Sei sulla pagina 1di 121

Microstructure evolution in crystal plasticity:

strain path effects and dislocation slip


patterning

This research was carried out under the project number MC2.03158 in the framework of
the Research Program of the Materials innovation institute M2i (www.m2i.nl), the former
Netherlands Institute for Metals Research.

CIP-DATA LIBRARY TECHNISCHE UNIVERSITEIT EINDHOVEN


Tuncay Yalcnkaya
Microstructure evolution in crystal plasticity: strain path effects and dislocation
slip patterning /
by T. Yalcnkaya Eindhoven : Technische Universiteit Eindhoven, 2011.
Proefschrift.
A catalogue record is available from the Eindhoven University of Technology
Library
ISBN: 978-90-386-2729-8
Subject headings: BCC metals / crystal plasticity / non-Schmid effects /
plastic anisotropy / strain path change effect / Bauschinger effect /
cross effect / microstructure evolution / non-convexity /
phase field modeling / dislocation patterning / finite element method /
non-convex free energy / strain gradient crystal plasticity
c
Copyright 2011
by Tuncay Yalcnkaya, all rights reserved.
This thesis was prepared with the LATEX 2 documentation system.
Reproduction: Universiteitsdrukkerij TU Eindhoven, Eindhoven, The
Netherlands.

Microstructure evolution in crystal plasticity:


strain path effects and dislocation slip
patterning

P ROEFSCHRIFT
ter verkrijging van de graad van doctor
aan de Technische Universiteit Eindhoven,
op gezag van de rector magnificus, prof.dr.ir. C.J. van Duijn,
voor een commissie aangewezen door het College voor Promoties
in het openbaar te verdedigen
op donderdag 20 oktober 2011 om 16.00 uur

door

Tuncay Yalcnkaya

geboren te Ankara, Turkije

Dit proefschrift is goedgekeurd door de promotor:


prof.dr.ir. M.G.D. Geers
Copromotor:
dr.ir. W.A.M. Brekelmans

Contents

Summary

ix

1 Introduction
1.1 Crystal plasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Objective and outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1
1
3

2 A finite strain BCC single crystal plasticity model and its experimental
identification
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Slip mechanisms in BCC metals . . . . . . . . . . . . . . . . . . . . . . .
2.3 Violation of Schmids law in BCC metals . . . . . . . . . . . . . . . . . .
2.4 A BCC crystal plasticity model at material point level . . . . . . . . . .
2.4.1 Kinematics in crystal plasticity . . . . . . . . . . . . . . . . . . .
2.4.2 Constitutive model . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5 Modeling some intrinsic properties of BCC single crystals . . . . . . . .
2.5.1 Orientation dependence . . . . . . . . . . . . . . . . . . . . . . .
2.5.2 Example: -Fe single crystal . . . . . . . . . . . . . . . . . . . .
2.5.3 Example: molybdenum single crystal . . . . . . . . . . . . . . .
2.5.4 Temperature dependence . . . . . . . . . . . . . . . . . . . . . .
2.6 Summary and Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . .

5
5
8
10
12
12
13
17
17
18
19
19
22

3 A composite dislocation cell model to describe strain path change effects in


BCC metals
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Dislocation substructure evolution . . . . . . . . . . . . . . . . . . . . .
3.3 Computational model . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4 Modeling of microstructure evolution . . . . . . . . . . . . . . . . . . .
3.4.1 Monotonic deformation . . . . . . . . . . . . . . . . . . . . . . .
3.4.2 Orthogonal change of deformation . . . . . . . . . . . . . . . . .
3.4.3 Reverse deformation . . . . . . . . . . . . . . . . . . . . . . . . .

25
25
28
31
34
34
36
37

vi

Contents

3.5

3.6

Numerical examples . . . . . . . . . . . . . . . . . . . . . . .
3.5.1 Example 1: monotonic deformation of single crystals
3.5.2 Example 2: strain path change of single crystals . . .
3.5.3 Example 3: strain path change of polycrystals . . . .
Summary and Conclusion . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

4 Deformation patterning driven by rate dependent non-convex strain gradient plasticity


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Macroscopic view: material instability and microstructure evolution
in inelastic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3 Thermodynamics of strain gradient plasticity . . . . . . . . . . . . . . .
4.4 Particular choices of free energy functions . . . . . . . . . . . . . . . . .
4.4.1 Slip based strain gradient plasticity . . . . . . . . . . . . . . . .
4.4.2 Slip based non-convex strain gradient plasticity . . . . . . . . .
4.5 Non-convexity and patterning in phase field modeling . . . . . . . . .
4.6 Numerical examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.6.1 Numerical example 1: convex case - monotonic loading . . . . .
4.6.2 Numerical example 2: non-convex case - monotonic loading . .
4.6.3 Numerical example 3: non-convex stress relaxation of a 1D bar
4.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.8 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.8.1 Finite element implementation of slip based strain gradient
plasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.8.2 Finite element implementation of slip based non-convex strain
gradient plasticity . . . . . . . . . . . . . . . . . . . . . . . . . .

38
38
39
40
43

45
45
49
50
54
54
55
57
58
59
60
66
68
68
68
71

5 Non-convex rate dependent strain gradient crystal plasticity and deformation patterning
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Strain gradient crystal plasticity and finite element implementation . .
5.3 Latent hardening based non-convex plastic potential . . . . . . . . . . .
5.3.1 Conditions for plastic slip patterning . . . . . . . . . . . . . . .
5.4 Numerical analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.4.1 Convex strain gradient crystal plasticity . . . . . . . . . . . . . .
5.4.2 Non-convex strain gradient crystal plasticity . . . . . . . . . . .
5.5 Summary and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . .

73
73
75
80
80
84
84
88
92

6 Discussion and conclusions

95

Bibliography

99

Contents

vii

Dankwoord / Acknowledgements

109

Curriculum Vitae

111

viii

Summary

During deformation polycrystalline metals tend to develop heterogeneous plastic


deformation fields at the microscopic scale, as the amount of plastic strain varies
spatially, depending on local grain orientation, geometry and defects. While grain
boundaries are natural places triggering plastic slip accumulation and geometrically
necessary dislocations that accommodate the gradients of the inhomogeneous plastic
strain, the deformation localizes within grains revealing dislocation cell structures or
micro slip bands (e.g. clear band formation in irradiated materials). Across grains,
macroscopic plastic slip bands (Luders

bands, etc.) exist as well. These intergranular and intragranular deformation patterns are stated to be inherent minimizers
of the free energy (including the microstructurally trapped plastic energy). These
microstructures may macroscopically manifest themselves through softening of the
material or through plastic anisotropy in hardening under strain path changes. These
effects are crucial with respect to the mechanics of the materials under consideration
and should be taken into account in the constitutive modeling.
In this thesis, the computational modeling of microstructure evolution (with softening or plastic anisotropy) is covered in different crystal plasticity frameworks. The
scope is basically two-fold. First, in the chapters 2 and 3 the plastic anisotropy of
Body Centered Cubic crystals is studied from the onset of deformation due to an
intrinsic orientation dependence from non-planar dislocation core structures, to the
anisotropy upon a strain path change owing to resulting dislocation cell formation.
In this part of the thesis, after developing a proper BCC crystal plasticity framework
taking into account the intrinsic anisotropy, a composite cell model was established
where the evolution of dislocation cells was modeled under monotonic and nonproportional loading histories. Here, the existence of a dislocation microstructure
is introduced into the model in terms of internal variables and the evolution was
described by phenomenologically based evolution equations. However, this phenomenological approach is not able to incorporate the formation stage of the microstructure. Hence, the crystal plasticity framework called for an extension in order
to capture the evolution of the microstructure driven by the free energy of the mateix

Summary

rial.
In order to complete the missing link between the formation of the microstructure and its evolution in crystal plasticity frameworks, the second part of the thesis concentrates on the development of a non-convex rate dependent crystal plasticity model, which reveals a rate dependent dislocation microstructure formation
and evolution together with macroscopic hardening-softening-plateau stress-strain
responses. To this end, non-convexity is treated as an intrinsic property of the plastic
free energy of the material. First, this non-convex contribution is incorporated into a
strain gradient crystal plasticity framework with a double-well character, which results in a computational routine partially dual to the Ginzburg-Landau type of phase
field modeling approaches (with high and low slipped regions representing the different phases). In this model, both the displacement and the plastic slip fields are
considered as primary variables. These fields are determined on a global level by
solving simultaneously the linear momentum balance and the slip evolution equation which is rederived in a thermodynamically consistent manner. In chapter 4, the
analysis is conducted in a 1D mathematical setting in order to illustrate the ability of
the model to capture the patterning of plastic slip. In chapter 5 and inspired by the
literature, the non-convexity originates from latent hardening in a multi-slip strain
gradient crystal plasticity framework. Hence, the 1D approach pursued in chapter 4
is extended to a 2D plane strain setting. Even though the phenomenological doublewell free energy function used in the 1D approach allows to track non-equilibrium
states during microstructure evolution, it does not rely on a physically based expression for non-convexity, but presents a generic formulation. Instead, chapter 5
concentrates more on the physical reasons of plastic slip localization, where a slip
interaction potential is analyzed and incorporated into the rate dependent strain gradient crystal plasticity framework. The non-convexity due to the slip interactions is
explicitly illustrated and the possibility of deformation patterning in the material is
discussed in a boundary value problem. The last part of the thesis, chapter 6, presents
a discussion and conclusions.

Chapter one

Introduction

Abstract / The physics and the basic principles behind the general crystal plasticity modeling are explained. An overview of strain path change related anisotropy and dislocation
microstructure evolution in crystal plasticity approaches is presented. The objectives and
outline of the thesis are given.

1.1 Crystal plasticity


A perfect metallic single crystal, which is characterized by a specific periodic arrangement of atoms, can respond only in a reversible elastic manner in thermal equilibrium with its surroundings when stressed monotonically well below the critical
levels that destabilize the crystal structure. Under an applied homogeneous stress,
the elastic response is homogeneous down to the atomic level. In contrast, the plastic
response is locally heterogeneous and requires crystal defects for its development.
The type and intensity of the plastic response depend on the character of the defect
state. For this purpose the crystal defects are introduced in a hierarchy of increasing
dimensionality from point, through line, to planar defects (see Argon (2008) for an
extensive overview). Among these, the line defects, i.e. dislocations, are regarded as
the principal carriers of plastic deformation. The crystallographic slip of dislocations
occurs on the most close-packed slip planes and in the most close-packed directions
which together form the slips systems. Depending on the specific arrangement of
atoms, each metal has specific slip systems. The first part of this thesis is focusing on
the body centered cubic (BCC) type of atomic arrangement.
Under an applied stress, the atomic lattice deforms elastically until the stretched
bonds near a dislocation break down and new bonds are formed. During this prop-

1 Introduction

CONSTITUTIVE MODELING OF DISLOCATION MOVEMENT

DISLOCATION CELLS

GRAINS

DISLOCATION CELL BLOCKS

Figure 1.1 / Crystal plasticity bridges the deformation of bulk material and the
movement of dislocations (chapter 2). Clustering of dislocations is accomplished
via non-convex strain gradient crystal plasticity (chapter 4 and 5). The phenomenological evolution of dislocation structures is simulated via a dislocation cell model
(chapter 3).

agating process, a part of the crystal gradually slips one interatomic distance with
respect to the other part. Instead of the ideal fictitious strength associated with the
movement of an entire slip plane, the dislocations enable only sections of the slip
plane to shear, resulting in the observed decimated strengths necessary for plastic
deformation.
The stress directly affecting the motion of dislocations is the projected shear stress
on the specific slip systems, which is also called the resolved Schmid stress. When
the resolved shear stress is larger than the resistance on the respective slip system,
glide is activated. The viscous type of crystal plasticity theories as used in this thesis
employs a power law relation between the rate of plastic slip and the ratio between
the shear stress and the slip resistance. Conceptually, all the slip systems are active
but the most favorable ones carry most plastic deformation. Crystal plasticity is a
mesoscopic modeling approach, bridging the stresses on the slip systems to the total
amount of plastic slip in the bulk material (see Fig. 1.1), eventually affecting the total
amount of macroscopic plastic strain.
In addition to their role of accommodating the plastic deformation in metals, the
work (strain) hardening behavior can also be attributed to the dislocations. The
distinct stages of strain hardening are related to their multiplication or mutual in-

1.2 Objective and outline

teraction processes. After moderate deformations, the formation of dislocation cell


structures (see Fig. 1.1) plays an important role in the hardening behavior of the
material.
Another important effect in the hardening of crystals is the gradient of the plastic
slip, requiring so-called geometrically necessary dislocations (GNDs). Once the applied load, or the material structure itself, triggers a gradient of the plastic deformation, a certain amount of GNDs will be necessary to preserve lattice compatibility
and to accomplish the required lattice rotation. Conventional crystal plasticity theories, as used in the chapters two and three of this thesis, however, do not take into
account these effects. Their strengthening mechanisms are therefore inherently incapable of predicting scale dependent behavior, i.e. different mechanical responses
due to varying plastic strain gradients. Hence, in the strain gradient crystal plasticity frameworks of the subsequent chapters, the gradients of the plastic slip enter the
plastic slip law together with a length scale parameter, where these do not only allow
for size effect predictions but also play a regularization role in the viscous formulation of non-convex strain gradient plasticity.

1.2 Objective and outline


During metal forming processes, materials experience complex strain path histories
which result in plastic anisotropy, i.e. transient hardening or softening regimes occurring in the macroscopic stress-strain response. This phenomenon plays an important role in the metal deformation and the effect should be included in the constitutive modeling. The physical origin resides in three distinct factors at three different
length scales: dislocation slip anisotropy, evolution of the dislocation microstructure
and textural anisotropy. The thesis focuses on the first two effects which are crucial
in early stages of the deformation and subsequent moderate straining.
In chapter two, a crystal plasticity model for body centered cubic (BCC) single crystals, taking into account the plastic anisotropy due to non-planar spreading of screw
dislocation cores is developed. A comprehensive summary of the intrinsic properties
of these materials is presented and incorporated into the framework through a modification in the plastic slip law. In the numerical examples section, emphasis is given
on the intrinsic orientation dependence of the flow stress due to the non-Schmid components of the stress field projected on the slip plane under monotonic deformation.
Attention is therefore given on the quantitative prediction of single crystal behavior,
for which experiments from the literature have been used.
Next, in chapter three, the anisotropy due to the dislocation cell structure evolution
is considered. A composite dislocation cell model has been combined with the BCC

1 Introduction

crystal plasticity framework to describe the dislocation cell structure evolution and
its macroscopic anisotropic effects. The computational framework departs from a
composite aggregate with a cell structure, consisting of a soft cell interior component
and hard cell wall components. The constitutive response of each component has
been obtained from crystal plasticity simulations, while a set of phenomenological
evolution equations for the cell size, the wall thickness and the dislocation density
captures the evolution of the microstructure under complex strain paths. Numerical
examples study both the intrinsic orientation dependence and the anisotropy due to
cell structure evolution.
Chapter four focuses on one of the origins of self-organizing dislocation structures
(driven by the deformation) rather than imposing the cell structure evolution as done
in the previous chapter. To this purpose, a rate dependent strain gradient plasticity
framework for the description of plastic slip patterning in a system with non-convex
energetic hardening is developed. Both the displacement field and the plastic slip
field are considered as primary variables. The slip law differs from classical ones in
the sense that it includes a stress term originating from a non-convex double-well free
energy, which enables patterning of the deformation field. The derivations and implementations are performed in a single slip 1D setting, which allows for a thorough
mechanistic understanding, not excluding its extension to multidimensional cases.
The numerical examples illustrate both the homogeneous and inhomogeneous plastic slip distributions as well as the stress-strain response in relation to the imposed
boundary conditions and the applied rate of deformation.
In chapter five the non-convex strain gradient crystal plasticity formulation is extended to the 2D plane strain case, including multiple slip systems. In order to
capture the effect of dislocation interactions on the non-convexity of the plastic slip
dependent free energy function, a more physically based free energy expression is
incorporated. Attention is focused on the inhomogeneous plastic slip distribution
and deformation patterning due to dislocation slip interactions.
The thesis concludes with a final chapter, summarizing the main achievements and
results obtained as well as an outlook to open challenges.

Chapter two

A finite strain BCC single crystal


plasticity model and its experimental
identification1

Abstract / A crystal plasticity model for body-centered-cubic (BCC) single crystals, taking into account the plastic anisotropy due to non-planar spreading of screw dislocation
cores is presented. In view of the long-standing contradictory statements on the deformation of BCC single crystals and their macroscopic slip planes, recent insights and developments are reported and included in this model. The flow stress of BCC single crystals
shows a pronounced dependence on the crystal orientation and the temperature, mostly
due to non-planar spreading of a/2h111i type screw dislocation cores. The main consequence here is the well-known violation of Schmids law in these materials, resulting in
an intrinsic anisotropic effect which is not observed in e.g. FCC materials. Experimental confrontations at the level of a single crystal are generally missing in the literature.
To remedy this, uniaxial tension simulations are done at material point level for -Fe,
Mo and Nb single crystals and compared with reported experiments. Material parameters, including non-Schmid parameters, are calibrated from experimental results using
a proper identification method. The model is validated for different crystal orientations
and temperatures, which was not attempted before in the open literature.

2.1 Introduction
In the present paper, attention is focused on the low temperature (room temperature and lower) properties of single BCC crystals including -Fe, metals of the group
VA (V, Nb, Ta) and of the group VIA (Mo, V, Cr) and certain alkali metals. These
materials show a peculiar mechanical behavior, mostly resulting from their screw
1

This chapter is reproduced from Yalcinkaya et al. (2008)

2 A finite strain BCC single crystal plasticity model and its experimental
identification

dislocation core configuration. They have a relatively high yield stress which is
strongly temperature, rate and orientation dependent. They exhibit complex slip
modes, dominated by the cross slip of a/2h111i screw dislocations. Due to their dislocation core structure, they show a severe glide direction sensitive behavior (slip
asymmetry) and the well-known Schmid law using the critical resolved shear stress
(CRSS) is violated. Another pronounced phenomenon is the anomalous slip (activation of an unexpected slip system at a certain orientation) observed in pure BCC
metals which, however, is only marginally dealt with in the present work.
Various discussions on the behavior of BCC crystals, reveal a number of contradictions with respect to the slip plane activity. Even though there is no generally accepted explanation to the dislocation behavior of these materials recent studies provide a good basis for the constitutive model presented in the current paper. Seeger
(2001) states that the slip nature of BCC crystals depends highly on the temperature, upon which dislocations may accommodate either a straight {110} slip or a
wavy type {112} cross slip pattern. At room temperature a/2h111i type of screw
dislocations move on {112} slip planes, which enables cross slip. The cross slip phenomenon will not be modeled explicitly but the resulting effects are taken into account in the slip and hardening laws (for example (Pichl (2002)), showing how cross
slip can be included in a model).
The value of the critical resolved shear stress (CRSS) is independent from the slip system and the sense of the slip of FCC metals. For these metals, it is generally accepted
that the only stress component affecting the glide is the Schmid stress. However,
BCC metals show an asymmetry in their slip behavior: the slip resistance in one
direction is different from the resistance in the opposite direction, indicated as the
twinning/anti-twinning asymmetry. Moreover, due to small edge fractional dislocation components in the screw dislocation core, stress components other than the
resolved Schmid stress affect the glide or the CRSS of the material. Both of these effects result from the non-planar spreading of the dislocation cores. For these reasons
Schmids law is not applicable to BCC metals. In the presently proposed crystal plasticity model these two types of intrinsic anisotropy effects will be taken into account.
The formulation of the constitutive crystal plasticity model departs from the papers
of Bronkhorst et al. (1992) and Kalidindi et al. (1992) related to FCC metals. Their developments are extended by including the intrinsic properties of BCC metals,using
a physical description of the slip law based on thermally activated dislocation kinetics. The barriers to dislocation movement are discriminated according to their
short-range or long-range nature. The short-range barrier can be overcome by thermal activation, whereas the long-range barrier is affected slightly through changes
of the elastic moduli. Non-Schmid effects are included in the model by incorporating non-Schmid terms in the slip activation. The model is implemented at a material

2.1 Introduction

point in matlab and compared with experimental results.


The phenomena incorporated in this work are basically the orientation and the temperature dependence of the flow stress and stress-strain behavior of single BCC crystals including the non-Schmid behavior. Each of these characteristics has been investigated extensively in the literature, and especially the temperature dependence of
the flow stress was an active research area until the 90s. At present, atomistic computer simulations studying screw dislocation cores are still an active area of research.
Many improvements have been achieved in this area and to our knowledge there is
no recent work combining these physical aspects of BCC structured materials with
crystal plasticity calculations. The objective of the present work is to exemplify this
combination.
The plan of this paper is as follows. Section 2 discusses the slip mechanisms in BCC
metals where the temperature dependence of the slip plane activation is strongly emphasized. Next, Schmids law and its violation in BCC crystals is handled in section
3, along with its connection to the non-planar spreading of screw dislocation cores.
In section 4, the crystal plasticity constitutive framework and its implementation is
outlined. Section 5 studies pronounced intrinsic properties, whereby examples are
presented and confronted with experimental results. Finally, concluding remarks are
given in section 6.
Cartesian tensors and associated tensor products will be used throughout this paper,
making use of a Cartesian vector basis {e1 e2 e3 }. Using the Einstein summation rule
for repeated indices, the following conventions are used in the notations of vectors,
tensors, related products and crystallography:
scalars a
vectors a = ai ei
second-order tensors A = Ai j ei e j
fourth-order tensors 4 A = Ai jkl ei e j ek el
C = a b = ai b j ei e j
C = A B = Ai j B jk ei ek
C = 4 A : B = Ai jkl Blk ei e j
crystallographic direction, family [uvw], huvwi
crystallographic plane, family (hkl ),{hkl }
slip system, family (hkl )[uvw], {hkl } huvwi

2 A finite strain BCC single crystal plasticity model and its experimental
identification

2.2 Slip mechanisms in BCC metals


First, some of the long-standing contradictions in the identification of the slip planes,
and the active slip mechanisms of BCC crystals are highlighted. The first attempt
goes back to the introduction of the pencil glide mechanism by Taylor and Elam
(1926), where the slip was assumed to be oriented in the h111i crystallographic direction while the mean plane of slip was the one having the maximal projected shear
stress. This plane might be a crystallographic but also a non-crystallographic plane.
After this pioneering research, there have been several of contradicting statements on
the active slip planes of BCC metals, for which an extended overview can be found
in Havner (1992). The different concepts will not be repeated here in detail, however,
a summary including the current developments will be presented in the following
paragraphs.
Gough (1928) and Barrett et al. (1937) state that the {110}, {112} and {123} families
contain the crystallographic slip planes during the deformation of BCC metals, an
assumption that is still being used in many crystal plasticity works. Another frequently used view is the participation of {110} and {112} slip planes only, whereby
it is assumed that {123} planes need a higher temperature for activation. Many researchers state that only the {110} slip planes are active at room temperature, based
on the argument that apparent slip on both {112} and {123} planes is actually composed of slip on two non-parallel {110} planes, e.g. Chen and Maddin (1954). Using
the latter argument, it would be physically more comprehensible to indeed model
only {110} planes in a crystal plasticity framework.
In this paper, attention is focused on an accurate description of the physical slip
mechanisms of BCC crystals, rather than re-advocating a discussion on the active
set of slip planes. As a result of their special slip mechanisms, BCC metals have
interesting intrinsic properties that are not observed in e.g. FCC metals. Many authors (e.g. Vitek et al. (2004b), Vitek (2004), Duesbery and Vitek (1998)) related most
of the phenomena to the core structure of screw dislocations, e.g. by performing
atomistic simulations. Slip system activation in BCC metals is highly dependent on
the crystal orientation and especially on the temperature. Seeger (2001) and Seeger
and Wasserbach (2002) (see Fig. 2.1) provide a detailed explanation of temperature
dependent slip for high purity BCC single crystals with an orientation inducing a
) and = 0.433 for the sysSchmid factor = 0.500 for the slip system [111](101
) and [111](211
). The work of Seeger and co-workers is adopted here,
tems [111](1 12
relying on the fact that BCC metals show different features and different slip mechanisms in different temperature ranges. The physical response below the so-called
knee temperature (which is around 0.2 times the melting temperature of the metal,
TK in Fig. 2.1) and above the knee temperature is thereby distinguished. Below the
knee temperature the slip is governed by the glide of a/2h111i type screw disloca-

2.2 Slip mechanisms in BCC metals

1000
(101)
slip

cross slip of [111] screws; elementary steps on (211) and (112)

800

Flow Stress (MPa)

wavy slip lines

600
straight
slip lines

400

200

100

200
300 T
Temperature T (K)

400

TK 500

Figure 2.1 / Flow stress vs. temperature curve for a pure Mo single crystal at a
plastic shear strain rate 8.6 104 s1 , from Seeger (2001).

tions in kink pairs as the mobility of screw dislocations is lower than the mobility
of edge components. In this temperature range, the flow stress (the stress to maintain plastic deformation after yield) of the metal is highly temperature and strain
rate dependent. The flow stress decreases with increasing temperature and increases
with increasing strain rate. Above the knee temperature, the flow stress decreases
considerably due to self diffusion and recovery processes and the mobilities of screw
and non-screw dislocations are no longer substantially different. The high temperature range is out of the scope of the present work. Attention will be focused on the
behavior at lower temperatures, including the behavior at room temperature.
The flow stress dependence on the temperature and the slip mechanism in this temperature range is visualized in Fig. 2.1, where the presented numerical values refer
to molybdenum single crystals. Below the lower limit T (70 K for Mo and 120 K for
-Fe) the dislocation glide is confined to {110} planes. Screw dislocation glide produces straight step patterns. In this temperature range, dislocations are stated to be
in their ground state. They show a threefold symmetry and they are able to slip on
any of the three {110} slip planes. As a result of mirror symmetry and the absence of
cross slip, no plastic anisotropy is observed.
the dislocation core configuration changes and dislocations undergo a tranAbove T,
sition from their low temperature configuration (slipping on {110} planes) to their
high temperature configuration (slipping on {112} planes). The dislocations show
a wavy type of structure which results from the cross slip, a characteristic property

2 A finite strain BCC single crystal plasticity model and its experimental
10
identification

associated with {112} slip. In BCC metals three {110} and three {112} slip planes intersect on a common h111i direction and screw dislocations can therefore distribute
their core on these planes. This spreading is non-planar and nearly all peculiarities
in the mechanical properties of BCC metals can be attributed to this phenomenon.
The cross slip accommodated by {112} slip planes is the main source of the nonSchmid behavior (orientation dependence of the flow stress). Similar explanations
of the change in the slip mechanism in BCC materials were presented by others. For
example Christian et al. (1990) focused on the mean jump distance of screw dislocations in BCC metals, which decreases considerably above a critical temperature. The
explanation is again a change in the slip mechanism of BCC metals.
The above paragraphs emphasized some recent developments in the understanding
of the slip system activity of BCC crystals and the connection between the core structure and the intrinsic properties. The development of an adequate model will be
based on these considerations.

