Zhang Bower Mishra Boyle IJP 2009

Embed Size (px)

Citation preview

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    1/21

    Numerical simulations of necking duringtensile deformation of aluminum single crystals

    F. Zhang a,*, A.F. Bower a, R.K. Mishra b, K.P. Boyle c

    a Division of Engineering, Brown University, Providence, RI 02912, USAb Materials and Processes Laboratory, GM R&D Center, Warren, MI 48090, USA

    c CANMET Materials Technology Laboratory, Ottawa, ON, Canada K1A 0G1

    Received 16 October 2007; received in final revised form 13 December 2007Available online 31 December 2007

    Abstract

    Finite element simulations are used to study strain localization during uniaxial tensile straining of

    a single crystal with properties representative of pure Al. The crystal is modeled using a constitutiveequation incorporating self- and latent-hardening. The simulations are used to investigate the influ-ence of the initial orientation of the loading axis relative to the crystal, as well as the hardening andstrain rate sensitivity of the crystal on the strain to localization. We find that (i) the specimen fails bydiffuse necking for strain rate exponents m< 100, and a sharp neck for m> 100. (ii) The strain tolocalization is a decreasing function of m for m< 100, and is relatively insensitive to m form> 100. (iii) The strain to localization is a minimum when the tensile axis is close to (but not exactlyparallel to) a high symmetry direction such as [100] or [111] and the variation of the strain to local-ization with orientation is highly sensitive to the strain rate exponent and latent-hardening behaviorof the crystal. This behavior can be explained in terms of changes in the active slip systems as theinitial orientation of the crystal is varied.

    2007 Elsevier Ltd. All rights reserved.

    Keywords: Localization; B. Crystal plasticity; C. Finite element

    0749-6419/$ - see front matter 2007 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.ijplas.2007.12.006

    *

    Corresponding author. Tel.: +1 401 863 2674; fax: +1 401 863 9009.E-mail address: [email protected](F. Zhang).

    Available online at www.sciencedirect.com

    International Journal of Plasticity 25 (2009) 4969

    www.elsevier.com/locate/ijplas

    mailto:[email protected]:[email protected]
  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    2/21

    1. Introduction

    Strain localization limits the formability of many ductile materials in processes such aspunching, drawing, and hemming. This is a particular concern to the automotive industry,

    where efforts to replace steel with lightweight Al and Mg alloys are impeded by the lowerformability of the replacement materials (Taub et al., 2007). Complex parts such as dooror lift-gate inner panels either need to be redesigned to accommodate the lower formabilityof Al and Mg alloys at room temperature, or else must be manufactured using moreexpensive and slower processes such as elevated temperature Quick Plastic Forming (Taubet al., 2007). These demands are inspiring a renewed interest in the problem of strain local-ization during plastic flow. A principal goal of this effort is to determine the roles of alloycomposition and microstructure in controlling flow localization, with a view to improvingformability through alloy and process design.

    Forming limits depend on the part geometry, loading conditions (especially deforma-tion rate, temperature, and the presence and type of lubrication), and the material prop-erties. Nevertheless, tensile tests are frequently used as a rapid means to estimate theability of a material to resist shear localization. The roles of material properties, specimengeometry, and loading conditions in controlling localization during a tensile test have beenextensively studied. Early work (Considere, 1888; Hill and Hutchinson, 1975; Hutchinsonand Neale, 1977) based on classical plasticity theory, focused on the influence of strainhardening and strain rate sensitivity of the specimen on the conditions necessary to initiatelocalization. Although these models oversimplify the constitutive behavior of the solid,their main conclusions remain broadly applicable even in more sophisticated analyses:

    namely, (i) materials with a high rate of strain hardening compared with their flowstrength will resist strain localization; (ii) plastic anisotropy can lead to the formationof vertices in the yield surface, which tend to make the material more prone to shear local-ization; and (iii) high strain rate sensitivity leads to the formation of diffuse necking andalso delays localization. More recent work has exploited the development of constitutivelaws for single crystals, which have been used to investigate the influence of plastic anisot-ropy resulting from texture formation in polycrystals on localization (Petryk, 1997; Shen,1993; Tvergaard and Needleman, 1993; Inal et al., 2002a,b), as well the influence of geo-metric softening resulting from crystallographic rotation during tensile deformation of sin-gle crystals (Asaro, 1979; Peirce et al., 1982, 1983; Balke and Estrin, 1994; Yang and Rey,

    1995; Stuwe and Toth, 2003).While these studies have provided insight into the mechanics of strain localization, we

    still lack the detailed connection between localization and alloy composition or micro-structure that is required for alloy design. This is partly because models of localizationare based on phenomenological constitutive laws, and it is not easy to relate constitutiveparameters in these models to the underlying processes that control deformation. Currentresearch is attempting to address this issue: studies range from efforts to establish the roleof alloy chemistry at the atomic scale on plastic flow (Curtin and Olmsted, 2006), toimproving the description of defects and microstructural features in constitutive models(Ma et al., 2006). As part of this effort,Boyle and Mishra (2005) and Boyle (2005, submit-

    ted for publication)have recently developed a hardening law for FCC crystals, with a par-ticular focus on pure Al and Cu, which includes a more direct description of themicroscopic processes that control hardening. This model is based on the standarddescription of rate dependent slip in single crystals, but extends prior models by incorpo-

    50 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    3/21

    rating a more detailed description of latent and self-hardening processes. In particular, itaccounts for the three stages of hardening, including the temperature dependence of StageIII; and also contains a particularly detailed model of the evolution of latent-hardeningwith plastic straining in the crystal. Although a large number of phenomenological param-

    eters remain in the constitutive equations, which are currently calibrated using experimen-tal data, many of these (such as the strengths of junctions between dislocations on thevarious active slip systems) can in principle be computed from atomic scale studies,thereby providing a more direct link between macroscopic flow and localization behaviorand microscopic deformation mechanisms.

