14.08.2013 Views

A comparative discrete-dislocation/nonlocal crystal-plasticity

A comparative discrete-dislocation/nonlocal crystal-plasticity

A comparative discrete-dislocation/nonlocal crystal-plasticity

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

Abstract<br />

Materials Science and Engineering A000 (2001) 000–000<br />

A <strong>comparative</strong> <strong>discrete</strong>-<strong>dislocation</strong>/<strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

analysis of plane-strain mode I fracture<br />

D. Columbus, M. Grujicic *<br />

Program in Materials and Engineering, Department of Mechanical Engineering, Clemson Uniersity, 241 Flour Daniel EIB, Clemson,<br />

SC 29634-0921, USA<br />

Received 22 January 2001; received in revised form 29 March 2001<br />

Crack growth associated with plane-strain mode I loading is studied computationally using both a <strong>discrete</strong>-<strong>dislocation</strong> approach<br />

and a <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> formulation. Within the <strong>discrete</strong>-<strong>dislocation</strong> approach, the material is modeled as a linear elastic<br />

solid which contains <strong>discrete</strong> <strong>dislocation</strong>s capable of moving and interacting with each other and with other lattice defects either<br />

at a distance through their stress fields, or through direct contact. In the case of <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong>, the effect of the plastic<br />

strain gradient on the materials’ behavior is incorporated through the contribution that the geometrically necessary <strong>dislocation</strong>s<br />

make to the rate of strain hardening. In both formulations, the behavior of the crack is modeled using a cohesive zone approach.<br />

The results show that while both approaches yield comparable stress and strain fields around the crack tip, the global (stress<br />

intensity versus crack extension) responses as well as the crack tip profiles can be significantly different in the two cases. These<br />

differences are related to the ways <strong>crystal</strong>lographic slip is modeled in the two approaches. © 2001 Elsevier Science B.V. All rights<br />

reserved.<br />

Keywords: Comparative <strong>discrete</strong>-<strong>dislocation</strong>/<strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> analysis; Plane-strain mode I; Fracture<br />

1. Introduction<br />

It is well established that the fracture behavior of<br />

structural materials is greatly affected by interactions<br />

between the material evolution in a region surrounding<br />

the crack tip and atomic-scale material separation processes.<br />

For example, the motion of <strong>dislocation</strong>s in the<br />

region near the crack tip results in plastic dissipation<br />

that increases the macroscopic work of separation, e.g.<br />

Refs. [1–3]. In sharp contrast, <strong>dislocation</strong> patterning,<br />

which can create a <strong>dislocation</strong>-free region ahead of the<br />

crack, and <strong>dislocation</strong> pile-ups, which give rise to<br />

strong stress concentration effects, can result in the<br />

onset of fracture at a lower level of the remotely<br />

applied stress [4,5].<br />

Most of the published work dealing with fracture of<br />

structural materials is based on the application of classical<br />

continuum <strong>plasticity</strong>. While such studies have<br />

UNCORRECTED PROOF<br />

* Corresponding author. Tel.: +1-864-6565639; fax: +1-864-<br />

6564435.<br />

E-mail address: mica@ces.clemson.edu (M. Grujicic).<br />

0921-5093/01/$ - see front matter © 2001 Elsevier Science B.V. All rights reserved.<br />

PII: S0921-5093(01)01397-1<br />

www.elsevier.com/locate/msea<br />

made a major contribution to improving the understanding<br />

of interactions between crack-tip deformation<br />

processes and separation/fracture processes, they are<br />

unable to fully account for atomic-scale separation.<br />

That is, the maximum stress level, at a blunted mode I<br />

crack tip, predicted by classical continuum <strong>plasticity</strong><br />

(3–5 times the material flow strength) is too low to<br />

cause bond breaking required for material separation<br />

[3,6]. Recently, several strain-gradient, continuum-<strong>plasticity</strong><br />

theories have been proposed, e.g. Refs. [7–9],<br />

which enabled significantly higher stress levels to be<br />

achieved in the vicinity of the crack tip. This was done<br />

by implicitly accounting for the effect of geometrically<br />

necessary <strong>dislocation</strong>s that form in the high strain-gradient<br />

region around the crack tip e.g. Ref. [10]. In<br />

addition to the strain-gradient <strong>plasticity</strong> theories, the<br />

existence of a <strong>dislocation</strong>-free elastic strip inside which<br />

the crack propagates was proposed to enable a stress<br />

build-up at the crack tip needed for atomic separation<br />

[11].<br />

In contrast to the continuum <strong>plasticity</strong> approaches<br />

discussed above, recently the interaction between dis-


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

2<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

crete <strong>dislocation</strong>s and material separation processes was<br />

modeled by Cleveringa et al. [12,13]. Cleveringa et al.<br />

[12,13] carried out a numerical analysis of the smallscale<br />

yielding for a plane-strain mode I crack under<br />

relatively high loading rates. In their analysis, <strong>discrete</strong><br />

<strong>dislocation</strong>s, represented as line discontinuities located<br />

within a linearly elastic, isotropic medium, are allowed<br />

to interact with each other through their long-range<br />

stress fields. In addition, constitutive relations are introduced<br />

to describe several short-range events such as<br />

<strong>dislocation</strong> nucleation, their annihilation and the pinning/unpinning<br />

of the <strong>dislocation</strong>s at obstacles. The<br />

velocity of <strong>dislocation</strong>s is assumed to be phonon-drag<br />

controlled and thus proportional to the Peach–Koehler<br />

force. The fracture behavior of the material is described<br />

using a cohesive-zone nonlinear constitutive model [6].<br />

Due to the nonlinear character of the cohesive-zone<br />

constitutive relation and the evolution of <strong>dislocation</strong><br />

structure with loading, the analysis is carried out in an<br />

incremental manner. At each time step, the total stress<br />

and strain fields are computed by superposing the ones<br />

associated with the <strong>discrete</strong> <strong>dislocation</strong>s and the complementary<br />

fields which enforce the imposed boundary<br />

conditions and the continuity conditions across the<br />

internal (symmetry) boundaries. The results of Cleveringa<br />

et al. [12,13] clearly demonstrated that the presence<br />

of <strong>discrete</strong> <strong>dislocation</strong>s in a region surrounding the<br />

crack tip is a prerequisite for attaining sufficiently high<br />

stress levels at the crack tip to cause atomic-scale<br />

material separation. In addition, Cleveringa et al.<br />

[12,13] showed that the fracture resistance of a material<br />

is controlled by the relative magnitudes of the loading<br />

rate, rate of <strong>dislocation</strong> nucleation, and <strong>dislocation</strong><br />

mobility. Consequently, <strong>dislocation</strong>s can either increase<br />

the fracture resistance through plastic dissipation or,<br />

conversely, promote fracture at a lower level of the<br />

remotely applied stress by giving rise to high local stress<br />

concentrations in the region ahead of the crack.<br />

In the present work, plane-strain mode I fracture is<br />

analyzed using both the <strong>discrete</strong>-<strong>dislocation</strong> approach<br />

of Cleveringa et al. [12,13], and a <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

formulation. While the <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

model can not account for the <strong>discrete</strong> nature of <strong>dislocation</strong>s<br />

and for various <strong>dislocation</strong>-based short-range<br />

events (e.g. <strong>dislocation</strong> nucleation, annihilation, etc.), it<br />

allows incorporation of the additional strain hardening<br />

effect that arises from the presence of geometrically<br />

necessary <strong>dislocation</strong>s in the regions associated with<br />

high plastic-strain gradients.<br />

The organization of the paper is as follows: In Section<br />

2.1, a detailed account is given of the boundary<br />

value problem and the computational procedure used<br />

UNCORRECTED PROOF<br />

in the <strong>discrete</strong>-<strong>dislocation</strong> analysis of plane-strain mode<br />

I fracture. A brief description of the <strong>nonlocal</strong> <strong>crystal</strong><strong>plasticity</strong><br />

model and its incorporation into the commercial<br />

finite element package Abaqus/Standard [14] is<br />

given in Section 2.2. The results of the two computational<br />

approaches are presented and discussed in Section<br />

3. The main conclusions resulting from the present<br />

study are given in Section 4.<br />

2. Analysis and computational procedure<br />

2.1. Discrete-<strong>dislocation</strong> formulation<br />

The <strong>discrete</strong>-<strong>dislocation</strong> analysis of mode I fracture,<br />

presented in the present paper, closely follows the analysis<br />

of Cleveringa et al. [12,13]. A schematic of the<br />

boundary-value problem at hand is shown in Fig. 1a.<br />

The following assumptions and simplifications are<br />

made:<br />

1. symmetry exists about the horizontal crack plane<br />

(located at x2=0) and hence only the upper or the<br />

lower half of the region around the crack tip needs<br />

to be analyzed, Fig. 1a;<br />

2. only small-scale yielding takes place and thus loading<br />

can be applied by prescribing the elastic mode I<br />

fields on the boundary of the computational<br />

domain;<br />

3. rotation effects can be neglected and hence a small<br />

strain formulation can be used;<br />

4. inertial effects can be neglected and hence a quasistatic<br />

analysis can be carried out;<br />

5. the material containing <strong>dislocation</strong>s is linearly elastic<br />

and isotropic and hence its constitutive behavior<br />

is completely defined by two elastic constants: the<br />

Young’s modulus, E, and the Poisson’s ratio, . The<br />

values of E and , as well as the values of other<br />

material parameters used in the present work, are<br />

given in Table 1; and<br />

6. the crack growth behavior can be represented using<br />

a cohesive zone crack-face traction versus crack-face<br />

separation constitutive relation.<br />

The analysis is carried out using a 1000×500 m<br />

computational domain while plastic deformation resulting<br />

from the motion of <strong>discrete</strong> <strong>dislocation</strong>s is primarily<br />

confinedtoa10×8.6 m process window (Fig. 1a).<br />

The Cartesian coordinate system is also denoted in Fig.<br />

1(a). The portion of the computational domain outside<br />

the process window is discretized using a finite element<br />

mesh consisting of 850 plane strain quadrilateral isoparametric<br />

elements (CPE4 in the Abaqus notation)<br />

(Fig. 1b). To reduce the number of degrees of freedom<br />

in the boundary value problem at hand, the size of the<br />

elements is progressively increased with their distance<br />

from the process window. The process window is discretized<br />

using a refined mesh consisting of 48×50<br />

CPE4 elements, all of the same size (Fig. 1c). In accordance<br />

with the small-scale yielding assumption, the<br />

loading is applied by imposing displacements consistent<br />

with the elastic mode I singular field on the outside


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 3<br />

Fig. 1. (a) The boundary value problem of a semi-infinite region subject to elastic mode I loading containing a cohesive zone at x 2=0 to enable<br />

crack formation and growth and a process window containing <strong>discrete</strong> <strong>dislocation</strong>s; (b) and (c) the finite element meshes for the outer region and<br />

the ‘process window’, respectively.<br />

Table 1<br />

Material, cohesive-zone, <strong>dislocation</strong> and loading parameters used in the <strong>discrete</strong>-<strong>dislocation</strong> analysis<br />

