than in the tension-shear

    case, cross hardening in the tensionetension case is also transient

    and tends to disappear with increasing loading after the loading

    path change. Consequently, for this case as well, dislocation struc-

    tures responsible for cross hardening appear to break down and

    disappear after change of loading path, i.e., as along as the loading

    path direction remains constant. The results in Fig. 6imply that the

    level of (pre)deformation during the rst stage of a multi-stage

    loading path inuences the amount of resulting cross hardening.

    As shown in Fig. 7, such a dependence is also reected in theconcomitant yield surface development. As before, the cross

    hardening evident in stress-deformation results like those in Fig. 6

    results in the LSM (Fig. 7, right), but not in the THM (Fig. 7, left), in a

    change in shape (and orientation) of the yield surface. In particular,

    as evident inFig. 7(right), the LSM predicts a rotation of the major

    axes of the (in the K11; K22 plane) elliptic yield surface away from

    the equibiaxial tension orientation towards K22 with increasing

    deformation.

    Compared to the case of plane-strain tension followed by simple

    shear (Fig. 1), the amount of cross hardening in the case of plane-

    strain tension in the rolling direction followed by uniaxial tension

    (Fig. 6) is generally much smaller (e.g., Clausmeyer et al., 2011a).

    This is correlated with the degree of orthogonality of the loading-

    path change. According to Schmitt et al. (1994), a two-stageloading path is characterized in this regard by the angle

    q arccos(NP1,NP2) determined by the direction of the rate of in-elastic ow NPiduring the rst (i 1) and second (i 2) stages ofthe two-stage path. In terms of this angle, plane-strain tension to

    simple shear is characterized by q 90, and plane-strain tension touniaxial tension byq 30. As has been discussed elsewhere (e.g.,Wang et al., 2008), this angle can be used to characterize the ex-

    pected degree of orthogonal loading-path change and similarly

    represent the strength of possible cross hardening as a function of

    position in sheet metal parts subject to forming operations like

    Fig. 5. Comparison of yield surface development inK12; K11space predicted by the THM and the LSM for DC06, and by the PM for IF steel, after uniaxial tension in the rolling

    direction (left) and simple shear in the rolling direction (right), up to 10% von Mises equivalent strain. As done for the results from the THM and the LSM via ( 19), those for Kfrom

    the PM were normalized using the initial yield stress of the material as determined by Peeters et al. (2002). See text for details and web version for color.

    Fig. 6. Development ofK11in DC06 during loading in plane-strain tension (in the rolling direction) to 5% (left) and 10% (right) deformation followed by unloading and reloading in

    uniaxial tension (again in the rolling direction) predicted by the THM and the LSM. For comparison, experimental and model results are also shown for monotonic uniaxial tension

    (Clausmeyer et al., 2011a) and web version for color.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131 127

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    9/12

    deep drawing. An extension to this concept is the recently proposed

    cross hardening indicator (Clausmeyer et al., 2014) which considersthe degree of loading-path change and the material specic ten-

    dency to exhibit cross hardening.

    As mentioned above and shown in Fig. 6 (right), more cross

    hardening is predicted by the THM than the LSM after 10% pre-

    tension. To examine this in more detail, different contributions to

    the general hardening level are displayed inFig. 8. As evident, the

    THM predicts a transient increase inX11(Fig. 8, left) not predicted

    by the LSM (Fig. 8, right) resulting from the coupling ofsxin (16) to

    S. For this loading path change, note that the change in the shape

    of the yield surface predicted by the LSM in the tensionetension

    case (Fig. 7, right) is notas pronounced as for the tension-shear case

    (Fig. 4, right).

