Você está na página 1de 30

International Journal of Plasticity 26 (2010) 12801309

Contents lists available at ScienceDirect

International Journal of Plasticity


journal homepage: www.elsevier.com/locate/ijplas

A perspective on trends in multiscale plasticity


David L. McDowell *
GWW School of Mechanical Engineering, School of Materials Science & Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405, USA

a r t i c l e

i n f o

Article history:
Received 28 November 2009
Available online 4 March 2010
Keywords:
Plasticity
Multiscale
Gradient
Scale effects
Grain boundaries

a b s t r a c t
Research trends in metal plasticity over the past 25 years are briey reviewed. The myriad
of length scales at which phenomena involving microstructure rearrangement during plastic ow is discussed, along with key challenges. Contributions of the authors group over
the past 30 years are summarized in this context, focusing on the statistical nature of
microstructure evolution and emergent multiscale behavior associated with metal plasticity, current trends and models for length scale effects, multiscale kinematics, the role of
grain boundaries, and the distinction of the roles of concurrent and hierarchical multiscale
modeling in the context of materials design.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Understanding of metal plasticity has advanced considerably in the past 25 years. In terms of topics, a quarter of a century
is long enough to observe both short wavelength cycles that involve revisiting certain topics and introducing improved
tools and frameworks as well as longer term transformational trends. Prior to 1980, the study of plasticity was approached
mainly through experimental means, with complementary, foundational theories based on the macroscopic theory of rateindependent plasticity and materials science foundations of dislocation mechanics. In deference to the many body nature of
defect elds in larger scale structural problems, focus was placed on issues such as the existence and role of vertices in yield
surfaces, yield surface anisotropy, associated versus nonassociated ow potentials, strain localization, and early work in coupling pressure dependence of yield with void nucleation and growth. Two transformational trends of the last 25 years have
moved the subject in new directions: (i) development of ubiquitous high resolution characterization tools that measure attributes of microstructures relating directly to plastic ow (e.g., orientation, disorientation, grain/phase size and shape distribution, dislocation density and conguration, etc.), and (ii) development of computational modeling and simulation tools
that address inelastic deformation phenomena over a broad range from atomistic to structural length scales. From this
authors perspective, signicant trends in research in metal plasticity over the past 25 years are listed as follows, with most
initiated in the 510 years preceding each respective era (McDowell, 2008a).
19801990
 Yield surface anisotropy
 Complex strain and temperature history effects
 Plastic spin: material spin versus substructure spin, objectivity
 Void growth and compressibility effects
 Localization conditions and criteria

* Tel.: +1 404 894 5128; fax: +1 404 984 0186.


E-mail address: david.mcdowell@me.gatech.edu
0749-6419/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2010.02.008

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1281

19851995
 Homogenization theories
 Eshelbian micromechanics and crystal plasticity
 Early computational micromechanics
 Gradient and nonlocal theories
 Cyclic hardening and transient behavior; ratcheting phenomena and thermomechanical behavior
 Initial forms of discrete dislocation plasticity
 Large scale crystal plasticity
19952005
 Discrete dislocation dynamics large scale
 Crystal plasticity studies of yield criteria
 Renement of computational micromechanics and integration with atomistics
 Scaling laws for grain renement
 Integration of dislocation theory and enhanced kinematical descriptions in crystal plasticity; nonlocal/gradient plasticity and damage evolution
 Hierarchical multiscale modeling methods
 Quasicontinuum method and coupled atomistics/continuum
2000present
 Crystal plasticity studies of yield criteria and ow
 Linkage of discrete dislocation theory and Nye tensor concepts in continuum eld models
 Atomistic-continuum coupling methods; atomistically informed models
 Statistical micromechanics
 Superposition of electron backscatter images with simulated elds of view for cyclic and monotonic plasticity
 Dislocation substructure and grain boundaries
 Statistical dislocation dynamics, scaling laws and self organized criticality
 Hierarchical and concurrent multiscale modeling
 Origins of scale effects
 Statistical micromechanics
 Accelerated MD and coupled MD-continuum methods
 Plasticity in materials design.
In the context of this article, liberty is taken to focus more on the work of the authors group than offering a broad review
of the literature in these areas. It is most interesting and relevant to point to some of the more compelling unresolved issues
that have been identied over the past 25 years in our research and in companion efforts of others that motivate continued
exploration. Topics to be considered include:
 Homogenization concepts for evolving hierarchical, heterogeneous microstructures
Isoscale non-associativity
Multiscale non-associativity





Self-organization and scaling laws in grain subdivision


Multiscale nature of the multiplicative decomposition of the deformation gradient
Grain boundaries in polycrystals
Plasticity in materials design.

2. Multiple scales and implications for modeling


Metal plasticity is fundamentally associated with the nucleation and migration of dislocations in crystals and polycrystals
as the underlying basis for microstructure rearrangement or evolution during plastic ow. Viewed at different length scales,
as shown in Fig. 1, it is clear that one moves from dynamics of a particle system (atomistics) to dislocation line segments
(discrete dislocations) to self-organized patterns of dislocations (substructure) to heterogeneous ow within sets of crystals
(polycrystal plasticity) to collective effects at hundreds of microns and above (macroscopic theory of plasticity). The latter
theories are of reduced order and are formulated on the basis of consistency of thermodynamics and kinetics with the underlying many body defect eld problem. As may be apparent, the aforementioned trends have more or less progressed with
time from right to left on this hierarchy, and are now focused on connecting them in both hierarchical and concurrent ways
(multiscale modeling). It should be borne in mind that models for the various levels of hierarchy in Fig. 1 are quite different,
often in fundamental ways. In fact, the corresponding suite of models is highly heterogeneous in its own right. The number of
degrees of freedom associated with description at each scale decreases for the same volume of material as one proceeds from
left to right in Fig. 1. The reduction of model degrees-of-freedom (DOF) for a given material volume with increasing level of

1282

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Statistical theories

Atomistic

Min. Length
Scale, L

O(10-10 m)

dynamic

Discrete
dislocations

O(10-8 m)

Dislocation
patterns

Polycrystal
plasticity

Macroscale
plasticity

O(10-7 m)

O(10-5 m)

O (10-3 m)

thermodynamic

Fig. 1. Hierarchy of length scales in metal plasticity ranging from atomic (dislocation cores) to patterns of dislocations to multiple grains to macroscopic
scale. The primary gap in modeling and simulation in multiscale modeling exists between the scales of atomistic simulations and dislocation pattern
modeling, with discrete dislocation simulation playing an increasing role but phenomenological statistical theories presently dominate.

hierarchy, as shown in Fig. 1, is a foundational characteristic of models framed at successively increasing levels of hierarchy.
Information necessary to calibrate model parameters can ow from bottomup, topdown, or in both directions. The latter is
typical of concurrent multiscale models, while bottomup and topdown ows are characteristic of hierarchical multiscale
models which employ various methods of coupling among scales, ranging from weak (information passing) to strong
(homogenization theory). Although the minimum length scale among windows in Fig. 1 increases by seven orders of magnitude, model degrees-of-freedom necessary to describe a material volume of 1 mm3 exhibit an inverse trend, with the
reduction of DOF collapsing by 12 orders of magnitude by moving from dynamical treatments such as molecular dynamics,
that account for particle position and momentum, to thermodynamic models based on the concept of a representative volume element (RVE). Reduced order models are sought for increasing scale of application for several reasons, not only to mitigate the intense nature of computation but also in view of the difculty of prescribing initial conditions and boundary
conditions for many-body dynamic defect elds.
The transition from dynamic to thermodynamic description pertains to the classication of the corresponding constitutive model representation of the evolution of microstructure associated with plastic ow. It turns out that simulations at the
atomic scale at 0 K (either based on some form of ground state density functional theory or atomistic simulations with energy minimization based on interatomic potentials) are useful in establishing equilibrium thermodynamics quantities,
including stable phases and related properties, energy functions for defects or interfaces, as well as transition states for unit
processes involved in evolution of microstructure. The atomistic to mesoscopic continuum description is characterized as
from dynamic to thermodynamic by virtue of discarding degrees of freedom that represent the complete dynamical state.
Beyond scientic curiosity, practical applications have driven the inquiry into higher delity models and ner scales. Progress in the last quarter century has been remarkable in this regard, with the momentum of continuum micromechanics in
the 1980s building towards integration with atomistic simulations in the 1990s to the present. Quantifying inelastic deformation at microstructure scales is key to next generation deformation and failure models with more predictive character.
Moreover, multi-mechanism and multi-physics phenomena require more rigorous linkage to unit process models and interactions that govern over the range of scales shown in Fig. 1. Prospects for manipulating microstructure to achieve enhanced
performance demand models that distinguish between nucleation, migration, absorption/desorption, trapping, and bypass or
annihilation of dislocations at various material length scales that manifest work hardening behavior. Most of these phenomena cannot be considered in isolation, which is a hallmark of plasticity it is a highly coupled phenomenon, in general, with
important effects of both short and long range character attributed both to the physics of dislocation cores and long range
interactions of dislocation arrays. Furthermore, it is often the case that behavior at any one of the scales in Fig. 1 depends on
phenomena occurring at yet ner scales.
There are several compelling reasons to develop multiscale models of plasticity (McDowell, 2001, 2008a,b; McDowell
et al., 2007; McDowell and Olson, 2008):
 to support design of polycrystalline and/or polyphase microstructures with tailored properties related to plastic deformation such as yield strength, true fracture ductility, fatigue and ductile fracture,
 to facilitate accelerated insertion of materials by coupling process route with properties in a more predictive way by
levering modeling and simulation,
 to support microstructure-sensitive failure prediction that may assist in prognosis and design of various products and
systems,
 to quantify inuence of environment or complicating contributions of impurities, manufacturing induced variability or
defects,
 to objectively allow for competition of distinct mechanical, chemical and transport phenomena without relying too heavily on intuition to guide solutions in the case of highly nonlinear interactions, and

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1283

 to develop self-consistent theories for the nucleation and evolution of defects during plastic ow at various length scales,
bridging domains of quantum physics and chemistry with engineering.
Bottom-up modeling of plasticity of realistic multiphase, hierarchical microstructures remains a grand challenge problem. Hierarchy plays a central role in addressing practical engineering applications. It is unlikely that analysis at a single window of resolution in Fig. 1 will sufce to describe all potential applications. For example, modeling the detailed
congurations of defect nucleation and migration in thin lms for electronic applications may warrant use molecular
dynamics or discrete dislocation models in view of the sensitivity of functionality to dislocation density. At the other end
of the spectrum, crystal plasticity framed at the micron scale sufces for many applications such as forging, sheet metal
forming, or wire drawing in view of the role of crystallographic texture; it is often applied in top-down fashion, calibrated
to experiments. Hence, the need to pursue bottom-up modeling depends on the application and degree of experimental support that can be provided.
2.1. Scale transition methods
Plasticity is a multiscale, multi-mechanism phenomenon (Ziegler, 1983; Drucker, 1984) manifested of irreversible
microstructure rearrangement (McDowell, 2005) associated with line and point defects in crystals. In this regard, our
interest is drawn to scaling up models for evolving rather than stationary microstructures, a goal that contrasts with
the notion of homogenization of effective properties such as elastic stiffness of composite materials (McDowell,
2008a,b).
Details of various schemes for multiscale model transitions in plasticity are described elsewhere (McDowell, 2008b). Most
have developed in the last 15 years, in some cases with the feasibility of concepts dating back perhaps 20 years. We point out
several important classes below with balance of emphasis on earlier contributions in each case, noting that our categorization is not necessarily unique:
(a) Discrete to continuous modeling approaches based on contiguous or overlapping domains (atomic structure representation to continuum representation)
 Domain decomposition, with atomistic domain abutted to continuum domain, partial overlap (Gumbsch, 1995;
Weinan and Huang, 2001; Shilkrot et al., 2002, 2004; Shiari et al., 2005; Qu et al., 2005), with both quasistatic
and dynamic applications involving dislocation nucleation and migration.
 Quasicontinuum (Tadmor et al., 1996a,b; Shenoy et al., 1998, 1999; Miller et al., 1998) approaches with the continuum elastic energy function informed by nonlocal atomistic interatomic potentials, suitable for framing long
range continuum formulation consistent with elds treated with atomistics near interfaces or crack tips.
 Dynamically equivalent continuum or quasicontinuum, consistent with fully dynamic atomistic simulations over
same overlapping domain (Zhou and McDowell, 2002; Curtarolo and Ceder, 2002; Muralidharan et al., 2003; Chung
and Namburu, 2003; Liu et al., 2003, 2006; Park et al., 2005; Zhou, 2005; Dupuy et al., 2005; Zhou et al., 2005;
Ramasubramaniam et al., 2007; Chen et al., 2009).
 Coarse-grained molecular dynamics (Rudd and Broughton, 1998, 2000, 2005; Rai-Tabar et al., 1998; Kulkarni
et al., 2008).
 Continuum phase eld theory (Khachaturyan et al., 2001; Chen, 2002; Shen and Wang, 2003), informed by ab initio
or atomistic simulations in terms of energy functions, with kinetics (mobilities) established by matching overlapping MD or experiments (Belak et al., 2009).
(b) Self-consistent micromechanics schemes for multiscale homogenization of continuum formulations based on
eigenstrain eld mechanics (Suquet, 1987; Mura, 1987; Lebensohn et al., 2004).
(c) Concurrent multiscale continuum modeling schemes for plastic deformation and failure processes (Ghosh et al., 2001,
2007; Liu et al., 2004, 2006; Nuggehally et al., 2007).
(d) Multiscale modeling schemes of hierarchical nature:
 Hierarchical, multilevel models for plasticity and fracture processes at multiple length scales of material heterogeneity based on informing model structure and parameters via optimization, informatics or handshaking (Hao et al.,
2003, 2004; Wang et al., 2006; Shenoy et al., 2008; Groh et al., 2009).
 Internal state variable models informed by atomistic/continuum simulations and coupled with micromechanics or
FE methods (Warner et al., 2006; Capolungo et al., 2007).
 Weak form variational principle of virtual velocities (PVV), including stochastic aspects associated with thermally
activated processes incorporated with addition of enhanced kinematics based on microstructure degrees of
freedom (Needleman and Rice, 1980; Cocks, 1996; Moldovan et al., 2002; Cleri et al., 2002).
 Extended weak form variational approaches of generalized continua or second gradient theory with conjugate
microstresses to represent gradient effects (Kouznetsova et al., 2002, 2004; Larsson and Diebels, 2007; Vernerey
et al., 2007; Luscher and McDowell, 2009; Luscher et al., submitted for publication).
(e) Statistical continuum theories that directly incorporate discrete dislocation dynamics (Amodeo and Ghoniem, 1988;
Kubin and Canova, 1992; Groma, 1997; Zaiser, 2001; Arsenlis and Parks, 1999, 2002; Zbib et al., 2002; Zbib and de
la Rubia, 2002; Zbib and Khaleel, 2004; Rickman and LeSar, 2006; Ohashi et al., 2007).