Parameter Symbol Units Magnitude<br />

Equation Where Used<br />

Young’s modulus<br />

E<br />

Gpa 70 Eqs. (5), (18) and (30)<br />

Poisson’s ratio<br />

N/A<br />

0.33<br />

Eqs. (5), (6), (12), (13), (18) and (30)<br />

Cohesive-zone strength<br />

max Gpa<br />

0.6<br />

Eqs. (1) and (30)<br />

Cohesive-zone separation n nm 1.0 Eq. (1)<br />

Slip resistance<br />

s MPa<br />

3<br />

Eq. (2)<br />

Burger’s vector magnitude b nm<br />

0.25<br />

Eqs. (2), (4), (5), (12), (13), (27) and (29)<br />

nucl m 24<br />

N/A<br />

−2<br />

Source density<br />

10E−4 Drag coefficient<br />

B Pa s<br />

Eq. (2)<br />

Mean nucleation shear stress ¯ nucl MPa<br />

25<br />

N/A<br />

S.D. of ¯ nucl<br />

0.2¯ nucl MPa<br />

5 N/A<br />

Nucleation time<br />

Loading rate<br />

tnucl K I<br />

s<br />

Gpa (m)<br />

0.01<br />

N/A<br />

Eq. (6)<br />

1/2 s−1 50<br />

UNCORRECTED PROOF


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

4<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

boundary of the computational domain (Fig. 1a). The<br />

x2=0 symmetry boundary of the computational domain<br />

is divided into three segments. The leftmost segment<br />

is modeled as the upper crack face and hence is<br />

subject to a traction free boundary condition. The middle<br />

segment is represented using cohesive-zone based interfacial<br />

elements while the rightmost segment of the x2=0 boundary is the plane of symmetry. Consequently, the<br />

middle and the rightmost segment are subject to zero<br />

x2-displacement and zero x1-traction boundary conditions.<br />

As discussed earlier, rather than employing a fracture<br />

criterion, crack growth is modeled using the cohesive<br />

zone framework [6]. The associated crack-face traction<br />

versus crack-face separation distance constitutive relation,<br />

used in the present work, is consistent with the<br />

exponential universal binding law [15] and is defined as:<br />

Un Tn(Un)=−emax exp<br />

n − Un (1)<br />

n where Tn is the normal traction, Un the normal crack-face<br />

separation distance and max and n are material-dependent<br />

cohesive-zone parameters. The cohesive-zone properties<br />

are assigned the same values as those used in Refs.<br />

[13,15] and listed in Table 1. The cohesive zone is<br />

modeled using two-node nonlinear spring-type interfacial<br />

elements whose stiffness matrix is derived using Eq.<br />

(1) and implemented in an Abaqus User Element Subroutine<br />

(UEL). Details of this procedure are given in our<br />

recent paper [16]. It should be noted that within the<br />

cohesive zone, the crack surface is defined by the top<br />

nodes of the interfacial elements.<br />

The <strong>dislocation</strong>s, treated as line singularities (i.e. point<br />

singularities within the present two-dimensional analysis),<br />

are assumed to be straight, infinitely long, of the edge<br />

character, and to have the same magnitude of Burgers<br />

vector b=0.25 nm with the line direction parallel to the<br />

x3-axis. Furthermore, they are assumed to be associated<br />

with one of the two slip systems symmetrically oriented<br />

relative to the crack plane. The process window is<br />

assumed to contain a f.c.c.-like single-<strong>crystal</strong>line material<br />

in which the slip takes place on the {1 1 1} 1 −1 0<br />

family of slip systems. The f.c.c.-like single <strong>crystal</strong> is<br />

assumed to be oriented for symmetric double slip i.e.: x1 [−1 1 0], x2 [0 0 1], and x3 [1 1 0]. Due to the<br />

plane-strain condition imposed in the x3-direction, no<br />

slip is allowed to take place on the (1 1 1) and (1 1 −1)<br />

slip planes. In addition, no slip occurs in the common [1<br />

1 0] direction of the remaining two slip planes, (1 −1<br />

1) and (−1 1 1), due to the zero resolved shear stress<br />

in this direction under uniaxial loading in the x2-direc tion. The (1 −11)and(−111) slip planes each contain<br />

UNCORRECTED PROOF<br />

two more 1 −1 0 slip directions which are symmetri-<br />

cally aligned relative to the loading (x 2) direction. Consequently,<br />

plastic deformations on the slip systems based<br />

on these directions can be assumed to be equal in<br />

magnitude; therefore, the resulting net ‘effective’ slip<br />

directions on the (1 −1 1) and (−1 1 1) slip planes are<br />

[−1 1 2] and [1 −1 2]. In addition, the crack is assumed<br />

to reside on a {0 0 1} plane. The [−1 12]and[1−1<br />

2] slip directions make an angle = tan−12 5.47° with the crack plane. To be consistent with the work<br />

of Cleveringa et al. [12,13], an angle of =60° is used<br />

in the present work.<br />

The slip-plane spacing is set to 130b resulting in 400<br />

equally spaced slip planes for each slip system. At the<br />

beginning of the simulation, the process window is<br />

assumed to be free of <strong>dislocation</strong>s, yet to contain a<br />

random distribution of <strong>dislocation</strong> sources. The density<br />

of the <strong>dislocation</strong> sources is given in Table 1. Details<br />

regarding the operation of <strong>dislocation</strong> sources, as well as<br />

the short-range <strong>dislocation</strong>-<strong>dislocation</strong> interactions<br />

which may lead to <strong>dislocation</strong> annihilation are given<br />

below.<br />

The mobility of <strong>dislocation</strong>s is generally controlled by<br />

their interactions with both short-range obstacles which<br />

can be overcome by the aid of thermal activation (thermal<br />

obstacles) and with long-range obstacles which can not<br />

be thermally overcome (athermal obstacles). In pure f.c.c.<br />

and h.c.p. materials, where the Peierls resistance is quite<br />

small, forest <strong>dislocation</strong>s are generally considered as the<br />

rate-controlling thermal obstacles and the resistance to<br />

<strong>dislocation</strong> motion shows a relatively weak temperature<br />

dependence. In b.c.c. materials, on the other hand, the<br />

Peierls resistance, which is substantial and increases<br />

rapidly with decreasing temperature, is considered to<br />

control the short-range <strong>dislocation</strong> interactions and the<br />

material strength shows a relatively strong temperature<br />

dependence. Large incoherent precipitates and groups of<br />

<strong>dislocation</strong>s are typical examples of long-range athermal<br />

obstacles.<br />

In our recent work [16], the mobility of <strong>dislocation</strong>s in<br />

b.c.c. metallic materials is studied at different temperatures.<br />

The results obtained show that thermal activation<br />

plays an important role only at temperatures a few<br />

hundred degrees above room temperature. Since all the<br />

computations in the present work are carried out for a<br />

f.c.c. material, at room temperature and at relatively high<br />

loading rates (which give rise to high stress levels), the<br />

effect of thermal activation is expected to be only minor<br />

[17,18] and a simple drag-controlled <strong>dislocation</strong>-velocity<br />

law can be used:<br />

V= effb (2)<br />

B<br />

where V is the magnitude of the <strong>dislocation</strong> velocity in<br />

the slip direction, eff the absolute value of the effective<br />

shear stress resolved on the slip plane and in the slip<br />

direction of the <strong>dislocation</strong> at hand, b the magnitude of<br />

the Burgers vector, and B is the drag coefficient. In<br />

contrast to the analysis of Cleveringa et al. [12,13], no<br />

localized obstacles to <strong>dislocation</strong> motion are considered


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 5<br />

in the present work. Instead, a constant slip resistance<br />

to <strong>dislocation</strong> motion, s, is assumed to arise from the<br />

Peierls stress and precipitates and/or preexisting <strong>dislocation</strong><br />

groups outside the process zone. In accordance<br />

with typical values for the slip resistance in f.c.c. materials,<br />

s=10 −4 ( is the shear modulus), s is set to 3<br />

MPa. The absolute value of the effective shear stress,<br />

eff, in Eq. (2) is then defined as the difference between<br />

the absolute value of the shear stress, , and s. When<br />

is less than s, the <strong>dislocation</strong> in question is assumed<br />

to be stationary.<br />

The computational analysis is carried out in an incremental<br />

manner by coupling a Fortran-based computer<br />

code, which updates the <strong>dislocation</strong> structure, with the<br />

commercial finite element program Abaqus/Standard<br />

[14], which is used to solve the resulting boundary value<br />

problem. Within each time step, the following computations<br />

are conducted.<br />

2.1.1. Update of the <strong>dislocation</strong> substructure<br />

At the beginning of a time step (time=t), the total<br />

(l) stress values at the position of each <strong>dislocation</strong> l, ij ,<br />

are determined by adding the stress values arising from<br />

(k) all remaining <strong>dislocation</strong>s kl, ˜ ij , and the stress<br />

values obtained at the end of the previous time step<br />

from the solution of the boundary value problem associated<br />

with the boundary conditions corrected for the<br />

effect of <strong>dislocation</strong>s, ˆ ij, as:<br />

(l) (k) ij (t)= ˜ ij (t)+ˆ ij(t) (3)<br />

kl<br />

Next, the shear stress at <strong>dislocation</strong> l, acting on the slip<br />

plane in the glide direction, is computed as:<br />

(l) (t)=n (l)<br />

(k) (l) i ˜ ij (t)+ˆ ij(t)b j /b (4)<br />

kl<br />

(l) (l) where n i and b j are components of the unit slip plane<br />

normal and the Burgers vector of <strong>dislocation</strong> l, respectively.<br />