    In this paper, we briefly review the main features of the refined model and the hardeninglaws and show that an appropriately calibrated constitutive law captures the principaltrends observed in experimental measurements of the tensile behavior of pure Al singlecrystals, including the influence of crystal orientation on the hardening response and thestrain to localization. We use the model to explore the influence of changes in the proper-ties of the crystal, with a particular focus on its strain rate sensitivity, orientation, andlatent-hardening behavior.

    2. Model

    The problem to be solved is illustrated inFig. 1. A tensile bar with lengthLand circularcross-section is subjected to prescribed displacements so as to extend the specimen in uni-axial tension at constant nominal strain rate. Specifically, we enforceu3= 0 on the plane atx3= 0 throughout straining and constrain u1=u2= 0 at the origin. Axial rotation of the

    solid is constrained by enforcing u2= 0 at x1= R, x3= 0. The end of the specimen atx3= L in the initial configuration is subjected to a constant velocity parallel to the x3direction. The loading is assumed to be quasi-static. Following Hutchinson and Neale(1977)we introduce a small geometric imperfection in the specimen, which serves to initi-ate localization in a controlled manner. This geometric imperfection takes the form of avariation in the cross-sectional area of the bar along its length given by

    Fig. 1. Schematic representation of a single-crystal tensile specimen, including the finite element mesh andgeometric parameters used in the simulation (symbols are defined in the text).

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 51

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    4/21

    A A0 1g cos 2px3

    L

    1

    where the A0 is the mean cross-sectional area of the bar and g is the amplitude of the

    variation in cross-sectional area.The model uses the standard description of the kinematics of slip in a crystal. Letxibethe position of a material particle in the undeformed crystal. The solid is subjected to adisplacement fieldui(xk), such that the point at ximoves toyi= xi+ uiafter deformation,as illustrated inFig. 2. The deformation gradient and its Jacobian are given by

    Fij dijoui

    oxj; JdetF 2

    The velocity gradient L, stretch rate Dand spin Ware given by

    Lij

    o _ui

    oyj _FikF

    1kj ; Dij Lij Lji=2; Wij Lij Lji=2 3

    The total deformation gradient is decomposed into elastic and plastic parts by assumingthat deformation takes place in two stages. The plastic strain is assumed to shear the lat-tice, without stretching or rotating it. The elastic deformation rotates and stretches the lat-tice. We think of these two events occurring in sequence, as illustrated inFig. 2, with theplastic deformation first, and the stretch and rotation second, giving

    Fij FeikF

    pkj 4

    The velocity gradient may then be decomposed into elastic and plastic parts by noting that

    Lij _FikF1kj _F

    eikF

    pkl F

    eik

    _Fpkl

    F

    p1lm F

    e1mj

    _FeikF

    e1kj F

    eik

    _FpklF

    p1lm F

    e1mj 5

    The velocity gradient contains two terms, one of which involves only measures of elasticdeformation, while the other contains measures of plastic deformation. We use this todecompose L into elastic and plastic parts as

    Lij Leij L

    pij L

    eij _F

    eikF

    e1kj L

    pijF

    eik

    _FpklF

    p1lm F

    e1mj 6

    Fig. 2. Sketch showing the decomposition of plastic strain into elastic and plastic parts.

    52 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    5/21

    Plastic flow in the crystal occurs by shearing a set ofNslip systems. An FCC crystal hasN= 12 slip systems, which are listed inTable 1and labeled according to the Schmidt andBoas convention. For computational purposes the slip systems are characterized by unitvectors parallel to slip directions sai and slip plane normals m

    ai in the undeformed solid.

    The rate of shear on the ath system is denoted by _ca. The velocity gradient due to thisshearing is

    _Fp

    ikFp1kj

    XNa1

    _casai maj 7

    The lattice shearing does not alter the slip plane normal or the slip direction. Conse-quently, sai and m

    ai are stretched and rotated only by the elastic part of the deformation

    gradient, so that in the deformed configuration

    sai Feiks

    akm

    ai m

    akF

    e1ki 8

    The plastic part of the velocity gradient can be expressed in terms of the shearing rates andthe deformed slip vectors as

    Lpij

    XNa1

    _casai maj 9

    Finally, the elastic and plastic parts of the velocity gradient can be decomposed into sym-metric and skew symmetric parts, representing stretching and spin, respectively, as

    Deij Leij L

    eji=2; W

    eij L

    eijL

    eji=2

    Dpij Lpij Lpji=2; Wpij LpijLpji=210

    In particular, the plastic stretching and spin can be expressed in terms of the latticeshearing as

    Dpij

    XNa1

    _casai maj s

    aj m

    ai =2; W

    pij

    XNa1

    _casai maj s

    aj m

    ai =2 11

    Table 1Slip systems for FCC single crystals

    Slip plane Slip direction

    (111) 011 a1101 a2110 a3

    111 [011] b1101 b2110 b3

    111 011 c1101 c2[110] c3

    111 [011] d1101 d2110 d3

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 53

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    6/21

    Internal forces in the crystal are characterized by the Cauchy (true) stress rij, repre-senting the force per unit deformed area in the solid. The Kirchoff, Nominal, and Materialstress tensors as follows:

    sijJrij SijJF1

    ikrkj; RijJF1

    ik rklF

    1

    jl 12respectively. Plastic shearing on the slip systems is driven by the resolved shear stress, de-fined as

    sa Jmai rijsaj 13

    The constitutive equations for the crystal must relate the elastic and plastic parts of thedeformation gradient (or strain rates derived from them) to the stresses acting on the crys-tal. The elastic constitutive equation relates the elastic stretch rate to the Kirchhoff stressas