2.3 Violation of Schmids law in BCC metals


Schmid and co-workers (e.g. Schmid and Boas (1935)) first recognized that the yield
stress of a metal crystal is strongly depending on the crystal orientation with respect
to the load direction. Yield on the slip plane of a crystallographic family occurs at
a constant projected shear stress, which is called the critical resolved shear stress
(CRSS), for a particular material. Constant should be understood here as independent from the slip system and the sense of slip. The resulting Schmid law assumes
that the only stress component triggering plastic flow of the material is the projected
shear stress on the slip system, in the direction of glide which is called the Schmid
stress. The other non-glide component, defined as the normal stress, does not have
any effect on the plastic deformation. These assertions are applicable on FCC metals,
however not on BCC metals. This violation becomes manifest through their plastic
anisotropy, revealing two distinct intrinsic non-Schmid effects in BCC metals.
The first intrinsic non-Schmid effect is the variation of the CRSS with the sense of the
shear. By definition shear on a certain plane of the family (e.g. {112}) produces shear
}) produces
in the twinning direction, whereas shear on another plane (e.g. {211
shear in the anti-twinning direction (see Table 2.1 for the respective slip systems).
In BCC metals, twin and slip mechanisms share common slip systems and twinning
is only observed when the temperature is very low and the strain rate is extremely
high. In this context, the main source of the plastic deformation is just the glide of
dislocations, but the resistance to this movement in the twinning and anti-twining
directions is asymmetric. This is called the twinning/anti-twinning asymmetry of
BCC crystals in the literature and the source of this asymmetry is the strong cou-

11

2.3 Violation of Schmids law in BCC metals

pling of the screw dislocation to the BCC lattice, which constrains the core and its
properties to adopt the symmetry of the lattice (Duesbery and Vitek (1998)). From
the modeling point of view, this phenomenon is included in the models by taking a
different slip resistance for the twinning and anti-twinning planes.

Table 2.1 / {112} slip systems of BCC crystals, T and A referring to twinning/antitwinning planes

(4) A (112)[111]

(1) A (112)[111]

(2) A (112)[1
11]
(5) A (121)[111]
111]

(3) A (112)[
(6) A (121)[11
1]

(7) A (121)[111]
(10)T
111]

(8) A (121)[
(11)T

(9) T (211)[111]
(12)T

(211)[111]

(211)[11
1]
11]

(211)[1

The second effect, which is sometimes denoted as an extrinsic non-Schmid effect,


e.g. Duesbery and Vitek (1998), is the sensitivity of the slip resistance to the nonglide components of the applied stress. This effect originates from the non-planar
spreading of a/2h111i type screw dislocation cores. Whereas this effect is often
called extrinsic in the literature in view of its relation to the applied stress, it essentially remains a physically intrinsic non-Schmid effect owing its existence to the
dislocation core structure. The difference between intrinsic and extrinsic is therefore
not made further in this contribution. In BCC metals, the non-glide component of
the applied stress tensor (in a direction perpendicular to the Burgers vector) is crucial. A Burgers vector in a BCC crystal can be decomposed into a screw component
and an edge component. The interaction of the applied stress field with the fractional edge component explains this peculiarity. This component of the stress does
not contribute to the movement of dislocations, and thereby induces many interesting features in BCC crystals. Especially in the last years many atomistic simulations
have been performed (e.g. Duesbery and Vitek (1998), Ito and Vitek (2001), Bassani et al. (2001), Duesbery et al. (2002), Vitek et al. (2004a), Groger

and Vitek (2005))


which support this effect. The hydrostatic pressure dependence of the flow stress (see
Spitzig (1979)), the strength differential effect or the tension-compression asymmetry
observed in BCC metals and intermetallic compounds (see Bassani (1994)) and critical conditions for the forming of shear bands and localization (e.g. Dao and Asaro
(1996)) typically result from this crystallographic non-Schmid effect, and its connection to non-associated plastic flow is well established by Racherla and Bassani (2007).
In the model presented next, the second intrinsic anisotropy effect will be included by
modifying the crystallographic flow rule which results in an update of the resolved
shear stress.

2 A finite strain BCC single crystal plasticity model and its experimental
12
identification

2.4 A BCC crystal plasticity model at material point level


2.4.1 Kinematics in crystal plasticity
In the classical crystal plasticity theory as developed by Lee (1969), Rice (1971), Hill
and Rice (1972) and Asaro and Rice (1977), the deformation gradient tensor is decomposed into an elastic part F e and a plastic part F p according to:
F = Fe F p

(2.1)

The tensor F p defines the stress-free intermediate configuration. In this configuration, resulting from plastic shearing along well-defined slip planes of the crystal lattice, the orientation of the crystal lattice is identical to the orientation in the reference
state (see Fig. 2.2). The tensor F e reflects the lattice deformation and local rigid body
rotations. The slip systems are labeled by a superscript , with = 1, 2..., ns where

m = Fe . m 0
T
n = Fe . n

F = Fe. Fp

Fe

n0
m 0

Fp

n0
m 0

Figure 2.2 / Multiplicative decomposition of the deformation gradient.

ns is the total number of slip systems. The vectors m0 and n0 denote the slip direction and the slip plane normal in the reference and intermediate configurations. In
the current state they are represented by m and n , respectively.
The crystallographic split of the plastic flow rate is given by
ns

Lp =

m0 n0

=1

with the individual slip rates on the slip systems.

(2.2)

2.4 A BCC crystal plasticity model at material point level

13

2.4.2 Constitutive model


The deformation is composed of an elastic contribution and a plastic contribution.
The elastic part is related to the stress, based on a hyper-elastic formulation while
the plastic part is determined by a physically based flow rule.
Elastic contribution
The second Piola-Kirchhoff stress tensor S is expressed in terms of the elastic GreenLagrange strain tensor E e , both relative to the intermediate state,
S = 4 C : E e and E e =

1
(C e I ) , C e = F eT F e
2

(2.3)

with C e the elastic right Cauchy-Green tensor and I the second order unity tensor.
The second Piola-Kirchhoff stress is the pull-back of the Kirchhoff stress tensor,
T
1
S = F
e Fe

(2.4)

where the Kirchhoff stress can be written in terms of the Cauchy stress using the
Jacobian according to

= J e with J e = det ( F e )

(2.5)

The fourth order tensor 4 C consists of the anisotropic elastic moduli.


The Schmid resolved shear stress is the projection of the Kirchhoff stress on the slip
systems, i.e.

= m n = m0 C e S n0

(2.6)

The slip systems of the {112} family in BCC crystals are given in Table 2.1, which is
the relevant set within the considered temperature range.
Plastic slip and hardening
In order to include the previously described crystallographic intrinsic properties, a
physical description of the slip law will be used instead of a classical phenomenological (power law) relation. Physically based slip laws have been formulated in
crystal plasticity models for materials with a symmetric planar slip dependency (e.g.
Kothari and Anand (1998)). The anisotropy in BCC crystals is included by adapting
the slip law accordingly.
The physical interpretation given hereafter, relies on the thermally activated dislocation kinetics. Plasticity occurs by dislocation motion on certain slip planes in an

2 A finite strain BCC single crystal plasticity model and its experimental
14
identification

energetically favorable direction. The actual flow stress is determined by the resistance to this dislocation motion. The motion is obstructed by short-range and longrange barriers. The short-range barriers in general are generated by the Peierls stress
(periodic resistance of the lattice) and the local forest of dislocations. The long-range
barriers originate from the elastic stress field due to grain boundaries, far field forests
of dislocations and other defects. The total resistance can be split accordingly
s = st + sa

(2.7)

where the short-range barriers are responsible for the first part st , referred to as the
thermal part since thermal activation is sufficient to overcome this resistance. The
athermal part of the resistance sa is related to the long-range barriers. Although
this contribution slightly decreases at higher temperatures (through a decrease of
the elastic moduli), this effect is negligible compared to the change of the thermal
resistance with varying temperature.
During their motion, dislocations are obstructed by a quasi-periodic short-range resistance. The Helmholtz free energy required to isothermally cross a barrier is denoted by F and the mechanical work of st can be written as W. The energy difference between these two,
G = F W

(2.8)

is the energy barrier to overcome by a dislocation through thermal activation. It is


well-known that the average dislocation velocity, v on slip system , can be estimated by,
v = l 0 exp {G /kT }

(2.9)

with l representing the distance between the barriers, 0 the attempt frequency, k
the Boltzmann constant and T the absolute temperature. The relation between the
slip rate and the average velocity is given by the Orowan relation, = bm v where
m and b represent the mobile dislocation density and Burgers vector, respectively.
Substituting the velocity expression (2.9) into the Orowan relation leads to the slip
law according to,

0
if ef f 0

=
(2.10)
G
0 exp { kT
} sign ( )
if 0 < ef f
where ef f = | | sa is the driving force for the dislocation motion and 0 =
bm l 0 is the reference strain rate, which is different for the different BCC slip plane
families since the distance between the barriers depends on the family. The energy
G to be supplied by the thermal fluctuations at constant temperature is calculated
as (see Kocks et al. (1975))
"
  p #q
e f f
G = G0 1
(2.11)
st

2.4 A BCC crystal plasticity model at material point level

15

where G0 is the activation free energy needed to overcome the obstacles without
the aid of an applied stress. The quantities p and q lie in the range 0 p 1
and 1 q 2, and in the numerical examples of this paper they are taken as 1.
The equations (2.10) and (2.11) constitute the slip law for materials that do not show
crystallographic asymmetry effects. The non-Schmid effects are included in equation
(2.10) by pursuing a similar strategy as introduced by Dao and Asaro (1993), where
the Schmid stress as defined by (2.6) is extended to account for the non-Schmid contribution:

n = + :

(2.12)

with the Kirchhoff stress and representing the tensor governing the non-Schmid
effects for slip system aligned with m , n and z = m n defined as,

= mm (m m) + nn (n n) + zz ( z z)+
mz (m z + z m) + nz (n z + z n)

(2.13)

In this framework, n enters the equations (2.10) and (2.11) via ef f = |n | sa .


The definition of the non-Schmid stress tensor and incorporation in the slip law is
phenomenological, and the physical meaning is not immediately trivial. The nonSchmid component is operative in the thermally activated process because of its effect on the fractional edge component of the screw dislocation cores.
For isothermal cases, the thermal part st of the slip resistance s is taken constant
and the athermal part of the slip resistance is evolving such that
s = sa = h | |

(2.14)

The hardening moduli h determine the rate of strain hardening on slip system
due to slip on slip system . This self and latent hardening are phenomenologically
described by (Asaro and Needleman (1985))
h = q h

(2.15)

where q and h are further detailed. As explained in the discussions in section


2.2, BCC metals show a {112} dominated slip pattern at room temperature. For the
{112}h111i slip system there are twelve different slip planes contributing to one slip
direction (see Table 2.1). This leads to the following definition of the q matrix (12
12),

1 qn . . . qn
qn 1 . . . qn

q = .. .. . .
(2.16)
.
. .
. ..
qn qn . . . 1

2 A finite strain BCC single crystal plasticity model and its experimental
16
identification

Fe

Fp

S, , n

Fc = Fe F p
R = F Fc

F e := F e + F e
Figure 2.3 / Schematic overview of the BCC crystal plasticity model.

where qn represents the ratio of the latent hardening with respect to the self hardening for non-coplanar slip systems.
Finally the specific form of the self hardening rate, which is motivated by Brown
et al. (1989), reads
a




sa
sa

h = h0 1 sign 1 ,
(2.17)
ss
ss

where h0 , sa and ss are the initial hardening rate, the actual athermal slip resistance
and the saturation value of the slip resistance, respectively. The exponent a is considered as a constant material parameter.
Implementation of the constitutive model

The implementation of the above model follows an incremental-iterative solution


procedure. The first step in this iteration is the initial estimate for the elastic part F e ,
resulting in a plastic part F p through (2.1). With the kinematics defined, the stress,
the Schmid and non-Schmid stresses, are calculated. From these, the slip rate on each
slip system is calculated by using the slip law (2.10). As this slip law is non-linear
in terms of the slip rates, a sublevel Newton-Raphson iteration process is adopted
to solve for the slip rates.The plastic part of the deformation gradient is obtained
from the calculated slip rates through a time integration scheme. An updated F p
is determined, and the associated deformation gradient is calculated according to

2.5 Modeling some intrinsic properties of BCC single crystals

17

F c = F e F p . Generally, the calculated F c and the imposed F will be different, which


results in a residual R. The linearization of the residual by computing the sensitivity with respect to F e leads to an update F e of the elastic part of the deformation
gradient. The elastic deformation gradient is updated and the process is repeated
until convergence is achieved. The procedure is performed for all time steps, which
results in a full history of stress and slip evolution. The main steps of the procedure
are summarized in Fig. 2.3.

2.5 Modeling some intrinsic properties of BCC single crystals


In this section, orientation and temperature dependence of BCC materials are simulated applying the presented crystal plasticity framework, and results are compared
with the single crystal experiments. To this purpose, material parameters are determined using a proper identification procedure. This direct confrontation has, to
the best of our knowledge, not been done before for single crystals. Although, polycrystal BCC crystal plasticity simulations and validations have been conducted in
the literature (e.g. Kothari and Anand (1998), Liao et al. (1998), Lee et al. (1999), Xie
et al. (2004), Ganapathysubramanian and Zabaras (2005)) and non-Schmid effects
have been incorporated into constitutive models (e.g. Qin and Bassani (1992)), the
results cannot directly be exploited in the context of actual model.

2.5.1 Orientation dependence


The hardening curves, work hardening rate, temperature and rate sensitivity, activity
of slip planes, CRSS and the flow stress of BCC metals all depend on the orientation
of the crystal. As emphasized before, two crucial important physical aspects controlling this pronounced orientation dependence are the so-called slip asymmetry (or
twinning/anti-twinning asymmetry) and the non-planar spreading of screw dislocation cores, which have been included in the model.
The twinning/anti-twinning asymmetry manifests itself on the {112} slip planes. On
these slip planes, the slip resistance in the anti-twinning direction is higher than in
the twinning direction, an effect which is difficult to quantify in experimental tests.
Guiu (1969) experimentally observed the asymmetry for Mo single crystals at different temperatures under direct shear. He concluded that the CRSS is roughly 1.5
times larger for the slip systems on the anti-twinning planes of {112}h111i compared
to the slip systems on the twining planes.
Due to the non-planar spreading of the screw dislocation cores, the non-glide component of the applied stress affects the dislocation core and hence, the sense of the

2 A finite strain BCC single crystal plasticity model and its experimental
18
identification

applied stress affects the yielding of a crystal. The associated non-Schmid parameters
will be identified together with the other parameters in the model.

2.5.2 Example: -Fe single crystal


In the first example (Fig. 2.4) the orientation dependence of -Fe single crystals under
uniaxial tension is examined. Here, and in the following examples the lattice vector
was aligned with the tensile direction in the undeformed configuration and the presented framework automatically accounts for the lattice rotations. The experimental
curves (Keh (1964)) were reproduced from the reported shear stress-strain curves
), (211
) and
which were initially determined from tensile data for the planes of (1 12
) with the [001], [011] and [111
] orientations respectively in the [111] direction.
(211
The experiments and simulations are performed at a strain rate of 3.3 104 s1 . The
180
160
140

[111]
[001]

Stress (MPa)

120
100
80
60

[011]

40
20
0

3
Strain %

] oriented Figure 2.4 / Tensile orientation dependence of [001], [011] and [111
Fe single crystals. Solid lines are the simulation results and dashed lines represent
experiments.

results show the pronounced influence of the orientation on the yielding and hardening behavior of the crystal. The orientations selected in the example constitute the
] directions. It is a wellcorners of a unit triangle mapping the [001], [011] and [111
established fact that at the corners and at the edges of the unit triangle multislip is
observed, while the deformation starts with single slip for directions mapped inside
the triangle. For these orientations the initial rate of work hardening is relatively
high and decreases rapidly with ongoing deformation.

2.5 Modeling some intrinsic properties of BCC single crystals

19

Material parameters have been identified using a least-square optimization procedure which minimizes an objective function that equals the sum of squares of the
differences between experimental and simulation results. Most of the parameters are
presented in Table 2.2. The remaining parameters are C11 = 236GPa, C12 = 134GPa,
Table 2.2 / Material parameters for -Fe single crystals

Initial hardening rate


Saturation value of slip resistance
Hardening rate exponent
Thermal slip resistance
Athermal slip resistance (atwin sense)
Athermal slip resistance (twin sense)
Non-Schmid parameter
Non-Schmid parameter
Non-Schmid parameter
Reference strain rate
Activation free energy

h0
ss
a
st
s a0
s a0
mm
nn
zz
0
G0

697.88 MPa
132.10 MPa
1.5
13.91 MPa
9.59 MPa
5.75 MPa
0.0544
-0.0293
-0.0267
1.07 106 s1
2.95 1018 J

C44 = 119GPa (Adams et al. (2006)), qn = 1.4, k = 1.3807 1023 J/K, while mz and
nz are taken zero..

2.5.3 Example: molybdenum single crystal


In the second example, the orientation dependence of molybdenum single crystals
under uniaxial tension is analyzed. The experimental curves for the [010], [101] and
[111] orientations are taken from the work of Irwin et al. (1974). They performed two
sets of experiments at 293K and 77K at a strain rate of 6 105 s1 . The results are
compared for the 293K case and presented in Fig. 2.5. The material parameters that
have been identified using the same least-square minimization process are presented
in Table 2.3. The remaining parameters are C11 = 469GPa, C12 = 167.6GPa, C44 =
106.8GPa (Bolef and Klerk (1962)), qn = 1.4, k = 1.3807 1023 J/K.

2.5.4 Temperature dependence


BCC metals exhibit a different mechanical response compared to FCC metals, in particular in the presence of temperature changes. The dependence of the flow stress
on the temperature, has been studied thoroughly in the literature and illustrated for
many BCC single crystals such as Mo (e.g. Hollang et al. (1997)), Nb (e.g. Ackermann

2 A finite strain BCC single crystal plasticity model and its experimental
20
identification

40

[111]

35

Stress (MPa)

30
25
20

[101]
15
10

[010]
5
0

0.2

0.4

0.6
0.8
Strain %

1.2

1.4

Figure 2.5 / Orientation dependence of [010], [101] and [111] oriented molybdenum single crystals. Solid lines are the simulation results and dashed lines represent
experiments.

Table 2.3 / Material parameters for Mo single crystals

Initial hardening rate


Saturation value of slip resistance
Hardening rate exponent
Thermal slip resistance
Athermal slip resistance (atwin sense)
Athermal slip resistance (twin sense)
Non-Schmid parameter
Non-Schmid parameter
Non-Schmid parameter
Reference strain rate
Activation free energy

h0
ss
a
st
s a0
s a0
mm
nn
zz
0
G0

251.37 MPa
76.90 MPa
1.05
11.89 MPa
7.23 MPa
4.34 MPa
-0.0528
0.0896
-0.0369
1.4 107 s1
0.1554 1018 J

et al. (1983)), Ta (e.g. Werner (1987)), -Fe (e.g. Brunner and Diehl (1987), Brunner
and Diehl (1997)). Below the so-called knee temperature, where the flow stress is
controlled by the mobility of screw dislocations, the critical shear stress of BCC metals rises progressively with decreasing temperature (Seeger (1981)). The flow stress
in FCC metals on the contrary, only shows a moderate increase when the temperature is lowered below room temperature. Most of the work reported on BCC metals
concentrates on the behavior of pure single crystals in order to eliminate secondary

2.5 Modeling some intrinsic properties of BCC single crystals

21

effects induced by interstitially dissolved foreign atoms. The purification process of


single crystals requires great effort, nevertheless, when tested usually impurities can
be identified.
The variation of the flow stress of a BCC metal under a temperature change has already been presented in Fig. 2.1 for Mo single crystals. For other BCC metals, the
behavior is qualitatively similar but quantitatively different. For the response below
the knee temperature TK , different regimes should be distinguished. Especially the
interval TK /2 < T < TK was analyzed by Seeger (1981) for which the rate and temperature dependence was explained through the formation of kink pairs in screw
dislocations without invoking impurity effects.
Because of the non-uniformity of the temperature dependence of the flow stress,
Seeger (1981) proposed different formulations for different temperature regimes. A
high-temperaturelow stress regime, an intermediate (diffusion controlled) regime
and a high-stress regime are distinguished. In the present paper, the flow is controlled by the slip rate equation (2.10), which roughly corresponds to the third
regime, where only kink pairs controlling dislocation movement in the direction of
applied stress are taken into account.

50
45
40

Stress (MPa)

35
30
25
20
15

77 K sim.
77 K exp.
113 K sim.
113 K exp.
175 K sim.
175 K exp.

10
5
0

0.2

0.4

0.6

0.8
1
Strain %

1.2

1.4

1.6

1.8

Figure 2.6 / Temperature dependence of [001] oriented niobium single crystals.

In Fig. 2.6, the tensile stress-strain curves for a [001] oriented Nb single crystal are
presented at three different temperature levels, at a strain rate of 1.3 104 s1 . The
experimental data is taken from Duesbery and Foxall (1969). The true stress and true
)[111
] primary
strain curves are reproduced from the shear values reported for (112

2 A finite strain BCC single crystal plasticity model and its experimental
22
identification

slip. The Schmid factor for the [001] orientation equals 0.471. The experiment and the
crystal plasticity simulations show an adequate agreement. The material parameters
for this material are: C11 = 250GPa, C12 = 135GPa, C44 = 30GPa (Carroll (1965)),
0 = 1.2 107 s1 , qn = 1.4, G0 = 1.12 1018 J, k = 1.3807 1023 J/K, a = 1.2,
s a0 = 4MPa (atwin sense) and s a0 = 2.4MPa (twin sense). Due to the lack of experimental evidence the identification of the non-Schmid parameters is disregarded and
the effect is excluded here. The variation of most parameters with the temperature is
negligible, with the exception of the parameters presented in Table 2.4.
Table 2.4 / Material parameters for Nb single crystals at three different temperatures

Temperature
h0
ss
st

77 K
1500 MPa
21.7 MPa
15.3 MPa

113 K
922 MPa
19.9 MPa
8.06 MPa

175 K
800 MPa
13.42 MPa
3.18 MPa

2.6 Summary and Conclusion


In the present paper, a crystal plasticity model has been proposed and implemented,
revealing the unique characteristics of BCC single crystals. A comprehensive summary of the intrinsic properties of these materials has been presented, including recent insights in the activation of different slip systems, violation of Schmids law,
temperature and orientation dependence of the flow stress and resulting stress-strain
curves. The parameters in the model are determined using a proper parameter identification process, relying on a least-square minimization procedure on the differences between the numerical and experimental uniaxial stress-strain responses.
Published results on the operational slip mechanisms in BCC crystals and the extended crystal plasticity models, reflect considerable contradictions. Most studies
take into account {110}, {112} and {123} type of slip planes without confronting
the model with the expected temperature and orientation dependence of the crystal.
Following Seeger (2001), this work uses {112} planes at moderate temperatures and
{110} planes at low temperatures. Contrary to most of the crystal plasticity elaborations, {123} type of slip planes are not taken into account. Cross slipping phenomena
and non-planar spreading of screw dislocation cores are implicitly incorporated in a
phenomenological manner.
The applied slip law plays an essential role in the present model since all the intrinsic

2.6 Summary and Conclusion

23

characteristics result from the actual formulation of the slip rate equation. Additional
to the pre-mentioned existing crystal plasticity frameworks, in this contribution the
non-Schmid behavior is introduced in the slip law by modifying the effective shear
stress, where the non-Schmid contribution represents the dislocation cores spreading
in a non-planar manner. Actually, the emphasis was put on this intrinsic anisotropy
effect, even though other anisotropy effects such as texture development and dislocation sub-structure evolution may be included in the model as well.
The necessity for this research results from the fact that there is a lack of published
works, that confront recent BCC single crystal plasticity models to single crystal experimental data that reveals the intrinsic orientation and temperature dependence of
these crystals.

24

Chapter three

A composite dislocation cell model to


describe strain path change effects in
BCC metals1

Abstract / Sheet metal forming processes are within the core of many modern manufacturing technologies, as applied in e.g. automotive and packaging industries. Initially flat sheet material is forced to transform plastically into a three dimensional shape
through complex loading modes. Deviation from a proportional strain path is associated
with hardening or softening of the material due to the induced plastic anisotropy resulting from the prior deformation. The main cause of these transient anisotropic effects at
moderate strains is attributed to the evolving underlying dislocation microstructures. In
this paper, a composite dislocation cell model, which explicitly describes the dislocation
structure evolution, is combined with a BCC crystal plasticity framework to bridge the
microstructure evolution and its macroscopic anisotropic effects. Monotonic and multistage loading simulations are conducted for a single crystal and polycrystal BCC metal,
and obtained macroscopic results and dislocation substructure evolution are compared
qualitatively with published experimental observations.

