multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal...

156
Multiscale modelling of grain boundary plasticity Citation for published version (APA): Beers, van, P. R. M. (2015). Multiscale modelling of grain boundary plasticity. Technische Universiteit Eindhoven. Document status and date: Published: 01/01/2015 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 20. Jan. 2021

Upload: others

Post on 24-Sep-2020

4 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Multiscale modelling of grain boundary plasticity

Citation for published version (APA):Beers, van, P. R. M. (2015). Multiscale modelling of grain boundary plasticity. Technische UniversiteitEindhoven.

Document status and date:Published: 01/01/2015

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 20. Jan. 2021

Page 2: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Multiscale modelling of

grain boundary plasticity

Paul van Beers

Page 3: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

This research was supported by the Dutch Technology Foundation STW, which is part ofthe Netherlands Organisation for Scientific Research (NWO), and which is partly fundedby the Ministry of Economic Affairs under Project Nr. 10109.

Multiscale modelling of grain boundary plasticity by Paul van Beers.Eindhoven : Technische Universiteit Eindhoven, 2015 – Proefschrift.

A catalogue record is available from the Eindhoven University of Technology Library.ISBN: 978-90-386-3773-0

Copyright © 2015 by P.R.M. van Beers. All rights reserved.

This thesis was prepared with the LATEX 2ε documentation system.Cover design: Paul van Beers and Stefanie van den HerikPrinted by: Proefschriftmaken.nl || Uitgeverij BOXPress

Page 4: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Multiscale modelling of

grain boundary plasticity

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan deTechnische Universiteit Eindhoven, op gezag van derector magnificus prof.dr.ir. C.J. van Duijn, voor een

commissie aangewezen door het College voorPromoties, in het openbaar te verdedigen

op donderdag 5 februari 2015 om 14:00 uur

door

Paulus Rene Maria van Beers

geboren te Heerlen

Page 5: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Dit proefschrift is goedgekeurd door de promotoren en de samenstelling van de pro-motiecommissie is als volgt:

voorzitter: prof.dr. L.P.H. de Goeypromotor: prof.dr.ir. M.G.D. Geerscopromotor: dr.ir. V.G. Kouznetsovaleden: prof.dr.ir. E. van der Giessen (Rijksuniversiteit Groningen)

prof.dr. B.J. Thijsse (Technische Universiteit Delft)Prof.Dr.-Ing.habil. T. Bohlke (Karlsruher Institut fur Technologie)prof.dr. P.M. Koenraad

adviseur: dr.ir. H. Vegter (Tata Steel R&D)

Page 6: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Contents

Summary v

Notation vii

1 Introduction 1

1.1 Background and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Objective and outline of the thesis . . . . . . . . . . . . . . . . . . . . . . 4

2 Grain boundary interface mechanics in strain gradient crystal plasticity 7

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1.1 Interactions between dislocations and grain boundaries . . . . . . . 8

2.1.2 Modelling grain boundary behaviour at the continuum scale . . . . 10

2.1.3 Higher-order interface BCs: review of existing approaches . . . . . 11

2.1.4 Outline and scope of the chapter . . . . . . . . . . . . . . . . . . . 13

2.2 Bulk crystal formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2.1 Crystal geometry and dislocation sign convention . . . . . . . . . . 14

2.2.2 Bulk crystal balance equations . . . . . . . . . . . . . . . . . . . . 15

2.2.3 Bulk constitutive equations . . . . . . . . . . . . . . . . . . . . . . 16

2.3 Interface formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3.1 Interface balance equations . . . . . . . . . . . . . . . . . . . . . . 18

2.3.2 Grain boundary constitutive model . . . . . . . . . . . . . . . . . . 18

2.3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.4 Numerical investigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.4.1 Model definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.4.2 Finite element implementation . . . . . . . . . . . . . . . . . . . . 24

2.4.3 Influence of interface properties . . . . . . . . . . . . . . . . . . . . 25

i

Page 7: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

ii Contents

2.4.4 Influence of crystal geometry . . . . . . . . . . . . . . . . . . . . . 29

2.4.5 Influence of multiple slip systems . . . . . . . . . . . . . . . . . . . 31

2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

3 A multiscale model of grain boundary structure and energy 37

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3.2 Fundamental GB structure concepts . . . . . . . . . . . . . . . . . . . . . 40

3.2.1 Grain boundary degrees of freedom . . . . . . . . . . . . . . . . . . 40

3.2.2 Lattice descriptions and grain boundary dislocations . . . . . . . . 40

3.2.3 Structural units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.2.4 Frank-Bilby equation . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.3 Atomistic simulations of GB structure and energy . . . . . . . . . . . . . 43

3.3.1 Simulation methodology . . . . . . . . . . . . . . . . . . . . . . . . 43

3.3.2 Simulation results . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.4 Multiscale approach to GB structure and energy . . . . . . . . . . . . . . 45

3.4.1 Details of the approach . . . . . . . . . . . . . . . . . . . . . . . . 45

3.4.2 Summary of the procedure and illustrative example . . . . . . . . 47

3.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.6.1 Frank-Bilby equation . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.6.2 Grain boundary structure . . . . . . . . . . . . . . . . . . . . . . . 57

3.6.3 Grain boundary energy . . . . . . . . . . . . . . . . . . . . . . . . 58

3.6.4 Model extension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

4 Grain boundary plasticity incorporating internal structure and energy 61

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.2 Bulk and interface frameworks . . . . . . . . . . . . . . . . . . . . . . . . 63

4.2.1 Bulk crystal formulation . . . . . . . . . . . . . . . . . . . . . . . . 63

4.2.2 Grain boundary interface formulation . . . . . . . . . . . . . . . . 65

4.3 Incorporation of GB structure and energy . . . . . . . . . . . . . . . . . . 66

4.3.1 Net defect description . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.3.2 Initial GB structure and energy . . . . . . . . . . . . . . . . . . . . 67

4.3.3 Evolution of GB structure and energy . . . . . . . . . . . . . . . . 67

4.4 Analysis of initial GB energetics . . . . . . . . . . . . . . . . . . . . . . . 69

4.5 Numerical analysis of GB plasticity . . . . . . . . . . . . . . . . . . . . . . 72

4.5.1 Model definition and finite element implementation . . . . . . . . . 72

4.5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

Page 8: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

iii

4.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

5 Defect redistribution in a continuum grain boundary plasticity model 83

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

5.2 Bulk and interface frameworks . . . . . . . . . . . . . . . . . . . . . . . . 86

5.2.1 Bulk crystal formulation . . . . . . . . . . . . . . . . . . . . . . . . 86

5.2.2 Grain boundary interface formulation . . . . . . . . . . . . . . . . 88

5.2.3 Grain boundary structure and energy . . . . . . . . . . . . . . . . 89

5.3 Defect redistribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

5.3.1 Derivation of the model . . . . . . . . . . . . . . . . . . . . . . . . 91

5.3.2 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

5.4 Solution procedure and finite element implementation . . . . . . . . . . . 97

5.5 Numerical study of defect redistribution . . . . . . . . . . . . . . . . . . . 99

5.5.1 Model problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

5.5.2 Material parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.5.3 Discretisation and boundary conditions . . . . . . . . . . . . . . . 102

5.5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

6 Conclusions and outlook 111

6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

6.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

A Finite element implementation 117

B Data per rotation axis 119

Bibliography 123

Samenvatting 139

Acknowledgements 141

Curriculum Vitae 143

Page 9: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary
Page 10: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Summary

Multiscale modelling of grain boundary plasticity

In polycrystalline materials, such as metals, different types of material defects occur,two of which are dislocations and grain boundaries. Dislocations are line defects thatperturb the regularity of the crystal lattice. The collective motion of dislocations inducesthe irreversible plastic deformation of the material. Grain boundaries are planar defects,constituting the interfaces between two adjacent crystals with a different orientationof the atomic lattice. Both of these microstructural defects have a major influence onthe resulting material properties and macroscopic behaviour, in particular when grainboundaries constitute a relatively large fraction of the material volume, e.g. in fine-grained metals, thin films and micro-devices. In these cases, the role of grain boundariesand their interaction with dislocations become of key importance. Various types ofinteraction mechanisms between dislocations and grain boundaries have been identifiedin the literature using small-scale experiments and computer simulations: dislocationscan be accumulated, transmitted, absorbed or nucleated at the interfaces. The accurateaccount of these grain boundary mechanisms is one of the current challenges in themodelling of the plasticity of polycrystalline metals, which is the subject of this PhDthesis.

The objective in this thesis is to bridge the scale between the fine-scale interaction mecha-nisms associated with grain boundary plasticity and the polycrystalline continuum level,where most engineering design methods rely on. This is established by the developmentof a multiscale grain boundary interface plasticity framework, to be used in the contextof dislocation based gradient enhanced crystal plasticity.

First, the foundations of the grain boundary plasticity model are developed through theformulation of interface balance equations and thermodynamically consistent constitutiveequations, which enter a microstructurally motivated strain gradient crystal plasticityframework through the higher-order boundary conditions. A key ingredient is the grainboundary flow rule, which is an evolution equation of a scalar continuum scale quantityon each slip system at the interface, representing the net effect of the different fine-

v

Page 11: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

vi Summary

scale phenomena taking place at the grain boundary. Only atomistic simulations canprovide access to the details of the grain boundary structure and energy. Hence, drivenby the need for a physically based continuum scale description of the grain boundaryenergy required in continuum modelling frameworks, a multiscale approach is proposedthat leads to a continuum representation of the initial grain boundary structure defectcontent and energy, based on the output of atomistic simulations. Next, this multiscaleatomistic-to-continuum approach is exploited in the grain boundary extended crystalplasticity model, by means of a generalisation of the continuum expression for the initialgrain boundary energy. Finally, a continuum model is developed that incorporates thedescription of the defect redistribution along the grain boundary into the grain boundaryplasticity model. A diffusion type evolution law is proposed for the redistribution of thenet defect within the grain boundary, which allows for the nonlocal relaxation of thelocally accumulated grain boundary net defect. This continuum model consists of aset of equations, including a balance and constitutive equations, for a vectorial grainboundary net defect field along the grain boundary surface.

In order to investigate the capabilities of the developed grain boundary plasticity frame-work at different stages of the model development, it is applied to study the microscopicand macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the influence of the grain boundary constitutive parameters and the crystal geome-try have revealed that the macroscopic behaviour is altered considerably with respect tothe conventional grain boundary modelling approaches that lack the critical interactionsbetween dislocations and grain boundaries.

The multiscale model of grain boundary plasticity presented in the thesis provides afoundation to a more thorough analysis of polycrystalline metals at the continuum level.Especially when large fractions of grain boundaries are present, the accurate descriptionof the role of these boundaries becomes increasingly important.

Page 12: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Notation

In the following definitions, a Cartesian coordinate system with unit vector base {e1, e2, e3}applies and following the Einstein summation convention, repeated indices are summedfrom 1 to 3. The definitions are used throughout the thesis, unless indicated otherwise.

Quantities

scalar α, a, A

vector a = aiei

second order tensor A = Aijeiej

higher (nth) order tensor nA = Aij . . . neiej . . . en

column˜a

matrix A

Operations

multiplication c = ab, c = ab, C = aB

dyadic product C = ab = aibjeiej

cross product c = a× b = εijkajbkei (Levi-Civita symbol εijk)

dot product c = a · b = aibi, C = A ·B = AijBjkeiek

double dot product C = 4A : B = AijklBlkeiej, c = A : B = AijBji

triple dot product c = 3A ∴ 3B = AijkBkji

conjugate/transpose AT = Ajieiej

gradient operator ∇ = ei∂∂xi

derivative of vector w.r.t. vector C = ∂a∂b = ∂ai

∂bjeiej

time derivative a, a, A

vii

Page 13: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

viii Notation

Crystallographic notation

crystallographic direction, family [uvw], 〈uvw〉crystallographic plane, family (hkl), {hkl}slip system, family (hkl)[uvw], {hkl}〈uvw〉

Page 14: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Chapter 1

Introduction

1.1 Background and motivation

Whatever application one considers, the engineering design thereof always involves dif-ferent types of materials, such as metals, polymers, ceramics, organic materials and com-posites. All these materials behave differently with respect to their material properties,e.g. their mechanical, thermal, magnetic, electrical and optical properties. Although, ap-parently invisible at an application level, the material properties are determined by theunderlying microstructure, which is strongly influenced by the chemistry and processing.By exploiting in-depth knowledge of the intrinsic relation between structure, proper-ties, processing and chemistry, materials and products can be engineered, endowed withspecific functionalities of interest.

Physical mechanisms that govern the material behaviour occur at different characteris-tic time and length scales [140], as schematically illustrated in Fig. 1.1 in the case ofplasticity of metallic materials, which have a crystalline structure. The study of plas-ticity, entailing permanent deformations, is of key importance, since a wide range oftechnologies depend upon the ability to deform metals into particular desirable shapes.Moreover, permanent and time-dependent deformations are often undesired for otherapplications, requiring precise control over the underlying plasticity processes. The plas-tic deformation of metals is accommodated by a collective motion of material defects,among which dislocations, which are line defects disturbing the regularity of the crystallattice. In the last few decades, different modelling approaches have been developed thatcan describe the plasticity of metals, each model limited to a particular range of timeand length scales. Typically, with decreasing length scale, the amount of underlying mi-crostructural features and phenomena to be resolved increases. For example, on the onehand, atomistic models can resolve the atomic structure of dislocations in the crystalline

1

Page 15: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2 1 Introduction

atomistics

discrete dislocationplasticity

crystalplasticity

continuumplasticity

Time (s)

Length (m)

10-12

10-9

10-6

10-3

100

103

10-9

10-6

10-3

100

bottom

-up

top-d

own

Figure 1.1: Schematic illustration of the different characteristic time and length scales in modelling theplasticity of metallic materials.

lattice, but the material volume under consideration is restricted to a few hundreds ofnanometers within a time span of nanoseconds. On the other hand, continuum plasticitymodels consider length and time scales up to the order of meters and hours, respectively.However, these models cannot resolve the microstructural phenomena that dictate thismacroscopic behaviour.

In order to accurately describe the overall response of a material subjected to a certainapplied load, the details of each of the processes at the various time and length scales needto be incorporated. The paradigm of multiscale modelling is an attempt to compromisein the trade-off between the resolution of the fine-scale discrete models and the efficiencyof the macroscopic continuum models. Different multiscale modelling techniques exist,which can be classified into three categories [116, 177]. The first class are the bottom-up approaches including sequential methods and coarse-graining (homogenisation). Inthe sequential methods, parameters obtained at smaller scales are passed to the largerscales, whereas in coarse-graining microscopic processes and features are homogenisedresulting in macroscopic constitutive descriptions. The second category are the top-down approaches, where physical phenomena are incorporated by adopting particularconstitutive formats based on observations at smaller scales. The third class are theconcurrent methods, in which multiple scales are considered simultaneously.

In polycrystalline metals, in addition to dislocations, other material defect exist. Animportant defect is the grain boundary, which is the interface between grains (crystals)

Page 16: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

1.1 Background and motivation 3

with a different orientation of the atomic lattice. Grain boundaries have a major influenceon the material properties and the macroscopic behaviour, in particular when theseinterfaces constitute a relatively large fraction of the material volume, e.g. in fine-grainedmetals, thin films and micro-devices. In these situations, so called size effects oftenemerge. This is typically the case when a characteristic length scale, such as the specimensize, approaches a microstructural length scale, such as the grain size. Another well-known example of a size effect is the Hall-Petch effect, inducing an increased materialstrength resulting from grain size refinement. However, when the average grain sizereduces, the role of grain boundaries and their interaction with dislocations become ofkey importance.

Various types of interaction mechanisms between dislocations and grain boundaries havebeen identified from small-scale experiments and computer simulations [145, 175]. Ingeneral, the details of the interaction mechanisms that take place depend upon the ini-tial structure of the grain boundary, the magnitude of applied forces, and the degreeof available thermal activation. Grain boundaries can act as obstacles to the motion ofdislocations in the crystal lattice leading to a dislocation pile-up, or as sources of dis-locations generating dislocations in the adjacent grains and at the interface. A latticedislocation may also be transmitted through the interface from one grain to the next in adirect or indirect manner, possibly leaving behind a residual defect in the grain boundary.Furthermore, a grain boundary can act as a sink that absorbs lattice dislocations, whichcan redistribute along the interface. All of these interaction mechanisms have a strongeffect on the material properties. The precise account of these grain boundary mecha-nisms is one of the current challenges in the modelling of the plasticity of polycrystallinemetals [117].

In terms of continuum scale modelling, the incorporation of the behaviour of grain bound-aries is still in its infancy. For example, conventional continuum plasticity models, eventhose enhanced with gradient effects and interface behaviour, do not explicitly accountfor the crystallographic misorientation across a grain boundary, which is an importantquantity defining micro-scale interactions between dislocations and grain boundaries.Classical crystal plasticity models overcome this deficiency, by incorporating the crystal-lographic structure in a natural way. However, they generally do not account for grainboundary phenomena, i.e. grain boundaries are only incorporated as planes where thecrystallographic orientation changes. Gradient enhanced crystal plasticity partially re-solves this limitation through higher-order boundary conditions, yet, mostly limited toeither impenetrable or completely transparant grain boundaries. More recently, thesemodels have been extended by the inclusion of grain boundary interface mechanics, butthe constitutive equations at the grain boundary, e.g. the continuum scale expressionof the grain boundary energy, still lack a solid physical background. At the same time,atomistic simulations have indicated that the grain boundary energy depends on the grainboundary structure and the defect content evolution. Hence, a multiscale description isrequired to establish a physically sound continuum scale description of the interactionbetween plasticity and grain boundaries.

Page 17: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4 1 Introduction

1.2 Objective and outline of the thesis

The main objective of this work is the development of a multiscale modelling frameworkfor the plasticity of grain boundaries within the context of dislocation based gradientenhanced crystal plasticity. To this end, the interactions between dislocations and grainboundaries enter the framework in a continuum fashion through the higher-order bound-ary conditions. The grain boundary plasticity model gradually increases in complexity ineach subsequent chapter. The model framework, which consists of a set of highly nonlin-ear coupled equations, is numerically implemented using the finite element method. Inorder to investigate the influence of the resulting model on the microscopic and macro-scopic behaviour at the different stages of the model development, a bicrystal benchmarkproblem is used throughout the thesis.

In Chapter 2 the foundations of the grain boundary interface framework are developedthrough the formulation of interface balance equations and thermodynamically consis-tent constitutive equations within the adopted microstructurally motivated strain gra-dient crystal plasticity model. A key ingredient is the grain boundary flow rule, whichis an evolution equation of a scalar continuum scale quantity on each slip system at theinterface, representing the net effect of the different fine-scale phenomena taking placeat the grain boundary. The plastic slip at the grain boundary is governed by the com-petition between three interface microforces, i.e. bulk-induced, energetic and dissipativemicroforces. In the numerical analysis, the interface constitutive parameters and crystalorientations are systematically varied, providing insights into the impact of the grainboundary behaviour on plastic deformation.

Driven by the need for physically based continuum scale descriptions of the grain bound-ary energy to be used in continuum modelling frameworks, Chapter 3 describes a mul-tiscale approach leading to a continuum representation of the initial grain boundarystructure defect content and energy based on the output of atomistic simulations. Dif-ferent atomistic and continuum concepts related to the grain boundary structure areconsolidated in the development of the model. The proposed approach introduces anintrinsic net defect density, representing the grain boundary structure, which is relatedto the atomistically calculated grain boundary energies through a logarithmic expression.The ability of the multiscale approach to capture the grain boundary energies in the com-plete misorientation range of three symmetric tilt grain boundary systems in aluminiumand copper is demonstrated.

The exploitation of the atomistic-to-continuum approach (Chapter 3) into the grainboundary extended crystal plasticity model (Chapter 2) is elaborated in Chapter 4.The incorporation requires a generalisation of the continuum expression for the initialgrain boundary energy. An analytical study of the resulting initial energetic interfacemicroforce is performed, followed by a numerical study examining the particular role andinfluence of the grain boundary energetic contributions on the macroscopic response. Inthis chapter it is assumed that the grain boundary net defect density remains a localproperty at a material point at the interface.

Page 18: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

1.2 Objective and outline of the thesis 5

Chapter 5 describes the development of a continuum model that incorporates the defectredistribution along the grain boundary into the grain boundary plasticity model of theprevious chapters. A diffusion type evolution law is proposed for the redistribution ofthe net defect within the grain boundary, which allows for the nonlocal relaxation of thelocal build-up of the grain boundary net defect. The continuum model consists of a set ofequations, including a balance and constitutive equations, for a vectorial grain boundarynet defect field along the grain boundary surface. The incorporation of a strongly nonlocalinteraction term leads to an integro-differential equation for the net defect evolution,which is further reformulated into a differential form. A numerical analysis investigatesthe effect of the defect redistribution and the influence of the constitutive parametersassociated with this relaxation mechanism on the plastic deformation of a bicrystal.

Finally, in Chapter 6, the main conclusions are summarised and an outlook is given forfuture work.

Page 19: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary
Page 20: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Chapter 2

Grain boundary interface mechanicsin strain gradient crystal plasticity1

Abstract

Interactions between dislocations and grain boundaries play an important role in theplastic deformation of polycrystalline metals. Capturing accurately the behaviour ofthese internal interfaces is particularly important for applications where the relativegrain boundary fraction is significant, such as ultra fine-grained metals, thin films andmicro-devices. Incorporating these micro-scale interactions (which are sensitive to anumber of dislocation, interface and crystallographic parameters) within a macro-scalecrystal plasticity model poses a challenge. The innovative features in this chapter include(i) the formulation of a thermodynamically consistent grain boundary interface modelwithin a microstructurally motivated strain gradient crystal plasticity framework, (ii) thepresence of intra-grain slip system coupling through a microstructurally derived internalstress, (iii) the incorporation of inter-grain slip system coupling via an interface energyaccounting for both the magnitude and direction of contributions to the residual defectfrom all slip systems in the two neighbouring grains, and (iv) the numerical implemen-tation of the grain boundary model to directly investigate the influence of the interfaceconstitutive parameters on plastic deformation. The model problem of a bicrystal de-forming in plane strain is analysed. The influence of dissipative and energetic interfacehardening, grain misorientation, asymmetry in the grain orientations and the grain sizeare systematically investigated. In each case, the crystal response is compared with ref-erence calculations with grain boundaries that are either ‘microhard’ (impenetrable todislocations) or ‘microfree’ (an infinite dislocation sink).

1Reproduced from [187]: P.R.M. van Beers, G.J. McShane, V.G. Kouznetsova, and M.G.D. Geers.Grain boundary interface mechanics in strain gradient crystal plasticity. Journal of the Mechanics andPhysics of Solids, 61:2659–2679, 2013.

7

Page 21: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

8 2 Grain boundary interface mechanics in strain gradient crystal plasticity

2.1 Introduction

Internal interfaces (such as grain and phase boundaries) play an important role in theplastic deformation of polycrystalline metals. They provide obstacles to dislocation glide,with a resistance that depends on a number of material and crystallographical parame-ters. As a consequence, they constrain the plastic deformation of adjacent grains, whosedifferent crystallographic orientations and preferential slip directions result in complexintra-granular strain fields. The high local stresses developing at the tip of a pile-up [105]can influence local plasticity and damage, and is obviously sensitive to the ability of dis-locations to transmit through the boundary. The defect structure of a grain boundary(GB) can also act as a dislocation source, promoting plastic slip [194].

The challenge addressed by the current work is the development of a macroscopic crystalplasticity model which accounts for the fine-scale grain boundary micromechanics. Anumber of applications motivate this. Metallic thin films and micro-devices can have agrain size comparable with some of the component dimensions, and accurate resolution ofinterface micromechanics may therefore be important for predicting device performance.For fine-grained and nano-crystalline metals, the large grain boundary area plays a pivotalrole in plastic deformation. As outlined in the review by Meyers et al. [121], a conventionalHall-Petch relationship breaks down for grain sizes below 1 µm, and the mechanisms bywhich plastic deformation progresses at these small grain sizes remains the subject ofcontinued investigation. Accurate models of grain boundary behaviour during plasticdeformation are required in order to capture these phenomena.

Experimental observations and, more recently, numerical simulations at the atomistic anddiscrete dislocation scales, indicate that the micro-scale interactions between dislocationsand grain boundaries are complex and diverse. To motivate the model development, firstthese experimental and model observations of the micro- and meso-scale behaviour ofgrain boundaries are briefly reviewed. Existing strategies for bridging the length scalesto capture the continuum level response are then summarised.

2.1.1 Interactions between dislocations and grain boundaries

Microscopic investigations of dislocation motion in the vicinity of grain boundaries havebeen used to develop an understanding of the interaction mechanisms and the criteria forthe propagation of slip across the interface. Livingston and Chalmers [111], studying pile-ups at a bicrystal interface, suggest that slip is induced in the adjacent grain in accordancewith a resolved shear stress criterion: slip propagates on the slip plane on which themagnitude of the resolved shear stress induced by the pile-up is greatest. Later studiesemployed transmission electron microscopy (TEM) to track the motion of individualdislocations during interaction with a grain boundary [100, 101, 158]. A number ofinteraction mechanisms were identified, most common of which is the initiation of slipnear the incoming pile-up on the preferentially oriented slip system, leaving a residualBurgers vector at the grain boundary. Three parameters emerge as most important fordetermining the slip plane and slip systems on which dislocations are emitted into an

Page 22: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.1 Introduction 9

adjacent grain: (i) minimising the angle between the projections onto the grain boundaryof the incoming and outgoing slip systems, (ii) maximising the resolved shear stress inthe slip direction and (iii) minimising the residual defect at the interface, required byconservation of the Burgers vector. These TEM investigations are generally restrictedto very local phenomena. Sun et al. [170, 171] and Zaefferer et al. [215] use orientationimaging microscopy (OIM) to measure spatial variations in the densities of geometricallynecessary dislocations (GNDs) over a larger area in the vicinity of an interface in adeformed aluminium bicrystal.

Although microscopy can provide insights into governing interaction mechanisms, it isdifficult to quantify the resistance to slip posed by an interface. This issue has beenaddressed using nano-indentation in the vicinity of interfaces. Hardness variations acrossinterfaces in various material and interface combinations have been reported [99, 168,204]. While providing a measure of the obstruction to slip provided by the interface,the results appear to be sensitive to experimental conditions, such as indentation forceand the orientation of the faces of the indenter tip relative to the interface. A number ofauthors have also studied ‘pop-ins’, interpreted as the sudden release of a pile-up througha grain boundary [167, 212]. Although challenging to interpret, these results indicate thatgrain boundaries pose an obstacle to slip of finite strength, and can be overcome withsufficient applied stress.

More recently, numerical approaches at the micro- and meso-scales have been used tostudy the interactions between dislocations and grain boundaries. At the atomistic scale,molecular dynamics (MD) has been used to study individual dislocation interactions. Thestudy of de Koning et al. [44] reveals that transmission of a dislocation loop across a tiltboundary is blocked if the misorientation angle exceeds 20◦. Furthermore, de Koninget al. [45] show that the resistance to transmission is higher for an asymmetric tiltboundary compared to a symmetric interface. The initial interface energy (which ishigher in the asymmetric case) appears to play a role. The calculations of Jin et al.[89] show that the recombination of partial dislocations prior to absorption into a grainboundary (which is influenced by the stacking fault energy) appears to influence thecritical applied stress for transmission.

At the meso-scale, discrete dislocation dynamics (DDD) has been used to study themotion of a large number of individual dislocations and their interactions with interfaces,for example [10, 17, 103]. In these studies, grain boundaries are treated as hard obstaclesto the motion of individual dislocations, but stresses local to the head of a pile-up mayactivate sources in the adjacent grain which reside on favourably oriented slip planes. Thepropagation of slip therefore depends on the activation strength of dislocation sources,their locations and the elastic stress fields which develop due to slip incompatibility.Recently Li et al. [106] have introduced penetrable interfaces within a DDD calculation.A criterion is specified by which a dislocation may transmit across a grain boundary,in the form of a threshold shear stress which accounts for the energy balance for thetransmission event.

Page 23: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

10 2 Grain boundary interface mechanics in strain gradient crystal plasticity

2.1.2 Modelling grain boundary behaviour at the continuum scale

The problem of interest is fundamentally multiscale in nature: dislocations interactingwith interfaces at the micro-scale influence the macroscopic response of the material.This poses a challenge in accounting for the fine-scale behaviour across different lengthscales. Three main strategies have emerged recently in order to capture grain boundarybehaviour at the continuum scale.

(i) Multiscale modelling via embedded, higher resolution domains. One approach to theproblem is to directly couple domains with different degrees of resolution (micro-, meso-and macro-scales), passing information between them at the boundaries. Dewald andCurtin [47] use a coupled molecular dynamics (MD) and discrete dislocation dynamics(DDD) approach to model the interactions between dislocations and a symmetric tiltboundary in aluminium. Individual dislocations exit the DDD domain and are modelledin greater refinement via MD as they approach the interface, and vice versa. Wallin et al.[196] couple discrete dislocation and continuum level calculations. These approaches havea high level of local refinement, providing insights into grain boundary micromechanicsfor multiple dislocation interactions. For example, Dewald and Curtin [47] reveal grainboundary defect accumulation and relaxation mechanisms. Refinements to the slip trans-mission criteria of Lee et al. [100, 101] are also proposed which account for the influence ofa multiple dislocation pile-up. In particular, the interface normal stress component andthe modification of the interface structure due to the absorption of multiple dislocationsare shown to have an influence. However, the method introduces significant modellingchallenges at the MD/DDD domain interfaces, to avoid introducing non-physical arti-facts.

(ii) Modelling grain boundaries as embedded continuum plasticity zones of finite volume.A lower resolution but computationally less demanding strategy (and therefore moresuited to larger boundary value problems) is to model grain boundaries as an embeddedcontinuum plasticity zone of finite volume, but having a distinct constitutive behaviourfrom the bulk crystal. Ashmawi and Zikry [6] adopt a conventional (i.e. not strain gra-dient dependent) crystal plasticity framework for the bulk crystal and grain boundaryzones featuring mobile and immobile dislocation densities as internal variables. A crite-rion is specified for transmitting dislocation density across the interface between the bulkand grain boundary zones, based on slip system misorientation. Ma et al. [113] adopt astrain gradient crystal plasticity framework. Slip resistance in the grain boundary zone isreformulated to include an energetic penalty to slip (entered via the activation energy inthe flow rule) which accounts for the misorientation of the adjacent grains. This energypenalty is representative of the residual defect which would remain at a grain boundaryafter absorption or transmission of a dislocation. A disadvantage of this type of approachis that a realistic grain boundary zone width (of the order of the Burgers vector) wouldrequire an impractically small finite element mesh size.

(iii) Specification of internal boundary conditions within strain gradient plasticity theo-ries. Continuum level plasticity theories which introduce strain gradient dependence inorder to account for experimentally observed size effects are well established (see [95] for arecent comparative review). These theories may be phenomenological in nature or, more

Page 24: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.1 Introduction 11

recently, microstructurally motivated, accounting directly for gradient dependence viathe evolution of populations of geometrically necessary dislocations (GNDs) [5]. Includ-ing strain gradient dependence requires ‘higher-order’ boundary conditions (BC) to bespecified in addition to the conventional ones. There is scope to use these boundary con-ditions to control plastic slip at interfaces, effectively reducing the grain boundary zoneto an interface layer. This is the approach adopted in the current investigation. In thefollowing, existing strategies for specifying higher-order boundary conditions that governinterface behaviour within strain gradient plasticity frameworks are briefly reviewed.

2.1.3 Higher-order interface BCs: review of existing approaches

To date, the majority of studies adopt ‘microscopically simple’ boundary conditions, i.e.conditions which expend zero interface work, of which there are two limiting cases:

1. A ‘microfree’ boundary, such as a free surface, which requires zero GND density onthe interface. Physically, this represents an interface at which dislocations are freeto escape, essentially an infinite sink.

2. A ‘microhard’ condition would correspond to zero plastic strain at the interface or,in the case of crystal plasticity frameworks, zero slip on systems interacting withthe interface. Refer to [95] and [54] for discussions on the geometric considera-tions for slip system-interface interactions. In other words, the interface poses animpenetrable obstacle to dislocations.

The ‘microhard’ case has been commonly applied to grain boundaries (see [16, 56] forexample). It is possible to relax this constraint, by specifying constitutive relationshipscoupling the jumps in higher-order tractions (moment tractions) and plastic strain acrossthe grain boundary to the interface work rate, consistent with appropriate higher-orderbalance equations. A third ‘intermediate’ condition, which also expends zero interfacework, is obtained when assuming continuity of both the higher-order tractions and plasticstrain, for example [38]. Since it is rather difficult, both practically and physically, toconstrain individual slip systems with such a global constraint, as would be required inthe context of the framework developed here, this condition has not been further exploredin this work.

Gudmundson [74] outlines such a treatment of boundary conditions within a phenomeno-logical strain gradient plasticity framework. Interface moment tractions are defined workconjugate to the local plastic strain. Both the moment traction and plastic strain arepermitted to be discontinuous across the interface. An energetic form for these momenttractions is proposed, with an interface free energy that is quadratic in the local plasticstrain at the interface. An interface length scale parameter (taken as a material constant)is specified which controls the rate of accumulation of interface free energy with plasticstraining. The investigation focuses on an elastic-plastic interface, whereby coupling ofslip activity on either side of an interface is not considered. The model is implementednumerically by Fredriksson and Gudmundson [62, 63], to analyse the problem of a plastic

Page 25: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

12 2 Grain boundary interface mechanics in strain gradient crystal plasticity

thin film on an elastic substrate. The authors develop the interface model further in [64].A more general power-law form for the interface energy is proposed, as well as dissipativemoment tractions. Aifantis and Willis [2] also propose energetic constitutive relationsfor internal interfaces which, similar to Fredriksson and Gudmundson, are functions ofthe local plastic strain. Solutions are obtained for the one-dimensional plastic deforma-tion of a medium containing internal interfaces. This is further developed in [3], withcalculations of the effective response (in 1D) of materials with a random distribution ofinternal interfaces accounting for the influence of the interface energy. The treatment ofcoupling between grain behaviour on either side of an interface is handled by constrain-ing plastic strain to be continuous across an interface. However, this may in general bean over-constraint for a grain boundary. Recently, Fleck and Willis [59] propose a moreelaborate form for the interface free energy which considers this cross-coupling further.Plastic strain is permitted to be discontinuous, with an interface energy that dependson both the jump in plastic strain and the average plastic strain across the interface.However, this is not pursued in detail in that work.

A drawback in the preceding studies is the inability to explicitly account for crystallo-graphic misorientation across a grain boundary, which is an important quantity to beconsidered in modelling micro-scale interactions between dislocations and grain bound-aries. Employing crystal plasticity overcomes this deficiency.