    Lastly, consider the FE-based simulation of the tension-bending

    of a DC06 sheet metal workpiece. For this purpose, the THM and theLSM were implemented into LS-DYNA (LSTC, 2012) via the user

    material interface. Implicit global time integration was used for the

    solution of the initial boundary value problem. The specimen of

    dimension 10 mm 10 mm 1 mm was discretized with a regularmesh of 2000 (20 20 5) tri-linear hexahedral nite elements(seeFig. 9) with reduced integration and hourglass control (LSTC,

    2012). Two tension-bending cases are considered: (i) plane strain

    tension in e1 (rolling direction), and (ii) equibiaxial tension in e1and e2, both followed by bending in the e2 direction (seeFig. 11

    below). After predeformation in tension, a bending moment is

    applied on the lower right hand edge of the specimen, and the left

    edge is xed. The sheet is bent until a bending angle of 90 is ob-tained. Bending moment-angle results are shown in Fig. 10; the

    deformed sheet and nal spatial stress distribution are displayed in

    Fig. 11.

    As evident in Fig.10, bending-moment results based on the THM

    and the LSM are qualitatively similar for both loading cases.

    Consideration of hardening stagnation in the THM results in faster

    saturation of the bending moment after about 60 in particular forthe equibiaxial (pre)tension. The expected trend in stress state and

    Fig. 7. Yield surface development in theK11; K22plane from the THM (left) and the LSM (right) due to plane-strain tension in the rolling direction to different levels of deformation.

    The symbols x, and *mark stress states on the yield surface at F1110.05,F1110.10, and F111 0.15, respectively.

    Fig. 8. Development ofrandX11predicted by the THM (left) and the LSM (right) during loading in plane-strain tension (in the rolling direction to 10%) to uniaxial tension ( Fig. 6).

    Fig. 9. Initial dimensions and FE discretization of sheet for tension-bending

    simulations.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131128

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    10/12

    stress distribution from tension (above) to compression (below) the

    midplane of the workpiece is clearly predicted by both models. In-

    plane inhomogeneity of the stress eld perpendicular to this is

    related to thenite dimensions of the specimen. As indicated by the

    bending moment results inFig. 10, the THM predicts more hard-ening in the specimen than the LSM for both plane-strain tension

    andequibiaxial tension followed by bending. This is consistent with

    the results forKIshown inFig. 11, i.e., the stress level predicted by

    the THM is higher than that predicted by the LSM. Since bending

    results in a transition from tension to compression, the above ob-

    servations for plane strain tension to uniaxial tension loading can

    be transferred to the bending case. Up to a bending angle of about

    8 bending of the sheet is mainly governed by elastic deformationbecause the release of the displacement constraints in the e1- and

    e2 directions is only partly compensated by the incipient straining

    in the e1-direction due to bending. Similar to the plane strain

    tension to uniaxial tension case, a transient hardening regime fol-

    lows. Due to the inhomogeneous nature of the bending deforma-

    tion, the duration of this regime (in terms of the bending angle) islarger compared to the homogeneous plane strain tension to uni-

    axial tension case. However, the same tendencies are observed. The

    level of stress predicted by the THM is larger compared to the LSM.

    Consequently, the resulting bending moment is also larger. For

    larger bending angles, the differences between the two models

    become smaller. Note, that the transition between the elastic

    regimeand the plastic transient hardening regime is sharper for the

    LSMcompared tothe THM. This is in agreement with the results for

    plane strain tension to uniaxial tension.

    7. Summary and discussion

    In the current work, two models for anisotropic, and in partic-

    ular cross, hardening in bccsheet steels subject to non-proportional

    loading have been compared with each other in detail. The rst

    model is the modied form (Wang et al., 2008) of that due to

    Teodosiu and Hu (1995, 1998). In this (modied) Teodosiu-Hu

    model (THM), cross hardening is assumed to affect (i) the yield

    stress and (ii) the saturation value of the back stress. The secondmodel is due toLevkovitch and Svendsen (2007)andNoman et al.