1284

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

(f) Nonlinear dislocation eld theory based on incompatibility with dislocation ux and Nye transport (Acharya, 2003;
Acharya et al., 2006, 2008; Roy and Acharya, 2006; Roy et al., 2007).
(g) Nonlocal reactiondiffusion and reactionconvection models directed towards formation of dislocation patterning and
substructure formation (Bammann and Aifantis, 1982; Walgraef and Aifantis, 1985; Aifantis, 1987, 1995).
(h) Dislocation substructure formation models based on fragmentation and blocky single slip or laminated ow characteristics, including non-convex energy functions (Kuhlmann-Wilsdorf, 1989; Kratochvil, 1990; Leffers, 1994; Nazarov
et al., 1996; Hughes et al., 1997; Horstemeyer and McDowell, 1998; Butler and McDowell, 1998; Ortiz and Repetto,
1999; McGinty and McDowell, 1999; Zaichenko and Glezer, 1999; Ortiz et al., 2000; Seefeldt et al., 2001; Seefeldt,
2001; Kalidindi, 2001; Carstensen et al., 2002; Pantleon, 2002; Seefeldt and Klimanek, 2002; Kratochvil and Sedlacek,
2003; Sedlacek and Kratochvil, 2005; Kratochvil et al., 2007).
(i) Transition state theory models, with atomistic unit process models (Li, 2007; Zhu et al., 2007, 2008) informing understanding and methods such as kinetic Monte Carlo (Deo and Srolovitz, 2002) and other coarse graining approaches (Li
et al., 2007).
We exclude in the above listing the many models of nonlocal and gradient type that have emerged from an understanding
of the role of ner scale heterogeneity in patterning of dislocations on response, but do not explicitly address the treatment
of averaging or homogenization processes in the scale transition. Such approaches are implicitly multiscale (as of course, are
most constitutive models of continuum type). These include, for example:
 Nonlocal, generalized continua stemming from the early work of Eringen applied to plasticity (Eringen and Claus, 1970;
Eringen, 1972, 1999; Forest et al., 2000; Forest and Sievert, 2003).
 Strain gradient theory (Dillon and Kratochvl, 1970; Dillon, 1977) framed in the context of compatibility of the total deformation gradient, with accommodation of incompatibility of the plastic deformation gradient arising from geometrically
necessary (net) Burgers vector of populations of dislocations, giving rise to lattice curvature (Nye, 1953; Bilby and Smith,
1956; Ashby, 1970; Zbib and Aifantis, 1992; Fleck and Hutchinson, 1993; Fleck et al., 1994; Nix and Gao, 1998; Gao et al,
1999; Acharya and Bassani, 2000; Hutchinson, 2000; Gurtin, 2002; Aifantis, 2003; Liu et al., 2003; Huang et al., 2004;
Evers et al., 2004; Bayley et al., 2006; Gurtin and Anand, 2007; Abu Al-Rub et al., 2007; Viatkina et al., 2007; Gerken
and Dawson, 2008).
Mesarovic (2005) has written cogently on the need for including elastic energy associated with long range interactions of
dislocations to frame more complete continuum theories in terms of size effects. Simple gradient theories offer a move in
this direction, but must be augmented by consideration of longer ranged interactions of net Burgers vector (Bayley et al.,
2006; Viatkina et al., 2007; Gerken and Dawson, 2008).
To summarize, we may consider that no single multiscale modeling strategy exists or sufces to wrap together the various scale transitions discussed here. As shown in Fig. 2, atomistics can be used to inform microscopic phase eld models as
an intermediate continuum description, or atomistic simulations can be used to support continuum criteria for the onset of

Atomistics
interfaces
Continuum

Phase Field
Models

Atomistics
Micromechanics;
Statistical mechanics

Continuum

Grain Size
GB network
Dislocations
Disclinations
Free Volume

Discrete defects;
generalized
continuum ISVs

Constitutive
Equations

Fig. 2. Connectivity of schema for bridging scales in Fig. 1 that integrates various types of heterogeneous models.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1285

plastic deformation and work hardening. Such criteria can either be framed in terms of phenomenological internal state variable descriptions at higher scale or used at a primal scale as part of a self consistent multiscale modeling strategy.
2.2. Nuances of homogenization of evolving microstructures and resulting scale effects
The majority of the aforementioned approaches for scale transition do not address invariance of all elements of balance of
mass, momentum and energy across scales of description shown in Fig. 1. For evolving microstructures, the rate of release
of free energy (or dissipation) is also important in order to sort out stored elastic energy (Luscher and McDowell, 2009).
Moreover, ux exchange of dislocations (Arsenlis and Parks, 1999, 2002; Acharya, 2003; Acharya et al., 2006) is relevant
to nonlocal constitutive descriptions of plasticity.
In some frameworks for scale transition, such as homogenization theory, many hierarchical or concurrent multiscale
models or internal state variable models, an implicit assumption is made that generalized normality holds at the aggregate
(highest scale) if it holds at the primitive level of a single crystal. This includes normality of the inelastic strain rate to the
yield function, which therefore also serves as a the (visco)plastic potential, as well as all generalized thermodynamic force
ux relations. However, there are important implications of microstructure heterogeneity and interdependence of mechanisms for the breakdown of the generalized normality hypothesis, even when conservative dislocation glide is dominant
at the primitive scale of response to be considered in a hierarchy of scales. Furthermore, if non-associativity breaks down
at any one scale, it is manifested as non-associativity at all higher scales to some degree. We shall discuss these aspects
briey in the following.
2.2.1. Isoscale non-associativity
An important case is characterized by a thermodynamic ux that is not uniquely driven by a conjugate thermodynamic
force at a given length scale. In the presence of cross-dependencies of the thermodynamic forces and uxes at a given scale, it
is not possible to construct a dissipation potential which serves as the basis for a generalized normality structure for plastic
ow and evolution of internal variables that reect evolving microstructure. In this case, non-associated microstructure evolution is manifested at all scales above this scale and cannot be diminished by increasing the size of the sampling volume. A
good example is dislocation core spreading (non-Schmid effect) in bcc, hcp and low stacking fault energy (SFE) fcc metals
the core must constrict for the dislocation to migrate (Paidar et al., 1984; Qin and Bassani, 1992; Dao and Asaro, 1993; Dao
et al., 1996; Bassani et al., 2001). Fig. 3 presents recent work of Racherla and Bassani (2007), framed at the polycrystalline
level. The yield surface F includes the third invariant of deviatoric stress to reect tension-compression asymmetry arising
from core constriction necessary to mobilize existing dislocations. Once constricted, the mobile dislocations behave in accordance with a simple von Mises ow potential G. The resulting non-associated structure leads to serrated ow in a problem
involving slip banding, in contrast to associated ow. This behavior is observed whether or not a second gradient formulation
is employed to regularize the solution by alleviating mesh dependence.
Another example is characteristic of problems in which defect nucleation (dislocations, voids) plays a comparable role to
growth in terms of contributing to the rate of inelastic strain evolution. Such cases arise, for example, when continuous
nucleation of voids plays an important role in coupling with plastic deformation. An example is that of void nucleation
and growth in steels under tensile stretching with signicant levels of superimposed hydrostatic compression. McDowell
et al. (1993) conducted experiments of this type on cold-formable steel (38C4), provided by IRSID/USINOR-SACILOR, measuring the number density of voids nucleated at MnS inclusions and void volume fraction at the center of the gage section
of specimens shown in Fig. 4 at levels of equivalent logarithmic tensile strain in the gage section ranging up to 100%. By modeling both nucleation and growth processes consistent with these experiments (Marin and McDowell, 1996, 1998), they
showed that the nucleation and growth must be addressed with distinct terms/potentials, comprising an inherently nonassociated structure to void growth and coupling with plasticity. It was not possible to model coupled plastic ow and void
growth using associated yield functions and ow potentials irrespective of nucleation. In essence, void nucleation is a phase
transition with different thermodynamics and kinetics than growth. Most associated plasticity theories for compressible
plasticity of metallic systems either neglect continuous void nucleation or somehow assume that the resulting non-associativity is negligible.
Yet another example is that of dislocation nucleation at small scales of highly stressed volume, as may arise in cases of
nanocrystalline materials with grain size below 5075 nm, incipient processes of plastic ow at crack tips, or in nanoindentation. In such cases, the nucleation of dislocations controls the onset of plastic ow in view of limited density of sources. In
the case of nanocrystalline materials, the nucleation of partial dislocations is dominant for grain sizes below 30 nm or so,
depending on stacking fault energy. In any case, dislocation nucleation has strong non-Schmid character (Spearot et al.,
2007a,b; Tschopp et al., 2007, 2008; Tschopp and McDowell, 2007a,b,c,d, 2008a,b,c; McDowell, 2008b) with coupling of resolved shear stress and normal stress the plane of the stacking fault or primary slip plane, contributing to isoscale non-associativity. The same results can be obtained using static methods based on hyperelastic lattice instability at the onset of
nucleation, for example the Interatomic Potential Finite Element Method (IPFEM) (Li et al., 2002; Zhu et al., 2004a,b). The
same type of result should apply to multilayered nanostructures (e.g., fccfcc or fcc/bcc multilayers, cf. Friedman and Chrzan
(1998)) and nanotwinned materials, namely that the response at all higher length scales will be non-associated in terms of
initial yield and continued ow. Some of these systems have relatively high ductility, and the result is therefore of practical
signicance. To date, however, very few multiaxial experiments have been conducted on these materials in bulk form to

1286

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Fig. 3. Non-associative law for bcc Mo proposed by Racherla and Bassani (2007), with distinct yield (F) and ow (G) surfaces (left) and strain burst
phenomena (right gures) modeled with (bottom right) and without (top right) the use of strain gradient theory to regularlize the solution.

quantify the degree of non-associative behavior, although shear localization has been commonly reported in nanocrystalline
materials (Cheng et al., 2003). It is well known that the activation volume tends towards small values on the order of the unit
cell size for unit processes of dislocation nucleation at grain boundaries in nanocrystals (Van Swygenhoven et al., 1999; Derlet and Van Swygenhoven, 2002; Zhu et al., 2004b; Asaro and Suresh, 2005; Derlet et al., 2009), at free surfaces in nanowires
or micropillars (Zhu et al., 2008), or for nanoindentation (Schuh et al., 2005; Wang et al., submitted for publication). The
associated activation energies are small as well, typically on the order of 0.30.5 eV. Of course, the corresponding applied
stress levels are very high as the associated transition state pathways are not associated with cooperative atomic motion
at larger scales. Certainly, the inelastic behavior of nanostructured materials is a rich new topic in which many of the classical notions of generalized normality break down with dominance of isoscale non-associativity.

2.2.2. Heterogeneity-induced non-associativity in scale transitions


This type of non-associated plastic ow is distinct from that of isoscale non-associativity outlined in the last section.
Numerous methods have been developed for homogenization of heterogeneous, multiphase materials to estimate statistically representative properties or responses. The notion of a representative volume element (RVE) is a core concept in
homogenization. A RVE is considered as a volume that is large enough relative to correlation lengths of microstructure heterogeneities such that further increases of volume do not change the resulting homogenized response over this volume, i.e.,
the condition of statistical homogeneity is satised for uniform macroelds. It is essential to specify the properties or responses of interest when discussing a RVE, because different categories of responses depend in a different manner on various
moments of the spatial distribution of microstructure attributes. In particular, responses that pertain to stationary, nonevolving microstructures, such as elastic stiffness or thermal conductivity, are less sensitive to higher order moments or uctuations of heterogeneity of multiple phases, orientation, etc. within the volume element. Other responses, such as high cycle
fatigue and true fracture ductility, reect the inuence of higher moments of heterogeneity on plastic deformation. They depend on extreme value attributes of microstructure, including size distributions and spatial correlations of phases. Homogenization methods have been applied mainly to properties of stationary microstructures. For evolving microstructures, it has
often been tacitly assumed that the RVE size is comparable to that for stationary properties such as elastic stiffness. However,
since evolution is highly sensitive to local congurations of microstructure attributes, the RVE for such evolutionary responses is often quite large compared to that of stationary properties.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1287

Fig. 4. Deformed specimens (initial geometry at left top) at various deformation levels (top images) and plots comparing experimentally measured void
volume fraction evolution (Cocks yield function) at the center of the specimen with predicted evolution for superimposed pressures of (a) 237 MPa and (b)
475 MPa. Calculations in each case are based on void growth alone (lower curve)) and void growth with continuous strain-driven void nucleation (upper
curve).

A related complication is the decision of where to set the primal scale in homogenization, whether separation of scales
can be achieved, and its inuence on the resulting structure of the homogenized relations. For example, Fig. 5 shows an idealized volume element or window with two microstructure scales that differ by an order of magnitude in size and spacing. It
could represent, for example, reinforcing particles embedded within a polycrystalline matrix. It may be convenient to proceed by considering the matrix as homogeneous and local in terms of constitutive relations, with the consideration of RVE
size depending only on the size and spacing of the large particles. This size will in turn depend on elastic and plastic phase

(a)

(b)

Fig. 5. (a) Two-scale hierarchical material with second phase embedded in polycrystalline material, with an order of magnitude difference in characteristic
lengths, and (b) homogenization step from coarse scale heterogeneity to equivalent continuum.

1288

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

contrast (including anisotropy). In other words, certain characteristics of the underlying ne scale constitutive description
are assumed a priori; this often includes an associated ow rule for the matrix, in accordance with expectations of a polycrystalline matrix deforming by conservative dislocation glide, absent twinning or dislocation core spreading, and presumed
to consist of grains large enough to avoid complications associated with grain boundary interface-mediated behaviors that
contribute to isoscale non-associativity outlined in the previous section. In such cases, however, the manifestation of plastic
ow on the traction-displacement response at the boundaries of the overall volume element will be affected by the ne-scale
phase contrast in elastic and plastic responses, and the relative ratios of (i) ne-scale to coarse scale size and spacing of heterogeneities and (ii) coarse scale particle correlation lengths relative to window size. If a window size for homogenization is
selected on the basis of statistical homogeneity of elastic stiffness, and it is concluded that sufcient separation of scales exists between ne and coarse to approximate the matrix response as homogeneous for the scale transition in Fig. 5(b), it must
be recognized that this does not apply to all relevant responses, including plastic ow and work hardening rate. In reality,
initial yielding, whether under monotonic loading or under change of strain path, will be particularly sensitive to correlation
length of extreme values of ne-scale disorientation distribution as well as coarse scale particle size and spacing. Moreover,
such effects may persist well into the regime of plastic ow if plastic strain localization occurs. If the size of the volume element for homogenization is sufciently large compared to correlation lengths for these localization events, then the resulting
volume element may sufce as a RVE.
In fact, these implications carry over to polycrystals even without large disparities in size and spacing. Rice (1971) plainly
showed that the associated structure of crystal plasticity theory arises from a unique relation between the critical resolved
shear stress conjugate to a given shearing rate on a crystallographic slip system in a single crystal. On the other hand, the
hypothesis of generalized normality has often been tacitly assumed for polycrystals, whereby it is assumed that the same
driving forces for irreversible evolution of microstructure also govern the threshold for incipient evolution. Historical focus
on microscrystalline materials with symmetric crystal structures and often relatively weak elastic anisotropy (e.g., fcc polycrystals) has perhaps suppressed progress in understanding more general situations involving bcc, hcp structures, heterogeneous systems with strong phase contrast. Furthermore, the need to address stress eld gradient problems such as notches
with homogenization methods places limitations on the maximum practical size of a volume element for modeling, which
may very well be less than the requisite RVE size; this size can be so large as to render use of a local constitutive law inapplicable when considering realistic stress/strain gradients in components.
As discussed by McDowell (2008b), multiscale non-associativity arises from the lack of unique parameterization between
boundary tractions and local driving forces for plastic ow in highly activated subvolumes within the considered volume of
material, and is particularly pronounced in cases involving high degrees of heterogeneity of the local evolution elds, such as
in initial yielding or localization. Accordingly, the notion of statistical homogeneity of the distribution of activation volumes
for microstructure evolution is relevant. In effect one distinguishes between micropotentials and macropotentials for inelastic ow, i.e., equations should be ensemble averaged in the phase space of the internal variables and their rates.
We next illustrate several examples in which statistical homogeneity breaks down for certain responses of interest, namely
the early stages of plastic ow in heterogeneous materials and evolution of elds of distributed microcracks. This gives rise to
a nonunique relation of the thermodynamic forcesuxes, which in turn manifests non-associativity of the ow potential
with the yield function and attendant size effects. This is particularly acute for describing evolution of microstructure, as
the RVE size is potentially much larger in this case for than for properties that depend only on low order moments of microstructure distribution at each step, such as elastic stiffness (Dvorak, 1993; McDowell, 1997a,b, 1999, 2005; Lacy et al., 1999;
Ostoja-Starzewski, 1998, 2005, 2006). Consider a simple multiphase system with strong phase contrast, for example metal
matrix composites (Dvorak, 1993; Ostoja-Starzewski, 2005), in which non-normality of initial plastic ow is exhibited with
regard to the yield surface. Fig. 6 shows a random checkerboard pattern of hard and soft phases introduced and analyzed by
Ostoja-Starzewski (2005). Essentially, as the ratio of the simulation window size to the size of the individual phases decreases,
an increase in departure of associativity of plastic ow with the yield surface normal direction is observed. Moreover, the differences in upper and lower bounds based on uniform displacement and traction boundary conditions, respectively, increases
as the scale of the simulation window relative to phase size decreases. Clearly, there are important implications for self-consistent homogenization methods that a priori assume an associative ow structure independent of scale.
As a second example, Lacy et al. (1997, 1999) considered the effect of window size relative to the scale of crack size and
spacing in brittle microcracked solids to draw a distinction between the RVE for elastic stiffness and for evolution of the
microcrack distribution. Fig. 7 shows (upper left) a periodic element with a randomly placed (position and orientation)
set of initial cracks with the same length. In the upper right of Fig. 7, it is shown that the elastic stiffness of this periodic cell
is a monotonic function of the microcrack density component in the uniaxial loading direction, D22, computed from various
realizations of the randomly microcracked system using the dyadic expression shown for the crack density tensor. In other
words, this window scale serves as a RVE for elastic stiffness. However, it is clearly shown in Fig. 7 at lower left that from the
onset of crack extension (as determined using a common mixed mode propagation criterion), the elastic energy release rate
Y22 versus D22 relation bifurcates into various branches, showing that this same window size is too small to serve as a RVE for
crack density evolution. This owes to highly heterogeneous crack extension at certain crack tips which interact strongly.
However, it is been common in damage mechanics to assume similitude of crack extension in such systems. For this to apply,
it is clear that the RVE would need to be very large, in fact large enough to constitute a complete statistical sampling of the
extreme valued interactions (e.g., distribution of maximum intensication of crack driving forces) of the crack distribution.
Such extreme value statistics considerations are common in evolution of systems with strong phase contrasts. Clearly,

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1289

Fig. 6. Illustration of breakdown of normality ow rule in two phase material (by Ostoja-Starzewski, 2005) with phase contrast and different size of
statistical volume element relative to primal phase cell size (top right), resulting stressstrain responses for displacement (upper bound) and traction
(lower bound) boundary conditions (top left), with plots of the window scale plastic strain rate vector superimposed in the stress space on the computed
yield locus (bottom).

generalized normality would not apply to this scale of window size. This issue is not completely addressed by introducing
higher order representations of moments of the crack distribution, or gradient type evolution equations (Lacy et al., 1999), as
it is still necessary for each window scale to contain a statistically homogeneous distribution (i.e., complete distribution) of
the microcrack driving force eld. Such formulations can improve that accuracy of modeling individual realizations of the
statistically inhomogeneous elds, but a set of multiple realizations/simulations is still required to build up the statistics.
In other words, stochastic/probabilistic approaches are necessary.