3.1 Introduction
For each car, the automotive industry manufactures more than 500 parts by multistage forming operations, involving complex deformation paths. Deviation from
a proportional strain path is commonly associated with a change in the hardening
(or softening) behavior of the material. In order to achieve a first-time-right design,
modern predictive tools relying on the finite element method are commonly used
nowadays. The anisotropy induced by complex deformation paths, which may lead
1

This chapter is reproduced from Yalcinkaya et al. (2009)

25

3 A composite dislocation cell model to describe strain path change effects


26
in BCC metals

to premature failure (e.g. Sang and Lloyd (1979)), is crucial in this sense and should
be included in the constitutive models used in the analysis.
The overall plastic anisotropy in BCC metals, induced by the imposed deformation,
originates from different sources at different length scales. Slip asymmetry and intrinsic anisotropy effects caused by the non-planar spreading of screw dislocation
cores are active at the micro level (e.g. Bassani et al. (2001), Duesbery and Vitek
(1998), Ito and Vitek (2001), Yalcinkaya et al. (2008)) whereas the development of dislocation substructures is relevant at the meso level (e.g. Rauch and Schmitt (1989),
Wagoner and Laukonis (1983), Rao and Laukonis (1983), Wilson and Bate (1994),
Gardey et al. (2005)). At the macro level, the texture development of polycrystalline
metal is contributing dominantly (e.g. Bacroix et al. (1994), Bacroix and Hu (1995),
Nesterova et al. (2001)). Upon switching strain paths, the intrinsic properties obviously have a substantial effect on the observed anisotropy due to changes in the
dislocation activity. However, the evolution of dislocation microstructures has been
recognized as a main driver triggering the observed anisotropic material behavior. In
a recent report Li et al. (2006) commented on the strong anisotropy, i.e., larger than
expected from the texture, induced by the dislocation structure in IF steel increasing
with the rolling prestrain. The prediction of dislocation microstructures within the
individual dislocation descriptions and continuum theories has been a challenging
subject in the last decades in the material science community (see Groma (1997) for an
overview). While transmission electron microscopy (TEM) observations have been a
powerful tool to understand their origin and to derive the physical parameters that
govern their evolution (e.g. Fernandes and Schmitt (1983)), discrete dislocation models and atomistic considerations improved the understanding of the formation and
the evolution of dislocation microstructures and the related plastic anisotropy. However, only a limited number of micromechanical modeling approaches have been addressing the anisotropy induced by evolving dislocation cells with a crystal plasticity
framework.
Among the attempts to develop plastic anisotropy models that incorporate the microstructure evolution for complex deformation histories, the most remarkable one is
the constitutive model proposed by Teodosiu and Hu (1995). This phenomenological
model uses the Hill criterion for the onset of yielding while the hardening is associated with the dislocation structures. The polarity of dislocation walls, the back-stress
and the strength of the dislocation structure are accounted for by internal variables.
Recently, Wang et al. (2008) presented an improvement of this model especially concentrating on continuous loading path changes from uniaxial tension to simple shear
without unloading the material.
Another attempt to describe the occurring phenomena is presented by Peeters et al.
(2000) dealing with a polycrystal plasticity model that incorporates more details of

27

3.1 Introduction

the microstructure evolution at the grain scale, where cell boundary dislocation densities, cell block boundary dislocation densities, and directionally movable dislocation densities are taken as internal variables. This model attributes a major part of the
strain path change effects to the evolution of cell block boundaries and polarization
of these structures. Additional to above mentioned models, Hoc and Forest (2001),
Mollica et al. (2001) and Tarigopula et al. (2008) presented some other approaches
dealing with the anisotropic strain path change effects. In the present paper the con-

w
w

r
Figure 3.1 / Composite representation of the cell structure.

centration is focused on a crystal plasticity model that incorporates the evolution of


dislocation cell structures. As originally introduced by Mughrabi (1987), a cell structure can be idealized as a two-component material, distinguishing cell walls and cell
interiors. It is characterized by the wall thickness w, the cell size r (see Fig. 3.1), the
dislocation densities in the cell walls w and the cell interiors c . The macroscopic
anisotropy effects are obtained by the evolution of these internal variables during
monotonic deformation and multi-stage loading processes. Inside the cell structure,
a BCC crystal plasticity framework (Yalcinkaya et al. (2008)) is incorporated, which
goes beyond the developments of Viatkina et al. (2003) for FCC metals in which a
classical von-Mises plasticity model was used. From this perspective, it is the first
example that incorporates a physically motivated constitutive model into the evolution of dislocation substructures for BCC metals, in order to model the anisotropy
due to strain path changes.
The plan of this paper is as follows. Section 2 discusses the evolution of dislocation
substructures under monotonic and multi-stage deformations. Next, in section 3 the
formulation of the BCC crystal plasticity framework is summarized. Section 4 handles the incorporation of the dislocation cell evolution model into the crystal plasticity framework, along with a summary of the numerical implementation. Further, in
section 5 computational results of single crystal and polycrystal tests are presented
on the basis of which the crystal anisotropy is distinguished from the dislocation

3 A composite dislocation cell model to describe strain path change effects


28
in BCC metals

cell anisotropy. The accordance of the results with respect to published experimental
results is discussed. Finally, concluding remarks are given in section 6.
Cartesian tensors and associated tensor products will be used throughout this paper,
making use of a Cartesian vector basis {e1 e2 e3 }. Using the Einstein summation rule
for repeated indices, the following conventions are used in the notations of vectors,
tensors, related products and crystallography:
scalars a
vectors a = ai ei
second-order tensors A = Ai j ei e j
fourth-order tensors 4 A = Ai jkl ei e j ek el
C = a b = ai b j ei e j
C = A B = Ai j B jk ei ek
C = 4 A : B = Ai jkl Blk ei e j
crystallographic direction, family [uvw], huvwi
crystallographic plane, family (hkl ),{hkl }
slip system, family (hkl )[uvw], {hkl } huvwi

3.2 Dislocation substructure evolution


Dislocation substructuring is characterized by the clustering of dislocations after a
certain amount of plastic deformation, where an initially statistically homogeneous
distribution of dislocations develops towards a dislocation pattern with high density
dislocation walls enveloping low density dislocation areas. This self-organization of
the microstructure in the grains is often referred to as the low-energy, steady state
configuration of dislocations (Kuhlmann-Wilsdorf (1989)). TEM analyses (e.g. Keh
et al. (1963)) revealed that for deformations larger than 3-4 % a well-developed dislocation cell structure forms in steel at ambient temperature. Further deformation
renders a polarized structure with dislocation sheets or cell block boundaries, which
envelope a number of dislocation cells. These structures are the result of the interactions between dislocations gliding on the most active slip planes and the secondary dislocations (Teodosiu (1992)). However, the occurrence of the dislocation

29

3.2 Dislocation substructure evolution

sheets is not always manifest and sometimes the microstructure is partitioned by ordinary cell boundaries having no particular crystallographic or macroscopic orientation (Hansen and Huang (1997)). For that reason, depending on the grain orientation
and the strain direction, either parallel dislocation walls or more equiaxed closed
cells are observed (e.g. Rauch and Schmitt (1989)) in low carbon steels. Besides, different materials end up in different type of microstructures. It is neither experimentally nor computationally an easy task to identify the type of evolving dislocation
microstructure, yet formation of dislocation cells is mostly observed. Thereof, this
paper concentrates on the formation and evolution of these dislocation cell structures.
MONOTONIC

CROSS TEST

STRESS REVERSAL

Figure 3.2 / Schematic evolution of a dislocation cell structure under strain path
change.

As discussed above, dislocation cell structures develop upon plastic strain in most
metals, and evolve in a distinct way depending on the applied strain path. The main
features are visualized in Fig. 3.2. Under monotonic deformation a dislocation cell
structure appears and evolves towards a decreasing cell size r, and wall thickness w
accompanied by an increasing dislocation density in the cell walls w (e.g. Fernandes

3 A composite dislocation cell model to describe strain path change effects


30
in BCC metals

and Schmitt (1983)). After a strain path change, the developed cell structure adjusts
to the new loading and the dislocation microstructure induced by the prestrain becomes unstable. It is disrupted and dissolved, and a new dislocation structure typical
of the new strain path forms (Barlat et al. (2003)). The characteristic features of the
initial cell structure disappear as the deformation proceeds in the new direction. Unfortunately, there is no clear interpretation of what is occurring with the dislocation
microstructure during the adaptation nor is there a unique terminology to describe
this evolution. Here we distinguish between two different scenarios; dissolution of
cells as in the cross test and disruption as occurring under reversed loading. There
appears to be no consistency in the literature in the use of the dissolution, disruption
and disintegration of dislocation cells, and most of the time any cell evolution after
a strain path change is described as a dissolution process (e.g. Rauch and Schmitt
(1989), Rauch (1992), Rauch (1991) Gardey et al. (2005), Rao and Laukonis (1983)).
Indeed, both dissolved and disrupted structures appear as disorganized structures
with a higher degree of homogeneity compared to the state before the strain path
change. Nevertheless, there are indications that there is a morphological difference
between the two microstructure evolution scenarios mentioned above (e.g. Gardey
et al. (2005)) in correspondence with the difference between the driving forces and
their physical origins.

Figure 3.3 / Left: Experimental results for mild steel DC06 subjected to monotonic
simple shear and simple shear followed by load reversal (10 % and 30 %) [Bouvier
et al. (2006)] (reversed loading) Right: Experimental results for IF-steel subjected to
monotonic simple shear and tensile tests (10 % and 20 %) followed by shear [Peeters
et al. (2000)] (orthogonal loading).

In the example shown in Fig. 3.2, two different strain path changes have been considered where the two types of evolution phenomena can be distinguished. A cross test,
e.g. tension followed by simple shear or a tension test followed by tension in a dif-

3.3 Computational model

31

ferent direction, reveals progressive cell evolution (e.g. Rao and Laukonis (1983)). It
has been observed that after a strain path change cell walls become thicker while the
dislocation density in the walls becomes smaller (e.g. Schmitt et al. (1991)). Hence,
the dislocation distribution is more uniform and the cell structure is less organized.
This evolution process can cause partial or complete dissolution of the existing cell
structure, while concurrently a new cell structure develops with a morphology related to the new loading direction. The cell structure evolution resulting from a
stress reversal has received more attention in the context of the analysis of the wellknown Bauschinger effect (e.g. Rauch (1991)). The evolution of the cell structure
under stress reversal can be characterized by the disruption of cell walls (e.g. Viatkina (2005), Christodoulou et al. (1986)). The thickness of the cell walls does not
change significantly, however, the walls tend to disconnect. Experimental observations (e.g. Christodoulou et al. (1986)) also report a strong flux of dislocations from
the walls to the cell interiors, decreasing the wall dislocation density and increasing
the density in the cell interiors. Accordingly, the descriptive modeling of the dislocation distribution relies on an increase of cell size and a dislocation redistribution
(Viatkina (2005)). With ongoing deformation cell walls reappear and a new cell structure originates.
Upon sustained loading after a strain path change, the microstructure always evolves
such that transient effects disappear and the macroscopic stress-strain curve saturates to the monotonic deformation curve (see Fig. 3.3).

3.3 Computational model


The constitutive behavior of each composite constituent (cell walls or cell interiors)
is modeled in a finite strain crystal plasticity framework with plastic slip governed
by the thermally activated motion of dislocations. The kinematics starts with the
multiplicative decomposition of the deformation gradient tensor into an elastic and
a plastic part of each component, as developed by Lee (1969), Rice (1971), Hill and
Rice (1972) in the classical plasticity theory,

F i = F ie F ip ,

(3.1)

where the superscript i indicates the specific component (w: wall, c: cell) and tensor
F ip defines the stress-free intermediate configuration. In this configuration, resulting
from plastic shearing along well-defined slip planes of the crystal lattice, the orientation of the slip systems is unaltered. The tensor F ie reflects the lattice deformation
and local rigid body rotations. The slip systems are labeled by a superscript , with
= 1, 2..., ns where ns is the total number of slip systems. The vectors m0 ,i and n0 ,i

3 A composite dislocation cell model to describe strain path change effects


32
in BCC metals

denote the slip direction and the slip plane normal in the reference and intermediate
configurations. In the current state they are represented by m ,i and n ,i , respectively.
i
The crystallographic split of the plastic flow rate Lip = F p F ip 1 is given by

Lip =

ns

,im0 ,i n0 ,i,

(3.2)

=1

with ,i the individual slip rate on the slip system .


The second Piola-Kirchhoff stress tensor Si is expressed in terms of the elastic GreenLagrange strain tensor E ie , both relative to the intermediate state,
Si = 4 C : E ie

with

E ie =

1 iT i
( F F e I ),
2 e

(3.3)

with I the second order unity tensor and 4 C the fourth order tensor consisting of
elastic moduli.
From the second Piola-Kirchhoff stress the Kirchhoff stress in the current configuration can be determined by a push-forward operation,
T

i = F ie Si F ie .

(3.4)

From the Kirchhoff stress the Cauchy stress can be derived according to,

i =

1 i

Jei

with

Jei = det( F ie ).

(3.5)

The Schmid resolved shear stress is the projection of the Kirchhoff stress on the slip
systems, i.e.

,i = m ,i i n ,i = m0 ,i C ie Si n0 ,i

with

C ie = F ie F ie ,

(3.6)

which is the driving force for the dislocation movement on a certain slip system .
There has been various discussions and contradictions considering the active slip
systems of BCC crystals, and recent studies shows that the slip system activation
is highly temperature dependent (see Yalcinkaya et al. (2008)). At room temperature the {112} slip system family is dominantly active. The effect of the non-Schmid
stresses on the non-planar screw dislocation cores which contributes to the orientation dependence of BCC crystals at single crystal level can be taken into account
as an additional contribution to the driving stress in equation (3.6) (Yalcinkaya et al.
(2008)), however, this contribution affects the initial anisotropy of these metals rather
than the transient effects observed during the strain path changes (see Fig. 3.3). Including this effect would increase the material parameters while it does not contribute to the aim of this paper. Hence it was decided not to account for this effect
here.

3.3 Computational model

33

The motion of dislocations is obstructed by thermal and a-thermal barriers which


are caused by the dislocation interactions upon flow, the elastic stress field due to
other dislocations and grain boundaries. Hence the slip resistance distinguishably
originates from a thermal part st ,i and an a-thermal part sa ,i. For the slip rates the
following slip law is adopted (see Yalcinkaya et al. (2008)),
(
!#)
"
ef,if
G0
,i
,i
= 0 exp
sign ( ,i ),
(3.7)
1

,i
kT
st
where ef,if = | ,i | sa ,i is the effective driving stress on the slip systems, G0 is the activation free energy, k is Boltzmanns constant and T is the absolute temperature, 0 ,i
is a reference strain rate. For isothermal cases, the thermal part st ,i of the slip resistance is taken constant and the a-thermal slip resistance is related to the dislocation
densities on all slip systems through,
s
sa ,i

ns

= Gb

Au |u,i |,

(3.8)

u=1

where G is the shear modulus, b is the magnitude of the Burgers vector, A u are
the interaction coefficients between the slip systems and u, and u,i the dislocation
density on the slip system u of component i. The dislocation interaction coefficients
of the matrix A u depend on the type of interaction between dislocations on different
slip systems (e.g. Franciosi and Zaoui (1982), Queyreau et al. (2008)). Because of the
lack of data on {112} slip systems, only the interactions between the dislocations
belonging to the same slip system, i.e. = u, and different slip systems, i.e. 6= u,
will be distinguished for A u .
The macroscopic mechanical response of the composite model is obtained by applying a Taylor averaging assumption where the deformation in each component is
assumed to be equal to the macroscopic deformation and where the rule of mixtures
gives the macroscopic stress from the local stresses in each component according to,

= f w + (1 f ) c .

(3.9)

In this equation f represents the actual volume fraction of the cell walls, expressed
in terms of the microstructural morphology parameters w and r according to:
 w 2  w 3
Vw
w
f =
= 3 3
+
,
V
r
r
r

(3.10)

where V and V w are the volumes of the entire composite and the wall component,
respectively.
Experimental studies (e.g. Fernandes and Schmitt (1983)) suggest that the wall thickness w, the cell size r (see Fig. 3.1), the dislocation densities in the cells ,c, and

3 A composite dislocation cell model to describe strain path change effects


34
in BCC metals

the walls ,w evolve with increasing applied strain. Moreover, these quantities are
dependent on the deformation history, and therefore they are taken into account as
internal variables in this framework. Corresponding evolution equations are to be
formulated that describe the cell structure development during monotonic loading
as well as complex strain path histories. This is done in the following section, where
the incorporation of a dislocation cell structure evolution model into the crystal plasticity framework is presented.

3.4 Modeling of microstructure evolution


In order to give a clear understanding of the model, three distinct types of loading
cases are considered, namely monotonic loading, orthogonal loading and reverse
loading. The purpose of the model consists in unifying these cases by capturing effects under continuous combinations of these deformations, through a single set of
evolution equations. It is assumed that the cell orientation is dictated by the loading,
and that there are always enough slip systems to accommodate that cell, independently of the crystal orientation.

3.4.1 Monotonic deformation


The evolution of a two-phase dislocation cell structure has been schematized in Fig.
3.2. During monotonic deformation, the cell size r and the wall thickness w decrease, yielding a decrease of the length scales of the spatial dislocation patterns that
is inversely proportional to the flow stress, often referred to as the law of similitude
(Kuhlmann-Wilsdorf (1962)). Experimental observations (e.g. Mughrabi et al. (1986))
suggest that the dislocation density inside the cell interiors does not change significantly. Accordingly, a constant dislocation density ,c in the cell interior component
is assumed here. The derivation of the evolution of the dislocation density in the
walls ,w departs from the frequently used balance between the multiplication of
mobile dislocations and annihilation events,
,w


1  ,w
,c ,w
,w
,w
=
I R
f
| | +
b
f

(3.11)

where R is the recovery length and I is a dislocation multiplication parameter. The


last term in the equation accounts for the change in volume occupied by the wall
component. The initial state of the composite is modeled as if the wall component
is occupying the entire volume; the initial value of its dislocation densities is determined by the value 0 of the initial uniform distribution. This in fact correctly
represents the case when no dislocation pattern is present.

3.4 Modeling of microstructure evolution

35

The following empirical relation between the cell size r and the flow stress y is
suggested in the literature (e.g. Barker et al. (1989), Mughrabi (1987)),

y =

CGb
,
rm

(3.12)

which is consistent with the experimental observations of Fernandes and Schmitt


(1983) and commonly used theoretical investigations (e.g. Mughrabi (1987)). The
parameter C, is a material constant and the exponent m is generally close to 1 for
cell structures. In the present framework, equation (3.12) is rewritten in terms of slip
variables at the slip system level instead of the continuum level yield stress y . This is
done by using evolving a-thermal slip resistances on the active slip systems, i.e. r
CGb/ y CGb/ s a , whereby the parameter C accounts for the scaling between
the two levels as well. The cell size evolution is approximated by incorporating the
rule of mixtures for the different components of the composite,
r=

,w
s a

CGb
+ (1 f ) sa ,c

(3.13)

The evolution of the wall thickness w is adopted from Viatkina et al. (2003) and assumed to be governed by an effective plastic strain rate measure | | according
to
w = km (win f w) | |

with

| | = f | ,w | + (1 f ) | ,c|.

(3.14)

In this evolution law a decrease of the wall thickness with a saturation factor km is
incorporated, with a final saturation value equal to win f , which is consistent with
experimental observations (e.g. Fernandes and Schmitt (1983)).
The implementation of the model presented above follows an incremental-iterative
solution procedure, which is applied for each of the composite components with the
same imposed deformation (Taylor approach). The first step in this procedure is the
initial estimate for the elastic part F ie , resulting in an estimate for the plastic part F ip
through (3.1). With the kinematics defined, both the stress and the Schmid stress is
calculated. These values together with the slip resistance (3.8) (which is calculated
from dislocation density evolution (3.11)) enter the slip law (3.7) resulting in the slip
rates on each slip system. The updated plastic part of the deformation gradient is
obtained from the calculated slip rates through a time integration scheme. Generally, the calculated and the imposed deformation will be different, which results in a
residual. Iteration on the residual leads to updated values of variables including F ip
and F ie . With the current values of r and w the volume fraction f is calculated with
(3.10), which is used to determine the macroscopic stress (3.9). The procedure is repeated for all time steps, which results in the entire history of stress, slip and internal
variable evolution.

3 A composite dislocation cell model to describe strain path change effects


36
in BCC metals

3.4.2 Orthogonal change of deformation


An orthogonal change of the deformation path leads to dissolution of dislocation
cell walls, which is captured through an increase of the wall thickness. In the limit
the wall occupies the whole material where w becomes equal to r, representing a
full recovery of a uniform dislocation configuration (i.e. no cells present). This limit
case is rarely observed in practice. After the dissolution process, new cells originate accommodating the new loading direction. Next, the cell size and dislocation
density are considered to evolve in the same way as for monotonic deformation, i.e.
the dislocation density increases in the walls and remains constant inside the cell.
More experimental evidence is needed to improve further on this phenomenological
relation.
The dissolution process is leading to a transient increase of the cell wall thickness
driven by the overall slip rate as given by
w = kd (r w) | |

(3.15)

where kd is the dissolution factor. The saturation value of the wall thickness logically
equals the cell size r corresponding to a complete dissolution of the cell. When both
the dissolution and the redevelopment processes are taken into account, the wall
evolution becomes
w = pkd (r w) | | + (1 p)km (win f w) | |,

(3.16)

where the first contribution on the right-hand side reflects equation (3.15) accounting
for the effect of the loading in the new direction, i.e. the dissolution process. The
second contribution on the right-hand side of (3.16) represents the development of
the wall structure according to equation (3.14). The transition parameter p, to be
specified in the following, defines the relative contribution of the dissolution process
in the evolution of w. It is characterized by taking into account: (i) the dissolution
process depends only on the angle between successive deformation paths and (ii)
the dissolution effect disappears as the deformation proceeds in the new direction.
In this context, the following expression for p is taken:




,
(3.17)
p = (1 ||) exp B | | | pre |

where is a scalar measure that identifies the strain path change and |pre | indicates the accumulated plastic deformation prior to the strain path change and B is a
material parameter. To characterize the strain path change measure , a commonly
used definition is adopted here:

L p1 : L p2
,
| L p1 || L p2 |

(3.18)

3.4 Modeling of microstructure evolution

37

where L p1 and L p2 are the macroscopic plastic velocity gradient tensors prior to and
after the strain path change. Here = 1 refers to monotonic deformation, = 0 to a
cross test, and = 1 to a reverse test.
Equations (3.16) and (3.17) describe the evolution of the wall thickness during the
whole deformation process. During monotonic deformation where = 1, equation (3.16) reduces to equation (3.14) describing cell wall thinning. After a strain
path change, the dissolution is initiated with an intensity proportional to (1 ||).
The dislocation walls start widening, governed by the competition between the new
structure development and the old structure dissolution. As the deformation proceeds in the new direction, the dissolution process fades out and accordingly p approaches 0 due to (3.17) and the wall thickness tends to decrease again: a new dislocation structure is developing.

3.4.3 Reverse deformation


The cell structure degeneration after a stress reversal, the so-called cell disruption,
is modeled by a temporary increase of the cell size. As already discussed in section
3.2, the thickness of the cell walls does not change significantly due to stress reversal.
Therefore, the other parameters, i.e. the wall thickness and the dislocation density,
are assumed to evolve in a similar way as under monotonic deformation. The disruption of cells is observed to be a rapid process in which the size of the cells rapidly
increases after a stress reversal and then slowly decreases (e.g. Viatkina et al. (2003),
Christodoulou et al. (1986)). As the deformation proceeds in the opposite direction
the cell size recovers to the level corresponding to the monotonic strain path (3.13).
In order to model this temporary increase of the cell size, an additional (transient)
term is incorporated in equation (3.13) for the cell size evolution:
r=

,w
s a

CGb
+ A exp [kc ( | | |pre |)],
+ (1 f ) sa ,c

(3.19)

where A defines the degree of disruption and kc is a constant reflecting the recovery
speed (governed by slip). The entire term in the right-hand side (added to (3.13))
determines the immediate increase of the cell size. This effect vanishes with ongoing
deformation depending on the value of the parameter kc . As soon as the disruption
contribution gradually disappears, the cell size again follows the evolution as given
for the monotonic deformation case. To reflect the fact that the cell disruption is
triggered by a stress reversal, the coefficient A depends on the strain path change
through

a||
if < 0
A=
(3.20)
0
if 0

3 A composite dislocation cell model to describe strain path change effects


38
in BCC metals

where a is a fitting parameter. Consequently, the Bauschinger test triggers the highest disruption. For more complex strain path changes with negative values of both
equations (3.16) and (3.19) have non-zero strain path change contributions (p > 0
and A > 0), describing a process with coexisting dissolution and disruption of cells.
When p = 0 and A = 0, the model describes the evolution under monotonic deformation. In this way the resulting system of equations effectively unifies different
types and combinations of strain path changes.

3.5 Numerical examples


Using reported experimental trends, this section presents and qualitatively validates
typical results: (i) the evolution of the internal variables during monotonic deformation and the intrinsic orientation effect during strain path changes of BCC single
crystals and (ii) the macroscopic stress-strain behavior of BCC polycrystals during
multi-stage loading processes. Due to the lack of quantitative experimental data,
however, an adequate quantitative comparison is not possible, preventing a reliable
quantification of the material parameters. Within a physically acceptable range of
material parameters, it will be shown that the presented model is well capable of
capturing all experimental trends. Youngs modulus E, Poissons ratio , the reference strain rate 0 , the shear modulus G, the magnitude of the Burgers vector b, the
interaction coefficients A and A u , the activation energy G0 , thermal slip resistance
st , the dislocation multiplication parameter I and the recovery length R are the standard parameters in the constitutive model describing the BCC material, whereas 0 ,
c , kd , km , win f , C, B, a, kc are the additional parameters associated with the dislocation cell structure evolution. The set of standard parameters are well documented in
the literature (see Table 3.1). The remaining parameters have a restricted range and
they are estimated to establish a qualitative agreement with experimental observations. The initial value of the dislocation densities in the walls 0 logically equals c
(no cells exist yet) and the initial value of the cell size r0 can be calculated by using
equation (3.13) where f = 1 and w = 0 .

3.5.1 Example 1: monotonic deformation of single crystals


Experimental observations of the microstructure evolution during monotonic deformation have been reported in section 3.2. In most cases, the length scales of the spatial patterns formed by the dislocations decrease with increasing strain. To analyze
this change in the microstructure, several studies have been conducted. For instance,
Sevillano et al. (1981) present several curves for the average cell size with respect to
the deformation during rolling and drawing processes for different materials, and

3.5 Numerical examples

39

Fernandes and Schmitt (1983) give data on the wall thickness and dislocation cell
size of low-carbon steel under various types of loading. In the current framework,
the parameters in the evolution equations are identified to retrieve this characteristic
behavior.