The magnitude of the glide velocity of <strong>dislocation</strong><br />

l is then computed using Eq. (2). The direction of<br />

<strong>dislocation</strong> glide is based on the sign of the resolved<br />

shear stress, , and the sign of the <strong>dislocation</strong>’s Burgers<br />

vector.<br />

Under the assumption that the glide velocities of<br />

each <strong>dislocation</strong> remains constant within the time step<br />

of duration t, the locations of all <strong>dislocation</strong>s at the<br />

end of the time step (time=t=t+t) could be readily<br />

computed. However, before <strong>dislocation</strong>s are assigned<br />

their new positions, short-range <strong>dislocation</strong><br />

interactions, which may result in annihilation and in<br />

nucleation of new <strong>dislocation</strong>s, need to be considered.<br />

Should at any time within the time step the distance<br />

UNCORRECTED PROOF<br />

between two <strong>dislocation</strong>s with opposite Burgers vectors<br />

gliding on the same slip plane become smaller than a<br />

material-dependent critical distance, L e, the two <strong>dislocation</strong>s<br />

are annihilated. Following Cleveringa et al.<br />

[12,13], L e is set equal to 6b.<br />

As stated earlier, the process window is initially free<br />

of <strong>dislocation</strong>s. However, as the loading proceeds, pairs<br />

of <strong>dislocation</strong>s of the opposite Burgers vectors are<br />

allowed to generate on Frank–Read sources. Within<br />

the two-dimensional framework used in the present<br />

work, the Frank–Read sources are simulated as point<br />

sources which generate a pair of <strong>dislocation</strong>s with opposite<br />

Burgers vectors.<br />

A critical shear stress for <strong>dislocation</strong> nucleation, nucl,<br />

is chosen at random from a Gaussian distribution with<br />

a mean value, ¯ nucl=25MPa, and a standard deviation<br />

of 0.2¯ nucl and assigned to each source. In addition, a<br />

fixed time period, t nucl, is selected to mimic the time<br />

required for a three-dimensional Frank–Read source to<br />

generate a new stable <strong>dislocation</strong> loop. The nucleation<br />

process is simulated by allowing a pair of <strong>dislocation</strong>s<br />

to form every time the resolved shear stress acting on a<br />

source exceeds nucl of that source over a time period<br />

greater than t nucl.<br />

The distance between the two newly nucleated <strong>dislocation</strong>s<br />

is set to a value L nucl at which the shear stress of<br />

one <strong>dislocation</strong> acting on the other <strong>dislocation</strong> is balanced<br />

by nucl as:<br />

Lnucl= E<br />

4(1− 2 b<br />

(5)<br />

) nucl It should be pointed out that contact interactions<br />

between intersecting <strong>dislocation</strong>s on different slip<br />

planes, which give rise to the formation of sessile<br />

<strong>dislocation</strong> junctions of different strengths that can play<br />

a significant role in the kinetics of <strong>dislocation</strong> motion,<br />

are not considered explicitly. Nevertheless, it was observed<br />

that noncontact interactions of intersecting <strong>dislocation</strong>s<br />

often yields <strong>dislocation</strong> configurations of very<br />

low mobility which act as <strong>dislocation</strong> locks (sessile<br />

junctions).<br />

As <strong>dislocation</strong>s exit the process window they are<br />

retained and allowed to move and interact with other<br />

<strong>dislocation</strong>s. The effect of loading on these <strong>dislocation</strong>s<br />

is simplified by using the analytical mode-I stress field.<br />

Finally, if a <strong>dislocation</strong> glides to the crack face, it is<br />

removed from the material and its prior existence represented<br />

by a displacement step between the free surface<br />

and the remaining complementary <strong>dislocation</strong>. However,<br />

if a <strong>dislocation</strong> leaves through the closed crack<br />

surface, because of the assumed symmetry, its mirror<br />

<strong>dislocation</strong>, which prior to this event was located below<br />

the x1-axis and only considered indirectly through the<br />

symmetry boundary conditions, enters the computational<br />

domain along the mirror slip plane.<br />

2.1.2. Computation of the boundary conditions<br />

In the absence of <strong>dislocation</strong>s, loading is accomplished<br />

by prescribing elastic mode I displacements on<br />

the outer-boundary of the computational domain as:


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

6<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

K I I<br />

r<br />

u 1= 2 cos1−2+sin2<br />

2<br />

2<br />

K I I<br />

r<br />

u 2= 2 sin2−2−cos2<br />

(6)<br />

2<br />

2<br />

where<br />

2 2 r=x 1+x2 =tan−1x2 (7)<br />

x1 and the shear modulus is defined as =E/2(1+)<br />

and KI is the mode I stress intensity factor.<br />

The crack tip is taken to be initially at the origin<br />

(x1=x 2=0) and hence, zero-traction boundary conditions<br />

are prescribed along the crack face (the first<br />

segment of the x2=0 boundary) as:<br />

I I T 1(x1, x2)=T 2(x1, x2)=0 for x10; x2=0 (8)<br />

Along the middle segment of the x2=0 boundary,<br />

which is modeled as a cohesive-zone, and along the<br />

rightmost segment of the x2=0 boundary, the symmetry<br />

conditions are imposed as:<br />

I u 2(x1, x2)=0, I T 1(x1,x2)=0 for x10; x2=0 (9)<br />

When <strong>dislocation</strong>s are present in the process window,<br />

they make contributions to the displacements and the<br />

tractions along the outer-boundary and the x2=0 crack<br />

surface, hence, the (<strong>dislocation</strong>-free) boundary conditions<br />