1290

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Fig. 7. Illustration from Lacy et al. (1997, 1999) regarding how a periodic window containing a eld of microcracks in a brittle solids with random position
and orientation (top left) can serve as a RVE for elastic stiffness (top right) but not evolution in terms of energy release rate as a function of crack density
(lower left).

2.2.3. Size effects also emerge from multiscale heterogeneity


As a corollary of these results, another manifestation of statistical inhomogeneity of the evolutionary response is that of
size effects. We offer two illustrative cases that embody different physics. The rst is heterogeneous microplasticity in high
cycle fatigue (HCF) of polycrystalline metallics. In HCF, the probability of failure is established by the extreme value interactions of multiple grains or phases that produce the strongest local slip conditions to drive crack formation and early
growth (Przybyla et al., 2010; Przybyla and McDowell, 2009). The RVE size necessary to capture these extreme value cumulative probability distributions, mathematically characterized by the Gumbel or Weibull distributions, is much larger in
many cases than the size of laboratory specimens or the critically stressed volume at notches in components. Accordingly,
it is essential to either build up statistics from large populations of laboratory experiments or to conduct large numbers of
simulations over random periodic ensembles of grains in polycrystalline microstructures. As a result, it would not be sensible
to pursue a deterministic damage mechanics model for HCF rooted in the mechanics of statistical homogeneity (see also
McDowell, 1999b).
The second example is that of dislocation plasticity in conned volumes. Generalized normality using classical potentials
(including internal variable evolution) can be realized only in averaging over large numbers of these subvolume elements,
which may be precluded in cases of strong gradients such as crack tip process zones, small scale indentation, nanostructured
multilayer fccbcc systems, etc. A strong size effect is imparted when the detailed conguration or arrangement of defects
that govern initial strength and evolution play a key role. The resulting evolution of microstructure depends heavily on
boundary conditions and dislocation sourceobstacle interactions, giving rise to strong size effects of the kind commonly
addressed using nonlocal or gradient methods; however, such generalized continuum methods may not have the capability
to precisely capture the dislocation conguration/interaction energies, as well as line energies (Sedlacek et al., 2003; Ohno
and Okumura, 2007; Akasheh et al., 2007; McDowell, 2008b). Application of an appropriate discrete or continuous eld theory for dislocations that considers details of conguration of defects addresses the geometry- and boundary condition-specic nature of this kind of problem is necessary (cf. Cleveringa et al., 1997; Kreuzer and Pippan, 2005; Nicola et al., 2006,
2007). In particular, strong interactions between heterogeneities that extend over the scale of the window of observation
diminishes the probability of a single convex dissipation potential and necessitates either a multi-potential formulation
or a nonlocal formulation for that scale. In dislocation plasticity, one approach to address these size effects that has become
popular in the mechanics literature over the last 15 years is that of geometrically necessary dislocations (GNDs). The basic
idea is that net dislocation density (net Burgers vector) is necessary to accommodate incompatibility associated with strain

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1291

gradients if other means are not available (e.g., elastic deformation or cracking) for ductile crystals. A number of GND theories have been advanced, starting with the original proposition of Ashby (1970). For more detailed discussion of limits of
second gradient (geometrically necessary dislocation) theories and other nonlocal approaches, the reader is referred to
McDowell (2008b).
There are, of course, other sources for size effects that manifest directly from a change of dominant mechanism for inelastic deformation as a function of scale, for example transition from twinning or partial dislocation-mediated behavior to that
of full dislocations, the changing role of grain boundary deformation mechanisms with grain size, and so forth.
2.3. A multiscale hierarchical thermodynamics and kinetics of evolving microstructure
Thermodynamics and kinetics relations are at the core of evolution equations for microstructure. They are distinct of
course, and the foregoing discussion of multiscale non-associativity arising from statistical inhomogeneity points to the need
to introduce a hierarchical description of thermodynamics and kinetics (cf. McDowell, 1997, 1999b):
Case A: The free energy function and dissipation potential are statistically homogeneous and the volume element under
consideration constitutes a RVE for both.
Example: Associative elasto-plasticity of conventional grain size equiaxed fcc polycrystals.
Case B: The free energy function is statistically homogeneous, but evolution is statistically inhomogeneous and the dissipation potential is distinct from the threshold surface for evolution.
Example: Microcracked ceramic polycrystals, ber reinforced metal matrix composites
Case C: Both the free energy function and the dissipation potential are statistically inhomogeneous, giving rise to the need
for a fully conguration and scale dependent model.
Example: Nanoindentation, sub-micron thin lms and multilayers.
In effect, as shown in Fig. 8, the homogeneity of the activation volume for sites of microstructure evolution is critical to
distinguishing these three cases. The correlation length of sites for evolution relative to the window size is critical to evaluating the degree of statistical inhomogeneity. Of particular practical relevance to homogenization of evolving microstructure are extended HillMandel approaches that augment conventional degrees of freedom with higher order deformation
gradients and associated inelastic uxes (Kouznetsova et al., 2002, 2004; Chambon et al., 2004; Larsson and Diebels,
2007). Many nonlocal, micropolar, gradient and micromorphic theories fall under this classication; typically, only the lower
order moments of the evolving microstructure are considered (cf. Krner, 1963, 1966, 1983). As previously mentioned, the
specication of boundary conditions has been problematic in such theories, as it is difcult to distinguish the incorporation
of additional kinematic degrees of freedom from specic congurations or deformation modes (defect distributions) in statistically inhomogeneous cases such as nanoindentation, torsion of micron scale wires, and so forth. When averaging is involved to reduce the number of degrees of freedom available to represent the internal structure, the appropriate averaging
equivalency for the thermodynamic forceux pairs should include the scaling of intrinsic dissipation (Luscher and McDowell, 2009) in generalized HillMandel approaches.
Reiterating, length scale effects are manifested in the statistically inhomogeneous (SI) case, along with strong dependence
on initial and boundary conditions. If a generalized continuum is introduced to close the gap on modeling scale dependence in the statistically inhomogeneous case, then minimally additional, explicit boundary conditions are necessary to address each case, since by denition SI response depends on boundary conditions.

Incr
e

asin
g

Distribution of
Activation volume
inho

mo g
ene
ity

RVE SH

SH

SI

Fig. 8. Distribution of activation volume through window considered for homogenization of effective evolutionary response. Increasing inhomogeneity is a
consequence of increase of the correlation length between activated sites for microstructure evolution relative to the overall window scale considered,
giving rise to statistical inhomogeneity. The statistically homogeneous case corresponding to a RVE is shown at far left.

1292

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

2.4. Internal state variables and material hierarchy


In framing constitutive models it is understood that coarse scale internal state variables emerge from processes occurring
at multiple length and time scales. These length scales may be separated by many orders of magnitude, some associated with
dislocation bypass of obstacles on slip planes, and others associated with internal stresses at intergranular scales. In conventional theories of internal state variable type, the hierarchical nature of internal variables is often obscured to some extent
due to their phenomenological character. However, kinematic hardening, isotropic hardening and even the ow rule can
have multiple contributions as evidenced from experimental study of the underlying phenomena. A good example is that
of evolution of back stress under cyclic loading in conventional polycrystalline materials, with contributions from:
 Anelastic bowing of dislocations (several hundred nm)
 Differential ow within crystals due to coherent or semi-coherent precipitates or higher strength particles (10500 nm)
 Interactions of slip in heterogeneous patterning of dislocation substructure (e.g., cells) with scale on the order of the mean
free path for dislocation glide (1001000 nm)
 Interaction energy of dislocation arrays with net Burgers vector in conned volumes (e.g., pile-ups in thin lms or at grain
or phase boundaries (100 nm several lm in continuum theory)
 Differential yielding and ow among a set of grains with orientation distribution (50 lm macro).
The latter two categories depend strongly on the polycrystal size, orientation and disorientation distribution, which
strongly affect initial yield behavior under initial and reversed loading, and is dominantly a thermal in character. The former
categories are of more intragranular character, are subject to thermal activation, and tend to dominant the work hardening
response. There are noteworthy statistically homogeneous cases for which relatively simple models sufce to approximate
the effects of differentially activated subvolumes. For example, in metal plasticity the multicomponent nonlinear kinematic
P
_ where e_ n e_ nM pN,
_
hardening rule of ArmstrongFrederick form a
ai ; a_ i Ai =Xilim Xilim N  ai p,
captures the macroscopic back stress evolution remarkably well for polycrystals under nonproportional loading when used with a smooth macroscopic yield surface and associative ow for initially isotropic polycrystals (Chaboche, 1989; Ohno, 1990; McDowell, 1985,
1994). It has much in common with the Mroz (1967) formalism, as discussed by Ohno and Wang (1991). Fig. 9 illustrates
how at least two (i = 1,2) back stress components are necessary to address different length scale phenomena that affect
the overall shape of the cyclic stress-strain response.
Since the back stress components have tensorial character, multiaxial nonproportional cyclic straining experiments are
most useful in sorting out evolution equations. McDowell (1985) showed that the ArmstrongFrederick form of back stress
evolution most closely agreed with measured evolution of back stress as determined from direct numerical differentiation of
data to obtain the inelastic strain rate, followed by backward extrapolation by a xed distance to achieve the most smooth
continuous trajectory of the yield surface for 304 stainless steel at room temperature, as shown in Fig. 10.
Nonproportional cyclic hardening for 304L SS was accentuated by formation of epsilon-martensite at slip band intersections, and increased multislip was observed relative to proportional loading (McDowell et al., 1988). Moosbrugger and

( A , ) ( A , )
1

1
lim

Intergranular interactions, particle


size short range back stress

lim

intragranular dislocation Interactions


long range back stress

Stress (MPa)

1400
1200
1000
800
600
400
200
0
-200
-400
-600
-800
-1000
-1200

Preliminary fitting
Experimental data

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08


Engineering Strain

Viscosity function in flow rule


governs relaxation behavior
Fig. 9. Relation of short and long range components of back stress to most closely associated regimes of stressstrain behavior under loadingunloading
reloading and stress relaxation events.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1293

Fig. 10. Comparison of experimental data from nonproportional cyclic deformation of 304 stainless steel in terms of back stress evolution with trajectory
predicted by (left) Armstrong-Frederick form of nonlinear kinematic hardening and (right) Prager kinematic hardening rule with back stress rate in the
direction of the plastic strain rate (McDowell, 1985). Near tangency of vectors at left indicate the superiority of the AF form.

McDowell (1989, 1990) showed the additional cyclic hardening under nonproportional loading for 304 SS and Ni-base superalloy Waspaloy was almost entirely kinematic in nature, at that time countering many propositions to the contrary that additional hardening was of isotropic character.
3. Self-organization in jerky ow and grain subdivision
Signicant advances have recently been gained by Koslowski et al. (2004), LeSar and Rickman (2004) in studies of statistical behavior of dislocations and by Uchic et al. (2004), Uchic and Dimiduk (2005), and Dimiduk et al. (2006, 2007) in understanding of scaling laws applicable to certain intermediate scale regimes of deformation of single crystal pure Ni micro-pillar
specimens with sizes ranging from several microns to tens of microns which are then subjected to quasi-static compression.
Observations of strain bursts and dislocation avalanches during intermittent nucleation and migration of dislocations indicate a regime of scale free plastic ow of dislocations within crystals that adheres to the tenets of similitude or self-organized
criticality. In other words, dislocation structures at one scale appear the same in a statistical sense as those at another scale,
independent of initial conditions. With increase of specimen diameter, the onset of formation of sessile junctions in the specimen interior can give rise to geometric sensitivity and specimen size effects on yield behavior and attendant work hardening. Moreover, different scaling with specimen diameter was observed for polycrystalline Ni with ne grains relative to
specimen size (Dimiduk et al., 2007) even in these compression experiments due to a change of rate-controlling deformation
mechanisms.
Another set of similitude or scaling relations pertains to the development dislocation substructure formation and associated strengthening effects in single crystals and at sub-grain scales of coarse grain polycrystals. The self-organization of
dislocations into periodic low energy substructures (Kuhlmann-Wilsdorf, 1989) resulting in formation of sub-grain boundaries was rationalized by Leffers (1994) in terms of limited slip accommodation of incompatible deformation modes and was
later characterized by Hughes et al. (1997) in terms of relevant scaling laws as a function of cumulative plastic strain. Several
authors (Butler and McDowell, 1998; Horstemeyer and McDowell, 1998) have employed these scaling relations regarding
the formation of low misorientation sub-grain boundaries with plastic strain in face-centered cubic (fcc) metals by modifying the plastic spin of the lattice to promote more diffuse texture evolution. A large body of work has addressed subgrain
formation as a function of strain by treating nucleation and growth of sub-grain boundaries as partial disclination dipoles/quadropoles, blocky dislocation substructures or microbands (Kratochvil, 1990; Seefeldt et al., 2001; Kalidindi,
2001; Seefeldt, 2001; Pantleon, 2002; Seefeldt and Klimanek, 2002; Kratochvil and Sedlacek, 2003; Sedlacek and Kratochvil,
2005; Kratochvil et al., 2007) in a statistical continuum sense, as well as deformation of nanocrystals (Nazarov et al., 1996;
Zaichenko and Glezer, 1999).
As shown in Fig. 11, the incorporation of simple scaling relations for dislocation boundary spacing and disorientation
accompanying renement of structure during nite strain plasticity of single and polycrystals can be employed in simplied
extended frameworks of crystal plasticity for fcc metals to enhance work hardening and diffuse texture. The substructure
spin in Fig. 11 is modied to retard texture evolution in simulations that employ the simple Taylor intergranular constraint
(Butler and McDowell, 1998; Butler et al., 2002). McGinty and McDowell (1999) used this kind of formulation, along with a

1294

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

{111} Pole Figures -50% Compression

Grain
Subdivision

Multiscale Model Predictions

Forming Limit Diagram of Al

Major True Strain, 1

Effective Stress (MPa)

400

300

Compression
200

Torsion Following
50% Compression
Torsion

100

Experiment
Model

0.3

0.2

0.1

Experimental Data
Initially Isotropic
Initially Textured

0
0

0.2

0.4

0.6

0.8

-0.2

-0.1

0.1

0.2

0.3

0.4

Minor True Strain, 2

Effective Strain

Fig. 11. Incorporation of scaling law for grain subdivision from Hughes et al. (1997) into modied relation for substructure spin (upper left) that diffuses
texture evolution when employed for OFHC Cu with a Taylor intergranular constraint assumption (upper right) (Butler et al., 2002), and associated sequence
effects in nite torsion following compression (lower left) (McGinty and McDowell, 1999) and prediction of forming limit diagram (lower right) (McGinty
and McDowell, 2004).

multiscale framework, to predict deformation path sequence effects for OFHC Cu subjected to large strain compression, torsion and torsion after compression. The model was also employed to predict effects of texture on the forming limit diagram
for Al (McGinty and McDowell, 2004), also shown in Fig. 11.
4. Multiscale kinematics of crystals and polycrystals
e

Use of the conventional multiplicative decomposition of the deformation gradient f  f  f has been pervasive in the
past 25 years in polycrystal plasticity. However, it is often overlooked that the multiplicative composition is also subject
e
p
to consideration of the inuence of scale transition. If f  f  f holds at the scale of a single crystal, why should it hold
at the polycrystal scale given the intervention of grain boundaries in accommodating incompatibility? To explore this issue,
Clayton and McDowell (2003) considered a statistical volume element of crystalline metal consisting of an arbitrary number
of grains. The SVE is not a representative volume element (RVE) since we do not assume a priori the statistical homogeneity
of response functions of interest, such as effective stiffness, as discussed in a previous section. Mesoscopic (e.g., point within
a grain) current and reference coordinates within the SVE are labeled x and x0 , respectively, with the mesoscopic motion
denoted by x /x0 ; t, and time denoted by t. Macroscopic current and reference coordinates for the SVE are labeled X
and X0 , respectively, with the macroscopic motion denoted by X UX0 ; t. The macroscopic coordinates describe the collective behavior of an aggregate of grains with dimension O(1 mm), while the mesoscopic coordinates represent the local
behavior of single grains and sub-grains within the aggregate of scale O(10 lm). Fig. 12 shows how this ne scale decomposition is embedded within the coarse scale elastoplastic deformation of an SVE.
Macroscopic deformation tensors are dened as volume averages of local, mesoscopic tensors dened at locations within
the SVE. Dening the macroscopic deformation gradient F by averaging the local single crystal deformation gradient,
e
p
@/
f  @x
f  f over the reference volume of the SVE, V SVE
ref , located at X0 , results in
0