Cermelli and Gurtin [30] propose constitutive relationships for internal interfaces withina strain gradient crystal plasticity framework incorporating densities of geometricallynecessary dislocations (GNDs). This approach facilitates comparison with experimentalinvestigations such as those of Sun et al. [170, 171]. Visco-plastic power-law relationsare proposed linking slip rates adjacent to an interface to the interface microtractions.These microtractions are in turn related to the GND densities on each slip system via thebulk crystal constitutive relations. Furthermore, constitutive relations are proposed formacroscopic slip at an interface, i.e. governing displacement discontinuities. Analyticalsolutions are obtained for the problem of a semi-infinite crystal with two active slipsystems bonded to a rigid substrate. The interface jump conditions are developed inmore detail by Gurtin and Needleman [78]. A more elaborate description of the interfacekinematics is proposed, employing an interface defect tensor, analogous to the Nye tensorfor the bulk crystal. Dissipative constitutive relations for plastic slip adjacent to theinterface are introduced, dependent on the rate of change of this interface defect tensor,physically representing the flow of Burgers vector in and out of the grain boundary.A consequence of this approach is that adjacent grains are coupled in their interfaceresponse: slip on either side of the boundary contributes to the rate of change of theinterface defect tensor. This model has been developed further by Gurtin [77]. Anenergetic contribution to the interface microtractions is introduced, based on an interfaceenergy that is quadratic in the interface defect tensor. This additional contributionis interpreted as an interface back stress, which also couples slip activity in the twograins. ‘Slip interaction moduli’, which characterise the slip system coupling, are derived,incorporating intra-grain and inter-grain contributions.

Although qualitative descriptions of the interface response are given in the precedingcrystal plasticity studies, the proposed models are not numerically implemented. How-

Page 26: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.1 Introduction 13

ever, it is desired to explore the capability of an interface model to capture complexbehaviour.

2.1.4 Outline and scope of the chapter

In the present chapter, the modelling of grain boundary behaviour through the develop-ment of microstructurally motivated higher-order internal boundary conditions within astrain gradient crystal plasticity framework is pursued. The following specific contribu-tions are made:

1. An interface constitutive model is developed which incorporates intra-grain slip sys-tem coupling consistent with the ‘full internal stress’ approach developed in the sin-gle crystal plasticity model of Bayley et al. [16]. This provides a microstructurallymotivated basis for slip system interactions at the grain boundary. Furthermore, asa non-local approach, it allows the cumulative effect of a dislocation pile-up to beaccounted for in determining slip activity at the interface. The resulting interfaceslip system couplings are distinct from the formulations of Gurtin and Needleman[78] and Gurtin [77].

2. The proposed interface constitutive model accounts in a transparent manner for thegeometric constraints on slip system-interface interactions identified by Kuroda andTvergaard [95] and Erturk et al. [54]. These interaction conditions influence both(i) the interface dissipation and (ii) the evolution of the interface energy.

3. The interface energy is formulated in terms of an accumulated interface defect,analogous to the interface net Burgers vector. This provides a micromechanicallysound basis for inter-grain slip system coupling at the interface. Consistent with theexperimental and numerical observations outlined above, the interface net defect isallowed to either retard or promote slip on intersecting slip systems, depending onthe defect magnitude and relative orientations of the Burgers vectors.

4. The interface constitutive model is numerically implemented in order to investigatedirectly the influence of the grain boundary constitutive model on plastic deforma-tion. The model problem of a bicrystal consisting of alternating misoriented grainsis considered. Interface constitutive parameters and crystal orientations are system-atically varied, providing insights into the influence of grain boundary behaviouron plastic deformation, at both macroscopic and microscopic levels.

The chapter is structured as follows. In Section 2 the bulk crystal strain gradient plas-ticity formulation is summarised. The grain boundary constitutive model is developed inSection 3. Section 4 describes the numerical implementation of the constitutive model.The results of an analysis of the role of grain boundaries during the plastic deformationof a bicrystal are described. Conclusions are provided in Section 5.

Page 27: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

14 2 Grain boundary interface mechanics in strain gradient crystal plasticity

2.2 Bulk crystal formulation

In this section, the strain gradient crystal plasticity model developed by Evers et al. [56]and Bayley et al. [16], shown to be thermodynamically consistent by Erturk et al. [54], isbriefly summarised. This framework is adopted for the bulk crystal in the current work.Connections are made between this constitutive model and the balance equations withina strain gradient crystal plasticity framework, to provide a more general basis for theinterface equations derived subsequently. In the current work, attention is restricted tosmall deformations.

2.2.1 Crystal geometry and dislocation sign convention

A crystal plasticity approach is adopted, with dislocation densities on individual slipsystems acting as key variables. Consider a grain A containing slip systems α, eachdefined by a slip plane normal mα and orthogonal unit vectors sα and lα which lie inthe slip plane. The Burgers vector is given by bsα. Edge and screw dislocations residingon α are considered to follow the ‘finish-start/right-handed’ (FS/RH) sign conventionwith respect to the reference crystal [82]: a Burgers circuit traversed in a right-handedmanner with respect to the dislocation line direction is closed by the Burgers vector,which is directed from the finish to the start of the circuit. An idealised dislocationloop is sketched in Fig. 2.1 consisting of four straight line segments (two of pure edgecharacter and two of pure screw) with the dislocation signs as indicated. If the slip planeexperiences a resolved shear stress as shown, the material within the loop undergoes apositive shear deformation γα as the loop expands. The glide directions of the varioussigned dislocation segments are shown in Fig. 2.1. The total distortion is given by thesum of elastic (e) and plastic (p) components

∇u = He + Hp . (2.1)

The plastic distortion is given by

Hp =∑α

γαsαmα . (2.2)

Geometrically necessary dislocation densities of edge and screw character, defined ρα(e)g

and ρα(s)g respectively, arise from gradients in slip as follows

ρα(e)g = ρ

α(e)g0 − 1

b∇γα · sα and ρ

α(s)g = ρ

α(s)g0 +

1

b∇γα · lα . (2.3)

Here, the unit vector lα = sα×mα and ρα(e)g0 and ρ

α(s)g0 represent the initial GND densities.

Page 28: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.2 Bulk crystal formulation 15

(+) edge

(+) screw

({) screw

({) edge

slip plane ®

glide directions dislocation linedirection

positive resolved shear stressdislocation loop

Figure 2.1: An idealised dislocation loop on slip plane α. The slip plane unit vectors and the adoptedsign convention for dislocation loop segments (here assumed to be either pure edge or pure screw) areshown.

2.2.2 Bulk crystal balance equations

First the principle of virtual work is employed to outline the balance relationships withinthe strain gradient crystal plasticity framework.2 Consider a bicrystal domain consistingof two grains A and B separated by an internal interface Γ with normal direction ngb,which points from A to B. The domain has total volume V = V A ∪ V B and externalsurface S = SA∪SB with outward normal n. Strain gradient dependence is incorporatedvia the gradient of the plastic slip on slip plane α, and hence the rate of internal workreads

W int =

∫V

σ : Ee dV +∑α

∫V

(πα γα + ξα ·∇γα) dV +

∫Γ

(φgbM + φgbm

)dΓ .

(2.4)

Here, the Cauchy stress σ is work conjugate to the elastic strain

Ee =1

2

(He + HeT

). (2.5)

The ‘microforce’ πα and ‘microstress’ ξα are work conjugate to the plastic slip andplastic slip gradient ∇γα, respectively. The rate of internal work at the grain boundaryconsists of two parts φgbM and φgbm that account for the work done due to macroscopic and

microscopic interface mechanisms, respectively. The role of φgbM and φgbm will be describedsubsequently. The rate of external work on the bicrystal domain is

W ext =

∫S

text · u dS +∑α

∫S

χαext γα dS , (2.6)

2A form for the virtual power balance is adopted consistent with the group of theories categorised byKuroda and Tvergaard [95] as ‘work-conjugate’ type. The equivalence between this and the ‘non-workconjugate’ frameworks of Evers et al. [56] and Bayley et al. [16], on which the current model is based,has been shown by Kuroda and Tvergaard [95] and Erturk et al. [54].

Page 29: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

16 2 Grain boundary interface mechanics in strain gradient crystal plasticity

where text and χαext are external tractions and microtractions, respectively. Equatinginternal and external work rates leads, after eliminating the elastic strain rate and ap-plication of the divergence theorem, to the following balance equations within the bulkcrystals V A and V B

∇ · σ = 0 and πα − σ : sαmα −∇ · ξα = 0 . (2.7)

2.2.3 Bulk constitutive equations

As has been shown by Kuroda and Tvergaard [95] and Erturk et al. [54], the bulk crystalbalance equation (2.7b) is equivalent to the definition of the effective shear stress ταeff ona slip plane α, as used in [56] and [16]

ταeff = τα − ταb , (2.8)

with

ταeff = πα , τα = σ : sαmα and ταb = −∇ · ξα . (2.9)

In the above relations τα and ταb are the Schmid and the back stress on a slip system α,respectively. The Schmid stress τα follows from an elastic relationship for the Cauchystress, projected on the slip plane α

σ = 4C : Ee , (2.10)

where 4C is the fourth order elasticity tensor and Ee the elastic strain.

The back stress, Eq. (2.9c), is derived from a microstructurally motivated internal stressσint

ταb = −σint : sαmα . (2.11)

The internal stress can be interpreted as a fine-scale correction, accounting for microstruc-tural effects not resolved by the macroscopic elastic strain field, e.g. dislocation pile-ups.In the frameworks of Evers et al. [56] and Bayley et al. [16], spatial gradients in thedensities of geometrically necessary dislocations (GNDs) are shown to induce a net re-solved shear stress on the slip plane. In this work the ‘full internal stress’ approach ofBayley et al. [16] is adopted, in which gradients in the GND densities on all slip systemsare accounted for in the formulation of the internal stress tensor, and hence the resolvedback stress. In the derivation, the stress fields of dislocations in an infinite elasticallyisotropic medium are used. Further, by assuming a linear variation in the GND densitiesof both edge and screw type within a radius R of the relevant material point, the follow-ing expression for the internal stress accounting for contributions from all slip systems ζis obtained [16]

σint =GbR2

8 (1− ν)

∑α

∇ρα(e)g · 3Cα(e) +

GbR2

4

∑α

∇ρα(s)g · 3Cα(s) . (2.12)

Page 30: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.2 Bulk crystal formulation 17

The third order slip system orientation tensors are

3Cα(e) = 3mαsαsα + mαmαmα + 4νmαlαlα − sαsαmα − sαmαsα ,

3Cα(s) = −mαlαsα −mαsαlα + lαmαsα + lαsαmα , (2.13)

Parameter G is the shear modulus and ν is Poisson’s ratio. Substituting Eqs. (2.11) and(2.12) into Eq. (2.9c) yields an expression for the microstress ξα [54]

ξα =

[GbR2

8 (1− ν)

∑ζ

ρζ(e)g

3Cζ(e) +GbR2

4

∑ζ

ρζ(s)g

3Cζ(s)

]: sαmα . (2.14)

The effective stress results from a visco-plastic power-law form

ταeff = sα( |γα|γ0

)msign (γα) . (2.15)

Parameters γ0 and m are positive constants, and sα is the slip resistance. The slipresistance sα for slip system α reflects the obstruction of dislocation glide through short-range interactions of all dislocations on coplanar and intersecting slip systems. Boththe densities of geometrically necessary dislocations (GNDs), ρξg, and statistically stored

dislocations (SSDs), ρξs , contribute to the slip resistance [5]

sα = Gb

√∑ξAαξ

∣∣ρξs ∣∣+∑

ξAαξ

∣∣ρξg∣∣ , (2.16)

where Aαξ represents the full dislocation interaction matrix as given by Franciosi andZaoui [61]. The generalised form of Essmann and Mughrabi [55] is employed to describe

the evolution of the SSD densities ρξs

ρξs =1

b

(1

Lξ− 2ycρ

ξs

) ∣∣γξ∣∣ with ρξs (t = 0) = ρξs0 . (2.17)

The dislocation accumulation is governed by the average dislocation segment length

Lξ =K√∑

ζ Hξζ∣∣ρζs ∣∣+

∑ζ H

ξζ∣∣ρζg∣∣ , (2.18)

where Hξζ are the immobilisation coefficients, structured analogously to the interactioncoefficients Aαξ. The dislocation annihilation is controlled by a critical annihilationlength yc.

Making use of Eq. (2.8) to substitute for the effective stress in Eq. (2.15) the followingslip system flow rule for the bulk crystal is obtained

γα = γ0

( |τα − ταb |sα

) 1m

sign (τα − ταb ) . (2.19)

Note that Evers et al. [56] include a thermal activation term in the effective stress,which is omitted here for brevity (isothermal conditions are assumed). Evers et al. [56]and Bayley et al. [16] describe the solution method for the strongly coupled system ofequations, (2.3), (2.10), (2.11), (2.15) and (2.19).

Page 31: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

18 2 Grain boundary interface mechanics in strain gradient crystal plasticity

2.3 Interface formulation

2.3.1 Interface balance equations

First, reconsider the statement of virtual work presented in Section 2.2.2, above. Equat-ing internal and external work rates for the bicrystal domain, Eqs. (2.4) and (2.6), andapplying the divergence theorem as before, the following interface balance relationshipsemerge. On the external surfaces SA and SB

σ · n = text and ξα · n = χαext . (2.20)

Across the internal interface Γ two balance equations hold:

〚t · u〛 + φgbM = 0 , (2.21)

〚∑α

ξα · ngb γα〛 + φgbm = 0 . (2.22)

Here, 〚·〛 = (·)|A−(·)|B , indicates the jump across Γ. In this work it will be assumed thatno displacement jump occurs at the internal interface, i.e. no grain boundary sliding oropening takes place. Therefore, the ‘macroscopic’ interface balance equation (2.21) doesnot need to be considered further. The ‘microscopic’ interface balance equation (2.22)sets the stage for the development of constitutive relationships enabling plastic slip atthe grain boundary, as will be developed next.

2.3.2 Grain boundary constitutive model

As a starting point the geometric arguments of Kuroda and Tvergaard [95] and Erturket al. [54], related to the interaction of slip systems at boundaries, are taken. Thefollowing parameters are defined, which will be referred to subsequently as ‘interfacenormal slip’ components of edge (e) and screw (s) dislocations at the interface Γ betweenadjacent grains A and B

grain A: qα(e) = γα sα · ngb and qα(s) = −γα lα · ngb ,

grain B: qβ(e) = −γβ sβ · ngb and qβ(s) = γβ lβ · ngb . (2.23)

The interface normal slip q is a meso-scale representation of different fine-scale phenom-ena taking place at the boundary, e.g. absorption, emission and transmission, that arenot resolved separately at the scale of the model. Positive q is defined in the sense ofgrowth of the net boundary defect. Note that, in what follows, the term ‘interface normalslip’ is reserved for dislocation-grain boundary interactions, and should not be confusedwith macroscopic cohesive interface slip or opening 〚u〛. Recall that in the bicrystal do-main outlined previously the grain boundary normal ngb is defined to point from A toB. The superscript α represents a slip system in grain A and β a slip system in grain

Page 32: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.3 Interface formulation 19

B. Note that having both qα(e) = 0 and qα(s) = 0 implies that either (i) sα · ngb = 0and lα · ngb = 0, in which case slip plane α lies parallel to the interface, or (ii) γα = 0.Either case corresponds to no interface normal slip, i.e. no interactions between latticedislocations on a slip system α and the grain boundary. Therefore the rate of microscopicinterface work is taken to depend only on these normal components of slip, accountingfor contributions from all slip systems in both grains A and B, giving

φgbm =∑α

(∂φgbm

∂qα(e)qα(e) +

∂φgbm

∂qα(s)qα(s)

)+∑β

(∂φgbm

∂qβ(e)qβ(e) +

∂φgbm

∂qβ(s)qβ(s)

).

(2.24)

On the other hand, elaborating the jump in the microscopic interface balance equation(2.22) gives

φgbm =∑β

ξβ · ngb γβ −∑α

ξα · ngb γα . (2.25)

Eqs. (2.24) and (2.25) may be satisfied by balancing terms for each slip system separately(making use of Eq. (2.23)), leading to the sufficient conditions

∂φgbm

∂qα(e)sα − ∂φgbm

∂qα(s)lα = −ξα ,

∂φgbm

∂qβ(e)sβ − ∂φgbm

∂qβ(s)lβ = −ξβ , (2.26)

for all slip systems α and β. Note that, although Eq. (2.26) refers to individual slipsystems, cross-coupling between slip systems may be accounted for via ξ and φgbm , asdescribed subsequently. Further manipulation, i.e. projection on the orthogonal unitvectors sα or lα respectively, leads to four separate balance equations for edge and screwdislocations in grains A and B:

∂φgbm

∂qα(e)= −ξα · sα and

∂φgbm

∂qα(s)= ξα · lα ,

∂φgbm

∂qβ(e)= −ξβ · sβ and

∂φgbm

∂qβ(s)= ξβ · lβ . (2.27)

The microscopic interface work rate may be decomposed into dissipative (φdm) and ener-getic (φem) parts which satisfy

φgbm = φdm + φem with φdm ≥ 0 , (2.28)

so that

∂φgbm

∂qα(e)=

∂φdm∂qα(e)

+∂φem∂qα(e)

, (2.29)

and likewise for screw dislocations. Next, the balance equations (2.27) are expressed interms of slip system microforces, as follows

Fα(e)d + Fα(e)

e = Fα(e)bk and Fα(s)

d + Fα(s)e = Fα(s)

bk ,

Fβ(e)d + Fβ(e)

e = Fβ(e)bk and Fβ(s)

d + Fβ(s)e = Fβ(s)

bk , (2.30)

Page 33: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

20 2 Grain boundary interface mechanics in strain gradient crystal plasticity

where the individual microforces are defined as follows for grain A (likewise for grain B,substitute B for A and β for α):

(i) Bulk-induced interface microforces, related to the microstress, and hence the distri-bution of GNDs in the bulk crystal

Fα(e)bk = −ξα · sα and Fα(s)

bk = ξα · lα . (2.31)

(ii) Energetic interface microforces, related to the interface free energy which in turnmay be linked to the grain boundary structure and its evolution

Fα(e)e =

∂φem∂qα(e)

and Fα(s)e =

∂φem∂qα(s)

. (2.32)

(iii) Dissipative interface microforces, required to satisfy

Fα(e)d qα(e) =

∂φdm∂qα(e)

qα(e) ≥ 0 and Fα(s)d qα(s) =

∂φdm∂qα(s)

qα(s) ≥ 0 . (2.33)

The constitutive relationships are completed by specifying an explicit form for theseinterface microforces, as follows.

Bulk-induced interface microforces

Adopting the form of the microstress related to the microstructurally motivated internalstress (Eq. (2.14)), the bulk-induced microforces (Eq. (2.31)) are determined. For slipsystems in grain A these are given by

Fα(e)bk = −

[GbR2

8 (1− ν)

∑ζ

ρζ(e)g

3Cζ(e) +GbR2

4

∑ζ

ρζ(s)g

3Cζ(s)

]∣∣∣∣A

∴ sαmαsα ,

Fα(s)bk =

[GbR2

8 (1− ν)

∑ζ

ρζ(e)g

3Cζ(e) +GbR2

4

∑ζ

ρζ(s)g

3Cζ(s)

]∣∣∣∣A

∴ sαmαlα . (2.34)

For grain B substitute B for A and β for α. The third order orientation tensors 3Cζ(e)

and 3Cζ(s) are defined in Eq. (2.13). Note that the GND densities on all slip systemsζ on the grain A side may influence the microforce on a particular system α. The effectof the pile-up structure with its internal stress enters the interface model through thesebulk-induced microforces that depend on ξα.

Energetic interface microforces

The structure and energy of a grain boundary depends on both its initial state, closelylinked to the misorientation of the adjacent grains, and the subsequent deformationhistory of these grains. When a lattice dislocation interacts with a grain boundary, the

Page 34: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.3 Interface formulation 21

structure of the interface is altered. A residual defect remains, which is necessary for theconservation of the Burgers vector [101]. Here, a simple form is proposed for the grainboundary structure which employs the normal slip components q as a convenient measureof interface normal slip activity. The following vectorial measure of the net defect at apoint on the grain boundary is defined

Bgb = B0 + Ceb∑α

(qα(e) + qα(s)

)sα + Ceb

∑β

(qβ(e) + qβ(s)

)sβ . (2.35)

The quantity B0 represents the initial grain boundary structure. In the low-angle limit,this initial defect could be obtained, for example, from an equivalent dislocation arraymodel for the grain boundary, such as the Read-Shockley model [150]. Physically, thisrepresents the contribution of ‘intrinsic’ grain boundary dislocations [13, 82]. Note thatin the low-angle case, the net Burgers vector is an equivalent measure to the surfacedislocation density, both related via the Frank-Bilby equation [175]. In the case of ahigh-angle boundary, the Read-Shockley model is no longer applicable and one has toresort to alternative descriptions. The detailed specification of B0, accounting for grainboundary microstructure for a range of grain boundary types, and the assessment of itsinfluence on interface normal slip activity, is the subject of a parallel investigation. Inthis chapter it is opted to omit the influence of the initial interface structure, and restrictattention to the influence of normal slip activity at a grain boundary. Hence, the formof the initial grain boundary defect is not elaborated further here.

The contribution of lattice dislocations interacting with the grain boundary from grainA (i.e. ‘extrinsic’ grain boundary dislocations) will have Burgers vector direction sα foreither edge or screw type. The cumulative defect magnitude will depend on the net flow,which scales with qα(e) and qα(s). Note that slip in either of the grains A and B whichborder the grain boundary are accounted for via qα and qβ , respectively. The interfacelength scale Ceb is a modelling parameter.

A free energy based on this interface defect is defined as follows

φem =1

2

G

CebBgb ·Bgb . (2.36)

Hence, the energetic microforces for slip system α in grain A are

Fα(e)e = GBgb · sα

= GB0 · sα + CeGbsα ·[∑

ξ

(qξ(e) + qξ(s)

)sξ]∣∣∣∣A

+

CeGbsα ·[∑

ζ

(qζ(e) + qζ(s)

)sζ]∣∣∣∣B

,

Fα(s)e = Fα(e)

e . (2.37)

For grain B, substitute B for A and β for α. The microforce therefore depends onthe orientations of the slip planes in both crystals (via the slip direction vectors s) andthe orientation of the interface plane Γ relative to these crystals (via the normal slipcomponents q).

Page 35: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

22 2 Grain boundary interface mechanics in strain gradient crystal plasticity

Dissipative interface microforces

A visco-plastic power-law form is chosen for the dissipative interface microforces, in asimilar manner to the effective stress in the bulk crystal. For edge dislocations on slipsystem α in grain A, let

Fα(e)d = Rα(e)

(∣∣qα(e)∣∣

q0

)nsign

(qα(e)

)(2.38)

For screw dislocations substitute (s) for (e), and for grain B, substitute B for A andβ for α. Note that the positive dissipation requirement, Eq. (2.33), is satisfied. Thequantity Rα(e) (units Nm−1) is the interface normal slip resistance for edge dislocationson slip system α. The parameters n and q0 are positive constants (equivalent to the bulkcrystal parameters m and γ0).

2.3.3 Discussion

Substituting the expressions for the slip system dissipative microforces (Eq. (2.38)) intothe balance equations (2.30) and rearranging yields the following interface flow rules, i.e.evolution equations for the normal slip components on slip system α on the grain A sideof Γ (the expressions for grain B are obtained by substituting β for α):

qα(e) = q0

(∣∣Fα(e)bk −Fα(e)

e

∣∣Rα(e)

) 1n

sign(Fα(e)

bk −Fα(e)e

),

qα(s) = q0

(∣∣Fα(s)bk −Fα(s)

e

∣∣Rα(s)

) 1n

sign(Fα(s)

bk −Fα(s)e

). (2.39)

From Eq. (2.39) it is apparent that the bulk-induced microforce Fbk can be interpretedphysically as the driving force for interface normal slip, with the energetic microforceFe acting as an interface ‘back stress’ that evolves with the defect content of the grainboundary. Qualitatively, slip will occur at the interface when the net driving force onthat slip system, given by Fbk − Fe, exceeds a threshold value, governed by the slipresistance R.

In its simplest form, the slip resistance R might be taken to be a material constant.However, there are a number of microstructural mechanisms that would justify morecomplex dependencies. Cermelli and Gurtin [30] propose a dependence on the slip sys-tem orientation, the orientation of the boundary with respect to the grains, the relativemisorientation of the grains, and the net accumulated slip from both grains at the grainboundary. Another model for slip transmission at interfaces relates to the activation ofsources in adjacent grains by the internal stress induced by a pile-up [10]. R could incor-porate the nucleation strength of dislocation sources close to the grain boundary. It mightalso be taken to capture the additional work required to overcome the configurational

Page 36: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.4 Numerical investigation 23

changes necessary for a dislocation to enter the grain boundary or to be emitted fromit. Such changes might include the recombination or dissociation of partial dislocations[89], which depends on the stacking fault energy of the crystal. There is scope to refinethese definitions through input from numerical calculations at smaller length scales, suchas MD. These elaborations will not be considered further in the current work.

The proposed form of the energetic microforce Fe, Eq. (2.37), provides a simple rep-resentation of the interface defect that captures both the magnitude and orientation ofcontributions from all interacting slip systems from both sides of the interface. It iscapable of responding in a manner that is representative of a number of experimentalobservations. Firstly, experiments show that slip activity on a given system is reducedif it results in the accumulation of an interface defect [101]. In the model, the energeticmicroforce will increasingly penalise slip on a particular slip system if the interface defectis well aligned with its Burgers vector. Secondly, an initial interface defect is included inthe model, which may initiate slip in the absence of a pile-up, effectively representing adislocation source. This has been observed experimentally by Venkatesh and Murr [194].However, some interface features are omitted by the current model. Firstly, the energeticmicroforce depends only on the interface defect at that point, i.e. the influence of thedefect remains local. However, it might be the case that distributions of defects maycumulatively exert a long range influence. Secondly, the defect structure itself remainslocal in nature: the interface defect at a point on the grain boundary depends only onslip activity at that point. Redistribution of the defect within the grain boundary byglide or climb (as observed experimentally by Lee et al. [101]) is not yet accounted for.

The bulk-induced microforce Fbk is here obtained from the bulk crystal internal stressderived by Bayley et al. [16]. This ‘full internal stress’ approach accounts for the resolvedcomponents of internal stress resulting from GNDs on all slip systems. Therefore, intra-granular slip system couplings within Fbk are incorporated in a physically viable manner.However, there remains scope to further refine this internal stress definition at (or near)an interface. For example, in the current approach Fbk at a point on the grain A side ofthe interface depends only on the GND distribution in grain A. A more elaborate internalstress might be conceived which would add the influence of the GNDs in grain B to thesegrain A microforces. The level of cross-interface influence from bulk GNDs would dependon the details of the grain boundary structure and its disruptive screening effect. Again,finer-scale modelling methods may be required to resolve this. In the current work,this coupling is omitted, and the interface is therefore considered to perfectly screen theindividual GND stress fields from each side of the grain boundary.

2.4 Numerical investigation

The interface model developed above is now implemented numerically in order to ex-plore the influence of the interface constitutive parameters and the crystal geometry. Tothis end, a model problem consisting of a periodic bicrystal loaded in plane strain isconsidered.

Page 37: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

24 2 Grain boundary interface mechanics in strain gradient crystal plasticity

B

U U

d

h

2d

mA

sA

mB

sB

e1

e2

AB

mB

sB

(a)

grainboundary

µB

µM

µA

µGB

grain A

grain B

(b)

Figure 2.2: (a) Geometry and boundary conditions for simple shear deformation of a periodic bicrys-tal. (b) Definition of slip plane orientations θa and θb, misorientation angle θm and grain boundaryorientation θgb.

2.4.1 Model definition

The geometry and boundary conditions of the periodic bicrystal are shown in Fig. 2.2(a).It consists of alternating columnar grains, designated A and B, separated by grain bound-aries (GBs). Each grain has width d in the e1-direction and is of infinite height in thee2-direction. A repeating unit cell of width 2d and height h is modelled, with periodicboundary conditions applied to the top and bottom boundaries and the left and rightboundaries, as shown in Fig. 2.2(a). Simple shear deformation is simulated by apply-ing a prescribed displacement U = γht on the top boundary. A constant shear rate ofγ = 10−3 s−1 is applied, up to a maximum of 0.1% macroscopic shear.

First grains are considered containing only a single slip system, whose orientation isdefined by the slip plane normal m and slip direction s, as shown in Fig. 2.2(b). Onlyedge dislocations are considered. The slip plane angle θa in grain A (θb in grain B)defines the anti-clockwise rotation of the slip direction relative to the e1-direction. Tospecify the geometric mismatch between the grains, two quantities are defined: (i) themisorientation angle

θm = θa − θb , (2.40)

and (ii) the rotation of the symmetry plane relative to the grain boundary

θgb =1

2(θa + θb) . (2.41)

2.4.2 Finite element implementation

The numerical implementation follows the approach of Evers et al. [56] and Bayley et al.[16], with the addition of the new grain boundary model. The weak forms of the equi-librium equation (2.7a) and the GND balance equations (2.3) are solved in a coupled

Page 38: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.4 Numerical investigation 25

Table 2.1: Material parameters representative of copper.

Parameter Description Magnitude

E Young’s modulus 144 GPaG Shear modulus 54.1 GPaν Poisson’s ratio 0.33b Burgers vector length 0.256 nmγ0 Reference slip rate 0.001 s−1

A11 = A22 Interaction coefficient 0.06A12 = A21 Interaction coefficient 0.612H11 = H22 Immobilisation coefficient 0.2H12 = H21 Immobilisation coefficient 0.4R GND evaluation radius 20 µmm Rate sensitivity exponent 0.1K Disl. segment length constant 26yc Critical annihilation length 1.6 nmρs0 Initial SSD density 7.0 µm−2

manner. Further details are provided in Appendix A. The periodic bicrystal is discre-tised using 24 elements per grain (48 elements in total) in the e1-direction, which hasbeen shown to provide sufficient resolution in a mesh sensitivity analysis. In the e2-direction a single element is adequate due to the periodicity, which has been verified byseparate calculations. Further details on the discretisation are provided in Appendix A.The material parameters used in the current work are similar to those of Bayley et al. [16]and are representative of copper. Individual values used in the numerical implementationare listed in Table 2.1. The initial GND densities are assumed to be zero.

2.4.3 Influence of interface properties

First the influence of the interface constitutive parameters on the shear response of thebicrystal with fixed geometry (grain size and misorientation) is considered. The influ-ence on the crystal deformation of dissipative and energetic interface contributions willbe systematically addressed. The results will be compared with limiting cases havingmicrohard (zero interface normal slip) and microfree (zero interface microforce) grainboundaries.

A bicrystal will be considered with grain size d = 100µm and a symmetric tilt boundarywith θa = −θb = 10◦ (θm = 20◦, θgb = 0◦). The interface reference slip rate q0 and ratesensitivity exponent n are taken to be the same as their bulk counterparts γ0 and m,respectively. Dissipative and energetic hardening at the interface are altered by varyingthe interface normal slip resistance R and grain boundary free energy parameter Ce,respectively. Both R and Ce are taken constant in the present calculations. Note thatthe limit of a microhard and microfree boundary corresponds to Ce = 0 and R → ∞

Page 39: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

26 2 Grain boundary interface mechanics in strain gradient crystal plasticity

microhard GB

microfree GB

R = 0.03 N/mm

R = 0.06 N/mm

R = 0.10 N/mm

R = 0.15 N/mm

(a)

microhard GB

microfree GB

R = 0.03 N/mm

R = 0.06 N/mm

R = 0.10 N/mm

R = 0.15 N/mm

(b)

microhard GB

microfree GB

R = 0.03 N/mm

R = 0.06 N/mm

R = 0.10 N/mm

R = 0.15 N/mm

(c)

microhard GB

microfree GB

R = 0.03 N/mm

R = 0.06 N/mm

R = 0.10 N/mm

R = 0.15 N/mm

(d)

Figure 2.3: The influence of dissipative interface hardening described by the interface normal slip re-sistance R. (a) Stress-strain responses, (b) interface normal slip rates in grain A (right interface inFig. 2.2(a)), (c) bulk-induced microforces in grain A and (d) GND density fields at 0.001 applied macro-scopic shear (the grey lines indicating the location of the grain boundaries).

and R → 0, respectively. In all calculations reported here, the role of the initial grainboundary defect is neglected, i.e. B0 = 0.

Interface normal slip with dissipative hardening

To assess the influence of dissipative hardening at the interface, the grain boundary freeenergy constant Ce is set to zero (such that the energetic interface microforce remainszero) and the interface normal slip resistance R is varied. The results are shown inFig. 2.3. Considering the normalised shear stress-shear strain results shown in Fig. 2.3(a),the responses for various interface normal slip resistances are bounded by the limiting

Page 40: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.4 Numerical investigation 27

Figure 2.4: Evolution of the GND density field with increasing applied macroscopic shear for R = 0.06N/mm.

cases of microhard and microfree interfaces. (Note that the reference shear stress τ0 inFig. 2.3(a) is taken to be the initial bulk crystal slip resistance, as defined Eq. (2.16).)For a given value of R, the shear stress initially follows the microhard grain boundarybehaviour as the deformation is increased. The interface normal slip rate, shown inFig. 2.3(b), remains zero, and the bulk microforce, Fig. 2.3(c), increases as the pile-upat the interface develops. At a critical value of shear, the interface yields and interfacenormal slip begins, Fig. 2.3(b). A larger interface resistanceR causes this transition pointto occur at a larger applied shear. As the deformation progresses, the interface normalslip reduces the development of strain gradients within the crystal, lowering the rate atwhich the GND density adjacent to the interface increases. This in turn reduces the rateof increase of both the driving force for continued slip (i.e. the bulk microforce) andthe strain hardening of the crystal. After further deformation a steady state is reached:the interface normal slip rate reaches a plateau and the rate of strain hardening remainsconstant, both equal to the value for a microfree interface. In this steady state, the rateat which GNDs join the pile-up exactly matches the rate at which they are eliminatedby interface normal slip, resulting in a constant pile-up, Fig. 2.4. This rate is dependentonly on the bulk crystal constitutive behaviour, and is independent of the interfacenormal slip resistance. However the bulk microforce required to achieve this steady-stateinterface normal slip rate increases withR, and likewise the magnitude of the steady-statepile-up, Fig. 2.3(d). These results are, indeed, in qualitative agreement with publishedexperimental observations. The bicrystal experiments of [170, 171] show the pile-up ofdislocations at the interface at small strains leading to locally high GND densities, thatevolve to a more uniform GND distribution at larger strains. This is consistent with thetransition in interface normal slip regimes predicted by the current model, as the bulkmicroforce builds up and overcomes the interface normal slip resistance, Figs. 2.3 and 2.4.A similar transition in interface slip activity has been identified experimentally in the‘pop-in’ phenomena observed during nano-indentation near to a grain boundary [167,212].

Page 41: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

28 2 Grain boundary interface mechanics in strain gradient crystal plasticity

microhard GB

microfree GB

CE = 0

CE = 5·104

CE = 2.5·105

CE = 106

(a)

microhard GB

microfree GB

CE

= 0

CE

= 5·104

CE

= 2.5·105

CE

= 106

(b)

microhard GB

microfree GB

CE

= 0

CE

= 5·104

CE

= 2.5·105

CE

= 106

(c)

microhard GB

microfree GB

CE

= 0

CE

= 5·104

CE

= 2.5·105

CE

= 106

(d)

Figure 2.5: The influence of energetic interface hardening described by the grain boundary free energyconstant Ce. (a) Stress-strain responses, (b) interface normal slip rates in grain A (right interface inFig. 2.2(a)), (c) bulk-induced microforces in grain A and (d) energetic microforces in grain A.