    sijr CeijklDekl 14

    where sijr

    is the Jaumann rate of Kirchhoff stress,

    sijr

    dsij

    dt Weikskj sikW

    ekj 15

    and

    CeijklFeinF

    ejmCnmpqF

    ekpF

    elq 16

    where Cijklare the components of the elastic stiffness tensor for the material with orienta-

    tion in the undeformed configuration. For an FCC crystal the elastic modulus tensor canbe characterized by the tensile and shear moduli c11C1111, c12C1122 and c44C1212with the remaining terms determined by symmetry. The modulic11,c12,c44are temperaturedependent, so we set

    cijc0ijc

    TijT 17

    where c0ij is the modulus at T= 0, cTij is the rate of change of modulus with temperature,

    and Tis absolute temperature. The values used for all the material constants are listedinTable 2. Elastic constants were fit to the data collected by Simmons and Wang (1971).

    The plastic constitutive equations specify the relationship between the stress on the

    crystal and slip rates _ca on each slip system. We adopt a power-law rate-dependent flowfor this purpose, with the form

    _ca _c0signsa

    jsaj

    ga

    m18

    where sa Jmai rijsaj is the resolved shear stress on the slip system, g

    a is its currentstrength (which evolves with plastic straining as defined below) and _c0, m are materialproperties.

    The constitutive law for describing the slip system hardening behavior of the bar is

    based on a model for fcc metals developed by Boyle and Mishra (2005) and Boyle(2005, submitted for publication). The details of the constitutive model and its calibrationare discussed in detail elsewhere (Boyle, submitted for publication). Here, we will merelysummarize the governing equations and give values for the relevant material parameters.

    54 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    7/21

    The hardening rule specifies the relationship between the slip system strengths ga andthe plastic strain. At time t= 0 each slip system has the same initial strength g0. Thereaf-ter, the slip systems evolve as a result of the plastic shearing. Theath system may harden asa result of shearing on the ath system itself or a coplanar system (self-hardening). It mayalso harden as a result of shearing on a non-coplanar system (latent-hardening). The twomodes of hardening are distinguished by setting

    _ga XNb1

    ~qabhbj_cbj 19

    where hb is a self-hardening modulus, and ~qab are a set of latent-hardening moduli. The

    latent-hardening moduli satisfy ~qab 1 for a = b. Fora6b, ~qab evolve with plastic strain-ing, as discussed below. It is convenient to discuss the constitutive equations for self- andlatent-hardening in turn.

    2.1. Self-hardening

    Strain hardening in crystals is conventionally divided loosely into three (or more) sep-arate stages, which correspond to characteristic features of the uniaxial tensile stressstraincurve, and are associated with different microscopic processes that contribute to strain

    Table 2List of values used for material parameters for Al

    c011 180395.51 MPacT11 46.71 MPa K

    1

    c0

    12 63255.16 MPacT12 9.52 MPa K1

    c044 32156.62 MPacT44 13.85 MPa K

    1

    l0 29720.70 MPalT 14.492 MPa K1

    h0 60.0 MPahI 15.0 MPas0 0.8 MPasI 1.0 MPasv0 115.5 MPaA/l 0.21019 m3

    _c1 107

    s1

    c0 0.005cII 0.0vNm=l 1.0vCm=l 0.0vGm=l 1.0vHm=l 0.75vSm=l 2.15qNm 2.0qCm 1.1qGm 2.0qH

    0

    m 1.6qSm 2.0qs 1.1

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 55

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    8/21

    hardening (Kocks and Mecking, 2003). Stage I hardening (or easy glide) is observed incrystals that are oriented to activate a single slip system. It is associated with a low, tem-perature independent hardening rate. Stage II hardening initiates when a second slip sys-tem is activated. It is associated with a high rate of hardening, with a roughly linear

    relationship between stress and strain. The hardening rate in Stage II is also insensitiveto temperature. Stage III hardening occurs for large strains, and corresponds to a progres-sively decreasing hardening rate. The rate of hardening in stage III is very sensitive to thetemperature and rate of deformation.

    To describe this behavior, the self-hardening modulushbfor the three stages (I, II, andIII) of hardening is expressed as

    hbhIIIb h

    Ib h

    IIb 20

    The stage I modulus is expressed as

    hIb h0 hIsech

    2

    h0hIcb=sI s0 hI 21

    where h0 and hI are the initial and final hardening rates during stage I, and

    cb

    Z t0

    j _cbjdt0 22

    is the accumulated plastic slip on the bth system.Stage II hardening is assumed to be activated by secondary slip with the stage II mod-

    ulus given by

    hIIb XNj1

    vbjfnacj=c0; fnacj=c0 0 c

    j

    cII

    23

    Here,cIIis a material constant that represents the amount of secondary slip necessary forthe initial activation of stage II,nais the number of active slip systems, and c0is a materialconstant, defined so that c0/na+cII is the amount of slip after which the interactionreaches 60% of peak strength. In this expression, a slip system is considered active ifthe fractional shearing rate is greater than a critical amount, i.e.

    naXna1

    naa ; n

    aa

    1 for _caj j=P _cbj jP 0:10 for _caj j=

    P _cbj j< 0:1

    24

    The coefficientsvbj in Eq.(23)are a set of material constants that control the magnitude ofthe dislocation interaction strengthening between slip systems bandj. The value ofvbj fora pair of slip systems depends on the type of dislocation reaction that occurs between thetwo systems. We consider the following types of interaction:

    (1) Systems with parallel slip directions, with jsb sj j= 1, form no junctions and havevbj =vN;

    (2) Coplanar systems with non-parallel slip directions (jsb

    s j

    j< 1, jmb

    mj

    j= 1) formcoplanar junctions and have vbj =vC;(3) Systems with mutually perpendicular slip planes and slip directions (mb sj = 0 or

    mj sb = 0) form a glissile lock, and have vbj =vG;

    56 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    9/21

    (4) The remaining systems, which must have sj mb 60 and sb mj 60 form a Hirthlock, with vbj =v H if the slip directions combine to produce a vector of [100] type(sj +sb [001]), and form a sessile lock, with vbj =vS, otherwise.