F

@U
1

@X0 V SVE
ref

Z
V SVE
ref

SVE

fdVref

1
V SVE
ref

Z
V SVE
ref

SVE

f  f dVref

1
V SVE
ref

Z
SSVE
ref

SVE

x  nref dSref

Next, the two-point deformation tensor Fe is introduced as the macroscopic elastic deformation gradient, not compatible in general with any prescribed deformation eld. We require that det Fe > 0, permitting unique left and right polar
decompositions, i.e., Fe Ve  Re Re  Ue : The elastic stretch tensors Ve and Ue are associated with application of macroscopic external stress. The conguration of the macroscopic SVE reached upon hypothetical instantaneous elastic unloading
from the current conguration via the inverse of the elastic deformation gradient, Fe1 , corresponds to null traction

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1295

Fig. 12. Distinction of coarse scale and ne scale multiplicative decompositions within a coarse scale SVE, with overall deformation gradient given by F
(Clayton and McDowell, 2003).

conditions on the boundary of the SVE (i.e., the traction ~tint 0 along SSVE
int ); for the present discussion, we merely require that
the boundaries of the SVE in the intermediate conguration be free of external traction. The elastic rotation Re is specied as
the solution of the integral-differential equation

T
1
R_ e  Re Xe  SVE
V ref

V SVE
ref

xe dVSVE
ref

1
V SVE
ref

V SVE
ref

SVE
r_ e  re dV ref ;

subject to initial conditions Re t 0 re t 0 1, where Xe is the macroscopic lattice spin tensor referred to the current
conguration. Innitesimal (elastic) lattice strain is tacitly assumed.
The local residual deformation gradient ~f is assumed to adhere to the decomposition ~f ~f e  ~f p , with ~f e and ~f p being the
local residual elastic and plastic deformation gradients, respectively. Hence, incompatible ~f e is not associated with the applied stress, but rather with accommodation of net dislocation density at the scale of grains. Residual elds ~f e and ~f p within
each crystal are entirely analogous to those addressed in works of Bilby et al. (1955, 1957), Bilby and Smith (1956), and others (Kondo, 1953; Krner, 1966; Fox, 1968; Eringen and Claus, 1970; Kossecka, 1974; Kossecka and de Wit, 1977), but they
associate in the present work with gradients of distributed plastic deformation at the mesoscale rather than discrete dislocations at the microscale.
~ is introduced for the intermediate conguraThe homogenized (macroscopic) residual deformation tensor for the SVE, F,
tion just described, i.e.,

~  Fe1  F  1
F
V SVE
ref

~fdVSVE 1
ref
V SVE
V SVE
ref
ref

Z
V SVE
ref

~f e  ~f p dVSVE
ref

~ may be expressed in terms of the coordinates of the external boundary of SVE only if ~f is compatible (i.e., intesuch that F
~p is specied as a volume average of the local residual plastic deforgrable). The homogenized residual plastic deformation F
mation over the SVE, i.e.,

~p  1
F
V SVE
ref

Z
V SVE
ref

~f p dVSVE :
ref

A multiplicative decomposition for the total deformation gradient follows as

~ Fe  F
~i  F
~p F
~L  F
~p :
F Fe  F

~L is the lattice deformation and meso-incompatibility tensor F


~i recties incompatibility arising due to heterogeneity
where F
of residual inelastic deformation within the grains of the polycrystal, and is given by

~i F
~F
~p1
F

1
V SVE
ref

Z
V SVE
ref

!
~f e  ~f p dVSVE
ref

1
V SVE
ref

Z
V SVE
ref

!1
~f p dVSVE
ref

~i 1 when residual lattice stretch and rotation vanish everywhere in the SVE (i.e., when ~f e 1). Otherwise
Note that F
~f e 1 in regions such as those near grain boundaries of polycrystals. In the most general case of heterogeneous residual elas~i is a function of both elastic and plastic residual deformations.
tic and plastic deformations in the SVE, F
~i is an outcome of an homogenization process, providing a SVE level
The meso-incompatibility deformation gradient F
~i
measure of the heterogeneity of the local, ne scale residual deformations ~f e and ~f p over the mesoscale. The tensor eld F
does not explicitly include higher-order macroscale deformation gradients in polycrystals, as in previous works (cf. Shizawa
and Zbib, 1999a,b; Naghdi and Srinivasa, 1993; Le and Stumpf, 1996; Menzel and Steinmann, 2000). The multiplicative
decomposition given by Eq. (5) implies the existence of a series of congurations for the SVE, as shown in Fig. 13. The interested reader is referred to Clayton and McDowell (2003) for a discussion of continuum balance laws in reference to this
~i as a missing link in the kinematdecomposition. Hartley (2003) and Gerken and Dawson (2008) have also commented on F
ical decomposition.

1296

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Fig. 13. Congurations associated with multiplicative decomposition of the deformation gradient at multiple scales; (left) schematic of congurations and
(right) more detailed rendering showing both local and global (homogenized) decompositions of the deformation gradient (Clayton and McDowell, 2003).

Clayton and McDowell (2003) pursued polycrystal plasticity nite element (FE) simulations to explore this decomposition
~i . Moreover, Clayton et al. (2005, 2006) discussed how that the generin terms of elastic energy storage and evolution of F
alization in Eq. (5) could be extended to incorporate micromorphic kinematics associated with microrotations associated
with disclinations and lattice distortion associated with point defects and general somigliana dislocations., i.e., the crystal
connection is generalized to

^ ::a F L a F L 1
C
cb
:a

a
:b;c

|{z}
crystal
connection
GNDs

C L 1 ad Q cbd
|{z}
microrotation
disclinations

dab Q c
|{z}

microexp ansion
isotropic point defects

C L 1 ad Q cbd  Q c C L bd
|{z}

microstrain
Somigliana dislocations

5. What about the grain boundaries?


One of the current frontiers in modeling plasticity of polycrystalline metals is more accurately accounting for the role of
grain boundaries (GBs). For some time experimental studies have shown that local crystal plasticity that treats grain boundaries only as surfaces of displacement continuity does not quantitatively account for grain boundaries. Moreover, gradient
and nonlocal theories of crystal plasticity that consider net dislocation density also encounter issues in treating high angle
boundaries, as do discrete dislocation approaches. For high angle boundaries, slip transfer reactions are highly sensitive to
GB structure, including absorption, desorption, reection and direct transmission. A number of recent works have considered
framing such reactions in simplied manner by considering slip transfer to be based on the jump of the Nye tensor (i.e., misorientation) across the boundaries (Evers et al., 2004; Gurtin, 2007; Acharya et al., 2006). In addition, micropolar or micromorphic plasticity models have potential applications with regard to modeling effects of grain boundaries on
accommodating incompatibility elds of adjacent crystals at grain boundaries (Forest et al., 2000; Forest and Sievert, 2003).
More detailed considerations of initial grain boundary structure and related slip transfer relations have been considered
in prior work:
 High angles GBs are typically composed of dissociated Shockley partial dislocations, equivalently represented by disconnections (Hirth et al., 2007), and varying fractions of disorder, depending on prior dislocation absorption.
 Relative activation of slip systems at the GB have been attributed in part to alignment of slip systems on either side of the
boundary (Lee et al., 1989, 1990a,b; Clark et al., 1992).
 Complexities of the structure of high angle GBs demand more detailed treatment than that offered by the jump of the Nye
tensor across the boundary since it carries only rst order strain gradient information. Transition states for unit processes
(reaction pathway and activation energy) of dislocation nucleation or other reactions depend on structure of the boundary, slip plane orientation to boundary, alignment with slip on other side of boundary, and the elastic energy of formation
of mist dislocations that remain as debris following slip penetration (Ma et al., 2006a,b; Capolungo et al., 2007).
In view of the complexity of slip transfer reactions and their many-body nature for common grain boundary networks,
mantle-and-core models (Kocks, 1970; Meyers and Ashworth, 1982) remain as plausible constructs to accept information
relevant to the statistical distribution of slip transfer reactions. The matter is still largely an open issue from a practical
perspective.

1297

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

5.1. Mesoscopic experiments


With the advent of EBSD methods and in situ optical measurement of slip traces, considerable effort has been recently
devoted to measurement of slip activity and lattice curvature in surface grains of (typically coarse grained) polycrystals. Such
information is useful to quantify concurrence of general trends of crystal plasticity models with experiments and to understand the effects of grain boundaries. Results are mixed in this regard, as might be expected. One drawback of these kinds of
experiments has been the focus on simple one dimensional applied stress states. Another class of experiments seeks to explore the issue of mesoscopic slip and rotation behavior over many grains under different nite strain states. To this end,
Schroeter and McDowell (2003) developed a methodology to measure in-plane material stretch and rotation elds at the
sub-grain scale for ensembles of 1020 grains in 62 lm average grain size OFHC Cu specimens subjected to nite compression and shear at a strain rate of 6  104 s1 at room temperature. Using a novel photolithographic technique, a series of 8
micron square grids were deposited onto the surfaces of specimens that were subsequently subjected to compression or torsion to uniaxial equivalent Mises strains up to unity. The resulting grid patterns for compression exhibited a tendency toward localization of stretch and rotation elds, with less deformation in the center of larger grains. For the case of shear,
there was a reduced tendency towards microstructural scale localization of the in-plane stretch and rotation elds compared
to compression. In general, compression to the same equivalent plastic strain level produced much greater heterogeneity of
deformation (including strong localized rotation elds near grain boundaries) than torsion, even when effects of enhanced
out-of-plane displacements were taken into account. In other words, the grain size distribution and grain boundary network
effects manifested much different details of stretch and rotation distributions under compression and shear, serving as a useful basis for benchmarking capabilities of mesoscopic gradient and nonlocal crystal plasticity theories, including those enhanced with mantle-and-core or other generalized continuum treatment of GB slip transfer relations. Fig. 14 shows back
scatter electron (BSE) compositional image renderings of the deformed surface grid patterns at two equivalent plastic strain
levels in compression and shear. The most striking outcome of this study was that the concentration of shear near grain
boundaries was much more pronounced at a uniaxial equivalent Mises strain of 50% in compression than in shear, indicating
that slip transfer relations between grains depend on the imposed long range deformation and stress state. The bulk deformation near the center of large grains was nearly uniform, whereas small grains exhibited substantial gradients of lattice
rotation and shear. With the notion that SSDs accumulate mainly in the grain interiors and GNDs manifest effects near grain
boundaries, such experiments can serve as validation of grain boundary-bulk slip interaction laws in crystal plasticity, particularly with regard to slip transmission through high angle boundaries. These results could not be replicated by simulation
using local crystal plasticity theory (Clayton et al., 2002).

2
100 m
m

2
100
m
100 m

Torsion, Eeff = 50%

Compression, Eeff = 50%

2
100 m
100
m
Torsion, Eeff = 100%

2
100 m
m
100
Compression, Eeff = 100%

Fig. 14. Back scatter electron compositional mode images of deformed surface grains under nite torsion (left column) and nite compression (right
column) for OHFC Cu with 8 lm lithographic grid patterns at 50% uniaxial equivalent Mises plastic strain (top row) and 100% uniaxial equivalent Mises
plastic strain (bottom row). Initial average grain size is 62 lm.

1298

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

This study of Schroeter and McDowell was complemented by the work of Bergugnat (2002) at Georgia Tech who employed the same methodology and considered differences in responses of Cu specimens with and without trace addition
of the grain boundary segregant Antimony (Sb). Effective plastic strain levels of 50% and 100% were examined in both simple
compression and shear as in the earlier study of Schroeter and McDowell (2003). OFHC copper and Cu0.2-wt% Sb specimens
were tested under the same conditions in order to assure valid comparison. The nite strain compression responses of Cu
specimens are compared to those of CuSb specimens in Fig. 15, and torsion responses for each are compared in Fig. 16.
Clearly, nonuniformity of the stretch and (particularly) rotation elds is decreased substantially for the CuSb specimens,
in accordance with expectations. The difference is quite marked in comparing pure Cu and CuSb in compression in Figs. 15
and 16. The rationale for Cu with trace Sb was that Sb facilitates grain boundary fracture to assist in relaxing intergranular
incompatibility, thereby mediating the role of grain boundaries in accommodating slip pileups. Rotation elds are expected
to vary signicantly near weakened grain boundaries with the addition of trace antimony, and stretch elds are expected to
be reduced in adjacent grains due to stress redistribution. Measurements of out-of-plane displacement distributions were
consistent with these notions. Lattice disorientation was additionally measured in these experiments using electron backscatter diffraction (EBSD) within grains to shed light on formation of low angle sub-grain boundaries. Within large domains
of cooperative deformation involving multiple grains, disorientation values found for CuSb were lower than the ones found
for Cu specimens at a given strain level. Values were the same for small domains for both Cu and CuSb specimens. Higher
values of the disorientation parameter in large domains seem to suggest that small domains accommodate the strain by
rotating rather than deforming, with large domains acting the opposite way. Moreover, the evolution of this disorientation
parameter proved to be proportional to E2=3
p , where Ep is the macroscopic uniaxial equivalent Mises effective plastic strain.
This behavior is consistent with one of the scaling laws enunciated by Hughes et al. (1997) concerning the average angle of
misorientation between low (or high) angle boundaries, and is supported by results of Butler et al. (2002) using synchrotron
radiation imaging of lattice orientations. Misorientation proles showed the appearance of high angle boundaries (1520)
at large strains.
In addition to the BSE compositional images, EBSD scans were performed on deformed specimens to measure the evolution of disorientation (ME) within grains as a function of applied strain. For all pixels of the grain, the disorientations with
respect to the average orientation were calculated and averaged; the resulting mean disorientation value for the ith grain
is dened as mi;E . Then, the mean disorientation parameter for the ensemble of grains in is dened as

P
ME

i mi;E

Ai;E

i Ai;E

where Ai;E is the area of the ith grain in the cross section at the given strain level. We use the term domain size here because
it is understood that grain size is not precise as a description of grains cut by an arbitrary plane. Domains with diameter

E p = 0.3

E p = 0.75

E p = 0.5

Cu

8 m
square
grids

100 m

100 m

100 m

Cu-Sb

100 m

100 m

100 m

Fig. 15. Back scatter electron compositional mode images of nite compression for (top row) OHFC Cu and (bottom row) Cu doped with Sb using 8 lm
lithographic square grid patterns at 30%, 50% and 75% uniaxial equivalent Mises plastic strain (three columns). Initial average grain size is 60 lm.

1299

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

E p =0.3

E p =0.75

E p =0.5

Cu

8 m
square
grids

100 m

100 m

100 m

100 m

100 m

Cu-Sb

100 m

Fig. 16. Back scatter electron compositional mode images of nite torsion for (top row) OHFC Cu and (bottom row) Cu doped with Sb using 8 lm
lithographic square grid patterns at 30%, 50% and 75% uniaxial equivalent Mises plastic strain (three columns). Initial average grain size is 60 lm.

between 10 lm and 30 lm are considered small; those with diameter between 30 lm and 100 lm are considered large.
Domains having a diameter less than 10 lm are assumed to arise from the non-indexed reduction process creation and domains larger than 100 lm are assumed to be the union of two smaller ones dened as the same grain by the grain reconstruction process; in both cases, these domains are not considered. The rst important result is that for OFHC Cu, large
domains/grains have larger values of M E than the smaller ones. This means that large domains exhibit more lattice disorientation than the smaller domains. Large domains have more neighbors, so it is preferable to accommodate neighboring
grains with gradients of strain within. Interestingly, the relation M E a E2=3
set forth by Hughes et al. (1997) ts the data
p
for both compression and torsion of pure Cu, as shown in Fig. 17. For pure Cu, at the same values of macroscopic strain
Ep, the mean disorientation M E is larger for torsion than for compression specimens. Hence, grain subdivision depends on
strain state and slip transfer through the grain boundary network, as well as compatibility requirements that depend on
grain size and perhaps shape.