    (2010). In this Levkovitch-Svendsen model (LSM), cross hardening

    is assumed to affect theow anisotropy. As attestedto by numerous

    applications of the THM (e.g., Boers et al., 2010; Bouvier et al., 2005,

    2003; Hiwatashi et al., 1997; Li et al., 2003; Thuillier et al., 2010;

    Wang et al., 2008) and the LSM (e.g., Barthel et al., 2013;

    Clausmeyer et al., 2014, 2011b; Noman et al., 2010), both models

    are able to account for the observed effect of cross hardening on

    experimental stress-deformation data. As investigated and docu-

    mented in the current work, there are otherwise a number of dif-

    ferences between the two. For example, in contrast to the THM, the

    LSM predicts no hardening stagnation after cross hardening as

    observed in experiments. On the other hand, in contrast to the LSM,

    the THM predicts no change in yield surface shape during non-proportional loading. As documented in the current work, this is

    in contrast to experimental results (for ferritic tube steel:

    Kowalewski and Sliwowski, 1997) as well as to predictions of more

    sophisticated micromechanical models (e.g.,Holmedal et al., 2008;

    Peeters et al., 2002).

    As discussed in many previous works as well as here, in the

    context of a yield functionfYof the form (1), the (ow) anisotropy

    tensor A determines the shape of the yield surface. In the THM, this

    shape is attributed to the effect of an (initial) rolling texture, i.e., the

    grain (orientation) structure, on the ow behavior. On the other

    hand, in the LSM, A is associated with both the grain orientation

    and dislocation structures, i.e., AA gra H . In the yield surfacecontext and model form (1) for fY, then, (13) and (17) imply the

    correspondence

    Fig. 10. Bending moment as a function of bending angle during tension-bending of DC06 predicted by FE simulations based on the THM and the L SM. Left: plane-strain (pre)tension

    toF111.2. Right: equibiaxial (pre)tension to F11 F221.1. See text for details.

    Fig. 11. Deformed sheet geometry (bending angle 90) after equibiaxial-tension-bending and spatial distribution of the rst principle componentKIof the Kirchhoff stress predicted

    by the THM (left) and the LSM (right). Refer to the web version for color.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131 129

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    11/12

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiS$A gra H

    S

    q

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiS$A gra S

    p fS

    (20)between H and S, where SMX. Consequently, in the THM, thecoupling between the state of stress Mon the yield surface and the

    cross hardening tensor S is linear and scalar. As in the case of

    standard isotropic hardening itself, then, it is not surprising that

    this model decouples any change in form of the yield surface from

    cross hardening effects. This is in contrast to the LSM, which cou-

    ples M and the contribution Hfrom cross hardening to the ow

    anisotropy tensor non-linearly and tensorially. In this case, a

    directional dependence is generic and inherent. As implied by the

    comparisons in Section 5 and Fig. 5, micromechanically based

    models like the PM (Peeters et al., 2001b,a; 2002) appear to support

    the idea of a change inyield surface shape due tocross hardeningas

    assumed in the LSM. Being physically based, the PM provides at

    least indirect physical evidence for the LSM, since it directly in-

    corporates the physical knowledge of the contribution of disloca-

    tion interaction to the evolution of the yield surface.

    With respect to the hardening modeling in general, note that

    isotropic, kinematic, and cross, hardening are decoupled in the

    LSM, but coupled in the THM. From the point of view of the ge-

    ometry and morphology of the yield surface, then, only in the LSMare isotropic, kinematic, and cross, hardening uniquely related to

    the size, center, and shape, respectively, of the yield surface. As

    implied by the current comparison, however, to account for pro-

    cesses such as hardening stagnation, coupling of hardening types

    such as that assumed in the THM may be necessary. Indeed, as

    implied by the results of the current workas well,except in the case

    of the coupling of yield surface morphology and development to

    anisotropic hardening in the sense of a change of yield surface

    shape, however, the THM is more sophisticated and micro-

    mechanical in nature.