given above need to be corrected to account for<br />

the effect of <strong>dislocation</strong>s such that the total stress and<br />

displacement fields agree with the boundary value problem<br />

shown in Fig. 1a. In the present work, the infinitemedium<br />

displacement and stress fields used are as<br />

follows:<br />

(l) (l) (l)<br />

ũ 1 (x1, x2))=cos()u 1 −sin()u 2<br />

(l) (l) (l) ũ 2 (x1, x2)=sin()u 1 +cos()u 2 (10)<br />

(l) (l) (l) ˜ 11(x1,x2)=(cos()<br />

11−sin()<br />

12)<br />

cos()<br />

(l) (l) −(cos() 12−sin()<br />

22)<br />

sin()<br />

(l) (l) (l) ˜ 22(x1,<br />

x2)=(sin() 11+cos()<br />

12)<br />

sin()<br />

(l) (l) +(sin() 12+cos()<br />

22)cos()<br />

(l) (l) (l) ˜ 12(x1,<br />

x2)=(cos() 11−sin()<br />

12)<br />

sin()<br />

(l) (l) +(cos() 12−sin()<br />

22)cos()<br />

(11)<br />

where<br />

(l) (l)<br />

b<br />

(l) (l) (l) x1 x2<br />

u 1 (x1 ,x2 )=<br />

2(1−) 2(r (l) ) 2+(1−) tan−1x(l) 2<br />

x(l)n 1<br />

(l) (l) (l) ũ2 (x1 , x2 )<br />

=<br />

b<br />

(l)<br />

x2 4(1−) r (l)2<br />

− 1<br />

2 ln(r(l) ) 2<br />

b 2 n (1−2) (12)<br />

(l) (l) 2 (l) 2 b<br />

(l) (l) (l) x2 (3(x1 ) +(x2 ) )<br />

11(x1<br />

, x2 )=−<br />

2(1−) (r (l) ) 4<br />

(l) (l) 2 (l) 2 b<br />

(l) (l) (l) x2 ((x1 ) −(x2 ) )<br />

22(x1<br />

, x2 )=<br />

2(1−) (r<br />

n<br />

(l) ) 4 n<br />

(l) (l) 2 (l) 2 b<br />

(l) (l) (l) x1 ((x1 ) −(x2 ) )<br />

12(x1<br />

, x2 )=<br />

2(1−) (r (l) ) 4 n (13)<br />

and<br />

(l) (l) (l) x1 (x1, x2)=cos()(x1−x 1 )+sin()(x2−x 2 )<br />

(l) (l) (l) x2 (x1, x2)=−sin()(x1−x 1 )+cos()(x2−x 2 )<br />

(14)<br />

and r (l) (l) 2 (l) 2 (l)<br />

=(x1 ) +(x2 ) , where x m (m=1, 2) denotes<br />

the location of <strong>dislocation</strong> l. As mentioned earlier, <br />

represents the angle that the two slip systems make with<br />

the crack plane (x2=0) and is set to 60°.<br />

The updated <strong>dislocation</strong> structure determined in Section<br />

2.1.1 can thus be used to compute the contributions<br />

of <strong>dislocation</strong>s to the displacement field as:<br />

(l) ũi(t)= ũ i (t)<br />

l<br />

(i=1, 2)<br />

and to the traction fields as:<br />

2<br />

(l) T i(t)= ˜ ij (t)hj<br />

l j=1<br />

(i=1,2); hj( j=1, 2);<br />

are the components of the outward surface normal)<br />

along the outer boundary of the computational domain<br />

and along the crack surface and the x2=0 symmetry<br />

boundary. The <strong>dislocation</strong>-free elastic mode I boundary<br />

conditions discussed above are then corrected for the<br />

presence of <strong>dislocation</strong>s as follows:<br />

1. −ũi (i=1, 2) displacements are added along the<br />

outer boundary;<br />

2. −T i(i=1, 2) tractions are added along the ‘traction-free’<br />

crack surface;<br />

3. −ũ2 and −T 1 are prescribed along the x2=0 symmetry plane<br />

4. zero-displacement conditions are prescribed along<br />

the bottom of the cohesive zone. In addition, the<br />

effective stiffness and the residual forces of the<br />

interfacial elements are modified to account for the<br />

contribution of <strong>dislocation</strong>s to the displacements<br />

and tractions along the top portion of the cohesive<br />

zone. A detailed account of this modification is<br />

given in our recent paper [16].<br />

2.1.3. Solution of the boundary alue problem<br />

Once the ‘corrected’ boundary conditions are determined<br />

at the end of the time step, Abaqus/Standard is<br />

used to solve the boundary value problem. The computed<br />

stresses ˆ ij(t) are then past back to Section 2.1.1<br />

where they are combined with the <strong>dislocation</strong>-based<br />

(k) stresses k ˜ ij (t) in accordance with Eq. (3). The<br />

effective shear stress acting on each <strong>dislocation</strong> is next<br />

computed and the procedure continued as discussed in<br />

Section 2.1.1.<br />

UNCORRECTED PROOF


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

2.2. Nonlocal <strong>crystal</strong>-<strong>plasticity</strong> formulation<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 7<br />

In this section, the same plane strain mode I crack<br />

boundary value problem shown in Fig. 1a is analyzed.<br />

However, while the crack behavior is again represented<br />

using the cohesive zone formalism, the bulk material is<br />

modeled using a rate-dependent, isothermal, elastic-viscoplastic,<br />

finite-strain, <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> formulation.<br />

The continuum mechanics and materials science<br />

foundations for this model can be traced to the work of<br />

Anand and coworkers [19,20], and Dao and Parks [10].<br />

Within the present model, the (initial) reference<br />

configuration is assumed to consist of a perfect, stressfree<br />

<strong>crystal</strong> lattice and the embedded material. The<br />

position of each material point in the reference configuration<br />

is given by its position vector X, while each<br />

material point in the current configuration is described<br />

by its position vector, x. The mapping of the reference<br />

configuration into the current configuration is described<br />

by the deformation gradient, F=x/X, a second-order<br />

tensor. Typically, in order to reach the current configuration,<br />

the reference configuration must be deformed<br />

both elastically and plastically. The total deformation<br />

gradient can be multiplicatively decomposed into its<br />

elastic, Fe , and plastic, Fp , parts as F=F e ·Fp .Inother<br />

words, the deformation of a single-<strong>crystal</strong> material<br />

point is considered to be the result of two independent<br />

atomic-scale processes: (i) an elastic distortion of the<br />

<strong>crystal</strong> lattice corresponding to the stretching of atomic<br />

bonds and; (ii) a plastic deformation associated with<br />

atomic plane slippage which leaves the <strong>crystal</strong> lattice<br />

undistorted.<br />

The present constitutive model is based on the following<br />

governing variables: (i) the Cauchy stress, T; (ii)<br />

the deformation gradient, F; (iii) <strong>crystal</strong> slip systems,<br />

labeled by integers . Each slip system is specified by a<br />

unit slip-plane normal n 0, and a unit vector m 0 aligned<br />

in the slip direction, both defined in the reference<br />

configuration; (iv) the plastic deformation gradient, Fp ,<br />

with det (Fp )=1 (plastic deformation by slip does not<br />

give rise to a volume change); (v) the slip system<br />

deformation resistance s 0 which has the units of<br />

stress; and (vi) the density of geometrically-necessary<br />

<strong>dislocation</strong>s associated with the slip systems , G.<br />

Based on the aforementioned multiplicative decomposition<br />

of the deformation gradient, the elastic deformation<br />

gradient Fe which describes the elastic<br />

distortions and rigid-body rotations of the <strong>crystal</strong> lattice,<br />

can be defined by:<br />

Fe =FFp−1 , det (Fe )0 (15)<br />

The plastic deformation gradient, Fp , on the other<br />

UNCORRECTED PROOF<br />

hand, accounts for the cumulative effect of shearing on<br />

all slip systems in the <strong>crystal</strong>.<br />

Since elastic stretches in metallic materials are generally<br />

small, the constitutive equation for stress under<br />

isothermal conditions can be defined by the linear<br />

relation:<br />

T*=C[Ee ] (16)<br />

where C is a fourth-order elasticity tensor, and E e and<br />

T * are, respectively, the Green elastic strain measure<br />

and the second Piola–Kirchoff stress measure relative<br />

to the intermediate, isoclinic configuration obtained<br />

after plastic shearing of the stress-free lattice as described<br />

by F p . E e and T * are respectively defined as:<br />

E e =(1/2){F eT F e −I}T*det(F e )F e−1 TF e−T (17)<br />

where I is the second order identity tensor.<br />

An isotropic-elasticity fourth-order tensor C, defined<br />

as:<br />

C= E<br />

1+<br />

I+ <br />

1+ IIn (18)<br />

is used in order to ensure consistency in the elastic<br />

materials response in the two (<strong>discrete</strong>-<strong>dislocation</strong> and<br />

<strong>crystal</strong>-<strong>plasticity</strong>) formulations. I in Eq. (18) is the<br />

fourth-order identity tensor and denotes the tensorial<br />

product.<br />

The evolution equation for the plastic deformation<br />

gradient is defined by the flow rule:<br />

Lp =F pFp−1 = <br />

S0 <br />

S0m0n0 (19)<br />

where Lp is the plastic velocity gradient and S0 the<br />

Schmid tensor. To obtain consistency in the plastic<br />

responses of the material in the two formulations, the<br />

same two slip systems discussed in the previous sections<br />

are used.<br />

The plastic shearing rate on a slip system is<br />

described using the following simple power-law<br />

relation:<br />

=˜ <br />

sign( ) (20)<br />

s 1/m<br />

where ˜ is a reference plastic shearing rate, and s <br />

are the resolved shear stress and the deformation resistance<br />

on slip system , respectively, and m is the<br />

material rate-sensitivity parameter.<br />

Since elastic stretches in metallic materials are generally<br />

small, the resolved shear stress on slip system can<br />

be defined as:<br />

=T* S0 (21)<br />

where the raised dot denotes the scalar product between<br />

two second order tensors.<br />

To complete the materials constitutive model discussed<br />

above, an evolution equation for the slip resistance,<br />

s , should be defined. Toward that end, the<br />

overall slip resistance of a given slip system is to have<br />

both a component arising from statistically stored <strong>dislocation</strong>s,<br />

s S , and a component due to geometrically


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

8<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

<br />

necessary <strong>dislocation</strong>s, s G. The following function for s<br />

is chosen:<br />

s 1<br />

k k =[(s S) +(s G) ] k (22)<br />

because it ensures that s s S whenever s S sG and<br />

that s s G whenever s S sG . Two values of k are<br />

generally investigated in the literature: (a) k=1 which<br />

corresponds to a superposition of the contributions of<br />

statistically stored and geometrically necessary <strong>dislocation</strong>s<br />

to the slip resistance; and (b) k=2 which, since<br />

slip resistance scales with a square root of <strong>dislocation</strong><br />

density, corresponds to a superposition of the densities<br />

of the two types of (statistically stored and geometrically<br />

necessary) <strong>dislocation</strong>s. In the present work however,<br />

only case (a), k=1, was considered.<br />

The evolution of the component of the slip system<br />

resistance associated with statistically stored <strong>dislocation</strong>s<br />

is taken to evolve as:<br />

s S= h (23)<br />

where h is an element of the matrix which describes<br />

the rate of strain hardening on the slip system due to<br />

shearing on the coplanar (self-hardening) and noncoplanar<br />

(latent-hardening) slip systems . Eq. (23) is<br />

often used within the framework of local <strong>crystal</strong> <strong>plasticity</strong>,<br />

e.g. Refs. [19,20]. Following Anand and coworkers<br />

[19,20], the following simple form for the h element of<br />

the slip system hardening matrix is adopted:<br />

h =q h (no summation on )<br />

Here, h denotes the single-slip hardening rate while<br />

q is a matrix describing the latent hardening behavior.<br />

Following Anand and coworkers [19,20], the matrix<br />

q and the single-slip hardening rate h are given as:<br />

q 1<br />

=<br />

q1=1.4 if and are coplanar slip systems<br />

otherwise<br />

(25)<br />

and<br />

h <br />

=h s S<br />

0 1− (26)<br />

s 1 r<br />

where, h 0, the initial hardening rate, s 1, the saturation<br />

slip resistance, and exponent r are respectively set equal<br />

for both slip systems (h 1 0=h 2 0, s 1 1=s 2 1).<br />

The component of the slip resistance arising from the<br />

presence of geometrically necessary <strong>dislocation</strong>s, s G,is assumed to be given by the following phenomenological<br />

equation [10]:<br />

s G=cbn G (27)<br />

where is the shear modulus, b the magnitude of the<br />

Burger’s vector, c a constant (set to 0.3 in accordance<br />

with Ashby [21]), and n G is the density of point obstacles<br />

(forest <strong>dislocation</strong>s) associated with geometrically<br />

necessary <strong>dislocation</strong>s. The short-range slip resistance<br />

arising from these point obstacles is a strong function<br />

of the type of <strong>dislocation</strong> junctions formed between<br />

mobile <strong>dislocation</strong>s on a given slip system and the<br />

forest <strong>dislocation</strong>s. Following Franciosi and co-workers<br />

[22,23], the dependence of the obstacle strength on the<br />

type of <strong>dislocation</strong> interaction is introduced through a<br />

set of interaction coefficients a . Consequently, the<br />

density of point obstacles, n G , can be expressed as:<br />

n G= a G (28)<br />

It is well established, e.g. Ref. [10], that the density of<br />

geometrically-necessary <strong>dislocation</strong>s scales with the<br />

Nye’s <strong>dislocation</strong> tensor, the tensor which is a measure<br />

of the tensorial curl of FP . Following the procedure<br />

proposed by Dao and Parks [10], which is briefly<br />

overviewed in the Appendix, an evolution equation for<br />

the density of geometrically necessary <strong>dislocation</strong>s associated<br />