Interface normal slip with energetic hardening

The influence of energetic hardening at the interface is now explored. In the following,the interface normal slip resistance remains constant at an intermediate value R = 0.03N/mm and the grain boundary free energy constant Ce is varied. Recall that Ce deter-mines the magnitude of the energetic microforce once interface normal slip occurs. Theresults are shown in Fig. 2.5. After interface normal slip begins (at an applied shearstrain independent of Ce), the interface normal slip rate increases and the rate of strainhardening decreases, both reaching a steady-state value (Figs. 2.5(a),(b)). However, thesteady-state interface normal slip rate no longer matches the microfree limit, reachinga plateau dependent on Ce. This can be explained as follows. As interface normal slipproceeds, the resultant grain boundary defect, and hence the energetic microforce, grows.

Page 42: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.4 Numerical investigation 29

At steady state, the bulk microforce must therefore continue to rise to balance this effect(Figs. 2.5(c),(d)). The rate at which GNDs join the pile-up (dictated by the bulk consti-tutive behaviour) therefore needs to be greater than the rate at which they are eliminatedby interface normal slip. This is achieved by a reduction in the steady-state interfacenormal slip rate compared to the microfree case. Note that the energetic microforceonly has a significant influence on the overall stress-strain response of the bicrystal whenCe & 104, i.e. when the reference length scale Ceb & 2.5 µm. A physical basis for theparameter Ce, and hence an indication of the importance of the energetic contributionfor realistic hardening parameters, is the subject of ongoing investigation and will not beelaborated further here.

2.4.4 Influence of crystal geometry

In the following, the effects of relative misorientation between adjacent grains (θm), grainboundary orientation (θgb) and grain size d are examined. In all simulations, an interfacewith fixed parameters (Ce = 0 and R = 0.03 N/mm), which will here be referred to as‘intermediate’, is considered together with the limiting cases of microhard and microfreegrain boundaries. For simplicity, here the influence of energetic hardening is omitted,facilitating the interpretation of the results.

Grain misorientation

To investigate the influence of misorientation, symmetric tilt boundaries are considered,i.e. θgb = 0◦, with misorientations in the range θm = 0◦ − 40◦. The results are shown inFig. 2.6. Considering first the microhard grain boundary, increasing the misorientationreduces the resolved shear stress on the slip planes and increases the incompatibilitybetween the slip directions and the applied shear. As a result, a larger misorientationincreases the yield strength and strain hardening, Fig. 2.6(a). The characteristics of themicrohard behaviour are the same as described before, but the rate of accumulation ofGNDs depends on the misorientation, Fig. 2.6(c). In the case of the intermediate strengthgrain boundary, a larger applied shear strain is required to induce grain boundary yieldingfor larger relative misorientations. This can again be attributed to the less favourableorientation of the slip systems with respect to the applied load. After interface yielding,a steady-state interface normal slip rate is reached which is equal to the value of themicrofree interface, see Fig. 2.6(b) (consistent with previous results in the absence ofenergetic hardening at the interface). This steady-state slip rate, dependent on the bulkcrystal slip rates, reduces with increasing misorientation due to the lower resolved shearstresses. The bulk-induced microforce also reaches a steady-state, Fig. 2.6(c), which isindependent of the misorientation. The steady-state GND density at the grain boundary,Fig. 2.6(d), is the same for each misorientation (intermediate case), as it depends on Ronly. The current model predicts an increased resistance to interface normal slip as themisorientation is increased. This is qualitatively consistent with the bicrystal experimentsof Zaefferer et al. [215], that show similar trends for increasing misorientation.

Page 43: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

30 2 Grain boundary interface mechanics in strain gradient crystal plasticity

µM = 0°

µM = 20°

µM = 40°

microhard GB

microfree GB

intermediate GB

(a)

µM

= 0°

µM

= 20°

µM

= 40°

microhard GB

microfree GB

intermediate GB

(b)

µM

= 0°

µM

= 20°

µM

= 40°

microhard GB

microfree GB

intermediate GB

(c)

µM = 0°

µM = 20°

µM = 40°

microhard GB

microfree GB

intermediate GB

(d)

Figure 2.6: The influence of grain misorientation. (a) Stress-strain responses, (b) interface normal sliprates in grain A (right interface in Fig. 2.2(a)), (c) bulk-induced microforces in grain A and (d) GNDdensity fields at 0.001 applied macroscopic shear (the grey lines indicating the location of the grainboundaries).

Grain boundary orientation

To investigate the effect of an asymmetric tilt boundary, the relative misorientation ofthe grains is now fixed at θm = 20◦ with their orientation relative to the grain boundaryvaried in the range θgb = 0◦−20◦ (Fig. 2.2(b)). The results are shown in Fig. 2.7. Due tothe asymmetric configuration, the two grains now reach a steady state at different levelsof applied shear strain. This is due to differences in their orientation with respect to thedirection of applied shear. As grain B becomes more favourably oriented than A withincreasing θgb, the pile-up and associated bulk-induced microforces in grain B increasefaster, Figs. 2.7(e),(f), and therefore overcome the interface resistance faster. The twograins also reach different steady-state conditions, with the preferential orientation of

Page 44: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.4 Numerical investigation 31

grain B, inducing a larger resolved shear stress, leading to greater interface normal sliprates, Figs. 2.7(c),(d).

Grain size d

The influence of the grain size is now explored for a symmetric tilt boundary (θgb = 0◦)with fixed relative misorientation θm = 20◦. Results are shown in Fig. 2.8 for differentvalues of the grain size d. It is noted that the stress-strain responses (Fig. 2.8(a)) arethe same for all grain sizes in the case of a microfree grain boundary. This is due tothe absence of slip gradients and the associated GNDs. In the case of a microhard grainboundary, larger slip gradients, higher GND densities (Fig. 2.8(b)) and higher ratesof strain hardening result as the grain size is reduced. For the intermediate strengthgrain boundary, yielding of the interface and the development of interface normal slip,Fig. 2.8(c), occurs at a lower applied shear strain but a higher applied shear stress as thegrain size reduces. This is due to the larger bulk-induced microforces associated withthe higher GND densities, Fig. 2.8(d). After interface yielding, the steady-state interfacenormal slip rate and strain hardening behaviour are identical for the different grain sizes.Consequently, the grain size dependence of the steady-state rates of strain hardening arediminished.

2.4.5 Influence of multiple slip systems

Finally, the effect of multiple slip systems is addressed. Two slip systems are consideredin each crystal. An important difference with respect to the previous simulations is thepresence of intra-granular slip system coupling that naturally influences both the bulkcrystal deformation and the interface normal slip via the bulk-induced microforce. Theslip system orientations, shown in Fig. 2.9, are: θα1 = 50◦, θα2 = −10◦, θβ1 = 30◦, θβ2 =−30◦. The material parameters used, including the slip system interaction coefficientsfor the bulk crystal constitutive model [16], can be found in Table 2.1. The grain sized = 100 µm and the interface parameters for the intermediate strength grain boundaryare R = 0.3 N/mm and Ce = 0. In this calculation, the interface normal slip resistanceR is increased compared to previous cases in order to clarify the role of multiple slipsystems.

The results for the different grain boundary conditions are shown in Fig. 2.10 for appliedmacroscopic shear strain up to 0.4%. Considering the normalised shear stress-shear strainresults in Fig. 2.10(a), kinks in the curve indicate the distinct yielding of different slipsystems in each grain, with preferentially oriented slip systems yielding first. For themicrofree and intermediate grain boundaries, the development of interface normal slip,Figs. 2.10(c),(d), also changes the rate of strain hardening. Again, interface normal slipbegins on preferentially oriented slip systems first. As observed previously, the grainboundary resistance influences the applied strain at which these slip events occur. Notethat the evolution of slip activity on the various slip systems shown in Figs. 2.10(c)-(f)is more complex than in previous cases due to the additional intra-granular slip systeminteractions in the bulk and at the interface.

Page 45: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

32 2 Grain boundary interface mechanics in strain gradient crystal plasticity

µGB = 0°

µGB = 10°

µGB = 20°

microhard GB

microfree GB

intermediate GB

(a)

µGB = 0°

µGB = 10°

µGB = 20°

microhard GB

microfree GB

intermediate GB

(b)

µGB

= 0°

µGB

= 10°

µGB

= 20°

microhard GB

microfree GB

intermediate GB

(c)

µGB

= 0°

µGB

= 10°

µGB

= 20°

microhard GB

microfree GB

intermediate GB

(d)

µGB = 0°

µGB = 10°

µGB = 20°

microhard GB

microfree GB

intermediate GB

(e)

µGB = 0°

µGB = 10°

µGB = 20°

microhard GB

microfree GB

intermediate GB

(f)

Figure 2.7: The influence of grain boundary orientation. (a) Stress-strain responses, (b) GND densityfields at 0.001 applied macroscopic shear, (c) interface normal slip rates in grain A (right interface inFig. 2.2(a)), (c) interface normal slip rates in grain B (right interface in Fig. 2.2(a)), (e) bulk-inducedmicroforces in grain A and (f) grain B. The grey lines in (b) represent the GBs.

Page 46: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.4 Numerical investigation 33

d ¹= 50 m

d ¹= 100 m

d ¹= 200 m

microhard GB

microfree GB

intermediate GB

(a)

d ¹= 50 m

d ¹= 100 m

d ¹= 200 m

microhard GB

microfree GB

intermediate GB

(b)

d ¹= 50 m

d ¹= 100 m

d ¹= 200 m

microhard GB

microfree GB

intermediate GB

(c)

d ¹= 50 m

d ¹= 100 m

d ¹= 200 m

microhard GB

microfree GB

intermediate GB

(d)

Figure 2.8: The influence of grain size. (a) Stress-strain responses, (b) GND density fields at 0.001applied macroscopic shear, (c) interface normal slip rates in grain A (right interface in Fig. 2.2(a)), and(d) bulk-induced microforces in grain A. The grey lines in (b) represent the GBs.

B

m®1

s®1

e1

e2

AB

m¯1

s¯1

m¯2

s¯2

m¯1

s¯1

m¯2

s¯2

m®2

s®2

Figure 2.9: Slip system orientations used for investigating the effect of multiple slip systems.

Page 47: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

34 2 Grain boundary interface mechanics in strain gradient crystal plasticity

microhard GB

microfree GB

intermediate GB

(a)

1 slip systemst

2 slip systemnd

microhard GB

microfree GB

intermediate GB

(b)

1 slip systemst

2 slip systemnd

microhard GB

microfree GB

intermediate GB

(c)

1 slip systemst

2 slip systemnd

microhard GB

microfree GB

intermediate GB

(d)

1 slip systemst

2 slip systemnd

microhard GB

microfree GB

intermediate GB

(e)

1 slip systemst

2 slip systemnd

microhard GB

microfree GB

intermediate GB

(f)

Figure 2.10: The response of multiple slip systems. (a) Stress-strain responses, (b) GND density fields at0.004 applied macroscopic shear, (c) interface normal slip rates in grain A (right interface in Fig. 2.2(a)),(d) interface normal slip rates in grain B (right interface in Fig. 2.2(a)), (e) bulk-induced microforces ingrain A and (f) grain B. The grey lines in (b) represent the GBs.

Page 48: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

2.5 Conclusions 35

2.5 Conclusions

Grain boundaries impede or constrain the motion of lattice dislocations, influencing theplastic deformation of polycrystals. They present an obstacle to slip whose strength isinfluenced by crystallographic orientations, the initial grain boundary structure and theaccumulation of a residual defect at the grain boundary due to dislocation interactions.In this chapter some key features of grain boundary behaviour at the continuum scale arecaptured, by presenting a constitutive model for the higher-order boundary conditionsat internal interfaces required within a strain gradient crystal plasticity theory. Theinnovative features in this chapter are:

• A thermodynamically consistent grain boundary interface model is formulatedwithin the microstructurally motivated strain gradient crystal plasticity frameworkdeveloped by Evers et al. [56] and Bayley et al. [16]. Plastic slip at the grainboundary is governed by the competition between three interface microforces. Abulk-induced microforce, derived from the bulk crystal GND densities, providesa driving force for interface normal slip which depends on the magnitude of theinterface pile-up. A dissipative microforce provides a resistance to slip at the inter-face. An energetic microforce acts as an interface ‘back stress’, tending to opposeaccumulation of the interface residual defect (i.e. the grain boundary net Burgersvector).

• By making a connection with the ‘full internal stress’ formulation of Bayley et al.[16], the bulk-induced microforces capture intra-granular slip system coupling in aphysically realistic manner.

• The energetic interface microforce provides inter-granular slip system coupling, ac-counting for both the magnitude and direction of contributions to the residualdefect from all slip systems in the two neighbouring grains. The geometric condi-tions controlling the interaction of a particular slip system with the grain boundaryare accounted for.

• The grain boundary model has been implemented numerically in order to studythe influence of a grain boundary on the shear deformation of a bicrystal.

The initiation of interface normal slip results in a departure from the ‘microhard’ bound-ary case, with a reduction in the rate of strain hardening. Higher interface normal slipresistance, i.e. a larger dissipative interface microforce, requires a larger applied shearstrain to induce interface yield. Post yield, interface normal slip reaches a steady-statelevel, achieving a balance between (i) the bulk crystal GND accumulation (dependenton the bulk crystal constitutive parameters) which drives interface normal slip via thebulk-induced microforce, (ii) the interface normal slip resistance and (iii) the developingenergetic interface microforce, linked to the accumulated interface defect. Increasing themisorientation of the adjacent grains delays the onset of interface normal slip and reducessteady-state interface normal slip rates, leading to higher rates of strain hardening. For

Page 49: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

36 2 Grain boundary interface mechanics in strain gradient crystal plasticity

asymmetric tilt boundaries, the grain preferentially oriented with the applied load ex-periences earlier interface yield and higher rates of interface normal slip. As the grainsize is reduced, larger strain gradients and higher GND densities result, which initiateinterface normal slip at a lower applied shear strain. For a bicrystal with two slip systemsper grain, the development of slip in the bulk and at the interface again depend on theorientation of the slip system with respect to the applied load. However, slip activityon each system now varies in a more complex way with increasing applied shear straindue to the intra-granular slip system coupling present in both the bulk and interfaceconstitutive models.

Qualitative agreement between the model prediction and the experimental observationsreported in the literature has been shown. Quantitative calibration and validation of themodel is presently hindered by a lack of suitable experimental data.

Possible directions for future development of the model include more complex microstruc-turally dependent interface microforces. The interactions between GNDs influence boththe bulk internal stress and the bulk-induced interface microforces. These formulationscan be refined to account for the presence of the grain boundary and the fact that stressfields of individual GNDs might interact across it. The interface normal slip resistancecould incorporate dependencies on local microstructure, for example the misorientationof the grains, the interface net defect or the nucleation strength of dislocation sourcesclose to the grain boundary. The energetic interface microforces could account for thenon-local influences between grain boundary dislocations, and the redistribution of theinterface defect within the grain boundary. Finer-scale modelling methods, such as MDor DDD, are required to support these elaborations.

Page 50: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Chapter 3

A multiscale model of grainboundary structure and energy1

Abstract

Grain boundaries play an important role in the mechanical and physical properties ofpolycrystalline metals. While continuum macro-scale simulations are appropriate formodelling grain boundaries in coarse-grained materials, only atomistic simulations pro-vide access to the details of the grain boundary structure and energy. Hence, a multiscaledescription is required to capture these grain boundary details. The research objectiveherein is to consolidate various approaches for characterising grain boundaries in an ef-fort to develop a multiscale model of the initial grain boundary structure and energy.The technical approach is detailed using various 〈100〉, 〈110〉 and 〈111〉 symmetric tiltgrain boundaries in copper and aluminium. Characteristic features are: (i) grain bound-ary energies and boundary period vectors as obtained from atomistic simulations, (ii)structural unit and dislocation descriptions of the grain boundary structure and (iii) theFrank-Bilby equation to determine the dislocation content. The proposed approach de-fines an intrinsic net defect density scalar that is used to accurately compute the grainboundary energy for these grain boundary systems. The significance of this work is thatthe developed atomistic-to-continuum approach is suitable for realistically inserting theinitial grain boundary structure and energy into continuum level frameworks.

1Reproduced from [188]: P.R.M. van Beers, V.G. Kouznetsova, M.G.D. Geers, M.A. Tschopp, andD.L. McDowell. A multiscale model of grain boundary structure and energy: From atomistics to acontinuum description. Acta Materialia, 82:513-529, 2015.

37

Page 51: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

38 3 A multiscale model of grain boundary structure and energy

3.1 Introduction

Grain boundaries (GB) have a major influence on the material properties of polycrys-talline metals [80]. The grain boundary behaviour contributes significantly to the macro-scopic response of metallic materials with large grain boundary-to-volume ratios. Ex-amples of these include fine-grained materials, thin films and micro-devices. In pastresearch, it has become clear that not only the original static grain boundary structure,but also the evolution of this structure, due to the interaction of grain boundaries withother lattice defects, plays an important role in the deformation behaviour [186]. Experi-mentally, it has been observed that many different interactions exist between dislocationsand grain boundaries during plastic deformation, see [60, 145, 175] and references therein.

One of the current challenges in modelling plasticity of polycrystalline metals is to accountmore accurately for the role of grain boundaries [116, 117]. In terms of continuumscale modelling, for example in classical crystal plasticity approaches, grain boundaryphenomena are often not accounted for and grain boundaries are only present as planeswhere the crystallographic orientation changes. Gradient enhanced crystal plasticityapproaches mostly incorporate only the limiting situations of either impenetrable orcompletely transparent grain boundaries through higher-order boundary conditions [16,56, 76, 95, 115]. More recently, several attempts have been made to elaborate theseframeworks to include grain boundary interface mechanics [21, 51, 77, 92, 110, 210].However, among the different choices that have to be made, the grain boundary energyoften follows a simple ad hoc expression that does not depend on the grain boundarystructure. Only a limited number of the above approaches (can) address the underlyinginitial and evolving grain boundary structure [66, 178, 187].

At a finer (discrete) scale, different grain boundary modelling approaches exist in litera-ture that differ in their ability to describe the initial structure, defect content and energy.These include: (i) dislocation and disclination models [4, 104, 150, 151, 161, 162], (ii) geo-metric theories, such as the coincidence site lattice (CSL), Frank-Bilby and O-lattice the-ory [20, 83, 148, 213], (iii) atomistic simulations, which led to the identification of struc-tural units and their relation with dislocation and disclination models [8, 68, 130, 176].Experiments can be used to validate these models, such as (a) conventional (brightand dark field) and high resolution transmission electron microscopy (HRTEM) for thegrain boundary structure at small length scales [39, 50, 97, 119, 120, 122], and (b) mea-surements of dihedral angles of surface grooves related to the grain boundary energy[70, 79, 127, 132, 133, 154].

Dislocation models are typically used for low-angle boundaries. For large misorientations(& 15◦) dislocation cores start to overlap and the model is only applicable in a formal way.The geometric theories elucidate many phenomena from a crystallographic perspective,but are unable to incorporate boundary core relaxation processes or describing boundaryenergy. Sutton and Balluffi [174] compared five geometric criteria for low interfacialenergies. They concluded that no general and useful criterion for low energy can ensuefrom a simple geometric framework. Any understanding of the evolution of interfacialenergy must account for the atomic structure and the details of the interatomic bonding at

Page 52: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.1 Introduction 39

the interface. Indeed, at present, atomistic simulations provide most detailed informationconcerning grain boundary structure and energy, but also these calculations face theirown difficulties, such as the choice of interatomic potential.

At the atomistic level, a considerable amount of investigations into the structure andenergy of symmetric and asymmetric grain boundaries has been performed [153, 179–181, 197, 198, 206]. However, at this point, no continuum net defect description of thegrain boundary structure exists that properly describes the boundary energy. Recently,Bulatov et al. [24] proposed a universal grain boundary energy function for face-centeredcubic (FCC) metals motivated by observations from extensive atomistic simulations todetermine grain boundary energies in the case of four metallic elements [84, 131]. It isassumed that the grain boundary energy is a continuous function of the five macroscopicdegrees of freedom of the grain boundary. A hierarchical interpolation is applied inwhich the full 5D energy function is built on a scaffolding consisting of lower-dimensionalsubsets of the 5D space. This approach, however, does not provide a correspondingnet defect description of the grain boundary structure suitable for implementation inadvanced continuum crystal plasticity models that can incorporate grain boundary in-terfacial plasticity, such as the model developed in Chapter 2.

In this chapter, the aim is to bridge the gap between the materials science and mechan-ics views on modelling an initial grain boundary structure and energy. The main novelcontribution of this chapter is the development of a multiscale approach leading to acontinuum representation of the initial grain boundary structure defect content and en-ergy based on the output of atomistic simulations through the consolidation of differentconcepts related to the grain boundary structure. An intrinsic net defect is attributed toa boundary period of a certain length that represents the initial grain boundary struc-ture. It is found that the corresponding net defect density can be related to the initialgrain boundary energy remarkably well. This is shown for the complete misorientationrange, i.e. both low- and high-angle boundaries, for three symmetric tilt grain bound-ary (STGB) systems in aluminium and copper. The developed multiscale approach canbe hierarchically incorporated in a grain boundary interface continuum model, therebyrealistically capturing the initial boundary structure and its energetics.

This work focuses on a special case of symmetric tilt boundaries. Although in some FCCmetals, it is experimentally found that asymmetric tilt grain boundaries and more generalmixed tilt-twist boundaries are preferred over symmetric tilt boundaries [149], symmetrictilt grain boundaries represent an idealized case to test the developed multiscale formal-ism, because of the extensive amount of experimental and computational data availableon the structure and energy of symmetric tilt grain boundaries, e.g. [120, 122, 153, 198].

The chapter is organized as follows. Different grain boundary structure concepts thatare used in the atomistic-to-continuum approach are summarized in Section 3.2. In Sec-tion 3.3, the methodology and the results of the atomistic simulations, used as the input,are briefly explained. Section 3.4 then describes the details of the multiscale approachfor deriving the corresponding continuum representations of the boundary structure andenergy. A summary of the procedure and an illustrative example are provided. In Sec-tion 3.5, the results for three symmetric tilt boundary systems in two metallic elements

Page 53: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

40 3 A multiscale model of grain boundary structure and energy

are presented, followed by a discussion in Section 3.6. Finally, conclusions are made inSection 3.7.

3.2 Fundamental GB structure concepts

Several fundamental concepts related to the structure of grain boundaries used in thedevelopment of the atomistic-to-continuum approach will be summarized here: (i) thedegrees of freedom that define the grain boundary geometry, (ii) lattice descriptions andgrain boundary dislocations, (iii) structural units and (iv) the Frank-Bilby equation.

3.2.1 Grain boundary degrees of freedom

The structure of a grain boundary depends on five macroscopic degrees of freedom(DOFs) [208]: grain misorientation (3 DOFs: rotation axis and angle) and boundaryplane inclination (2 DOFs: plane normal). For the grain boundaries studied in thiswork, the rotation axis is contained within the grain boundary plane (i.e. tilt grainboundaries) and the misorientation of the adjacent crystal lattices is symmetric aboutthe grain boundary plane (i.e. STGBs). An alternative representation of the five macro-scopic DOFs relies on the interface plane normals in both crystals (4 DOFs) and a twistrotation of one crystal with respect to the other about the interface normal (1 DOF). Inaddition, three microscopic DOFs involve translations parallel (2 DOFs) and perpendic-ular (1 DOF) to the boundary. These microscopic DOFs provide an important relaxationmechanism for reducing the grain boundary energy.

3.2.2 Lattice descriptions and grain boundary dislocations

In analysing interfacial structures of grain boundaries, useful descriptions are provided bythe coincident site lattice (CSL) and displacement shift complete (DSC) lattice construc-tions [20, 73]. For certain discrete misorientations between two (virtually) interpenetrat-ing lattices a fraction of the lattice sites will coincide, see Fig. 3.1(a). The coincidentatom positions themselves form a lattice, called the coincident site lattice (CSL), seeFig. 3.1(b). The CSL is characterised by the parameter Σ, which is the inverse densityof coincident sites, equivalently expressed as the ratio of the volume of the unit cell ofthe CSL to that of the crystal lattice. The actual grain boundary plane runs through theCSL (Fig. 3.1(c)). The displacement shift complete (DSC) lattice is the coarsest latticethat contains all of the sites of both crystals in the coincidence orientation as sublattices,see Fig. 3.1(b). It is noted that all translation vectors of the CSL and the crystal latticesare also vectors of the DSC lattice, but the elementary vectors of the DSC lattice aremuch smaller.

Different types of dislocations exist that are part of the grain boundary, two of whichare (i) primary and (ii) secondary grain boundary dislocations [175]. A dislocation is

Page 54: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.2 Fundamental GB structure concepts 41

θ

(a) (b) (c)

Figure 3.1: Lattice descriptions in the analysis of grain boundaries. (a) Two interpenetrating latticesrotated 36.87◦ with respect to each other about [001]. (b) The Σ5 CSL (thick lines) and the DSC lattice(thin lines). (c) A possible (3 1 0) boundary generated from the CSL. After [58].

characterised by its Burgers vector, which can be determined by a Burgers circuit con-struction [82]. A primary GB dislocation has a Burgers vector that is a translation vectorof one of the crystal lattices. A secondary GB dislocation has a Burgers vector that is atranslation vector of a DSC lattice, excluding those vectors that are also crystal latticetranslation vectors. Whereas primary interface dislocations conserve the periodicity ofthe crystal lattice by accommodating small deviations from the perfect single crystal(Σ1), secondary interface dislocations conserve the structure of the coincidence orien-tation boundary by the accommodation of small deviations from the nearest CSL [85].Furthermore, secondary GB dislocations can be considered as perturbations in a uniformdistribution of primary GB dislocations.

3.2.3 Structural units

The grain boundary structure can be described in terms of the structural unit model(SUM) [18, 19, 176]. For any misorientation axis, short-period boundaries exist thatconsist of a contiguous sequence of only one type of structural element. These bound-aries are called ‘favoured’. Non-favoured boundaries in the misorientation range betweentwo adjacent favoured boundaries are composed of a mixture of these two structural el-ements. The proportion and sequence of the structural units are derived from a rule ofmixture. The favoured boundary may be regarded as being composed of a uniformly dis-tributed array of primary (lattice) dislocations. Secondary grain boundary dislocationsare periodic perturbations of this distribution and are localized at the minority struc-tural units, which are the units smallest in number. A hierarchy of structural unit anddislocation descriptions exists, where each description corresponds to a different choiceof reference boundaries (termed ‘delimiting’, i.e. favoured or multiple unit referencestructures (MURS)) that provide the basic units [11, 12, 129]. The selection of delimit-ing boundaries may vary with the physical property under study [175] or the ability tovisualize dislocations using TEM diffraction-contrast [11, 97]. Although the predictive

Page 55: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

42 3 A multiscale model of grain boundary structure and energy

capability of the structural unit model is limited for grain boundaries with relativelyhigh-index rotation axes, i.e. 〈uvw〉 with some indices larger than one, and/or for mixedtilt and twist boundaries [173], the model is particularly effective in the case of pure tiltand twist boundaries with low-index rotation axes, such as 〈100〉, 〈110〉 and 〈111〉. Thestructural unit model is not found to be useful for describing the grain boundary struc-tures in low stacking fault energy (SFE) FCC metals, due to the extensive local atomicrelaxations and delocalized or dissociated grain boundary dislocations [153]. Nonetheless,even when the structural unit model does not fully apply, boundaries are still composedof structural units and boundaries that are comprised of only one type of structural unitdo exist.

3.2.4 Frank-Bilby equation

The Frank-Bilby equation is a widely used approach to determine the dislocation contentof a general boundary, including both grain and interphase boundaries. This equation,based on a Burgers circuit construction, relates the net Burgers vector Bp of all interfacialdislocations crossing any vector p in the interface and the lattice deformations SA andSB (each including rotation and elastic deformation) converting the reference lattice,from which the bicrystal is created, into real lattices A and B [33, 176]

Bp =(S−1A − S−1

B

)· p (3.1)

The net Burgers vector Bp is defined as the closure failure in the perfect reference lattice.This reference lattice, used to construct the bicrystal, could be chosen to be one of thetwo crystal lattices A or B. When the two lattices can be related by a pure rotation,which is the case for grain boundaries, the same lattice at a median orientation, calledthe median lattice, could be taken as the reference lattice. The two crystal lattices arethen obtained from the median lattice by equal and opposite rotations. The Frank-Bilbyequation can be employed to determine both primary and secondary GB dislocations bycarrying out a Burgers circuit construction using the single crystal and DSC referencelattices, respectively [12, 60].

The net lack of closure resulting from the Burgers circuit construction for high-angle grainboundaries has a significant contribution from dissociated partial dislocations within theinterface, in accordance with the elevated excess energy. From a continuum perspec-tive, one may kinematically represent the closure fault of a Burgers circuit by using acombination of translational dislocations and rotational lattice defects known as par-tial disclination dipoles, which are characterised by a Frank vector. Some continuousfield theories have employed this combined dislocation-disclination representation withassociated interface energy functions, cf. [35, 36] and [65, 66].

Page 56: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.3 Atomistic simulations of GB structure and energy 43

3.3 Atomistic simulations of GB structure and energy

In this section, based on the atomistic simulations in [179, 180], the methodology tocompute the structure and energy of grain boundaries using atomistic simulations willbe briefly described, followed by the simulation results. Emphasis is given to the requiredelements from the atomistic output to be used in the multiscale approach developed inthis chapter.

3.3.1 Simulation methodology

The equilibrium structures and energies at 0 K of three symmetric tilt grain boundarysystems with the 〈100〉, 〈110〉 and 〈111〉 rotation axes in Al and Cu have been calcu-lated using a bicrystal computational cell with 3D periodic boundary conditions. A largenumber of initial configurations with different in-plane rigid body translations and anatom deletion criterion were used to access the minimum energy structures. A nonlinearconjugate gradient algorithm was used for energy minimization. The embedded-atommethod interatomic potentials for Al and Cu [123, 124] were used that reproduce thestacking fault energies consistent with available experimental data and ab initio calcula-tions. Further details of the simulation methodology and parameters used can be foundin [179].

3.3.2 Simulation results

The results of the atomistic simulations are shown in Fig. 3.2. The grain boundaryenergies versus the misorientation angle for the three rotation axes in Al and Cu aregiven in Fig. 3.2(a). In particular, the grain boundary energy varies strongly with themisorientation angle for the 〈110〉 rotation axis. The two deep cusps are the Σ11 (1 1 3)θ = 50.48◦ symmetric tilt boundary and the Σ3 (1 1 1) θ = 109.47◦ coherent twinboundary in the 〈110〉 STGB system. Examples of grain boundary structures associatedwith minimum energies in Cu are shown in Fig. 3.2(b)-(d) for each rotation axis. Thedifferent coloured circles represent atomic positions on consecutive planes perpendicularto the rotation axis. In each grain boundary, the structural units characterising the grainboundary structure are indicated. These structural units, consistent with the structuralunits of nearby favoured boundaries, are kite-shaped. Note, however, that the definitionof structural units is not unique, i.e. different shapes of structural units can be defined aslong as these are used consistently. In general, multiple structural units are present withinone boundary period. The grain boundary structure is composed of repeating boundaryperiods, which are characterised by the boundary period vector p perpendicular to therotation axis, see Fig. 3.2(d). In contrast to the other two rotation axes, the 〈110〉 axisshows grain boundary dissociation by the emission of stacking faults reducing the grainboundary energy, see Fig. 3.2(c). Finally, as can be seen in Fig. 3.2(b), a pair of atomicplanes terminates at a minority unit that forms the core of an edge dislocation.

Page 57: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

44 3 A multiscale model of grain boundary structure and energy

0 20 40 60 80 100 120 140 160 1800

100

200

300

400

500

600

700

800

900

1000

Misorientation angle, θ (degrees)

Gra

in b

ou

nd

ary

en

erg

y (

mJ/m

2)

(a) (b)

(c)

p

(d)

Figure 3.2: Results of the atomistic simulations. (a) Symmetric tilt GB energy versus misorientationangle for the 〈100〉 (blue), 〈110〉 (red) and 〈111〉 (green) rotation axes in Al (open circles) and Cu (closedcircles). (b)-(d) Examples of grain boundary structures in Cu for the 〈100〉, 〈110〉 and 〈111〉 rotationaxes, respectively. The black, white and red circles represent atomic positions on consecutive {200},{220} and {333} planes in (b)-(d), respectively. Structural units in each grain boundary are indicated.Two atomic planes (dashed) terminating at a minority unit corresponding to an edge dislocation aredepicted in (b). The period vector of the grain boundary p is illustrated in (d).

From these atomistic simulations the grain boundary energy data required in the atomistic-to-continuum multiscale approach has been extracted. The grain boundary structureplots can also be used to analyse the dislocation content and the boundary period vec-tors.

Page 58: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.4 Multiscale approach to GB structure and energy 45

3.4 Multiscale approach to GB structure and energy

In order to incorporate the initial grain boundary structure and energy, obtained fromthe simulations at the discrete atomistic scale, in continuum level frameworks, continuumequivalent measures should be defined that realistically reflect the energetics and thestructure of grain boundaries. After explaining the details of the proposed atomistic-to-continuum approach, a summary of the procedure and an illustrative example arepresented next.

3.4.1 Details of the approach

The key idea behind the proposed multiscale approach is to attribute an intrinsic netdefect, representing the initial grain boundary structure, to a boundary period for eachboundary in the whole misorientation range around a certain rotation axis. A corre-sponding intrinsic net defect density is then related to the grain boundary energy. In thedevelopment of this multiscale approach, the following information is consolidated:

1. The grain boundary energies and boundary period vectors obtained from the atom-istic simulations.

2. The structural unit and dislocation descriptions of the grain boundary structure.

3. The Frank-Bilby equation to determine the effective net dislocation content.

Consider a grain boundary period of a certain length. An intrinsic net defect Bp withinthis period is defined as the sum of the Burgers vectors of the primary and secondarydislocations in that period. In general, the intrinsic net defect comprises both primaryand secondary dislocations, but only the contribution of one type in a certain misorien-tation range is considered in this work. The term ‘intrinsic’ in the definition of Bp is toindicate that it is the net defect representing the initial boundary structure. The corre-sponding dislocations form a part of the equilibrium structure of the boundary [13, 81].In contrast, ‘extrinsic’ dislocations, which are not part of the equilibrium structure, arelattice dislocations that interact with the grain boundary during plastic deformation.Consideration of the extrinsic dislocations is outside the scope of this chapter.

The Frank-Bilby equation, Eq. (3.1), can be used to determine the primary or secondarydislocation content. In the case of (i) homophase interfaces, i.e. grain boundaries wherethe two adjoining lattices are related by a rotation only, and (ii) taking the median(perfect crystal) lattice as the reference lattice (for primary dislocations), this equationreduces to Frank’s formula [82, 175]

Bp = 2 sin

(∆θ

2

)(p× c) with |Bp| = 2|p| sin

(∆θ

2

)(3.2)

Here p is the boundary period vector, c is the unit vector along the rotation axis and∆θ is the misorientation angle between the two adjacent grains with respect to the

Page 59: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

46 3 A multiscale model of grain boundary structure and energy

SF

bicrystal

S F

median reference lattice

p

c

Bp

Δθ2

Δθ2

p

Figure 3.3: Illustration of Frank’s formula in the case of a symmetric tilt grain boundary. Starting witha closed loop of length |p| around the boundary in the bicrystal (left) and applying equal and oppositerotations of ∆θ/2 around unit vector c back to the median reference lattice, the resultant Burgers vectorBp is defined by the closure fault (right).

perfect reference lattice. Bp is the resultant Burgers vector of the primary dislocationspierced by the boundary period vector p. An illustration of Frank’s formula in the caseof a symmetric tilt grain boundary is shown in Fig. 3.3. For the determination of thesecondary dislocation content, the median DSC lattice is used as the reference latticeand ∆θ in Eq. (3.2) then reduces to the misorientation from the nearby CSL orientation.Furthermore, the ‘finish-start/right-handed’ (FS/RH) convention is used.