    The authors are not aware of comprehensive experimental

    evidence which shows the evolution of the yield surface of DC06

    for larger (pre)strains (z10%). Experiments on tubular specimens

    have shown that there is an evolution of the yield surface (Phillipset al., 1974). In the case of other steels (e.g., SUS304: Ishikawa,

    1997), experimentally observed changes in the yield surface

    shape during tension-torsion loading are accounted for by the

    modeling of cross hardening in the LSM as based on Hand related

    to dislocation structure evolution. Kuwabara et al. (2000) and

    Kuwabara (2007)investigated the effect of uniaxial, plane strain

    and biaxial tension on the yield surface of IF steel up to 1.0% strain.

    This work focused on the initial yield surface alone. In any case,

    more experimental work on sheet steels along these lines is

    clearly needed and required, and will hopefully be available in the

    future.

    Acknowledgments

    The authors thank the reviewers of the rst version of this work

    for their constructive criticism and comments which have lead to

    major improvement. The authors would like to thank Ton van den

    Boogaard from the Faculty of Engineering Technology, University

    Twente, The Netherlands, for providing the opportunity to use the

    biaxial tester, and Alper Gner from the Institute of Forming Tech-

    nology and Lightweight Construction, TU Dortmund University,

    Dortmund, Germany, for providing the uniaxial tension test data.

    Partial nancial support from the German Research Foundation

    (DFG) under contract PAK 250 Identikation und Modellierung der

    Werkstoffcharakteristik fr die Finite-Element-Analyse von Ble-

    chumformprozessen - TP4 is gratefullyacknowledged. The material

    investigated was provided by ThyssenKrupp Steel Europe AG.

    References

    Bacroix, B., Hu, Z., 1995. Texture evolution induced by strain path changes in lowcarbon steel sheets. Metall. Mater. Trans. A 26 (3), 601e613.

    Baltov, A., Sawczuk, A., 1965. A rule of anisotropic hardening. Acta Mech. I (2),81e92.

    Barlat, F., Ha, J., Gracio, J.J., Lee, M.-G., Rauch, E.F., Vincze, G., 2013. Extension ofhomogeneous anisotropic hardening model to cross-loading with latent effects.Int. J. Plast. 46, 130e142.

    Barthel, C., Klusemann, B., Denzer, R., Svendsen, B., 2013. Modeling of a thermo-mechanical process chain for sheet steels. Int. J. Mech. Sci. 74, 46e54.

    Boers, S.H.A., Schreurs, P.J.G., Geers, M.G.D., Levkovitch, V., Wang, J., Svendsen, B.,2010. Experimental characterization and model identication of directionalhardening effects in metals for complex strain path changes. Int. J. Solids Struct.47 (10), 1361e1374.

    Bouvier, S., Alves, J., Oliveira, M., Menezes, L., 2005. Modelling of anisotropic work-hardening behaviour of metallic materials subjected to strain-path changes.Comput. Mater. Sci. 32, 301e315.

    Bouvier, S., Gardey, B., Haddadi, H., Teodosiu, C., 2006a. Characterization of thestrain-induced plastic anisotropy of rolled sheets by using sequences of simpleshear and uniaxial tensile tests. J. Mater. Process. Technol. 174 (1e3), 115e126.

    Bouvier, S., Haddadi, H., Levee, P., Teodosiu, C., 2006b. Simple shear tests: experi-mental techniques and characterization of the plastic anisotropy of rolledsheets at large strains. J. Mater. Process. Technol. 172, 96e103.

    Bouvier, S., Teodosiu, C., Haddadi, H., Tabacaru, V., 2003. Anisotropic work-hardening behaviour of structural steels and aluminium alloys at largestrains. J. Phys. IV Fr. 105, 215e222.

    Butuc, M.C., Teodosiu, C., Barlat, F., Gracio, J.J., 2011. Analysis of sheet metal form-

    ability through isotropic and kinematic hardening models. Eur. J. Mech. A/Solids30 (4), 532e546.

    Carvalho Resende, T., Bouvier, S., Abed-Meraim, F., Balan, T., Sablin, S.S., 2013.Dislocation-based model for the prediction of the behavior of b.c.c. materials egrain size and strain path effects. Int. J. Plast. 47, 29e48.