with each slip system, under the plane-strain<br />

condition, is derived as:<br />

1 d<br />

G= +F21 )]<br />

[ bp 0(3) dx2 P (F 11n<br />

0(1)<br />

− d<br />

[<br />

dx1 P P (F 12n<br />

0(1) +F22n<br />

0(2)<br />

P n 0(2)<br />

)] (29)<br />

where n 0(1) and n 0(2) are components of the slip plane<br />

normal n 0, and p 0(3) is a component of p 0, a vector<br />

defined as p 0=m 0×n 0. The geometrically necessary<br />

<strong>dislocation</strong>s are of the edge character, with their line<br />

directions parallel to the x3-axis and their Burgers<br />

vectors collinear with the corresponding slip direction<br />

m 0.<br />

The stress strain relationship, Eq. (16), the flow rule,<br />

Eq. (20), and the evolution equations for the slip resistance<br />

due to statistically stored <strong>dislocation</strong>s, Eq. (23),<br />

and for the geometrically-necessary <strong>dislocation</strong>s, Eq.<br />

(27), constitute the non-local <strong>crystal</strong>-<strong>plasticity</strong> material<br />

constitutive model. The integration of the model along<br />

the loading path is carried out within the User Material<br />

Subroutine (UMAT) of Abaqus/Standard. With the<br />

exception of the density of geometrically-necessary <strong>dislocation</strong>s,<br />

all other material state variables are integrated<br />

using a forward Euler approach. Since the<br />

integration of the density of geometrically necessary<br />

<strong>dislocation</strong>s requires the computation of the gradient of<br />

state variables, which in turn depends on the material<br />

state at surrounding integration points, a backward<br />

Euler integration scheme was used to update G. Since<br />

the same integration scheme has been applied in a<br />

number of our recent studies, e.g. Ref. [24] it will not be<br />

discussed here.<br />

UNCORRECTED PROOF


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 9<br />

Fig. 2. The variation of the normalized mode I stress intensity factor<br />

with crack extension for the <strong>discrete</strong> <strong>dislocation</strong> analysis (Discrete<br />

Dislocation) and two <strong>crystal</strong> <strong>plasticity</strong> analyses (Model-I and Model-<br />

II).<br />

3. Results and discussion<br />

3.1. Discrete-<strong>dislocation</strong> analysis<br />

In this section, selected results are shown for the<br />

<strong>discrete</strong>-<strong>dislocation</strong> analysis of plane-strain mode I<br />

crack growth. The material, cohesive-zone, <strong>dislocation</strong>,<br />

and loading parameters used in this analysis are summarized<br />

in Table 1.<br />

The variation of crack extension with the applied<br />

stress intensity factor (the curve denoted ‘Discrete Dislocation’)<br />

is shown in Fig. 2. The stress intensity factor<br />

is normalized with respect to its critical value corresponding<br />

to unstable crack growth without any <strong>dislocation</strong><br />

activity, K I0, defined as [25]:<br />

KI0= En 1− 2<br />

(30)<br />

where n=e max n is the work of separation.<br />

The results shown in Fig. 2 indicate that the crack<br />

begins to extend at the normalized stress intensity factor<br />

of K I/K I0=1.10. The initial crack extension (up to<br />

ca. 0.3 m) is characterized with a very slow rise in the<br />

normalized stress intensity factor. This regime of crack<br />

extension is followed by one in which the normalized<br />

stress intensity factor rises very steeply. When K I/K I0<br />

reaches a value of approximately 2.0, the crack extension<br />

regime characterized by a slow increase in K I/K I0<br />

The <strong>dislocation</strong> structure, a contour plot of the total<br />

normal stress, 22, and the distorted finite element mesh<br />

in the process window at the onset of crack extension<br />

are shown in Fig. 3a–c, respectively.<br />

To distinguish between positive and negative <strong>dislocation</strong>s<br />

residing on the two slip systems shown in Fig. 3a,<br />

solid and open symbols are used to represent <strong>dislocation</strong>s<br />

on the =60° and =120° slip systems, respectively;<br />

while, delta and gradient symbols denote,<br />

respectively, positive and negative <strong>dislocation</strong>s on both<br />

slip systems. The results shown in Fig. 3a can be<br />

summarized as follows:<br />

1. the <strong>dislocation</strong> activity is localized on slip systems<br />

originating from the region near the crack tip;<br />

2. a substantially larger number of <strong>dislocation</strong>s are<br />

associated with the =60° slip system than with the<br />

=120° slip system;<br />

3. there is some evidence of <strong>dislocation</strong> polarization.<br />

For example, mostly negative <strong>dislocation</strong>s are found<br />

near the crack tip on both slip systems. Conversely,<br />

near the outer-boundary of the process window the<br />

<strong>dislocation</strong>s are mostly of the positive sign.<br />

In the total 22 stress field contour plot shown in Fig.<br />

3b, the darkest shade of gray is used to denote the<br />

highest positive stress values. It should be noted that<br />

despite the fact that the <strong>dislocation</strong>-based 22 stress is<br />

singular, no stress singularities are present in Fig. 3b<br />

since the stresses had to be extrapolated to the nodal<br />

points for plotting purposes. The results shown in Fig.<br />

3b can be summarized as follows:<br />

1. the stress field is roughly partitioned into three<br />

regions, with little stress variation within each region.<br />

The region ahead of the crack tip is characterized<br />

by relatively large tensile 22 stress, while the<br />

region in the crack wake contains small (mostly<br />

negative) 22 stress values;<br />

2. at many places 22 is dominated by the <strong>dislocation</strong>based<br />

˜ 22 stress which gives rise to small (diamondshaped)<br />

regions of significantly different value of 22 than that of its surroundings.<br />

The distorted finite element mesh in the process<br />

window of the computational domain at the onset of<br />

crack extension is shown in Fig. 3c. To enhance clarity,<br />

a magnification factor of 30 is applied to the nodal<br />

displacements. The location of the crack tip is denoted<br />

by an arrow. A careful examination of Fig. 3c reveals<br />

the presence of two not fully developed deformation<br />

bands emanating from the crack tip and aligned with<br />

the respective slip systems. These bands are clearly the<br />

result of the localized <strong>dislocation</strong> structure observed in<br />

Fig. 3a.<br />

The <strong>dislocation</strong> structure, the contour plot of the<br />

total normal stress, 22, and the distorted finite element<br />

mesh in the process window at the normalized stress<br />

intensity factor KI/KI0=1.48 are shown respectively in<br />

Fig. 4a–c. Based on the results shown in Fig. 2, at<br />

UNCORRECTED PROOF<br />

resumes. This period of crack growth is again pro-<br />

ceeded by the regime within which the crack-extension<br />

resistance (change of K I/K I0 per change in crack extension)<br />

is relatively high.


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

10<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

K I/K I0=1.48, the crack extension is characterized by a<br />

high fracture resistance (i.e. K I/K I0 increases sharply<br />

with crack extension).<br />

A comparison of the results shown in Fig. 3a and<br />

Fig. 4a indicates that the density of <strong>dislocation</strong>s on<br />

both slip systems has increased while their activity<br />

remains localized within two broad bands emanating<br />

from the crack tip region.<br />

By comparing the results shown in Fig. 3b and Fig.<br />

4b the following observations can be made:<br />

1. the three 22 stress regions observed at the onset of<br />

crack extension are maintained at higher values of<br />

K I/K I0;<br />

2. due to a large density of <strong>dislocation</strong>s at higher<br />

K I/K I0 values, the <strong>dislocation</strong>s’ effects on the 22<br />

stress field become more pronounced, Fig. 4b;<br />

3. the region of the highest tensile stress ahead of the<br />

crack tip, responsible for crack advance (the region<br />

denoted by the darkest shade of gray), is substantially<br />

reduced at K I/K I0=1.48, Fig. 4b, in comparison<br />

to that at the onset of crack extension, Fig. 3b.<br />

This finding is fully consistent with the results<br />

shown in Fig. 2, which indicate a sharp increase in<br />

K I/K I0. This implies that higher loading is needed to<br />

extend the region of maximum tensile stress ahead<br />

of the crack tip before significant crack extension<br />

can take place.<br />

A comparison of the results shown in Fig. 3c and<br />

Fig. 4c indicates that as loading proceeds the deformation<br />

bands become more developed. As in the case of<br />

Fig. 3c, the position of the crack tip is denoted by an<br />

arrow.<br />

The <strong>dislocation</strong> structure, a contour plot of the total<br />

normal stress, 22, and the distorted finite element mesh<br />

in the process window at the normalized stress intensity<br />

factor K I/K I0=2.10 are shown in Fig. 5a–c, respectively.<br />

At this level of K I/K I0, the crack extends relatively<br />

easily (i.e. the K I/K I0 versus a curve has a small<br />

slope, Fig. 2).<br />

Comparison of the results shown in Fig. 5a with their<br />

counterparts shown in Fig. 3a and Fig. 4a indicates that<br />

<strong>dislocation</strong> density increases fairly monotonically with<br />

UNCORRECTED PROOF<br />

Fig. 3. (a) The <strong>dislocation</strong> structure; (b) the 22 stress contour plot and (c) the deformed finite element mesh for the <strong>discrete</strong> <strong>dislocation</strong> analysis<br />

at the onset of crack extension (K I/K I0=1.10, arrows indicate position of crack tip).