Cu compression

Cu torsion

9
8

Large grains

12

Small grains

10

ME (degrees)

M E (degrees)

5
4
3

Large grains
Small grains

8
6
4

2
2

1
0

0
0

0.2

0.4

0.6

Plastic strain Ep

0.8

1.2

0.2

0.4

0.6

0.8

1.2

Plastic strain Ep

Fig. 17. Minimum misorientation at subgrain boundaries for large and small surface grains in OFHC Cu subjected to (left) nite compression and (right)
nite shear. All curves are t with different values of a in the scaling relation ME aE2=3
(Bergugnat, 2002).
p

1300

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Cu large

Cu-Sb large

M E (degrees)

5
4

Cu and CuSb small

3
2
1
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Ep
Large grains Cu

Small grains Cu-Sb

Large grains Cu-Sb

Small grains Cu

Fig. 18. Evolution of M E , as a function of the uniaxial equivalent strain Ep for torsion tests for Cu, and CuSb t of the relation M E aE2=3
p , a = 4.9 for small
domains, a = 12 for large ones (Bergugnat, 2002).

A comparison between the M E values for torsional Cu specimens with and without Sb is shown in Fig. 18. The values of M E
for small domains are the same for Cu and CuSb. This implies that small domains accommodate only limited deformation,
independent of the presence of Sb. But values of M E for large domains are different for Cu and CuSb; M E values for Cu specimens are higher than for the CuSb specimens for large domains. Antimony has played its segregant role, favoring accommodation of grain boundary incompatibility via intergranular fracture; hence, less strain and lower strain gradients are
experienced by large grains in the CuSb alloy. Some other perhaps subtle key points are as follows:
 From imaging of the deformed lithographic grids, the large, relatively unstrained areas observed by Schroeter (2001) for
pure Cu, corresponding to the cooperative deformation of groups of grains (Sakai and Muto, 1998), do not appear on the
CuSb grid images, replaced by domains at the scale of grain size.
 From 3-D out-of-plane measurements, the size of the unstrained, dominantly rotating regions changed from the scale of
several grains for the Cu specimens to the size of a grain for the CuSb ones; this was also supported by in-plane deformation as determined from distortion of the lithographic grids.
 By considering disorientation parameter evolution versus the macroscopic equivalent strain, the values found for CuSb
in large domains were lower than those found for Cu specimens for large domains. Indeed rigid body rotation is favored in
accommodating the imposed deformation rather than material deformation of large CuSb domains. Meanwhile, values
were the same for small domains for both Cu and CuSb specimens.
5.2. Atomistic simulations
Work within the past ve years by the author and his graduate students has sought to explore structureproperty relations for fcc Cu and Al bicrystal interfaces over a wide parametric range of misorientation angles for symmetric and asymmetric (inclination angle) tilt boundaries for [1 0 0] and [1 1 0] tilt axes. These simulations have employed the EAM
potentials of Mishin et al. (1999, 2001) for both Cu and Al, which are particularly accurate with regard to dislocation behavior. The gist of the results is summarized in Tschopp et al. (2008) as well as McDowell (2008b). Details can be found in additional references (Spearot et al., 2005, 2007a,b; Tschopp and McDowell, 2007a,b,c,d, 2008a,b,c; Tschopp et al., 2007, 2008)
and are summarized as follows:
 Nucleation of partial dislocations from the interface due to tensile stress applied normal to the boundary (NPT ensemble)
is generally of non-Schmid character (the same is true of homogeneous nucleation). The stress resolved normal to the slip
plane plays a dominant role in correlating nucleation over a signicant portion of the standard stereographic triangle.
The nucleation stress may be summarized in a relatively simple model for a wide range of symmetric tilt boundaries,
excluding h1 1 0i tilt axis boundaries that contain the E structural unit. The model for non-Schmid dislocation nucleation
at interfaces has been extended to include nite temperature kinetics applicable to longer time scales estimated using
MD.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1301

 Strong tensioncompression asymmetry is evident for dislocation nucleation strength and the character of the emission
of the trailing partial dislocation from the interface, with substantially constricted stacking fault widths in compression
assisting loop formation.
 Apart from delocalization of boundaries of moderate SFE Cu in certain ranges, the structure unit model of Sutton and
Vitek (1983a,b,c) offers an appropriate description of general high angle boundaries, including asymmetric boundaries
(boundary plane asymmetry relative to the two adjacent lattices).
 Asymmetric boundaries can give rise to a substantial number density of ledges and profuse nucleation, with the resulting
nucleation stress for the rst partial dislocation dependent on the distribution of ledges and free volume in the interface.
The high levels of computed stress required for nucleation of the rst partial dislocation and tensioncompression asymmetry of nucleation stress are conrmed by quasistatic simulations, as well as independent calculations based on the hyperelastic Interatomic Potential Finite Element Method (Liu et al., 2009). Such high levels of stress (510 GPa) pertain to
homogeneous nucleation or nucleation of dislocations at near equilibrium boundaries without prior deformation. Examples
of application include nanoindentation, behavior in the front of strong shock waves, nanocrystals with initial low dislocation
density, and nano-thickness multilayers.
Another important aspect, overlooked in most MD calculations to date, is the role of excess defects (dislocations and
vacancies) at grain boundaries in cold worked polycrystals. Ungar et al. (2005, 2007), Ungar (2006), and Zehetbauer et al.
(2005) have shown increasing dislocation density levels and vacancy concentrations with plastic deformation, and have inferred that grain and subgrain structures are the vacancy accumulation sites. It is important to understand the atomic structure in the grain boundary regions following extensive deformation, including the relative degree of disorder, in contrast to
the near equilibrium, highly ordered structures computed using energy minimization via molecular statics. The complexity
in representing heavily deformed nonequilibrium grain boundary structures with atomistics hinders exploration of effects of
prior cold work, since the time scales involved in establishing these structures greatly exceed MD capabilities. Hence, some
sort of biased Monte Carlo scheme is likely necessary. This is an important area for future work, along with the following:
 Effects of distribution of free volume on nucleation and slip transfer reactions, including exchange of free volume with
grain boundary triple point regions and congurational entropy.
 Effects of shear on atomic shufing, boundary migration, nucleation and slip transfer reactions.
 Effects of impurities, segregants and quantum-engineered GB structures.

6. Systems-based, integrated design of materials and products


Recent research trends have focused heavily on bottom-up multiscale modeling of materials, as outlined earlier in this
paper. However, many applications require tailoring of microstructure morphology to achieve performance requirements
at higher length and time scales. This is a top-down directed activity. As stated by the author in his 2001 article Materials
design: a useful research focus for inelastic behavior of structural metals, . . .a realistic approach to designing durable microstructures at the outset calls for application of plasticity theory at the microstructural level, with its many implications.
Materials design calls for a greater quantitative understanding of microstructureproperty relations, principally achieved
through simulations of representative (realistic) microstructures. The existing paradigm of materials selection is being replaced by the emergence of materials design (Olson, 1997, 2000; Seepersad et al., 2004; McDowell, 2007; McDowell and Olson, 2008), which seeks to concurrently design materials and products by considering material microstructure attributes as
design variables as part of the design system.
The insertion of new and/or improved materials in products typically requires costly and time-consuming certication
testing in addition to the material development phase. Physically based, microstructure-sensitive models are key elements
in supporting strategies for accelerated discovery, design and insertion of materials into products. The integration of multiscale modeling of materials with concurrent design of materials and products is being actively pursued by industry and government initiatives, as outlined in the recent report of the National Academy of Engineerings National Materials Advisory
Board study group on Integrated Computational Materials Engineering (Pollock and Allison, 2007). The heterogeneous microstructure is viewed as a multilevel hierarchy with each level as a subsystem conferring design degrees of freedom. The clear
aim of modeling is then to effectively address details of microstructure and associated phenomena at each appropriate level
to provide decision support, often without the necessity of transitioning among models across scales. In this sense, the aim is
much different than that of multiscale modeling discussed earlier in this overview.
As in Olsons (1997, 2000) inductive strategy for materials design, we seek top-down (goals/means), inductive methods
to design the process route and resulting hierarchical microstructure to tailor the material to meet a set of specied performance requirements. This is shown schematically in Fig. 19. Perhaps the most daunting challenge is the lack of inverse relations in moving from top-down in Fig. 19 for many complex irreversible processes associated with evolutionary material
responses. Most modeling and simulation tools for irreversible processes such as metal plasticity, phase transformation,
and damage evolution are of bottom-up character. A related challenge is the degree of uncertainty associated with forms
of models, model parameters, microstructure stochasticity, and microstructure hierarchy, in addition to the transfer of information from ne grain, high resolution models to coarse grain models with reduced degree of freedom at higher scales.

1302

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

System
New
New area
area

ds
Assembly
th o
e
sM
ysi
s
nal
od
Part
A
eth
ect
f
M
f
/E
gn
use
esi
D
Ca
t ed
Continuum
ien
Or
l
a
Go
Mesoscale
Material
Design
Design methods
methods
Selection
are
are available
available
Atomistic
Limitation in
Inverse
problem

Quantum

High Degree of Uncertainty

Fig. 19. Extension of systems-based, top-down materials design from part, sub-assemblies, assemblies and components to hierarchical levels of material
structure, treating levels of material structure and associated responses effectively as sub-systems.

With regard to materials design objectives, concurrent multiscale modeling schemes or homogenization concepts may
not be necessary in many cases, because often the goal is not to accurately predict mean properties but rather to understand
their sensitivity to microstructure and to capture essential dominant mechanisms and their transitions with applied loading
and environment. Given the uncertainty of model forms, model parameters, and schemes for linking various scales in Fig. 19,
the notion of design optimization using hierarchical or concurrent multiscale models is not particularly useful in many cases.
Instead, extension of concepts of systems-based robust design introduced by Taguchi (1993) to multilevel integrated design
of materials and products (Pollock and Allison, 2007; McDowell, 2007; McDowell and Olson, 2008; McDowell et al., 2009) is
more practical, with sensitivity of various responses to microstructure variation playing a central role. The challenge is to
tailor microstructures that deliver required performance requirements with consideration of such aspects as:








Phase morphologies of alloy systems for multifunctional applications


Evolution of microstructure (e.g. plasticity, phase transformation, diffusion, etc.)
Resistance to or preference for shear banding
Formability enhanced by transformation- or twinning-induced plasticity
Processing with targeted porosity control
Surface treatments, inclusions, and residual stresses
Fatigue and fracture resistance achieved with multiple phases, precipitate strengthening, crystallographic texture, and
grain boundary networks.

The latter categories are good examples of responses that depend on extremal characteristics of microstructure (largest
grains, inclusions, large inclusions near free surfaces, and so forth), rather than average microstructure attributes such as
mean grain size, phase volume fractions, averaging spacing, etc. Particularly critical issues in materials design revolve around
identication of the feasible design space, how to pursue top-down systems design given a lack of invertibility of most processstructure and structureproperty relations, how to couple often asynchronous processstructure and structureproperty experiments and models, and how to search for candidate solutions well outside the domain of prior experience or
material systems. A range of strategies have emerged to pursue materials design (cf. Zohdi, 2003; Ganapathysubramanian
and Zabaras, 2004; Li et al., 2005; Sankaran and Zabaras, 2007; Knezevic et al., 2008), some of which target problems for
which direct analytical or computational inverse problems can be effectively pursued, such as texture control of elastic properties (Adams et al., 2004; Lyon and Adams, 2004; Kalidindi et al., 2004, 2005).
The practical and potentially quite attractive goal of this kind of systems strategy is to replace a fraction of the number of
decisions in the materials development and certication cycle made on the basis of empiricism with those informed by modeling and simulation. If at present 95% of decisions are made on the basis of empirical development and experiments, can we
decrease this to 85% or 70%? Such advances are essential to reducing the cycle time for materials development to match the
computer-assisted design and prototyping cycle time for complex products such as airplanes, automobiles, and consumer
goods. Emerging experimental tools with various levels of resolution play an important role in model calibration and validation, as shown in Fig. 20. For a more detailed discussion of systems design of materials, the reader is referred to a recent
monograph by McDowell et al. (2009).

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1303

Fig. 20. Companion scale appropriate experimental methods for model calibration and validation in extending systems design to hierarchy of material
length scales.

7. Closure
There have been a number of exciting developments during the last 25 years with the advent of ubiquitous computational
modeling and simulation and with the use of computers in test control and data acquisition, advanced material characterization, and in situ experiments. During this period, the author has had the opportunity to contribute to a number of related
areas, some of which are not addressed in this article for lack of space these additional areas include:
 Multiaxial fatigue and early stages of fatigue crack formation and growth
 Microstructure-sensitive modeling of cyclic deformation of microstructures to support fatigue life estimation/variability,
including hierarchical models
 Thermomechanical deformation and fatigue
 Creep crack growth mechanics in homogeneous and bi-material interfaces
 Viscoplasticity of interconnect materials in circuit boards
 Plasticity in repeated rolling and fretting contacts
 Pseudoelastic shape memory alloy and polymer transformations, including texture and cyclic loading effects
 Plasticity of cellular materials and structures
 Computational algorithms for hierarchical crystal plasticity and internal state variable models
 Molecular modeling of polymer systems and linkage to internal state variable network models
 Dynamic plasticity of metals and multi-phase particle mixtures.

Acknowledgements
This work represents a synthesis and integration of ideas and developments supported in part by the US National Science
Foundation, the Army Research Ofce, the Ofce of Naval Research, the Air Force Ofce of Scientic Research and NASA. Support of DARPA in synthetic multifunctional materials, as well as the DARPA AIM (GEAC subcontract) and DARPA Engine Systems Prognosis (Pratt & Whitney subcontract) programs since 2000 was foundational to several subjects, including materials
design. The author is grateful for the support of the Center for Computational Materials Design, a joint Georgia Tech-Penn
State NSF I/UCRC with member companies such as GE, Pratt & Whitney, Timken, and Ford Motor Company taking a keen
interest in topics presented here. Sandia National Laboratories in Livermore California provided a rewarding environment
for interaction with the authors research program from 1990 to 2004. In addition, support of the Carter N. Paden, Jr. Distinguished Chair in Metals Processing has provided valuable support in new and emerging topics since 1998.
The author is most fortunate to have had the opportunity to work with enormously talented and enthusiastic graduate
students on various aspects of inelastic behavior of materials, many of whom have continued technical contributions in these
subjects: K.-I. Ho (1987), J. Moosbrugger (1988), C.-P. Leung (1988), K.B. Yoon (1990), J.-Y. Berard (1992), N.B. Adefris (1993),
R. Shah (1994), M.P. Miller (1994), E.B. Marin (1995), D.E. Hall (1995), M.F. Horstemeyer (1995), T.E. Lacy (1998), A.B. Tanner
(1998), A. Siu (1998), G.C. Butler (1999), V.P. Bennett (1999), S. Graham (1999), C.-Y. Fu (1999), T.J. Lim (1999), R. Morrissey
(2000), R.D. McGinty (2001), B. Schroeter (2001), J.D. Clayton (2003), C.-H. Goh (2003), J.-B. Bergugnat (2003), C.-C. Seepersad
(2004), H. Choi (2005), J. Shepherd (2006), M. Shenoy (2006), A.P. Gordon (2006), D.E. Spearot (2006), M.A. Tschopp (2007),
M. Zhang (2008), R. Prasannavenkatesan (2009), and current students D.J. Luscher, R.A. Austin, J. Mayeur, C. Przybyla,

1304

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

G. Tucker, N. Salajegheh, J. Lloyd, G. Castellucio, W. Musinski, J. Song, A. Patra, S. Tiwari, S. Britt, and B. Ellis. Visiting collaborators in related subjects have stimulated much work, including D.J. Bammann, J. Fan, F.P.E. Dunne and doctoral or post doctoral fellows J. Zhang, A. Wang, R. Kumar, K. Haldrup, F. Bridier and M. Owolabi. From his early days as a graduate student at
the University of Illinois at Urbana-Champaign working initially with F.A. Leckie and then with D.F. Socie, the author was
exposed to the value of a combined strategy of experiments and modeling. Of course, the Dean of Engineering at UIUC at
the time was none other than D.C. Drucker, which served to reinforce the importance of inelastic behavior of solids at all
levels of the organization! The past 26 years have been spent at Georgia Tech framing a complementary approach between
applied mechanics and materials science that has only been achieved with the generous spirit and high levels of expertise of
fellow faculty research collaborators through the years such as S.D. Antolovich, A. Saxena, R. Talreja, R.W. Neu, M. Zhou, K.
Jacob, J. Qu, T. Zhu, C. Ume, C. Lynch, N. Thadhani, T. Sanders, J. Cochran, K. Gall, M. Cherkaoui, J.K. Allen and F. Mistree. The
author is also grateful for productive collaborations with D.J. Benson of USCD and G.B. Olson of Northwestern University.