    The stage III modulus is sensitive to temperature and strain rate and is given by

    hIIIb 1tanh4 2

    gb

    sv

    25

    where, followingKocks and Mecking (2003), we take the saturation stresssvto be given by

    sv

    l

    sv0

    l0

    _cb

    _c1

    kTA

    26

    where l = l0+ lTTis a temperature dependent average shear modulus for the crystal,l0

    and sv0 are the shear modulus and staturation stress at 0 K, _c1 is a representative strainrate,kis Boltzmanns constant,Tis the absolute temperature, and A is a material param-eter which depends on the stacking fault energy of the crystal. In our computations, Aistaken to be a fixed fraction of the shear modulus, as listed in Table 2.

    2.2. Latent-hardening

    The latent-hardening ratios ~qab specify the rate of strain hardening due to shearing onnon-co-planar slip systems. FollowingBoyle (submitted for publication), the latent-hard-ening coefficient ~qab are expressed in terms of the slip system strengths as

    ~qab 2

    sI s0sIq

    abm qs s0q

    abs 1sech

    2 2

    sI s0gb s0

    qs 27

    whereb= 1, . . .,N, and no summation over b is implied. Among the internal variables,s0andsIare the initial and back-extrapolated critical yield stresses in shear for stage I, whileqabm and qsare maximum and saturation latent-hardening ratios. The values ofq

    abm depend

    on the types of dislocation reaction that occur between slip systems a and b, and are clas-sified using the same convention that was used to assign values to vab above. For example,systems which form no junction have qabm q

    Nm ; systems which form coplanar junctions

    have qabm qCm; systems which form glissile locks have qabm qGm; systems which form Hirthlocks have qabm q

    Hm ; and systems which form sessile locks have q

    abm q

    Sm.

    Parameter values for material constants representing pure Al are listed in Table 2. Thecalibration of the model is described in more detail inBoyle (submitted for publication).

    3. Finite element implementation

    The boundary value problem was solved using the commercial finite element codeABAQUS, augmented by a user-subroutine that implements the constitutive equations

    discussed above.Fig. 1illustrates a representative finite element mesh, which consists of8-noded reduced integration brick elements. Mesh sensitivity was tested by repeatingselected simulations with four times the number of elements: our results indicate that local-ization strains are changed by less than 2.5% with the refined mesh.

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 57

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    10/21

    4. Results and discussion

    4.1. Orientation dependence of limit strain

    Fig. 3shows a set of representative nominal stressnominal strain curves for uniaxialtensile deformation of model single crystals with properties intended to represent pureAl at a temperature of 273 K.Table 2lists the values used for parameters in the constitu-tive model. The parameter g characterizing the variation of the cross-sectional area of thespecimen (Eq.(1)) wasg = 0.005 and the solid was deformed at a constant nominal strainrate of 0.05 s1. Results are shown for several initial orientations of the tensile axis withrespect to the crystal, including loading along [123], [112], [100], and [111] directions.The behavior of crystals loaded near a high symmetry orientation such as [100] and[1 1 1] is extremely sensitive to small misorientations. Consequently, results are shown bothfor perfect [100] and [111] orientated crystals, and also for specimens in which the loadingaxis was given a small (0.05) random misorientation away from the high-symmetrydirection.

    In our simulations, crystals that were loaded along or near the [100] and [111] direc-tions failed by strain localization. For these cases,Fig. 3shows the necking strain, whichis defined as the strain at the instant that the strain rate near the ends of the specimendrops to zero.

    The orientation dependence of the localization strain is simply a consequence of thehigh initial rate of hardening that occurs when the crystal is loaded along a symmetry axis,due to slip on multiple systems. This results in a higher flow stress at strains exceeding

    10%; at the same time the hardening rate decreases with continued straining which makes

    Fig. 3. Comparison of predicted nominal stressstrain curves for pure Al deformed in uniaxial tension with thetensile axis initially parallel to [1 0 0], [11 0], [1 1 1] and [1 2 3] orientations. Necking strains are marked by a dot oncurves. FEM results are shown as solid lines while experimental results are fromHosford et al. (1960). The FEMresults for perfect [100] and [111] orientations are shown as dotted lines, while solid lines for these orientationscorrespond to a deviation of 0.05 from the ideal direction.

    58 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    11/21

    the material more prone to localization, as predicted by first approximation calculations(Considere, 1888) over a century ago.

    For comparison,Fig. 3also shows the experimental stressstrain curves for Al singlecrystals obtained byHosford et al. (1960). Although our simulations do not give a perfect

    match to the experimental stressstrain curves, the model clearly reproduces the principalfeatures of experimental data, including (i) the qualitative dependence of the flow stress ofthe crystal on the direction of straining: the [111] orientation has the highest strength; fol-lowed by [100], [112], and [123] oriented crystals as well as the crossover of the stressstrain behavior at high strains for [100] and [112] orientations; (ii) the model correctlypredicts that crystals with [100], [111], and [112] orientations localize, whereas that with[123] orientation show no evidence of localization up to 50% nominal strain. In addition,the localization strains predicted by the simulations are close to those observed experimen-tally. The experiments and computations show detailed features that are not in perfectagreement, however. For example, the model predicts a strong increase in hardening rateat approximately 40% strain for the [123] oriented crystal: this is associated with the stageIstage II transition at the onset of secondary slip in the crystal. Hosford et al. (1960)found no evidence of a stage III transition in their experiments.