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 11<br />

Fig. 4. (a) The <strong>dislocation</strong> structure; (b) the 22 stress contour plot<br />

and (c) the deformed finite element mesh for the <strong>discrete</strong> <strong>dislocation</strong><br />

analysis at the normalized stress intensity factor K I/K I0=1.48 (arrows<br />

indicate position of crack tip).<br />

The 22 stress contour plot shown in Fig. 5b displays<br />

the three familiar stress regions, and, as a consequence<br />

of a higher <strong>dislocation</strong> density, a more pronounced<br />

contribution from ˜ 22. The main change in the stress<br />

field, however, is the establishment of a larger high<br />

tensile stress region ahead of the crack tip. The lower<br />

crack resistance at this value of K I/K I0 appears to be<br />

related to the presence of this stress region.<br />

The deformed finite element mesh shown in Fig. 5c<br />

indicates a larger crack opening displacement in the<br />

crack wake. However, very little change can be observed<br />

in the shape of the crack tip in comparison to<br />

the one shown in Fig. 4c. This finding is consistent with<br />

the fact that between K I/K I0=1.48 and K I/K I0=2.10,<br />

there is very little crack extension, thus, the imposed<br />

loading is mainly absorbed by <strong>dislocation</strong>-based deformation<br />

of the material.<br />

3.2. Nonlocal <strong>crystal</strong>-<strong>plasticity</strong> analysis<br />

In this section selected results are shown for the<br />

<strong>crystal</strong>-<strong>plasticity</strong> analysis of plane strain mode I crack<br />

growth. The cohesive-zone and the loading parameters<br />

are identical to the ones listed in Table 1. As far as the<br />

<strong>crystal</strong>-<strong>plasticity</strong> parameters are concerned, it was not<br />

possible to determine a set of parameters which would<br />

replicate the KI/KI0 versus a resistance curve obtained<br />

using the <strong>discrete</strong>-<strong>dislocation</strong> approach, Fig. 2. The<br />

KI/KI0 versus a curves obtained using the <strong>crystal</strong>-<strong>plasticity</strong><br />

analysis show a monotonically increasing normalized<br />

stress intensity factor, KI/KI0, with a (as opposed<br />

to having a step-like behavior obtained in the <strong>discrete</strong><strong>dislocation</strong><br />

analysis). In addition, the overall slope of<br />

the KI/KI0 versus a curves obtained using the <strong>crystal</strong><strong>plasticity</strong><br />

analysis is generally quite small in comparison<br />

to the overall slope of the corresponding KI/KI0 versus<br />

a curve obtained using the <strong>discrete</strong>-<strong>dislocation</strong> approach.<br />

To overcome this problem, two sets of <strong>crystal</strong><strong>plasticity</strong><br />

parameters were assessed (Table 2). The first<br />

set gives rise to a KI/KI0 versus a response which best<br />

matches the initial portion of the KI/KI0 versus a curve<br />

obtained using the <strong>discrete</strong>-<strong>dislocation</strong> analysis. The<br />

second set of <strong>crystal</strong>-<strong>plasticity</strong> parameters is determined<br />

in such a way that the KI/KI0 versus a response best<br />

matches the overall KI/KI0 versus a response obtained<br />

using the <strong>discrete</strong>-<strong>dislocation</strong> approach. The resulting<br />

<strong>crystal</strong>-<strong>plasticity</strong> based KI/KI0 versus a curves are denoted<br />

in Fig. 2 as Model-I and Model-II, respectively.<br />

The equivalent plastic strain, normal, 22, stress contour<br />

plot, and the deformed finite element mesh at<br />

KI/KI0=1.18, the onset of crack extension for the<br />

Model-I <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> analysis, are shown<br />

in Fig. 6a–c, respectively. Again, the regions associated<br />

with the largest values of the equivalent plastic strain<br />

and 22 stress are denoted by the darkest shade of gray<br />

in Fig. 6a and b.<br />

UNCORRECTED PROOF<br />

K I/K I0. However, no major changes in the width of<br />

<strong>dislocation</strong> bands and the extent of <strong>dislocation</strong> polarization<br />

can be observed at high values of K I/K I0.


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

12<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

The results shown in Fig. 6a indicate that plastic<br />

deformation is localized into two deformation bands,<br />

emanating from the region surrounding the crack tip.<br />

The deformation band associated with the =60° slip<br />

system is more developed and contains the region of<br />

highest equivalent plastic strain. This finding is consis-<br />

tent with the <strong>dislocation</strong> structure at a comparable<br />

K I/K I0 level, Fig. 3a.<br />

The contour plot of the 22 stress shown in Fig. 6b<br />

indicates that the stress field around the crack tip is<br />

partitioned into three regions, each characterized by a<br />

different level of average stress. The region ahead of the<br />

Fig. 5. (a) The <strong>dislocation</strong> structure; (b) the 22 stress contour plot and (c) the deformed finite element mesh for the <strong>discrete</strong> <strong>dislocation</strong> analysis<br />

at the normalized stress intensity factor K I/K I0=2.10 (arrows indicate position of crack tip).<br />

Table 2<br />

Material parameters used in the <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> analyses (Model-I and Model-II)<br />

Parameter<br />

Symbol Units Magnitude (Model-I) Magnitude (Model-II) Equation where used<br />

s 0.001<br />

0.010<br />

−1 Reference plastic shearing rate<br />

0<br />

Eq. (20)<br />

Material rate sensitivity parameter m N/A 0.01 0.01 Eq. (20)<br />

Initial local slip resistance sS,0 MPa 11.0<br />

11.0<br />

Eq. (23)<br />

Latent hardening parameter<br />

ql N/A 1.4 1.4<br />

Eq. (25)<br />

Initial local hardening rate h MPa 380<br />

100<br />

0<br />

Eq. (26)<br />

Local saturation slip resistance s MPa<br />

1 160 160 Eq. (26)<br />

Local hardening exponent r N/A 1.2<br />

1.2<br />

Eq. (27)<br />

Ashby’s constant c N/A 0.3 0.3 Eq. (28)<br />

Interaction coefficient a11=a 22 N/A 0.45<br />

0.20<br />

Eq. (29)<br />

Interaction coefficient<br />

a12=a 21 N/A 0.45 0.20 Eq. (29)<br />

UNCORRECTED PROOF


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 13<br />

Fig. 6. (a) The equivalent plastic strain contour plot; (b) the 22 stress<br />

contour plot and (c) the deformed finite element mesh for the<br />

<strong>nonlocal</strong> <strong>crystal</strong> <strong>plasticity</strong> (Model-I) analysis at the onset of crack<br />

extension (K I/K I0=1.18, arrows indicate position of crack tip).<br />

addition, the stress levels in the corresponding three<br />

regions are comparable for the two types of analyses. A<br />

comparison of the distorted finite element mesh resulting<br />

from the <strong>crystal</strong>-<strong>plasticity</strong> analysis, Fig. 6c, with the<br />

corresponding mesh resulting from the <strong>discrete</strong>-<strong>dislocation</strong><br />

analysis, Fig. 3c, indicates very similar crack profiles<br />

in the two cases. However, the deformation bands<br />

observed in Fig. 3c, can not be readily identified in Fig.<br />

6c.<br />

To show the role of the <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

phenomenon on mode I crack growth behavior, the<br />

equivalent plastic strain, the normal, 22, stress contour<br />

plot, and the distorted mesh obtained from a <strong>crystal</strong>-<strong>plasticity</strong><br />

analysis based on Model-I parameters, but with the<br />

<strong>nonlocal</strong> parameters (c, a0 and a1) turned off, at the same<br />

level of stress intensity factor as in Fig. 6a–c (KI/KI0= 1.18), are shown in Fig. 7a–c. This analysis is referred<br />

to as a local <strong>crystal</strong>-<strong>plasticity</strong> analysis.<br />

The main differences in the distribution of equivalent<br />

plastic strain in local (Fig. 7a) and <strong>nonlocal</strong> (Fig. 6a)<br />

<strong>crystal</strong>-<strong>plasticity</strong> analyses can be summarized as follows:<br />

1. higher values of equivalent plastic strain are observed<br />

in the local-<strong>plasticity</strong> case and the highest<br />

plastic-strain region is located at the crack tip. In<br />

the <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> case, the highest plastic-strain<br />

region is slightly removed from the crack<br />

tip;<br />

2. the plastic strain bands, while well formed in both<br />

cases, appear to be narrower in the local-<strong>plasticity</strong><br />

case.<br />

A comparison of the local, Fig. 7b, and <strong>nonlocal</strong>,<br />

Fig. 6b, 22 stress distributions indicates that while the<br />

distributions are quite similar, ca. 8% higher tensile<br />

stress levels are attained in the <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

case.<br />

A comparison of the local, Fig. 7c, and <strong>nonlocal</strong>,<br />

Fig. 6c, deformed finite element meshes indicates significant<br />

differences in the crack profile especially near<br />

the crack tip. In the local <strong>crystal</strong>-<strong>plasticity</strong> case, the<br />

crack tip is quite blunted and no crack advance is<br />

observed. In sharp contrast, the crack tip is considerably<br />

sharp and crack extension is observed in the case<br />

of <strong>nonlocal</strong> <strong>crystal</strong> <strong>plasticity</strong>.<br />

The equivalent plastic strain, the normalized 22 stress contour plot, and the deformed finite element<br />

mesh at the onset of crack extension (KI/KI0=1.53) for<br />

the Model-II <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> analysis are<br />

shown in Fig. 8a–c, respectively.<br />

A comparison of the equivalent plastic strain distributions<br />

obtained using Model-II (Fig. 8a) and Model-I<br />

(Fig. 6a) <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong> analyses indicates<br />

that higher plastic strain levels are attained in the<br />

former case. In addition, the highest plastic strain region<br />

in the case of Model-II is located at the crack tip.<br />

These findings are consistent with the fact that in order<br />

to delay the onset of crack extension in the Model-II<br />

UNCORRECTED PROOF<br />

crack tip is associated with the highest tensile stress<br />

values while the region behind the crack tip contains<br />

low-magnitude primarily negative 22 stress values.<br />

These observations are fully consistent with the ones<br />

made during the <strong>discrete</strong>-<strong>dislocation</strong> analysis, Fig. 3b. In