The intrinsic net defect density |Bp|/|p| is determined and related to the grain boundaryenergy φgb obtained from the atomistic simulations. It is found that for each rotationaxis considered here, a logarithmic relationship matches

φgb = M|Bp||p| −N

|Bp||p| ln

( |Bp||p|

)(3.3)

where M and N are the two continuum parameters. A standard least squares fittingprocedure will be applied. This logarithmic relationship shows resemblance with theRead-Shockley equation [150, 151] and, as will be discussed later in Section 3.6, con-stitutes a generalization to the whole misorientation range, including high-angle grainboundaries. For the purpose of subsequent analyses, Eq. (3.3) can be split into twocontributions

φgb = φcore+φstrain with φcore = M|Bp||p| and φstrain = −N |Bp|

|p| ln

( |Bp||p|

)(3.4)

This split resembles the contributions of the core and strain energy to the total grainboundary energy [150, 205, 207]. This will be discussed further in Section 3.6.

As will be shown in the following sections, for the 〈100〉 and 〈111〉 rotation axes primarydislocation descriptions are used in the full misorientation ranges. In the case of the〈110〉 rotation axis, the primary and secondary dislocation descriptions are employed indifferent parts of the misorientation range. It may be argued that the primary dislocationdescription has little physical meaning in the case of high-angle grain boundaries, becausedislocation cores overlap and dislocations can no longer be resolved discretely. However,

Page 60: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.4 Multiscale approach to GB structure and energy 47

in this case the boundary period can be considered to be a continuous distribution ofinfinitesimal dislocations [33], the net effect of which leads to the same intrinsic result.Under this assumption, the use of the primary dislocation description for high-angle grainboundaries is justified.

3.4.2 Summary of the procedure and illustrative example

To summarize, the atomistic-to-continuum approach follows a three-step procedure:

1. For each of the grain boundaries in the complete misorientation range around agiven rotation axis, determine the boundary period vector p from the atomisticsimulations and employ Frank’s formula, Eq. (3.2), to determine the intrinsic netdefect Bp. Calculate the corresponding intrinsic net defect density |Bp|/|p|, to-gether with Bp providing the continuum measures of the grain boundary defectstructure.

2. Determine the continuum parametersM andN in the logarithmic relation, Eq. (3.3),between the intrinsic net defect density |Bp|/|p| and the grain boundary energyφgb as obtained from the atomistic simulations.

3. Combine the previous two steps to obtain a continuum scale relation between thegrain boundary energy φgb and the misorientation angle θ. If needed, the boundaryenergy can be split into representative core and strain energy terms, Eq. (3.4).

The output of this multiscale approach is a continuum representation of the initial grainboundary structure and energy, i.e. Bp and φgb (θ), that can be used in continuum levelframeworks.

As an example, this procedure is demonstrated for the grain boundaries with the [0 0 1]rotation axis in Al. For this rotation axis, the misorientation angle range is θ ∈ [0◦, 90◦],in which the end points, θ = 0◦ and θ = 90◦, are perfect crystals.

Step 1. From the grain boundaries within the misorientation domain, the Σ41 (5 4 0)θ = 77.32◦ boundary is selected to be further elaborated as a particular example, seeFig. 3.4. The orientations of both crystals are shown together with the boundary periodvector p and the atomic planes terminating at the minority structural units, which formthe dislocation cores. In this example, the (perfect) Σ1 (1 1 0) θ = 90◦ crystal boundaryis used as the reference. Hence, the misorientation angle ∆θ = 77.32◦ − 90◦ = −12.68◦.

It is pointed out that symmetric tilt grain boundaries can be described with an arrayconsisting of one type of edge dislocations each having Burgers vector b [86]. Hence, theintrinsic net defect Bp = nb, where n denotes the number of edge dislocations crossedby the boundary period vector p. These Burgers vectors b are translation vectors of thereference lattice. Furthermore, for symmetric tilt boundaries, the Burgers vectors b areparallel to the boundary normal unit vector n [86]. This corresponds to b = a

2 [1 1 0]in this example, with a the lattice parameter. In order to determine the number of

Page 61: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

48 3 A multiscale model of grain boundary structure and energy

p

[110]A

12

[110]B

12

[110]A

12

[110]B

12

[100]A

[100]B

[010]A

[010]B

n

c !nc

Figure 3.4: A grain boundary in Al that is used to illustrate the first step in the atomistic-to-continuumapproach. The crystal orientations, the boundary unit normal n, the unit vector c along the rotationaxis, the boundary period vector p and pairs of atomic planes terminating at the minority structuralunits are shown.

dislocations pierced, the boundary period vector p is expressed as p = (c× n)|p|. ThenFrank’s formula (Eq. (3.2)) yields

nb = 2 sin

(∆θ

2

)[(c× n)× c] |p| = 2 sin

(∆θ

2

)|p|n (3.5)

where c · n = 0 is used in the second equality. Since b = −|b|n (in the reference lattice)and sin(−∆θ) = − sin(∆θ), the number of pierced dislocations reads

n = 2 sin

(−∆θ

2

) |p||b| (3.6)

For the [001] rotation axis and associated (h k 0) boundary planes, the boundary periodlength |p| = a

√h2 + k2. With |b| = a

2

√2, Eq. (3.6) then gives n = 2 dislocations crossed

by vector p, as can be verified in Fig. 3.4. Hence, the intrinsic net defect Bp = 2b =2a2 [1 1 0] and the intrinsic net defect density |Bp|/|p| = 0.22.

This procedure is repeated for all the grain boundaries within the complete misorientationrange. The resulting intrinsic net defect density versus misorientation angle is shown inFig. 3.5(a). The primary dislocations are those related to the (perfect) Σ1 (1 0 0) θ = 0◦

and Σ1 (1 1 0) θ = 90◦ crystal boundaries, i.e. b = a[1 0 0] and b = a2 [1 1 0], respectively.

As will be discussed in Section 3.6, when employing Frank’s formula, the multiplicity ofdislocation descriptions of the same boundary has to be dealt with. The misorientationangle at which the preferred description changes can be determined from the atomisticdata on the boundary energy. A natural choice would be the boundary associated witha cusp in the energy versus misorientation angle profile. These boundaries typicallycorrespond to favoured boundaries that have a simple boundary structure as observedcomputationally and experimentally [120, 122, 153, 180, 198]. Furthermore, automatingthe demarcation angle based on the coefficient of determination R2 in step 2 predicts

Page 62: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.4 Multiscale approach to GB structure and energy 49

0 10 20 30 40 50 60 70 80 900

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Misorientation angle, θ (degrees)

Intr

insic

net defe

ct density (

−)

Al <100>

(a)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90

100

200

300

400

500

600

Intrinsic net defect density (−)

Gra

in b

ou

ndary

energ

y (

mJ/m

2)

Al <100>

(b)

0 10 20 30 40 50 60 70 80 900

100

200

300

400

500

600

Misorientation angle, θ (degrees)

Gra

in b

oundary

energ

y (

mJ/m

2)

Al <100>

(c)

Figure 3.5: Symmetric tilt grain boundaries with the 〈100〉 rotation axis in Al. (a) The intrinsic netdefect density versus misorientation angle. (b) The grain boundary energy versus intrinsic net defectdensity with the fit in red. (c) The grain boundary energy versus misorientation angle with the energyprofile resulting from the multiscale approach in red; the boundary energy is split into representativecore (solid blue) and elastic strain (dashed blue) terms.

angles that are favoured boundaries or boundaries reasonably close to these and, hence,confirms the natural choice. Here, the demarcation angle for the 〈100〉 rotation axiscorresponds to the Σ5 (3 1 0) θ = 36.87◦ boundary.

Step 2. The grain boundary energies, obtained from the atomistic simulations, versusthe intrinsic net defect densities, acquired in the first step of the procedure, are shownin Fig. 3.5(b). Applying a standard least squares fit, the parameters M and N in thelogarithmic relation, Eq. (3.3), read M = 425 mJ/m2 and N = 842 mJ/m2. At thispoint, it is emphasised that the complete misorientation range, including both low- and

Page 63: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

50 3 A multiscale model of grain boundary structure and energy

high-angle grain boundaries, can be captured remarkably well (R2 = 0.949) using thelogarithmic relationship.

Step 3. Upon combination of the two previous steps, the grain boundary energy versusmisorientation angle profile resulting from the multiscale procedure is shown in Fig. 3.5(c)along with the atomistically computed grain boundary energy data points. Furthermore,the split of the boundary energy into representative core and strain energy terms isshown. As can be observed, the continuum energy profile and its split are discontinuous,which is related to the change in dislocation description, see Fig. 3.5(a).

3.5 Results

The multiscale atomistic-to-continuum approach, as described above, is applied to theatomistic data sets of symmetric tilt boundaries with the 〈100〉, 〈111〉 and 〈110〉 rotationaxes in aluminium and copper. The results are shown in Figs. 3.5-3.10. In each figure onecan find for Al or Cu (i) the intrinsic net defect density versus the misorientation angle,(ii) the boundary energy versus the intrinsic net defect density with its correspondinglogarithmic fit and (iii) the boundary energy versus misorientation angle profile followingfrom the multiscale approach along with the split of the grain boundary energy intorepresentative core and strain energy terms. The values of the continuum parametersM and N (Eq. (3.3)) and the coefficient of determination R2 of each fit are provided inTable 3.1 for all three rotation axes in both Al and Cu. Further detailed information perrotation axis can be found in Appendix B.

For copper, Fig. 3.6, the relation between the grain boundary energy and the intrinsicnet defect density matches adequately (R2 = 0.976) using the logarithmic expression, seeFig. 3.6(b). For the 〈111〉 rotation axis in Al and Cu, see Figs. 3.7 and 3.8, respectively,the misorientation angle range is θ ∈ [0◦, 60◦]. The primary dislocations are the onescorresponding to the (perfect) Σ1 (0 1 1) θ = 0◦ reference boundary, i.e. b = a

2 [0 1 1]. Itis emphasised that the boundary energy versus intrinsic net defect density relations areagain accurately represented (R2 = 0.996 and 0.998 for Al and Cu, respectively) in thiscase, see Figs. 3.7(b) and 3.8(b).

The results in the case of the 〈110〉 rotation axes are shown in Figs. 3.9 and 3.10 for Al andCu, respectively. For this rotation axis, the misorientation angle range is θ ∈ [0◦, 180◦].This is a more difficult case, because of the presence of the special Σ3 (1 1 1) θ = 109.47◦

coherent twin boundary. To address this, the following considerations have been applied.Coherent twin boundaries can be described by simple shear, not requiring any dislocations[33], although a dislocation description can also be associated with it [81]. Furthermore,the Σ3 coherent twin boundary has a low energy. Therefore, it may be considered as a‘perfect’ crystal boundary with no dislocation content and zero energy. (Note that theΣ11 (1 1 3) θ = 50.48◦ also has a relatively low energy; however, in this case, a distributionof dislocations, and thus a non-zero net defect, is still attributed.) The misorientationdomain can then be divided into two parts delineated by the Σ3 (1 1 1) boundary, i.e.θ ∈ [0◦, 109.47◦] and θ ∈ [109.47◦, 180◦]. In the first misorientation regime, the primary

Page 64: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.5 Results 51

0 10 20 30 40 50 60 70 80 900

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Misorientation angle, θ (degrees)

Intr

insic

net defe

ct density (

−)

Cu <100>

(a)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90

100

200

300

400

500

600

700

800

900

1000

Intrinsic net defect density (−)

Gra

in b

ou

nd

ary

en

erg

y (

mJ/m

2)

Cu <100>

(b)

0 10 20 30 40 50 60 70 80 900

100

200

300

400

500

600

700

800

900

1000

Misorientation angle, θ (degrees)

Gra

in b

oundary

energ

y (

mJ/m

2)

Cu <100>

(c)

Figure 3.6: Symmetric tilt grain boundaries with the 〈100〉 rotation axis in Cu. (a) The intrinsic netdefect density versus misorientation angle. (b) The grain boundary energy versus intrinsic net defectdensity with the fit in red. (c) The grain boundary energy versus misorientation angle with the energyprofile resulting from the multiscale approach in red; the boundary energy is split into representativecore (solid blue) and elastic strain (dashed blue) terms.

dislocations are those related to the (perfect) Σ1 (0 0 1) θ = 0◦ reference boundary andthe (secondary) dislocations having the Σ3 (1 1 1) θ = 109.47◦ boundary as the reference,i.e. b = a[0 0 1] and b = a

3 [1 1 1], respectively. In the second misorientation regime,the (secondary) dislocations again correspond to the Σ3 (1 1 1) θ = 109.47◦ referenceboundary, whereas the primary dislocations take the Σ1 (1 1 0) θ = 180◦ boundary asthe reference, i.e. b = a

3 [1 1 1] and b = a2 [1 1 0], respectively. The demarcation angles in

the two misorientation regimes correspond to the Σ11 (1 1 3) θ = 50.48◦ and Σ9 (2 2 1)θ = 141.06◦ boundaries, respectively. It is observed that the resulting boundary energyversus the intrinsic net defect density in each of the two misorientation regimes matches

Page 65: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

52 3 A multiscale model of grain boundary structure and energy

0 10 20 30 40 50 600

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Misorientation angle, θ (degrees)

Intr

insic

net defe

ct density (

−)

Al <111>

(a)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90

50

100

150

200

250

300

350

400

450

500

Intrinsic net defect density (−)

Gra

in b

ou

ndary

energ

y (

mJ/m

2)

Al <111>

(b)

0 10 20 30 40 50 600

50

100

150

200

250

300

350

400

450

500

Misorientation angle, θ (degrees)

Gra

in b

oundary

energ

y (

mJ/m

2)

Al <111>

(c)

Figure 3.7: Symmetric tilt grain boundaries with the 〈111〉 rotation axis in Al. (a) The intrinsic netdefect density versus misorientation angle. (b) The grain boundary energy versus intrinsic net defectdensity with the fit in red. (c) The grain boundary energy versus misorientation angle with the energyprofile resulting from the multiscale approach in red; the boundary energy is split into representativecore (solid blue) and elastic strain (dashed blue) terms.

well (R2 > 0.9), see Figs. 3.9(b),(c) and 3.10(b),(c). Similar to the 〈100〉 rotation axis,discontinuities are present in the continuum boundary energy versus misorientation angleprofiles and in the decompositions into representative core and strain energy terms, seeFigs. 3.9(d) and 3.10(d), related to the change in dislocation descriptions, see Figs. 3.9(a)and 3.10(a).

The total number of continuum parameters in the present work, examining symmetrictilt grain boundaries in the case of three rotation axes in aluminium and copper, equals8 for one material, for the full misorientation ranges, see Table 3.1. Inspired by the

Page 66: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.5 Results 53

0 10 20 30 40 50 600

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Misorientation angle, θ (degrees)

Intr

insic

net defe

ct density (

−)

Cu <111>

(a)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90

100

200

300

400

500

600

700

800

900

Intrinsic net defect density (−)

Gra

in b

ou

ndary

energ

y (

mJ/m

2)

Cu <111>

(b)

0 10 20 30 40 50 600

100

200

300

400

500

600

700

800

900

Misorientation angle, θ (degrees)

Gra

in b

oundary

energ

y (

mJ/m

2)

Cu <111>

(c)

Figure 3.8: Symmetric tilt grain boundaries with the 〈111〉 rotation axis in Cu. (a) The intrinsic netdefect density versus misorientation angle. (b) The grain boundary energy versus intrinsic net defectdensity with the fit in red. (c) The grain boundary energy versus misorientation angle with the energyprofile resulting from the multiscale approach in red; the boundary energy is split into representativecore (solid blue) and elastic strain (dashed blue) terms.

work in [84], it has been investigated if a scaling relationship exists between the grainboundary energies of the metallic elements considered in this work. This would implythat the number of continuum parameters for different FCC metals can be reduced.The result is shown in Fig. 3.11. As pointed out in [84], the linear trend suggests thatthe boundary energy can be scaled by an element-dependent proportionality constant.Material parameters can be taken for this constant. In atomistic simulations of materialsystems, appropriate interatomic potentials need to be used that reproduce materialparameters consistent with experimental data and ab initio calculations. The lines inFig. 3.11 indicate the scaling for three different material parameters following from the

Page 67: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

54 3 A multiscale model of grain boundary structure and energy

0 20 40 60 80 100 120 140 160 1800

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Misorientation angle, θ (degrees)

Intr

insic

ne

t d

efe

ct

de

nsity (

−)

Al <110>

(a)

0 0.2 0.4 0.6 0.8 10

50

100

150

200

250

300

350

400

450

Intrinsic net defect density (−)

Gra

in b

ou

nd

ary

en

erg

y (

mJ/m

2)

Al <110>

(b)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.70

50

100

150

200

250

300

350

400

450

500

Intrinsic net defect density (−)

Gra

in b

ounda

ry e

nerg

y (

mJ/m

2)

Al <110>

(c)

0 20 40 60 80 100 120 140 160 1800

50

100

150

200

250

300

350

400

450

500

Misorientation angle, θ (degrees)

Gra

in b

ou

nd

ary

en

erg

y (

mJ/m

2)

Al <110>

(d)

Figure 3.9: Symmetric tilt grain boundaries with the 〈110〉 rotation axis in Al. (a) The intrinsic netdefect density versus misorientation angle. (b),(c) The grain boundary energy versus intrinsic net defectdensity with the fit in red for the two misorientation regimes, θ ∈ [0◦, 109.47◦] (b) and θ ∈ [109.47◦, 180◦](c). (d) The grain boundary energy versus misorientation angle with the energy profile resulting fromthe multiscale approach in red; the boundary energy is split into representative core (solid blue) andelastic strain (dashed blue) terms.

interatomic potentials for Al and Cu [123, 124] employed in this work. The ratios of thestacking fault energy γSF, the elastic constant aC44 and the Voigt average shear modulusaµ are used for the scaling, the latter two being multiplied by the lattice constant fordimensional consistency with the grain boundary energy. Here, the Voigt average shearmodulus seems to be a reasonable candidate for the scaling, although the scaling is lessgood for higher energy boundaries.

Page 68: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.6 Discussion 55

0 20 40 60 80 100 120 140 160 1800

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Misorientation angle, θ (degrees)

Intr

insic

ne

t d

efe

ct

de

nsity (

−)

Cu <110>

(a)

0 0.2 0.4 0.6 0.8 10

100

200

300

400

500

600

700

800

Intrinsic net defect density (−)

Gra

in b

ou

nd

ary

en

erg

y (

mJ/m

2)

Cu <110>

(b)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.70

100

200

300

400

500

600

700

800

900

Intrinsic net defect density (−)

Gra

in b

ounda

ry e

nerg

y (

mJ/m

2)

Cu <110>

(c)

0 20 40 60 80 100 120 140 160 1800

100

200

300

400

500

600

700

800

900

Misorientation angle, θ (degrees)

Gra

in b

ou

nd

ary

en

erg

y (

mJ/m

2)

Cu <110>

(d)

Figure 3.10: Symmetric tilt grain boundaries with the 〈110〉 rotation axis in Cu. (a) The intrinsic netdefect density versus misorientation angle. The grain boundary energy versus intrinsic net defect densitywith the fit in red for the two misorientation regimes, θ ∈ [0◦, 109.47◦] (b) and θ ∈ [109.47◦, 180◦] (c).(d) The grain boundary energy versus misorientation angle with the energy profile resulting from themultiscale approach in red; the boundary energy is split into representative core (solid blue) and elasticstrain (dashed blue) terms.

3.6 Discussion

3.6.1 Frank-Bilby equation

As already discussed extensively in the literature, two sources of ambiguity exist whenapplying the Frank-Bilby equation to determine the dislocation content of a grain bound-ary [175]. The first source of ambiguity is due to the rotational symmetries in the crystal

Page 69: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

56 3 A multiscale model of grain boundary structure and energy

Table 3.1: Values of continuum parameters M and N (in mJ/m2) for each rotation axis in Al and Cu.The coefficient of determination R2 and the misorientation domains are indicated.

〈100〉 axis 〈111〉 axis 〈110〉 axis

Al Cu θ Al Cu θ Al Cu θ

M 425 901 [0◦,90◦] 400 766 [0◦,60◦]177 358 [0◦,109.47◦]395 797 [109.47◦,180◦]

N 842 1329 [0◦,90◦] 803 1331 [0◦,60◦]905 1434 [0◦,109.47◦]755 1263 [109.47◦,180◦]

R2 0.949 0.976 [0◦,90◦] 0.996 0.998 [0◦,60◦]0.905 0.983 [0◦,109.47◦]0.966 0.984 [109.47◦,180◦]

that can lead to a multiplicity of dislocation descriptions of the same grain boundary.However, Balluffi and Olson [12] argued that no problems of real physical significancearise due to this lack of uniqueness, whereby any of the various possible descriptionscan be used. Comparison of the dislocation descriptions to the image of the boundaryobtained by electron microscope diffraction contrast may guide the selection of the de-scription. Furthermore, Cahn et al. [27, 28] considered the migration of high-angle grainboundaries coupled to shear deformation. The grain boundaries are shifted by shearstresses and their motion agrees very accurately with the one or the other (sometimesboth) solution of the Frank-Bilby equation over the entire misorientation range, whichthey verified by atomistic simulations. Similarly, in this work, in the case of the 〈100〉and 〈110〉 rotation axes, the multiplicity is utilized in the relation between the grainboundary energy and the intrinsic net defect density.

0 100 200 300 400 500 6000

100

200

300

400

500

600

700

800

900

1000

1100

Grain boundary energy Al (mJ/m2)

Gra

in b

ou

nd

ary

en

erg

y C

u (

mJ/m

2)

<100> axis

<110> axis

<111> axis

ratio of ac44

ratio of aµ

ratio of γsf

Figure 3.11: Comparison of the grain boundary energies for Al and Cu, having high and moderatestacking fault energies, respectively. The lines are the scaling using three different material parameters.

Page 70: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.6 Discussion 57

The other source of ambiguity is the choice of a reference structure in which the dislo-cation content is defined. It has been argued that this choice may be arbitrary, oftentaking one of the adjacent crystals or the median lattice as the reference lattice. How-ever, as discussed in [193], who included the consideration of elastic fields along with thegeometry, the choice of the reference lattice is not arbitrary, since it influences the valuesof the far-field stresses, strains and rotations associated with the interface dislocations.These, in turn, are usually subject to constraints, namely that the long-range stresses areabsent, which is a requirement for the application of the Frank-Bilby equation, and thatthe far-field rotations are consistent with the prescribed misorientation. They showedthat for symmetric tilt and twist boundaries only the median reference lattice leads tozero long-range stresses. In the present work, the median lattice is therefore taken as thereference lattice.

3.6.2 Grain boundary structure

Rittner and Seidman [153] investigated the grain boundary structures of 〈110〉 symmetrictilt grain boundaries in FCC metals with low stacking fault energies using atomistic sim-ulations. One of the main conclusions in their work is that the grain boundary structuresin low stacking fault energy materials may be quite wide, i.e. on the order of 1 nm, dueto delocalization or dissociation of the primary or secondary grain boundary dislocations.These structures tend to have stronger local atomic relaxations than those typical forgrain boundaries in higher stacking fault energy materials. Therefore, a structural unitmodel is not found to be useful for describing low to moderate stacking fault energy FCCmetals, such as copper. However, in this work the interest is in obtaining an average,continuum scale, representation of the initial grain boundary structure without consid-ering small-scale details, such as steps, microfacets and partial dislocations, which aretypically not resolved in continuum level frameworks. The effect of these local atomicrelaxations is captured implicitly by relating the corresponding net defect to the relaxedboundary energy. Indeed, the total net defect will be conserved, even if dislocations dis-sociate into partials. Therefore, the applicability of the developed multiscale approachis not restricted by atomistic limitations of the structural unit model.

In order to analyse relaxed grain boundary structures, Wolf [206, 207] defined a coor-dination coefficient that represents a convenient measure of how well on average thegrain boundary is coordinated. The coordination coefficient equals the number of bro-ken nearest-neighbor bonds per unit area and is closely related to the radial distributionfunction, which contains all the detailed information on the distribution of interatomicseparations. It was found that a good correlation exists between the coordination co-efficient and the grain boundary energy. Examining the coordination coefficient versusmisorientation plots for the 〈100〉, 〈110〉 and 〈111〉 rotation axes in [206], it is observedthat similar trends are present in the intrinsic net defect density versus misorientationangle profiles in this work, see Figs. 3.5(a), 3.6(a), 3.7(a), 3.8(a), 3.9(a), 3.10(a). Wolf[206] also pointed out that a connection seems to exist between the measure of coordina-tion and geometric models describing the (more local) structure of symmetric tilt grainboundaries, without further investigation. This work presents an approach that supportsthis statement.

Page 71: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

58 3 A multiscale model of grain boundary structure and energy

3.6.3 Grain boundary energy

The results in Section 3.5 show that the atomistic-to-continuum approach, employingdislocation descriptions representing the initial grain boundary structure, adequatelymatches the atomistically computed grain boundary energies (R2 > 0.9). As mentionedin Section 3.4, for high-angle grain boundaries dislocation cores start to overlap anddislocations can no longer be resolved discretely. Yet, remarkably good results are stillobtained, even for the 〈110〉 rotation axis including the special case of the Σ3 coherenttwin boundary. Moreover, the adopted logarithmic relation (Eq. (3.3)) is similar to theRead-Shockley equation applicable for low misorientation boundaries [150, 151]. Theboundary energy, according to the Read-Shockley equation, consists of contributions dueto the core energy and elastic strain energy. For increasing misorientation, the disloca-tion spacing decreases and the stress fields associated with isolated dislocations canceleach other over larger areas. Hence, for high-angle grain boundaries it is expected thatthe relative contribution of the core energy increases. This behaviour has been observedin [205, 207], who proposed a strictly empirical extension of the Read-Shockley model tohigh-angle grain boundaries by replacing θ with sin(θ) and tested its validity by compar-ison with energies of some twist and tilt boundaries in copper and gold obtained fromatomistic simulations. In this work, the boundary energy is related to a continuum mea-sure of the boundary structure. The total grain boundary energy has been decomposedinto terms representative of the core and elastic strain energy (Eq. (3.4)). Indeed, itis seen in Figs. 3.5(c), 3.6(c), 3.7(c), 3.8(c), 3.9(d), 3.10(d) that for high-angle bound-aries the relative contribution of the core energy increases and can dominate. Hence,this suggests that the energetics of high-angle boundaries is represented in a physicallycorrect manner by the proposed multiscale approach. Note, however, that the decompo-sition is less ‘clean’ when grain boundary dissociation occurs, which is the case for someboundaries with the 〈110〉 rotation axis in Cu and to a much lesser extent in Al due tothe higher stacking fault energy. The contributions to the total grain boundary energythen also include terms resulting from the excess energy associated with the stackingfaults created from dislocation core spreading and the interaction energy between thetwo closely spaced sub-boundaries that have been formed [153].

It is of interest to compare the current findings with experimental work in the literature.However, a direct comparison regarding the grain boundary energies cannot be done fortwo reasons [197]. First, the experimentally measured values are not the internal ener-gies, but free energies. Whereas the atomistic simulations in this work are at 0 K, theexperimental measures are usually at much higher temperatures, such that the entropiccontribution due to atomic thermal vibrations is not negligible. Secondly, atomistic stud-ies employing empirical interatomic potentials cannot be expected to give exact numericalvalues of boundary energies in a given material. In the work in [182], a comparison wasmade between the geometric structures observed in HRTEM images and those expectedfrom the O-lattice theory. They found a good correlation between the grain boundarydislocation density and the grain boundary energy obtained from experiments [133] for〈110〉 symmetric tilt boundaries in aluminium for the whole misorientation range, i.e.low- and high-angle grain boundaries. Using three types of Burgers vectors, two of whichare equal to the ones in this work, they found, in contrast to the Read-Shockley theory,

Page 72: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

3.7 Conclusions 59

that the data points could be fit irrespective of the specific choice of the Burgers vectors.The boundary energy as a function of the grain boundary dislocation density parameterwas well described by a relation similar to the Read-Shockley equation and Eq. (3.3)in this work. In order to facilitate comparison with their fitting parameters (α = 837mJ/m2 and B = 0.20), a single fit in the whole misorientation range has been performed,resulting in N = 934 mJ/m2 and M/N = 0.20, which are in good agreement. Note thatif we would restrict ourselves to small misorientation angles, more dedicated fits could bemade that would enable a direct comparison with experimental data of 〈100〉 low-anglegrain boundaries in Cu [70] and Al [132]. However, this was not the focus here.

3.6.4 Model extension

Randle [149] reviewed experimental data on the statistics of grain boundary plane crys-tallography in polycrystals. Asymmetric tilt grain boundaries prevail over symmetric tiltgrain boundaries in FCC metals, such as nickel, copper and gold. Only few twist grainboundaries were reported. Furthermore, over half of the 900 grain boundaries char-acterised in annealed high-purity nickel and copper consisted of 〈110〉 tilt boundaries.Symmetric tilt boundaries that have relatively large energies may reduce their energy byrotation of the grain boundary plane. This inclination of the boundary plane rendersthe boundary asymmetric. These conclusions suggest to also extend the developed mul-tiscale approach to asymmetric tilt boundaries in the future, using published atomisticdata [179, 180]. In this case, two sets of dislocations are required to describe the bound-ary. The determination of the dislocation content again employs Frank’s formula, butnow following the techniques as detailed in [60, 82, 175]. It is expected from [150, 161]that the form of the grain boundary energy in Eq. (3.3) is still applicable, but the pa-rameters M and N will then become dependent on the inclination angle. Extension ofthe developed approach towards more complex grain boundaries, including asymmetry,twist and general boundaries, poses extra challenges related to the details of the grainboundary structure.

Finally, comparing the grain boundary energies of aluminium and copper (Fig. 3.11) alinear trend is observed, but some outliers and scatter are present. As discussed in [84],this is due to the increased differences in grain boundary structure when materials differstrongly in stacking fault energy. Aluminium and copper have a high and moderatestacking fault energy, respectively. From the observations in [84], it is expected thata strong correlation exists when comparing metallic elements of similar stacking faultenergy that will enable a reduction in the number of continuum parameters for capturingthe grain boundary energy.

3.7 Conclusions

In this chapter, a multiscale approach to obtain a continuum representation of the initialgrain boundary structure and energy has been developed based on the output of atomistic

Page 73: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

60 3 A multiscale model of grain boundary structure and energy

simulations. An intrinsic net defect density, representing the initial boundary structure,is related to atomistically computed grain boundary energies for symmetric tilt grainboundaries with the 〈100〉, 〈110〉 and 〈111〉 rotation axes in aluminium and copper. Thekey features of the proposed approach are:

1. Grain boundary energies and boundary period vectors are obtained from atomisticsimulations.

2. Structural unit and dislocation descriptions of the grain boundary structure areestablished.

3. The Frank-Bilby equation to determine the dislocation content is exploited.

The developed atomistic-to-continuum approach can be directly incorporated in grainboundary interface continuum models, realistically reflecting the initial boundary struc-ture and energetics. Note that the current approach is by no means unique and differentmethods to capture structure and energy may be applied.

The results have shown that the multiscale approach yields adequate descriptions (R2 >0.9) of the atomistically computed grain boundary energies for both low- and high-anglegrain boundaries in the complete misorientation range. A comparison with literaturebased on experimental results for the 〈110〉 rotation axis in Al shows a good agreementand supports the findings in this work.

Possible future directions are twofold. Firstly, from a materials science perspective, anextension of the current work may include the application and elaboration of the de-veloped atomistic-to-continuum approach to (i) the full space of macroscopic DOFs ofa grain boundary, starting with asymmetric tilt boundaries and twist boundaries, (ii)grain boundaries in body-centered cubic and hexagonal close-packed metals, and (iii)interphase boundaries. Furthermore, metallic elements of similar stacking fault energycould be investigated to verify the possibility of the reduction of the number of contin-uum parameters through scaling. Secondly, from a mechanics perspective, the multiscaleapproach is to be implemented into a continuum level extended crystal plasticity frame-work to investigate its influence of the atomistically informed grain boundary energeticson the macroscopic response, which will be the topic of the next chapter.

The issue of the evolution of the grain boundary structure and energy upon plasticdeformation remains a challenge. This will demand methods that account for the changein the elastic energy of the boundary in the continuum bulk-interface representation, inaddition to the interface-specific evolution of the dislocation core energy as dislocationsfrom the bulk react with the interface.

Page 74: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Chapter 4

Grain boundary plasticityincorporating internal structure

and energy1

Abstract

Modelling the behaviour of grain boundaries in polycrystalline metals using macroscopiccontinuum frameworks demands a multiscale description of the underlying details of thestructure and energy of grain boundaries. The objective in this chapter is the incor-poration of a multiscale atomistic-to-continuum approach of the initial grain boundarystructure and energy into a grain boundary extended crystal plasticity framework in orderto investigate the role of the grain boundary energetics on the macroscopic response. Tothis end, the methodology includes: (i) the generalisation of the atomistic-to-continuumresults of the initial grain boundary structure and energy, (ii) an analytical analysis ofthe resulting grain boundary energetics in the continuum framework, (iii) the numericalimplementation of the developed framework in the case of a periodic bicrystal subjectedto simple shear deformation considering a symmetric tilt boundary system in the fullmisorientation range. This work provides a step forward towards the physically basedcontinuum modelling of grain boundary interfacial plasticity.

1Reproduced from [189]: P.R.M. van Beers, V.G. Kouznetsova, and M.G.D. Geers. Grain boundaryinterfacial plasticity with incorporation of internal structure and energy. Submitted.

61

Page 75: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

62 4 Grain boundary plasticity incorporating internal structure and energy

4.1 Introduction

In polycrystalline materials, differences in the orientation of the crystals that constitutethe material play a significant role in the macroscopic behaviour. In spite of numerousinvestigations in the past, the physical understanding of the underlying deformationmechanisms defining this macroscopic behaviour and its incorporation into continuumlevel modelling are still the subject of ongoing studies. A well-known example is theincreased strength and hardening due to grain size refinement, i.e. the Hall-Petch effect,which has led to advanced material models capturing these size effects. However, whenthe average grain size reduces, also the behaviour of the grain boundaries (GB) and theirinteraction with crystal defects become of key importance. Various types of interactionsbetween dislocations and grain boundaries have been reported experimentally [67, 101,159, 166, 186] and in atomistic simulations [9, 45, 139, 156, 216]. Dislocations can beaccumulated, transmitted, absorbed or nucleated at the interfaces.