Figure 3.4 / Evolution of the cell structure variables during monotonic deformation
for -Fe. Left: wall thickness w and cell size r. Right: volume fraction of the walls.

In Fig. 3.4 the evolution of the microstructure of an -Fe single crystal during a
uniaxial tension simulation at 298 K and a strain rate of 5 104 s1 is presented
with the material parameters given in Table 3.1 2 . The cell size shows a decreasing
trend with increasing strain as expected. The wall thickness decreases quickly and
stabilizes at a constant value. The volume fraction approaches a value of around 0.1
after a sharp decrease. This trend can also be found in the literature, where it is stated
that the volume fraction of dislocation walls remains at a constant value (e.g. Peeters
(2002)).

3.5.2 Example 2: strain path change of single crystals


In this example the effect of the intrinsic anisotropy of single BCC crystals during
multi-stage loading is illustrated. In Fig. 3.5 the results of a cross loading simulation
is presented where the crystal was first loaded in the [001] direction and next in the
[011] direction after unloading at a strain value of 0.15. For comparison purposes
a reference calculation is performed in which the evolution of the microstructure
2

Note that the number of papers focusing on the determination of the interaction coefficients in
BCC single crystals are very limited. Moreover, the actual value strongly depends on the actual BCC
crystal considered and impurities present and the initial dislocation density. Values may therefore
differ considerably, where a difference of a factor 10 can be easily find throughout the literature.

3 A composite dislocation cell model to describe strain path change effects


40
in BCC metals

Youngs modulus
Shear modulus
Poissons ratio
Reference strain rate
Burgers vector length
Interaction coefficient (self)
Interaction coefficient (latent)
Initial dislocation density
Activation free energy
Dislocation multiplication parameter
Dislocation annihilation rate parameter
Boltzmann constant
Thermal dislocation resistance
Saturation factor
Saturation value
Material constant

E
G

0
b
A
A u
0
G0
I
R
k
st
km
win f
C

139 GPa
64 GPa
0.362
1.07 106 s1
0.248 109 m
0.00072
0.001
0.18 1014 m2
2.95 1018 J
0.228
5.1 109
1.3807 1023 J/K
15 MPa
150
0.18 106 m
20

Table 3.1 / Material parameters for monotonic loading case. Some of the parameters, i.e. 0 , G0 and st have already been identified in previous work (Yalcinkaya
et al. (2008)). The parameters b and G are taken from Frost and Ashby (1982). The
value of 0 is presented by Krejci and Lukas (1971). The latent interaction coefficient
A u is identified and the self interaction coefficient A is obtained by assuming a
ratio of 1.4 between latent and self hardening. The remaining parameters are identified by comparing the experimental trends with computational results. Additional
parameters needed for complex strain paths are commented in the text.

is not incorporated, i.e. transient hardening and softening effects due to the strain
path change (see solid line in Fig. 3.5) are absent. The dashed line presents the
outcome of the full microstructure evolution computation, in which both the crystal slip anisotropy and the dislocation cell anisotropy are revealed. The purpose of
this example is to discriminate these two intrinsic sources of anisotropy at the single
crystal level. Obviously, both mechanisms here contribute to a larger yield stress after reloading. Whereas this increase is systematic for dislocation cell contribution, it
is obviously orientation dependent for the slip anisotropy contribution.

3.5.3 Example 3: strain path change of polycrystals


In this subsection, the performance of the model in the context of complex deformation histories is evaluated by determining the response of a BCC polycrystal with a
random texture under a sequence of: (i) two uniaxial tension tests in different directions to obtain the cross effect; (ii) simple shear and reversal in order to capture the
Bauschinger effect.

41

3.5 Numerical examples

60

50

Stress (MPa)

40

30

20

10

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

Strain

Figure 3.5 / Stress-strain curve for [001] uniaxial tension followed by [011] uniaxial
tension with (dashed line) and without (solid lines) microstructure evolution effect
for -Fe single crystal.

As explained previously, at a single crystal level the effect of intrinsic crystallographic anisotropy during a strain path change is noticeably high. The initial
anisotropy, the so-called orientation dependence of BCC single crystals has been
studied before (e. g. Yalcinkaya et al. (2008)) and this effect adds up to anisotropy
due to the dislocation microstructure evolution. In this example the main interest
focuses on the anisotropy due to substructure evolution during multi-stage loading
processes for the case where the intrinsic orientation effect is known to contribute
less. To this purpose polycrystal simulations have been conducted, where 100 randomly oriented crystals are considered interacting according to a Taylor averaging
scheme.
First, the cell dissolution process and its macroscopic cross effect are analyzed. The
characteristic feature of the stress-strain curve in Fig. 3.3 is the transient change induced by a change of the deformation path. An initial increase in the yield stress is
followed by moderate softening. The cross effect vanishes gradually and the curve
saturates towards to the monotonic case. In order to measure this effect experimentally either tension followed by simple shear or two successive orthogonal tensile
experiments need to be conducted. With respect to the latter approach Schmitt et al.
(1991) presented clear experimental results where various tensile sequences were examined, with different angles between the succeeding tensile directions equal to 15 ,
45 and 90 with different amounts of prestrain. In Schmitt et al. (1991), it was reported that no evolution of cell-blocks was observed, supporting the case examined
here, where the cell structure development is assumed to be the main mechanism

3 A composite dislocation cell model to describe strain path change effects


42
in BCC metals

accompanying the strain path change. The sequence of two uniaxial tests with 45
between the tensile axes is, according to equation (3.18), characterized by = 0.25,
and it is rather close to a cross test exhibiting the highest cell dissolution. Additional
to the parameters used during monotonic deformation (see Table 3.1), kd and B are
identified as 300 [-] and 20 [-] respectively. The obtained typical cross effect is presented in Fig. 3.6. This transient effect is observed in many materials and the size
of the effect is determined by the amount of applied prestrain while the shape of the
hardening and softening zones depends on the material.

60

50

Stress (MPa)

40

30

20

10

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

Strain

Figure 3.6 / Predicted stress-strain curve of a 45 cross tensile test of an -Fe polycrystal.

The second example concerns the Bauschinger effect, which yields a reduction of the
yield strength of the material after a load reversal. A simple shear and reversal simulation is presented in Fig. 3.7 , where the evolution of the cell size r is dominantly
contributing to the anisotropy at this continuum level. Additional to the parameters used during monotonic deformation in Table 3.1, the parameters a and kc are
identified as 1 104 m and 14 [-] respectively, to validate this part of the model.
Even though both the cross effect and the Bauschinger effect are extensively documented in the literature, quantitative data on the evolution of dislocation cell structure during strain path changes remains hard to find. For this reason, a qualitative
analysis rather than a quantitative study has been conducted here.

43

3.6 Summary and Conclusion

50
45
40

Stress (MPa)

35
30
25
20
15
10
5
0
0

0.05

0.1

0.15

0.2

0.25
Strain

0.3

0.35

0.4

0.45

Figure 3.7 / Bauschinger effect for shear - reverse shear test of an -Fe polycrystal.

3.6 Summary and Conclusion


This paper has presented a computational study on the anisotropy effects induced
by strain path changes for BCC structured metals. For this purpose a composite
dislocation cell model, which describes the dislocation substructure evolution, has
been combined with a BCC crystal plasticity framework to bridge the dislocation cell
structure evolution and its macroscopic anisotropic effects. The BCC crystal plasticity framework was based on Yalcinkaya et al. (2008) and the composite cell model
was built upon the contribution of Viatkina et al. (2003), who analyzed strain path
dependency phenomena in a phenomenological plasticity framework at small strains
for FCC structured materials.
The presented computational framework assumed a composite aggregate, in which
the material with a cell structure was considered to consist of two components: a soft
cell interior component and hard cell wall components. The constitutive response of
each component has been obtained from crystal plasticity simulations, while a set of
phenomenological evolution equations for the cell size, the wall thickness and the
dislocation density inside the walls captured the evolution of the microstructure.
The numerical examples of this work have revealed an adequate qualitative agreement between the simulations and the experimental trends for strain path change
tests, i.e. a cross test and a Bauschinger test. Further quantitative analyses call for
more extensive and more qualitative experimental results to compare with.
The paper clearly forwards a number of original contributions:
A phenomenological cell structure evolution model embedded into a crystal

3 A composite dislocation cell model to describe strain path change effects


44
in BCC metals

plasticity framework is well able to reproduce all essential characteristics of


strain path changes reported, consistently with experimental observations at
two scales.
The model proposed allows to study the interaction between different sources
of anisotropy, where a clear example at the single crystal and polycrystal has
been given.
The level at which the enrichment of the crystal plasticity model was made,
enables its use in more complex microstructures as e.g. multi-phase steels.

Chapter four

Deformation patterning driven by rate


dependent non-convex strain gradient
plasticity1

Abstract / A rate dependent strain gradient plasticity framework for the description
of plastic slip patterning in a system with non-convex energetic hardening is presented.
Both the displacement and the plastic slip fields are considered as primary variables.
These fields are determined on a global level by solving simultaneously the linear momentum balance and the slip evolution equation which is postulated in a thermodynamically consistent manner. The slip law differs from classical ones in the sense that it includes a non-convex free energy term, which enables patterning of the deformation field.
The formulation of the computational framework is at least partially dual to a GinzburgLandau type of phase field modeling approach. The essential difference resides in the
fact that a strong coupling exists between the deformation and the evolution of the plastic slip, whereas in the phase field type models the governing fields are only weakly
coupled. The derivations and implementations are done in a transparent 1D setting,
which allows for a thorough mechanistic understanding, not excluding its extension to
multidimensional cases.

4.1 Introduction
During forming processes most metals develop cellular dislocation structures due to
dislocation slip patterning from moderate strains onwards. Typical examples of dislocation microstructures are dislocation cells and dislocation walls (see e.g. Young
et al. (1986), Mughrabi (1987), Yalcinkaya et al. (2009)). Patterning typically refers
to the self organization of dislocations with formation of regions of high dislocation
1

This chapter is reproduced from Yalcinkaya et al. (2011a)

45

46

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

density (dislocation walls) which envelop areas of low dislocation density (dislocation cell interiors), also to be regarded as domains of high plastic slip and low plastic
slip activity. Due to the induced macroscopic anisotropic effects (see e.g. Peeters
et al. (2000), Li et al. (2006), Wang et al. (2008), Yalcinkaya et al. (2009)) the occurrence
of dislocation microstructures and their evolution have been an interesting topic
for the materials science community for decades. Starting with the studies on the
cold-worked sub-structure of polycrystals using transmission electron microscopy
in 1960s (e.g. Bailey and Hirsch (1960), Keh et al. (1963) and Swann (1963)), a vast
amount of experimental results have been collected, and several promising theoretical models were presented dealing with dislocation (or slip) patterning. Nevertheless, a complete descriptive understanding of the occurring phenomena has never
been reached and the necessary input for computational models is still subject of
ongoing discussions.
In the context of the computational modeling of plastic slip pattering (or dislocation
sub-structure formation), different approaches have been pursued in the literature
that can be categorized into three main groups: (i) models using directly the mechanics of single dislocations or populations of dislocations, (ii) phase field modeling
of dislocation patterning, (iii) the incremental variational formulation of inelasticity
which has the advantage of applying the concept of relaxation resulting in fine scale
microstructure evolution.
In the first group, scientists deal either with the problem at a discrete dislocation level
by solving a system with only a limited number of dislocations (e.g. Lubarda et al.
(1993), Groma and Pavley (1993), Kubin and Canova (1992)), or they approach the
problem from a continuum point of view by using homogenized variables like dislocation densities and internal stress fields in a system of coupled balance equations.
One of the fundamental studies in this class was introduced by Kuhlmann-Wilsdorf
(see e.g. Kuhlmann-Wilsdorf and Van der Merwe (1982)) within the so-called lowenergy dislocation structure (LEDS) approach, seeking for dislocation configurations
of minimal free energy under given constraints. In this framework, pattern formation
is driven by the reduction of the system energy. Holt (1980) improved this static description by introducing dynamics into the model, which is based on a conservation
law for the dislocation density by setting up a relation for the evolution of the dislocation density in terms of elastic energy changes related to the dislocation density
fluctuations. Another approach in this class is the reaction-diffusion model (e.g. Walgraef and Aifantis (1985) and Aifantis (1987)), where mobile and immobile dislocations are distinguished and the dynamics of the system is governed by diffusion and
reaction terms. The competition between the mobility and the non-linear interactions
(creation, annihilation and pinning) causes the instability of uniform dislocation distributions versus inhomogeneous ones and leads to the formation and persistence of
dislocation patterns. Even though there have been some improvements on the men-

4.1 Introduction

47

tioned models, they generally rely on assumptions that are difficult to validate at the
discrete dislocation level.
In the second group, there are the phase field modeling approaches of dislocation
patterning, which offer a valuable alternative for discrete dislocation dynamics simulations. These models have the advantage of making less small scale assumptions
on the dislocation interactions, such as multiplication/annihilation of dislocations.
Instead, elastic interactions of numerously interacting individual segments of dislocations are considered, dealing with only a few density functions (see Wang et al.
(2001a), Wang et al. (2001b) and Wang et al. (2001c)). However, these models require
the use of large computational grids and time integration over large numbers of time
steps. These difficulties are overcome by the model of Koslowski et al. (2002) presenting an analytically tractable theory which determines the value of phase field at
point-obstacle sites without using any grid. Even though the framework is successful
in predicting dislocation (slip) patterns, it is rather complicated to incorporate into
a finite-deformation formulation of single-crystal elastic-plasticity, as needed in the
context of large-scale finite element calculations of macroscopic samples.
The third category of methods aims to capture the phenomenological evolution of
dislocation microstructures by modeling a physically deformed crystal in finite plasticity frameworks, which are the most relevant ones considering the present paper.
Ortiz and Repettto (1999) regard the dislocations as manifestations of the incompatibility of the plastic deformation gradient field. Within this framework the incremental displacements of inelastic solids follow as minimizers of a suitably defined
pseudoelastic energy function. Miehe et al. (2004) propose an approach based on
finite-step-sized incremental energy minimization. The boundary value problems
were recast into a principle of minimum incremental energy for standard dissipative solids. The general concepts are applied to analyze evolving deformation microstructures in single-slip plasticity. The derivations of the incremental variational
formulation of inelasticity in both frameworks are conceptionally parallel to each
other. Hackl and Kochmann (2008) employ energy principles as well to analyze the
microstructure formation and evolution as a result of energy minimization and relaxation via lamination. The idea is that, for non-quasiconvex energy potentials the
minimizers are no longer continuous deformation fields but small-scale fluctuations
related to probability distributions of deformation gradients to be calculated via energy relaxation.
The objective of the present paper consists in bridging models of the last two categories described above. To achieve this, a continuum plastic slip field model in a
non-convex strain gradient plasticity framework is proposed to simultaneously predict microstructure formation and the overall macroscopic elastic-plastic response in
a computationally low cost setting. A viscous relaxation scheme in a system with en-

48

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

ergetic hardening is used, which makes the method comparable to non-local phase
field models (e.g. Ubachs et al. (2004)) as used in diffusional evolving microstructures. Using the adequate formulations for the physics, thermodynamics and kinematics, phase field models are able to predict the evolution of microstructures and
morphologies without explicit tracking of the positions of the phases and interfaces.
Inspired by their success, the present paper extends this concept to the formation
and evolution of plastic slip patterns based on a consistent derivation of the thermodynamic relations and evolution equations. Plastic slip patterns in metals are
typically linked to the evolution of dislocation sub-structures, i.e. dislocation cell
walls and cell interiors. The computational framework yields a stable algorithm for
which the finite element formulation and implementation has been performed in a
fictitious 1D setting in order to preserve simplicity and transparency, and to easily
interpret the results. Moreover, in this simple presentation the theory stays general
and depending on the parameters and their physical interpretation, the model can
potentially describe different dislocation slip microstructures (from Luders

bands
to dislocation cells) at different length scales. The formulation can be extended to
2D or 3D, where the multi-slip character of crystals is often considered as a natural
source of non-convexity due to latent hardening mechanism. In line with models of
the third category, a double-well free energy function is assumed here, which makes
the approach rather phenomenological for the considered 1D case, whereas the nonconvex nature in a single slip deformation state becomes more natural through the
interaction of multiple slip systems (e.g. through latent hardening). In the model, the
plastic slip and the displacement are taken as degrees of freedoms. These fields are
obtained on a global level by solving simultaneously the linear momentum balance
and the slip evolution equation. The thermodynamically consistent slip law is the
crucial part of the model. At first glance, the slip law (see equation (4.18)) is similar
to the one encountered in a classical rate dependent crystal plasticity approach (e.g.
Hutchinson (1976), Peirce et al. (1982), Yalcinkaya et al. (2008)), however the stress
expression in the numerator differs. Three participating stress contributions can be
distinguished: (1) the conventional resolved stress directly related to the external
loading, (2) a surface-like stress depending on the gradient of plastic slip which is
characteristic for strain gradient crystal plasticity models (e.g. Evers et al. (2004b))
and (3) the stress that emanates from a non-convex free energy, which is omitted
in classical convex theories and which triggers the patterning of the dislocation slip
field.
The paper is organized as follows. First, in section 2, the incremental variational approach for microstructure evolution is discussed shortly including some comments
on its connection with the present framework. Then, in section 3, the thermodynamical consistency of the constitutive model and the derivation of the equations to be
solved in a finite element context are studied. Next, in section 4, the finite element

4.2 Macroscopic view: material instability and microstructure evolution in


inelastic materials
49

formulation for the 1D strain gradient plasticity framework and the incorporation of
the non-convexity into the problem are summarized. Further, section 5 addresses the
link between the non-convexity and the resulting patterning in phase field models in
general and in the presented model in particular. In section 6 numerical examples are
presented in order to demonstrate the performance of the proposed model. Finally,
some concluding remarks are given in section 7.

4.2 Macroscopic view: material instability and microstructure


evolution in inelastic materials
The mechanical response of many engineering components is often influenced by
an existing or an emerging microstructure (martensite, dislocation sub-structures,
voids, shear bands, etc.). To account for the actually relevant microstructure, there
are different approaches to model the formation and the development of the microstructure as additional fields in the material. The difficulty in the solution of
these conventional (local) multi-field problems is the localization of the corresponding field and local strain hardening-softening elastic-plastic behavior, which yields
numerical instabilities. The boundary value problem becomes ill-posed and in the
context of finite element formulations it gives mesh dependent results (e.g. de Borst
(1987)). Regarding localized failure mechanisms and softening problems, a broad
range of regularizing approaches have been developed as viscoplasticity, non-local
continuum theories and Cosserat theories. Additional to these classical techniques,
another recent approach in the context of emerging and evolving microstructures formulates the inelastic boundary value problems by using direct methods of calculus
of variations in an incremental setting (e.g. Ortiz and Repettto (1999), Miehe (2002),
Miehe et al. (2002), Lambrecht et al. (2003), Miehe et al. (2004), Svendsen (2004)).
Variational calculus principles are well-understood in the theory of elasticity. Material stability conditions, and well-posedness of the problem specifications have been
studied extensively. The conclusion is that for a globally stable material, the free
energy function should be convex. However, due to its strong restrictions weak convexity concepts (polyconvexity, quasiconvexity and rank-one convexity) have been
used in stability analyses and fine scale microstructure evolution descriptions in the
material. If the elastic energy function is non-convex the minimization problem does
not give a solution in the classical sense and fine scale microstructures are generated. Boundary conditions play also an important role in this procedure and the
variational problem might have infinitely many, a unique or no solution at all.
The deformation theory of plasticity, where all material points are assumed to follow
certain optimal deformation paths, can be elaborated by pursuing the well-defined

50

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

variational principles of elasticity as well. However, the flow theory of plasticity


and many other inelastic material descriptions cannot be handled in the same way
due to the intrinsic path dependence. Therefore, incremental variational principles
have been applied to remedy the numerical stability issues and to model the formation and evolution of microstructures in dissipative scale-invariant inelastic materials
(e.g. Miehe (2002), Miehe et al. (2002), Lambrecht et al. (2003), Miehe et al. (2004)).
The problem can be formulated as,
W (n+1 ) = inf
I

tn +1
tn



+ D dt

(4.1)

(I , I ) is the dissipation with I standing for the set of internal variables,


where D = D
is the strain, is the free energy functional and W is the minimum energy at the solution point. The minimization problem is solved at each time increment (tn+1 tn )
allowing for path dependent problems (e.g. plasticity) revealing microstructures.
The local minimization problem becomes non-local or scale variant when a length
scale is incorporated by supplying an additional energy component in the equation
(4.1) reflecting the size effect (e.g. Conti and Ortiz (2005)). The mathematical method
for the solution of the above problem is well known under the name relaxation. This
method produces a new well-posed relaxed problem in the sense of existence of solutions and implicitly allows for the formation of microstructures. The relaxation is
associated with quasi-convexification of the non-convex function W.
The main disadvantages of the incremental variational formulation are its high computational cost and the fact that it allows one to obtain only equilibrium states of
microstructures. Computational cost can be reduced by implementing analytical relaxation solutions to the problems (e.g. Conti et al. (2007)) which might be limited
for realistic calculations. The issue of tracking the actual non-equilibrium evolution
of microstructures is the aim of the framework that is presented in the following
sections. The viscous nature of the model allows the prediction of non-equilibrium
states of microstructure evolution depending on the rate of deformation.

4.3 Thermodynamics of strain gradient plasticity


In this section, the thermodynamical consistency and the derivation of the governing
system equations are discussed shortly. For conceptual simplicity, all formulations
are casted in a 1D setting (where a coordinate x is used to indicate the position), not
limiting their extension to 2D or 3D. In the geometrically linear small strain context
the time dependent displacement field is denoted by u = u( x, t), the strain is given
The strain is assumed to be decomposed
by = u/x and the velocity is v = u.

4.3 Thermodynamics of strain gradient plasticity

51

additively,

= e + p

(4.2)

into an elastic part e and a plastic part p . In the single slip 1D case considered here,
the total amount of plastic slip and the plastic strain are identical. Note that in
general the plastic slip may be derived from a collection of internal variables, even
for a 1D problem (in particular if multiple slip systems are present). The free energy
is assumed to be a function of the state variables,
state = e , ,

(4.3)

where = /x is the gradient of the plastic slip. In the multi slip crystal plasticity context the definition of plastic slip gradients is not trivial, and Cermelli and
Gurtin (2001) summarize the different options and requirements for such a formulation. Following the arguments of Gurtin (e.g. Gurtin (2000), Gurtin (2002)), the
power expended by each independent rate-like kinematical descriptor is expressible
in terms of an associated force system consistent with its own balance. Yet, the basic
kinematical rate variables, namely e , u and are not independent. It is therefore not
apparent what forms the associated force balances should take, and, for that reason,
these balances are established using the principle of virtual power.
e and
Assuming that at a fixed time the fields u, e and are known, we consider u,
e , ).
as virtual rates which are collected in the generalized virtual velocity V = (u,
The force systems are characterized through their work-conjugated nature with respect to the state variables. Pext is the power expended on the domain P and Pint a
concomitant expenditure of power within P,

Pext ( P, V) =
Pint ( P, V) =

R
R

t(n)udS
+

e
P dV +

(n) dS

dV +
P

(4.4)

dV
P

where , and are the thermodynamical forces conjugate to the internal state variables e , and respectively. In Pext , t(n) is the macroscopic surface traction while
represents the microscopic surface traction conjugate to at the boundary P with
n indicating the normal direction.
Postulation of the principle of virtual power states that given any generalized virtual
velocity V the corresponding internal and external powers are balanced,

Pext ( P, V) = Pint ( P, V)

(4.5)

52

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

Considering a generalized virtual velocity without slip, the virtual velocity field can
be chosen arbitrarily and this leads to the classical macroscopic force balance

=0
x

(4.6)

and considering the virtual slip field arbitrarily without generalized virtual velocity
leads to the microscopic force balance,

+ = 0
x

(4.7)

In Gurtin (2000), Gurtin (2002), Gurtin et al. (2007) the previous derivations are fully
detailed in a generalized 3D setting while Del Piero (2009) presents an alternative
way to derive similar balance equations.
The local internal power expression can simply be written as,
Pi = e + +

(4.8)

Hence, the local dissipation inequality can be expressed as


D = Pi 0

or

D = e + + 0

(4.9)

The free energy is assumed to take the following form,

= e + +

(4.10)

where e is the elastically stored energy, is the microstructurally stored energy


due to plastic slip and is the energy associated to the gradients of slip.
By exploiting (4.10), the dissipation (4.9) can be rewritten as,

D = e + + e e







de
d
d
e
= e +
+
0
d
d
d
|
{z
}
{z
}
|
0

(4.11)

Additional to the stress , the microforce conjugate to the slip gradient is also
assumed to be energetic:

de
de

d
=
d
The remaining term in the dissipation expression (4.11) reads,

(4.12)

4.3 Thermodynamics of strain gradient plasticity


d
D=
0
d
|
{z
}

53

(4.13)

dis

where the term conjugated to the slip rate is identified as the dissipative stress dis,

dis =

d
d

(4.14)

The following constitutive equation which satisfies the inequality (4.13) is next proposed,

dis = sign ( )

(4.15)

where represents the actual slip resistance,


 m
| |
=s
0

(4.16)

In here, 0 and m are the reference slip rate and the rate sensitivity exponent respectively. Furthermore, s is the resistance to dislocation slip which is assumed to
be constant for simplicity. Substitution of equation (4.16) into equation (4.15) and
extracting the slip rate yields the slip law

= 0

| dis |
s

 m1

sign ( dis )

(4.17)

As a result, a thermodynamically consistent constitutive relation for the slip evolution is obtained. By using the definition of dis in equation (4.14) and using the
microforce balance according to equation (4.7), equation (4.17) can be written as,
m1

d
| x + d |
| {z }

sign ( d )
= 0

s
d

(4.18)

Each of the stress contributions in the slip law is derived from the free energy, i.e.
= /e , = / and any change in the free energy directly affects the slip
law.

