Thermodynamic coarsening of dislocation mechanics and the size-dependent continuum crystal plasticity

https://doi.org/10.1016/j.jmps.2009.12.002Get rights and content

Abstract

Starting from the standard coarsening of dislocation kinematics, we derive the size-dependent continuum crystal plasticity by systematic thermodynamic coarsening of dislocation mechanics.

First, we observe that the energies computed from different kinematic descriptions are different. Then, we consider systems without boundary dissipation (relaxation) and derive the continuum approximation for the free energy of elastic–plastic crystals. The key elements are: the two-dimensional nature of dislocation pile-ups at interfaces, the localized nature of the coarsening error in energy, and, the orthogonal decomposition theorem for compatible and incompatible elastic strain fields. Once the energy landscape is defined, the boundary dissipation is estimated from the height of energy barriers.

The characteristic lengths are the average slip plane spacing for each slip system. They may evolve through the double-cross slip mechanism. The theory features the slip-dependent interface free energy and interface dissipation for penetrable interfaces.

The main constitutive parameters are derived from elasticity. The exception is the dependence of interface energy on slip plane orientation, which is determined from numerical results.

The theory requires no higher order boundary conditions. The only novel boundary conditions are kinematic, involving slip relaxation on the two sides of an interface.

Introduction

Classical crystal plasticity is size-invariant, and incapable of incorporating boundary conditions for slip. Yet, numerous experiments1 indicate pronounced size dependence, once the experimental length scale is below about 100 μm. Moreover, the need to impose boundary conditions on slip at an interface between two crystals is evident (Sun et al., 2000).

The physical mechanism for the observed size-dependence has been elucidated by Ashby (1970): In small volumes the motion of dislocations is constrained, resulting in pile-ups of dislocations with identical Burgers vector—the geometrically necessary dislocations. Following Ashby's physical argument, a number of heuristic mathematical formulations have been proposed during the last two decades.2 The theories include higher order kinematic gradients, which, in turn, require characteristic length(s) (Mindlin, 1964; Aifantis, 1987), and inevitably produce qualitative prediction of size effect. The higher order gradients are typically related to Nye's (1953) dislocation density tensor. While this choice is intuitive, it is by no means unique. Moreover, the characteristic lengths are typically phenomenological, without clear physical meaning that can be connected to dislocation mechanics.

More recently, the focus has been shifted to interfaces3, with particular emphasis on dissipative models. However, analyses are mostly on the continuum level; starting from phenomenological assumptions, mathematical consequences are explored.

Attempts to fit the predictions of a theory to either experiments4 or simulations5 indicate a need for more rigorous approach, i.e., a derivation of a continuum theory by systematically coarsening dislocation mechanics.

Advances in statistical analysis of collective dislocation behavior6 have resulted in statistical continuum dislocation dynamics formulations7 (SCDD). The SCDD formulations emphasize dynamics of dislocation motion, while representing dislocations as continuum density fields. The key ingredient stems from the attempt to resolve the fundamental problem of discrete dislocation dynamics: long-range interactions. This is achieved by computing a local approximation to the driving force (the Peach–Koehler force). The basis for this approximation is the assumption that dislocation–dislocation correlation function is short-range (Groma, 1997). There are several important differences between these formulations and the present approach.

  • (1)

    In this paper, we take a point of view that size effects in plasticity are the result of dislocation pile-ups against the boundaries, and consider highly correlated stacked dislocation pile-ups.

  • (2)

    In our formulation, the long-range forces are the part of the continuum problem (Kroner, 1958), while the additional free energy is the error caused by kinematic coarsening (Roy et al., 2007; Geers et al., 2009). We show that this error is localized. This leads to natural treatment of all possible boundary conditions that can be applied to a continuum crystal plasticity problem.

  • (3)

    We aim to develop a more economical (and inevitably, less accurate) quasi-static formulation, in which the number of degrees of freedom will be comparable to the one in the standard crystal plasticity. In contrast to SCDD formulations, we will not attempt to follow the dynamics of dislocation fields, but concentrate on constrained thermodynamics. Moreover, we will assume that the phenomenological treatment of statistical hardening in standard crystal plasticity is sufficiently accurate for our purposes, and focus on the thermodynamic differences that arise from ordering of dislocations in pile-ups against the boundaries.

It should be stated at the outset, that the complexities of dislocation mechanics (complex geometry and long-range interactions) all but preclude a mathematically rigorous approach. Nevertheless, while some simplifying assumptions are necessary, the errors should be, in principle, quantifiable. Moreover, the requirement that all new quantities, fields and constitutive parameters, have a clear physical meaning at the level of dislocation mechanics is not contradictory to quantifiable assumptions.