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

14<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

analysis, the <strong>nonlocal</strong> effects were reduced. The similarities<br />

between the results shown in Fig. 8a with those<br />

shown in Fig. 7a suggest that <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

effects play a smaller role in Model II.<br />

A comparison of Model-II (Fig. 8b) and Model-I<br />

(Fig. 6b) normal, 22, stress distributions indicates that<br />

while the two stress fields are very similar, and in<br />

particular, the shapes of the regions of highest tensile<br />

stress are almost identical, somewhat (ca. 5%) lower<br />

stress levels are observed in the case of the Model-II<br />

UNCORRECTED PROOF<br />

Fig. 7. (a) The equivalent plastic strain contour plot; (b) the 22 stress<br />

contour plot and (c) the deformed finite element mesh for the local<br />

<strong>crystal</strong> <strong>plasticity</strong> analysis at the onset of crack extension (KI/KI0= 1.18, arrows indicate position of crack tip).<br />

Fig. 8. (a) The equivalent plastic strain contour plot; (b) the 22 stress<br />

contour plot and (c) the deformed finite element mesh for the<br />

<strong>nonlocal</strong> <strong>crystal</strong> <strong>plasticity</strong> (Model-II) analysis at the onset of crack<br />

extension (K I/K I0=1.53, arrows indicate position of crack tip).


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 15<br />

<strong>crystal</strong>-<strong>plasticity</strong> analysis. This finding can be attributed<br />

to the aforementioned less pronounced role of the<br />

<strong>nonlocal</strong> effects in the Model-II analysis.<br />

The crack profile, in Fig. 8c, is more blunted in<br />

comparison to that for Model-I, Fig. 6c. This is another<br />

indication of the less pronounced role of <strong>nonlocal</strong><br />

<strong>crystal</strong>-<strong>plasticity</strong> effects in the former case.<br />

Above, results shown in Fig. 6 and Fig. 8 were used<br />

to compare the corresponding fields and deformed finite<br />

element meshes from Model-I and Model-II <strong>crystal</strong><strong>plasticity</strong><br />

analyses at the respective onsets of crack<br />

extension. Therefore, the K I/K I0 values where different<br />

in the two cases. The results shown in Fig. 9a–c are<br />

used to show the effect of <strong>nonlocal</strong> <strong>crystal</strong>-<strong>plasticity</strong><br />

parameters at the same level of K I/K I0. That is, the<br />

results shown in Fig. 9a–c are obtained using the<br />

Model-I <strong>crystal</strong>-<strong>plasticity</strong> analysis at a level of the stress<br />

intensity factor of 1.53 (level of K I/K I0 at which crack<br />

advance begins for the Model-II analysis).<br />

The equivalent plastic strain distribution results<br />

shown in Fig. 9a can be summarized as follows:<br />

1. The deformation band associated with the =60°<br />

slip system broadens as the crack extends. Band<br />

broadening takes place mainly in the direction of<br />

crack extension.<br />

2. The highest plastic strain region remains removed<br />

from the crack tip, while the plastic deformation<br />

band as a whole, expands and increases in<br />

magnitude.<br />

3. The region just ahead of the crack tip is characterized<br />

by low plastic strain values. It appears that the<br />

crack tip is advancing into an elastic region.<br />

The normal, 22, stress contour plot shown in Fig. 9b<br />

indicates that the stress field possesses the same basic<br />

features previously identified in Fig. 6b, Fig. 7b, and<br />

Fig. 8b. The new feature of the stress field shown in<br />

Fig. 9b is the lateral movement of the region of maximum<br />

tensile stress with the crack tip.<br />

The deformed mesh shown in Fig. 9c, as well as the<br />

one shown in Fig. 7c indicate that the crack profile<br />

remains sharp as it advances. Again, no well-defined<br />

deformation bands can be observed.<br />

3.3. Comparison of <strong>discrete</strong>-<strong>dislocation</strong> and<br />

<strong>crystal</strong>-<strong>plasticity</strong> analyses<br />

A comparison of the corresponding results shown in<br />

Figs. 3–9 suggests that the <strong>dislocation</strong>/plastic strain<br />

and 22 stress fields as well as distorted finite element<br />

meshes and crack profiles obtained using the <strong>discrete</strong><strong>dislocation</strong><br />

and <strong>crystal</strong>-<strong>plasticity</strong> analyses are quite<br />

comparable. Nevertheless, some significant differences<br />

can be identified:<br />

1. the 22 stress fields, which in both analyses consists<br />

of three well defined regions, can be locally very<br />

different due to the ˜ 22 stress contribution in the<br />

<strong>discrete</strong>-<strong>dislocation</strong> analysis. This is particularly<br />

pronounced within the two deformation bands emanating<br />

from the crack tip region which contain high<br />

densities of <strong>dislocation</strong>s;<br />

UNCORRECTED PROOF<br />

Fig. 9. (a) The equivalent plastic strain contour plot; (b) the 22 stress<br />

contour plot and (c) the deformed finite element mesh for the<br />

<strong>nonlocal</strong> <strong>crystal</strong> <strong>plasticity</strong> (Model-I) analysis at the normalized stress<br />

intensity factor KI/KI0=1.53 (arrows indicate position of crack tip).


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

16<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

2. localization of plastic deformation in the form of<br />

deformation bands emanating from the crack tip is<br />

substantially more pronounced in the <strong>discrete</strong>-<strong>dislocation</strong><br />

case (e.g. Fig. 4c) than in the <strong>crystal</strong>-<strong>plasticity</strong><br />

case (e.g. Fig. 8c). This difference can be readily<br />

explained. In the <strong>crystal</strong>-<strong>plasticity</strong> analysis, while<br />

plastic deformation takes place on specific <strong>crystal</strong>lographic<br />

slip systems, the systems are distributed in a<br />

continuous manner throughout the material. Consequently,<br />

plastic deformation varies continuously<br />

throughout the material. In sharp contrast, in the<br />

<strong>discrete</strong>-<strong>dislocation</strong> analysis, the same <strong>crystal</strong>lographic<br />

slip systems are represented in a <strong>discrete</strong><br />

fashion. Consequently, plastic deformation is the<br />

result of the nucleation and motion of <strong>discrete</strong> <strong>dislocation</strong>s<br />

on <strong>discrete</strong> slip planes;<br />

3. while the overall (stress and strain) fields which<br />

resulted from the <strong>discrete</strong>-<strong>dislocation</strong> and the <strong>nonlocal</strong><br />

<strong>crystal</strong>-<strong>plasticity</strong> analyses are qualitatively (and<br />

to a good extent quantitatively) similar, the global<br />

(stress intensity versus crack extension) responses<br />

show significant differences. Primarily, the response<br />

obtained using the <strong>crystal</strong>-<strong>plasticity</strong> analysis is<br />

monotonic, while the one resulting from the <strong>discrete</strong>-<strong>dislocation</strong><br />

analysis has a step-wise character.<br />

In addition, the overall slopes of the K I/K I0 versus<br />

a curves differ in the two cases by a factor of three<br />

to four which can give rise to a substantial difference<br />

in the stress intensity factor at the crack extensions<br />

significantly larger than the ones analyzed in<br />

the present work. These differences can be attributed<br />

to the fact that within the <strong>discrete</strong>-<strong>dislocation</strong><br />

approach, the <strong>dislocation</strong> structure is<br />

undergoing a continuous evolution. Consequently,<br />

various <strong>dislocation</strong> configurations with greatly different<br />

net stress and displacement fields are being<br />

generated. Some of these configurations impede<br />

crack extension resulting in a steep K I/K I0 versus a<br />

response. Others, on the other hand interfere with<br />

plastic deformation promoting crack extension. In<br />

the case of <strong>crystal</strong> <strong>plasticity</strong>, due to the continuum<br />

nature of <strong>crystal</strong>lographic slip, the fields and the<br />

overall material response evolve only in a continuous<br />

manner.<br />

4. Conclusions<br />

Based on the results obtained in the present work,<br />

the following main conclusions can be drawn: (1) Mode<br />

I fracture problem in the presence of <strong>crystal</strong>lographic<br />

sip can be analyzed in a consistent manner using both<br />

UNCORRECTED PROOF<br />

<strong>discrete</strong>-<strong>dislocation</strong> and continuum <strong>crystal</strong>-<strong>plasticity</strong><br />

formulations. Within both formulations, the cohesive<br />

zone approach can be used to represent the force-displacement<br />

crack constitutive behavior. (2) When com-<br />

pared at comparable levels of the applied stress<br />

intensity factor, the <strong>dislocation</strong>/strain and stress fields,<br />

as well as distorted meshes and crack profiles are quite<br />

comparable for the two approaches. The observed differences<br />

in these fields can be readily attributed to the<br />

<strong>discrete</strong> nature of <strong>dislocation</strong>s and to the <strong>discrete</strong> nature<br />

of slip systems in the <strong>discrete</strong>-<strong>dislocation</strong> approach.<br />

(3) The main difference in the results obtained using the<br />

two formulations pertains to the global stress intensity<br />

versus crack extension response. The continuum-nature<br />

of the <strong>crystal</strong>-<strong>plasticity</strong> analysis yields a monotonic<br />