    Our main objective in the remainder of this paper is to study in more detail the influenceof geometric and material parameters in controlling the localization strain for these modelcrystals.

    4.2. Effects of specimen geometry

    It is important to note that critical localization strains predicted by the simulations areinfluenced by specimen geometry, especially geometric imperfections. For example,Fig. 4shows the critical localization strain as a function of the dimensionless defect parameterg,which characterizes the initial geometric imperfection in the specimen as defined in Eq.(1).The results are shown for [100] loading of an Al crystal at 273 K, with two values of the

    Fig. 4. Necking strains (defined to be the critical nominal strain where the strain rate outside the localized regionof the specimen drops to zero; see text for details) as a function of the dimensionless geometric defect g defined inEq.(1). Results are shown for [100] (with deviation) loading of Al at 273 K.

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 59

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    12/21

    slip rate exponentm = 20 and m= 150 in Eq.(18). With a low rate exponentm = 20, thelocalization strain is evidently proportional to 1/log(g). For the high rate exponentm = 150, the localization strain still decreases with the increasing value of g, but is notdescribed by a simple linear relationship with log(g). The sensitivity of localization strain

    to specimen geometry makes it difficult to compare quantitatively the predicted localiza-tion strains with experiment. However, the qualitative influences of crystal orientation,temperature and material properties on localization strain are independent of geometry.

    In the remaining simulations reported in this paper, a small, fixed geometric defect isused to initiate localization in all simulations. This is not strictly necessary: the inhomoge-neity associated with the finite element discretization and rounding errors during the com-putation will cause even a geometrically perfect specimen to localize. In this case,localization strains would, however, be sensitive to mesh design and machine precision,so it is preferable to introduce a controlled defect. In actual experiments, localizationmay be triggered by a variety of processes, but the qualitative variation of localizationstrains with material and loading parameters is not sensitive to the details of the initiationprocess.

    4.3. Effects of strain rate sensitivity

    Fig. 5shows the results of simulations in which the strain rate exponent min Eq.(18)was systematically varied. Results are shown for [100] loading of Al at 273 K, with a geo-metric defect g= 0.005. We observe two regimes of behavior: for m< 100 the specimenfails by forming a relatively broad neck near the center of the specimen, and in this regime

    the localization is highly sensitive to the strain rate exponentm. Form> 100, the specimenforms a diffuse neck during the early stages of deformation, but then localizes near theedges of this necked region. In this regime both the stressstrain curves and the criticalstrain to localization show a reduced sensitivity to m. Similar trends have been reportedby earlier studies, e.g.Hutchinson and Neale (1977), Peirce et al. (1983).

    Fig. 5. Necking strains as a function of strain rate exponentmas defined in Eq.(18). Results are shown for [1 0 0](with deviation) loading of Al at 273 K, with a geometric defectg = 0.005.

    60 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    13/21

    The localization modes for perfectly oriented [100] crystals are different in these tworegimes. When m< 100, localization occurs by the formation of a single neck, as shownin Fig. 6a. In contrast, for m> 100, we observe two necking regions, as shown inFig. 6b. However, the double neck is observed only for a crystal with perfect [100] orien-

    tation: a very small deviation of the loading axis can change the deformation mode back tothe standard, single neck, as shown inFig. 6c. In contrast, specimens with lower rate sen-sitivity (m< 100) always fail by formation of a single neck, regardless of orientation. Theneck becomes progressively more diffuse as m is reduced.

    The results show that strain rate sensitivity clearly plays a critical role in governingmaterial behavior, and from the standpoint of alloy design, modifying alloy chemistryand microstructure to increase strain rate sensitivity may be a more effective way toimprove formability than attempting to modify hardening behavior. Assuming that theabove trend for a single crystal holds true for polycrystalline alloys where m valuesbetween 10 and 250 in AlMg alloys (Yao et al., 2000) is achievable, this result can betaken as a guide to designing high ductility Al alloys.

    4.4. Effects of initial orientation of the crystal

    In this section, we study in more detail the influence of the initial orientation of the sin-gle crystal on its strain to localization. Throughout this section, the orientation of the ten-sile axis with respect to the crystal will be displayed graphically on an inverse pole figure,such as that shown inFig. 7. A detailed description of the construction of an inverse pole

    Fig. 6. Localization modes for Al loaded along [1 0 0] at 273 K: (a)m= 5; (b)m= 150; and (c)m= 150 but with asmall deviation from the perfect [1 0 0] orientation.

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 61

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    14/21

    figure can be found inHosford, 1993). Briefly, the figure displays stereographic projections

    of characteristic crystallographic directions in the convention adopted here, the projec-tions of the [010] and [001] directions form a Cartesian coordinate system in the plane ofthe figure, while the [100] direction points out of the plane. The orientation of the tensileaxis with respect to the crystal can then be specified by the projection of a unit vector par-allel to the tensile axis, which corresponds to a. The full inverse pole figure for an fcc crys-tal can be divided into 24 triangular regions, each of which (100), (011), and (111)directions at the three vertices.Fig. 7shows two such regions. Each point within these tri-angles corresponds to a particular loading direction for example, the vertices indicateloading parallel to [100], [110], or [111] directions. Due to the symmetry of the crystal,each of the 24 triangular regions in the pole figure are crystallographically identical,

    and it is only necessary to explore behavior with the tensile axis in a single representativetriangle. Accordingly,Fig. 7shows a contour map of the strain to localization as a func-tion of the initial orientation of the tensile axis relative to the [100], [110], or [111] direc-tions. The results are for Al deformed at 273 K with strain rate sensitivity m= 150 and aspecimen defect parameter value of 0.005. The necking strain is smallest for initial orien-tations near the two high symmetry orientations [111] and [100].