54

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

4.4 Particular choices of free energy functions


4.4.1 Slip based strain gradient plasticity
In this section the algorithm to solve equations (4.6) and (4.18) as a rate dependent
(convex) strain gradient plasticity problem over a domain 0 < x < L is outlined.
Treating this case provides a natural setting for the extension to the non-convex case
and its proper interpretation. The displacement field u and the plastic slip field
are selected as primary variables, which leads to the simultaneous solution of the
balance of linear momentum and the evolution of slip.
A convex free energy function,

= e + =

1 e2 1
E + A( )2
2
2

(4.19)

leads to classical strain gradient crystal plasticity frameworks, where E is Youngs


modulus and A (which includes an internal length parameter) is governing the effect of the slip gradient contribution to the internal stress field. Its expression in 3D
would be A = ER2 /(16(1 2 )) as e.g. used in Evers et al. (2004b), Bayley et al.
(2006), Geers et al. (2007). In this expression R physically represents the radius of the
dislocation domain contributing to the internal stress field, is Poissons ratio. Note
that in equation (4.19) does not include the term. Adding a convex energy term
would result in a stress-like expression similar to the a-thermal slip resistance in
the numerator of the slip law (4.18) as used in some BCC crystal plasticity frameworks (e.g. Yalcinkaya et al. (2008)). The governing system of equations is given by
the strong form of the linear momentum balance (equation (4.6)) and the plastic slip
evolution (equation (4.17)),

| dis |
s

 m1

sign (

dis

= 0
(4.20)

) = 0

For further elaborations it is more convenient to write the plastic slip evolution in the
following format,

| |m sign ( )

0m dis
=0
s

Implicit time integration gives,






m
n m

sign n 0 dis = 0
t
t
s

(4.21)

(4.22)

55

4.4 Particular choices of free energy functions

with t representing the time step and n is the plastic slip at the end of the previous
time step.
The dissipative stress in equation (4.22) can be written as

dis = +

(/ )
2
=+
=+A 2
x
x
x

Substituting equation (4.23) into equation (4.22) gives,








2
m
n m
sign n 0 + A = 0

t
t
s
x2

(4.23)

(4.24)

The weak forms of the equations are obtained in a standard manner, using a Galerkin
procedure. Both the balance of linear momentum and the slip equation are tested
with virtual displacement u and virtual slip fields, respectively, and integrated
over the domain 0 < x < L. The obtained variational Galerkin functionals are solved
in a fully coupled manner (monolithic) by means of a Newton-Raphson scheme after
linearization and discretization procedures as outlined in the appendix section. Linear interpolation functions for the slip field and quadratic interpolation functions
for the displacement field u are used. The resulting system of equations is solved
for the increments of the displacement u and the plastic slip . The formulation is
numerically implemented and solutions for the nodal displacements and plastic slips
are obtained in a standard incremental iterative manner. Examples are presented in
section 4.6.

4.4.2 Slip based non-convex strain gradient plasticity


The non-convex case is recovered by adding a non-convex contribution to the free
energy (4.19). The additional term is a polynomial function of the plastic slip

= e + +

1 e2 1
E + A( )2 + (C1 4 + C2 3 + C3 2 + C4 + C5 )
2
2

(4.25)

where non-convexity is obtained by specific values of the polynomial coefficients.


The above procedure is similar to the inclusion of non-convexity in the configurational free energy of phase field models which have been studied extensively in the
literature. In many of the cases the non-convexity is approximated by a double well
potential (capturing the patterned field) where the wells correspond to the states
governing the minimum energy configuration of the mixture.
Considering the 3D evolution of the dislocation microstructures (e.g. labyrinth, mosaic, fence and carpet structures) driven by the imposed deformation, Ortiz and

56

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

Repettto (1999) present a comprehensive summary on the relation between the nonconvexity of the energy function and the dislocation microstructure evolution. In
crystals exhibiting latent hardening, the energy function is non-convex and has wells
corresponding to single-slip deformations. This favors microstructures consisting locally of single slip. Looking from thermodynamics perspective Kuhlmann-Wilsdorf
LEDS (low energy dislocation structure) theory (e.g. Kuhlmann-Wilsdorf (2001))
completes the explanation by stating that among all potentially accessible dislocation structures, plastic deformation will generate the one with the lowest free energy.
This means that, limited only by dislocation mobility, availability of slip system and
insignificant entropy, dislocation structures always approach the lowest possible mechanical energy of the present dislocation population.
The present study clearly reveals the intrinsic role of the non-convexity in a 1D problem, even though it is a simplification of the underlying (more complex) 3D reality.
The non-convex energy which is coming from the accumulation of trapped dislocations is introduced phenomenologically in terms of a dislocation slip potential representing the microstructurally trapped energy. It is introduced in terms of plastic
slip representing the dislocation movement looking for the minimum energy configuration (in analogy to patterning in phase field models). The convex gradient term
represents the surface energy (penalizing spatial transitions from low to high values
of slip), which in fact regularizes the problem in the mathematical sense. Depending
on the value of the R (hidden in parameter A standing in front of the gradient term),
the presented theory is able to explain the formation and evolution of microstructures at different length scales.
The non-convex function does not have to be a polynomial one. The free energy function adopted is just a simple mathematical representation for a double-well function
motivated from phase field models. The parameters C1 , C2 and C3 are chosen in a
way that they introduce a small modulation from a convex plastic slip potential (see
Fig. 4.3 (a)). Their exact values are therefore not essential if the non-convex modulation is superimposed on a free energy function used for classical hardening laws.
It is clear that if a convex plastic slip potential enters equation (19) it would result in
only hardening (no softening branch) behavior and a homogeneous (constant) distribution of the plastic slip would be obtained.
The strong form of the system description according to the equations (4.20) remains
identical. The dissipative stress term in equation (4.23) can be rewritten as,

2
= + A 2 (4C1 3 + 3C2 2 + 2C3 + C4 )
x

dis = +

(4.26)

57

4.5 Non-convexity and patterning in phase field modeling

Substituting equation (4.26) into equation (4.22) gives,






n m

sign n
t
t


2
m
0

3
2
+ A 2 (4C1 + 3C2 + 2C3 + C4 ) = 0

s
x

(4.27)

The same procedure as in the previous section is followed to obtain the set of equations to be solved for the displacement and plastic slip increments (see the Appendix
for details).

4.5 Non-convexity and patterning in phase field modeling


In phase field models (e.g. Ubachs et al. (2004), Kuhl and Schmid (2007)), the configurational energy of a two-phase material, is often represented by a (non-convex)
double-well potential (Fig. 4.1). It can be constructed from the configurational free
energy of the individual phases by assuming that at a certain composition only the
phase with the lowest energy will exist. Equilibrium is reached when the chemical
potential becomes homogeneous throughout the system. The configuration with the
lowest possible free energy is found when, due to phase separation, each material
point has the composition of either of the two phases, corresponding with the binodal points. In the context of plastic slip patterning there are no physically distinct
phases but rather distinct regions in which the slip is either high (large dislocation
density) or low (small dislocation density). Two other points which are typically of

Figure 4.1 / Free energy curves of a two-phase material, and binodal and spinodal
points.

interest are the so-called spinodal points. These points mathematically reflect the

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

58

state (e.g. composition) for which the second derivative of the configurational free
energy becomes zero. At these points the sign of the curvature (2 c /c2 ) changes.
In the region where the curvature is positive the system is stable with respect to
small fluctuations in composition. However in the regions where the curvature is
negative (between the spinodal points) the system is locally unstable with respect to
small fluctuations and for that reason phase separation occurs (see e.g. Cahn (1961),
Langer (1971), Nauman and He (2001)). For the case considered here, this issue is
easily shown on the basis of a stability analysis of the steady state solution of the
plastic slip distribution, implying that = 0,

dis = + A

=0
x2

(4.28)

which can be considered as a nonlinear equation in terms of . A spatial wave perturbation is applied to the distribution = 0 + = 0 + gei(kxt) with g the
complex amplitude, k the wave number and the frequency. Linearization of (4.28)
around 0 yields the behavior in terms of the perturbation
A

2 2
2
= 0
x2
=0

Substitution of the spatial perturbation results in,




2
2
Ak +
= 0

2 =0
Real values for the wave number k are found if
2
<0

2 =0

(4.29)

(4.30)

(4.31)

Indicating that physical solutions for such spatially non-homogenous perturbations


exist (corresponding to patterning of the plastic slip). For a double-well function
according to = C1 4 + C2 3 + C3 2 + C4 + C5 , the spinodal points are the roots
of
6C1 2 + 3C2 + C3 = 0

(4.32)

and between these two spinodal points the system is unstable and any given perturbation will give rise to patterning of the plastic slip.

4.6 Numerical examples


In this section, three different numerical cases of a fictitious 1D bar with length L
under tension are presented to study the behavior of the discussed strain gradient

4.6 Numerical examples

59

models. The first example deals with a conventional convex free energy with slip
gradients (see section 4.4), where the influence of the internal length parameter determining the value of the constant A (see equation (4.19)) is elucidated. The second
example studies the absence, onset and evolution of patterning of the plastic slip, departing from a homogeneous distribution, depending on the deformation rate during
monotonic loading of a bar. The homogeneous (non-patterned) case is also compared
with the analytical solution. The third example studies plastic slip patterning in a relaxation test where the 1D bar is deformed to a certain state mapped between the
spinodal points. Constraining the deformation at this point leads to viscous relaxation of the microstructure evolution. The effect of the boundary conditions on the
slip patterning is analyzed in each example by considering hard ( = 0) and soft
( /x = 0) cases.

4.6.1 Numerical example 1: convex case - monotonic loading


In this example, the influence of the internal length parameter R present in the material constant A is briefly addressed, in terms of the plastic slip distribution and the
stress strain behavior. The (convex) strain gradient plasticity framework of section
4.4.1 is used to this purpose. The displacement at x = 0 is suppressed while displacement at x = L is prescribed such that the average strain increases incrementally up
to 0.05. Hard boundary conditions ( = 0) are applied at both ends of the bar. The
material parameters are: Youngs modulus E = 210 GPa, Poissons ratio = 0.33,
slip resistance s = 15 MPa, reference slip rate 0 = 5 102 s1 , the rate sensitivity
exponent m = 1 and the length of the bar is L = 1 mm. The bar is discretized into
100 finite elements and deformed with a rate of = 5 102 s1 . It is important
to note that, the material parameters do not represent a certain material, and taking
m=1 in the numerical examples may seem too simple at first sight. This choice was
of course made for simplicity only, yet with a large similarity to discrete dislocation
studies using linear drag relations. Simulations with other values of m do not change
the qualitative nature of the examples, yet they will affect the rate dependent (time
dependent) behavior. Considering the parameter s, it is remarked that its constant
value does result in redundancy, however, the introduction makes the stress term in
the slip evolution equation (4.22) dimensionless.
As pointed out in various strain gradient crystal plasticity models (e.g. Gurtin (2002),
Svendsen (2002), Bardella (2006), Bardella (2007), Bayley et al. (2006)), the effect of
the internal length scale enters the formulations via the (internal) back stress. In the
current framework, the general expression A2 /x2 in equation (4.23) gives rise to
an internal back stress as occurring in a microstructure with dislocations.
The results presented in Fig. 4.2 show accordance with the solutions from the litera-

60

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity
40
35

Stress (MPa)

30

R = 0.01 mm
R = 0.02 mm
R = 0.05 mm

25
20
15
10
5

0.2

0.4

0.6

0.8

0
0

0.01

0.02

0.03

0.04

0.05

Strain

Figure 4.2 / 1D bar in tension with hard slip boundaries.

ture studying energetic size effects (e.g. Shu et al. (2001), Evers et al. (2004a), Dunstan
and Bushby (2004)). Because of the constrained slip at each end of the bar, plastic slip
gradients develop, resulting in inhomogeneous deformation with the occurrence of
boundary layers. The boundary layer thickness is typically increasing with R. As expected, Fig. 4.2 also reflects the size dependence of the stress vs. strain curve which
is consistent with the literature on this aspect.

4.6.2 Numerical example 2: non-convex case - monotonic loading


Theoretical homogenous solution
In this example we investigate homogeneous deformation (i.e. without patterning)
for the non-convex strain gradient plasticity framework and compare the results with
the analytical solution. The non-convexity is introduced through a small modulation
on top of a convex energy function (representing a classical hardening behavior in a
qualitative sense). The effect of this small non-convex modulation is discussed further on. The rate dependent character of the model renders the possibility to stabilize
homogeneous deformation states at high rates of loading since microstructures do
not have enough time to evolve. A 1D bar is constrained at the ends (suppressed displacement at x = 0 and prescribed displacement at x = L) where soft boundary conditions ( /x = 0) are applied. The results of the finite element computation will be
compared with the analytical solution by solving the slip equation without the nonlocal effects in the steady state limit for a set of applied strain values. The material
parameters are: Youngs modulus E = 210 GPa, Poissons ratio = 0.33, slip resistance s = 35 MPa, reference slip rate 0 = 5 s1 , internal length parameter R = 0.1

61

4.6 Numerical examples

mm, the rate sensitivity exponent m = 1 and the length of the bar is L = 1 mm. The
overall deformation rate used in this calculation is = 2 s1 . This rather high rate
prevents microstructures from evolving and thereby favors a merely homogenous
solution even for the non-convex case considered here. The plastic slip dependent
free energy is taken, in analogy with phase field approaches, as a double-well
non-convex potential = 1.525 108 4 5.2 106 3 + 5 104 2 MPa. Note,
however that with respect to this item there is insufficient experimental data to specify the precise form and nature of this non-convex contribution. The spinodal points
are identified by 2 / 2 = 0 yielding sp1 = 0.0042 and sp2 = 0.0131. The binodal
points are obtained by extracting the points where the tangent line touches the free
energy curve, bp1 = 0.001 and bp2 = 0.0163 (see Fig. 4.3 (a)).

350
300

Stress (MPa)

250
200
150
100
50
0
0

0.005

0.01

0.015

0.02

Strain

(a) Plastic slip dependent free energy.

(b) Stress vs. strain.

Figure 4.3 / (a) Applied plastic slip dependent part of the free energy density function for the non-convex case (solid line) with spinodal (stars) and binodal (polygons)
points and for the convex case (dotted line). (b) Stress vs. strain response for a homogeneous deformation (solid line: non-convex case, dotted line: convex case).

For the given free energy ( shown in Fig. 4.3 (a)) and the other selected parameters
a homogeneous plastic slip distribution all along the bar is recovered. The resulting
stress vs. strain response for this particular case is shown in Fig. 4.3 (b). In Fig. 4.3
the effect of the convex plastic slip free energy density is also presented with material
parameters: C1 = 0.72 108 MPa, C2 = 2 106 MPa and C3 = 2.1 104 MPa. This
set of parameters leads to a small change in the free energy curve, however results
in a significantly different behavior in the stress-strain response, i.e. the convex case
does not show a softening branch. The peculiar behavior described by Fig. 4.3 (b)
at this scale can be related to Luders

bands and the Portevin-Le Chatelier effect (e.g.


Sun et al. (2003) and Halim et al. (2007)). However, note that this example represents

62

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

Free energy density (MPa)

2.5
Total
Plastic
Elastic

1.5

0.5

0
0

0.005

0.01

0.015

0.02

Strain

Figure 4.4 / Energy plots of the homogeneous case (analytical solution).

the idealized homogeneous plastic slip distribution case which is not easy to obtain
in experiments due to the fact that plastic deformation always imposes evolution of
dislocation slip microstructures. Therefore, stress vs. strain curves generally include
a plateau corresponding to microstructure evolution (see following examples).
The analytical solution for this case is obtained assuming a steady state, = 0 revealing dis = 0 and using the homogeneity condition A2 /x2 = 0. The slip equation
then simplifies to:
E( ) 4C1 3 3C2 2 2C3 = 0

(4.33)

Equation (4.33) is solved analytically for a set of strain values ranging from 0 to 0.02
which is a different way to recover the homogeneous solution. This solution is equal
to the homogeneous continuous loading case if the strain rates are not very high
(viscosity effects are minor). The contributing terms to the free energy densities are
plotted with respect to strain in Fig. 4.4. The obtained stress vs. strain response is
identical with the one from the finite element calculation (see Fig. 4.3 (b)). Within
the applied strain range, the total free energy also shows a double-well non-convex
behavior, which is an important prerequisite for patterning of the plastic slip.
Patterned solution and rate dependency
In this example, the rate dependent evolution of the plastic slip patterns during
monotonic loading is dealt with. The geometry and material parameters are identical to the ones in the previous example. First, results of hard and soft boundary
conditions are respectively presented in Figs. 4.5 and 4.6 at a strain rate of = 0.02
s1 which is low compared to the value used in the previous subsection. The effect

63

4.6 Numerical examples

of the loading rate on the stress vs. strain response and slip patterning is shown in
Figs. 4.7 and 4.8.

400
Patterned solution
Homogenous solution

350

Stress (MPa)

300
250
200
150
100
50
0
0

0.005

0.01

0.015

0.02

0.025

Strain

(a) Stress vs. global strain.

(b) Plastic slip evolution.

Figure 4.5 / Stress vs. strain response and plastic slip evolution (legend presenting
the global strain) for a low monotonic loading rate with hard boundary conditions.

350
Patterned solution
Homogenous solution

300

Stress (MPa)

250
200
150
100
50
0
0

0.005

0.01

0.015

0.02

0.025

Strain

(a) Stress vs. global strain.

(b) Plastic slip evolution.

Figure 4.6 / Stress vs. strain response and plastic slip evolution (legend presenting
the global strain) for a low monotonic loading rate with soft boundary conditions.

In Figs. 4.5 (a) and 4.6 (a), the stress vs. strain curve is presented for the patterned solution together with the corresponding homogeneous solution. The stress response

64

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

shows a typical plateau for the values of strain where plastic slip patterning is observed. Due to the low values of the elastic strain the total strain in the stress vs.
strain curves can roughly be linked to the plastic slip values. The plastic slip limits of the plateau regime in the stress vs. strain curves coincide with the spinodal
points in the free energy and the distribution (decomposition) of the plastic slip during the inhomogeneous plastic slip evolution matches with the binodal points (see
Fig. 4.3). The constant stress plateau corresponds to the convexified solution of the
non-convex stress potential, where a linear convex envelope yields constant stress
values (see e.g. Lambrecht et al. (2003) and Miehe et al. (2004)).
The patterned solutions in 4.5 (a) and 4.6 (a) converge in the rate independent limit
to Maxwell-lines which can be associated with a global convexification of rate independent problems, such as discussed in Lambrecht et al. (2003). However, while
a convexification approach is able to resolve full Maxwell-lines similar to classical
treatments in phase-decompositions of real gases, the presented non-convex gradient
theory catches the peaks which can be directly related to lower and upper yield phenomena observed experimentally during the formation of microstructures in metals,
as in the case of Luders

band formation and movement (e.g. Hahner (1994), Sun et al.


(2003), Halim et al. (2007), Yoshida et al. (2008)), considering the length scale in the
presented examples. The upper yield point (the peak in the stress-strain curve) can
be related to the formation of a microstructure (e.g. as an appearing Luders

band)
and the plateau can be linked to evolution or the movement of the band. However,
the relaxed stress obtained in the convexification approach represents only the perfectly plastic response.
Regarding the distribution of the plastic slip, in case of both hard and soft boundary conditions at low and at high strain levels (outside the binodal region) the same
stable behavior as in the case of convex strain gradient plasticity (see example 1) is
logically recovered. Hard boundary conditions typically trigger slip gradients at the
ends of the bar, resulting in characteristic boundary layers. Soft boundaries on the
other hand induce an initially homogeneous distribution of plastic slip. At strain
levels corresponding to the levels where the stress plateau is observed (between the
spinodal points) pronounced slip patterns develop depending on the boundary conditions (see Fig. 4.5 (b) and 4.6 (b)). This is consistent with the expectations based on
phase field modeling on the one hand, and the incremental minimization procedure
on the other hand. What is new with respect to the latter, is the fact that transitory
regimes are obviously captured as well, highlighting the role of the rate dependent
character of the model and the spinodal characteristic of the free energy. The stress
drop in Fig. 4.6 (a) is essentially due to the sudden (rate dependent) loss of stability
in the spinodal regime. This drop is sharp for the considered idealized case, and it
is not expected to occur if a more gradual loss of stability is invoked through the
intrinsic statistically non-homogeneous nature of the material.

65

4.6 Numerical examples

The analysis in the present paper concerns single laminates only. Multiple laminates can be observed as in the energy-convexification methods, e.g. Lambrecht
et al. (2003) if a spatial fluctuation or random distribution is given to the material
parameters in the model. Departing from such fluctuations the obtained laminates
would coarsen in time, depending on the stabilizing gradient term in the free energy. Changes of the surface (gradient) term or introduction of higher gradient terms
(as in Cahn-Hilliard models) could give different response to a random fluctuations,
i.e. laminates may become more pronounced and/or stabilize at a particular size,
depending on the stabilizing (gradient) term in the slip evolution equation.
4000

1500

450

3500

400
350

2000
1500

1000

Stress (MPa)

2500

Stress (MPa)

Stress (MPa)

3000

500

1000

200
150

50
0.005

0.01

0.015

0.02

0
0

0.025

Strain

0.005

0.01

0.015

0.02

0
0

0.025

(b) = 200 s1.


350

300

300

250

250

Stress (MPa)

250
200
150
100
50
0.01

0.015

Strain

(d) = 2 s1 .

0.02

0.025

Stress (MPa)

350

300

200
150

100
50
0.01

0.015

0.025

0.02

0.025

150

50
0.005

0.02

200

100

0
0

0.015

(c) = 20 s1 .

350

0.005

0.01

Srain

400

0
0

0.005

Strain

(a) = 2000 s1.

Stress (MPa)

250

100

500
0
0

300

0.02

Strain

(e) = 0.2 s1 .

0.025

0
0

0.005

0.01

0.015

Strain

(f) = 0.02 s1 .

Figure 4.7 / Rate dependent stress vs. global strain response.

The examples in Figs. 4.7 and 4.8 deal with the dependence of stress vs. strain response and the microstructure (plastic slip) evolution on the loading rate under soft
boundary conditions. A high loading rate (i.e. 2 s1 ) tends to inhibit the microstructure evolution leading to a homogenous distribution of plastic slip. The stress vs.
strain response corresponds to the steady state analytical solution presented in section 4.6.2. Higher loading rates (i.e. 20 s1 , 200 s1 , 2000 s1 ) result in a homogeneous plastic slip distribution with a more stiff stress vs. strain response, approaching the elastic limit behavior. Lower loading rates (i.e. 0.2 s1 , 0.02 s1 ) typically
result in patterned plastic slip distributions, which implies that the patterns evolve
only close to the rate independent limit. The slip pattern development as a function

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

66

of the deformation rate is shown in Fig. 4.8. The deformation is initially homogeneous and the slip patterns after passing the first spinodal point. This slip patterning evolves further with time and finally vanishes after passing the second spinodal
point. The transition from the homogeneous to the inhomogeneous slip distribution
is clearly rate dependent. For low loading rates (Fig. 4.8(b)), the transition from
the homogeneous to the patterned state (after the first spinodal point) and from the
patterned to the homogeneous state (after the second spinodal point) appears considerably faster compared to higher loading rates (Fig. 4.8(a)). This is most obvious
at an overall displacement of 7.2 m and a displacement equal to 14.8 m, where the
different loading rate cases clearly reveal different regimes.

u = 20m [t=0.1 s]

u = 20m [t=1 s]

u = 14.8m [t=0.074 s]

u = 14.8m [t=0.74 s]

u = 13.6m [t=0.068 s]

u = 13.6m [t=0.68 s]

u = 7.8m [t=0.0392 s]
u = 7.8m [t=0.392 s]
u = 7.2m [t=0.36 s]

u = 7.2m [t=0.036 s]

0.2

0.4

0.6

u = 4m [t=0.02 s]

u = 4m [t=0.2 s]

u = 0 [t=0]

u = 0 [t=0]

0.8

(a) evolution at = 0.2 s1 .

0.2

0.4

0.6

0.8

(b) evolution at = 0.02 s1 .

Figure 4.8 / Rate dependent microstructure evolution for the same imposed displacement.

4.6.3 Numerical example 3: non-convex stress relaxation of a 1D bar


In the final example, we illustrate the evolution of slip patterning in a stress relaxation test, again including both hard and soft boundary conditions. The displacement at x = 0 is suppressed and a displacement at x = L is applied with a rate of
= 0.02 s1 until the homogeneous plastic slip reaches a value between the spinodal
points. Next, the displacement at x = L is kept constant, leading to relaxation of
the stress in the bar. In this way, one can observe the evolution of the plastic slip
microstructure at a constant (macroscopic) average strain level, accompanied by relaxation of the stress.
The material parameters and the plastic slip dependent non-convex free energy are

67

4.6 Numerical examples

(a) Stress vs. global strain.

0.2

0.4

0.6

0.8
0.2

(b) Plastic slip evolution.

Figure 4.9 / Stress relaxation test: stress vs. strain response and plastic slip evolution
with hard boundary conditions.

(a) Stress vs. global strain.

0.2

0.4

0.6

0.8

(b) Plastic slip evolution.

Figure 4.10 / Stress relaxation test: stress vs. strain response and plastic slip evolution with soft boundary conditions.

the same as used in section 4.6.2. The relaxation starts immediately after the displacement yielding an average overall strain equal to 0.0049 has been reached for the hard
boundary case and 0.0054 for the soft boundary case. The stress vs. strain response
and the evolution of the plastic slip are presented for hard and soft boundaries in
Figs. 4.9 and 4.10, respectively. Note that the soft boundary conditions lead to an initially homogeneous distribution, for which any perturbation may trigger patterning.
In order to stabilize this a small spatial fluctuation is applied to the Youngs modulus E along the bar, which restores uniqueness and triggers a stable evolution of the

68

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

microstructure. The use of a small initial defect or a perturbation is also commonly


done within the analysis of non-local damage models in 1D systems.
In both relaxation and monotonic tension tests with soft boundaries the full binodal
decomposition of the plastic slip is observed, because of the fact that the slip is not
constrained contrary to the hard boundary case where the full decomposition of the
field is naturally prohibited.