References
Abu Al-Rub, R.K., Voyiadjis, G.Z., Bammann, D.J., 2007. A thermodynamic based higher-order gradient theory for size dependent plasticity. International
Journal of Solids and Structures 44, 28882923.
Acharya, A., 2003. Driving forces and boundary conditions in continuum dislocation mechanics. Proceedings of the Royal Society of London A 459, 1343
1363.
Acharya, A., Bassani, J.L., 2000. Lattice incompatibility and a gradient theory of crystal plasticity. Journal of the Mechanics and Physics of Solids 48, 1565
1595.
Acharya, A., Roy, A., Sawant, A., 2006. Continuum theory and methods for coarse-grained, mesoscopic plasticity. Scripta Materialia 54, 705710.
Acharya, A., Beaudoin, A.J., Miller, R., 2008. New perspectives in plasticity theory: dislocation nucleation, waves and partial continuity of the plastic strain
rate. Mathematics and Mechanics of Solids 13 (34), 292315.
Adams, B.L., Lyon, M., Henrie, B., 2004. Microstructures by design: Linear problems in elastic-plastic design. International Journal of Plasticity 20 (89),
15771602.
Aifantis, E.C., 1987. The physics of plastic deformation. International Journal of Plasticity 3, 211247.
Aifantis, E.C., 1995. Pattern formation in plasticity. International Journal of Engineering Science 33 (15), 21612178.
Aifantis, E.C., 2003. Update on a class of gradient theories. Mechanics of Materials 35, 259.
Akasheh, F., Zbib, H.M., Hirth, J.P., Hoagland, R.G., Misra, A., 2007. Dislocation dynamics analysis of dislocation intersections in nanoscale metallic
multilayered composites. Journal of Applied Physics 101, 084314.
Amodeo, R.J., Ghoniem, N.M., 1988. A review of experimental-observations and theoretical-models of dislocation cells and subgrains. Res Mechanica 23 (2
3), 137160.
Arsenlis, A., Parks, D.M., 1999. Crystallographic aspects of geometrically-necessary and statistically-stored dislocation density. Acta Materialia 47 (5), 1597
1611.
Arsenlis, A., Parks, D.M., 2002. Modeling the evolution of crystallographic dislocation density in crystal plasticity. Journal of the Mechanics and Physics of
Solids 50, 19792009.
Asaro, R.J., Suresh, S., 2005. Mechanistic models for the activation volume and rate sensitivity in metals with nanocrystalline grains and nano-scale twins.
Acta Materialia 53, 33693382.
Ashby, M.F., 1970. The deformation of plastically non-homogeneous materials. Philosophical Magazine 21, 399424.
Bammann, D.J., Aifantis, E.C., 1982. On a proposal for a continuum with microstructure. Acta Mechanica 45, 91121.
Bassani, J.L., Ito, K., Vitek, V., 2001. Complex macroscopic plastic ow arising from non-planar dislocation core structures. Materials Science and Engineering
A 319321, 97101.
Bayley, C.J., Brekelmans, W.A.M., Geers, M.G.D., 2006. A comparison of dislocation-induced back stress formulations in strain gradient crystal plasticity.
International Journal of Solids and Structures 43, 72687286.
Belak, J., Turchi, P.E.A., Dorr, M.R., Richards, D.F., Fattebert, J.-L., Wickett, M.E., Streitz, F.H., 2009. Coupling of atomistic and meso-scale phase-eld modeling
of rapid solidication. In: 2009 APS March Meeting, Pittsburgh, PA, March 1620.
Bergugnat, J.-B., 2002. Strain and lattice rotation elds of deformed polycrystals. MS Thesis, GWW School of Mechanical Engineering, Georgia Institute of
Technology, 2002.
Bilby, B.A., Smith, E., 1956. Continuous distributions of dislocations III. Proceedings of the Royal Society of London A 236, 481505.
Bilby, B.A., Bullough, R., Smith, E., 1955. Continuous distributions of dislocations: a new application of the methods of non-Riemannian geometry.
Proceedings of the Royal Society of London A 231, 263273.
Bilby, B.A., Gardner, L.R.T., Stroh, A.N., 1957. Continuous distributions of dislocations and the theory of plasticity. In: Proc. 9th Int. Congr. Appl. Mech.
Bruxelles, vol. 8, Universit de Bruxelles, pp. 3544.
Butler, G.C., McDowell, D.L., 1998. Polycrystal constraint and grain subdivision. International Journal of Plasticity 14 (8), 703717.
Butler, G.C., Stock, S.R., McGinty, R.D., McDowell, D.L., 2002. X-ray microbeam Laue pattern studies of the spreading of orientation in OFHC copper at large
strains. ASME Journal of Engineering Materials and Technology 124, 4854.
Capolungo, L., Spearot, D.E., Cherkaoui, M., McDowell, D.L., Qu, J., Jacob, K., 2007. Dislocation nucleation from bicrystal interfaces and grain boundary ledges:
relationship to nanocrystalline deformation. Journal of the Mechanics and Physics of Solids doi:10.1016/j.jmps.2007.04.001.
Carstensen, C., Hackl, K., Mielke, A., 2002. Non-convex potentials and microstructures in nite-strain plasticity. Proceedings of the Royal Society of London
A458, 299317.
Chaboche, J.-L., 1989. Constitutive equations for cyclic plasticity and cyclic viscoplasticity. International Journal of Plasticity 5 (3), 247302.
Chambon, R., Caillerie, D., Tamagnini, C., 2004. A strain space gradient plasticity theory for nite strain. Computer Methods in Applied Mechanics and
Engineering 193 (2729), 27972826.
Chen, L.-Q., 2002. Phase-eld models for microstructure evolution. Annual Review of Materials Research 32, 113140.
Chen, Y., Lee, J., Xiong, L., 2009. A generalized continuum theory and its relation to micromorphic theory. ASCE Journal of Engineering Mechanics 135 (3),
149155.
Cheng, S., Spencer, J.A., Milligan, W.W., 2003. Strength and tension/compression asymmetry in nanostructured and ultrane-grain metals. Acta Materialia
51 (15), 45054518.
Chung, P.W., Namburu, R.R., 2003. On a formulation for a multiscale atomistic-continuum homogenization method. International Journal of Solids and
Structures 40, 25632588.
Clark, W.A.T., Wagoner, R.H., Shen, Z.Y., 1992. On the criteria for slip transmission across interfaces in polycrystals. Scripta Metallurgica et Materialia 26,
203206.
Clayton, J.D., McDowell, D.L., 2003. A multiscale multiplicative decomposition for elastoplasticity of polycrystals. International Journal of Plasticity 19,
14011444.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1305

Clayton, J., Schroeter, B., Graham, S., McDowell, D.L., 2002. Distributions of stretch and rotation in polycrystalline OFHC Cu. ASME Journal of Engineering
Materials and Technology 124 (3), 302313.
Clayton, J.D., Bammann, D.J., McDowell, D.L., 2005. A geometric framework for the kinematics of crystals with defects. Philosophical Magazine 85 (3335),
39834010.
Clayton, J.D., McDowell, D.L., Bammann, D.J., 2006. Modeling dislocations and disclinations with nite micropolar elastoplasticity. International Journal of
Plasticity 22 (2), 210256.
Cleri, F., DAgostino, G., Satta, A., Colombo, L., 2002. Microstructure evolution from the atomic scale up. Computational Materials Science 24, 2127.
Cleveringa, H., Van der Giessen, E., Needleman, A., 1997. Comparison of discrete dislocation and continuum plasticity predictions for a composite material.
Acta Materialia 45 (8), 31633179.
Cocks, A.C.F., 1996. Variational principles, numerical schemes and bounding theorems for deformation by NabarroHerring creep. Journal of the Mechanics
and Physics of Solids 44 (9), 14291452.
Curtarolo, S., Ceder, G., 2002. Dynamics of an inhomogeneously coarse grained multiscale system. Physical Review Letters 88 (25), 255504.
Dao, M., Asaro, R.J., 1993. Non-Schmid effects and localized plastic ow in intermetallic alloys. Materials Science and Engineering A 170, 143160.
Dao, M., Lee, B.J., Asaro, R.J., 1996. Non-Schmid effects on the behavior of polycrystals with applications to Ni3Al. Metallurgical Transactions A 27A, 8199.
Deo, C.S., Srolovitz, D.J., 2002. First passage time Markov chain analysis of rare events for kinetic Monte Carlo: double kink nucleation during dislocation
glide. Modeling and Simulation in Materials Science and Engineering 10, 581596.
Derlet, P.M., Van Swygenhoven, H., 2002. Length scale effects in the simulation of deformation properties of nanocrystalline metals. Scripta Materialia 47,
719724.
Derlet, P.M., Gumbsch, P., Hoagland, R., Li, J., McDowell, D.L., Van Swygenhoven, H., Wang, J., 2009. Atomistic simulations of dislocations in conned
volumes. MRS Bulletin 34 (3), 184189.
Dillon, O.W., 1977. Strain gradients in plasticity. Acta Mechanica 26, 189200.
Dillon, O.W., Kratochvl, J., 1970. A strain gradient theory of plasticity. International Journal of Solids and Structures 6, 15131533.
Dimiduk, D.M., Koslowski, M., LeSar, R., 2006. Preface to the viewpoint set on: statistical mechanics and coarse graining of dislocation behavior for
continuum plasticity. Scripta Materialia 54, 701704.
Dimiduk, D.M., Uchic, M.D., Rao, S.I., Woodward, C., Parthasarathy, T.A., 2007. Overview of experiments on microcrystal plasticity in fcc-derivative
materials: selected challenges for modeling and simulation of plasticity. Modelling and Simulation in Materials Science and Engineering 15, 135146.
Drucker, D.C., 1984. Material response and continuum relations: or from microscales to macroscales. ASME Journal of Engineering Materials and Technology
106 (4), 286289.
Dupuy, L.M., Tadmor, E.B., Miller, R.E., Phillips, R., 2005. Finite-temperature quasicontinuum: molecular dynamics without all the atoms. Physical Review
Letters 95 (6), 060202.
Dvorak, G.J., 1993. Micromechanics of inelastic composite materials: theory and experiment. Nadai Lecture. ASME Journal of Engineering Materials and
Technology 115 (4), 327338.
Eringen, A.C., 1972. Nonlocal polar elastic continua. International Journal of Engineering Science 10, 116.
Eringen, A.C., 1999. Microcontinuum eld theories. I. Foundations and Solids. Springer-Verlag, New York, Inc.
Eringen, A.C., Claus, Jr., W.D., 1970. A micromorphic approach to dislocation theory and its relation to several existing theories. In: Simmons, J.A., de Wit, R.,
Bullough, R. (Eds.), Fundamental Aspects of Dislocation Theory 2, pp. 10231040.
Evers, L.P., Brekelmans, W.A.M., Geers, M.G.D., 2004. Non-local crystal plasticity model with intrinsic SSC and GND effects. Journal of the Mechanics and
Physics of Solids 52, 23792401.
Fleck, N.A., Hutchinson, J.W., 1993. A phenomenological theory for strain gradient effects in plasticity. Journal of the Mechanics and Physics of Solids 41,
18251857.
Fleck, N.A., Muller, G.M., Ashby, M.F., Hutchinson, J.W., 1994. Strain gradient plasticity: theory and experiments. Acta Metallurgica Materialia 42, 475487.
Forest, S., Sievert, R., 2003. Elastoviscoplastic constitutive frameworks for generalized continua. Acta Mechanica 160, 71111.
Forest, S., Barbe, F., Cailletaud, G., 2000. Cosserat modelling of size effects in the mechanical behaviour of polycrystals and multi-phase materials.
International Journal of Solids and Structures 37, 71057126.
Fox, N., 1968. On the continuum theory of dislocations and plasticity. The Quarterly Journal of Mechanics and Applied Mathematics 21, 6775.
Friedman, L.H., Chrzan, D.C., 1998. Scaling theory of the HallPetch relation for multilayers. Physical Review Letters 81 (13), 27152718.
Ganapathysubramanian, S., Zabaras, N., 2004. Design across length scales: a reduced-order model of polycrystal plasticity for the control of microstructuresensitive material properties. Computer Methods in Applied Mechanics and Engineering 193 (4547), 50175034.
Gao, H., Huang, Y., Nix, W.D., Hutchinson, J.W., 1999. Mechanism-based strain gradient plasticityI. Theory. Journal of the Mechanics and Physics of Solids
47, 12391263.
Gerken, J.M., Dawson, P.R., 2008. A crystal plasticity model that incorporates stresses and strains due to slip gradients. Journal of the Mechanics and Physics
of Solids 56 (4), 16511672.
Ghosh, S., Lee, K., Raghavan, P., 2001. A multi-level computational model for multi-scale damage analysis in composite and porous materials. International
Journal of Solids and Structures 38 (14), 23352385.
Ghosh, S., Bai, J., Raghavan, P., 2007. Concurrent multi-level model for damage evolution in micro structurally debonding composites. Mechanics of
Materials 39 (3), 241266.
Groh, S., Marin, E.B., Horstemeyer, M.F., Zbib, H.M., 2009. Multiscale modeling of the plasticity in an aluminum single crystal. International Journal of
Plasticity 25 (8), 14561473.
Groma, I., 1997. Link between the microscopic and mesocopic length-scale description of the collective behavior of dislocations. Physical Review B 56 (10),
58075813.
Gumbsch, P., 1995. An atomistic study of brittle fracture: toward explicit failure criteria from atomistic modeling. Journal of Materials Research 10, 2897
2907.
Gurtin, M.E., 2002. A gradient theory of single-crystal viscoplasticity that accounts for geometrically necessary dislocations. Journal of the Mechanics and
Physics of Solids 50 (1), 532.
Gurtin, M.E., 2007. A theory of grain boundaries that accounts automatically for grain misorientation and grain-boundary orientation. Journal of the
Mechanics and Physics of Solids, doi:10.1016/j.jmps.2007.05.002.
Gurtin, M.E., Anand, L., 2007. A gradient theory for single-crystal plasticity. Modelling and Simulation in Materials Science and Engineering 15, S263270.
Hao, S., Moran, B., Liu, W.K., Olson, G.B., 2003. A hierarchical multi-physics model for design of high toughness steels. Journal of Computer-Aided Materials
Design 10, 99142.
Hao, S., Liu, W.K., Moran, B., Vernerey, F., Olson, G.B., 2004. Multi-scale constitutive model and computational framework for the design of ultra-high
strength, high toughness steels. Computer Methods in Applied Mechanics and Engineering 193, 18651908.
Hartley, C.S., 2003. A method for linking thermally activated dislocation mechanisms of yielding with continuum plasticity theory. Philosophical Magazine
83 (31/34), 37833808.
Hirth, J.P., Pond, R.C., Lothe, J., 2007. Spacing defects and disconnections in grain boundaries. Acta Materialia 55, 54285437.
Horstemeyer, M.F., McDowell, D.L., 1998. Modeling effects of dislocation substructure in polycrystal elastoviscoplasticity. Mechanics of Materials 27, 145
163.
Huang, Y., Qu, S., Hwang, K.C., Li, M., Gao, H., 2004. A conventional theory of mechanism-based strain gradient plasticity. International Journal of Plasticity
20, 753782.