    The influence of initial orientation on necking strain is intimately connected to the evo-lution of the crystals orientation with straining.Fig. 8shows the progressive orientationchange of the crystal as a function of straining for crystals with a range of initial orienta-tion plotted on an inverse-pole figure. The initial orientations correspond to points that lie

    along lines marked a, b and c inFig. 7. Each line inFig. 8is color-coded to indicate thenominal strain variation with the orientation of the loading axis.

    Fig. 8shows that the crystal rotation may follow one of three general trends dependingon the initial orientation:

    Fig. 7. A contour map showing the strain to localization as a function of orientation of the initial loading axis inthe standard triangle. The results are for Al, deformed at 273 K with strain rate sensitivity m= 150 and aspecimen defect parameterg= 0.005. Lines marked a, b and c refer to deformation paths referred to inFigs. 8and 9. The simulations were terminated when the nominal strain reached 60%, so no localization occurred in thered regions of the map.

    62 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    15/21

    (1) If the crystal is loaded with the tensile axis perfectly parallel to the [100] or [111]axis, it does not rotate at all during deformation.

    (2) If the initial loading axis is far from the [1 0 0] and [1 1 1] vertices, the crystal initiallydeforms in single slip, and consequently the loading axis rotates along a great circletowards the [101] orientation. Shortly after the tensile axis leaves the standard trian-gle, a second slip system is activated: as a result, the crystal then begins to rotate soas to bring the tensile axis towards the stable [112] orientation. If the crystal couldcontinue to deform indefinitely without localization, the loading axis would eventu-ally converge to the [112] direction. However, our simulations show that in all caseslocalization sets in shortly after the onset of double slip. We note in passing that thecritical strain to activate double slip depends strongly on the latent-hardening behav-ior of the crystal. This behavior is explored in more detail in the next section.

    (3) If the tensile axis lies very close to, but not exactly on, the [100] or [111] loadingdirection, our simulations predict extremely complex behavior, as illustrated inFig. 8. In this case, multiple slip systems are activated, and the loading axis showsa tendency to spiral around the high-symmetry orientation. In most cases, the crystaleventually settles to a state of double slip, but the choice of slip systems in this state ishighly sensitive to the initial misorientation of the crystal, as well as its latent-hard-ening behavior.

    Fig. 9shows the influence of initial orientation for crystals loaded near [100] and [111]orientations in more detail. Fig. 9a and b corresponds to crystal orientation variations

    along lines a and b in Fig. 7, respectively. Fig. 9a can be thought of as representingthe effects of misaligning the loading axis from the [100] direction by rotating the crystalabout the 0 6 1 1 direction; similarly, Fig. 9b represents the influence of misaligningthe loading axis from the [111] direction by rotating the crystal about its 1 5 4direction.

    Fig. 8. Inverse pole figures showing the evolution of the tensile axis for crystals with various initial orientations inthe standard triangle. The figure shows results for initial orientations along the lines marked a, b and c inFig. 7. The curves are color coded to indicate the variation of nominal strain along the orientation path. Mostinitial orientations deform initially in single slip, then transition to double slip after the loading axis leaves thestandard triangle, and approach the stable orientation with loading axis parallel to the [112] direction. If theloading axis is initially close to a high symmetry orientation, however, it may spiral around the high symmetrydirection.

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 63

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    16/21

    Several features of these results are worth noting. Firstly, the strain to localization is amonotonically increasing function of the misorientation angle for angles exceeding 4.For smaller misorientations, however, the variation of localization strain with misorienta-tion is complex. A very small misorientation away from the perfect [100] or [111] orien-

    tation causes a rapid decrease in localization strain: the effect is particularly noticeablenear the [111] orientation in Fig. 9b, where the localization strain drops from 0.38 to0.22 as a result of a misorientation of only 0.04. The drop is less pronounced for crystalsloaded near [100], but in this case, the necking strain is a fluctuating function of the mis-orientation. These fluctuations are not numerical noise: they are associated with changes inthe history of rotation of the crystal (seeFig. 8). This behavior occurs because the activeslip systems are highly sensitive to small changes in orientation when the crystal is loadednear a high symmetry orientation. The hardening behavior of the crystal and its rate sen-sitivity play a critical role in controlling the active slip systems, as discussed in more detailbelow.

    4.5. Effects of rate sensitivity and latent-hardening on variation of localization strain

    The behavior of crystals loaded near high symmetry orientations (e.g. with loading axesclose to [100] or [111] directions) is determined by the number of active slip systems andthe relative shear rates on each active system. These, in turn, are strongly sensitive to therate sensitivity and strain hardening behavior of the crystal.

    The influence of strain rate sensitivity and latent-hardening on the orientation depen-dence of localization strain is illustrated inFig. 10, which shows the variation of localiza-tion strain with the misorientation of the tensile axis for a crystal loaded near the [111]

    direction. (Fig. 10b shows the same data as (a), but re-plotted to show the behavior atsmall misorientations more clearly.) Results are shown for crystals with isotropic harden-ing andm = 150 as well as with latent-hardening with rate exponent m = 20 and m= 150.To understand the implications of this comparison, it is helpful to recall that Al has

    Fig. 9. Variation of necking strain with misorientation of the tensile axis from a high symmetry direction: (a)results for tensile axis near [100], following the path marked a inFig. 7and (b) results for the tensile axis near[11 1], along the path b in Fig. 7.

    64 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    17/21

    ~qab >1 (the value ofq evolves with straining), which implies that slip on the ath systemcauses inactive systems (b6a) to harden more rapidly than the active system. In caseof isotropic hardening, with ~qab 1, all the systems harden at the same rate.