The purely kinematic coarsening has been accomplished half century ago (Nye, 1953; Kroner, 1958). In this work we focus on thermodynamic coarsening. In particular, we aim to accurately express the free energy of the elastic–plastic crystals in terms of continuum fields, but on the basis of the accurate free energy, as computed using the discrete dislocation model.

The dissipation is, in general, a more difficult problem. Elementary dissipative events occur at the atomic length at time scales, so that the discrete dislocation model is of limited use. Nevertheless, dissipated energy can often be estimated if the energy landscape is known. The typical event consists of a peak energy configuration, followed by rapid relaxation, and the dissipated energy is approximated as the difference between the two energy levels.

The paper is organized as follows. We summarize different kinematic descriptions in Section 2. Then, in Section 3, we briefly review some relevant elements of the standard crystal plasticity and Kroner's (1958) continuum theory of dislocations.

In Section 4, we first note that the energies computed from actual and coarsened kinematics must be different, as illustrated earlier by Roy et al. (2007). Then, we define the framework in which the correction to the energy, i.e., the error resulting from kinematic coarsening, can be introduced into the framework of crystal plasticity. The orthogonal decomposition theorem for compatible and incompatible strain fields guarantees that this correction is orthogonal to the compatible strain, and therefore additive. It is a novel result, possibly with broader implications.

In Section 5, we consider several dislocation mechanics problems with simple geometries and thus amenable to analysis. The key observation is that the relevant configurations are the stacked pile-ups in parallel slip planes, and those—much like crack tips, have predominantly two-dimensional nature. We conclude that for slowly varying slip gradients, the coarsening error in energy density can be represented as a local function of slip gradients. Sharp boundary layers, resulting from pile-ups of dislocations, violate the slow variation condition, and are best modeled by a slip-dependent surface energy. The characteristic lengths that emerge from the analysis—one for each activated slip system, are the average spacings of slip planes in each slip system.

In Section 6, we formulate the continuum theory for single slip, and in Section 7, we generalize the results for multiple slip problems.

Consistent with our view that a continuum expression for the free energy is the primary goal from which the dissipation model can be deduced, we considered first the interfaces impenetrable to dislocations. The penetrable interfaces, interface relaxation, and the accompanying dissipation are discussed in Section 8. The summary and discussion are given in Section 9.

At the outset, we limit our analysis to: linearized kinematics theory with additive decomposition of elastic and plastic deformation, elastic isotropy, and isothermal deformation.

Section snippets

Representations of dislocations and kinematic coarsening

The densities of geometrically necessary (GN) dislocations8

Standard crystal plasticity and the Kroner's theory

For the reasons that will become clear later, we use the notation different from the standard crystal plasticity notation.9 The slip system α is defined by the unit normal mα, and two orthogonal in-plane unit vectors: ζα, parallel to the Burgers vector and in the direction of a screw segment, and, ζα, orthogonal to the Burgers vector and in the direction of an edge segment, so that (Fig. 3):ζα×ζα=mα.

The slips γα, on all active slip

Thermodynamics of crystal plasticity and microstructural energy

Consider the finite volume V, bounded by the surface S. The weak form (principle of virtual work) of the classical crystal plasticity can be written as:δW=V(σ:δe+αταδγα)dV=St·δudS.

The first term with elastic strain e, and the corresponding stress σ=C:e, is the variation of the elastic strain energy:Φ=12Ve:C:edV,

The second term in (16) includes dissipation, as well as the statistically stored energy (in the sense of Ashby, 1970), which is considered mechanically irreversible. This is

2D problems and the local approximation

Consider the configuration shown in Fig. 4, with inclined slip planes and pile-ups of dislocations at the boundary. In isotropic elasticity, the problems of edge (plane strain) and screw (anti-plane) dislocations are orthogonal and stresses, strains and energies are additive (Hirth and Lothe, 1992). The general case with arbitrary Burgers vectors is obtained by superposition of the two basic problems.

We consider equally spaced slip planes. This is, of course, an approximation. We expect that it

Continuum crystal plasticity for single slip

Consider a crystal deforming on a single slip system and embedded into elastic matrix with identical elastic properties (Fig. 7). The continuum slip γα, and the slip gradient components gα and gα (12), are represented in terms of their counterparts in the semi-discrete model, sα and B=B⊥/∘:γα=sα/hα;gα=B/hα;gα=B/hα.