From a continuum modelling point of view, conventional continuum plasticity models,even those enhanced to include strain gradients and interface behaviour [41, 64, 114, 141,195], do not explicitly account for crystallographic misorientation across a grain bound-ary, which is an important quantity defining micro-scale interactions between dislocationsand grain boundaries. Classical crystal plasticity models overcome this deficiency, butthey generally do not account for grain boundary phenomena, i.e. grain boundaries areonly incorporated as planes where the crystallographic orientation changes. Gradientenhanced crystal plasticity resolves this limitation. While these approaches mostly in-corporate the limiting situations of either impenetrable or completely transparent grainboundaries through higher-order boundary conditions [16, 56, 76, 95, 115], several at-tempts have been made to elaborate these frameworks to include grain boundary inter-face mechanics [51, 77, 92, 110, 134, 210]. Among the different constitutive choices thathave to be made, the grain boundary energy often follows a simple ad hoc expression.However, atomistic simulations indicate that the grain boundary energy is dependent onthe grain boundary structure [79, 153, 180, 198, 206]. Currently, a limited number of thecontinuum modelling approaches (can) address the underlying initial and evolving grainboundary structure [66, 178, 187].

In the previous chapter, a multiscale approach has been developed that leads to a con-tinuum representation of the initial grain boundary structure and energy based on theoutput of atomistic simulations. The proposed approach introduces an intrinsic net defectdensity, representing the grain boundary structure, which is related to the atomisticallycalculated grain boundary energies through a logarithmic expression. It was shown thatthis multiscale method can accurately describe the grain boundary energies in the com-plete misorientation range of symmetric tilt grain boundary systems in aluminium andcopper. In this chapter, this atomistic-to-continuum approach is incorporated into thegrain boundary extended crystal plasticity model, developed in Chapter 2. The resultingmodel is implemented numerically in order to examine the particular role and influ-ence of the grain boundary energetic contributions on the macroscopic response. Themodelling approach is illustrated on a periodic copper bicrystal subject to simple sheardeformation, considering symmetric tilt boundaries with the [001] rotation axis in thefull misorientation range, including both low- and high-angle grain boundaries.

Page 76: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.2 Bulk and interface frameworks 63

The outline of this chapter is as follows. In Section 4.2, a summary of the grain bound-ary extended crystal plasticity framework is provided. Section 4.3 briefly describes themultiscale atomistic-to-continuum approach of the initial grain boundary structure andenergy, and covers the incorporation of the atomistic-to-continuum approach into thegrain boundary interface model within the strain gradient crystal plasticity framework.In Section 4.4, the resulting initial grain boundary energetics is analysed analytically.The numerical implementation and the results of the effect of the energetic contributionsof the grain boundaries on the macroscopic response of copper bicrystals are describedin Section 4.5. This is followed by a discussion in Section 4.6. Finally, conclusions aremade in Section 4.7.

4.2 Bulk and interface frameworks

The microstructurally motivated strain gradient crystal plasticity framework including agrain boundary interface model as developed in Chapter 2 is summarised next.

4.2.1 Bulk crystal formulation

The bulk crystal formulation originates from the one proposed by Evers et al. [56, 57]and Bayley et al. [16], restricted to small deformations. The total distortion is given bythe sum of elastic (e) and plastic (p) components

∇u = He + Hp . (4.1)

The plastic distortion is given by

Hp =∑α

γαsαmα . (4.2)

Here, γα represents slip on slip system α having slip plane normal mα and slip direc-tion sα. In the crystal plasticity approach adopted, dislocation densities on individualslip systems act as key variables. Geometrically necessary dislocation (GND) densities

of edge and screw character, denoted ρα(e)g and ρ

α(s)g , respectively, arise from gradients

in slip as follows

ρα(e)g = ρ

α(e)g0 − 1

b∇γα · sα and ρ

α(s)g = ρ

α(s)g0 +

1

b∇γα · lα , (4.3)

with unit vector lα = sα×mα. The quantities ρα(e)g0 and ρ

α(s)g0 represent the initial GND

densities and b is the Burgers vector magnitude. For a given slip system α the resolvedshear stress or Schmid stress τα is given by

τα = σ : sαmα , (4.4)

Page 77: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

64 4 Grain boundary plasticity incorporating internal structure and energy

where the Cauchy stress tensor σ is determined through the conventional elastic rela-tionship

σ = 4C : Ee with Ee =1

2

(He + HeT

). (4.5)

Here, 4C is the fourth order elasticity tensor and Ee the elastic strain. The evolution ofslip is accomplished through a visco-plastic power-law by

γα = γ0

( |τα − ταb |sα

) 1m

sign (τα − ταb ) . (4.6)

Parameters γ0 and m are positive constants. The back stress ταb is obtained by theprojection of the internal stress σint onto slip system α

ταb = −σint : sαmα . (4.7)

The internal stress can be interpreted as a fine-scale correction, accounting for microstruc-tural effects not resolved by the long-range elastic strain field, e.g. through dislocationpile-ups. Spatial gradients in the densities of geometrically necessary dislocations inducelong- and short-range non-vanishing stress fields. In the derivation, the stress fields ofdislocations in an infinite elastically isotropic medium are used. Further, by assuming alinear spatial variation of the GND densities of both edge and screw type within a radiusR around a material point, the following expression for the internal stress accounting forcontributions from all slip systems α is obtained (see [16] for more details)

σint =GbR2

8 (1− ν)

∑α

∇ρα(e)g · 3Cα(e) +

GbR2

4

∑α

∇ρα(s)g · 3Cα(s) , (4.8)

with the third order slip system orientation tensors

3Cα(e) = 3mαsαsα + mαmαmα + 4νmαlαlα − sαsαmα − sαmαsα ,

3Cα(s) = −mαlαsα −mαsαlα + lαmαsα + lαsαmα . (4.9)

The length scale parameter R determines the extent of the nonlocal influence, G is theshear modulus and ν is the Poisson’s ratio. Due to the focus on the grain boundarybehaviour in this chapter, the back stress, Eq. (4.7), is simplified, taking into accountonly the ‘self’ internal stress components [16], i.e. only the last two terms in the thirdorder slip system orientation tensors in Eq. (4.9) are considered.

In Eq. (4.6), the slip resistance sα for slip system α reflects the obstruction of dislocationglide through short-range interactions of all dislocations on coplanar and intersecting slipsystems. It depends on both the densities of geometrically necessary (GNDs), ρξg, and

statistically stored dislocations (SSDs), ρξs , [5]

sα = Gb

√∑ξAαξ

∣∣ρξs ∣∣+∑

ξAαξ

∣∣ρξg∣∣ . (4.10)

Page 78: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.2 Bulk and interface frameworks 65

Here, Aαξ represents the full dislocation interaction matrix as given by Franciosi andZaoui [61]. The components Aαξ are the coefficients a0, a1, a2 and a3, which representthe strength of six types of interactions: self hardening (a0), coplanar (a1), cross slip(a1), Hirth lock (a1), glissile junction (a2) and Lomer-Cottrell lock (a3). The evolution

of the SSD densities ρξs is described by the generalised form of Essmann and Mughrabi[55]

ρξs =1

b

(1

Lξ− 2ycρ

ξs

) ∣∣γξ∣∣ with ρξs (t = 0) = ρξs0 . (4.11)

The dislocation accumulation is governed by the average dislocation segment length

Lξ =K√∑

ζ Hξζ∣∣ρζs ∣∣+

∑ζ H

ξζ∣∣ρζg∣∣ , (4.12)

where Hξζ are the immobilisation coefficients, structured analogously to the interactioncoefficients Aαξ. The dislocation annihilation is controlled by a critical annihilationlength yc. More details on this gradient crystal plasticity framework can be found in[56, 57] and [16].

4.2.2 Grain boundary interface formulation

The interface framework, developed in Chapter 2, allows for the incorporation of macro-scopic sliding and opening of the interface. However, this is not considered here sincecontinuity of displacements across the interface will be assumed. The different fine-scalephenomena taking place at the boundary, e.g. absorption, emission and transmission, arenot resolved separately at the scale of the model, but the net effect of these interactionsis lumped into a continuum scale quantity termed ‘interface normal slip’ q. The interfacenormal slip components of edge (e) and screw (s) dislocations at the interface Γ betweenadjacent grains A (slip systems α) and B (slip systems β) are defined as

grain A: qα(e) = γα sα · ngb and qα(s) = −γα lα · ngb ,

grain B: qβ(e) = −γβ sβ · ngb and qβ(s) = γβ lβ · ngb . (4.13)

The grain boundary normal ngb points from grain A to grain B. In a similar mannerto slip in the bulk crystal, evolution of the interface normal slip on slip system α at thegrain A side of Γ is modelled with a visco-plastic power-law (the expressions for grain Bare obtained by substituting β for α)

qα(e) = q0

(∣∣Fα(e)bk −Fα(e)

e

∣∣Rα(e)

) 1n

sign(Fα(e)

bk −Fα(e)e

),

qα(s) = q0

(∣∣Fα(s)bk −Fα(s)

e

∣∣Rα(s)

) 1n

sign(Fα(s)

bk −Fα(s)e

). (4.14)

Page 79: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

66 4 Grain boundary plasticity incorporating internal structure and energy

The parameters n and q0 are positive constants (similar to the bulk crystal parametersm and γ0). The quantities Rα(e) and Rα(s) are the interface normal slip resistances foredge and screw dislocations on slip system α, respectively. The bulk-induced interface

microforces Fα(e)bk and Fα(s)

bk are the projections of the slip system microstress ξα

Fα(e)bk = −ξα · sα and Fα(s)

bk = ξα · lα . (4.15)

The microstress ξα can be determined from the microstructurally motivated internal

stress σint, see Chapter 2 for the details. The energetic interface microforces Fα(e)e and

Fα(s)e , which can be interpreted as interface ‘back stresses’, are determined from the

grain boundary free energy φe

Fα(e)e =

∂φe

∂qα(e)and Fα(s)

e =∂φe

∂qα(s). (4.16)

The grain boundary free energy φe is linked to the grain boundary structure and itsevolution

φe = φe(Bgb

(qα(e,s), qβ(e,s)

)). (4.17)

The quantity Bgb is the grain boundary net defect, representing the grain boundarystructure, that evolves upon interaction with lattice dislocations through the interfacenormal slip measures q. The grain boundary structure and energy, as well as theirevolution will be further elaborated in the next section.

4.3 Incorporation of GB structure and energy

The incorporation of the internal structure and energy of grain boundaries into the grainboundary extended crystal plasticity framework is now addressed. Departing from thegrain boundary net defect description, the multiscale approach for a continuum repre-sentation of the initial structure and energy of grain boundaries, developed in Chapter3, is summarised. Next, the atomistic-to-continuum approach is incorporated into theextended crystal plasticity framework.

4.3.1 Net defect description

The structure and energy of a grain boundary depends on its initial state and the sub-sequent deformation history of the adjacent grains. When a lattice dislocation interactswith a grain boundary, the structure of the interface is altered. A residual defect re-mains, which is necessary for the conservation of the Burgers vector [101]. The followingvectorial measure of the average defect at a point on the grain boundary is defined

Bgb = B0 + B . (4.18)

B0 represents the initial grain boundary defect. Physically, it corresponds to the contri-bution of ‘intrinsic’ grain boundary dislocations [13, 82]. These dislocations form part of

Page 80: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.3 Incorporation of GB structure and energy 67

the ‘equilibrium’ structure of the boundary.2 The vector B represents the defect resultingfrom the interaction of lattice dislocations with the grain boundary during plastic de-formation. These ‘extrinsic’ grain boundary dislocations are not part of the equilibriumstructure of the boundary. The equations characterising these two contributions to thegrain boundary net defect Bgb will be detailed in the subsequent sections.

4.3.2 Initial GB structure and energy

The continuum representation of the initial boundary structure and energy is based onthe output of atomistic simulations (see Chapter 3) and incorporates: (i) grain boundaryenergies and boundary period vectors obtained from atomistic simulations, (ii) structuralunit and dislocation descriptions of the grain boundary structure and (iii) the Frank-Bilbyequation to determine the dislocation content.

The key concept in the multiscale approach is to attribute an intrinsic net defect Bp,representing the initial grain boundary structure, to a boundary period p for each bound-ary in the whole misorientation range around a certain rotation axis. A correspondingintrinsic net defect density |Bp|/|p| is related to the initial grain boundary energy φe0through a logarithmic relationship

φe0 = M|Bp||p| −N

|Bp||p| ln

( |Bp||p|

), (4.19)

where M and N are two parameters to be identified. This logarithmic relationshipresembles the Read-Shockley equation for low-angle boundaries [150, 151] and can beinterpreted as a generalisation to the complete misorientation range, including high-anglegrain boundaries. This atomistic-to-continuum approach has been applied to symmetrictilt grain boundaries with the 〈100〉, 〈110〉 and 〈111〉 rotation axes in aluminium andcopper. Fairly accurate descriptions of the grain boundary energy were thereby obtained.The reader is referred to Chapter 3 for more details. An example of symmetric tiltboundaries with the 〈100〉 rotation axis in Cu is shown in Fig. 4.1. Low- and high-anglegrain boundaries typically have (i) small and large intrinsic net defect densities and (ii)low and high initial grain boundary energies, respectively, see Fig. 4.1.

4.3.3 Evolution of GB structure and energy

As discussed in Section 4.2.2, the energetic interface microforces (Eq. (4.16)) are derivedfrom the grain boundary energy, which depends on the evolution of the grain boundarystructure upon plastic deformation. However, a reliable physically based description ofthe evolution of the structure and energy of a grain boundary upon plastic deformationremains a challenge at present. Even at the atomistic level, i.e. the only length scale at

2Note that any grain boundary in itself is a non-equilibrium crystal defect. In the strict thermody-namic sense a bicrystal containing a grain boundary may be in local equilibrium, i.e. a metastable stateonly [186].

Page 81: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

68 4 Grain boundary plasticity incorporating internal structure and energy

0 10 20 30 40 50 60 70 80 900

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Misorientation angle, θ (degrees)

Intr

insic

ne

t d

efe

ct

de

nsity (

−)

Cu <100>

low high low

(a)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90

100

200

300

400

500

600

700

800

900

1000

Intrinsic net defect density (−)

Initia

l G

B e

ne

rgy (

mJ/m

2)

Cu <100>

low high

(b)

Figure 4.1: Symmetric tilt grain boundaries with the 〈100〉 rotation axis in Cu (see Fig. 3.6). (a) Theintrinsic net defect density versus misorientation angle. (b) The initial grain boundary energy versusintrinsic net defect density, and the model according to Eq. (4.19), in blue. Symbols are the atomisticdata. The vertical lines are rough indications of regions corresponding to low- and high-angle grainboundaries.

which the grain boundary structure can be physically resolved, no systematic simulationsexist that track the evolution of the boundary structure and energy as a consequence ofthe interaction with multiple dislocations. Yet, the papers of Sangid et al. [155, 156],Tucker and McDowell [183], Tucker et al. [184] provide clear steps in this direction.Furthermore, the quasi-continuum (QC) [160, 164, 214] and the coupled atomistic discretedislocation (CADD) methods [46–48, 163] are multiscale techniques that couple atomisticand continuum descriptions enabling analyses of the collective interaction of dislocationswith grain boundaries at larger scales. In the future, such simulations are expected tofurther resolve the changes in the grain boundary energy as required at the continuumlevel. In this work, a generalisation of the expression of the initial grain boundary energy(Eq. (4.19)) is proposed.

Consider a length ` = k|p| along the grain boundary, where |p| is the size of boundaryperiod p and k is the number of boundary periods in `, not necessarily an integer. Thelength ` can be interpreted as the size of a Burgers circuit around the grain boundary.Since the intrinsic net defect Bp is defined within one boundary period, the intrinsiccontribution B0

` to the net defect Bgb (Eq. (4.18)) within length ` reads

B0` = kBp . (4.20)

A simple form of the extrinsic contribution B` within the segment of length ` due to theinteraction with lattice dislocations of edge and screw character from both sides of thegrain boundary is defined as

B` = `∑α

(qα(e) + qα(s)

)sα + `

∑β

(qβ(e) + qβ(s)

)sβ . (4.21)

Page 82: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.4 Analysis of initial GB energetics 69

The following defect densities per unit length of grain boundary are defined

B0 =B0`

`=

Bp

|p| and B =B`

`. (4.22)

With the net defect density defined by (4.18), the initial grain boundary energy φe0 isthen generalised to a general grain boundary energy φe that depends on the net defectdensity Bgb

φe = M |Bgb| −N |Bgb| ln (|Bgb|) + 12CeN |Bgb −B0|2 , (4.23)

where |Bgb| =√Bgb ·Bgb and Ce is a positive constant. Finally, using Eqs. (4.16),

(4.18) and (4.20)-(4.23), the energetic interface microforces for slip system α in grain Aresult as

Fα(e)e = [M −N −N ln (|Bgb|)] cos (ωα) + CeN

(Bgb −B0

)· sα

with cos (ωα) =Bgb · sα|Bgb| . (4.24)

where ωα is the angle between the net defect density Bgb and the slip direction sα.

Furthermore, Fα(e)e = Fα(s)

e and for grain B, substitute β for α.

In the case of a zero net defect, i.e. Bgb = B0 = 0 (a perfect crystalline state), thegrain boundary energy, Eq. (4.23), is zero, and the energetic microforce, Eq. (4.24), isundefined. However, the presented continuum model does not entail small or zero valuesof the net defect, due to the discrete nature of the problem. It is not possible to haveinfinitely small continuous variations of the net defect around zero, because at least oneBurgers vector in a finite boundary period is required to obtain the smallest possiblemisorientation.

4.4 Analysis of initial GB energetics

First, the initial energetic interface microforces are analysed analytically. In the initialstate, the interface normal slip measures q equal zero and the relations for the net defectBgb (Eq. (4.18)) and grain boundary energy φe (Eq. (4.23)) consistently reduce to theircorresponding initial expressions B0 (Eq. (4.20)) and φe0 (Eq. (4.19)), respectively. The

energetic microforce Fα(e)e (Eq. (4.24)) reduces to

Fα(e)e0 =

[M −N −N ln

( |Bp||p|

)]cos (ωα0 ) with cos (ωα0 ) =

Bp · sα|Bp|

, (4.25)

where ωα0 is the angle between the intrinsic net defect Bp and the slip direction sα.

The resulting initial energetic interface microforce Fα(e)e0 as a function of the intrinsic

net defect density |Bp|/|p| and angle ωα0 is shown in Fig. 4.2, for the case of symmetrictilt grain boundaries with the 〈100〉 rotation axis in Cu. For this rotation axis, the

Page 83: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

70 4 Grain boundary plasticity incorporating internal structure and energy

00.2

0.40.6

0.81

−180−90

090

180−0.006

−0.004

−0.002

0

0.002

0.004

0.006

Intrinsic net defect density (−)high

Cu <100> [0°,90°]

low

Angle ωα

0 (degrees)

Initia

l energ

etic m

icro

forc

e (

N/m

m)

Figure 4.2: The initial energetic interface microforce as a function of the intrinsic net defect density|Bp|/|p| and the angle ωα0 between the initial net defect and the slip direction for symmetric tilt bound-aries with the 〈100〉 rotation axis in Cu. The grey line indicates regions corresponding to low- andhigh-angle grain boundaries.

misorientation angle range is θ ∈ [0◦, 90◦], in which the end points, θ = 0◦ and θ = 90◦,are perfect crystals. From Chapter 3, the resulting energetic interface parameters equalM = 901 mJ/m2 and N = 1329 mJ/m2. Note that similar surface profiles result forthe 〈100〉, 〈110〉 and 〈111〉 rotation axes in Al and Cu. Depending on the sign of the

bulk-induced interface microforce, Fα(e)bk , positive and negative values of the energetic

interface microforce correspond to interface normal slip being retarded or promoted ona slip system, see Eq. (4.14).

As apparent from Fig. 4.2, the low-angle grain boundaries, characterised by a low intrinsicnet defect density, have larger (absolute) values of the energetic microforce compared tothe high-angle grain boundaries. This is consistent with the normalised shear stress fieldsof arrays of edge dislocations with two different spacings D, i.e. D = 30b and D = 10b,illustrated in Fig. 4.3. The stress fields of the dislocations in the array with smallerdislocation spacing, i.e. higher defect density in high-angle grain boundaries, decay overa shorter distance than in the array with larger spacing, i.e. a lower defect density.

Two configurations exist in which the energetic microforce equals zero, as can be identifiedfrom Eq. (4.25). The first configuration occurs when the angle ωα0 = ±90◦, correspond-ing to the intrinsic net defect Bp and the slip direction sα being orthogonal. For smalldeviations around these angles, the magnitude of the energetic microforce increases (pos-itive or negative) due to the better alignment of the interface defect and slip direction.Note that for symmetric tilt boundaries, where the median lattice in Frank’s formula ishere used to determine the net defect (see Chapter 3), the interface net defect is parallel

Page 84: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.4 Analysis of initial GB energetics 71

Normalised x/b distance

Norm

alis

ed y

/b d

ista

nce

−30 −20 −10 0 10 20 30

−30

−20

−10

0

10

20

30

Norm

alis

ed s

hear

str

ess σ

xy /S

−0.5

−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

0.5

(a)

Normalised x/b distance

Norm

alis

ed y

/b d

ista

nce

−30 −20 −10 0 10 20 30

−30

−20

−10

0

10

20

30

Norm

alis

ed s

hear

str

ess σ

xy /S

−0.5

−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

0.5

(b)

Figure 4.3: The normalised shear stress profiles of dislocation walls with different dislocation spacingsD, (a) D = 30b and (b) D = 10b. The parameter S = Gb

2(1−ν)D. The stress fields decay over a shorter

distance in the case of higher-angle boundaries, which have a smaller dislocation spacing.

to the grain boundary plane normal [82]. Hence, the orthogonality of the interface netdefect and slip direction is in this case equivalent to the slip plane α being parallel to theboundary plane. The second configuration that gives a zero energetic microforce appearswhen the term within brackets in Eq. (4.25) equals zero, i.e.

|Bp||p| = exp

(M −NN

). (4.26)

Calculating this value of the intrinsic net defect density yields 0.72, which correspondsto 2813b per µm length of grain boundary.

Finally, for low-angle boundaries (i.e. small intrinsic net defect density), considerablylarger fluctuations per slip direction (via angle ωα0 ) occur compared to high-angle bound-aries (i.e. large intrinsic net defect density). Due to the smaller screening of the stressfields in the case of low-angle grain boundaries, a variation of the slip direction inducesa larger variation of the energetic microforce and, hence, has an increased effect on theinterface normal slip. In the case of high-angle grain boundaries these differences are lesspronounced.

Page 85: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

72 4 Grain boundary plasticity incorporating internal structure and energy

U U

d

h

2d

grain Agrain B grain B=[010]e2

=[100]e1

Figure 4.4: Geometry and boundary conditions for simple shear deformation of a periodic bicrystal.

4.5 Numerical analysis of GB plasticity

The incorporation of the multiscale atomistic-to-continuum approach of the initial grainboundary structure and energy into the grain boundary extended crystal plasticity model,as described above, is now implemented numerically in order to examine the particularrole and influence of the energetic contributions on the macroscopic response for incip-ient plastic deformation of grain boundaries. This is illustrated for a periodic bicrystalloaded in plane strain in the case of 〈100〉 symmetric tilt boundaries throughout themisorientation range including both low- and high-angle grain boundaries.

4.5.1 Model definition and finite element implementation

The geometry and boundary conditions of the periodic bicrystal loaded in plane strainare shown in Fig. 4.4. It consists of alternating columnar grains, designated A (in blue)and B (in red), separated by grain boundaries. Each grain has width d in the e1-directionand is of infinite height in the e2-direction. A repeating unit cell of width 2d and height his modelled, with periodic boundary conditions applied to the top-bottom and left-rightboundaries. Simple shear deformation is simulated by applying a prescribed displacementU = γht on the top-left node. A constant shear rate of γ = 10−3 s−1 is applied, up toa maximum of 1% macroscopic shear. A bicrystal with a grain size d = 1 µm will beconsidered and only edge dislocations are taken into account in this planar configuration.

The response of 8 symmetric tilt boundaries in the complete misorientation range withthe [001] rotation axis in FCC copper is investigated. Table 4.1 shows the characteristicsof the studied grain boundaries, i.e. the misorientation angle θ, the grain boundary plane,the Σ value, the boundary period vector p, the intrinsic net defect Bp in a boundaryperiod and the intrinsic net defect density |Bp|/|p|; the latter three parameters havebeen computed using the atomistic-to-continuum approach presented in Chapter 3 assummarised in Section 4.3.2. Adopting the commonly used demarcation of about 15◦

with respect to the perfect crystal misorientation angles (θ = 0◦ and θ = 90◦), grain

Page 86: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.5 Numerical analysis of GB plasticity 73

Table 4.1: Symmetric tilt grain boundaries with the [001] rotation axis in Cu analysed here. Given arethe misorientation angle θ (in degrees), the grain boundary plane, the sigma value Σ characterizing theCSL, the boundary period vector p, the intrinsic net defect Bp in a boundary period and the intrinsicnet defect density |Bp|/|p|; a is the lattice parameter.

GB θ GB plane Σ p Bp |Bp|/|p|1 5.72 (20 1 0) 401 a[1 20 0] 2a[1 0 0] 0.102 14.25 (8 1 0) 65 a[1 8 0] 2a[1 0 0] 0.253 28.07 (4 1 0) 17 a[1 4 0] 2a[1 0 0] 0.494 43.60 (5 2 0) 29 a[2 5 0] 6a

2 [1 1 0] 0.795 53.13 (2 1 0) 5 a[1 2 0] 2a

2 [1 1 0] 0.636 61.93 (5 3 0) 17 a

2 [3 5 0] 2a2 [1 1 0] 0.49

7 73.74 (4 3 0) 25 a[3 4 0] 2a2 [1 1 0] 0.28

8 83.97 (10 9 0) 181 a[9 10 0] 2a2 [1 1 0] 0.11

boundaries 1, 2, 7 and 8 can be classified as low-angle, while the remaining boundarieshave a high-angle character. The orientations of the grains in the bicrystal are emphasisedin Fig. 4.5. Starting from the reference state, corresponding to the fixed coordinatesystem, the grains are first rotated (vector transformation) 180◦ counterclockwise tothe state where no misorientation between the adjacent grains is present, i.e. a perfectcrystal. A particular grain boundary is recovered by applying a rotation of θ/2 in oppositedirections for grains A and B, respectively. The orientations of the slip plane normalsmα and slip directions sα of the 12 FCC slip systems α are given in Table 4.2.

The grain boundary extended strain gradient crystal plasticity framework consisting of aset of highly nonlinear coupled equations along with the boundary conditions are solvedin a finite element setting (see Chapter 2). The periodic bicrystal is discretised using16 elements per grain (32 elements in total) in the e1-direction. In a mesh sensitivityanalysis, this has shown to provide sufficient resolution. In the e2-direction a single ele-ment is adequate due to the periodicity, which has been verified by separate calculations.The initial GND densities are assumed to be zero. The bulk material parameters usedin the calculations are representative of copper and are similar to those of Evers et al.[57] and Bayley et al. [16]. The values are listed in Table 4.3. The interface parameters

[100]

[110]12

[110]12

[010]

[100]B

[110]B

12

[110]B

12

[010]B

[100]A

[110]A

12

[110]A

12

[010]A

[100]B

[110]B

12

[110]B

12

[010]B

[100]A

[110]A

12

[110]A

12

[010]A

perfect state ( 0°)θ= misoriented state ( 0°)θ>reference state

θ2

θ2

Figure 4.5: Orientations of the grains in the bicrystal starting from the reference state (left) via theperfect state (middle) to the misoriented state (right).

Page 87: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

74 4 Grain boundary plasticity incorporating internal structure and energy

Table 4.2: Slip systems (α) with their corresponding Miller indices of the normal (mα) and slip direction(sα) unit vectors for an FCC crystal in the reference state.

α 1 2 3 4 5 6

sα [110] [101] [011] [110] [101] [011]mα (111) (111) (111) (111) (111) (111)

α 7 8 9 10 11 12

sα [110] [101] [011] [110] [101] [011]mα (111) (111) (111) (111) (111) (111)

are listed in Table 4.4. The interface reference slip rate q0 and rate sensitivity exponentn are taken to be the same as their bulk counterparts γ0 and m. The energetic interfaceparameters M and N have been computed in Chapter 3.

In the following subsections, the behaviour of the grain boundary model will be referredto as ‘intermediate’ and compared to the two limiting cases of microhard (impenetrableto dislocations) and microfree (an infinite dislocation sink) grain boundaries. The lattertwo are obtained by setting Fe = 0 with R → ∞ and R → 0, respectively. Finally,

Table 4.3: Bulk material parameters representative of copper.

Parameter Description Magnitude

E Young’s modulus 144 GPaG Shear modulus 54.1 GPaν Poisson’s ratio 0.33b Burgers vector length 0.256 nmγ0 Reference slip rate 0.001 s−1

a0 Interaction coefficient 0.06a1/a0 Interaction coefficient 5.7a2/a0 Interaction coefficient 10.2a3/a0 Interaction coefficient 16.6h0 Immobilisation coefficient 0.2h1 Immobilisation coefficient 0.3h2 Immobilisation coefficient 0.4h3 Immobilisation coefficient 1.0R GND evaluation radius 500 nmm Rate sensitivity exponent 0.1K Dislocation segment length constant 26yc Critical annihilation length 1.6 nmρs0 Initial SSD density 7.0 µm−2

Page 88: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.5 Numerical analysis of GB plasticity 75

Table 4.4: Interface parameters used in the numerical simulations.

Parameter Description Magnitude

q0 Interface reference slip rate 0.001 s−1

n Interface rate sensitivity exponent 0.1Ce Interface energy constant 105

R Interface normal slip resistance 0.005 N mm−1

M Interface energy constant 901 mJ m−2

N Interface energy constant 1329 mJ m−2

‖(·)α‖ is introduced to obtain an effective measure and denotes the Euclidean norm ofthe contributions of all slip systems α = 1, . . . , 12. Note that the effective measures maysometimes mask the behaviour of individual slip systems. However, the main featuresare usually captured well by the effective measures.

4.5.2 Results

For each grain boundary studied in this chapter, the initial value of the energetic mi-croforce for each slip system is calculated using Eq. (4.25), as shown in Fig. 4.2. Theeffective measure of the energetic microforce as a function of misorientation angle isshown in Fig. 4.6. The resulting curve in Fig. 4.6 is consistent with Fig. 4.2, wherelow- and high-angle grain boundaries have relatively large and small magnitudes of theenergetic microforce, respectively, as discussed in Section 4.4.

0 10 20 30 40 50 60 70 80 900

0.001

0.002

0.003

0.004

0.005

0.006

Initia

l e

ne

rge

tic m

icro

forc

e ||F

α E|| (

N/m

m)

Misorientation angle (degrees)

Figure 4.6: The effective measure of the initial energetic interface microforce for the 8 symmetric tiltgrain boundaries.

Page 89: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

76 4 Grain boundary plasticity incorporating internal structure and energy

0 0.002 0.004 0.006 0.008 0.010

50

100

150

200

250

300

350

She

ar

str

ess (

MP

a)

Total shear (−)

θ=14.25°

microhard GB

microfree GBintermediate GB (F

E=0)

intermediate GB

(a)

0 0.002 0.004 0.006 0.008 0.010

50

100

150

200

250

300

350

She

ar

str

ess (

MP

a)

Total shear (−)

θ=28.07°

microhard GB

microfree GBintermediate GB (F

E=0)

intermediate GB

(b)

0 0.002 0.004 0.006 0.008 0.010

50

100

150

200

250

300

350

She

ar

str

ess (

MP

a)

Total shear (−)

θ=53.13°

microhard GB

microfree GBintermediate GB (F

E=0)

intermediate GB

(c)

0 0.002 0.004 0.006 0.008 0.010

50

100

150

200

250

300

350

She

ar

str

ess (

MP

a)

Total shear (−)

θ=73.74°

microhard GB

microfree GBintermediate GB (F

E=0)

intermediate GB

(d)

Figure 4.7: The macroscopic stress-strain responses for the intermediate boundary behaviour and thelimiting microhard and microfree grain boundaries for four grain misorientations.

The effect of the initial energetic microforce and its evolution on the macroscopic stress-strain response is shown in Figs. 4.7 and 4.8. First, the contribution of the bulk behaviouris assessed. Fig. 4.7 shows the macroscopic response in the case of four symmetric tiltboundaries. In order to clarify the effect, also simulations with zero energetic microforceare included, i.e. the energetic microforce is neglected in the entire simulation. Further-more, the limiting situations of the microhard and microfree grain boundary conditionsare shown for each case. For increasing misorientation angle θ, the onset of plastic de-formation occurs at a lower applied shear stress and upon continued deformation a lowerhardening rate results. This is due to the favourable orientation of the slip systems withrespect to the applied load. For the microfree limit, dislocations can freely leave the

Page 90: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.5 Numerical analysis of GB plasticity 77

system at a grain boundary, leading to a constant slip profile across the bicrystal. Hence,no dislocation pile-ups occur at the grain boundary. On the other hand, significantdislocation pile-ups arise in the case of the microhard limit.

Next, the focus is on the effect of the intermediate boundary behaviour on the bicrystalresponse. When the energetic microforce is neglected, the shear stress initially followsthe microhard behaviour as the deformation increases and dislocations pile up at theboundary. At a critical value, determined by the magnitude of the interface normalslip resistance, the interface yields and interface normal slip starts. The bulk-inducedmicroforce has to maintain a constant level, dictated by the interface resistance value(taken constant in these simulations). This means that the rate at which GNDs join thepile-up balances the rate at which they are evacuated by interface normal slip. Consid-ering the intermediate behaviour with the energetic microforce evolving from its initialvalue to a steadily growing value, it is seen that the microhard behaviour is tracked upto a certain point, after which the bicrystal response depends on the evolution of theenergetic microforce on the individual slip systems. This evolution influences the bulk-induced microforces and hence the dislocation pile-up at the grain boundary evolves aswell. This in turn affects the GND density distribution in the bulk and consequentlythe overall strain hardening behaviour. For example, compared to the zero energeticmicroforce situation, a stronger response may occur, Figs. 4.7(a) and (d), or a mixedresponse, Fig. 4.7(b), or a weaker response, Fig. 4.7(c). The overall behaviour clearlydepends on the misorientation.

The influence of the energetic microforce on the bicrystal response is further elucidatedthrough the shear stress difference, which is defined as the difference between the micro-hard and intermediate grain boundary response, in Fig. 4.8. The black and green markersindicate the yield points of the bulk and the grain boundary, respectively. The former isdetermined using the stress difference between the elastic and microhard grain boundaryresponse, whereas the determination of the latter employs the stress difference definedabove. Since a rate-dependent formulation is used for both the bulk and interface normalslip, a threshold value of 0.01 MPa has been used to define both bulk and grain boundaryyield points. Other values for this threshold do not affect the conclusions qualitatively.The yield point shear difference, which is defined as the difference in applied shear atwhich the bulk and the grain boundary yield, is shown for each misorientation in theinset.

Fig. 4.8 reveals that for increasing misorientation angle, the bulk yields at a smallerapplied shear, consistent with Fig. 4.7. Moreover, the difference between the yield pointsof the bulk and the grain boundary are lower for the low-angle boundaries than for thehigh-angle boundaries. This corresponds to the initial energetic microforce profile inFig. 4.6. Low-angle boundaries have larger initial energetic microforces causing interfacenormal slip to occur earlier. Finally, it is observed that for high-angle boundaries theshear stress difference reaches higher values than for low-angle boundaries. However, thisdoes not have to be the case in general.