    Cazacu, O., Barlat, F., 2004. A criterion for description of anisotropy and yield dif-ferential effects in pressure-insensitive metals. Int. J. Plast. 20 (11), 2027e2045.

    Clausmeyer, T., Bargmann, B., Gershteyn, G., Schaper, M., Svendsen, B., 2011a. Finite-element Simulation of the Evolution of Plastic Anisotropy. Steel Research In-ternational ICTP 2011 Special Issue, pp. 836e841.

    Clausmeyer, T., Gerstein, G., Bargmann, S., Svendsen, B., van den Boogaard, A.,Zillmann, B., 2012. Experimental characterization of microstructure develop-ment during loading path changes in bcc sheet steels. J. Mater. Sci. 48 (2),674e689.

    Clausmeyer, T., Gner, A., Tekkaya, A.E., Levkovitch, V., Svendsen, B., 2014. Modelingand nite element simulation of loading-path-dependent hardening in sheetmetals during forming. Int. J. Plast. 63, 64e93.

    Clausmeyer, T., van den Boogaard, A.H., Noman, M., Gershteyn, G., Schaper, M.,Svendsen, B., Bargmann, S., 2011b. Phenomenological modeling of anisotropy

    induced by evolution of the dislocation structure on the macroscopic andmicroscopic scale. Int. J. Mater. Form. 4 (2), 141e154.

    Ghosh, A.K., Backofen, W.A., 1973. Strain-hardening and instability in biaxiallystretched sheets. Metall. Trans. 4, 1113e1123.

    Haddadi, H., Bouvier, S., Banu, M., Maier, C., Teodosiu, C., 2006. Towards an accuratedescription of the anisotropic behaviour of sheet metals under large plasticdeformations: modelling, numerical analysis and identication. Int. J. Plast. 22(12), 2226e2271.

    Haddag, B., Balan, T., Abed-Meraim, F., 2007. Investigation of advanced strain-pathdependent material models for sheet metal forming simulations. Int. J. Plast. 23(6), 951e979.

    Hill, R., 1948. A theory of the yielding and plasticow of anisotropic metals. Proc. R.Soc. Lond. Ser. A. Math. Phys. Sci. 193 (1033), 281e297.

    Hiwatashi, S., van Bael, A., van Houtte, P., Teodosiu, C., 1997. Modelling of plasticanisotropy based on texture and dislocation structure. Comput. Mater. Sci. 9,274e284.

    Holmedal, B., van Houtte, P., An, Y., 2008. A crystal plasticity model for strain-pathchanges in metals. Int. J. Plast. 24 (8), 1360e1379.

    Hu, Z., Rauch, E.F., Teodosiu, C., 1992. Work-hardening behavior of mild steel understress reversal at large strains. Int. J. Plast. 8 (7), 839e856.

    Ishikawa, H., 1997. Subsequent yield surface probed from its current center. Int. J.Plast. 13 (6e7), 533e549.

    Klepaczko, J.R., Nguyen, H.V., Nowacki, W.K., 1999. Quasi-static and dynamicshearing of sheet metals. Eur. J. Mech. A/Solids 18 (2), 271e289.

    Kowalewski, Z.L., Sliwowski, M., 1997. Effect of cyclic loading on the yield surfaceevolution of 18G2A low-alloy steel. Int. J. Mech. Sci. 39 (1), 51e68.

    Kuwabara, T., 2007. Advances in experiments on metal sheets and tubes in supportof constitutive modeling and forming simulations. Int. J. Plast. 23 (3), 385e419.

    Kuwabara, T., Kuroda, M., Tvergaard, V., Nomura, K., 2000. Use of abrupt strain pathchange for determining subsequent yield surface: experimental study withmetal sheets. Acta Mater. 48 (9), 2071e2079.