1306

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Hughes, D.A., Liu, Q., Chrzan, D.C., Hansen, N., 1997. Scaling of microstructural parameters: misorientations of deformation induced boundaries. Acta
Materialia 45 (1), 105112.
Hutchinson, J.W., 2000. Plasticity at the micron scale. International Journal of Solids and Structures 37, 225238.
Kalidindi, S., 2001. Modeling anisotropic strain hardening and deformation textures in low stacking fault energy fcc metals. International Journal of
Plasticity 17, 837860.
Kalidindi, S.R., Houskamp, J.R., Lyon, M., Adams, B.L., 2004. Microstructure sensitive design of an orthotropic plate subjected to tensile load. International
Journal of Plasticity 20 (89), 15611575.
Kalidindi, S.R., Houskamp, J., Proust, G., Duvvuru, H., 2005. Microstructure sensitive design with rst order homogenization theories and nite element
codes. In: Materials Science Forum, v 495497, n PART 1, Textures of Materials, ICOTOM 14 Proceedings of the 14th International Conference on
Textures of Materials, pp. 2330.
Khachaturyan, A.G., Wang, Y.U., Jin, Cuitino, A.M., 2001. Nanoscale phase eld microelasticity theory of dislocations: model and 3D simulations. Acta
Materialia 49, 18471857.
Knezevic, M., Kalidindi, S.R., Mishra, R.K., 2008. Delineation of rst-order closures for plastic properties requiring explicit consideration of strain hardening
and crystallographic texture evolution. International Journal of Plasticity 24 (2), 327342.
Kocks, U.F., 1970. The relation between polycrystal deformation and single-crystal deformation. Metallurgical Transactions 1, 11211143.
Kondo, K., 1953. On the geometrical and physical foundations of the theory of yielding. In: Proc. 2nd Japan Congr. Appl. Mech., Japan National Committee for
Theoretical and Applied Mechanics, Science Council of Japan, Tokyo, pp. 4147.
Koslowski, M., LeSar, R., Thomson, R., 2004. Dislocation structures and the deformation of materials. Physical Review Letters 93, 265503.
Kossecka, E., 1974. Mathematical theory of defects. Part I. Statics. Archives of Mechanics 26, 9951010.
Kossecka, E., de Wit, R., 1977. Disclination kinematics. Archives of Mechanics 29, 633651.
Kouznetsova, V., Geers, M.G.D., Brekelmans, W.A.M., 2002. Multi-scale constitutive modelling of heterogeneous materials with a gradient-enhanced
computational homogenization scheme. International Journal for Numerical Methods in Engineering 54 (8), 12351260.
Kouznetsova, V.G., Geers, M.G.D., Brekelmans, W.A.M., 2004. Multi-scale second-order computational homogenization of multi-phase materials: a nested
nite element solution strategy. Computer Methods in Applied Mechanics and Engineering 193 (48/51), 55255550.
Kratochvil, J., 1990. Instability origin of dislocation cell misorientation. Scripta Metallurgica et Materialia 24, 12251228.
Kratochvil, J., Sedlacek, R., 2003. Pattern formation in the framework of the continuum theory of dislocations. Physical Review B 67 (9), 094105.
Kratochvil, J., Kruzik, M., Sedlacek, R., 2007. Statistically based continuum model of misoriented dislocation cell structure formation. Physical Review B 75,
064104.
Kreuzer, H.G.M., Pippan, R., 2005. Discrete dislocation simulation of nanoindentation: the inuence of obstacles and a limited number of dislocation sources.
Philosophical Magazine 85 (28), 33013319.
Krner, E., 1963. On the physical reality of torque stresses in continuum mechanics. International Journal of Engineering Science 1, 261278.
Krner, E., 1966. Theory of Crystal Defects. Academic Press, NY. 231.
Krner, E., 1983. Field theory of defects in Bravais crystals. In: Paidar, V., Lejcek, L. (Eds.), The Structure and Properties of Crystal Defects. Elsevier, pp. 357
369.
Kubin, L.P., Canova, G., 1992. The modeling of dislocation patterns. Scripta Metallurgica 27, 957962.
Kuhlmann-Wilsdorf, D., 1989. Theory of plastic deformation: properties of low energy dislocation structures. Materials Science and Engineering A 113, 1
41.
Kulkarni, Y., Knap, J., Ortiz, M., 2008. A variational approach to coarse-graining of equilibrium and non-equilibrium atomistic description at nite
temperature. Journal of the Mechanics and Physics of Solids 56 (4), 14171449.
Lacy, T., McDowell, D.L., Talreja, R., 1997. Effects of damage distribution on evolution. In: McDowell, D.L. (Ed.), Applications of Continuum Damage
Mechanics to Fatigue and Fracture, ASTM STP 1315, pp. 131149.
Lacy, T.E., McDowell, D.L., Talreja, R., 1999. Gradient concepts for evolution of damage. Mechanics of Materials 31, 831860.
Larsson, R., Diebels, S., 2007. A second-order homogenization procedure for multi-scale analysis based on micropolar kinematics. International Journal of
Numerical Methods in Engineering 69, 24852512.
Le, K.C., Stumpf, H., 1996. On the determination of the crystal reference in nonlinear continuum theory of dislocations. Proceedings of the Royal Society of
London A 452, 359371.
Lebensohn, R.A., Liu, Y., Ponte Castaeda, P., 2004. Macroscopic properties and eld uctuations in model power-law polycrystals: full-eld solutions versus
self-consistent estimates. Proceedings of the Royal Society of London A 460, 13811405.
Lee, T.C., Robertson, I.M., Birnbaum, H.K., 1989. Prediction of slip transfer mechanisms across grain boundaries. Scripta Metallurgica 23, 799803.
Lee, T.C., Robertson, I.M., Birnbaum, H.K., 1990a. An in situ transmission electron microscope deformation study of the slip transfer mechanisms in metals.
Metallurgical Transactions A 21A, 24372447.
Lee, T.C., Robertson, I.M., Birnbaum, H.K., 1990b. TEM in situ deformation study of the interaction of lattice dislocations with grain boundaries in metals.
Philosophical Magazine A 62, 131153.
Leffers, T., 1994. Lattice rotations during plastic deformation with grain subdivision. Materials Science Forum 157162, 18151820.
LeSar, R., Rickman, J.M., 2004. Incorporation of local structure in continuous theory of dislocations. Physical Review B 69, 172105.
Li, J., 2007. The mechanics and physics of defect nucleation. MRS Bulletin 32, 151159.
Li, J., Van Vliet, K.J., Zhu, T., Yip, S., Suresh, S., 2002. Atomistic mechanisms governing elastic limit and incipient plasticity in crystals. Nature 418 (6895),
307310.
Li, D.S., Bouhattate, J., Garmestani, H., 2005. Processing path model to describe texture evolution during mechanical processing. In: Materials Science Forum,
v 495497, n PART 2, Textures of Materials, ICOTOM 14 Proceedings of the 14th International Conference on Textures of Materials, pp. 977982.
Li, J., Kevrekidis, P.G., Gear, C.W., Kevrekidis, I.G., 2007. Deciding the nature of the coarse equation through microscopic simulations: the baby-bathwater
scheme. SIAM Review 49, 469487.
Liu, B., Qiu, X., Huang, Y., Hwang, K.C., Li, M., Liu, C., 2003. The size effect on void growth in ductile materials. Journal of the Mechanics and Physics of Solids
51, 11711187.
Liu, W.K., Karpov, E.G., Zhang, S., Park, H.S., 2004. An introduction to computational nanomechanics and materials. Computer Methods in Applied Mechanics
and Engineering 193 (1720), 15291578.
Liu, W.K., Park, H.S., Qian, D., Karpov, E.G., Kadowaki, H., Wagner, G.J., 2006. Bridging scale methods for nanomechanics and materials. Computer Methods in
Applied Mechanics and Engineering 195, 14071421.
Liu, X.-M., Liu, Z.-L., You, X.-C., Nie, J.-F., Zhuang, Z., 2009. Theoretical strength of face-centred-cubic single crystal copper based on a continuum model.
Chinese Physics Letters 26 (2), 026103.
Luscher, D.J., McDowell, D.L., 2009. An extended multiscale principle of virtual velocities approach for evolving microstructure. Mesomechanics 2009,
Elsevier, Procedia Engineering, University of Oxford, June 2426.
Luscher, D.J., McDowell, D.L., Bronkhorst, C.A., 2009. A second gradient theoretical framework for hierarchical multiscale modeling of materials.
International Journal of Plasticity, submitted for publication.
Lyon, M., Adams, B.L., 2004. Gradient-based non-linear microstructure design. Journal of the Mechanics and Physics of Solids 52 (11), 25692586.
Ma, A., Roters, F., Raabe, D., 2006a. A dislocation density based constitutive model for crystal plasticity FEM including geometrically necessary dislocations.
Acta Materialia 54, 21692179.
Ma, A., Roters, F., Raabe, D., 2006b. On the consideration of interactions between dislocations and grain boundaries in crystal plasticity nite element
modeling theory, experiments and simulations. Acta Materialia 54, 21812194.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1307

Marin, E.B., McDowell, D.L., 1996. Associative versus non-associative porous viscoplasticity based on internal state variable concepts. International Journal
of Plasticity 12 (5), 629669.
Marin, E.B., McDowell, D.L., 1998. Models for compressible elasto-plasticity based on internal state variables. International Journal of Damage Mechanics 7
(1), 4783.
McDowell, D.L., 1985. An experimental study of the structure of constitutive equations for nonproportional cyclic plasticity. ASME Journal of Engineering
Materials and Technology 107, 307315.
McDowell, D.L., 1994. Multiaxial effects in metallic materials. In: Symposium on Durability and Damage Tolerance, ASME AD vol. 43, ASME Winter Annual
Meeting, Chicago, IL, November 611, pp. 213267.
McDowell, D.L., 1997. Evolving structure and internal state variables. Nadai Award Lecture, ASME Materials Division, ASME IMECE, Dallas, TX, November 20.
McDowell, D.L., 1999a. Non-associative aspects of multiscale evolutionary phenomena. In: Picu, R.C., Krempl, E. (Eds.), Proceedings 4th International
Conference on Constitutive Laws for Engineering Materials, pp. 5457.
McDowell, D.L., 1999b. Damage mechanics in metal fatigue: a discriminating perspective. International Journal of Damage Mechanics 8, 377403.
McDowell, D.L., 2001. Materials design: a useful research focus for inelastic behavior of structural metals. In: Sih, G.C., Panin, V.E. (Eds.), Special Issue of the
Theoretical and Applied Fracture Mechanics, Prospects of Mesomechanics in the 21st Century: Current Thinking on Multiscale Mechanics Problems, vol.
37, pp. 245259.
McDowell, D.L., 2005. Internal state variable theory. In: Sidney Yip, Horstemeyer, M.F. (Eds.), Handbook of Materials Modeling, Part A: Methods, Springer,
The Netherlands, pp. 11511170.
McDowell, D.L., 2007. Simulation-assisted materials design for the concurrent design of materials and products. JOM 59 (9), 2125.
McDowell, D.L., 2008a. A perspective on trends in multiscale plasticity. Khan International Medal Lecture, Plasticity 2008, St. Thomas, Virgin Islands, January
8.
McDowell, D.L., 2008b. Viscoplasticity of heterogeneous metallic materials. Materials Science and Engineering R: Reports 62 (3), 67123.
McDowell, D.L., Olson, G.B., 2008. Concurrent design of hierarchical materials and structures. Scientic Modeling and Simulation (CMNS) 15 (1), 207.
McDowell, D.L., Stock, S.R., Stahl, D., Antolovich, S.D., 1988. Biaxial path dependence of deformation substructure of type 304 stainless steel. Metallurgical
Transactions 19A, 12771293.
McDowell, D.L., Marin, E., Bertoncelli, C., 1993. A combined kinematic-isotropic hardening theory for porous inelasticity of ductile metals. International
Journal of Damage Mechanics 2, 137161.
McDowell, D.L., Choi, H.-J., Panchal, J., Austin, R., Allen, J.K., Mistree, F., 2007. Plasticity-related microstructureproperty relations for materials design. Key
Engineering Materials 340341, 2130.
McDowell, D.L., Panchal, J.H., Choi, H.-J., Seepersad, C.C., Allen, J.K., Mistree, F., 2009. Integrated design of multiscale, multifunctional materials and products.
Elsevier, p. 392, ISBN-13:978-1-85617-662-0.
McGinty, R.D., McDowell, D.L., 1999. Multi-scale polycrystal plasticity. ASME Journal of Engineering Materials and Technology 121, 203209.
McGinty, R.D., McDowell, D.L., 2004. Application of multiscale crystal plasticity models to forming limit diagrams. ASME Journal of Engineering Materials
and Technology 126 (3), 285291.
Menzel, A., Steinmann, P., 2000. On the formulation of higher gradient plasticity for single and polycrystals. Journal of the Mechanics and Physics of Solids
48 (8), 17771796.
Mesarovic, S.Dj., 2005. Energy, congurational forces and characteristic lengths associated with the continuum description of geometrically necessary
dislocations. International Journal of Plasticity 21, 18551889.
Meyers, M.A., Ashworth, E., 1982. A model for the effect of grain size on the yield stress of metals. Philosophical Magazine A 46, 737759.
Miller, R., Tadmor, E.B., Phillips, R., Ortiz, M., 1998a. Quasicontinuum simulation of fracture at the atomic scale. Modelling and Simulation in Materials
Science and Engineering 6 (5), 607638.
Miller, R., Ortiz, M., Phillips, R., Shenoy, V., Tadmor, E.B., 1998b. Quasicontinuum models of fracture and plasticity. Engineering Fracture Mechanics 61 (34),
427444.
Mishin, Y., Farkas, D., Mehl, M.J., Papaconstantopoulos, D.A., 1999. Interatomic potentials for monoatomic metals from experimental data and ab initio
calculations. Physical Review B 59, 3393.
Mishin, Y., Mehl, M.J., Papaconstantopoulos, D.A., Voter, A.F., Kress, J.D., 2001. Structural stability and lattice defects in copper. Ab initio, tight-binding and
embedded atom calculations. Physical Review B 63, 224106.
Moldovan, D., Wolf, D., Phillpot, S.R., Haslam, A.J., 2002. Role of grain rotation during grain growth in a columnar microstructure by mesoscale simulation.
Acta Materialia 50, 33973414.
Moosbrugger, J.C., McDowell, D.L., 1989. On a class of kinematic hardening rules for nonproportional cyclic plasticity. ASME Journal of Engineering Materials
and Technology 111, 8798.
Moosbrugger, J.C., McDowell, D.L., 1990. A rate dependent bounding surface model with a generalized image point for cyclic nonproportional viscoplasticity.
Journal of the Mechanics and Physics of Solids 38 (5), 627656.
Mroz, Z., 1967. On the description of anisotropic work hardening. Journal of the Mechanics and Physics of Solids 15, 163175.
Mura, T., 1987. Micromechanics of Defects in Solids, second ed. Kluwer Academic Publishers, The Netherlands.
Muralidharan, K., Deymier, P.A., Simmons, J.H., 2003. A concurrent multiscale nite difference time domain/molecular dynamics method for bridging an
elastic continuum to an atomic system. Modelling and Simulation in Materials Science and Engineering 11 (4), 487501.
Naghdi, P.M., Srinivasa, A.R., 1993. A dynamical theory of structured solids. I. Basic developments. Philosophical Transactions of the Royal Society of London
A 345, 425458.
Nazarov, A.A., Romanov, A.E., Valiev, R.Z., 1996. Random disclination ensembles in ultrane-grained materials produced by severe plastic deformation.
Scripta Materialia 34 (5), 729734.
Needleman, A., Rice, J.R., 1980. Plastic creep ow effects in the diffusive cavitation of grain boundaries. Acta Metallurgica 28 (10), 13151332.
Nicola, L., Xiang, Y., Vlassak, J.J., Van der Giessen, E., Needleman, A., 2006. Plastic deformation of freestanding thin lms: experiments and modeling. Journal
of the Mechanics and Physics of Solids 54, 20892110.
Nicola, L., Bower, A.F., Kim, K.-S., Needleman, A., Van der Giessen, E., 2007. Surface versus bulk nucleation of dislocations during contact. Journal of the
Mechanics and Physics of Solids 55, 11201144.
Nix, W.D., Gao, H., 1998. Indentation size effects in crystalline materials: a law for strain gradient plasticity. Journal of the Mechanics and Physics of Solids
46, 411425.
Nuggehally, M.A., Shephard, M.S., Picu, C.R., Fish, J., 2007. Adaptive model selection procedure for concurrent multiscale problems. Journal of Multiscale
Computational Engineering 5 (5), 369386.
Nye, J.F., 1953. Some geometrical relations in dislocated crystals. Acta Metallurgica 1, 153162.
Ohashi, T., Kawamukai, M., Zbib, H., 2007. A multiscale approach for modeling scale-dependent yield stress in polycrystalline metals. International Journal
of Plasticity 23 (5), 897914.
Ohno, N., 1990. Recent topics in constitutive modeling of cyclic plasticity and viscoplasticity. Applied Mechanics Reviews 43 (11), 283295.
Ohno, N., Okumura, D., 2007. Higher-order stress and grain size effects due to self-energy of geometrically necessary dislocations. Journal of the Mechanics
and Physics of Solids 55, 18791898.
Ohno, N., Wang, J.-D., 1991. Transformation of a nonlinear kinematic hardening rule to a multisurface form under isothermal and nonisothermal conditions.
International Journal of Plasticity 7, 879891.
Olson, G.B., 1997. Computational design of hierarchically structured materials. Science 277, 12371242.
Olson, G.B., 2000. Designing a new material world. Science 288, 993998.