Page 91: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

78 4 Grain boundary plasticity incorporating internal structure and energy

0

0.002

0.004

0.006

0.008

0.01

01020304050607080900

10

20

30

40

50

60

Total shear (

−)

Misorientation angle (degrees)

Shear

str

ess d

iffe

rence (

MP

a)

0 10 20 30 40 50 60 70 80 900.5

1

1.5

2

2.5x 10

−3

Misorientation angle (degrees)

Yie

ld p

oin

t sh

ea

r d

iffe

ren

ce

(−

)

Figure 4.8: The effect of the energetic interface microforce at the macro-scale for the eight symmetrictilt grain boundaries. The shear stress difference is the difference between the hard and intermediategrain boundary behaviour. The black and green markers indicate the yield points of the bulk and thegrain boundary, respectively. The yield point shear difference in the inset is the difference in appliedshear at which the bulk and the grain boundary yield.

4.6 Discussion

The energetic interface microforce can be interpreted as the effect of the elastic interactionof the grain boundary with crystal dislocations, comparable to the back stress in thebulk. However, no systematic studies of the elastic interaction due to the stress fieldof the grain boundary as a function of misorientation and slip system orientation fromeither experiments or atomistic simulations were found in the literature to compare with.Appealing atomistic simulations in this direction have been presented by Pestman et al.[138] and Saraev and Schmauder [157], where a bicrystal system containing a singledislocation in the vicinity of the grain boundary is relaxed in the absence of an appliedexternal load. However, care should be taken in the determination of the origin of theinteraction effect, because additional contributions due to image forces arising from thepresence of free surfaces and elastic anisotropy across the grain boundary also play a role[145, 175].

Qualitative agreement seems to exist with the results of Sangid et al. [155], who deter-mined energy barriers for the transmission and nucleation of a single dislocation at agrain boundary as a function of the initial grain boundary energy in nickel. For high ini-tial boundary energies, the energy barriers decrease following a power-law fit. This wouldcorrespond to the high-angle grain boundaries in this work, see Figs. 4.1 and 4.2, which

Page 92: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.6 Discussion 79

have relatively large intrinsic net defect densities and corresponding boundary energies,and relatively low initial energetic microforces. However, it should be pointed out thatthe data of Sangid et al. [155] considers only a few grain boundaries (both twist and tilt)for different rotation axes. The initial grain boundary energy levels may deviate consid-erably for different rotation axes. Hence, this may lead to misleading conclusions whencompared to the results for the full misorientation range of only one rotation axis. Fur-thermore, in the case of transmission, the dislocation interaction with the grain boundaryalready initiates at a much lower energy. It is not trivial how to incorporate data obtainedfrom atomistic simulations, such as the energy barriers reported by Sangid et al. [155],into continuum level modelling of grain boundary plasticity. Moreover, it is recalled thatat the continuum scale, where the proposed model applies, the individual interactionsbetween a dislocation and a grain boundary cannot be resolved, i.e. no distinction can bemade between the energy levels required for individual phenomena such as absorption,nucleation and transmission as observed in atomistic simulations.

In the current work, the interface normal slip resistance R is taken constant in all sim-ulations, but in general R is expected to also depend on the misorientation betweenthe grains and the orientation of the grain boundary. Only few experimental and atom-istic studies exist that (systematically) investigate the influence of misorientation onthe interaction (mostly transmission) of dislocations with grain boundaries. Based onexperimental observation (optical and scanning electron microscopy), Zhang and Wang[217] and Zhang et al. [218] ascertained that persistent slip bands can transmit throughlow-angle grain boundaries, whereas they are blocked at general high-angle boundaries.Indentation studies [23, 93, 199, 204] have shown that high-angle boundaries pose largerresistance to transmission than low-angle boundaries. In particular, Kobayashi et al. [93]found that the degree of grain boundary resistance to dislocation interaction depends onthe orientation of the grain boundary rotation axis. Also from atomistic simulations itcan be inferred that the resistance diminishes in the case of low-angle grain boundaries[9, 44, 216]. Low-angle grain boundaries may act as strong obstacles to dislocation pen-etration depending on the interaction strength between the incident and grain boundarydislocations [107, 108]. In the present continuum model, the increase in resistance forlarger misorientations would be modelled through the dissipative interface normal slipresistance R. A simplified expression capturing the essential features observed is givenby:

Rα(e) = Cd|Bgb| and Rα(e)0 = Cd

|Bp||p| , (4.27)

where Cd is a positive constant for a given misorientation axis, analogous to the ener-getic interface parameters M and N . In combination with Fig. 4.1(a), it can be seenthat this equation yields a larger initial resistance for increasing misorientation angle.Eq. (4.27) can be extended to incorporate a dependence on the slip system orientationas well. Nano-indentation experiments could provide insight into the magnitude of theinterface normal slip resistance [1, 195]. Moreover, atomistic simulations [155] as well asthe quasi-continuum [214] and coupled atomistic discrete dislocation [47] methods mayfacilitate the quantification of this resistance. It should be emphasised that in modellingat the (polycrystalline) continuum level care should be taken to avoid double countingof mechanisms in the bulk and interface descriptions, e.g. through lumping of bulk be-

Page 93: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

80 4 Grain boundary plasticity incorporating internal structure and energy

haviour near the boundary into the interface model. A further analysis of the interfacenormal slip resistance, however, is not pursued in the present work.

The expression for the evolution of the grain boundary energy, Eq. (4.23), is a simplegeneralisation of the initial grain boundary energy. The grain boundary energy consis-tently reduces to its initial value for zero plastic deformation at the interface, i.e. zerointerface normal slip, and the initial energetic microforce recovers the physically basedconfiguration of the stress fields of dislocation arrays. In this chapter it is assumed thatthe grain boundary defect remains local. However, grain boundary relaxation throughthe redistribution of defects within the grain boundary can occur. This is the scope ofChapter 5.

When the grain size decreases, a transition occurs from dislocation mediated plasticityin the bulk of the crystals to grain boundary controlled plasticity [42, 72, 191, 209, 219].The effect of grain size d on the strength of metallic materials can roughly be split intofour regimes, characterized by different corresponding deformation mechanisms, althoughsome specific mechanisms are still controversial [37, 94, 121, 220]: (i) d & 1 µm, (ii) 1 µm& d & 100 nm, (iii) 100 nm & d & 20 nm and (iv) d . 20 nm. Grain-size hardening occursin regimes (i)-(iii), whereas softening occurs in regime (iv). The deformation structure inthe first regime consists of dislocation cells where dislocations intersect, multiply and pileup; in the second regime, the dislocations are mostly restricted to their slip planes withfew dislocation sources, immobilisation may exist due to reactions and dislocation pile-upat boundaries; in the third regime, the bulk crystals are perfect and a change occurs in theemission and annihilation of full dislocations at grain boundaries to partial dislocationemission and deformation twinning; in the fourth regime, computer simulations haveindicated that dislocations are absent and that deformation occurs by the shearing ofatoms in the grain boundary. In view of this classification, the proposed model frameworkis applicable in regimes (i) and (ii) only. The nanocrystalline regime (d . 100 nm)cannot be described properly in a continuum framework. However, existing examplesof continuum modelling approaches in this regime include modified conventional crystalplasticity [202, 203] and two-phase (grain interior and grain boundaries) composites [29].

Finally, as mentioned in Section 4.3.3, the challenging task of tracking the evolution ofboundary structure and energy is imperative for accurately modelling grain boundarybehaviour upon plastic deformation, including grain sizes beyond the nanocrystallineregime for which small-scale simulation techniques, such as molecular dynamics simula-tions, are hardly feasible considering the time and length scales to be tackled. Multiscalesimulation techniques, such as the coupled atomistic discrete dislocation method [163],are promising. However, due to the unique character of individual interactions betweengrain boundaries and dislocations, detailed studies of many grain boundaries are required,leading to a massive amount of data that may be difficult to interpret towards the de-formation in polycrystalline aggregates [46]. Nevertheless, despite this unique character,generic continuum ‘building blocks’ can be identified that are representative of differentcollective interactions of dislocations with a grain boundary having a similar macroscopiceffect.

Page 94: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

4.7 Conclusions 81

4.7 Conclusions

Grain boundaries play a predominant role in the deformation behaviour of polycrystallinemetals. Grain boundaries become particularly important when the volume fraction ofinterface regions increases. At present, (polycrystalline) continuum modelling incorpo-rating the underlying initial and evolving structure and energy of grain boundaries is stillin its infancy. The main contribution of this chapter and conclusions are:

1. The multiscale atomistic-to-continuum approach of the initial structure and energyof grain boundaries of Chapter 3 has been incorporated into a grain boundaryextended crystal plasticity framework in order to examine the particular role andinfluence of the energetic contributions on the macroscopic response due to theplastic deformation of grain boundaries.

2. A generalised grain boundary energy as a function of the evolution of an interfacenet defect, representing the grain boundary structure upon plastic deformationat the interface, has been proposed and used to determine an expression of theenergetic interface microforce.

3. An analytical study of the developed initial energetic interface microforce revealeda physically sound behaviour as to be expected for low- and high-angle symmetricaltilt grain boundaries.

4. The complete framework has been implemented numerically and applied for theinvestigation of a periodic copper bicrystal subjected to simple shear deformation.Eight low- and high-angle symmetric tilt boundaries with the [001] rotation axisthroughout the misorientation range have been analysed.

5. It has been observed that the effect of the energetic interface microforce on themacroscopic response depends strongly on the misorientation between adjacentgrains.

The resulting initial energetic interface microforce profiles can be verified using atomisticsimulations specifically focussing on the elastic interactions between grain boundariesand dislocations in the absence of external loads in future work. However, qualitativeagreement with the available atomistic data has already been observed.

Future work should focus on the effect of the evolution of the energetic and dissipativeinterface contributions on the macroscopic response. A significant amount of work inthis direction will be required, which is impossible without multiscale simulation tech-niques, such as the quasi-continuum or coupled atomistic discrete dislocation methods.Furthermore, this chapter assumes the locality of the grain boundary defect. The nextchapter will incorporate the effect of defect redistribution along the grain boundary.

Page 95: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary
Page 96: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Chapter 5

Defect redistribution in a continuumgrain boundary plasticity model1

Abstract

The mechanical response of polycrystalline metals is significantly affected by the be-haviour of grain boundaries, in particular when these interfaces constitute a relativelylarge fraction of the material volume. One of the current challenges in the modelling ofgrain boundaries at a continuum (polycrystalline) scale is the incorporation of the manydifferent interaction mechanisms between dislocations and grain boundaries, as identifiedfrom fine-scale experiments and simulations. In this chapter, the objective is to developa model that accounts for the redistribution of the defects along the grain boundary inthe context of gradient crystal plasticity. The proposed model incorporates the nonlocalrelaxation of the grain boundary net defect density. A numerical study on a bicrystalspecimen in simple shear is carried out, showing that the spreading of the defect contenthas a clear influence on the macroscopic response, as well as on the microscopic fields.This work provides a basis that enables a more thorough analysis of the plasticity ofpolycrystalline metals at the continuum level, where the plasticity at grain boundariesmatters.

1Reproduced from [190]: P.R.M. van Beers, V.G. Kouznetsova, and M.G.D. Geers. Defect redistri-bution within a continuum grain boundary plasticity model. Submitted.

83

Page 97: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

84 5 Defect redistribution in a continuum grain boundary plasticity model

5.1 Introduction

Based on the knowledge acquired in the last few decades on the relation between materialproperties and the underlying microstructure, it has become possible to process metalssuch that specific functionality can be obtained. The understanding of this interactionbetween structure and property expands even further as computational and experimentalresources increase, enabling studies at different time and length scales. An example isthe mechanical behaviour of metallic materials. In particular, topics of interest relevantto this work include (i) the miniaturisation in industrial applications involving metalswith dimensions approaching microstructural length scales, such as the grain size, and(ii) the ultra-high strength of fine-grained metals. In both cases the behaviour of grainboundaries (GBs) is essential, since they constitute a relatively large fraction of thematerial volume. Accordingly, one of the current challenges in modelling plasticity ofpolycrystalline metals is to account more accurately for the role of grain boundaries[116, 117].

Different interaction mechanisms between dislocations and grain boundaries have beenidentified from experiments and simulations, see [60, 145, 175, 186] and references therein.In general, the details of the interaction mechanisms that take place depend on theinitial structure of the grain boundary, the magnitude of the applied forces, and thedegree of thermal activation. Grain boundaries can act as obstacles to the motion oflattice dislocations that accumulate into a dislocation pile-up, or as sources of dislocationsgenerating dislocations in the adjacent grains and at the interface. A lattice dislocationmay also be transmitted through the interface from one grain to the other in a direct orindirect manner, in the latter case leaving behind a residual defect in the grain boundary.

Another mechanism which has a strong effect on the mechanical properties of materials,is the redistribution of the defect content along the grain boundary, as observed in exper-iments [34, 40, 43, 49, 88, 91, 102, 109, 118, 142, 143, 147, 192] and fine-scale simulations[32, 45, 47, 87, 90, 96, 139, 164, 201, 214]. If the grain boundary acts as a sink and alattice dislocation is absorbed into the grain boundary, it can (i) remain localised, or (ii)dissociate and spread along the boundary plane. The spreading observed by transmis-sion electron microscopy (TEM) consists in a widening and subsequent disappearanceof the dislocation diffraction contrast. However, high-resolution TEM has revealed thatnon-negligible decomposition products may exist, which are not resolved by conventionalTEM. Three types of models have been proposed to describe these accommodation phe-nomena [112, 128, 144]: (i) delocalisation models, which assume continuous spreadingof the dislocation cores; (ii) dissociation models, according to which the dislocation de-composes into two or more discrete dislocations that have smaller Burgers vectors; (iii)incorporation models, which consider the accommodation of the dislocation into theintrinsic grain boundary defect structure. The three models yield similar kinetics ofspreading.

The individual interactions discussed above cannot be resolved in continuum (crystal)plasticity frameworks. This demands the development of a new model that describes thedefect redistribution along the grain boundary in an average sense. In recent years, from

Page 98: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.1 Introduction 85

the perspective of continuum scale modelling, first steps have been made to incorporategrain boundary interface mechanics into strain gradient plasticity [41, 64, 114, 141, 195]and strain gradient crystal plasticity [21, 51, 77, 110, 134, 210] frameworks. The latterhas the advantage of naturally incorporating the crystallographic misorientation acrossa grain boundary, which is of particular importance when considering the interactionsbetween dislocations and grain boundaries. However, in these frameworks, constitu-tive equations at the grain boundary, e.g. the continuum scale expression of the grainboundary energy, still lack a solid physical background. At the same time, atomisticsimulations have indicated that the grain boundary energy is dependent on the grainboundary structure [79, 153, 180, 198, 206].

In Chapter 3, based on the output of atomistic simulations, a multiscale approach hasbeen developed that leads to a continuum representation of the initial grain boundarystructure and energy. In this approach, the grain boundary structure is represented by anintrinsic net defect density, which is related to the atomistically calculated grain boundaryenergies through a logarithmic expression. It was shown that the grain boundary energiesin the complete misorientation range of symmetric tilt grain boundaries in aluminiumand copper can be accurately described by this multiscale method. In the previouschapter this atomistic-to-continuum approach was incorporated into the grain boundarycrystal plasticity model developed earlier. The resulting framework was numericallyimplemented in order to examine the particular role and influence of the grain boundaryenergetic contributions on the macroscopic response. It was thereby observed that theeffect of the energetic contribution on the macroscopic response depends strongly on themisorientation between adjacent grains, whereby it was assumed that the grain boundarynet defect remains localised.

The novel contribution in this chapter is the development of a model that incorporatesthe defect redistribution along the grain boundary into the continuum grain boundaryplasticity framework, as described in Chapters 2 and 4. A diffusion type evolution law isproposed for the redistribution of the net defect within the grain boundary, which allowsfor a nonlocal relaxation of the grain boundary net defect. The continuum model consistsof a set of equations, including a balance and constitutive equations, for a vectorial grainboundary net defect field along the grain boundary surface. The incorporation of astrongly nonlocal interaction term leads to an integro-differential equation for the netdefect evolution, which is further reformulated into a differential form. A numericalimplementation of the resulting model framework investigates the effect of the defectredistribution on the macroscopic response in the case of a copper bicrystal loaded insimple shear.

The chapter is organised as follows. Section 5.2 summarises the grain boundary crystalplasticity framework and the incorporation of the internal grain boundary structure andenergy. In Section 5.3, the continuum model of the defect redistribution is developed anddiscussed. Section 5.4 covers the solution procedure and the finite element implemen-tation of the framework. The model problem and the numerical study of the influenceof the defect redistribution on the macroscopic response are described in Section 5.5.Finally, conclusions are provided in Section 5.6.

Page 99: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

86 5 Defect redistribution in a continuum grain boundary plasticity model

5.2 Bulk and interface frameworks

This section summarises the dislocation based strain gradient crystal plasticity frameworkextended with a grain boundary interface model as developed in Chapter 2, followed bya brief description of the incorporation of the grain boundary structure and energy asestablished in Chapters 3 and 4.

5.2.1 Bulk crystal formulation

This work departs from the bulk crystal formulation originally proposed by Evers et al.[56, 57] and Bayley et al. [16]. Attention is restricted to a small deformation framework.The total distortion is decomposed into its elastic (e) and plastic (p) contributions

∇u = He + Hp , (5.1)

where He represents the stretching and rotation of the lattice, while the plastic distortionHp describes the local deformation due to slip given by

Hp =∑α

γαsαmα . (5.2)

Here, γα is the slip on slip system α having slip plane normal mα and slip direction sα.The symmetric part of the elastic distortion He gives the elastic strain Ee

Ee =1

2

(He + HeT

). (5.3)

For a given slip system α the resolved shear stress or Schmid stress τα is given by

τα = σ : sαmα with σ = 4C : Ee . (5.4)

The latter equation is the conventional linear elastic relationship, which relates the stresstensor σ to the elastic strain Ee via the fourth order elasticity tensor 4C.

In the crystal plasticity approach adopted, dislocation densities on individual slip systemsact as key variables. Gradients in slip induce geometrically necessary dislocation (GND)

densities of edge (e) and screw (s) character, denoted ρα(e)g and ρ

α(s)g , respectively

ρα(e)g = −1

b∇γα · sα and ρ

α(s)g =

1

b∇γα · lα , (5.5)

with unit vector lα = sα ×mα and b the Burgers vector magnitude. The evolution ofslip is accomplished through a visco-plastic power-law given by

γα = γ0

( |τα − ταb |sα

) 1m

sign (τα − ταb ) . (5.6)

Page 100: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.2 Bulk and interface frameworks 87

Parameters γ0 and m are positive constants. The slip resistance sα for slip system αreflects the obstruction of dislocation glide through short-range interactions of all dis-locations on coplanar and intersecting slip systems. Both the densities of geometricallynecessary dislocations (GNDs), ρξg, and statistically stored dislocations (SSDs), ρξs , con-tribute to the slip resistance [5]

sα = Gb

√∑ξAαξ

∣∣ρξs ∣∣+∑

ξAαξ

∣∣ρξg∣∣ , (5.7)

where Aαξ represents the full dislocation interaction matrix as given by Franciosi andZaoui [61]. The components Aαξ are the coefficients a0, a1, a2 and a3, which representthe strength of six types of interactions: self hardening (a0), coplanar (a1), cross slip(a1), Hirth lock (a1), glissile junction (a2) and Lomer-Cottrell lock (a3). The generalisedform of Essmann and Mughrabi [55] is employed to describe the evolution of the SSD

densities ρξs

ρξs =1

b

(1

Lξ− 2ycρ

ξs

) ∣∣γξ∣∣ with ρξs (t = 0) = ρξs0 . (5.8)

The dislocation accumulation is governed by the average dislocation segment length

Lξ =K√∑

ζ Hξζ∣∣ρζs ∣∣+

∑ζ H

ξζ∣∣ρζg∣∣ , (5.9)

where Hξζ are the immobilisation coefficients, structured analogously to the interactioncoefficients Aαξ. The dislocation annihilation is controlled by a critical annihilationlength yc.

In Eq. (5.6), the back stress ταb is obtained by the projection of the internal stress σint

onto slip system α

ταb = −σint : sαmα . (5.10)

The internal stress can be interpreted as a fine-scale correction accounting for microstruc-tural dislocation interactions that are not resolved in the long-range elastic strain field,e.g. due to dislocation pile-ups. Spatial gradients in the densities of geometrically nec-essary dislocations induce long- and short-range non-vanishing stress fields. The stressfields of dislocations in an infinite elastically isotropic medium are used in the derivationof this internal stress. Furthermore, a linear spatial distribution of the GND densitieswithin a volume with radius R around a material point is assumed. The length scaleparameter R determines the extent of the nonlocal influence. Accounting for the contri-butions from all slip systems α, the following expression for the internal stress is obtained(see [16] for more details)

σint =GbR2

8 (1− ν)

∑α

∇ρα(e)g · 3Cα(e) +

GbR2

4

∑α

∇ρα(s)g · 3Cα(s) , (5.11)

Page 101: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

88 5 Defect redistribution in a continuum grain boundary plasticity model

where the third order slip system orientation tensors

3Cα(e) = 3mαsαsα + mαmαmα + 4νmαlαlα − sαsαmα − sαmαsα ,

3Cα(s) = −mαlαsα −mαsαlα + lαmαsα + lαsαmα , (5.12)

with G the shear modulus and ν the Poisson’s ratio. Since the focus of this chapter is onthe grain boundary behaviour, a simplified expression is used here, taking into accountonly the ‘self’ internal stress components [16], i.e. only the last two terms in the thirdorder slip system orientation tensors in Eq. (5.12). More details on this gradient crystalplasticity framework can be found in [56, 57] and [16].

5.2.2 Grain boundary interface formulation

The interface modelling framework, developed in Chapter 2, incorporates the possibilityto describe grain boundary behaviour other than the conventional higher order boundaryconditions of either impenetrable or completely transparent grain boundaries. Althoughthe framework allows for the incorporation of macroscopic sliding and opening of theinterface, these mechanisms are not incorporated here, i.e. continuity of displacementsacross the interface is assumed. The different fine-scale phenomena taking place at theboundary, e.g. absorption, emission and transmission, are not resolved separately at thescale of the model, but the net effect of these interactions is lumped into a continuumscale quantity termed ‘interface normal slip’ q. The interface normal slip components ofedge (e) and screw (s) dislocations at the interface Γ between adjacent grains A (slipsystems α) and B (slip systems β) are defined as

grain A: qα(e) = γα sα · ngb , qα(s) = −γα lα · ngb ,

grain B: qβ(e) = −γβ sβ · ngb , qβ(s) = γβ lβ · ngb . (5.13)

The grain boundary normal ngb points from grain A to grain B. The evolution of theinterface normal slip qα on slip system α at the grain A side of Γ is modelled with avisco-plastic power-law, similar to slip in the bulk crystal (the expressions for grain Bare obtained by substituting β for α)

qα(e) = q0

(∣∣Fα(e)bk −Fα(e)

e

∣∣Rα(e)

) 1n

sign(Fα(e)

bk −Fα(e)e

),

qα(s) = q0

(∣∣Fα(s)bk −Fα(s)

e

∣∣Rα(s)

) 1n

sign(Fα(s)

bk −Fα(s)e

). (5.14)

The parameters n and q0 are positive constants (similar to the bulk crystal parameters

m and γ0). The quantities Fα(e)bk and Fα(s)

bk are the bulk-induced interface microforcesand follow from the projections of the slip system microstress ξα

Fα(e)bk = −ξα · sα , Fα(s)

bk = ξα · lα . (5.15)

Page 102: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.2 Bulk and interface frameworks 89

The microstress ξα can be determined from the microstructurally resolved internal stressσint and is given by (see Chapter 2 for the details)

ξα =

[GbR2

8 (1− ν)

∑ζ

ρζ(e)g

3Cζ(e) +GbR2

4

∑ζ

ρζ(s)g

3Cζ(s)

]: sαmα , (5.16)

and similarly for grain B. Quantities Rα(e) and Rα(s) represent the interface normal slipresistances for edge and screw dislocations on slip system α, respectively. The energetic

interface microforces Fα(e)e and Fα(s)

e are determined from the grain boundary energy φe

Fα(e)e =

∂φe

∂qα(e), Fα(s)

e =∂φe

∂qα(s), (5.17)

and can be interpreted as interface ‘back stresses’. The grain boundary energy φe dependson the grain boundary structure and its evolution, as discussed in the next section.

5.2.3 Grain boundary structure and energy

The structure and energy of a grain boundary depends on its initial state and the sub-sequent deformation history of the adjacent grains. In general, when lattice dislocationsinteract with a grain boundary, the structure of the interface and the defect content arealtered. In order to obtain a continuum representation of the grain boundary structure,in Chapters 2 and 4 the following vectorial measure of the grain boundary net defectdensity at a material point on the interface is proposed

Bgb = B0 + B . (5.18)

Here, B0 represents the initial grain boundary defect and corresponds to the contributionof ‘intrinsic’ grain boundary dislocations [13, 81]. These dislocations form the ‘equilib-rium’ structure of the boundary. The vector B represents the defect resulting from theinteraction of bulk lattice dislocations with the grain boundary during plastic defor-mation. These ‘extrinsic’ grain boundary dislocations are not part of the equilibriumstructure of the boundary.

The following form of the extrinsic net defect density B due to the interaction with latticedislocations of edge and screw character from both sides of the grain boundary is defined

B =∑α

(cα(e)qα(e) + cα(s)qα(s)

)sα +

∑β

(cβ(e)qβ(e) + cβ(s)qβ(s)

)sβ , (5.19)

with the coefficients

cα(e) = lα · (ngb × k) = (sα · ngb)(mα · k)− (sα · k)(mα · ngb)

cα(s) = sα · (ngb × k) = (mα · ngb)(lα · k)− (mα · k)(lα · ngb) (5.20)

and similar for cβ(e) and cβ(s) by substituting β for α. Here, k = p/|p| is the unitboundary period vector. Definition (5.19) is a general expression that is consistent with

Page 103: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

90 5 Defect redistribution in a continuum grain boundary plasticity model

an ‘infinitesimal’ Burgers circuit at the grain boundary [77]. In Chapters 2 and 4 asimplified expression for B was used, with coefficients cα equal to 1, which correspondsto a planar configuration.

The grain boundary energy φe is linked to the grain boundary structure and its evolutionvia the relation

φe = φe(Bgb

(qα(e,s), qβ(e,s)

)), (5.21)

which depends on the grain boundary net defect density Bgb that evolves through theinterface normal slip measures q. In order to move towards a physically based expressionfor the grain boundary energy in Eq. (5.21), a continuum representation of the initialboundary structure and energy based on the output of atomistic simulations has beendeveloped in Chapter 3. It incorporates: (i) grain boundary energies and boundaryperiod vectors obtained from atomistic simulations, (ii) structural unit and dislocationdescriptions of the grain boundary structure and (iii) the Frank-Bilby equation to deter-mine the grain boundary dislocation content. In this atomistic-to-continuum approach,the key concept is the attribution of an intrinsic net defect Bp, representing the initialgrain boundary structure, to a boundary period p. A corresponding scalar intrinsic netdefect density |B0|, with B0 = Bp/|p|, is related to the initial grain boundary energyφe0 through a logarithmic relationship valid for the whole misorientation range around acertain rotation axis

φe0 = M |B0| −N |B0| ln(|B0|) , (5.22)

where M and N are two parameters to be identified by fitting atomistic simulationresults. This logarithmic relationship resembles the Read-Shockley equation for low-angle boundaries [150, 151] and can be interpreted as a generalisation to the completemisorientation range, including high-angle grain boundaries. This approach has beenapplied in Chapter 3 to symmetric tilt grain boundaries with the 〈100〉, 〈110〉 and 〈111〉rotation axes in aluminium and copper, thereby obtaining fairly accurate descriptions ofthe continuum scale grain boundary energy.

As indicated in Chapter 4, a reliable physically based description of the evolution of thestructure and energy of a grain boundary upon plastic deformation remains a challengeat present. No systematic (atomistic scale) simulations exist that track the evolutionof the boundary structure and energy as a consequence of the interaction with multipledislocations. Therefore, a straightforward generalisation of the expression of the initialgrain boundary energy (Eq. (5.22)) is here proposed

φe = M |Bgb| −N |Bgb| ln (|Bgb|) + 12CeN |Bgb −B0|2 , (5.23)

where Ce is a positive constant and the net defect density is defined in Eq. (5.18) withEq. (5.19).

More details on the continuum representation of the grain boundary structure and energycan be found in Chapters 3 and 4. So far, only local expressions for the grain boundarynet defect have been considered. The subsequent section will incorporate nonlocal effectsdue to the defect redistribution along the grain boundary.

Page 104: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.3 Defect redistribution 91

5.3 Defect redistribution

The development of a continuum model that takes into account the redistribution of thedefect content along the grain boundary within the context of strain gradient crystalplasticity is elaborated in this section. After detailing the derivation of the model, theresulting model is discussed.

5.3.1 Derivation of the model

The continuum redistribution model derived here consists of a defect balance equationand constitutive equations (cf. [185]). Here, a redistribution of a vectorial quantity ata surface is considered, which differs from the consideration of a scalar field within avolume, as done for scalar diffusion problems.

Balance equation

The grain boundary net defect balance equation reads

d

dt

∫Γ

Bgb dΓ = −∫LnΓ · J dL+

∫Γ

S dΓ , (5.24)

which states that the change of the grain boundary net defect density Bgb defined onthe grain boundary surface Γ is equal to (i) what is lost or gained by flow J through theedge L of the surface (having unit normal nΓ pointing outward, which lies in the GBplane (if planar)) and (ii) what is created or consumed by sources and sinks S at thegrain boundary surface (see Fig. 5.1). The source/sink term S of the net defect densityis due to the in- and outflow of bulk lattice dislocations in the normal direction ngb onthe lateral faces of the grain boundary and reads

S =∑α

(cα(e)qα(e) + cα(s)qα(s)

)sα +

∑β

(cβ(e)qβ(e) + cβ(s)qβ(s)

)sβ , (5.25)

which corresponds to the time derivative of the extrinsic net defect density in Eq. (5.19).The surface divergence theorem will be employed and reads [22, 169]∫

Γ

∇s · a dΓ =

∫Γ

(∇s · ngb)ngb · a dΓ +

∫LnΓ · a dL , (5.26)

where ∇s is the surface gradient ∇s = ∇ · (I−ngbngb), a is an arbitrary vector or tensorand ∇s · ngb is a measure of the curvature of the grain boundary surface. Applicationof the surface divergence theorem, assuming a flat grain boundary, leads to the followinglocal balance equation for the evolution of the grain boundary net defect density Bgb ata material point x at the grain boundary surface Γ

dBgb

dt= −∇s · J + S . (5.27)

Page 105: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

92 5 Defect redistribution in a continuum grain boundary plasticity model

nGB

grain A

grain B

¡

L

Figure 5.1: Schematic of the grain boundary surface Γ having unit normal ngb pointing from grain Ato grain B on the lateral face. The edge L of the grain boundary surface has unit normal nΓ pointingoutward, which lies in the grain boundary plane (if planar).

Constitutive equations

As a constitutive assumption, the flow J of the grain boundary net defect density isrelated to a driving force given by the surface gradient of a ‘chemical’ potential µ

J = − 4η : ∇s µ . (5.28)

Here, 4η is a fourth order mobility tensor, which is assumed to be isotropic for simplicity,i.e. 4η = η4I, with the fourth order unit tensor 4I = δilδjkeiejekel, and η a positivemobility parameter. Substituting Eq. (5.28) into the net defect density evolution equation(5.27) gives

dBgb

dt= η∆s µ + S , (5.29)

with the surface Laplacian operator ∆s = ∇s ·∇s. Note that no change in the grainboundary net defect density occurs due to redistribution, if the chemical potential ishomogeneous along the grain boundary.

The chemical potential µ is obtained by taking the variation of the grain boundary energydensity ψe with respect to Bgb at the point x

δψe =∂ψe

∂Bgb(x)· δBgb(x) with µ =

∂ψe

∂Bgb(x). (5.30)

The following two contributions to the grain boundary energy density ψe at a grainboundary material point x as a function of the grain boundary net defect density Bgb

are considered

ψe(Bgb(x),Bgb(y)) = ψel (Bgb(x)) + ψe

nl(Bgb(x),Bgb(y)) . (5.31)

Page 106: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.3 Defect redistribution 93

Here, the local energy density ψel is given by Eq. (5.23), repeated here for completeness

ψel (Bgb(x)) = M |Bgb(x)|−N |Bgb(x)| ln (|Bgb(x)|)+

1

2CeN |Bgb(x)−B0(x)|2 . (5.32)

The nonlocal energy density ψenl, which accounts for the effect of the net defect density

Bgb(y) at grain boundary material points y on the grain boundary energy density at apoint x, is inspired by [14, 69, 185] for two-phase systems, and given as

ψenl(B

gb(x),Bgb(y)) =C

4

∫Γ

W (x,y)|Bgb(x)−Bgb(y)|2 dΓ . (5.33)

where W (x,y) is a weight function that weighs the contribution of the grain boundarynet defect density Bgb(y) and C is a positive constant. Calculating the chemical potentialµ (Eq. (5.30)) gives

µ =∂ψe

l (Bgb(x))

∂Bgb(x)+∂ψe

nl(Bgb(x),Bgb(y))

∂Bgb(x), (5.34)

with the derivatives of the local and nonlocal energy densities reading

∂ψel (Bgb(x))

∂Bgb(x)=[M −N −N ln(|Bgb(x)|)

] Bgb(x)

|Bgb(x)| +CeN(Bgb(x)−B0(x)) , (5.35)

∂ψenl(B

gb(x),Bgb(y))

∂Bgb(x)=C

2

∫Γ

W (x,y)(Bgb(x)−Bgb(y)) dΓ . (5.36)

This latter expression can be rewritten into (see [185] for another nonlocal diffusionproblem)

∂ψenl(B

gb(x),Bgb(y))

∂Bgb(x)= λ

(Bgb(x)−B

gb(x))

, (5.37)

where the nonlocal grain boundary net defect density Bgb

(x), a line interface coefficientλ and a scaled weight function W ∗(x,y) are defined as

Bgb

(x) =

∫Γ

W ∗(x,y)Bgb(y) dΓ , λ =C

2

∫Γ

W (x,y) dΓ and W ∗(x,y) =W (x,y)

2λ.