4.7 Conclusion
Inspired by the efficiency of phase field models for microstructure formation and
evolution, we developed a non-convex strain gradient plasticity model in a system with energetic hardening-softening, which takes a conceptually dual structure
to Ginzburg-Landau type of phase field models with high and low slipped regions
representing different phases. Focusing on the basics and simplicity, the derivation
and the implementation are conducted in a 1D setting in order to illustrate the ability
of the model to capture the patterning of plastic slip similar to a phase decomposition mechanism. The destabilizing non-convex term in the free energy is stabilized
through the gradient term in the free energy and the viscous nature of the thermodynamically consistent slip law. Note that the resulting model typically yields a rate
dependent microstructure evolution. The framework can capture both homogeneous
and inhomogeneous deformations depending on the rate of the applied deformation.
The model is conceptually capable of covering the entire microstructure evolution
process, depending on the externally applied load and boundary conditions. Nonequilibrium microstructures are thereby well at reach.
The non-convexity in the presented model is incorporated by a double-well function,
not addressing particular materials yet. We present a generic formulation in order to
demonstrate the microstructure evolution in a thermodynamical and mathematical
rigorous setting. A more practical emphasis applied to a particular material is obviously needed in future work, provided reliable experimental data to recover the
non-convex term are available.

4.8 Appendix
4.8.1 Finite element implementation of slip based strain gradient plasticity
The weak forms of the equations are obtained in a standard manner, using a Galerkin
procedure. Firstly, the balance of linear momentum is tested with a field of virtual

69

4.8 Appendix

displacements u and integrated over the domain 0 < x < L, which results in Gu ,
Gu =

dx = 0
x

(4.34)

The slip equation (4.24) is tested with a field of virtual slips and integrated towards
G
m



Z L 
n
n
0m
0m 2

G =


sign

A
dx = 0 (4.35)
t
t
s
s x2
0
Using integration by parts equation (4.34) can be written as
Gu =

L
0

du
dx + [u ]|0L = 0
dx

Similarly, equation (4.35) can be rewritten as







n m
RL
n
d( )


G = 0
sign
Ca + Cb
dx
t
t
dx x

Cb [ ]|0L = 0
x

(4.36)

(4.37)

with the abbreviations Ca = 0m /s and Cb = 0m A/s.


The variational functionals Gu and G are solved in a fully coupled manner (monolithic) by means of a Newton-Raphson scheme. For this reason Gu and G are linearized with respect to the variations of the primary variables u and .
LinGu = u Gu + Gu + Gu = 0
LinG = G + u G + G = 0

(4.38)

where Gu and G stand for the values of the previous estimate and with
u Gu =
Gu =
G

R L du

dx

dx

R L du

dx
dx
m 



RL
n n

= 0 sign
m
dx
t
t
n

RL
0

u G =

C a

RL
0

d( )
dx + 0L Cb
dx

dx x

C a

dx

(4.39)

70

4 Deformation patterning driven by rate dependent non-convex strain


gradient plasticity

The equations (4.38) are discretized according to a finite element approach. Linear
interpolation functions N are used for the slip field , and quadratic interpolation
functions N u for the displacement field u.

u = J N JuuJ

u = K NKu uK
(4.40)

L NL L

M N M M

Substitution of the finite element interpolations into the linearized forms results in
the global system of coupled linear equations with element tangent matrices,
R

kuu =
k u
k

dN Ju dNKu
dx
Be J K
dx dx

dN Ju
= Be J M
N dx
dx M
m 





R
n n
m

N M dx
= B e L M NL sign
t
t
n


R

dNL dN M

B e L M N L Ca N M
Cb
dx

dx
dx
R

k u =

Be

L K N L Ca

(4.41)

dNKu
dx
dx

where / = E, / = E due to = E( ) and B e is the element domain. Element residual vectors are calculated by using the values of and from
the previous estimate
dN Ju
dx
Be J
dx

m





R
n
n
sign
dx
= B e L M NL N M
t
t



R
dNL dN M

M
B e L NL Ca L M
Cb

dx
dx
dx

ru =
r

(4.42)

The assembly operation gives the global tangent and residual i.e.,
K uu = A {kuu }

K u = A { k u }

R u = A {r u }

K u = A {k u }

K = A {k }

R = A {r }

and the system of equations to be solved reads:


"
# "
#
"
#
Ru + Rext
K uu K u
u
u
=
K u K

R + Rext

(4.43)

(4.44)

ext
where Rext
u and R originate from the boundary terms in the equilibrium (4.36) and
slip (4.37) weak forms.

71

4.8 Appendix

4.8.2 Finite element implementation of slip based non-convex strain gradient plasticity
Pursuing the same procedure as in the previous section, the weak form for equilibrium is written as
Gu =

L
0

du
dx [u ]|0L = 0
dx

and for the plastic slip evolution as







n m
RL

d
(

sign
G = 0
Ca + Cb
dx
t
t
dx x
R

+ 0L [ Ca (4C1 3 + 3C2 2 + 2C3 + C4 )] dx Cb [ ]|0L = 0


x

(4.45)

(4.46)

The variational functionals Gu and G are again solved in a fully coupled manner
(monolithic) by means of a Newton-Raphson scheme. The corresponding element
tangent matrices are identical to (4.41), except the k term which becomes,
m 





R
n n
m

k = B e L M NL sign
N M dx
t
t
n


R

dNL dN M
Cb
B e L M NL Ca N M

(4.47)

dx
dx


2
+ NL N M Ca (12C1 + 6C2 + 2C3 ) dx
and r
r =

Be

Be

Be





n m

sign n
dx
t
t



dNL dN M

M
Cb
dx
L NL Ca L M
dx
dx



3
2
L NL Ca (4C1 + 3C2 + 2C3 + C4 ) dx



L M NL N M

(4.48)

which are calculated for the values of and from the previous estimate. After assembly the resulting system of equations in (4.44) is solved in a standard incremental
iterative manner.

72

Chapter five

Non-convex rate dependent strain


gradient crystal plasticity and
deformation patterning1

Abstract / A rate dependent strain gradient crystal plasticity framework is presented


where the displacement and the plastic slip fields are considered as primary variables.
These coupled fields are determined on a global level by solving simultaneously the linear momentum balance and the slip evolution equation, which is derived in a thermodynamically consistent manner. The formulation is based on the 1D theory presented in
Yalcinkaya et al. (2011a), where the patterning of plastic slip is obtained in a system with
non-convex energetic hardening through a phenomenological double-well plastic potential. In the current multi-dimensional multi-slip analysis the non-convexity enters the
framework through a latent hardening potential presented in Ortiz and Repettto (1999)
where the microstructure evolution is obtained explicitly via lamination procedure. The
current study aims the implicit evolution of deformation patterns due to the incorporated
non-convex potential.

5.1 Introduction
At the microscopic scale, deformed crystalline materials usually show heterogeneous
plastic deformation, where the amount of plastic strain varies spatially. At moderate
strain levels, regular cellular dislocation structures have been observed. Typical examples of dislocation microstructures are dislocation cells and dislocation walls (see
e.g. Rauch and Schmitt (1989), Gardey et al. (2005), Yalcinkaya et al. (2009)). Patterning typically refers to the self organization of dislocations, yielding regions with a
high dislocation density (dislocation walls) that envelop areas with a low dislocation
1

This chapter is reproduced from Yalcinkaya et al. (2011b)

73

74

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

density (dislocation cell interiors), also regarded as domains of high and low plastic
slip activity, respectively. In addition to the cellular microstructures at meso-scale,
clear band formation and related plastic flow localization in irradiated materials (see
e.g. Sauzay et al. (2010)) at lower scales and macroscopic plastic slip bands such as
Luders

bands (see e.g. Shaw and Kyriakides (1998)) are also commonly observed
structures due to plastic deformation. These microstructures macroscopically e.g.
manifest themselves through softening of the material or through plastic anisotropy
under strain path changes.
The softening of the material and macroscopic anisotropic effects under strain path
changes (see e.g. Peeters et al. (2000), Yalcinkaya et al. (2009)), resulting from evolving dislocation microstructures, have been an interesting topic for the materials science community and the metal forming industry for decades. Starting with the studies on the cold-worked sub-structure of polycrystals using transmission electron microscopy in 1960s (e.g. Bailey and Hirsch (1960), Keh et al. (1963) and Swann (1963)),
a vast amount of experimental results have been collected, and several promising
theoretical models were presented dealing with dislocation (or slip) patterning. Nevertheless, a complete descriptive understanding of the occurring phenomena has not
been reached and the necessary input for computational models is still subject of
ongoing discussions.
In the context of the computational modeling of plastic slip pattering (or dislocation sub-structure formation), different approaches have been pursued in the literature which can be categorized into three main groups: (i) models using directly
the mechanics of single dislocations or populations of dislocations, (ii) phase field
modeling of dislocation patterning, (iii) the incremental variational formulation of
inelasticity by applying relaxation concepts for fine scale microstructure evolution.
See Yalcinkaya et al. (2011a) for a global overview of the available approaches.
The objective of the present paper is to develop a rate dependent strain gradient crystal plasticity finite element framework for the simulation of dislocation microstructure evolution, where the non-convexity is treated as an intrinsic property of the plastic free energy of the material. The constitutive model aims to simulate deformation
patterning and the macroscopic material behavior in a thermodynamically consistent
manner. Hence, the influence of latent hardening on the dislocation microstructure
evolution is studied through a physically based latent hardening potential proposed
by Ortiz and Repettto (1999). The physically based non-convex formulation is relevant in multi-slip deformation states, accounting for the interaction of slip systems.
In the model, the plastic slip and the displacement are taken as degrees of freedoms.
These fields are determined on a global level by solving simultaneously the linear
momentum balance and the slip evolution equation. The latter, expressed through a
thermodynamically consistent slip law is a crucial part of the model. At first glance,

5.2 Strain gradient crystal plasticity and finite element implementation

75

the slip law (see equation (5.18)) is similar to the one encountered in classical rate
dependent crystal plasticity approaches (e.g. Hutchinson (1976), Peirce et al. (1982),
Yalcinkaya et al. (2008)), however the contribution of the actual stress state differs.
Three stress contributions can be distinguished: (1) the conventional resolved shear
stress directly related to the external loading, (2) an internal stress depending on the
gradient of the plastic slip, which is characteristic for strain gradient crystal plasticity models (e.g. Evers et al. (2004b), Yefimov et al. (2004), Bayley et al. (2006)) and
(3) the stress that emanates from the non-convex part in the free energy. The latter
contribution eventually triggers the patterning of the dislocation slip field.
The paper is organized as follows. First, in section 2, the rate dependent strain gradient crystal plasticity and its finite element solution is briefly discussed. Then, in
section 3, the incorporation of non-convexity into the model is presented, using a
physically based latent hardening potential. A detailed analysis of the latent hardening based non-convex function is performed in this section in order to clarify the
conditions enabling microstructure evolution. In section 4, numerical examples are
presented in order to demonstrate the capability of the proposed model. First, the
size effect related to plastic slip gradients and rate dependent deformation evolution
in the context of convex strain gradient crystal plasticity is studied. Then, the rate
dependent microstructure evolution via the physically motivated latent hardening
non-convex potential is addressed in this section. Finally, some concluding remarks
are given in section 5.

5.2 Strain gradient crystal plasticity and finite element implementation


In this section, the theoretical framework of the slip based strain gradient crystal
plasticity is presented and its incorporation into a finite element formulation is addressed briefly. First, the thermodynamical consistency and the derivation of the
governing system equations are discussed. In a geometrically linear context, with
small displacements, strains and rotations, the time dependent displacement field
is denoted by u = u( x, t), where the vector x indicates the position of a material
point. The strain tensor is defined as = 12 ( u + (u)T ), and the velocity vector
is represented as v = u.
The strain is decomposed additively as

= e + p

(5.1)

into an elastic part e and a plastic part p . The plastic strain rate can be written as
a summation of plastic slip rates on the individual slip systems, p = P with
P = 12 (s n + n s ) the symmetrized Schmid tensor, where s and n are the

76

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

unit slip direction vector and unit normal vector on slip system , respectively. The
state variables are chosen to be given by the set
state = e , ,

(5.2)

where contains the plastic slips on the different slip systems and represents
the gradient of the slips on these slip systems. Following the arguments of Gurtin
(e.g. Gurtin (2000), Gurtin (2002)), the power expended by each independent ratelike kinematical descriptor is expressible in terms of an associated force consistent
with its own balance. However, the basic kinematical fields of rate variables, namely
e , u and are not spatially independent. It is therefore not immediately clear how
the associated force balances are to be formulated, and, for that reason, these balances
are established using the principle of virtual power.
e
Assuming that at a fixed time the fields u, e and are known, we consider u,
and as virtual rates, which are collected in the generalized virtual velocity V =
e , ). Pext is the power expended on the domain and Pint a concomitant
(u,
expenditure of power within , given by

Pext (, V) =
Pint (, V) =

t (n ) u dS +

e d +
:

( (n) ) dS

( )d +

(5.3)

( )d

where the stress tensor , the scalar internal forces and the microstress vectors
are the thermodynamical forces conjugate to the internal state variables e ,
and , respectively. In Pext , t (n ) is the macroscopic surface traction while
represents the microscopic surface traction conjugate to at the boundary S with n
indicating the boundary normal.
The principle of virtual power states that for any generalized virtual velocity V the
corresponding internal and external power are balanced, i.e.

Pext (, V) = Pint (, V)

(5.4)

Considering a generalized virtual velocity without slip, the virtual velocity field can
be chosen arbitrarily and this leads to the classical macroscopic force balance,

= 0

(5.5)

and considering arbitrary virtual slip fields without a generalized virtual velocity
leads to the microscopic force balances,

+ = 0

(5.6)

5.2 Strain gradient crystal plasticity and finite element implementation

77

on each slip system , where is the resolved Schmid stress given by = : P .


The local internal power expression can be written as
Pi = : e + ( + )

(5.7)

The local dissipation inequality results in


D = Pi = : e + ( + ) 0

(5.8)

The material is assumed to be endowed with a free energy with different contributions according to

= e + +

(5.9)

The time derivative of the free energy is expanded and equation (5.8) is elaborated
to
e


)
e :


(5.10)
de

= ( e ) : e + ( ) + (
) 0

| {zd }
|
{z
}

D = : e + ( +

The stress and the microstress vectors are regarded as energetic quantities having no contribution to the dissipation

de
d e

(5.11)

whereas does have a dissipative contribution,


D=

( ) 0

(5.12)

The multipliers of the plastic slip rates are identified as the set of dissipative stresses
dis

dis =

(5.13)

In order to satisfy the reduced dissipation inequality at the slip system level the following constitutive equation is proposed

dis
= sign ( )

(5.14)

78

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

where represents the mobilized slip resistance of the slip system under consideration
s
= | |
(5.15)
0
with s is the resistance to dislocation slip which is assumed to be constant and 0 is
the reference slip rate. Substituting (5.15) into (5.14) gives

s dis

Substitution of dis
according to (5.13) into (5.16) reveals,


0

=
s

Using the microforce balance (5.6) results in the plastic slip equation


0

= +
s

(5.16)

(5.17)

(5.18)

In addition to the explicit contribution of , other contributions of the free energies


defined in (5.9) enter the slip equation via (5.11) with = de /de : P and =
/ . Quadratic forms are used for the elastic free energy e and the plastic
slip gradients free energy contribution , i.e.
1
e = e : 4 C : e
2

(5.19)

1
A
2

where A is a scalar quantity, which includes an internal length scale parameter, governing the effect of the plastic slip gradients on the internal stress field. It may be
expressed as A = ER2 /(16(1 2 )) as e.g. used in Bayley et al. (2006) and Geers
et al. (2007), where R physically represents the radius of the dislocation domain contributing to the internal stress field, is Poissons ratio and E is Youngs modulus.
The plastic slip dependent free energy will be defined in the following section.
In order to solve the initial boundary value problem for this rate dependent strain
gradient crystal plasticity framework, a fully coupled finite element solution algorithm is used in which both the displacement u and plastic slips are considered
as primary variables. These fields are determined in the solution domain by solving
simultaneously the linear momentum balance (5.5) and the slip evolution equation
(5.18), which constitute the local strong form of the balance equations:

= 0

0 0
0

+
s
s
s

= 0

(5.20)

5.2 Strain gradient crystal plasticity and finite element implementation

79

In order to obtain variational expressions representing the weak forms of the governing equations given above, these equations are multiplied by weighting functions u
and and integrated over the domain . Using the Gauss theorem (S is the boundary of ) results in
Gu =
G

u : d

u tdS
0
d +
s

0
d
s

0
A d
s

(5.21)

dS

0
where t is the external traction vector on the boundary S, and = A n.
s
The domain is subdivided into finite elements, where the unknown fields of the
displacement and slips and the associated weighting functions within each element
are approximated by their nodal values multiplied with the interpolation shape functions stored in the N u and N matrices, using a standard Galerkin approach.

u = N u u

u = Nuu

= N

(5.22)

= N

with u, u , and are columns containing the nodal variables. Bilinear interpolation functions for the slip field and quadratic interpolation functions for the displacement field are used. An implicit backward Euler time integration scheme is
used for in a typical time increment [tn , tn+1 ] which gives = [n+1 n ]/t.
The discretized element weak forms read

Z
Z
T
u
u
e
e
e
Gu = u
B d
N t dS
e

G e

"Z

Z

Se

n+1 n
t

0
A B T B de +

0 T
N de
s

0 T
N
de
s

Se

(5.23)

dSe

The weak form of the balance equations (5.23) are linearized with respect to the variations of the primary variables u and and solved by means of a Newton-Raphson
solution scheme for the increments of the displacement field u and the plastic slips
. The procedure results in a system of linear equations which can be written in
the following matrix format,
" uu
# "
#
"
#
u
Ru + Rext
K
K u
u
=
(5.24)

K u K
R + Rext

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

80

where K uu , K u , K u and K represent the global tangent matrices while Ru and R


ext
are the global residual columns. The contributions Rext
u and R originate from the
boundary terms.

5.3 Latent hardening based non-convex plastic potential


In this section, a latent hardening based non-convex plastic potential proposed by
Ortiz and Repettto (1999) is examined, whereby the conditions for the occurrence of
plastic slip patterning is studied.
In physically deformed crystals one of the main presumed reasons for dislocation
microstructure formation, is latent hardening accompanied with non-convexity of
the energy function due to slip system interactions. Such a function is proposed by
Ortiz and Repettto (1999) which gives parabolic-like hardening in single slip and
latent (off-diagonal) hardening in multi-slip (see Fig. 5.1 for two slip systems)
2
= 0 0
3

"

| | | |
a 0 0

# 3/ 4

(5.25)

Here 0 and 0 are a reference resolved shear stress and a reference slip value, respectively, and a are interaction coefficients. For the values of the matrix a , a simple
geometrical model is used proposed by Cuitino and Ortiz (1993),

2
=

1 (n n )2

(5.26)

where n and n are normals on the slip planes of the systems considered. Using
(5.26), slip systems do not self-harden in a multi-slip context. The reasoning leading
to (5.26) is based on the fact that the typical resolved shear stress required to deform a well-annealed crystal in single slip tends to be small compared to the stress
required for multi-slip. For the purpose of understanding the morphology of dislocation structures, self-hardening can be neglected at first instance. Ortiz and Repettto
(1999) and Ortiz et al. (2000) employ (5.25) and (5.26) in the context of crystal plasticity to obtain lamellar dislocation structures via a sequential lamination procedure.
The purpose here is to study patterning driven by this latent hardening based nonconvex potential.

5.3.1 Conditions for plastic slip patterning


Plastic deformation tends to generate dislocation microstructures that minimize the
free energy (see e.g. Kuhlmann-Wilsdorf (2001)). From a thermodynamics point of

81

5.3 Latent hardening based non-convex plastic potential

0.012

0.01

0.012
0.01

0.008

0.008
0.006
0.004

0.006

0.004

0.002

0.002

0
0.06
0.06

0.04
0.04
0.02

0.02
0

0
0.06

0.04

0.02

0.06

0.05

0.04

0.03

0.02

0.01

Figure 5.1 / Two different views of (MPa) for different amounts of slip on two
slip systems oriented with respect to x axis as 60 and 120 where 0 = 1 and 0 = 1
MPa.

view, a patterned microstructure will develop if it has a lower energy than a state
with a homogeneous plastic deformation distribution. Yalcinkaya et al. (2011a) studied this aspect on a double-well potential and showed that plastic slip patterning
occurs at relatively low strain rates due to the existence of an unstable regime in
the double-well potential, where the microstructure evolves in a patterned way to
lower its free energy. Convex energy potentials preserve stability and do not trigger
patterning. Therefore, the presence of non-convexity is the first condition for a heterogeneous microstructure evolution to develop, yielding a lower energy than the
homogeneous state. In what follows, the latent hardening function (5.25), which is
assumed to be the driving force for the evolution of dislocation microstructures, is
examined in this sense.
For this purpose, a single crystal in 2D having 2 slip systems oriented 1 and 2
with respect to x axis is considered. A pure shear deformation is applied with a
macroscopic strain field M , whose components are written in Cartesian coordinates
as
"
#
0

M =
(5.27)

0
which satisfies the plastic incompressibility condition tr ( M ) = 0. Assume that the
plastic deformation is patterned in the solution domain and that there exist two states
with volume fraction f and 1 f . The weighted average strain of the two phases
should be equal to the macroscopic strain, i.e.
f 1 + (1 f ) 2 = M

(5.28)

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

82

The contributing strains 1 and 2 are calculated according to,

1 = 2 =1 1 P
2 = 2 =1 2 P

(5.29)

with both patterned regions having the same crystal orientation and hence the same
P . The strain tensors, 1 = 1 (11 , 12 , 1 , 2 ) and 2 = 2 (21 , 22 , 1 , 2 ) are symmetric
and incompressible. Using (5.29), equation (5.28) reduces to a set of 2 linear equations with 5 unknowns 11 , 12 , 21 , 22 , f . For a fixed value of f and given two of the
unknown slips 11 and 12 the value of the other plastic slips 21 and 22 can be calculated according to

21 =
22

f
P21
1
11 +

1 f
P22 P11 P12 P21 1 f

1
f
P11
=
12 +

1 f
P22 P11 P12 P21 1 f

(5.30)

with P11 = P1 (1, 1), P12 = P1 (1, 2), P21 = P2 (1, 1), P22 = P2 (1, 2). The energy of this
assumed patterned state can be calculated as,

= f 1 (11 , 12 , 1 , 2 ) + (1 f )2 (21 , 22 , 1 , 2 )

(5.31)

which is plotted in Fig. 5.2. The latent hardening energy corresponding to the macroscopic shear strain M is calculated via (5.25) in terms of the plastic slips 1M and 2M
on given slip systems as,

= ( 1M , 2M , 1 , 2 )

(5.32)

The purpose of this analysis is to investigate if there exists a domain of a lower latent
hardening based energy for the assumed patterned case than the macroscopic
latent hardening based energy (which reflects the homogeneous non-patterned
state). To this end, is calculated for different values of 11 and 12 , and for specific
(chosen) orientations and volume fractions.
If a macroscopic shear deformation tensor M with = 0.02 is applied on a combinations of slip systems with orientations 1 = 60 and 1 = 120 , the amount of
slips on the two slip systems for the local homogeneous state equals 0.04. Using
these values for the slips and the orientations, the locally homogeneous latent hardening plastic potential is calculated via Equation (5.25). Assuming 0 = 1 MPa and
0 = 1 the characterizing latent hardening energy for the non-patterned state equals
= 0.0057 MPa. Presuming the existence of patterned states for different values of
f and (11 , 12 ), the values of 21 and 22 are calculated via (5.30) and the latent hardening energy follows from (5.31). In order to obtain a patterned microstructure the

83

5.3 Latent hardening based non-convex plastic potential

f = 0.1

x 10

f = 0.2

x 10

x 10

x 10

6.4

6.5
6

6.2
6

6.5
6

5.8

5.6

5.5

5.5

5.8
5.6

5.4

5.4

4.5
5.2

5.2

4.5
4

5
5
0.1

0.08

0.06

11

0.04

0.02

0.05

0.1

0.1

12

f = 0.3

x 10

11

0.05

x 10

12

0.1

x 10
12

0.014

7.5
7

0.05

f = 0.5

0.012

10

6.5
0.01

5.5
5

0.008

3.5

4.8

0.006
4.5
4

0.004

3.5

0.002

3
0

2
0.1

11

0.05

12

0.05

0.1

f = 0.6

2.5

0.1

11

0.05

16

0.04 2 0.06
1

0.08

0.1

f = 0.7

x 10

0.018

0.02

0.022

0.025

0.02

0.016
14

0.018

0.02

0.014
12

0.016

10

0.01

0.008

0.006

0.014

0.015

0.012

0.012

0.01

0.01

0.004

0.008
0.006

0.005

0.004

0.002
2
0

0.002

0
0.1

11

0.05

0.05

0.1

0.1

12

11

0.05

f = 0.8

0.05

12

0.1

f = 0.9
0.05

0.045

0.045

0.04

0.03

0.035
0.03

0.025

0.025

0.04

0.035

0.035
0.03

0.02

0.02
0.015

0.015
0.01

0.01

0.03
0.025

0.025
0.02

0.02

0.015

0.015

0.01
0.005

0.005

0.01

0.005
0.005

0
0.1

11

0.05

0.05

0.1

12

0.1

0.05

11

0.05

0.1

12

Figure 5.2 / (MPa) in terms of a set of given 11 and 12 for different values of f
with 1 = 60 and 1 = 120 .

84

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

energy of the patterned state should be smaller than the energy of the local homogeneous state < . As seen in Fig. 5.2 there are many combinations satisfying this
inequality for different values of the volume fraction f and the plastic slip values 11
and 12 .

5.4 Numerical analysis


In this section, two different numerical examples are presented to study the behavior
of the proposed rate dependent strain gradient crystal plasticity models. The first
example deals with a conventional convex free energy in terms of plastic slip gradients and elastic strains, where the effect of the applied shear rate and the internal
length parameter R are analyzed. Then, the influence of the physically based nonconvex latent hardening plastic potential on the mechanical behavior of the material
is discussed.