    Reducing the strain rate exponent m retards localization for all orientations, asexpected from the results shown inFig. 5. It also changes the sensitivity of the necking

    strain to misorientation. The underlying reason is straightforward: with high strain ratesensitivity (m = 150), the active slip systems change very rapidly as the loading axis isrotated away from the perfect [11 1] orientation. With a lower strain rate exponent(m = 20), the slip activity changes more gradually with misorientation. The influence ofthis behavior on the rotation of the crystal is shown inFig. 11, which should be contrastedwith the results shown inFig. 8for a crystal with m = 150. InFig. 8, the loading axis spi-rals around the high symmetry orientation only if the tensile axis is very close to the [111]

    Fig. 11. An inverse pole figure plot showing the rotation of the tensile axis for a crystal with various initialorientations near the [111] direction. The results are for the model including latent-hardening with strain ratesensitivitym = 20 and a specimen defect parameter g= 0.005.

    Fig. 10. (a) The variation of strain to localization with misorientation for crystals loaded with the tensile axis nearthe [111] direction with isotropic hardening and latent-hardening and (b) an enlarged view of the results for smallmisorientations.

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 65

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    18/21

    direction. In contrast, with m= 20, this behavior is observed for larger misorientations.Consequently, with m= 20 we find that the localization strain is minimized for a misori-entation around 2.

    Latent-hardening has little influence on the strain to localization if the crystal is loadedalong a high symmetry orientation. In this case, latent-hardening merely increases theoverall strain hardening behavior of the crystal, and its effects can be understood using

    the classical Considere analysis of necking (Considere, 1888). For a misoriented crystal,however, latent-hardening has a much more significant effect. For small misorientations,the material with Boyles model of latent-hardening localizes earlier than an isotropicallyhardened material as shown inFig. 12. This behavior is again a consequence of the influ-ence of hardening on the active slip systems: latent-hardening tends to prevent multipleslip in crystals that are loaded close to a high symmetry orientation. In contrast, whenthe initial loading is not close to a high symmetry orientation, latent-hardening increasesthe overshoot prior to the onset of secondary slip, and consequently increases the strain tolocalization.

    While the above results are for aluminum single crystal, they provide important clues

    about designing aluminum alloys to increase the localization strain and enhance ductility.Reducing geometric imperfections has implications for control of second phase intermetal-lic distribution. Manipulating orientation is related to controlling texture. Changing strainrate exponent can be accomplished through solid solution strengthening and precipitationhardening, which also alter latent-hardening. Assuming that the above trends for singlecrystal also holds for aluminum alloys, a combination of these microstructural featurescan lead to a high ductility material.

    5. Summary

    Finite element simulations have been used to study localization in single crystals withproperties representing Al under uniaxial tensile straining. A refined constitutive lawwhich includes a detailed description of latent and self-hardening in the crystal (Boyleand Mishra, 2005; Boyle, 2005, submitted for publication) was used in the simulation.

    Fig. 12. An inverse pole figure plot showing the rotation of the tensile axis for a crystal with various initialorientations near the [111] direction. The results are for an isotropically hardening material, deformed at 273 K

    with strain rate sensitivity m = 150 and a specimen defect parameter g= 0.005.

    66 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    19/21

    The simulations were used to investigate the influence of specimen geometry, the orienta-tion of the loading axis with respect to crystallographic directions, and the properties ofthe material, in particular, the slip rate sensitivity exponent m and the latent-hardeningratio on the strain-to-localization in uniaxial tension. Our simulations showed that:

    (1) Two general forms of localization may occur, depending on the value of the strainrate exponentm. Form< 100, failure occurs by the gradual development of a diffuseneck at the center of the specimen; for m > 100, a sharp neck forms, and if the load-ing axis is exactly parallel to the [100] direction, two necks may form. In the diffusenecking regime (m< 100) the strain to localization decreases rapidly with increasingm, while for values of strain rate exponent m> 100 the strain to localization is onlyweakly sensitive to m.

    (2) The strain to localization is sensitive to specimen geometry: for example, the strain tolocalization varies inversely with log(g), where g is a dimensionless measure of thelocal thinning at the center of the specimen, defined in Eq. (1).

    (3) The influence of the orientation of the loading axis with respect to the crystal onstrain to localization was explored in detail. Our simulations showed that crystalslocalized most readily when the tensile axis is close to (but not exactly parallel to)a high symmetry orientation such as [100] or [111]. Broadly, this behavior occursbecause a larger number of slip systems are active when the crystal is loaded nearthe [100] or [111] direction, which causes a higher rate of hardening. For larger mis-orientations, the crystal initially deforms by single slip, and rotates until a secondaryslip system is activated, whereupon the crystal tends to rotate so as to align the [112]

    direction with the loading axis. In these cases we found that localization invariablyoccurred shortly after secondary slip was activated.(4) Strain rate sensitivity has a strong influence on the orientation dependence of the

    localization strain. Small values of the strain rate exponent m (e.g. m= 20) tendto promote multiple slip, and the strain-to-localization has a minimum value whenthe loading axis is misoriented by approximately 2from a high symmetry orienta-tion. For high values of rate sensitivity, the minimum value of strain-to-localizationoccurs when the loading axis is misaligned by approximately 0.05 to the loadingaxis.