(5.38)

The integral formulation, Eq. (5.38), can be reformulated into a differential form [53,135, 185] using a Taylor series expansion around material point x at the grain boundaryplane truncated after the quadratic term

Bgb(y) = Bgb(x) + (y− x) ·∇sBgb(x) +

1

2(y− x)(y− x) : ∇s∇sB

gb(x) . (5.39)

Substituting this expansion into the integral definition of the nonlocal grain boundarynet defect density (Eq. (5.38)) gives

Bgb

(x) =

∫Γ

W ∗(x,y)Bgb(x) dΓ +

∫Γ

W ∗(x,y)(y− x) ·∇sBgb(x) dΓ+

1

2

∫Γ

W ∗(x,y)(y− x)(y− x) : ∇s∇sBgb(x) dΓ . (5.40)

Page 107: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

94 5 Defect redistribution in a continuum grain boundary plasticity model

An isotropic weight function is assumed, i.e. it depends only on the distance ρ = |x− y|between the points y and x, for which a Gaussian function is taken [15]

W (ρ) =1

2πL2exp

(− ρ2

2L2

). (5.41)

Here, L is the length scale that determines the extent of the nonlocal spatial interactionsin the grain boundary. Note that for an infinite boundary plane∫

Γ

W (ρ) dΓ = 1 . (5.42)

Elaboration of the integral expression (Eq. (5.40)) for an infinite grain boundary planegives

Bgb

(x) = Bgb(x) + c∆sBgb(x) , (5.43)

where the terms involving odd derivatives vanish due to the isotropy of the weight func-tion and the anti-symmetry of the terms y− x with respect to x. The term with ∇s∇s

condensates to a single term involving the surface Laplacian operator ∆s with the pa-rameter c = 1

2L2. Expression (5.43) is also called an ‘explicit’ gradient approximation of

the nonlocal integral (5.38), since Bgb

(x) is given explicitly in terms of Bgb(x) and itsderivatives. At this point, note that substituting the gradient approximation (5.43) andnonlocal expression (5.37) into the chemical potential (5.34) yields

µ =∂ψe

l (Bgb(x))

∂Bgb(x)− κ∆sB

gb(x) . (5.44)

where κ = λc is a gradient energy coefficient. Expression (5.44) resembles the chemicalpotential derived for two-phase systems [75, 137], and can be considered as a weaklynonlocal formulation. The strongly nonlocal formulation, which will be used in thiswork, is found by applying the surface Laplacian operator to Eq. (5.43), multiplying itwith c, subtracting the result from Eq. (5.43) and neglecting the higher-order terms.This leads to the relation [135, 185]

Bgb

(x)− c∆sBgb

(x) = Bgb(x) , (5.45)

which is a Helmholtz equation for Bgb

. In contrast to Eq. (5.43), expression (5.45) iscalled an ‘implicit’ gradient approximation.

The Helmholtz equation of the type (5.45), here used for the determination of the nonlocalgrain boundary net defect density, implies a strongly nonlocal formulation, as shown by[136] in the context of damage mechanics. It retains the truly nonlocal character of thenonlocal integral expression (5.38), in contrast to the explicit gradient approximationin Eq. (5.43), and it implies another weight function than the one given in Eq. (5.41).Although the length scale L governs the intensity of the spatial interactions within thegrain boundary, the explicit gradient approximation is only weakly nonlocal, i.e. theinteractions are limited to an infinitesimal neighbourhood. In this case, the truly nonlocalcharacter is lost due to the truncation of the Taylor series expansion. The reader isreferred to [136] for more details on weakly and strongly nonlocal formulations.

Page 108: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.3 Defect redistribution 95

Summary continuum redistribution model

The continuum model for defect redistribution along the grain boundary consists of thefollowing set of equations. The grain boundary net defect balance equation

dBgb

dt= −∇s · J + S with Bgb(t = 0) = B0 , (5.46)

where the flow J of the net defect is given by

J = −η∇s µ = −η∇s

[∂ψe

l

∂Bgb + λ(Bgb −B

gb)]

. (5.47)

Here, the nonlocal grain boundary net defect density is obtained from the Helmholtzequation

Bgb − c∆sB

gb= Bgb , (5.48)

the derivative of the local energy density reads

∂ψel

∂Bgb =[M −N −N ln(|Bgb|)

] Bgb

|Bgb| + CeN(Bgb −B0) , (5.49)

and the source/sink term is given by

S =∑α

(cα(e)qα(e) + cα(s)qα(s)

)sα +

∑β

(cβ(e)qβ(e) + cβ(s)qβ(s)

)sβ . (5.50)

The continuum defect redistribution model summarised here introduces the need foradditional boundary and initial conditions for the solution of Eqs. (5.46) and (5.48). Theboundary condition for the redistribution equation (5.46) is the zero flow of net defectdensity at the edge L, i.e. the homogeneous (natural) Neumann condition nΓ · J = 0,which means that there is no loss or gain of the net defect density through L. Moreadvanced triple junction models could be used here as well. For the Helmholtz equation(5.48), a homogeneous Neumann boundary condition is used, i.e. nΓ ·∇sB

gb= 0 (cf.

[136, 185]). With this condition the global net defect density in the grain boundary ispreserved in the nonlocal averaging∫

Γ

BgbdΓ =

∫Γ

Bgb dΓ . (5.51)

5.3.2 Discussion

The continuum model consisting of the set of equations (5.46)-(5.50) incorporates theredistribution of the defect content within the grain boundary. It is an extension ofthe model for the grain boundary net defect as summarised in Section 5.2.3. Note thatthe model without redistribution is recovered if the mobility η and the length scale L

Page 109: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

96 5 Defect redistribution in a continuum grain boundary plasticity model

are set to zero. The former means that the dislocations cannot move along the grainboundary, whereas the latter implies that the nonlocal grain boundary energy densityψe

nl is disregarded. Eq. (5.46) then represents the rate form of Eqs. (5.18) and (5.19).Furthermore, the spreading of the defect content occurs via a diffusion-like equationand shows similarities with phase field modelling [31, 125, 146]. In fact, the evolutionequation (5.46) with (5.47) is of the Cahn-Hilliard type [25, 26]. The grain boundarydefect content is conserved if the source/sink term S = 0.

The mobility of dislocations in the grain boundary η is proportional to DgbbkT [71], where

Dgb is the grain boundary self-diffusion, k is the Boltzmann constant, T is the absolutetemperature and b is, as before, the magnitude of the Burgers vector. The grain boundaryself-diffusion follows an Arrhenius form [175]

Dgb = Dgb0 exp

(− Q

gb

RgT

), (5.52)

where Dgb0 is a pre-exponential factor, Qgb is the grain boundary self-diffusion activation

energy and Rg is the gas constant. In this work, the mobility parameter η is thereforetaken as

η =δDgbb

kT=δDgb

0 b

kTexp

(− Q

gb

RgT

), (5.53)

with δ the grain boundary width. The quantities δ and Dgb combined can be mea-sured experimentally [172]. In general, the grain boundary self-diffusion depends onthe grain boundary structure, i.e. the less ordered the grain boundary, the higher theself-diffusion, due to the more ‘open’ structure for diffusional transport. Hence, high-angle grain boundaries typically have larger values of self-diffusion than low-angle grainboundaries. Furthermore, the grain boundary diffusivity can be anisotropic, which leadsto different values of self-diffusion in different directions in the boundary plane [7]. Thisanisotropy can be incorporated via the mobility tensor 4η, but is not considered here.

It has been reported from experiments [52, 102, 126] and fine-scale simulations [47, 87,139] that an absorbed lattice dislocation may decompose into two smaller dislocations(possibly after reacting with other defects first), one of which glides along the grainboundary plane, whereas the other remains at the point of impingement. Glide is possiblewhen the Burgers vector of the product is parallel to the grain boundary plane, whichshould be a physically possible slip plane. Movement of a product with a Burgers vectorcomponent lying out of the boundary plane requires both climb and glide. Climb is athermally activated process and is controlled by grain boundary diffusion. In general,due to the rather special conditions for glide, it has been concluded in [71] that, from theexperimental point of view, independently of the orientation of the Burgers vector withrespect to the grain boundary plane, the mobility of dislocations is controlled by grainboundary diffusion. In this work, at the continuum scale, no distinction is made betweenglissile and sessile dislocations in the redistribution of the net defect. A net defect densityis considered and it is assumed that all redistribution processes occur through physicallydifferent types of atomic shuffling collectively described by the evolution equation (5.46).

Page 110: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.4 Solution procedure and finite element implementation 97

grain boundary extendedcrystal plasticity framework

continuum net defectredistribution

interface normal slip rate

GB net defect density

equilibrium equation

GND balance equations

+ IC, BC and GB model

implicit time integration

incremental-iterative procedure

3D/2D mesh

+ IC and BC

redistribution equation

Helmholtz equation

2D/1D mesh along GB

explicit time integration

at start of increment

Figure 5.2: Schematic of the staggered solution scheme.

Finally, in general, the (coupled) processes of grain boundary sliding, grain boundary mi-gration and grain rotation are accompanied with the defect redistribution and the changeof the grain boundary structure [71, 144, 186]. It is expected that these phenomena havea significant influence on the macroscopic behaviour. In the case of a polycrystal, thegrain boundary is subjected to additional constraints imposed by the microstructure,such as triple junctions. Other stress relaxation mechanisms such as cavitation and de-cohesion can also occur. The presence of impurities and grain boundary segregationfurther increases the complexity of the problem. These phenomena and features are notmodelled in this work.

5.4 Solution procedure and FE implementation

The grain boundary extended strain gradient crystal plasticity framework combined withthe continuum defect redistribution model, complemented by appropriate initial (IC)and boundary conditions (BC), comprises a set of highly nonlinear coupled equations.Different solution strategies exist, one of which is to solve the set of equations in afinite element (FE) setting using a staggered solution scheme, schematically shown inFig. 5.2. The staggered solution procedure consists of the sequential solution of twoparts of the model, namely (i) the bulk and grain boundary model and (ii) the grainboundary redistribution model. The grain boundary redistribution model is solved on adomain of one dimension lower than the bulk and grain boundary model.

The finite element implementation of the strain gradient crystal plasticity frameworkenhanced with a grain boundary model follows the procedure outlined in Appendix A. Itconsists of the equilibrium equation and the GND balance equations in the weak form.The displacements and the GND densities act as the degrees of freedom. The grainboundary is modelled by uncoupling the finite elements adjacent to the grain boundaryby inserting double nodes. It is assumed that the displacements are continuous acrossthe grain boundary, which is accomplished by tying the corresponding displacementdegrees of freedom. Implicit time integration is employed, which results in a standard

Page 111: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

98 5 Defect redistribution in a continuum grain boundary plasticity model

incremental-iterative procedure for the solution of this part of the framework. For thesolution of this part the grain boundary net defect density distribution along the grainboundary needs to be known, which is obtained from the second part of the framework,see Fig. 5.2.

The finite element implementation of the continuum defect redistribution model solvesthe weak forms of the redistribution equation (5.46) and the Helmholtz equation (5.48),with the local and nonlocal grain boundary net defect densities acting as degrees of free-dom. Considering the grain boundary net defect balance equation (5.46), multiplicationwith a weight function w, integration over the grain boundary, application of the productrule and employing the surface divergence theorem (5.26), while considering a flat grainboundary and zero flow of the net defect density at the edge L, i.e. nΓ · J = 0, leads tothe weak form∫

Γ

w · dBgb

dtdΓ =

∫Γ

(∇sw)T

: J dΓ +

∫Γ

w · S dΓ ∀w . (5.54)

Substituting the flow J = −η∇s µ and applying explicit time integration for tn+1 =tn + ∆t yields after elaboration∫

Γ

w ·Bgb(tn+1) dΓ =

∫Γ

w ·Bgb(tn) dΓ +

∫Γ

∆tw · S(tn) dΓ

−∫

Γ

η∆t

(∇sw ·

∂µ

∂Bgb

∣∣∣∣tn

)T

: ∇sBgb(tn) dΓ ∀w , (5.55)

where the derivative of the chemical potential with respect to the local grain boundarynet defect density reads

∂µ

∂Bgb =

(M −N −N ln(|Bgb|)

|Bgb| + CeN + λ

)I−(M −N ln(|Bgb|)

|Bgb|3)BgbBgb . (5.56)

Next consider the Helmholtz equation (5.48). The weak form is obtained by multiplyingwith a weight function w, integrating over the grain boundary, applying the productrule and the surface divergence theorem (5.26) for a flat grain boundary, while using the

homogeneous Neumann condition nΓ ·∇sBgb

= 0 at the edge L∫Γ

w ·Bgb(tn) dΓ +

∫Γ

c (∇sw)T

: ∇sBgb

(tn) dΓ =

∫Γ

w ·Bgb(tn) dΓ ∀w . (5.57)

Applying the standard finite element discretisation leads to two sets of discretised linearequations

KBB˜Bgb(tn) =

˜fB(tn) and KBB

˜Bgb(tn+1) =

˜fB(tn) , (5.58)

which are solved at the start of each increment in the incremental-iterative solutionprocedure of the grain boundary crystal plasticity part of the framework. In order tosolve this second part of the framework, the distribution of the interface normal slip rate

Page 112: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.5 Numerical study of defect redistribution 99

at the grain boundary q(tn), as required for Bgb, needs to be transferred from the firstpart, see Fig. 5.2.

The finite element meshes for the GND density fields and the local and nonlocal grainboundary net defect densities do not necessarily have to be conforming. The grainboundary integration point values of the interface normal slip rates of the GND densitymesh are passed to the net defect density mesh through linear interpolation. For thetransfer of the grain boundary net defect density nodal values from the net defect densitymesh to the grain boundary integration points of the GND density mesh, the sameprocedure is applied.

5.5 Numerical study of defect redistribution

The strain gradient crystal plasticity framework enhanced with a grain boundary model,as summarised in Section 5.2, extended with the net defect redistribution along the grainboundary developed in Section 5.3, are numerically implemented next, using the solutionprocedure described in the previous section. The influence of the defect redistributionon the macroscopic behaviour is investigated in the case of a bicrystal problem in simpleshear. To this end, first the model problem is explained and the material parameters aregiven. Then the discretisation and the boundary conditions are described, after whichthe results are presented.

5.5.1 Model problem

In order to examine the effect of the redistribution of the net defect within the grainboundary on the macroscopic response, an academic problem of a bicrystal under planestrain conditions is considered, as shown in Fig. 5.3. The bicrystal consists of grainsA (blue) and B (red), which have width d in the e1-direction and height h in the e2-direction, and are separated by a grain boundary. A grain size d = h = 2 µm will beconsidered and only edge dislocations are taken into account in this planar configuration.

The material used in the simulations is copper. Copper has a face-centered cubic (FCC)crystalline structure, in which 12 slip systems exist, denoted by {111}〈110〉, where {111}corresponds to the family of slip planes and {110} to the family of slip directions. It hasbeen shown [98, 152, 165] that plane strain conditions emanate when certain slip systemsact cooperatively, for which the 12 slip systems associated with the FCC crystal reduceto three effective slip systems. This is illustrated in Fig. 5.4 for the state of plane strainin the [110]-[001] plane. The (111)[110] and (111)[110] slip systems can be combined toform an effective in-plane slip system, called the complex system [152]. In this work,the effective complex slip system acts in the [110] direction and is referred to as slipsystem α = 1. The (111)[011] and (111)[101] slip systems operate in equal amounts toform an effective plane strain slip system on the (111) plane in the [112] direction andis here referred to as slip system α = 2. Similarly, the slip system α = 3 is the effective

Page 113: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

100 5 Defect redistribution in a continuum grain boundary plasticity model

U

h

2d

grain A grain B

[001]e2 =

1√2

[110]e1 =

[110]e31√2

=

Figure 5.3: Geometry and boundary conditions for simple shear deformation of a bicrystal. Stronginhomogeneities are introduced in the form of slip bands at the grey regions. Two of the three meshesemployed in the staggered solution procedure are depicted, which correspond to the bilinear quadrilater-als (solid squares) for the GND densities (the displacement finite elements are biquadratic quadrilaterals)and the linear elements (solid circles) for the local and nonlocal GB net defect densities.

plane strain slip system on the (111) plane and in the [112] direction as a combinationof (111)[101] and (111)[011]. Note that the three effective in-plane slip systems can beobtained by applying counterclockwise rotations of 0◦, 54.74◦ and 125.26◦, respectively,relative to the [110] direction. Furthermore, from geometrical considerations it followsthat the effective lengths of the Burgers vectors for these three slip systems α = 1, 2 and3 are b, 1

2

√3b and 1

2

√3b, respectively [165, 200].

[010]

[001]

[100]

(111) with [011] and [101]

(111) with [101] and [011]

(111) and (111) with [110]

s1

m1

m2 s

2

m3

s3

φ

ω

φ = 70.53°

ω = 54.74°

[110]

[001]

[110]

Figure 5.4: An FCC crystal with the simultaneous action of slip systems to resemble the plane straincondition (left) and the three effective plane strain slip systems in the [110]-[001] plane (right).

Page 114: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.5 Numerical study of defect redistribution 101

[001]

perfect state ( 0°)θ= misoriented state ( °)θ=70.53reference state

θ2

θ2

[110]

s1

m1

m2 s

2

m3

s3

m1A

s1A

m2A

s2A

m3A

s3A

m1B

s1B

m2B

s2B

m3B

s3B

m1A

s1A

m2A

s2A

m3A

s3A

m1B

s1B

m2B

s2B

m3B

s3B

Figure 5.5: Orientations of the grains in the bicrystal starting from the reference state (left) via theperfect state (middle) to the misoriented state (right).

The orientations of the grains in the bicrystal are explained in Fig. 5.5. Starting fromthe reference state, corresponding to the fixed coordinate system, the grains are firstrotated (vector transformation) 90◦ counterclockwise to the state where no misorientationbetween the adjacent grains is present, i.e. a perfect crystal. A target grain boundaryof misorientation θ is recovered by applying a rotation of θ/2 in opposite directionsfor grains A and B, respectively. The grain boundaries that are produced by theseoperations are symmetric tilt boundaries. For the particular case considered in thepresent work, the rotation axis is parallel to the [110] direction. The structure and energyof symmetric tilt grain boundaries with this rotation axis in aluminium and copper in theentire misorientation range have been investigated in detail in Chapter 3. In this chapter,a misorientation θ = 70.53◦ between the grains is considered, which corresponds to theΣ3 (112) symmetric tilt boundary. Note that for this misorientation the third effectiveslip system in grain A and the second effective slip system in grain B are horizontal, seeFig. 5.5.

5.5.2 Material parameters

The bulk material parameters used in the calculations are representative of copper andare similar to those of Evers et al. [57] and Bayley et al. [16]. The values are listedin Table 5.1, where the interaction and immobilisation matrices Aαξ and Hξζ resultfrom the interaction and immobilisation coefficients a0 and h0 on the diagonal, and theoff-diagonal entries a3 and h3. The interface parameters are listed in Table 5.2. Theinterface reference slip rate q0 and rate sensitivity exponent n are taken to be the sameas their bulk counterparts γ0 and m. The energetic interface parameters M and N havebeen determined in Chapter 3. The lumped interface self-diffusion parameter δDgb

0 andthe grain boundary self-diffusion activation energy Qgb for high-purity copper are takenfrom [172].

In order to illustrate the effect of the defect redistribution along the grain boundary,strong inhomogeneities are introduced in both crystals, as shown in Fig. 5.3. The greyregions represent slip bands, in which the slip resistance sα (Eq. (5.7)) for slip in the bulk(Eq. (5.6)) is lowered for the third and second effective slip system (oriented horizontally)in grains A and B, respectively. For this purpose, the initial SSD density ραs0 (Eq. (5.8))of the other slip systems is reduced with a factor 10. Furthermore, the interface normal

Page 115: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

102 5 Defect redistribution in a continuum grain boundary plasticity model

Table 5.1: Bulk material parameters representative of copper.

Parameter Description Magnitude

E Young’s modulus 144 GPaG Shear modulus 54.1 GPaν Poisson’s ratio 0.33b Burgers vector length 0.256 nmγ0 Reference slip rate 0.001 s−1

a0 Interaction coefficient 0.06a3 Interaction coefficient 1.0h0 Immobilisation coefficient 0.2h3 Immobilisation coefficient 1.0R GND evaluation radius 1 µmm Rate sensitivity exponent 0.1K Dislocation segment length constant 26yc Critical annihilation length 1.6 nmρs0 Initial SSD density 7.0 µm−2

Table 5.2: Interface parameters used in the numerical simulations.

Parameter Description Magnitude

q0 Interface reference slip rate 0.001 s−1

n Interface rate sensitivity exponent 0.1Ce Interface energy constant 103

R Interface normal slip resistance 0.25 N mm−1

M Interface energy constant 358 mJ m−2

N Interface energy constant 1434 mJ m−2

k Boltzmann constant 1.38 · 10−23 J K−1

Rg Gas constant 8.314 J mol−1K−1

Qgb Interface self-diffusion activation energy 72.47 kJ mol−1

δDgb0 Lumped interface self-diffusion parameter 3.98 · 10−16 m3 s−1

slip activity (Eq. (5.14)) of these two slip systems inside the slip bands (at the GB) isincreased by lowering the interface normal resistance R by a factor 25.

5.5.3 Discretisation and boundary conditions

The numerical implementation of the framework follows the solution procedure elabo-rated in Section 5.4. The specific initial and boundary conditions for the model problemconsidered in this work along with the particular discretisation of the unknown variablesare described next for the two parts of the staggered solution scheme, see Fig. 5.2.

Page 116: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.5 Numerical study of defect redistribution 103

Considering first the equilibrium equation, the displacements at the top and bottomboundary are fully prescribed, see Fig. 5.3. Simple shear deformation is simulated byapplying a prescribed displacement U = γht on the top boundary in the e1-direction,while the bottom boundary is fully fixed. A constant shear rate of γ = 10−3 s−1 isapplied, up to a maximum of 1% macroscopic shear. Traction-free boundary conditionsare imposed on the left and right boundary. For the GND balance equations, microhard(no slip, homogeneous Neumann) boundary conditions are applied to the top and bottomboundary, where the specimen is clamped. The left and right boundary are subjected tomicrofree (zero GND density, homogeneous Dirichlet) boundary conditions. The GNDdensity field is initially zero. The constitutive behaviour at the grain boundary is dictatedby the grain boundary model. Each grain in the bicrystal is discretised with 4 elementsin the e1-direction and 16 elements in the e2-direction, see Fig. 5.3. The displacementfield is discretised with 8-node quadrilaterals using 2×2 reduced integration, whereas theGND density field is interpolated with 4-node bilinear shape functions employing 2 × 2full integration. The slip bands are located at the 13th and 14th row of elements in grainA and at the 3rd and 4th element rows in grain B.

The boundary condition for the redistribution equation is the zero flow of net defectdensity out of the clamped specimen where the top and bottom boundary meet the grainboundary, i.e. the homogeneous (natural) Neumann condition nΓ ·J = 0. The initial localgrain boundary net defect density is uniform and equals the intrinsic defect density B0,as follows from the atomistic simulations (see Chapter 3). For the Helmholtz equation,

the homogeneous Neumann boundary condition is used, i.e. nΓ ·∇sBgb

= 0. The grainboundary is discretised with a 1D mesh of 32 elements in the e2-direction, see Fig. 5.3.The local and nonlocal net defect density fields along the grain boundary are discretisedwith 1D linear elements using 2 integration points.

5.5.4 Results

To assess the effect of the defect redistribution along the grain boundary on the macro-scopic response, as well as the microscopic fields, three different mobilities η are consid-ered by selecting the temperatures T = 293, 323 and 353 K. In this set of simulations,the line interface redistribution coefficient λ and the nonlocal interface length parameterL are fixed at λ = 0.6 J/m2 and L = 100 nm. Furthermore, additional simulations thatserve as reference cases are included, where (i) the grain boundary acts as an impene-trable boundary, i.e. a microhard grain boundary with Fe = 0 and R → ∞, and (ii) noredistribution occurs within the grain boundary, i.e. η = 0 and L = 0.

The results of the simulations are shown in Fig. 5.6. Fig. 5.6(a) shows the resultingmacroscopic stress-strain response, while Fig. 5.6(b) plots the scalar local grain boundary

net defect density, defined as |Bgb| =√Bgb ·Bgb, along the grain boundary. In the

case of the microhard grain boundary, from the onset of plasticity in the bulk crystals,dislocations start to pile up at the interface inducing GND densities that contribute to therate of strain hardening, see Fig. 5.6(a), through the slip resistance sα and back stress ταb(Eq. (5.6)). The local grain boundary net defect density is not altered and remains equal

Page 117: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

104 5 Defect redistribution in a continuum grain boundary plasticity model

0 0.002 0.004 0.006 0.008 0.010

20

40

60

80

100

120

She

ar

str

ess (

MP

a)

Applied shear (−)

microhardno redistributionT = 293 KT = 323 KT = 353 K

(a)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Scala

r lo

cal G

B n

et defe

ct density (

−)

Dimensionless vertical position (−)

microhardno redistributionT = 293 KT = 323 KT = 353 K

(b)

Figure 5.6: The influence of the net defect redistribution along the grain boundary on the macroscopicresponse (a) and the distribution of the scalar local grain boundary net defect density at the final appliedload (b) for three different temperatures. The reference cases of a microhard grain boundary and noredistribution are included.

to the initial intrinsic net defect density throughout the simulation. For the situationincluding the grain boundary plasticity model, but excluding the defect redistribution,the behaviour initially follows the microhard result (Fig. 5.6(a)) up to the point wherethe interface normal slip resistance is reached and dislocations are absorbed into the grainboundary, where they continue to accumulate locally due to the absence of redistributionalong the grain boundary, see Fig. 5.6(b). The absorption of dislocations causes the rateat the which dislocations pile up at the grain boundary to decrease, whereas the interfaceback stress grows due to the local build-up of a net defect in the grain boundary at theregions where the slip bands impinge on the grain boundary. The net effect of these twomechanisms yields a decrease in the macroscopic rate of strain hardening, see Fig. 5.6(a).Incorporating the redistribution of the grain boundary net defect, the initial behaviouris close to the microhard result up to the point of dislocation absorption into the grainboundary, see Fig. 5.6(a). For increasing temperature, and hence mobility, the extent towhich the net defect content is redistributed increases, see Fig. 5.6(b). Due to the defectredistribution along the grain boundary, the local accumulation of the net defect relaxes,which lowers the interface back stress at the points of impingement. This decreases therate of pile-up accumulation of the bulk dislocations at these interface spots and hencereduces the rate of strain hardening, see Fig. 5.6(a), the effect of which is larger for highertemperature.

In order to obtain a better understanding of the influence of temperature on the defectredistribution, the evolution of the local grain boundary net defect density is plotted inFig. 5.7 for the three temperatures and the case without redistribution of the net defect.The evolution results provide insight in the competition between the local net defectaccumulation through the source term and the redistribution of the net defect along thegrain boundary via the diffusion term (see Eq. (5.46)). In the absence of defect redis-

Page 118: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.5 Numerical study of defect redistribution 105

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

no redistribution

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(a)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

T = 293 K

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(b)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

T = 323 K

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(c)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7S

ca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

T = 353 K

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(d)

Figure 5.7: The evolution of the scalar local net defect density along the grain boundary for (a) the noredistribution case, (b) redistribution at temperature T = 293 K, (c) redistribution at T = 323 K and(d) redistribution at T = 353 K.

tribution the evolution is completely governed by the local net defect accumulation, seeFig. 5.7(a). For the lowest temperature, Fig. 5.7(b), the evolution is still dominated bythe local build-up of the grain boundary net defect. A transition occurs for the inter-mediate temperature, where the net defect accumulation and redistribution are presentin approximately equal amounts (Fig. 5.7(c)). In the case of the highest consideredtemperature, the redistribution of the net defect controls the evolution, see Fig. 5.7(d).

Next, the effect of the line interface redistribution coefficient λ and the nonlocal interfacelength L on the macroscopic and microscopic behaviour is examined. In the first seriesof simulations, the temperature and the nonlocal interface length are fixed at T = 293K and L = 100 nm, whereas the interface redistribution coefficient λ is varied. Theresults are shown in Fig. 5.8 for λ = 0.6, 6 · 103, 6 · 104 and 3 · 105 J/m2. The influenceof the line interface redistribution coefficient λ on the macroscopic response is minor,

Page 119: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

106 5 Defect redistribution in a continuum grain boundary plasticity model

0 0.002 0.004 0.006 0.008 0.010

20

40

60

80

100

120

Sh

ea

r str

ess (

MP

a)

Applied shear (−)

λ = 6e−1 J/m2

λ = 6e3 J/m2

λ = 6e4 J/m2

λ = 3e5 J/m2

(a)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

λ = 6e3 J/m2

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(b)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

λ = 6e4 J/m2

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(c)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

λ = 3e5 J/m2

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(d)

Figure 5.8: The influence of the line interface redistribution coefficient λ in the net defect redistributionalong the grain boundary on the macroscopic response (a) and the evolution of the scalar local net defectdensity for the line interface redistribution coefficients (b) λ = 6 · 103 J/m2, (c) λ = 6 · 104 J/m2 and(d) λ = 3 · 105 J/m2.

even for the large range of six orders of magnitude, see Fig. 5.8(a). A more pronouncedinfluence is observed microscopically when considering the evolution of the scalar localgrain boundary net defect density in Fig. 5.7(b) and Fig. 5.8(b)-(d). For increasing lineinterface redistribution coefficient λ, the diffusion term becomes stronger with respectto the source term, resulting in a more homogeneous distribution of the local net defectdensity.

In the second series of simulations, the nonlocal interface length L is varied, while thetemperature and the line interface redistribution coefficient are fixed at T = 293 K andλ = 3 · 105 J/m2. The results are shown in Fig. 5.9 for L = 100, 200, 400 and 800nm. Again, for the considered parameter range, the macroscopic influence is limited, seeFig. 5.9(a). Microscopically, considering the evolution of the scalar local grain boundary

Page 120: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.5 Numerical study of defect redistribution 107

0 0.002 0.004 0.006 0.008 0.010

20

40

60

80

100

120

Sh

ear

str

ess (

MP

a)

Applied shear (−)

L = 100 nmL = 200 nm

L = 400 nmL = 800 nm

(a)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

L = 200 nm

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(b)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7

Sca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

L = 400 nm

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(c)

0 0.2 0.4 0.6 0.8 10.64

0.65

0.66

0.67

0.68

0.69

0.7S

ca

lar

loca

l G

B n

et

de

fect

de

nsity (

−)

Dimensionless vertical position (−)

L = 800 nm

Ap

plie

d s

he

ar

(−)

0

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.01

(d)

Figure 5.9: The effect of the nonlocal interface length L in the net defect redistribution along the grainboundary on the macroscopic response (a) and the evolution of the scalar local net defect density forthe nonlocal interface lengths (b) L = 200 nm, (c) L = 400 nm and (d) L = 800 nm.

net defect density in Fig. 5.8(d) and Fig. 5.9(b)-(d), the variation of L has a clear effect.The diffusion contribution becomes even more dominating for increasing L, leading toa larger annihilation of the net defect at the center of the grain boundary, where thepositive and negative defect content from the slip bands in grains A and B, respectively,cancel in the redistribution.

To summarise, the results indicate that the grain boundary net defect redistribution hasa clear macroscopic influence, but is mainly due to the temperature directly affectingthe mobility η of the dislocations within the grain boundary. For the current set ofparameters, only a limited effect on the macroscopic response was observed for varyingthe line interface redistribution coefficient λ and the nonlocal interface length L. All threeparameters (η, λ and L) have, however, a significant microscopic effect. Note that bysubstituting the Helmholtz equation (5.48) into the redistribution equation (5.46), similar

Page 121: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

108 5 Defect redistribution in a continuum grain boundary plasticity model

as in expression (5.44), the gradient energy coefficient κ = λc = 12λL

2 is identified, whichcharacterises the relative contribution of the nonlocal term into the redistribution. Thenonlocal interface length L changes κ quadratically, for this reason having a larger effectas compared to the line interface redistribution coefficient λ. For the bicrystal problemconsidered here, it is expected that further increase of the parameters η, λ and L willincrease the influence on the macroscopic and microscopic behaviour, up to the pointwhere the diffusion contribution completely dominates the evolution, whereby absorbedlattice dislocations directly redistribute and annihilate.

As a final remark, in general, the nonlocal term containing the interface redistributioncoefficient λ and the nonlocal interface length L, does not only incorporate spatial inter-actions, but also serves as a regularising term in the cases when the local energy densityis nonconvex, e.g. in spinodal decomposition [185] or dislocation patterning [211], bypenalising strong gradients in the quantity of interest. The local grain boundary energydensity chosen in this work (Eq. (5.32)), based on a straightforward generalisation ofthe initial grain boundary energy determined from atomistic simulations, already hasa convex character depending on the value of Ce. In the specific planar configurationexamined here, this can be verified analytically by calculating the eigenvalues of the sec-ond derivative of the local energy density, given in Eq. (5.56) without the λ, which arepositive, and hence indicate convexity by positive-definiteness of the second derivative.Hence, pattern formation in the redistributed defects along the grain boundary is notexpected.

5.6 Conclusions

Many different interaction mechanisms between dislocations and grain boundaries existthat have an important effect on the mechanical behaviour of polycrystalline materials.To date, significant challenges are present in physically based modelling of these grainboundary plasticity features in macroscopic continuum frameworks. The main contribu-tion of this chapter and conclusions are:

1. A novel model that incorporates the nonlocal relaxation of the grain boundary netdefect due to the redistribution of the defect content along the interface, which isdeveloped within the context of strain gradient crystal plasticity.

2. The continuum redistribution model consists of a set of equations, including abalance and constitutive equations, which lead to an evolution equation of the netdefect density at the grain boundary and a Helmholtz equation that accounts fornonlocal spatial interactions within the interface.

3. The net defect evolution at a grain boundary material point is determined by: (i)flow of the net defect along the grain boundary and (ii) a source/sink term dueto the in- and outflow of lattice dislocations at the grain boundary. Furthermore,defect content of opposite sign can annihilate in the redistribution.

Page 122: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

5.6 Conclusions 109

4. The flow of the interface defect resulting in the redistribution is driven by thegradient of a chemical potential, which depends on local and nonlocal defect con-tributions to the grain boundary energy density.

5. The defect redistribution model, based on a vectorial quantity defined at a sur-face, has a diffusional character and shows similarities with phase field modellingapproaches.

6. The complete framework is implemented numerically employing a staggered solu-tion scheme in order to investigate the influence of the redistribution of the grainboundary defect content on the macroscopic response and the microscopic fields inthe case of a copper bicrystal subjected to simple shear.

7. The resulting defect redistribution has a clear macroscopic and microscopic effect,which becomes stronger for increasing temperature, as a result of the increasedmobility of the defect content, and for increasing nonlocal interface parameters,which enhance the nonlocal contribution to the redistribution.

Different processes that accompany the defect redistribution and the change of grainboundary structure, such as grain boundary sliding, grain boundary migration and grainrotation, are not taken into account in the current modelling framework. These processesmainly occur at higher temperatures at which other mechanisms in the bulk, such as dis-location climb, become thermally activated. This would also necessitate adjustments ofthe crystal plasticity model, since these mechanisms are not incorporated in the currentbulk framework. In polycrystals, the presence of triple junctions poses additional con-straints. It is expected that these mechanisms and features can change the microscopicand macroscopic behaviour considerably, the investigations of which are topics for futurework.

Page 123: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary
Page 124: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Chapter 6

Conclusions and outlook

6.1 Conclusions

The objective in this thesis was to bridge the scale between the fine-scale interactionmechanisms associated with grain boundary plasticity and the polycrystalline continuumscale, where most engineering design methods rely on. This has been established by thedevelopment of a multiscale grain boundary interface framework within the context ofdislocation based gradient enhanced crystal plasticity.

First, in Chapter 2, a thermodynamically consistent grain boundary model has beenformulated, which is incorporated in a microstructurally motivated strain gradient crystalplasticity framework. The main results and conclusions of Chapter 2 are:

• The net effect of the different interactions between dislocations and grain bound-aries is represented by an interface normal slip measure, for which an evolutionequation is proposed constituting the grain boundary flow rule.

• Plastic slip at the grain boundary is governed by the competition between threeinterface microforces, i.e. a bulk-induced microforce, an energetic microforce and adissipative microforce.

• The bulk-induced microforce, derived from geometrically necessary dislocation den-sities in the bulk crystal, provides a driving force for interface normal slip, whichdepends on the magnitude of the dislocation pile-up at the interface. By makinga connection with a full internal stress formulation, the bulk-induced microforcescapture the intra-granular coupling in a physically realistic manner.