    Lee, E., 1969. Elastic-plastic deformation at nite strains. J. Appl. Mech. 36, 1e6.Levkovitch, V., Svendsen, B., 2007. Accurate hardening modeling as basis for the

    realistic simulation of sheet forming processes with complex strain-pathchange. In: Cesar de Sa, J.M.A. (Ed.), 9th International Conference on Numeri-cal Methods in Industrial Forming Processes (NUMIFORM 2007), AIP Confer-ence Proceedings, Porto, Portugal, vol. 2, pp. 1331e1336.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131130

  • 7/26/2019 European-A-2015-Comparison of two models for anisotropic hardening and yield surface.pdf

    12/12

    Li, S., Hoferlin, E., van Bael, A., van Houtte, P., Teodosiu, C., 2003. Finite elementmodeling of plastic anisotropy induced by texture and strain-path change. Int. J.Plast. 19, 647e674.

    LSTC, 2012. LS-DYNA. Keyword User's Manual, vol. I. Livermore Software Tech-nology Corporation, Livermore, USA.

    Mandel, J., 1971. Plasticite Classique et Viscoplasticite. CISM Course Lecture Notes97. Springer, Berlin, Germany.

    Mandel, J., 1974. Thermodynamics and Plasticity. In: Foundations of ContinuumThermodynamics Delgado Domingos. MacMillan.

    Miyauchi, K., 1987. On the simple shear deformation. Sci. Pap. Inst. Phys. Chem. Res.

    81, 57e67.Nesterova, E.V., Bacroix, B., Teodosiu, C., 2001a. Experimental observation of

    microstructure evolution under strain-path changes in low-carbon IF steel.Mater. Sci. Eng. A309e310, 495e499.

    Nesterova, E.V., Bacroix, B., Teodosiu, C., 2001b. Microstructure and texture evolu-tion under strain-path changes in low-carbon interstitial-free steel. Metall.Mater. Trans. 32A, 2527e2538.

    Noman, M., 2010. Characterization and model identication for the simulation ofthe forming behavior of ferritic steels (Ph.D. thesis). Fakultat Maschinenbau, TUDortmund University.

    Noman, M., Clausmeyer, T., Barthel, C., Svendsen, B., Huetink, J., van Riel, M., 2010.Experimental characterization and modeling of the hardening behavior of thesheet steel LH800. Mater. Sci. Eng. A 527 (10e11), 2515e2526.

    Peeters, B., Bacroix, B., Teodosiu, C., van Houtte, P., Aernoudt, E., 2001a. Work-hardening/softening behaviour of b.c.c. polycrystals during changing strain:part i. an integrated model based on substrucure and texture evolution, and itsprediction of the stress-strain behaviour of an IF steel during two-stage strainpaths. Acta Mater. 49 (9), 1607e1619.

    Peeters,B., Kalidindi, S.R.,Teodosiu,C., vanHoutte, P.,Aernoudt, E.,2002. A theoreticalinvestigation of the inuenceof dislocation sheets on evolution of yieldsurfacesin single-phase b.c.c. polycrystals. J. Mech. Phys. Solids 50 (4), 783e807.

    Peeters, B., Seefeldt, M., van Houtte, P., Aernoudt, E., 2001b. Taylor ambiguity in bccpolycrystals: a non-problem if substructural anisotropy is considered. Scr.Mater. 45 (12), 1349e1356.

    Phillips, A., Tang, J.L., Ricciuti, M., 1974. Some new observations on yield surfaces.Acta Mech. 20 (1), 23e39.

    Pietryga, M.P., Vladimirov, I.N., Reese, S., 2012. A nite deformation model forevolving ow anisotropy with distortional hardening including experimentalvalidation. Mech. Mater. 44, 163e173.

    Rauch, E.F., Schmitt, J.H., 1989. Dislocation substructures in mild steel deformed insimple shear. Mater. Sci. Eng. A 113, 441e448.

    Rauch, E.F., Thuillier, S., 1993. Rheological behaviour of mild steel under monotonicloading conditions and cross-loading. Mater. Sci. Eng. A 164, 255e259.