K I/K I0 versus a response. In sharp contrast, the <strong>discrete</strong><br />

nature of <strong>dislocation</strong>s and their sources, as well as<br />

the interaction between <strong>dislocation</strong>s and the advancing<br />

crack tip, gives rise to a step-like K I/K I0 versus a<br />

behavior.<br />

5. Uncited references<br />

[26].<br />

Acknowledgements<br />

The material presented here is based on work supported<br />

by the National Science Foundation, Grant<br />

Numbers DMR-9906268 and CMS-9531930 and by the<br />

US Army Grant Number DAAH04-96-1-0197. The authors<br />

are indebted to Drs Bruce A. MacDonald and<br />

Daniel C. Davis of NSF and Dr David M. Stepp of<br />

ARO for the continuing interest in the present work.<br />

The help of Dr Erik van der Giessen in providing the<br />

preprints of his work and in clarifying some aspects of<br />

the computational procedure is greatly appreciated.<br />

The authors also acknowledge the support of the Office<br />

of High Performance Computing Facilities at Clemson<br />

University.<br />

Appendix A. Derivation of an evolution equation for<br />

the density of geometrically necessary <strong>dislocation</strong>s.<br />

If dX represents an infinitesimal vector connecting<br />

two material points in the reference configuration, then<br />

its image in the intermediate configuration obtained<br />

after plastic deformation is given as:<br />

dY=F p dX (A1)<br />

Then, for an infinitesimal surface S in the reference<br />

configuration which has a unit normal r 0 and is enclosed<br />

counterclockwise by a circuit C, the total Burgers<br />

vector content of all the <strong>dislocation</strong>s piercing the<br />

area S can be evaluated using the standard right-hand<br />

Burgers circuit. The image of C in the intermediate<br />

configuration obtained after plastic shearing of the


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000 17<br />

<strong>crystal</strong> lattice, C 1 , is generally an open contour, and<br />

hence, the line integral C 1dY defines the closure failure.<br />

Since the total Burgers vector B completes the circuit<br />

C 1 , it can be defined as:<br />

B=− dY=− FpdX (A2)<br />

C 1<br />

C<br />

Eq. (A2) can be transformed into:<br />

B=− (×F pT ) Tr0 dS= r0 dS (A3)<br />

S<br />

using a generalized Stokes theorem. in Eq. (A2) is the<br />

so-called Nye’s tensor [26], and is defined as the tensorial<br />

curl of F p as:<br />

−(×F pT ) T (A4)<br />

The material derivative of the Nye’s tensor can be<br />

defined as:<br />

−(×F pT ) T (A5)<br />

which through the use of Eq. (19), becomes:<br />

=−[×(LpFp ) T ] T =−× <br />

m pnTT<br />

0n0 F<br />

(A6)<br />

Next, the portion of associated with activity of slip<br />

system is defined as:<br />

−{×[( m 0 n0 )F p ] T } T<br />

S<br />

=−m 0 [×( F pT n0 )] (A7)<br />

Thus<br />

= <br />

<br />

(A8)<br />

Next, a phenomenological equation for is defined as<br />

follows. is assumed to scale with the accumulation<br />

rate of three sets of geometrically necessary <strong>dislocation</strong>s<br />

all with Burgers vectors in the m0 direction: (a) a set of<br />

screw <strong>dislocation</strong>s with tangent vector in the m0-direc tion; (b) a set of edge <strong>dislocation</strong>s with tangent vector<br />

in the x3-direction; and (c) a set of edge <strong>dislocation</strong>s<br />

with tangent vector in the n0-direction. Hence, for a<br />

cubic lattice with a common Burgers vector magnitude,<br />

b, can be written as:<br />

=m0b( G m0+ G p0+ G n0) (A9)<br />

(s )<br />

(e )1<br />

(e )2<br />

where G , G , and G are the accumulation<br />

(s ) (e )1<br />

(e )2<br />

rates of density of the three aforementioned sets of<br />

geometrically necessary <strong>dislocation</strong>s, respectively. A<br />

comparison of the second Eq. (A7) and Eq. (A9) yields:<br />

UNCORRECTED PROOF<br />

−×( F pT n 0 )=b( G (s)m 0 + G (e)1p 0 + G (e)2n 0 ) (A10)<br />

It can be easily shown that for in-plane-strain deformation,<br />

(F P 13=F P 23=F P 31=F P 32=0, F P 33=1), Eq. (A10)<br />

yields:<br />

G (s)=0, G (e)2=0<br />

and<br />

G (e )1<br />

= 1<br />

d<br />

[ P P (F 11n<br />

0(1) +F21n<br />

0(2) )]<br />

bp 0(3) dx2 − d<br />

[<br />

dx1 P P (F 12n<br />

0(1) +F22n<br />

0(2) )] (A11)<br />

Eq. (A11) is an evolution equation for the density of<br />

geometrically necessary <strong>dislocation</strong>s and is used in Section<br />

2.2 to evaluate the non-local contribution to the<br />

slip resistance.<br />

References<br />

[1] G.R. Irwin, Fracture, in: S. Flugge (Ed.), Encyclopedia of<br />

Physics, vol. 6, Springer New York, New York, 1958, pp.<br />

551–590.<br />

[2] L.B. Freund, The mechanics of <strong>dislocation</strong>s in strained-layer<br />

semiconductor-materials, Adv. Appl. Mech. 30 (1994) 1–66.<br />

[3] V. Tvergaard, J.W. Hutchinson, The relation between crack<br />

growth resistance and fracture process parameters in elastic-plastic<br />

solids, J. Mech. Phys. Solids 40 (1992) 1377–1397.<br />

[4] I.-H. Lin, R. Thomson, Cleavage, <strong>dislocation</strong> emission, and<br />

shielding for cracks under general loading, Acta Metall. 34<br />

(1986) 187–206.<br />

[5] V. Shastry, P.M. Anderson, R. Thomson, Nonlocal effects of<br />

existing <strong>dislocation</strong>s on crack-tip emission and cleavage, J.<br />

Mater. Res. 9 (1994) 1–16.<br />

[6] A. Needleman, A continuum model for void nucleation by<br />

inclusion debonding, J. Appl. Mech. 54 (1987) 525–531.<br />

[7] N.A. Fleck, J.W. Hutchinson, Strain gradient <strong>plasticity</strong>, Adv.<br />

Appl. Mech. 33 (1997) 295–361.<br />

[8] N.A. Fleck, G.M. Muller, M.F. Ashby, J.W. Hutchinson, Strain<br />

gradient <strong>plasticity</strong>: theory and experiment, Acta Metall. Mater.<br />

42 (1994) 475–487.<br />

[9] A. Acharya, J.L. Bassani, Incompatibility and <strong>crystal</strong> <strong>plasticity</strong>,<br />

J. Mech. Phys. Solids (1998) (in press).<br />

[10] H. Dao, Parks, Geometrically necessary <strong>dislocation</strong> density in<br />

continuum <strong>plasticity</strong> theory, FEM implementation and application,<br />

PhD Thesis, MIT, August 1997.<br />

[11] Z. Suo, C.F. Shih, A.G. Varias, A theory for cleavage cracking<br />

in the presence of plastic flow, Acta Metall. Mater. 41 (1993)<br />

1551–1557.<br />

[12] H.H. M. Cleveringa, E. van der Giessen, A. Needleman, A<br />

<strong>discrete</strong> <strong>dislocation</strong> analysis of mode I crack growth, J. Mech.<br />

Phys. Solids (in press)<br />

[13] H.H. M. Cleveringa, E. van der Giessen, A. Needleman, A<br />

<strong>discrete</strong> <strong>dislocation</strong> analysis of rate effects in mode I crack<br />

growth, Mater. Sci. Eng. (in print).<br />

[14] ABAQUS Theory Manual, Version 5.8, Habbilt, Karlsson, and<br />

Soreassen, Providence, RI, 1999.<br />

[15] J.H. Rose, J. Ferrante, J. Smith, Universal binding energy curves<br />

for metals and bimetallic interfaces, Phys. Rev. Lett. 47 (1981)<br />

675–678.<br />

[16] M. Grujicic, D. Columbus, Discrete-<strong>dislocation</strong> analysis of ductile-to-brittle<br />

transition in BCC-like metallic materials, J. Mater.<br />

Sci. (2000) (in print).<br />

[17] A.S. Krausz, H.E. Eyring, Deformation Kinetics, Wiley, New<br />

York, 1975.<br />

[18] V.F. Kocks, A.S. Argon, M.F. Ashby, Thermodynamics and<br />

Kinetics of Slip, Pergamon, London, 1975 in Progress in Materials<br />

Science.


typeset2:/sco3/jobs1/ELSEVIER/msa/week.17/Pmsa15088y.001 Wed May 16 07:53:37 2001 Page Wed<br />

18<br />

D. Columbus, M. Grujicic / Materials Science and Engineering A000 (2001) 000–000<br />

[19] S.R. Kalidindi, C.A. Bronkhorst, L. Anand, J. Mech. Phys.<br />

Solids 40 (1992) 537–569.<br />

[20] C.A. Bronkhorst, S.R. Kalidindi, L. Anand, Phil. Trans. R. Soc.<br />

Lond. A341 (1992) 443–477.<br />

[21] M.F. Ashby, The deformation of plastically non-homogeneous<br />

materials, Phil. Mag. 21 (1970) 399–424.<br />

[22] P. Franciosi, M. Berveiller, A. Zaoui, Latent hardening in copper<br />

and aluminum single <strong>crystal</strong>s, Acta Metall. 28 (1980) 273–<br />

283.<br />

[23] P. Franciosi, A. Zaoui, Multislip in FCC <strong>crystal</strong>s: a theoretical<br />

approach compared with experimental data, Acta Metall. 30<br />

(1982) 1627–1637.<br />

[24] M. Grujicic, Y. Zhang, J. Mater. Sci. 34 (1999) 1419–1437.<br />

[25] J.R. Rice, A path independent integral and the approximate<br />

analysis of strain concentration by notches and cracks, J. Appl.<br />

Mech. 35 (1968) 379–386.<br />

[26] J.F. Nye, Some geometrical relations in dislocated <strong>crystal</strong>s, Acta<br />

Metall. 1 (1953) 153–162.<br />

UNCORRECTED PROOF<br />

.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!