5.4.1 Convex strain gradient crystal plasticity


In this subsection, the plastic potential is assumed to be zero therefore having no
influence on patterning and the hardening of the material. The incorporated hardening is only due to the gradients of the plastic slip. This is referred to as convex
strain gradient crystal plasticity because there is no plastic potential inducing a lack
of convexity.
To reveal the main characteristics of this convex strain gradient viscous crystal plasticity model, a constrained plane strain shear problem of an infinite strip, induced by
periodic boundary conditions, is studied in the first example. A strip, with height H
(in y direction) is bounded by rigid walls that are impenetrable for dislocations, i.e.
a no slip condition applies ( = 0) at top and bottom edges. These so-called micro
clamped boundary conditions for the plastic slips invoke an inhomogeneous plastic
deformation state (e.g. as present near grain boundaries in polycrystals or at the surface of a thin film). The displacements at y = 0 are suppressed (ux = 0 and u y = 0)
and prescribed at y = H as ux = u(t) and u y = 0. In addition, in x direction all field
quantities are taken to be independent of x. Consequently, the field quantities on the
left side are assumed to be identical to those on the right side ul = ur and l = r .
Locally, two slip systems with orientations 60 and 120 with respect to the horizontal axis are considered. The material is assumed to be elastically isotropic with
Youngs modulus E = 210 GPa, Poissons ratio = 0.33, slip resistance for both slip
systems s = 35 MPa, and a reference slip rate 0 = 0.15 s1 . Results presented in
the following correspond to a discretization with 1x100 rectangular elements. One

85

5.4 Numerical analysis

element in the x-direction is sufficient, given the uniformity in this direction.


= 2.5s1

900

= 0.25s1

140

800

120

Shear stress [MPa]

Shear stress [MPa]

700
600
500
400
300

100

80

60

40

200
20

100
0

0.005

0.01

0.015

0.02

Applied shear = u/H

3.5

12

Shear stress [MPa]

Shear stress [MPa]

0.015

0.02

= 0.0025s1

14

10
8
6
4
2
0

0.01

Applied shear = u/H

= 0.025s1

16

0.005

2.5
2
1.5
1
0.5

0.005

0.01

0.015

Applied shear = u/H

0.02

0.005

0.01

0.015

0.02

Applied shear = u/H

Figure 5.3 / Rate dependent shear stress vs. the applied shear for the plane strain
shear problem of an infinite strip.

The examples addressed in this subsection are carried out for different applied shear
rates and R/ H ratios for varying R and constant height H, where the relative effect
of the internal length scale parameter R determining A, acting on the higher order
microstresses (see equations (5.11) and (5.19)), is analyzed.
First, the resulting shear stress versus the applied macroscopic shear is presented
in Fig. 5.3 for different overall shear rates , using R = 0.35 m, and H = 20 m.
is the macroscopic shear defined as = u/ H. In this case, the average of the local
strain 12 should be half of the macroscopic shear . If the elastic deformation is
p
small, the average value of the local plastic shear strain 12 will be roughly equal to
half of the macroscopic shear as well. When the applied macroscopic shear rate
increases, the material shows a stiffer (dominantly elastic) response. In Fig. 5.4, the
local rate dependent plastic shear strain evolution is presented at a macroscopic shear
level of = 0.02. There is a clear boundary layer width dependence on the applied
shear rate. For high values of the applied shear rate, e.g. = 2.5s1 we observe
a sharp boundary layer, while for low values of the shear rate, e.g. = 0.0025s1

86

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

the boundary layer thickness is more diffuse. This type of rate dependent dislocation
slip profile is also observed in Yalcinkaya et al. (2011a) where a high strain rate causes
that the plastic slip has not enough time to evolve.
= 2.5s1

= 0.25s1
3

x 10
4.5

x 10
9

4
8
3.5
7
3

2.5

1.5

0.5

= 0.025s1

= 0.0025s1
3

x 10
12

x 10

10

12

14

10

8
6
6
4
4
2

Figure 5.4 / Rate dependent plastic shear strain 12 distribution for the plane strain
shear problem of an infinite strip at a global shear level of = 0.02.

The next example in this subsection concerns the influence of the internal length scale
parameter R on the mechanical behavior for a fixed height H = 20 m. The value of R
is taken as 0.35 m, 0.7 m and 1.75 m corresponding to a R/ H ratio equal to 0.0175,
0.035 and 0.0875 respectively. The shear stress versus the applied macroscopic shear
response is plotted in Fig. 5.5 for different values of R at = 0.025s1 . Note that
the formulation is intrinsically viscous, corresponding to models using a linear drag
law for dislocation motion to determine the slip. Fig. 5.5 illustrates a significant
effect of the internal length parameter where the strip is exhibiting a stiffer response
for larger values of R. The results are consistent with many strain gradient models
(e.g. Shu et al. (2001), Evers et al. (2004b), Yalcinkaya et al. (2011a)) in which the
influence of the length scale was shown. Physically, this effect results from the fact
that large R values induce a large internal stress and hence penalize high plastic slip

87

5.4 Numerical analysis

gradients (see Fig. 5.6) spreading the geometrically necessary dislocation densities.
In this convex example, the slip gradient is the dominating factor in the plastic and
hardening behavior of the material and therefore a clear size effect is observed. In
Fig. 5.6 a clear dependence of the boundary layer evolution on R is presented at
= 0.02. With increasing R an increased influence of the boundary conditions is
observed with a more diffuse boundary layer.

70
R = 0.35 m
R = 0.7 m
R = 1.75 m

Shear stress [MPa]

60

50

40

30

20

10

0.005

0.01

0.015

0.02

Applied shear = u/H

Figure 5.5 / The effect of the internal length scale parameter R on the stress vs.
applied shear response for the plane strain shear problem of an infinite strip.

R = 0.7m

R = 0.35m

x 10
12

R = 1.75m

x 10

x 10
14

12
10

10
8

12
10
8

6
4
2

Figure 5.6 / The effect of the internal length scale parameter R on the plastic shear
p
strain 12 distribution for the plane strain shear problem of an infinite strip at a global
shear level of = 0.02.

The distribution of the plastic slips on each of the slip systems is illustrated in Fig. 5.7
for the same example with R = 0.35 m and = 0.025s1 at = 0.02. The amount

88

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning
= 60

= 120

0.0121

0.0121

0.0242

0.0242

Figure 5.7 / The distribution of the plastic slip on both slip systems at a global shear
level of = 0.02.

of the slip on both slip systems is identical and the evolution is similar to the plastic
shear strain evolution, exhibiting a boundary layer.

5.4.2 Non-convex strain gradient crystal plasticity


In this subsection the influence of the latent hardening plastic potential (5.25) on
the mechanical behavior and the deformation patterning is illustrated in the present
rate dependent strain gradient crystal plasticity framework. In addition to , an
elastic strain energy potential e and a plastic slip gradient free energy potential
are incorporated as well. The viscous formulation of the problem and the gradient
free energy potential regularize the problem.
The paper of Ortiz and Repettto (1999) shows that in crystals exhibiting latent hardening the energy function is non-convex, which favors the development of microstructures. Therefore, uniform deformation fields are not the minimizers of the incremental work of deformation. In other words, in the context of classical variational
formulations the crystals exhibiting latent hardening might not reach the expected
solution, i.e. the minimum free energy is not achieved. It is possible to construct
deformation mappings to recover the minimum value. Such deformation mappings
make use of the existence of fine microstructures. Using a sequential lamination
method, Ortiz and Repettto (1999) were able to characterize analytically several dislocation structures. The same procedure is followed later in Ortiz et al. (2000) where
microstructures are regarded as instances of sequential lamination during deformation. The microstructures are explicitly constructed by recursive lamination and their
subsequent equilibration.
The numerical study in section 5.3 proves that the latent hardening potential (5.25)
is non-convex and satisfies the energetic conditions for plastic slip phase separation

5.4 Numerical analysis

89

in the context of the present small strain rate dependent crystal plasticity formulation. Compared to the previous work of Ortiz and Repettto (1999) and Ortiz et al.
(2000) relying on the explicit construction of cellular dislocation microstructures via
a lamination procedure, the present framework is based on the implicit evolution of
microstructures driven by the deformation. Throughout the incremental deformation process the state of the plastic slip enters energetically favorable regimes, (as
illustrated in section 5.3) which might eventually result in deformation heterogeneity.
The numerical study in this subsection concerns a plane strain pure shear problem of
a square representative volume element (RVE) in order to have a direct link with the
study in section 5.3. Locally, two slip systems with orientations 60 and 120 with
respect to the x axis are considered. The material is assumed to be elastically isotropic
with Youngs modulus E = 210 GPa, Poissons ratio = 0.33, slip resistance for both
slip systems s = 35 MPa, and a reference slip rate 0 = 0.15 s1 . The reference slip
strain and resolved shear stress in (5.25) are assumed to be 0 = 0.015 and 0 = 50
MPa, respectively.
In the square RVE the displacements and the plastic slips on the left edge are tied to
the ones on the right edge and the ones on the bottom edge are tied to the ones on the
top edge which makes it a fully periodic configuration. The vertical displacement at
the right bottom corner and the horizontal displacement at the left top corner are prescribed, both equal to u(t). The displacements at left bottom corner are suppressed
together with the horizontal displacement at the right bottom corner and vertical displacement at left top corner. These boundary and loading conditions result in a pure
shear deformation mode. The length of each edge of the square is H = 20 m. Results presented in the following correspond to a discretization with 20x20 rectangular
elements.
In Fig. 5.8 the shear stress versus applied macroscopic shear, defined as = 2u/ H,
is plotted for an applied shear rate = 0.005s1 and an internal length scale parameter R = 0.01 m. In this case, the average of the local strain 12 should be half of
the macroscopic shear . If the elastic deformation is small, the average value of the
p
local plastic shear strain 12 will be roughly equal to half of the macroscopic shear
as well. The corresponding shear strain, plastic shear strain and plastic slips of
the slip systems are plotted in Fig. 5.9. All plotted fields exhibit a strong patterned
response due to the incorporated latent hardening potential. Note that the applied
periodic boundary conditions lead to an initially homogeneous distribution of the
strain and plastic slip fields. In order to properly trigger a small perturbation a small
spatial fluctuation is applied to the Youngs modulus E in the RVE. In order to illustrate the influence of the selected spatial fluctuation of E on the evolution of the
plastic microstructure, two different types of fluctuations are applied in the next two

90

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning
100
90

Shear stress [MPa]

80
70
60
50
40
30
20
10
0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

Applied shear = 2u/H

Figure 5.8 / Latent hardening based non-convex shear stress vs. applied macroscopic shear for a plane strain pure shear problem of a fully periodic RVE.

p12

12
0.022

0.021

0.021

0.02

0.02

0.019

0.019

0.018

0.018

0.017

0.017

0.016

0.016
0.015
0.015
0.014
0.014
0.013
0.013
0.012

= 60

= 120
0.01

0.01

0.015
0.02

0.02

0.025
0.03

0.03

0.035
0.04

0.04

0.045
0.05

0.05

0.055

Figure 5.9 / Shear strain (first figure), plastic shear strain (second figure), and plastic
slip (last two figures) distributions at = 0.034 for a plane strain pure shear problem
of a fully periodic RVE.

examples. In Fig. 5.10 the field quantities are plotted for two additional cases with
different fluctuations applied in Youngs modulus. First, the fluctuation is given only

91

5.4 Numerical analysis

p12

12

0.02
0.02
0.019
0.019
0.018
0.018
0.017
0.017
0.016
0.016
0.015
0.015
0.014
0.014
0.013

= 60

= 120
0.015

0.015

0.02

0.02

0.025

0.025

0.03

0.03

0.035

0.035

0.04

0.04

0.045

0.045

0.05

0.05

p1 2

12

0.02
0.02
0.019
0.019
0.018
0.018
0.017
0.017
0.016
0.016
0.015
0.015
0.014
0.014
0.013

= 60

= 120
0.015

0.015

0.02

0.02

0.025

0.025

0.03

0.03

0.035

0.035

0.04

0.04

0.045

0.045

0.05

0.05

Figure 5.10 / Shear strain, plastic shear strain, and plastic slip distributions at =
0.034 for a plane strain pure shear problem of a fully periodic RVE with a fluctuation
applied to the central element (first four figures) and to the middle element on the
right edge (last four figures).

92

5 Non-convex rate dependent strain gradient crystal plasticity and


deformation patterning

to an element in the center of the RVE and in the second case the fluctuation is given
to an element in the middle of the right edge. The applied fluctuation has certainly
an effect on the distribution of the deformation patterns, however, the spacing of
high and low strain areas and the amplitude do not depend on the applied fluctuation. The macroscopic stress versus strain diagram is not plotted in these two cases
because it is identical to the previous case plotted in Fig. 5.8.
It is remarked that the obtained deformation patterns depend considerably on the
applied rate of deformation. Increasing the strain rate results in a stiffer stress versus
strain response while the amplitude of the obtained patterns decreases. Another
important parameter affecting the patterning of deformation fields is the internal
length scale R which should be small enough to obtain a stable numerical solution
and pronounced patterns. These two factors and other material parameters play
an important role in the convergence of the numerical solution as well. Due to the
non-convexity of the latent hardening potential at each increment of the deformation
some combinations of the parameters might give convergence problems and it is not
always possible to reach the same state of deformation with different combinations
of the rate of deformation and material parameters.
As stated before the non-convex potential has been used to recover a cellular kind of
dislocation microstructures by Ortiz and Repettto (1999) and Ortiz et al. (2000) via
an external lamination procedure, while we try to obtain deformation patterns via
a non-convex evolution problem. In the current examples we observe deformation
patterning at low loading rates and a small internal length scale parameter due to
the latent hardening based non-convexity. This shows agreement with the study in
section 5.3 illustrating the capability of the latent hardening function for deformation
patterning.

5.5 Summary and conclusion


A plastic slip based rate dependent non-convex strain gradient crystal plasticity
model is proposed and embedded in a FEM solution framework using displacements
and plastic slips as degree of freedoms. A physically based latent hardening nonconvex plastic potential (Ortiz and Repettto (1999)) is incorporated into the thermodynamically consistent viscous strain gradient crystal plasticity model based on the
1D formulation as presented in Yalcinkaya et al. (2011a) in order to obtain deformation and plastic slip patterns due to slip system interactions in physically deforming
crystals. The presented approach models the implicit evolution of deformation patterns through the intrinsic non-convexity of the free energy function. This kinetics
driven method offers an alternative to the explicit construction of the microstructure
evolution (e.g. in Ortiz and Repettto (1999)).

5.5 Summary and conclusion

93

Selected examples demonstrate the ability of the model to obtain a deformation


driven plastic slip microstructure evolution. While a convex theory explicitly illustrates the size dependent and rate dependent boundary layer development, the
non-convex formulation, originating from the slip interaction phenomena in crystals, allows for deformation and plastic slip patterning.
In the model the destabilizing non-convex term in the free energy is mathematically
stabilized through the gradient term in the free energy and the viscous nature of the
thermodynamically consistent slip law. The obtained microstructure evolution is rate
dependent, where the homogeneity or inhomogeneity of the deformation basically
depends on the applied rate. Due to the non-convexity at each increment of the
applied deformation, the convergence and the patterning of the field are sensitive to
many parameters, where we have shown only the effect of the applied fluctuation.
A more detailed investigation of the dependence of the results on the loading type,
loading rates, slip system orientations, material parameters, the mesh and boundary
conditions would help for a deeper understanding of the observed microstructure
evolution phenomena.

94

Chapter six

Discussion and conclusions

Abstract / The main conclusions drawn from the thesis are summarized in this concluding chapter and an outlook to further work related to microstructure evolution in crystal
plasticity models is presented.

The objective of this thesis was the investigation of the macroscopic transient hardening and softening effects due to the dislocation microstructure evolution in BCC metals. The project started with the development of a large strain BCC crystal plasticity
framework where the intrinsic anisotropy due to non-planar spreading of screw dislocation cores was addressed. Next, the evolution of an existing dislocation cell structure has been investigated in a composite cell model which was incorporated into
the BCC crystal plasticity model. This framework starts with an assumed dislocation
structure and studies the evolution of this microstructure via phenomenological dislocation evolution equations. However, the formation of dislocation microstructures
or deformation patterning cannot be simulated by these kind of models. Therefore,
in order to model the formation and evolution of inhomogeneous deformation patterns, a non-convex rate dependent strain gradient crystal plasticity model has been
developed in a small strain context in the last part of the thesis.
The BCC crystal plasticity framework, presented in Yalcinkaya et al. (2008) (chapter
2), reveals some unique characteristics of BCC crystals. A comprehensive summary
of intrinsic properties of these materials are presented, including recent insights in
the activation of different slip systems, violation of Schmids law, temperature and
orientation dependence of the flow stress and resulting stress-strain curves. With
respect to the BCC crystal plasticity model the following conclusions are drawn:
The applied slip evolution equation which is based on the thermally activated
dislocation kinetics plays an essential role in the model since all the intrinsic
characteristics result from the actual formulation of this equation.
95

96

6 Discussion and conclusions

Contrary to the common assumption of using the {110}, {112} and {123} type
of slip planes in all conditions, it is explained that the activation of slip systems
depends on the temperature. At moderate temperatures only {112} slip planes
are activated which affect considerably the mechanical behavior of BCC metals.
The non-Schmid behavior is introduced in the slip law by modifying the effective shear stress, where the non-Schmid contribution represents the dislocation
cores spreading in a non-planar manner.
Deviation from a proportional strain path is associated with hardening or softening
of the material due to the induced plastic anisotropy. At moderate strains the dominating effect is attributed to the evolving underlying dislocation microstructures.
Chapter 3 (Yalcinkaya et al. (2009)) deals with a combination of a composite dislocation cell model, which explicitly describes the dislocation structure evolution, with
the BCC crystal plasticity framework from chapter 2 to bridge the microstructure
evolution and its macroscopic anisotropic effects. The main conclusions from this
part are:
A phenomenological cell structure evolution model embedded into a crystal
plasticity framework is well able to reproduce all essential characteristics of
strain path changes reported, consistently with experimental observations at
two scales.
The model proposed allows to study the interaction between different sources
of anisotropy, where a clear example at the single crystal and polycrystal has
been given.
The level at which the enrichment of the crystal plasticity model was made,
enables its use in more complex microstructures as e.g. multi-phase steels.
In order to complete the missing link between the formation of the microstructure
and its evolution in crystal plasticity frameworks, the second part of the thesis
concentrated on the development of a non-convex rate dependent crystal plasticity model, which may eventually simulate rate dependent dislocation microstructure formation and evolution together with macroscopic hardening-softening stressstrain responses. In chapter 4 (Yalcinkaya et al. (2011a)), inspired by the efficiency
of phase field models for microstructure formation and evolution, a 1D non-convex
strain gradient plasticity model has been developed yielding combined hardeningsoftening. The resulting formulation takes a conceptually dual structure to the
Ginzburg-Landau type of phase field models where high and low slipped regions
represent different phases. The main conclusions related to this part are:

97

The thermodynamically consistent constitutive model is able to capture patterning of plastic slip, qualitatively similar to a phase decomposition mechanism.
The mathematical structure of the employed double-well plastic potential allows to control the amount and the timing of deformation patterning.
The framework can capture both homogeneous and inhomogeneous deformations depending on the rate of the applied deformation.
The model is conceptually capable of covering many aspects of microstructure
evolution processes, depending on the externally applied load and boundary
conditions. Non-equilibrium microstructures are thereby well at reach.
The model could be used for various materials. Especially the clear band formation observed in irradiated materials or Luders

band formation and motion


in low carbon steels exhibit one-to-one correspondence with the obtained results from the numerical examples.
Chapter 5 (Yalcinkaya et al. (2011b)) extends the 1D non-convex strain gradient crystal plasticity formulation of the previous chapter to 2D crystals with multi-slip systems. In this chapter a more physically based non-convex plastic potential is considered which originates from the slip interactions in crystals. The main conclusions
are:
Studying the effect of the latent hardening based non-convexity (Ortiz and
Repettto (1999)) on the deformation patterning revealed that the latent hardening potential satisfies the energetic conditions for patterning.
In the numerical examples, the convex theory explicitly illustrates the effect of
the internal length scale parameter and the deformation rate on the microstructure evolution due to the hard boundary conditions. The results show agreement with chapter 4 and the literature.
The non-convex formulation presents the possibility of deformation patterning depending on the loading rate. Strong deformation patterns are obtained
during monotonic pure shear deformation.
The non-convex strain gradient plasticity frameworks in the last two chapters of the
thesis present an energetic approach to deformation patterning and related softening of the materials. While in many crystal plasticity models the softening effect is

98

6 Discussion and conclusions

introduced externally, in the current approach it becomes a natural product. Moreover, compared to incremental variational principles and relaxation theories for microstructure evolution it is simpler to implement and computationally less expensive. Capturing both equilibrium and non-equilibrium states of the microstructures
makes it a unique framework in computational plasticity. The crucial aspect in the
present models is the plastic potential entering the formulation in a thermodynamically consistent manner.
In this thesis two different plastic potentials have been formulated. The double-well
based non-convexity in the fourth chapter presents a phenomenological description
of plastic slip patterning. The coefficients of the potential and the rate of deformation control the whole patterning mechanism. Even though the material parameters are not identified for real materials it offers a broad range of application. A
more practical emphasis applied to a particular material is obviously needed in future work, provided reliable experimental data to recover the non-convex term are
available. On the other hand, the latent-hardening based non-convex potential used
in the last chapter addresses a physical phenomenon, i.e. slip interactions in crystals.
It induces a clear implicit deformation patterning in the present non-convex strain
gradient crystal plasticity framework, compared to its previous usage for external
microstructure evolution via lamination procedure. A more detailed study on the
available functions and possible other descriptions of the non-convexity for specific
materials is obviously needed as a future work.

Bibliography

A CKERMANN , F., M UGHRABI , H., and S EEGER , A. (1983). Temperature and strainrate dependence of the flow stress of ultrapure niobium single crystals in cyclic
deformation. Acta Metall., 31, 13531366.
A DAMS , J. J., A GOSTA , D. S., L EISURE , R. G., and L EDBETTER , H. (2006). Elastic
constant of monocrystal iron from 3 to 500 K. J. Appl. Phys., 100, 113530.
A IFANTIS , E. C. (1987). The Physics of plastic deformation. Int. J. Plast., 3, 211247.
A RGON , A. S. (2008). Strengthening mechanisms in crystal plasticity. Oxford University Press.
A SARO , R. and N EEDLEMAN , A. (1985). Texture Development and Strain Hardening
in Rate Dependent Polycrystals. Acta Metall., 33, 923953.
A SARO , R. J. and R ICE , J. R. (1977). Strain localization in ductile single crystals. J.
Mech. Phys. Solids, 25, 309338.
B ACROIX , B. and H U , Z. (1995). Texture evolution induced by strain path changes
in low carbon steel sheets . Metall. Mater. Trans. A, 26, 601613.
B ACROIX , B., G ENEVOIS , P., and C., T EODOSIU (1994). Plastic anisotropy in low
carbon steels subjected to simple shear with strain path changes. Eur. J. Mech.
A/Solids, 13, 661675.
B AILEY , J. E. and H IRSCH , P. B. (1960). The dislocation distribution, flow stress, and
stored energy in cold-worked polycrystalline silver. Philos. Mag., 5, 485.
B ARDELLA , L. (2006). A deformation theory of strain gradient crystal plasticity that
accounts for geometrically necessary dislocations. J. Mech. Phys. Solids, 54, 128160.
B ARDELLA , L. (2007). Some remarks on the strain gradient crystal plasticity modelling, with particular reference to the material length scales involved. Int. J. Plast.,
23, 296322.
B ARKER , I., H ANSEN , N., and R ALPH , B. (1989). The development of deformation
substructures in face-centered cubic metals. Mat. Sci. Eng. A, 113, 449454.

99

100

Bibliography

B ARLAT , F., F ERREIRA D UARTE , J. M., G RACIO , J. J., L OPES , A. B., and R AUCH ,
E. F. (2003). Plastic flow for non-monotonic loading conditions of an aluminum
alloy sheet sample. Int. J. Plast, 19, 12151244.
B ARRETT , C. S., A NSEL , G., and M EHL , R F (1937). Slip, twinning and cleavage in
iron and silicon ferrite. Trans. Am. Soc. Met., 25, 702733.
B ASSANI , J. L. (1994). Plastic flow of crystals. Adv. Appl. Mech., 30, 191257.
B ASSANI , J. L., I TO , K., and V ITEK , V. (2001). Complex macroscopic plastic flow
arising from non-planar dislocation core structures. Mat. Sci. Eng. A, 319-321, 97
101.
B AYLEY , C. J., B REKELMANS , W. A. M., and G EERS , M. G. D. (2006). A comparison
of dislocation induced back stress formulations in strain gradient crystal plasticity.
Int. J. Solids Struct., 43, 72687286.
B OLEF , D. I. and K LERK , J. D. (1962). Elastic Constants of Single-Crystal Mo and W
between 77 and 500 K. J. Appl. Phys., 33, 23112314.
B OUVIER , S., G ARDEY , B., H ADDADI , H., and C., T EODOSIU (2006). Characterization of the strain-induced plastic anisotropy of rolled sheets by using sequences of
simple shear and uniaxial tensile tests. J. Mater. Process. Tech., 174, 115126.
B RONKHORST , C. A., K ALIDINDI , S. R., and A NAND , L. (1992). Polycrystalline
plasticity and the evolution of crystallographic texture in FCC metals. Philos. T.
Roy. Soc. A, 341, 443477.
B ROWN , S. B., K IM , K. H., and A NAND , L. (1989). An internal variable constitutive
model for hot working of metals. Int. J. Plast., 5, 95130.
B RUNNER , D. and D IEHL , J. (1987). The use of stress-relaxation measurements for
investigations on the flow stress of alpha-iron. Phys. Stat. Sol. A, 104, 145155.
B RUNNER , D. and D IEHL , J. (1997). The effect of atomic lattice defects on the softening phenomena of high-purity alpha-iron. Phys. Stat. Sol. A, 160, 355372.
C AHN , J. W. (1961). On spinodal decomposition. Acta Metall., 9, 795801.
C ARROLL , K. J. (1965). Elastic Constants of Niobium from 4.2 to 300 K. J. Appl. Phys.,
36, 36893690.
C ERMELLI , P. and G URTIN , M. E. (2001). On the characterization of geometrically
necessary dislocations in finite plasticity. J. Mech. Phys. Solids, 49, 15391568.
C HEN , N. K. and M ADDIN , R. (1954). Slip planes and energy of dislocations in a
body centered cubic structure. Acta Metall., 2, 4951.
C HRISTIAN , A., K ANERT, O., and D E H OSSON , J. T H . M. (1990). Dislocation dynamics in Vanadium: A nuclear magnetic and transmission electron microscopic
study. Acta Metall. Mater., 38, 24792484.