    Schmitt, J.H., Shen, E.L., Raphanel, J.L., 1994. A parameter for measuring themagnitude of a change of strain path: validation and comparison with exper-iments on low carbon steel. Int. J. Plast. 10 (5), 535e551.

    Shi, B., Mosler, J., 2012. On the macroscopic description of yield surface evolution bymeans of distortional hardeningmodels: application to magnesium. Int. J. Plast. 44,1e22.

    Shinohara, Y., Madi, Y., Besson, J., 2010. A combined phenomenological model forthe representation of anisotropic hardening behavior in high strength steel linepipes. Eur. J. Mech. A/Solids 29 (6), 917e927.

    Svendsen, B., 2001. On the modeling of anisotropic elastic and inelastic materialbehaviour at large deformation. Int. J. Solids Struct. 38, 9579e9599.

    Tarigopula, V., Hopperstad, O.S., Langseth, M., Clausen, A.H., 2008. Elastic-plasticbehaviour of dual-phase, high-strength steel under strain-path changes. Eur. J.

    Mech. A/Solids 27 (5), 764e782.Tarigopula, V., Hopperstad, O.S., Langseth, M., Clausen, A.H., 2009. An evaluation of a

    combined isotropic-kinematic hardening model for representation of complexstrain-path changes in dual-phase steel. Eur. J. Mech. A/Solids 28 (4), 792e805.

    Teodosiu, C., Hu, Z., 1995. Evolution of the intragranular microstructure at moderateand large strains: modelling and computational signicance. In: Shen, S.F.,Dawson, P.R. (Eds.), Simulation of Materials Processing: Theory, Methods andApplications. Balkema, Rotterdam, pp. 173e182.

    Teodosiu, C., Hu, Z., 1998. Microstructure in the continuum modelling of plasticanisotropy. In: Proceedings of 19th Ris International Symposium on Material'sScience: Modelling of Structure and Mechanics of Materials from Microscale toProduct. Ris National Laboratory, Roskilde, Denmark, pp. 149e168.

    Thuillier, S., Manach, P.Y., Menezes, L.F., 2010. Occurrence of strain path changes in atwo-stage deep drawing process. J. Mater. Process. Technol. 210 (2), 226e232.

    Thuillier, S., Rauch, E.F.,1994. Development of microbands in mild steel during crossloading. Acta Metall. Mater. 42, 1973e1983.

    van Riel, M., van den Boogaard, A.H., 2007. Stress-strain responses for continuousorthogonal strain path changes with increasing sharpness. Scr. Mater. 57 (5),381e384.

    Verma, R.K., Kuwabara, T., Chung, K., Haldar, A., 2011. Experimental evaluation andconstitutive modeling of non-proportional deformation for asymmetric steels.Int. J. Plast. 27 (1), 82e101.

    Viatkina, E.M., Brekelmans, W.A.M., Geers, M.G.D., 2007. Modelling of the internalstress in dislocation cell structures. Eur. J. Mech. A/Solids 26 (6), 982e998.

    Wagoner, R.H., Lim, H., Lee, M.-G., 2013. Advanced issues in springback. Int. J. Plast.45, 3e20.

    Wang, J., Levkovitch, V., Reusch, F., Svendsen, B., Huetink, J., van Riel, M., 2008. Onthe modeling of hardening in metals during non-proportional loading. Int. J.Plast. 24 (6), 1039e1070.

    Wu, H.-C., 2003. Effect of loading-path on the evolution of yield surface foranisotropic metals subjected to large pre-strain. Int. J. Plast. 19 (10), 1773e1800.

    Zienkiewicz, O.C., Taylor, R.L., Zhu, J.Z., 2010. The Finite Element Method e its Basis& Fundamentals, sixth ed. Elsevier e Butterworth Heinemann, Amsterdam.

    T. Clausmeyer, B. Svendsen / European Journal of Mechanics A/Solids 54 (2015) 120e131 131