1308

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

Ortiz, M., Repetto, E.A., 1999. Nonconvex energy minimization and dislocation structures in ductile single crystals. Journal of the Mechanics and Physics of
Solids 47, 397462.
Ortiz, M., Repetto, E.A., Stainier, L., 2000. A theory of subgrain dislocation structures. Journal of the Mechanics and Physics of Solids 48, 20772114.
Ostoja-Starzewski, M., 1998. Random eld models of heterogeneous materials. International Journal of Solids and Structures 35, 24292455.
Ostoja-Starzewski, M., 2005. Scale effects in plasticity of random media: status and challenges. International Journal of Plasticity 21, 11191160.
Ostoja-Starzewski, M., 2006. Material spatial randomness: from statistical to representative volume element. Probabilistic Engineering Mechanics 21, 112
132.
Paidar, V., Pope, D.P., Vitek, V., 1984. A theory of the anomalous yield behaviour in L12 ordered alloys. Acta Metallurgica 32, 435448.
Pantleon, W., 2002. Formation of disorientations in dislocation structures during plastic deformation. Solid State Phenomena 87, 7392.
Park, H.S., Karpov, E.G., Klein, P.A., Liu, W.K., 2005. The bridging scale for two-dimensional atomistic/continuum coupling. Philosophical Magazine 85 (1),
79113.
Pollock, T.M., Allison, J., 2007. Committee on Integrated Computational Materials Engineering: developing a roadmap for a grand challenge in materials.
National Materials Advisory Board, National Academy of Engineering <http://www7.nationalacademies.org/nmab/CICME_home_page.html>.
Przybyla, C.P., McDowell, D.L., 2009. Microstructure-sensitive extreme value probabilities for high cycle fatigue of Ni-Base superalloy IN100. International
Journal of Plasticity, doi:10.1016/j.ijplas.2009.08.001.
Przybyla, C., Prasannavenkatesan, R., Salajegheh, N., McDowell, D.L., 2010. Microstructure-sensitive modeling of high cycle fatigue. International Journal of
Fatigue 32 (3), 512525.
Qin, Q., Bassani, J.L., 1992. Non-Schmid yield behavior in single crystals. Journal of the Mechanics and Physics of Solids 40, 813833.
Qu, S., Shastry, V., Curtin, W.A., Miller, R.E., 2005. A nite-temperature dynamic coupled atomistic/discrete dislocation method. Modelling and Simulation in
Materials Science and Engineering 13 (7), 11011118.
Racherla, V., Bassani, J.L., 2007. Strain burst phenomena in the necking of a sheet that deforms by non-associated plastic ow. Modelling and Simulation in
Materials Science and Engineering 15, S297S311.
Rai-Tabar, H., Hua, L., Cross, M., 1998. A multi-scale atomistic-continuum modeling of crack propagation in a two-dimensional macroscopic plate. Journal
of Physics Condensed Matter 10 (11), 23752387.
Ramasubramaniam, A., Carter, E.A., 2007. Coupled quantum-atomistic and quantum-continuum mechanics methods in materials research. MRS Bulletin 32
(11), 913918.
Rice, J.R., 1971. Inelastic constitutive relations for solids: an internal variable theory and its application to metal plasticity. Journal of the Mechanics and
Physics of Solids 19, 433455.
Rickman, J.M., LeSar, R., 2006. Issues in the coarse-graining of dislocation energetics and dynamics. Scripta Materialia 54, 735739.
Roy, A., Acharya, A., 2006. Size effects and idealized dislocation microstructure at small scales: predictions of a phenomenological model of mesoscopic eld
dislocation mechanics. Journal of the Mechanics and Physics of Solids 54, 17111743.
Roy, A., Puri, S., Acharya, A., 2007. Phenomenological mesoscopic dislocation mechanics, lower-order gradient plasticity and transport of mean excess
dislocation density. Modelling and Simulation in Materials Science and Engineering 15, S167S180.
Rudd, R.E., Broughton, J.Q., 1998. Coarse-grained molecular dynamics and the atomic limit of nite elements. Physical Review B 58 (10), R5893R5896.
Rudd, R.E., Broughton, J.Q., 2000. Concurrent coupling of length scales in solid state systems. Physica Status Solidi B Basic Research 217 (1), 251291.
Rudd, R.E., Broughton, J.Q., 2005. Coarse-grained molecular dynamics: nonlinear nite elements and nite temperature. Physical Review B 72 (14), 144104.
Sakai, M., Muto, H., 1998. A novel deformation process in an aggregate: a candidate for superplastic deformation. Scripta Materialia 38 (6), 909915.
Sankaran, S., Zabaras, N., 2007. Computing property variability of polycrystals induced by grain size and orientation uncertainties. Acta Materialia 55 (7),
22792290.
Schroeter, B.M., 2001. Techniques for micro- and meso-scale measurements of rotation and stretch processes in metallic polycrystals. MS Thesis, GWW
School of Mechanical Engineering, Georgia Institute of Technology.
Schroeter, B.M., McDowell, D.L., 2003. Measurement of deformation elds in polycrystalline OFHC copper. International Journal of Plasticity 19 (9), 1355
1376.
Schuh, C.A., Mason, J.K., Lund, A.C., 2005. Quantitative insight into dislocation nucleation from high-temperature nanoindentation experiments. Nature
Materials 4 (8), 617621.
Sedlacek, R., Kratochvil, J., 2005. Variational approach to subgrain formation. Zeitschrift Fur Metallkunde 96, 602607.
Sedlacek, R., Kratochvil, J., Werner, E., 2003. The importance of being curved: bowing dislocations in a continuum description. Philosophical Magazine 83
(313), 37353752.
Seefeldt, M., 2001. Disclinations in large-strain plastic deformation and work-hardening. Review on Advanced Materials Science 2, 4479.
Seefeldt, M., Klimanek, P., 2002. Modeling of plastic deformation by means of dislocation-disclination dynamics. Solid State Phenomena 87, 93112.
Seefeldt, M., Delannay, L., Peeters, B., Aernoudt, E., Van Houtte, P., 2001a. Modeling the initial stage of grain subdivision with the help of a coupled
substructure and texture evolution algorithm. Acta Materialia 49, 21292143.
Seefeldt, M., Delannay, L., Peeters, B., Kalidindi, S.R., Van Houtte, P., 2001b. A disclination-based model for grain subdivision. Materials Science and
Engineering A 319321, 192196.
Seepersad, C.C., Fernandez, M.G., Panchal, J.H., Choi, H.-J., Allen, J.K., McDowell, D.L., Mistree, F., 2004. Foundations for a systems-based approach for
materials design. In: 10th AIAA/ISSMO Multidisciplinary Analysis and Optimization Conference. AIAA MAO, Albany, NY, pp. AIAA-20044300.
Shen, C., Wang, Y., 2003. Modeling dislocation network and dislocationprecipitate interaction at mesoscopic scale using phase eld method. International
Journal of Multiscale Computational Engineering 1 (1), 91104.
Shenoy, V.B., Miller, R., Tadmor, E.B., Phillips, R., Ortiz, M., 1998. Quasicontinuum models of interfacial structure and deformation. Physical Review Letters 80
(4), 742745.
Shenoy, V.B., Miller, R., Tadmor, E.B., Rodney, D., Phillips, R., Ortiz, M., 1999. An adaptive nite element approach to atomic-scale mechanics the
quasicontinuum method. Journal of the Mechanics and Physics of Solids 47 (3), 611642.
Shenoy, M., Tjiptowidjojo, Y., McDowell, D.L., 2008. Microstructure-sensitive modeling of polycrystalline IN 100. International Journal of Plasticity 24 (10),
16941730.
Shiari, B., Miller, R.E., Curtin, W.A., 2005. Coupled atomistic/discrete dislocation simulations of nanoindentation at nite temperature. ASME Journal of
Engineering Materials and Technology 127 (4), 358368.
Shilkrot, L.E., Curtin, W.A., Miller, R.E., 2002. A coupled atomistic/continuum model of defects in solids. Journal of the Mechanics and Physics of Solids 50,
20852106.
Shilkrot, L.E., Miller, R.E., Curtin, W.A., 2004. Multiscale plasticity modeling: coupled atomistics and discrete dislocation mechanics. Journal of the Mechanics
and Physics of Solids 52, 755787.
Shizawa, K., Zbib, H.M., 1999a. A thermodynamical theory of gradient elastoplasticity with dislocation density tensor. I: fundamentals. International Journal
of Plasticity 15 (9), 899938.
Shizawa, K., Zbib, H.M., 1999b. ASME Journal of Engineering Materials and Technology 121, 247253.
Spearot, D., Jacob, K.I., McDowell, D.L., 2005. Nucleation of dislocations from [0 0 1] bicrystal interfaces in aluminum. Acta Materialia 53 (13), 35793589.
Spearot, D.E., Jacob, K.I., McDowell, D.L., 2007a. Dislocation nucleation from bicrystal interfaces with dissociated structure. International Journal of Plasticity
23, 143160.
Spearot, D.E., Tschopp, M.A., Jacob, K.I., McDowell, D.L., 2007b. Tensile strength of h1 0 0i and h1 1 0i tilt bicrystal copper interfaces. Acta Materialia 55 (2),
705714.
Suquet, P.M., 1987. Homogenization Techniques for Composite Media. Lecture Notes in Physics, vol. 272. Springer, Berlin.

D.L. McDowell / International Journal of Plasticity 26 (2010) 12801309

1309

Sutton, A.P., Vitek, V., 1983a. On the structure of tilt grain boundaries in cubic metals I. Symmetrical tilt boundaries. Philosophical Transactions of the Royal
Society of London A 309, 136.
Sutton, A.P., Vitek, V., 1983b. On the structure of tilt grain boundaries in cubic metals II. Asymmetrical tilt boundaries. Philosophical Transactions of the
Royal Society of London A 309, 3754.
Sutton, A.P., Vitek, V., 1983c. On the structure of tilt grain boundaries in cubic metals. III. Generalizations of the structural study and implications for the
properties of grain boundaries. Philosophical Transactions of the Royal Society of London A 309, 5568.
Tadmor, E.B., Ortiz, M., Phillips, R., 1996a. Quasicontinuum analysis of defects in solids. Philosophical Magazine A 73 (6), 15291563.
Tadmor, E.B., Phillips, R., Ortiz, M., 1996b. Mixed atomistic and continuum models of deformation in solids. Langmuir 12 (19), 45294534.
Taguchi, G., 1993. Taguchi on Robust Technology Development: Bringing Quality Engineering Upstream. ASME Press, New York.
Tschopp, M.A., McDowell, D.L., 2007a. Structures and energies of R3 asymmetric tilt grain boundaries. Philosophical Magazine 87 (22), 31473173.
Tschopp, M.A., McDowell, D.L., 2007b. Tension-compression asymmetry in homogeneous dislocation nucleation in single crystal copper. Applied Physics
Letters 90, 121911121916.
Tschopp, M.A., McDowell, D.L., 2007c. Asymmetric tilt grain boundary structure and energy in copper and aluminium. Philosophical Magazine 87, 3871
3892.
Tschopp, M.A., McDowell, D.L., 2007d. Structural unit and faceting description of summation R3 asymmetric tilt grain boundaries. Journal of Materials
Science 42, 78067811.
Tschopp, M.A., McDowell, D.L., 2008a. Dislocation nucleation in R3 asymmetric tilt grain boundaries. International Journal of Plasticity 24 (2), 191217.
Tschopp, M.A., McDowell, D.L., 2008b. Grain boundary dislocation sources in nanocrystalline copper. Scripta Materialia 58 (4), 299302.
Tschopp, M.A., McDowell, D.L., 2008c. Inuence of single crystal orientation on homogeneous dislocation nucleation under uniaxial loading. Journal of the
Mechanics and Physics of Solids 56 (5), 18061830.
Tschopp, M.A., Spearot, D.E., McDowell, D.L., 2007. Atomistic simulations of homogeneous dislocation nucleation in single crystal copper. Modeling and
Simulation in Materials Science and Engineering 15 (7), 693709.
Tschopp, M.A., Spearot, D.E., McDowell, D.L., 2008. Inuence of grain boundary structure on dislocation nucleation in fcc metals. In: Dislocations in Solids, A
Tribute to F.R.N. Nabarro (Ed.), J.P. Hirth, vol. 14, Elsevier Publ., pp. 43139.
Uchic, M., Dimiduk, D., 2005. A methodology to investigate size scale effects in crystalline plasticity using uniaxial compression testing. Materials Science
and Engineering A 400401, 268278.
Uchic, M., Dimiduk, D., Florando, J., Nix, W., 2004. Sample dimensions inuence strength and crystal plasticity. Science 305, 986989.
Ungar, T., 2006. Subgrain size-distributions, dislocation structures, stacking- and twin faults and vacancy concentrations in SPD materials determined by Xray line prole analysis. Materials Science Forum 503504, 133140.
Ungar, T., Schaer, E., Hanak, P., Bernstorff, S., Zehetbauer, M., 2005. Vacancy concentrations determined from the diffuse background scattering of X-rays in
plastically deformed copper. Zeitschrift fur Metallkunde 96, 578583.
Ungar, T., Schaer, E., Hanak, P., Bernstorff, S., Zehetbauer, M., 2007. Vacancy production during plastic deformation in copper determined by in situ X-ray
diffraction. Materials Science and Engineering A 462, 398401.
Van Swygenhoven, H., Spaczer, M., Caro, A., 1999. Microscopic description of plasticity in computer generated metallic nanophase samples: A comparison
between Cu and Ni. Acta Materialia 47, 31173126.
Vernerey, F., Liu, W.K., Moran, B., 2007. Multi-scale micromorphic theory for hierarchical materials. Journal of the Mechanics and Physics of Solids 55 (12),
26032651.
Viatkina, E.M., Brekelmans, W.A.M., Geers, M.G.D., 2007. Modelling of the internal stress in dislocation cell structures. European Journal of Mechanics A/
Solids 26, 982998.
Walgraef, D., Aifantis, E.C., 1985. On the formation and stability of dislocation patterns, IIII. International Journal of Engineering Science 12, 13511372.
Wang, A.-J., Kumar, R.S., Shenoy, M.M., McDowell, D.L., 2006. Microstructure-based multiscale constitutive modeling of cc0 Ni-base superalloys.
International Journal of Multiscale Computational Engineering 4 (56), 663692.
Wang, W., Zhong, Y., Lu, K., Lu, L., McDowell, D.L., Zhu, T., 2009. Strength uctuations and size Effects in nanoindentation. Physical Review Letters, submitted
for publication.
Warner, D.H., Sansoz, F., Molinari, J.F., 2006. Atomistic based continuum investigation of plastic deformation in nanocrystalline copper. International Journal
of Plasticity 22 (4), 754774.
Weinan, E., Huang, Z., 2001. Matching conditions in atomistic-continuum modeling of materials. Physical Review Letters 8713 (13), 135501.
Zaichenko, S.G., Glezer, A.M., 1999. Disclination mechanism of plastic deformation of nanocrystalline materials. Interface Science 7, 5767.
Zaiser, M., 2001. Statistical modeling of dislocation systems. Material Science and Engineering A 309310, 304315.
Zbib, H.M., Aifantis, E.C., 1992. On the gradient-dependent theory of plasticity and shear banding. Acta Mechanica 92, 209225.
Zbib, H.M., de la Rubia, T.D., 2002. A multiscale model of plasticity. International Journal of Plasticity 18 (9), 11331163.
Zbib, H.M., Khaleel, M.A., 2004. Recent advances in multiscale modeling of plasticity. International Journal of Plasticity 20 (6). 979-979.
Zbib, H.M., de la Rubia, T.D., Bulatov, V., 2002. A multiscale model of plasticity based on discrete dislocation dynamics. ASME Journal of Engineering
Materials and Technology 124 (1), 7887.
Zehetbauer, M., Schaer, E., Ungar, T., 2005. Vacancies in plastically deformed copper. Zeitschrift fur Metallkunde/Materials Research and Advanced
Techniques 96, 10441048.
Zhou, M., 2005. Thermomechanical continuum representation of atomistic deformation at arbitrary size scales. Proceedings of the Royal Society A 461
(2063), 34373472.
Zhou, M., McDowell, D.L., 2002. Equivalent continuum for dynamically deforming atomistic particle systems. Philosophical Magazine A 82 (13), 25472574.
Zhou, H., Feng, R., Diestler, D.J., Zeng, X.C., 2005. Coarse-grained free-energy-functional treatment of quasistatic multiscale processes in heterogeneous
materials. Journal of Chemical Physics 123 (16), 164109.
Zhu, T., Li, J., Yip, S., 2004a. Atomistic study of dislocation loop emission from a crack tip. Physical Review Letters 93 (2), 025503.
Zhu, T., Li, J., Van Vliet, K.J., Ogata, S., Yip, S., Suresh, S., 2004b. Predictive modeling of nanoindentation-induced homogeneous dislocation nucleation in
copper. Journal of the Mechanics and Physics of Solids 52 (3), 691724.
Zhu, T., Li, L., Samanta, A., Kim, H.G., Suresh, S., 2007. Interfacial plasticity governs strain rate sensitivity and ductility in nanostructured metals. Proceedings
of the National Academy of Sciences of the United States of America 104, 30313036.
Zhu, T., Li, L., Samanta, A., Leach, A., Gall, K., 2008. Temperature and strain-rate dependence of surface dislocation nucleation. Physical Review Letters 100,
025502.
Ziegler, H., 1983. An introduction to thermomechanics. In: Becker, E., Budiansky, B., Koiter, W.T., Lauwerier, H.A. (Eds.), North Holland Series in Applied
Mathematics and Mechanics, vol. 21, second ed. North Holland, Amsterdam, New York.
Zohdi, T.I., 2003. Constrained inverse formulations in random material design. Computer Methods in Applied Mechanics and Engineering 192 (2830),
31793194.

Você também pode gostar