Bibliography

101

C HRISTODOULOU , N., W OO , O. T., and M AC E WEN , S. R. (1986). Effect of stress


reversals on the work hardening behaviour of polycrystalline copper. Acta Metall.,
34, 15531562.
C ONTI , S. and O RTIZ , M. (2005). Dislocation microstructures and the effective behavior of single crystals. Arch. Ration. Mech. Anal., 176, 103147.
C ONTI , S., H AURET , P., and O RTIZ , M. (2007). Concurrent multiscale computing of
deformation microstructure by relaxation and local enrichment with application to
single-crystal plasticity. Multiscale Model. Simul., 6, 135157.
C UITINO , A. M. and O RTIZ , M. (1993). Constitutive modeling of L12 intermetallic
crystals. Mater. Sci. Eng. A, 170, 111123.
D AO , M. and A SARO , R, J. (1993). Non-Schmid effects and localized plastic flow in
intermetallic alloys. Mater. Sci. Eng., 170, 143160.
D AO , M. and A SARO , R. J. (1996). Localized deformation modes and non-Schmid
effects in crystalline solids, Part I and II. Adv. Appl. Mech., 23, 71132.
D EL P IERO , G. (2009). On the method of virtual power in continuum mechanics.
Journal of Mechanics of Materials and Structures, 4, 281292.
D UESBERY , M. S. and F OXALL , R. A. (1969). A detailed study of the deformation of
high purity niobium single crystals. Philos. Mag., 20, 719751.
D UESBERY , M. S. and V ITEK , V. (1998). Plastic anisotropy in BCC transition metals.
Acta Mater., 46, 14811492.
D UESBERY , M. S., V ITEK , V., and C SERTI , J. (2002). Understanding Materials, 165.
D UNSTAN , D. J. and B USHBY , A. J. (2004). Theory of deformation in small volumes
of material. Proc. R. Soc. Lond. A, 460, 27812796.
E VERS , L. P., B REKELMANS , W. A. M., and G EERS , M. G. D. (2004a). Non-local
crystal plasticity model with intrinsic SSD and GND effects. J. Mech. Phys. Solids,
52, 23792401.
E VERS , L. P., B REKELMANS , W. A. M., and G EERS , M. G. D. (2004b). Scale dependent crystal plasticity framework with dislocation density and grain boundary
effects. Int. J. Solids Struct., 41, 52095230.
F ERNANDES , J. V. and S CHMITT , J. H. (1983). Dislocation Microstructures in steel
during deep drawing. Philos. Mag. A, 48, 841870.
F RANCIOSI , P. and Z AOUI , A. (1982). Multislip in FCC crystals; a theoretical approach compared with experimental data. Acta Metall., 30, 16271637.
F ROST , H. J. and A SHBY , M. F. (1982). Deformation-Mechanism Maps: The plasticity
and creep of metals and ceramics. Pergamon Pr.
G ANAPATHYSUBRAMANIAN , S. and Z ABARAS , N. (2005).
Modeling the
thermoelastic-viscoplastic response of polycrystals using a continuum representation over the orientation space. Int. J. Plast., 21, 119144.

102

Bibliography

G ARDEY , B., B OUVIER , S., R ICHARD , V., and B ACROIX , B. (2005). Texture and
dislocation structures observation in a dual-phase steel under strain-path changes
at large deformations. Mater. Sci. Eng., A, 400-401, 136141.
G EERS , M. G. D., B REKELMANS , W. A. M., and B AYLEY , C. J. (2007). Second-order
crystal plasticity: internal stress effects and cyclic loading. Model. Simul. Mater. Sci.
Eng., 15, 133145.
G OUGH , H. J. (1928). The behavior of a single crystal alpha-iron subjected to alternating torsional stresses. P. Roy. Soc. Lond. A, 118, 498534.

G R OGER
, R. and V ITEK , V. (2005). Breakdown of the Schmid law in BCC molybdenum related to the effect of shear stress perpendicular to the slip direction. Mater.
Sci. Forum, 482, 123.
G ROMA , I. (1997). Link between the microscopic and mesoscopic length scale description of the collective behavior of dislocations. Phys. Rev. B, 56, 58075813.
G ROMA , I. and PAVLEY , G. S. (1993). Role of the secondary slip system in a computer
simulation model of the plastic behaviour of single crystals. Mater. Sci. Eng. A, 164,
306311.
G UIU , F. (1969). Slip asymmetry in molybdenum single crystals deformed in direct
shear. Scr. Metall., 3, 449454.
G URTIN , M. E. (2000). On the plasticity of single crystals: free energy, microforces,
plastic-strain gradients. J. Mech. Phys. Solids, 48, 9891036.
G URTIN , M. E. (2002). A gradient theory of single-crystal viscoplasticity that accounts for geometrically necessary dislocations. J. Mech. Phys. Solids, 50, 532.
G URTIN , M. E., A NAND , L., and L ELE , S. P. (2007). Gradient single-crystal plasticity
with free energy dependent on dislocation densities. J. Mech. Phys. Solids, 55, 1853
1878.
H ACKL , K. and K OCHMANN , D. M. (2008). Relaxed potentials and evolution equations for inelastic microstructures. IUTAM Symposium on Theoretical, Computational
and Modelling Aspects of Inelastic Media, 2741.

H AHNER
, P. (1994). Theory of solitary plastic waves. Part I: Luders

bands in polycrystals. Appl. Phys. A, 58, 4148.


H ALIM , H., W ILKINSON , D. S., and N IEWCZAS , M. (2007). The PortevinLe Chatelier (PLC) effect and shear band formation in an AA5754 alloy. Acta Metall., 55,
41514160.
H ANSEN , N. and H UANG , X. (1997). Dislocation Structures and Flow Stress. Mat.
Sci. Eng. A, 234, 602605.
H AVNER , K. S. (1992). Finite Plastic Deformation of Crystalline Solids. Cambridge
Monographs on Mechanics and Applied Mathematics.

Bibliography

103

H ILL , R. and R ICE , J. R. (1972). Constitutive analysis of elastic-plastic crystals at


arbitrary strains. J. Mech. Phys. Solids, 20, 401413.
H OC , T. and F OREST , S. (2001). Polycrystal modelling of IF-Ti steel under complex
loading path. Int. J. Plast, 17, 6585.
H OLLANG , L., H OMMEL , M., and S EEGER , A. (1997). The flow stress of ultra-highpurity molybdenum single crystals. Phys. Stat. Sol. A, 160, 329354.
H OLT , D. L. (1980). Dislocation cell formation in metals. J. Appl. Phys., 41, 31973201.
H UTCHINSON , J. W. (1976). Bounds and self-consistent estimates for creep of polycrystalline materials. Proc. R. Soc. London, A, 348, 101127.
I RWIN , G. J., G UIU , F., and P RATT , P. L. (1974). The influence of orientation on slip
and strain hardening of molybdenum single crystals. Phys. Stat. Sol. A, 22, 685698.
I TO , K. and V ITEK , V. (2001). Atomistic study of non-Schmid effects in the plastic
yielding of BCC metals. Philos. Mag., 81, 13871407.
K ALIDINDI , S. R., B RONKHORST , C. A., and A NAND , L. (1992). Crystallographic
texture evolution in bulk deformation processing of FCC metals. J. Mech. Phys.
Solids, 40, 537569.
K EH , A. S. (1964). Work hardening and deformation sub-structure in iron single
crystals deformed in tension at 298 K. Philos. Mag., 12, 930.
K EH , A. S., W EISSMAN , S., T HOMAS , G., and WASHBURN , J. (1963). Electron Microscopy and Strength of Crystals. Interscience.
K OCKS , U. F., A RGON , A. S., and A SHBY , M. F. (1975). Thermodynamics and
Kinetics of Slip. Prog. Mater Sci., 19, 1291.
K OSLOWSKI , M., C UITINO , A. M., and O RTIZ , M. (2002). A phase field theory of
dislocation dynamics, strain hardening and hysteresis in ductile single crystals. J.
Mech. Phys. Solids, 50, 25972635.
K OTHARI , M. and A NAND , L. (1998). Elasto-viscoplastic constitutive equations for
polycrystalline metals: Application to tantalum. J. Mech. Phys. Solids, 46, 5183.
K REJCI , J. and L UKAS , P. (1971). Dislocation substructure in fatigued -iron single
crystals. Phys. Status Solidi A, 5, 315325.
K UBIN , L. P. and C ANOVA , G. (1992). The modelling of dislocation patterns. Scr.
Metall. Mater., 27, 957962.
K UHL , E. and S CHMID , D. W. (2007). Computational modeling of mineral unmixing
and growth. Comput. Mech., 39, 439451.
K UHLMANN -W ILSDORF , D. (1962). A new theory of work hardening. Trans. Am.
Inst. Min. Metall. Eng., 224, 10471061.
K UHLMANN -W ILSDORF , D. (1989). Theory of Plastic Deformation: properties of low
energy dislocation structures. Mat. Sci. Eng. A, 113, 141.

104

Bibliography

K UHLMANN -W ILSDORF , D. (2001). Q: Dislocation structures - how far from equilibrium? A: Very close indeed. Mater. Sci. Eng. A, 315, 211216.
K UHLMANN -W ILSDORF , D. and VAN DER M ERWE , J. H. (1982). Theory of dislocation cell sizes in deformed metals. Mater. Sci. Eng., 55, 7983.
L AMBRECHT , M., M IEHE , C., and D ETTMAR , J. (2003). Energy relaxation of nonconvex incremental stress potentials in a strain-softening elastic-plastic bar. Int. J.
Solids Struct., 40, 13691391.
L ANGER , J. S. (1971). Theory of spinodal decomposition in alloys. Ann. Phys., 65,
5386.
L EE , E. H. (1969). Elastic-plastic deformation at finite strains. J. Appl. Mech., 36, 16.
L EE , Y. J., S UBHASH , G., and R AVICHANDRAN , G. (1999). Constitutive modeling of
textured body-centered-cubic (BCC) polycrystals. Int. J. Plast., 15, 625645.
L I , Z. J., W INTHER , G., and H ANSEN , N. (2006). Anisotropy in rolled metals induced by dislocation structure. Acta Mater., 54, 401410.
L IAO , K. C., F RIEDMAN , P. A., and PAN , J. (1998). Texture development and plastic
anisotropy of BCC strain hardening sheet metals. Int. J. Solids Struct., 35, 5205
5236.
L UBARDA , V. A., B LUME , J. A., and N EEDLEMAN , A. (1993). An Analysis of Equilibrium Dislocations Distributions. Acta Metall. Mater., 41, 625642.
M IEHE , C. (2002). Strain-driven homogenization of inelastic microstructures and
composites based on an incremental variational formulation. Int. J. Numer. Methods
Eng., 55, 12851322.
M IEHE , C., S CHOTTE , J., and L AMBRECHT, M. (2002). Homogenization of inelastic
solid materials at finite strains based on incremental minimization principles. Application to the texture analysis of polycrystals. J. Mech. Phys. Solids, 50, 21232167.

M IEHE , C., L AMBRECHT, M., and G URSES


, E. (2004). Analysis of material instabilities in inelastic solids by incremental energy minimization and relaxation methods:
evolving deformation microstructures in finite plasticity. J. Mech. Phys. Solids, 52,
27252769.
M OLLICA , F., R AJAGOPAL , K. R., and S RINIVASA , A. R. (2001). The inelastic behaviors of metals subject to loading reversal. Int. J. Plast, 17, 11191146.
M UGHRABI , H. (1987). A Two-parameter Description of heterogeneous dislocation
Distributions in Deformed Metal Crystals. Mater. Sci. Eng., 85, 1531.
M UGHRABI , H., U NGAR , T., and W ILKENS , W. (1986). Long range internal stresses
and asymmetric X-ray line-broadening in tensile-deformed [001]-oriented copper
single crystals. Philos. Mag. A, 53, 793813.
N AUMAN , E. B. and H E , D. Q. (2001). Nonlinear diffusion and phase separation.
Chem. Eng. Sci., 56, 19992018.

Bibliography

105

N ESTEROVA , E. V., B ACROIX , B., and T EODOSIU , C. (2001). Microstructure and


texture evolution under strain-path changes in low-carbon interstitial-free steel.
Metall. Mater. Trans. A, 32, 25272538.
O RTIZ , M. and R EPETTTO , E. A. (1999). Nonconvex energy minimization and dislocation structures in ductile single crystals. J. Mech. Phys. Solids, 47, 397462.
O RTIZ , M., R EPETTO , E. A., and S TAINER , L. (2000). A theory of subgrain dislocation structures. J. Mech. Phys. Solids, 48, 20772114.
P EETERS , B. (2002). Multiscale modelling of the induced plastic anisotropy in IF steel
during sheet forming. Ph.D. thesis, Katholieke Universiteit Leuven.
P EETERS , B., K ALIDINDI , S. R., VAN H OUTTE , P., and A ERNOUDT , E. (2000). A
Crystal Plasticity Based Work-Hardening/Softening Model for B.C.C. Metals Under Changing Strain Paths. Acta Mater., 48, 21232133.
P EIRCE , D., A SARO , R. J., and N EEDLEMAN , A. (1982). An analysis of non-uniform
and localized deformation in ductile single crystals. Acta Metall., 30, 10871119.
P ICHL , W. (2002). Slip geometry and plastic anisotropy of body-centered cubic metals. Phys. Stat. Sol. A, 189, 525.
Q IN , Q. and B ASSANI , J. L. (1992). Non-Schmid yield behavior in single crystals. J.
Mech. Phys. Solids, 40, 813833.
Q UEYREAU , S., M ONNET , G., and D EVINCRE , B. (2008). Slip system interactions
in alpha-iron determined by dislocation dynamics simulations. Int. J. Plast, 25,
361377.
R ACHERLA , V. and B ASSANI , J. L. (2007). Strain burst phenomena in the necking
of a sheet that deforms by non-associated plastic flow. Modelling Simul. Mater. Sci.
Eng., 15, 297311.
R AO , B. V. N. and L AUKONIS , J. V. (1983). Microstructural mechanism for the
anomalous tensile behavior of aluminum-killed steel prestrained in plane tension.
Mater. Sci. Eng., 60, 125135.
R AUCH , E. F. (1991). Stress reversal tests imposed by shear on mild steel. n: Strength
of Metals and Alloys, (D.G. Brandon, R. Shaim, A. Rosen, Eds.), Haifa, Israel, July 14-19,
1991, (Proc. of ICSMA 9, Freund Publ. House, Ltd., London), 1, 187194.
R AUCH , E. F. (1992). The flow law of mild steel under monotonic or complex strain
path. Solid State Phenom., 23-24, 317333.
R AUCH , E. F. and S CHMITT , J. H. (1989). Dislocation Substructures in Mild Steel
Deformed in Simple Shear. Mat. Sci. Eng. A, 113, 441448.
R ICE , J. R. (1971). Inelastic constitutive relations for solids: an internal variable
theory and its application to metal plasticity. J. Mech. Phys. Solids, 19, 433455.
S ANG , H. and L LOYD , D. J. (1979). The influence of biaxial prestrain on the tensile
properties of three aluminum alloys. Metal. Trans. A, 10, 17731776.

106

Bibliography

S AUZAY , M., B AVARD , K., and K ARLSEN , W. (2010). TEM observations and finite
element modelling of channel deformation in pre-irradiated austenitic stainless
steels Interactions with free surfaces and grain boundaries. J. Nucl. Mater., 406,
152165.
S CHMID , E. and B OAS , W. (1935). Kristallplastizitat. Springer.
S CHMITT , J. H., F ERNANDES , J. V., G RACIO , J. J., and F., V IEIRA M. (1991). Plastic
behavior of copper sheets during sequential tension tests. Mat. Sci. Eng. A, 147,
143154.
S EEGER , A. (1981). Temperature and strain-rate dependence of the flow stress of
body-centered cubic metals. A theory based on kink-kink interactions. Z. Metallkd.,
72, 369380.
S EEGER , A. (2001). Why anomalous slip in body-centered cubic metals? Mat. Sci.
Eng. A, 319-321, 254260.
S EEGER , A. and WASSERBACH , W. (2002). Anomalous slip - A feature of high-purity
body-centered cubic metals. Phys. Stat. Sol. A, 189, 2750.
S EVILLANO , J. G., VAN H OUTTE , P., and A ERNOUDT , E. (1981). Large strain work
hardening and textures. Prog. Mater Sci., 25.
S HAW , J. A. and K YRIAKIDES , S. (1998). Initiation and propagation of localized
deformation in elasto-plastic strips under uniaxial tension. Int. J. Plast., 13, 837
871.
S HU , J. Y., F LECK , N. A., VAN DER G IESSEN , E., and N EEDLEMAN , A. (2001).
Boundary layers in constrained plastic flow: comparison of nonlocal and discrete
dislocation plasticity. J. Mech. Phys. Solids, 49, 13611395.
S PITZIG , W. A. (1979). Effect of hydrostatic pressure on plastic properties - flow
properties of iron single crystals. Acta Metall., 27, 523534.
S UN , H. B., Y OSHIDA , F., M A , X., K AMEI , T., and O HMORI , M. (2003). Finite element simulation on the propagation of Luders

band and effect of stress concentration. Matter. Lett., 57, 32063210.


S VENDSEN , B. (2002). Continuum thermodynamic models for crystal plasticity including the effects of geometrically-necessary dislocations. J. Mech. Phys. Solids, 50,
12971329.
S VENDSEN , B. (2004). On thermodynamic- and variational-based formulations of
models for inelastic continua with internal length scales. Comput. Meth. Appl. Mech.
Eng., 193, 54295452.
S WANN , P. R. (1963). Electron Microscopy and Strength of Crystals. G. Thomas and J.
Washburn, eds. Interscience, New York, 131.
TARIGOPULA , V., H OPPERSTAD , O. S., L ANGSETH , M., and H., C LAUSEN A. (2008).
Elastic-plastic behavior of dual-phase, high-strength steel under strain path
changes. Eur. J. Mech. A/Solids, 27, 764782.

Bibliography

107

TAYLOR , G. I. and E LAM , C. F. (1926). The distortion of iron crystals. P. Roy. Soc.
Lond. A, 112, 337361.
T EODOSIU , C. (1992). Materials Science Input to Engineering Models. Modeling of
Plastic Deformation and its Engineering Applications (Proc. 13th Riso), 125146.
T EODOSIU , C. and H U , Z. (1995). Simulation of Materials Processing: Theory and Applications (Proc. of NUMIFORM 95), 173182.
U BACHS , R. L. J. M., S CHREURS , P. J. G., and G EERS , M. G. D. (2004). A nonlocal phase field model for microstructure evolution of Sn-Pb solder. J. Mech. Phys.
Solids, 52, 17631792.
V IATKINA , E. (2005). Micromechanical modelling of strain path dependency in FCC
metals. Ph.D. thesis, Eindhoven University of Technology.
V IATKINA , E. M., B REKELMANS , W. A. M., and G EERS , M. G. D. (2003). Strain path
dependency in metal plasticity. J. Phys. IV, 105, 355362.
V ITEK , V. (2004). Core structure of screw dislocations in body-centred cubic metals:
Relation to symmetry and interatomic bonding. Philos. Mag., 84, 415428.

, R., B ASSANI , J. L, R ACHERLA , V., and Y IN ,


V ITEK , V., M ROVEC , M., G R OGER
L. (2004a). Effects of non-glide stresses on plastic flow of single and polycrystals
of molybdenum. Mater. Sci. Eng. A, 387, 138142.
V ITEK , V., M ROVEC , M., and B ASSANI , J. L. (2004b). Influence of non-glide stresses
on plastic flow: from atomistic to continuum modeling. Mater. Sci. Eng. A, 365,
3137.
WAGONER , R. H. and L AUKONIS , J. V. (1983). Plastic behavior of aluminum-killed
steel following plane-strain deformation. Metal. Trans. A, 14, 14871495.
WALGRAEF , D. and A IFANTIS , E. C. (1985). Dislocation patterning in fatigued metals
as a result of dynamical instabilities. J. Appl. Phys., 58, 688691.
WANG , J., L EVKOVITCH , V., R EUSCH , F., S VENDSEN , B., H UETINK , J., and
VAN R IEL , M. (2008). On the Modeling of Hardening in Metals During NonProportional Loading. Int. J. Plast., 24, 10391070.
WANG , Y. U., J IN , Y. M., C UITINO , A. M., and K HACHATURYAN , A. G. (2001a).
Application of phase field microelasticity theory of phase transformations to dislocation dynamics: Model and three dimensional simulations in a single crystal.
Philos. Mag. Lett., 81, 385393.
WANG , Y. U., J IN , Y. M., C UITINO , A. M., and K HACHATURYAN , A. G. (2001b).
Nano scale phase field microelasticity theory of dislocations: Model and 3D simulations. Acta Mater., 49, 18471857.
WANG , Y. U., J IN , Y. M., C UITINO , A. M., and K HACHATURYAN , A. G. (2001c).
Phase field microelasticity theory and modeling of multiple dislocation dynamics.
Appl. Phys. Lett., 78, 23242326.

108

Bibliography

W ERNER , M. (1987). Temperature and strain-rate dependence of the flow stress of


ultrapure tantalum single crystals. Phys. Stat. Sol. A, 104, 6378.
W ILSON , D. V. and B ATE , P. S. (1994). Influence of cell walls and grain boundaries
on transient response of an IF steel to changes in strain path. Acta Metall. Mater.,
42, 10991111.
X IE , C. L., G HOSH , S., and G ROEBER , M. (2004). Modeling Cyclic Deformation of
HSLA Steels Using Crystal Plasticity. J. Eng. Mater. Technol., 126, 339352.
YALCINKAYA , T., B REKELMANS , W. A. M., and G EERS , M. G. D. (2008). BCC single
crystal plasticity modeling and its experimental identification. Model. Simul. Mater.
Sci. Eng., 16, 085007.
YALCINKAYA , T., B REKELMANS , W. A. M., and G EERS , M. G. D. (2009). A composite dislocation cell model to describe strain path change effects in BCC metals.
Model. Simul. Mater. Sci. Eng., 17, 064008.
YALCINKAYA , T., B REKELMANS , W. A. M., and G EERS , M. G. D. (2011a). Deformation patterning driven by rate dependent nonconvex strain gradient plasticity. J.
Mech. Phys. Solids., 59, 117.
YALCINKAYA , T., B REKELMANS , W. A. M., and G EERS , M. G. D. (2011b). Nonconvex rate dependent strain gradient crystal plasticity and deformation patterning. In Prep.
Y EFIMOV , S., G ROMA , I., and VAN DER G IESSEN , E. (2004). A comparison of a
statistical-mechanics based plasticity model with discrete dislocation plasticity calculations. J. Mech. Phys. Solids, 52, 279300.
Y OSHIDA , F., K ANEDA , Y., and YAMAMOTO , S. (2008). A plasticity model describing
yield-point phenomena of steels and its application to FE simulation of temper
rolling. Int. J. Plast., 24, 17921818.
Y OUNG , C. T., H EADLEY , T. J., and LYTTON , J. L. (1986). Dislocation substructures
formed during the flow stress recovery of high purity aluminum. Mater. Sci. Eng.,
81, 391407.
B ORST , R. (1987). Computation of post-bifurcation and post-failure behaviour of
strain-softening solids. Comput. Struct., 25, 211224.

DE

Acknowledgements

Firstly, I would like to express my deep gratitude to my promoter Marc Geers, who
gave me the opportunity to start my PhD research in his group and who has been
continuously supporting me scientifically and socially throughout the whole PhD
process. He was not only the person to ask questions about the convergence problems in the models or the thermodynamical relations but also he has been the example researcher widening my horizon considerably. His positive approach during the
stressful times has always ended up with high motivation. Marc, thank you.
I would like to thank my co-promoter Marcel Brekelmans for his valuable input in the
process of developing each chapter and his fruitful discussions and criticism during
the development phase of the models and also on writing articles. Even though he
has been partially stopped with the university he never gave up supporting me.
I sincerely thank to the committee members; Prof.Dr. Klaus Hackl, Prof.Dr. Fionn
Dunne, Prof.Dr. Bob Svendsen, Prof.Dr. Marc Peletier and Dr. Henk Vegter for their
interest and constructive comments on the thesis.
A special thanks goes to Dr. Izzet Ozdemir who has always been there for discussions
and input. He has taught me how to be stubborn on debugging process and how to
overcome numerical problems during the implementation of some of the models.
I am grateful to Prof.Dr. Jeff de Hosson whom I visited in Groningen to discuss on
BCC crystals and who supported me via emails when I needed physical insight for
my computational models. Even though I have never met him, Prof.Dr. Ali Argon
from MIT has supported me remotely by answering my emails on the difficulties I
have faced in crystal plasticity modeling. I would like to thank him for his valuable
comments.
My sincere thanks goes to the IT administrators Patrick and Leo for their neverending support related to computer issues. And I would like to thank Alice van
Litsenburg in Mate and Monika Hoekstra from M2i for the support on solving practical matters faced throughout the PhD years in the Netherlands.

109

110

Acknowledgements

I would like to thank Karl Fredrik Nilsson, Peter Haehner and Vesselina Rugnelova
at the Institute of Energy in Joint Research Centre of European Commission, where I
have already started my post-doctoral research, for their understanding and support
on wrapping up my thesis. Considering the fact that chapter 5 of the thesis has
been completed at the Institute of Energy, the support of the institute is gratefully
acknowledged.
Finally, I am extremely grateful to my parents Duran and Tazegul
and my sister
Derya. I am such a lucky person to have such lovely and open-minded family. They
have always supported me for all the decisions I have taken.

Tuncay Yalcnkaya,
Petten, August 2011.

Curriculum Vitae

Tuncay Yalcnkaya was born on January 31, 1980 in Ankara, Turkey. After completing
his high school education in Batkent High School in 1998, he studied Aerospace
Engineering at Middle East Technical University, Ankara and got his B.Sc. degree in
2003. Directly thereafter, he continued his studies with a full DAAD scholarship at
the University of Stuttgart and got his M.Sc. degree on computational mechanics of
materials and structures (COMMAS) in 2005. After his graduation he was employed
by the Materials Innovation Institute (M2i) in Delft and has been working on his PhD
thesis on strain path change effects and crystal plasticity modeling of BCC metals at
the Mechanical Engineering Department of the Eindhoven University of Technology
under the supervision of Prof.Dr.Ir. M.G.D. Geers and Dr.Ir. W.A.M. Brekelmans.
He is currently a post-doctoral researcher at the Institute for Energy - Joint Research
Centre of European Commission.

111

Potrebbero piacerti anche