    (5) The influence of latent-hardening on the variation of strain-to-localization with mis-

    orientation was probed by contrasting the behavior of the model material withparameters representing Al with an isotropically hardening solid. In the model Almaterial, the hardening is such that slip on a single system tends to harden inactivesystems more rapidly than the active system; whereas in the isotropically hardeningmaterial, all slip systems harden at the same rate. We found that if the loading axis isperfectly aligned with a high symmetry orientation such as [100] or [111], latent-hardening has almost no effect on strain to localization. If the loading axis is misa-ligned with the high symmetry orientation by less than 2, the isotropically hardeningmaterial has a higher strain to localization than the model material with propertiesrepresenting Al. In contrast, if the loading axis is misoriented by more then 2, the

    trend is reversed. This behavior is a consequence of the influence of latent-hardeningon the rotation of the crystal. When the loading axis is very near the high-symmetryorientation, the isotropically hardening solid deforms by slip on more than two sys-tems, while the material with properties representing Al tends to deform in double

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 67

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    20/21

    slip, and localizes more readily. When the initial loading axis is misoriented by morethan 2, both materials initially deform in single slip, before eventually transitioningto double slip just prior to localization. Under these conditions higher latent-hard-ening tends to delay the onset of double slip, and consequently delays localization.

    Acknowledgement

    This work was supported by General Motors at Brown University as part of GM-Brown Collaborative Research Laboratory (CRL).

    References

    Asaro, R.J., 1979. Geometrical effects in the inhomogeneous deformation of ductile single crystals. Acta Metall.

    27 (3), 445453.

    Balke, H., Estrin, Y., 1994. Micromechanical modeling of shear banding in single crystals. Int. J. Plasticity 10 (2),

    133147.

    Boyle, K.P., 2005. Latent hardening in copper and copper alloys. Materials Science Forum, vols. 495497. Trans

    Tech Publications, Zurich-Uetikon, Switzerland, pp. 10431048.

    Boyle, K.P., submitted for publication. A quantitative crystal plasticity hardening law. Modell. Simul. Mater. Sci.

    Eng.

    Boyle, K.P., Mishra, R.K., 2005. Latent hardening in crystal plasticity. In: Dislocations, Plasticity, Damage and

    Metal Forming: Material Response and Multiscale Modeling. Neat Press, Fulton, pp. 616618.

    Considere, M., 1888. Die Anwendung von Eisen und Stahl bei Konstruktionen. Gerold-Verlag, Wien.

    Curtin, W.A., Olmsted, D.L., 2006. A predictive mechanism for dynamic strain ageing in aluminiummagnesiumalloys. Nat. Mater. 5 (11), 875880.

    Hill, R., Hutchinson, J.W., 1975. Bifurcation phenomena in the plane tension test. J. Mech. Phys. Solids 23 (45),

    239264.

    Hosford, R.L., Fleischer, R.L., Backofen, W.A., 1960. Tensile deformation of aluminum single crystals at low

    temperatures. Acta Metall. 8, 187199.

    Hosford, W.F., 1993. The Mechanics of Crystals and Textured Polycrystals. Oxford University Press, Oxford.

    Hutchinson, J.W., Neale, K.W., 1977. Influence of strain-rate sensitivity on necking under uniaxial tension. Acta

    Metall. 25 (8), 839846.

    Inal, K., Wu, P.D., Neale, K.W., 2002a. Instability and localized deformation in polycrystalline solids under

    plane-strain tension. Int. J. Solids Struct. 39 (4), 9831002.

    Inal, K., Wu, P.D., Neale, K.W., 2002b. Finite element analysis of localization in fcc polycrystalline sheets under

    plane stress tension. Int. J. Solids Struct. 39 (1314), 34693486.Kocks, U.F., Mecking, H., 2003. Physics and phenomenology of strain hardening: the fcc case. Prog. Mater Sci.

    48 (3), 171273.

    Ma, A., Rosters, F., Raabe, D., 2006. On the consideration of interactions between dislocations and grain

    boundaries in crystal plasticity finite element modeling theory, experiments and simulations. Acta Mater. 54

    (8), 21812194.

    Peirce, D., Asaro, R.J., Needleman, A., 1982. Analysis of nonuniform and localized deformation in ductile single

    crystals. Acta Metall. 30 (6), 10871119.

    Peirce, D., Asaro, R.J., Needleman, A., 1983. Material rate dependence and localized deformation in crystalline

    solids. Acta Metall. 31 (12), 19511976.

    Petryk, H., 1997. Plastic instability: criteria and computational approaches. Arch. Comput. Meth. Eng. 4 (2),

    111151.

    Shen, Y.L., 1993. An analysis of shear band formation in (100)(011) textured fcc metals. Phil. Mag. Lett. 67 (2),

    7376.

    Simmons, G., Wang, H., 1971. Single Crystal Elastic Constants and Calculated Aggregate Properties. MIT Press,

    Cambridge, MA.

    68 F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969

  • 8/13/2019 Zhang Bower Mishra Boyle IJP 2009

    21/21

    Stuwe, H.P., Toth, L.S., 2003. Plastic instability and Luders bands in the tensile test: the role of crystal

    orientation. Mater. Sci. Eng. A 358 (12), 1725.

    Taub, A.I., Krajewski, P.E., Luo, A.A., Owens, J.N., 2007. Yesterday, today and tomorrow: the evolution of

    technology for materials processing over the last 50 years: The automotive example. JOM 59 (2), 4857.

    Tvergaard, V., Needleman, A., 1993. Shear-band development in polycrystals. Proc. Roy. Soc. Lond. A 443

    (1919), 547562.

    Yang, S., Rey, C., 1995. Analysis of shear band development in a viscoplastic double slip, single crystal in tension.

    J. Appl. Mech. 62 (4), 827833.

    Yao, X.X., Zajac, S., Hutchinson, B., 2000. The strain-rate sensitivity of flow stress and work-hardening rate in a

    hot deformed Al1.0Mg alloy. J. Mater. Sci. Lett. 19 (9), 743744.

    F. Zhang et al. / International Journal of Plasticity 25 (2009) 4969 69