Since the singular distribution of gα at the boundary will be replaced by the appropriate boundary value of slip, the geometry at the intersection of slip plane and the boundary

Continuum crystal plasticity for multiple slip

The problem of interaction energies between dislocations on multiple slip systems is significantly simplified by the local nature of the coarsening error in microstructural energy; the orientation dependence of interaction energy between two segments can be separated from its distance dependence. Such orientation dependence can be obtained either from interaction of discrete dislocation segments (Hirth and Lothe, 1992, Eqs. (4), (5), (6), (7), (8), (9), (10), (11), (12), (13), (14), (15), (16),

Penetrable boundaries: free energy and dissipation

When two plastically deforming crystals are interfaced, dislocation segments in boundary layers can undergo complex reactions.12 The typical result is partial annihilation of Burgers vectors on the two sides of the interface and formation of extra dislocation walls. From the continuum point of view, the conditions that preserve the total Burgers vector have been studied by Gurtin and Needleman (2005). In the present case, the net

Summary and discussion

Starting from the continuous kinematics of Nye (1953) and Kroner (1958), and by noticing that the energies computed from discrete and continuous kinematics are different, we have derived the continuum approximation for the free energy of elastic–plastic crystals.

At first, we considered systems without boundary dissipation (relaxation). Once the energy landscape is defined, the dissipation model for the boundaries is defined from the height of energy barriers (peak energy configurations).

The

Acknowledgements

This work was supported by the US National Science Foundation, grant # CMS-0528550, and, the US Department of Energy, grant # DE-FG02-05ER25709.

References (59)

  • N.A. Fleck et al.

    Strain gradient plasticity: theory and experiment

    Acta metal. mater.

    (1994)
  • N.A. Fleck et al.

    A mathematical basis for strain-gradient plasticity theory—part I: scalar plastic multiplier

    J. Mech. Phys. Solids

    (2009)
  • H. Gao et al.

    Mechanism-based strain gradient plasticity. I. Theory

    J. Mech. Phys. Solids

    (1999)
  • I. Groma et al.

    Spatial correlations and higher-order gradient terms in a continuum description of dislocation dynamics

    Acta Mater.

    (2003)
  • P. Gudmundson

    A unified treatment of strain gradient plasticity

    J. Mech. Phys. Solids

    (2004)
  • M.E. Gurtin

    A gradient theory of single-crystal viscoplasticity that accounts for geometrically necessary dislocations

    J. Mech. Phys. Solids

    (2002)
  • M.E. Gurtin

    On a framework for small-deformation viscoplasticity: free energy, microforces, strain gradients

    Int. J. Plasticity

    (2003)
  • M.E. Gurtin et al.

    Boundary conditions in small-deformation, single crystal plasticity that account for the Burgers vector

    J. Mech. Phys. Solids

    (2005)
  • J.P. Hirth et al.

    Disconnections in tilt walls

    Acta Mater.

    (2006)
  • T.A. Khraishi et al.

    Dislocation dynamics simulations of the interaction between a short rigid fiber and a glide circular dislocation pile-up

    Comp. Mater. Sci.

    (2002)
  • J.W. Kysar et al.

    High strain gradient plasticity associated with wedge indentation into face-centered cubic single crystal: Geometrically necessary dislocation densities

    J. Mech. Phys. Solids

    (2007)
  • S. Limkumnerd et al.

    Study of size effects in thin films by means of a crystal plasticity theory based on DFT

    J. Mech. Phys. Solids

    (2008)
  • S.Dj. Mesarovic

    Energy, configurational forces and characteristic lengths associated with the continuum description of geometrically necessary dislocations

    Int. J. Plasticity

    (2005)
  • W.D. Nix et al.

    Indentation size effects in crystalline materials: a law for strain gradient plasticity

    J. Mech. Phys. Solids

    (1998)
  • J.F. Nye

    Some geometrical relations in dislocated crystals

    Acta Metall.

    (1953)
  • T. Ohashi

    Three dimensional structures of the geometrically necessary dislocations in matrix-inclusion systems under uniaxial tensile loading

    Int. J. Plasticity

    (2004)
  • W.J. Poole et al.

    Micro-hardness of annealed and work-hardened copper polycrystals

    Scripta Mater.

    (1996)
  • K. Shizawa et al.

    A thermodynamical theory of gradient elastoplasticity with dislocation density tensor. I: Fundamentals

    Int. J. Plasticity

    (1999)
  • J.Y. Shu et al.

    Boundary layers in constrained plastic flow: comparison of nonlocal and discrete dislocation plasticity

    J. Mech. Phys. Solids

    (2001)
  • Cited by (0)

    View full text