111

Page 125: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

112 6 Conclusions and outlook

• The energetic microforce acts as an interface back stress, restricting the accumu-lation of the residual interface defect. The energetic interface microforce providesinter-granular slip system coupling, accounting for both the magnitude and direc-tion of contributions to the residual defect from all slip systems in the two neigh-bouring grains. The geometric conditions controlling the interaction of a particularslip system with the grain boundary are implicitly accounted for.

• The dissipative microforce provides a resistance to slip at the interface.

• A detailed numerical analysis of the influence of the interface constitutive param-eters and the crystal geometry in the case of a simplified periodic bicrystal hasrevealed that the macroscopic behaviour is altered considerably with respect to theconventional limiting grain boundary conditions. Qualitative agreement betweenthe model predictions and the experimental observations reported in the literaturehas been shown.

Only atomistic simulations can provide access to the details of the grain boundary struc-ture and energy. This demands a multiscale description to upscale these fine-scale grainboundary details to the continuum scale. In an attempt to bridge the gap between thematerials science and mechanics views on modelling of an initial grain boundary structureand energy, Chapter 3 proposes a multiscale approach for resolving the initial interfacestructure and energy on the continuum scale based on the output of atomistic simula-tions, thereby consolidating various approaches for characterising grain boundaries. Themain results and conclusions are:

• An intrinsic net defect density, representing the initial boundary structure, is re-lated to atomistically computed grain boundary energies for symmetric tilt grainboundaries with the 〈100〉, 〈110〉 and 〈111〉 rotation axes in aluminium and copper.

• Characteristic features of the approach include: (i) grain boundary energies andboundary period vectors as obtained from atomistic simulations, (ii) structural unitand dislocation descriptions of the grain boundary structure and (iii) the Frank-Bilby equation to determine the dislocation content.

• The results have shown that the proposed multiscale approach yields adequate de-scriptions of the atomistically computed grain boundary energies for both low- andhigh-angle grain boundaries in the complete misorientation range for the consid-ered type of grain boundaries. A comparison with the experimental results for the〈110〉 rotation axis in aluminium, available in literature, shows a good agreementand supports the findings.

• The developed atomistic-to-continuum approach can be directly incorporated ingrain boundary interface continuum models, realistically reflecting the initial bound-ary structure and energetics.

In Chapter 4, the multiscale atomistic-to-continuum model of the initial structure andenergy of grain boundaries, developed in Chapter 3, has been incorporated into the grain

Page 126: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

6.1 Conclusions 113

boundary extended crystal plasticity framework of Chapter 2 in order to examine theparticular role and influence of the energetic contributions to the grain boundary plasticflow on the resulting macroscopic response. The main results and conclusions of this partare:

• A generalised grain boundary energy as a function of the evolution of the interfacenet defect, representing the grain boundary structure upon plastic deformation,has been proposed and used to determine an expression for the energetic interfacemicroforce.

• An analytical study of the developed initial energetic interface microforce revealeda physically sound behaviour for low- and high-angle symmetrical tilt grain bound-aries.

• The resulting framework has been implemented numerically and applied for theinvestigation of a periodic bicrystal, including the complete set of 12 slip systems,subjected to simple shear deformation. Eight low- and high-angle symmetric tiltboundaries with the [001] rotation axis have been analysed throughout the fullmisorientation range. It has been observed that the effect of the energetic interfacemicroforce on the macroscopic response depends strongly on the misorientationbetween adjacent grains.

Up to and including Chapter 4, it was assumed that the grain boundary net defectremained a local property at a material point at the grain boundary. In the literature,there have been many fine-scale observations of redistribution of the defect content alongthe grain boundary. Chapter 5 presents an approach describing the incorporation ofthis relaxation mechanism through the formulation of a continuum defect redistributionmodel along the interface within the grain boundary plasticity framework developed inthe previous chapters. The main results and conclusions of this chapter are:

• The continuum redistribution model consists of a set of equations, including abalance and constitutive equations, which lead to an evolution equation of thelocal net defect density at the grain boundary and a Helmholtz equation for thenonlocal defect density that accounts for spatial interactions within the interface.

• The net defect evolution at a grain boundary material point is determined by: (i)flow of the net defect along the grain boundary and (ii) a source/sink term dueto the in- and outflow of lattice dislocations at the grain boundary. Furthermore,defect content of opposite sign can annihilate in the redistribution.

• The flow of the interface defect resulting in the redistribution is driven by thegradient of a chemical potential, which depends on local and nonlocal defect con-tributions to the grain boundary energy density.

• The resulting defect redistribution model, based on a vectorial quantity defined at asurface, has a diffusional character and shows similarities with phase field modellingapproaches.

Page 127: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

114 6 Conclusions and outlook

• In a numerical study of a bicrystal problem, using the entire framework, it isfound that the defect redistribution has a clear macroscopic and microscopic effect,which becomes stronger for increasing temperature, as a result of the increaseddefect mobility, and for increasing nonlocal interface parameters, which enhancethe energetic interactions driving the redistribution.

The work presented in the thesis provides a foundation to a more thorough analysis ofpolycrystalline metals at the continuum level. Especially when large fractions of grainboundaries are present, the accurate description of the behaviour at these boundariesbecomes increasingly important.

6.2 Outlook

In this work, the developed multiscale model has only been applied to relatively simplebicrystal configurations, which served the exploration of the new grain boundary interfaceframework. However, the model itself is formulated in a general 3D setting. As the nextstep, it is of interest to apply this framework to more complex (academic and industrial)problems at the polycrystalline level.

At the same time, it should be pointed out that the current framework does have itslimitations with respect to the underlying physical phenomena. Moreover, the quantifi-cation of some of the modelling parameters involved is still an open question. Basedon these limitations an outlook for future research directions is given next. The focushere is mainly on the continuum modelling of grain boundaries rather than the bulkmaterial, the latter having its own list of research topics, although the two are intimatelyconnected.

• Whereas qualitative agreement with experiments has been achieved in this work,the quantitative calibration and validation of the model is presently hampered bya lack of suitable quantitative experimental data. In characterising the dislocationcontent in the bulk near the grain boundary, refinement of electron backscatterscanning microscopy results to slip system based data can be a promising route.Furthermore, computer simulations at smaller length scales can assist in estimatingappropriate values for the model parameters.

• In the current approach, the bulk-induced microforce related to one of the adjacentgrains at the interface only depends on the geometrically necessary dislocationdistribution in that grain, i.e. it is assumed that the internal stress field fromthe neighbouring grain is entirely ‘shielded’ by the interface. A more elaborateexpression might be conceived which accounts for the influence of the geometricallynecessary dislocations from the other grain. The level of cross-interface influencefrom geometrically necessary dislocations in the bulk would depend on the detailsof the grain boundary structure and its disruptive screening effect.

Page 128: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

6.2 Outlook 115

• The initial energetic interface microforce profiles can be verified using atomisticsimulations specifically focussing on the elastic interactions between grain bound-aries and dislocations in the absence of external loads. However, care should betaken in the determination of the origin of the interaction effect, because addi-tional contributions due to image forces arising from the presence of free surfacesand elastic anisotropy across the grain boundary also play a role.

• The interface normal slip resistance has been taken constant in the simulationsthroughout the thesis. However, in general, it is expected that the interface nor-mal slip resistance, similar to the energetic microforce, depends on the slip systemorientation, the misorientation between the grains and the orientation of the grainboundary. This can be incorporated into the model through a dependence of theinterface normal resistance on the grain boundary net defect density. The interfacenormal slip resistance could also incorporate the nucleation strength of dislocationsources close to the grain boundary or the additional work required to overcomeconfigurational changes necessary for a dislocation to interact with the grain bound-ary. It should be emphasised that in modelling at the (polycrystalline) continuumlevel, double counting of mechanisms in the bulk and interface descriptions shouldbe avoided, e.g. through lumping of bulk behaviour near the boundary into theinterface model.

• An extension of the developed atomistic-to-continuum multiscale approach for theinitial structure and energy of a grain boundary may include the application to (i)the full space of macroscopic degrees of freedom of a grain boundary, starting withasymmetric tilt boundaries and twist boundaries, (ii) grain boundaries in body-centered cubic and hexagonal close-packed metals, and (iii) interphase boundaries.Furthermore, metallic elements of similar stacking fault energy could be inves-tigated to verify the potential of reducing the number of continuum parametersthrough scaling.

• One of the most important challenges is the development of an adequate contin-uum description for the evolution of the grain boundary structure and energy uponplastic deformation. This in turn will directly influence the effect of the evolu-tion of the energetic and dissipative interface contributions on the macroscopic andmicroscopic behaviour. Tracking the evolution of the grain boundary structureand energy is impossible without fine-scale experiments and multiscale simulationtechniques. Moreover, due to the unique character of the individual interactionsbetween grain boundaries and dislocations, detailed studies of many grain bound-aries are required. This complicates the possible identification of generic continuumscale ‘building blocks’ that are representative of different collective interactions ofdislocations with a grain boundary having a similar macroscopic effect.

• The (coupled) processes of grain boundary sliding, grain boundary migration andgrain rotation accompany the defect redistribution and the change of the grainboundary structure. Other stress relaxation mechanisms such as cavitation and de-cohesion can also occur. These phenomena are not modelled in this work, althoughpossibilities already exist for the straightforward incorporation of some of theseprocesses. For example, grain boundary sliding and opening could be accounted

Page 129: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

116 6 Conclusions and outlook

for by allowing displacement discontinuities at the grain boundary. Furthermore,grain rotation could be included by extending the framework to finite deformations.However, it is noted that most of the phenomena mentioned here mainly occur athigher temperatures at which other mechanisms in the bulk become thermally ac-tivated, also necessitating adjustments of the bulk model.

• In a polycrystalline case, the grain boundary is subjected to additional constraintsimposed by the microstructure, such as triple junctions where three grains meetalong a line. As a first approach, triple junctions can be modelled using limitinghigher-order boundary conditions, such as treating the triple junction as a hardobstacle. At a later stage specific constitutive models should be developed also forthese grain boundary characteristics.

• The presence of impurities would further increase the complexity of the problemby altering both the bulk and interface behaviour. For example, grain boundarysegregation can occur, whereby the impurity atoms accumulate (segregate) at grainboundaries to such an extent that the boundary may become different in chemicalnature, which will change the grain boundary properties and hence the materialproperties. Appropriate constitutive models will have to be developed to describethe resulting behaviour.

Page 130: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Appendix A

Finite element implementation

The numerical implementation follows the procedures outlined by Evers et al. [56] andBayley et al. [16], with all necessary modifications to incorporate the grain boundaryconstitutive model. A summary is presented here with emphasis on the new interfacecontributions. The boundary value problem consists of the conventional equilibriumequation (2.7a) and the GND balance equations (2.3) in the weak form, with the dis-placements and GND densities acting as degrees of freedom. The weak forms for a singlecrystal (within a polycrystal) are∫

V

(∇w)T

: σ dV =

∫S

w · text dS +

∫Γ

w · t dΓ ∀ w , (A.1)

∫V

(vαρ

α(e)g − 1

b∇vα · sα γα

)dV =

∫V

vαρα(e)g0 dV−

1

b

∫S

vαqα(e) dS − 1

b

∫Γ

vαqα(e) dΓ ∀ vα , (A.2)

∫V

(wαρ

α(s)g +

1

b∇wα · lα γα

)dV =

∫V

wαρα(s)g0 dV−

1

b

∫S

wαqα(s) dS − 1

b

∫Γ

wαqα(s) dΓ ∀ wα , (A.3)

with α denoting the slip system. Furthermore, qα(e) = γαsα · n, qα(e) = γαsα · ngb,qα(s) = −γαlα · n and qα(s) = −γαlα · ngb. The displacement field is discretised with8-node quadrilaterals using 2 × 2 reduced integration, whereas the GND density fieldis interpolated with 4-node bilinear shape functions employing 2 × 2 full integration.The grain boundaries are modelled by uncoupling the local finite elements by means of

117

Page 131: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

118 A Finite element implementation

placing double nodes. It is assumed that the displacements are continuous across thegrain boundaries, which is accomplished by tying the displacements of the double nodes.Keeping track of the crystals that share a grain boundary, the system of equations canbe assembled. Considering only edge dislocations, the final set of discretised linearisedequations for the unknown iterative corrections δ

˜u and δ

˜ρg reads[

Kuu Kuρ

Kρu Kρρ

] [δ˜u

δ˜ρg

]+

[0 0

Kgbρu Kgb

ρρ

] [δ˜u

δ˜ρg

]=

[˜fuext

˜fρext

]−[

˜fuint

˜fρint

]−[

˜0

˜fρgb

](A.4)

which is solved in a standard manner. The contributions of the grain boundary modelare clearly visible. In the case of a single grain boundary Γ separating grains A and B,the contributions have the following form

Kgbρu =

∑e

∫ΓeA

1

bNeT

ρ

(∂˜qA

∂˜Lu

)eBeu dΓeA +

∑e

∫ΓeB

1

bNeT

ρ

(∂˜qB

∂˜Lu

)eBeu dΓeB , (A.5)

Kgbρρ =

∑e

∫ΓeA

1

bNeT

ρ

((∂˜qA

∂˜ρg

)eNeρ +

(∂˜qA

∂˜Lρ

)eBeρ

)dΓeA+

∑e

∫ΓeB

1

bNeT

ρ

((∂˜qB

∂˜ρg

)eNeρ +

(∂˜qB

∂˜Lρ

)eBeρ

)dΓeB , (A.6)

˜fρgb =

∑e

∫ΓeA

1

bNeT

ρ˜qA dΓeA +

∑e

∫ΓeB

1

bNeT

ρ˜qB dΓeB . (A.7)

Page 132: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Appendix B

Data per rotation axis

The detailed information for the symmetric tilt grain boundaries with the 〈111〉, 〈100〉and 〈110〉 rotation axes is provided in Tables B.1, B.2 and B.3, respectively.

Table B.1: Data of the symmetric tilt grain boundaries with the [111] rotation axis in Al and Cu.Given are the misorientation angle θ (in degrees), the grain boundary energy φgb (in mJ/m2), the grainboundary plane, the parameter Σ characterizing the CSL, the boundary period vector p, the intrinsic netdefect Bp in a boundary period and the intrinsic net defect density |Bp|/|p|; a is the lattice parameter.

θ φgb Al φgb Cu GB plane Σ p Bp |Bp|/|p|0.00 0 0 (0 1 1) 1 a

2 [2 1 1] a[0 0 0] 0.007.30 270 437 (1 13 14) 183 a

2 [9 5 4] a2 [0 1 1] 0.13

17.90 419 710 (1 5 6) 31 a2 [11 7 4] 3a

2 [0 1 1] 0.3121.80 450 775 (1 4 5) 21 a

2 [3 2 1] a2 [0 1 1] 0.38

27.80 471 842 (1 3 4) 13 a2 [7 5 2] 3a

2 [0 1 1] 0.4832.20 489 889 (2 5 7) 39 a

2 [4 3 1] 2a2 [0 1 1] 0.55

38.20 461 867 (1 2 3) 7 a2 [5 4 1] 3a

2 [0 1 1] 0.6546.80 469 852 (2 3 5) 19 a

2 [8 7 1] 6a2 [0 1 1] 0.79

50.60 460 822 (3 4 7) 37 a2 [11 10 1] 9a

2 [0 1 1] 0.85

119

Page 133: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

120 Data per rotation axis

Table B.2: Data of the symmetric tilt grain boundaries with the [001] rotation axis in Al and Cu.Given are the misorientation angle θ (in degrees), the grain boundary energy φgb (in mJ/m2), the grainboundary plane, the parameter Σ characterizing the CSL, the boundary period vector p, the intrinsic netdefect Bp in a boundary period and the intrinsic net defect density |Bp|/|p|; a is the lattice parameter.

θ φgb Al φgb Cu GB plane Σ p Bp |Bp|/|p|0.00 0 0 (1 0 0) 1 a[0 1 0] a[0 0 0] 0.005.72 302 470 (20 1 0) 401 a[1 20 0] 2a[1 0 0] 0.108.17 362 578 (14 1 0) 197 a[1 14 0] 2a[1 0 0] 0.1411.42 419 686 (10 1 0) 101 a[1 10 0] 2a[1 0 0] 0.2014.25 454 757 (8 1 0) 65 a[1 8 0] 2a[1 0 0] 0.2518.92 483 837 (6 1 0) 37 a[1 6 0] 2a[1 0 0] 0.3322.62 489 878 (5 1 0) 13 a

2 [1 5 0] a[1 0 0] 0.3925.06 499 910 (9 2 0) 85 a[2 9 0] 4a[1 0 0] 0.4328.07 486 914 (4 1 0) 17 a[1 4 0] 2a[1 0 0] 0.4931.89 497 939 (7 2 0) 53 a[2 7 0] 4a[1 0 0] 0.5534.21 495 940 (13 4 0) 185 a[4 13 0] 8a[1 0 0] 0.5936.87 465 905 (3 1 0) 5 a

2 [1 3 0] 2a2 [1 1 0] 0.89

39.97 505 964 (11 4 0) 137 a[4 11 0] 14a2 [1 1 0] 0.85

43.60 513 983 (5 2 0) 29 a[2 5 0] 6a2 [1 1 0] 0.79

47.92 531 991 (9 4 0) 97 a[4 9 0] 10a2 [1 1 0] 0.72

50.40 529 984 (17 8 0) 353 a[8 17 0] 18a2 [1 1 0] 0.68

53.13 494 951 (2 1 0) 5 a[1 2 0] 2a2 [1 1 0] 0.63

56.14 506 941 (15 8 0) 289 a[8 15 0] 14a2 [1 1 0] 0.58

59.49 488 901 (7 4 0) 65 a[4 7 0] 6a2 [1 1 0] 0.53

61.93 466 856 (5 3 0) 17 a2 [3 5 0] 2a

2 [1 1 0] 0.4963.22 468 852 (13 8 0) 233 a[8 13 0] 10a

2 [1 1 0] 0.4667.38 433 790 (3 2 0) 13 a[2 3 0] 2a

2 [1 1 0] 0.3971.08 413 732 (7 5 0) 37 a

2 [5 7 0] 2a2 [1 1 0] 0.33

73.74 387 677 (4 3 0) 25 a[3 4 0] 2a2 [1 1 0] 0.28

77.32 351 595 (5 4 0) 41 a[4 5 0] 2a2 [1 1 0] 0.22

79.61 321 533 (6 5 0) 61 a[5 6 0] 2a2 [1 1 0] 0.18

81.20 296 484 (7 6 0) 85 a[6 7 0] 2a2 [1 1 0] 0.15

82.37 275 444 (8 7 0) 113 a[7 8 0] 2a2 [1 1 0] 0.13

83.27 257 411 (9 8 0) 145 a[8 9 0] 2a2 [1 1 0] 0.12

83.97 241 383 (10 9 0) 181 a[9 10 0] 2a2 [1 1 0] 0.11

90.00 0 0 (1 1 0) 1 a2 [1 1 0] a[0 0 0] 0.00

Page 134: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

121

Table B.3: Data of the symmetric tilt grain boundaries with the [110] rotation axis in Al and Cu.Given are the misorientation angle θ (in degrees), the grain boundary energy φgb (in mJ/m2), the grainboundary plane, the parameter Σ characterizing the CSL, the boundary period vector p, the intrinsic netdefect Bp in a boundary period and the intrinsic net defect density |Bp|/|p|; a is the lattice parameter.

θ φgb Al φgb Cu GB plane Σ p Bp |Bp|/|p|0.00 0 0 (0 0 1) 1 a

2 [1 1 0] a[0 0 0] 0.008.09 314 424 (1 1 20) 201 a[10 10 1] 2a[0 0 1] 0.1413.44 396 566 (1 1 12) 73 a[6 6 1] 2a[0 0 1] 0.2320.05 424 657 (1 1 8) 33 a[4 4 1] 2a[0 0 1] 0.3526.53 387 719 (1 1 6) 19 a[3 3 1] 2a[0 0 1] 0.4631.59 356 697 (1 1 5) 27 a

2 [5 5 2] 2a[0 0 1] 0.5438.94 331 652 (1 1 4) 9 a[2 2 1] 2a[0 0 1] 0.6744.00 302 600 (2 2 7) 57 a

2 [7 7 4] 4a[0 0 1] 0.7550.48 151 310 (1 1 3) 11 a

2 [3 3 2] 4a3 [1 1 1] 0.98

53.59 255 437 (5 5 14) 123 a[7 7 5] 18a3 [1 1 1] 0.94

58.99 320 527 (2 2 5) 33 a2 [5 5 4] 6a

3 [1 1 1] 0.8565.47 369 604 (5 5 11) 171 a

2 [11 11 10] 12a3 [1 1 1] 0.75

70.53 355 592 (1 1 2) 3 a[1 1 1] 2a3 [1 1 1] 0.67

77.88 414 661 (4 4 7) 81 a2 [7 7 8] 6a

3 [1 1 1] 0.5482.95 403 657 (5 5 8) 57 a[4 4 5] 6a

3 [1 1 1] 0.4686.63 370 633 (2 2 3) 17 a

2 [3 3 4] 2a3 [1 1 1] 0.40

93.37 344 594 (3 3 4) 17 a[2 2 3] 2a3 [1 1 1] 0.28

97.05 324 549 (4 4 5) 57 a2 [5 5 8] 2a

3 [1 1 1] 0.22102.12 275 427 (7 7 8) 81 a[4 4 7] 2a

3 [1 1 1] 0.13109.47 75 22 (1 1 1) 3 a

2 [1 1 2] a[0 0 0] 0.00114.53 229 292 (11 11 10) 171 a[5 5 11] 2a

3 [1 1 1] 0.09121.01 319 511 (5 5 4) 33 a[2 2 5] 2a

3 [1 1 1] 0.20126.41 377 655 (7 7 5) 123 a

2 [5 5 14] 4a3 [1 1 1] 0.29

129.52 388 702 (3 3 2) 11 a[1 1 3] 2a3 [1 1 1] 0.35

131.55 419 760 (11 11 7) 291 a2 [7 7 22] 8a

3 [1 1 1] 0.38136.00 452 829 (7 7 4) 57 a[2 2 7] 6a

3 [1 1 1] 0.46141.06 447 834 (2 2 1) 9 a

2 [1 1 4] 2a3 [1 1 1] 0.54

144.36 494 855 (11 11 5) 267 a2 [5 5 22] 10a

2 [1 1 0] 0.61148.41 486 858 (5 5 2) 27 a[1 1 5] 4a

2 [1 1 0] 0.54153.47 431 845 (3 3 1) 19 a

2 [1 1 6] 2a2 [1 1 0] 0.46

159.95 416 795 (4 4 1) 33 a2 [1 1 8] 2a

2 [1 1 0] 0.35166.56 361 678 (6 6 1) 73 a

2 [1 1 12] 2a2 [1 1 0] 0.23

171.91 285 505 (10 10 1) 201 a2 [1 1 20] 2a

2 [1 1 0] 0.14180.00 0 0 (1 1 0) 1 a[0 0 1] a[0 0 0] 0.00

Page 135: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary
Page 136: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Bibliography

[1] K.E. Aifantis and A.H.W. Ngan. Modeling dislocation-grain boundary interactionsthrough gradient plasticity and nanoindentation. Materials Science and Engineer-ing A, 459:251–261, 2007.

[2] K.E. Aifantis and J.R. Willis. The role of interfaces in enhancing the yield strengthof composites and polycrystals. Journal of the Mechanics and Physics of Solids,53:1047–1070, 2005.

[3] K.E. Aifantis and J.R. Willis. Scale effects induced by strain-gradient plasticity andinterfacial resistance in periodic and randomly heterogeneous media. Mechanics ofMaterials, 38:702–716, 2006.

[4] S. Amelinckx and W. Dekeyser. The structure and properties of grain boundaries.Solid State Physics, 8:325–499, 1959.

[5] M.F. Ashby. The deformation of plastically non-homogeneous materials. Philo-sophical Magazine, 21:399–424, 1970.

[6] W.M. Ashmawi and M.A. Zikry. Prediction of grain-boundary interfacial mecha-nisms in polycrystalline materials. Journal of Engineering Materials and Technol-ogy, 124:88–96, 2002.

[7] A. Atkinson. Grain boundary diffusion - Structural effects and mechanisms. Journalde Physique (Colloque C4), 46:397–391, 1985.

[8] D.V. Bachurin, R.T. Murzaev, and A.A. Nazarov. Atomistic computer and discli-nation simulation of [001] tilt boundaries in nickel and copper. The Physics ofMetals and Metallography, 96:555–561, 2003.

[9] D.V. Bachurin, D. Weygand, and P. Gumbsch. Dislocation-grain boundary inter-action in 〈111〉 textured thin metal films. Acta Materialia, 58:5232–5241, 2010.

[10] D.S. Balint, V.S. Deshpande, A. Needleman, and E. van der Giessen. Discretedislocation plasticity analysis of the grain size dependence of the flow strength ofpolycrystals. International Journal of Plasticity, 24:2149–2172, 2008.

123

Page 137: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

124 Bibliography

[11] R.W. Balluffi and P.D. Bristowe. On the structural unit/grain boundary dislocationmodel for grain boundary structure. Surface Science, 144:28–43, 1984.

[12] R.W. Balluffi and G.B. Olson. On the hierarchy of interfacial dislocation structure.Metallurgical Transactions A, 16:529–541, 1985.

[13] R.W. Balluffi, Y. Komem, and T. Schober. Electron microscope studies of grainboundary dislocation behavior. Surface Science, 31:68–103, 1972.

[14] P.W. Bates and J. Han. The Neumann boundary problem for a nonlocal Cahn-Hilliard equation. Journal of Differential Equations, 212:235–277, 2005.

[15] Z.P. Bazant and M. Jirasek. Nonlocal integral formulations of plasticity and dam-age: Survey of progress. Journal of Engineering Mechanics, 128:1119–1149, 2002.

[16] C.J. Bayley, W.A.M. Brekelmans, and M.G.D. Geers. A comparison of dislocationinduced back stress formulations in strain gradient crystal plasticity. InternationalJournal of Solids and Structures, 43:7268–7286, 2006.

[17] S.B. Biner and J.R. Morris. A two-dimensional discrete dislocation simulation ofthe effect of grain size on strengthening behavior. Modelling and Simulation inMaterials Science and Engineering, 10:617–635, 2002.

[18] G.H. Bishop and B. Chalmers. A coincidence - ledge - dislocation description ofgrain boundaries. Scripta Metallurgica, 2:133–140, 1968.

[19] G.H. Bishop and B. Chalmers. Dislocation structure and contrast in high anglegrain boundaries. Philosophical Magazine, 24:515–526, 1971.

[20] W. Bollmann. Crystal Defects and Crystalline Interfaces. Springer, Berlin, 1970.

[21] U. Borg. A strain gradient crystal plasticity analysis of grain size effects in poly-crystals. European Journal of Mechanics A/Solids, 26:313–324, 2007.

[22] L. Brand. Vector and Tensor Analysis. John Wiley & Sons, New York, 1947.

[23] T.B. Britton, D. Randman, and A.J. Wilkinson. Nanoindentation study of sliptransfer phenomenon at grain boundaries. Journal of Materials Research, 24:607–615, 2009.

[24] V.V. Bulatov, B.W. Reed, and M. Kumar. Grain boundary energy function for fccmetals. Acta Materialia, 65:161–175, 2014.

[25] J.W. Cahn. On spinodal decomposition. Acta Metallurgica, 9:795–801, 1961.

[26] J.W. Cahn and J.E. Hilliard. Free energy of a nonuniform system. I. Interfacialfree energy. The Journal of Chemical Physics, 28:258–267, 1958.

[27] J.W. Cahn, Y. Mishin, and A. Suzuki. Duality of dislocation content of grainboundaries. Philosophical Magazine, 86:3965–3980, 2006.

[28] J.W. Cahn, Y. Mishin, and A. Suzuki. Coupling grain boundary motion to sheardeformation. Acta Materialia, 54:4953–4975, 2006.

Page 138: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

125

[29] L. Capolungo, D.E. Spearot, M. Cherkaoui, D.L. McDowell, J. Qu, and K.I. Jacob.Dislocation nucleation from bicrystal interfaces and grain boundary ledges: Rela-tionship to nanocrystalline deformation. Journal of the Mechanics and Physics ofSolids, 55:2300–2327, 2007.

[30] P. Cermelli and M.E. Gurtin. Geometrically necessary dislocations in viscoplasticsingle crystals and bicrystals undergoing small deformations. International Journalof Solids and Structures, 39:6281–6309, 2002.

[31] L.Q. Chen. Phase-field models for microstructure evolution. Annual Review ofMaterials Research, 32:113–140, 2002.

[32] Y. Cheng, M. Mrovec, and P. Gumbsch. Atomistic simulations of interactions be-tween the 1/2〈111〉 edge dislocation and symmetric tilt grain boundaries in tung-sten. Philosophical Magazine, 88:547–560, 2008.

[33] J.W. Christian. The Theory of Transformations in Metals and Alloys. Pergamon,Oxford, 2002.

[34] W.A.T. Clark and D.A. Smith. Interaction of lattice dislocations with periodicgrain boundary structures. Journal of Materials Science, 14:776–788, 1979.

[35] J.D. Clayton, D.J. Bammann, and D.L. McDowell. A geometric framework for thekinematics of crystals with defects. Philosophical Magazine, 85:3983–4010, 2005.

[36] J.D. Clayton, D.L. McDowell, and D.J. Bammann. Modeling dislocations anddisclinations with finite micropolar elastoplasticity. International Journal of Plas-ticity, 22:210–256, 2006.

[37] H. Conrad. Grain-size dependence of the flow stress of Cu from millimeters tonanometers. Metallurgical and Materials Transactions A, 35:2681–2695, 2004.

[38] N.M. Cordero, S. Forest, E.P. Busso, S. Berbenni, and M. Cherkaoui. Grain sizeeffects on plastic strain and dislocation density tensor fields in metal polycrystals.Computational Materials Science, 52:7–13, 2012.

[39] F. Cosandey and C.L. Bauer. Characterization of [001] tilt boundaries in gold byhigh-resolution transmission electron microscopy. Philosophical Magazine A, 44:391–403, 1981.

[40] J.P. Couzinie, B. Decamps, and L. Priester. On the interactions between dislo-cations and a near-Σ = 3 grain boundary in a low stacking-fault energy metal.Philosophical Magazine Letters, 83:721–731, 2003.

[41] C.F.O. Dahlberg, J. Faleskog, C.F. Niordson, and B.N. Legarth. A deformationmechanism map for polycrystals modeled using strain gradient plasticity and in-terfaces that slide and separate. International Journal of Plasticity, 43:177–195,2013.

[42] M. Dao, L. Lu, R.J. Asaro, J.T.M. de Hosson, and E. Ma. Toward a quantitativeunderstanding of mechanical behavior of nanocrystalline metals. Acta Materialia,55:4041–4065, 2007.

Page 139: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

126 Bibliography

[43] T.P. Darby, R. Schindler, and R.W. Balluffi. On the interaction of lattice disloca-tions with grain boundaries. Philosophical Magazine A, 37:245–256, 1978.

[44] M. de Koning, R. Miller, V.V. Bulatov, and F.F. Abrahams. Modeling grain-boundary resistance in intergranular dislocation slip transmission. PhilosophicalMagazine A, 82:2511–2527, 2002.

[45] M. de Koning, R.J. Kurtz, V.V. Bulatov, C.H. Henager, R.G. Hoagland, W. Cai,and M. Nomura. Modeling of dislocation-grain boundary interactions in FCC met-als. Journal of Nuclear Materials, 323:281–289, 2003.

[46] M. Dewald and W.A. Curtin. Multiscale modeling of dislocation/grain-boundaryinteractions: III. 60◦ dislocations impinging on Σ3, Σ9 and Σ11 tilt boundaries inAl. Modelling and Simulation in Materials Science and Engineering, 19:055002–1–36, 2011.

[47] M.P. Dewald and W.A. Curtin. Multiscale modelling of dislocation/grain-boundaryinteractions: I. Edge dislocations impinging on Σ11 (113) tilt boundary in Al.Modelling and Simulation in Materials Science and Engineering, 15:S193–S215,2007.

[48] M.P. Dewald and W.A. Curtin. Multiscale modelling of dislocation/grain boundaryinteractions. II. Screw dislocations impinging on tilt boundaries in Al. PhilosophicalMagazine, 87:4615–4641, 2007.

[49] D.J. Dingley and R.C. Pond. On the interaction of crystal dislocations with grainboundaries. Acta Metallurgica, 27:667–682, 1979.

[50] O.H. Duparc, S. Poulat, A. Larere, J. Thibault, and L. Priester. High-resolutiontransmission electron microscopy observations and atomic simulations of the struc-tures of exact and near Σ = 11, {332} tilt grain boundaries in nickel. PhilosophicalMagazine A, 80:853–870, 2000.

[51] M. Ekh, S. Bargmann, and M. Grymer. Influence of grain boundary conditions onmodeling of size-dependence in polycrystals. Acta Mechanica, 218:103–113, 2011.

[52] M. Elkajbaji and J. Thibault-Desseaux. Interactions of deformation-induced dislo-cations with Σ = 9 (122) grain-boundaries in Si studied by HREM. PhilosophicalMagazine A, 58:325–345, 1988.

[53] R.A.B. Engelen, M.G.D. Geers, and F.P.T. Baaijens. Nonlocal implicit gradient-enhanced elasto-plasticity for the modelling of softening behaviour. InternationalJournal of Plasticity, 19:403–433, 2003.

[54] I. Erturk, J.A.W. van Dommelen, and M.G.D. Geers. Energetic dislocation inter-actions and thermodynamical aspects of strain gradient crystal plasticity theories.Journal of the Mechanics and Physics of Solids, 57:1801–1814, 2009.

[55] U. Essmann and H. Mughrabi. Annihilation of dislocations during tensile andcyclic deformation and limits of dislocation densities. Philosophical Magazine A,40:731–756, 1979.

Page 140: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

127

[56] L.P. Evers, W.A.M. Brekelmans, and M.G.D. Geers. Non-local crystal plasticitymodel with intrinsic SSD and GND effects. Journal of the Mechanics and Physicsof Solids, 52:2379–2401, 2004.

[57] L.P. Evers, W.A.M. Brekelmans, and M.G.D. Geers. Scale dependent crystal plas-ticity framework with dislocation density and grain boundary effects. InternationalJournal of Solids and Structures, 41:5209–5230, 2004.

[58] H.F. Fischmeister. Structure and properties of high angle boundaries. Journal dePhysique (Colloque C4), 46:3–23, 1985.

[59] N.A. Fleck and J.R. Willis. A mathematical basis for strain-gradient plasticitytheory. Part I: Scalar plastic multiplier. Journal of the Mechanics and Physics ofSolids, 57:161–177, 2009.

[60] C.T. Forwood and L.M. Clarebrough. Electron Microscopy of Interfaces in Metalsand Alloys. Adam Hilger, New York, 1991.

[61] P. Franciosi and A. Zaoui. Multislip in FCC crystals: a theoretical approachcompared with experimental data. Acta Metallurgica, 30:1627–1637, 1982.

[62] P. Fredriksson and P. Gudmundson. Size-dependent yield strength and surfaceenergies of thin films. Materials Science and Engineering A, 400-401:448–450,2005.

[63] P. Fredriksson and P. Gudmundson. Size-dependent yield strength of thin films.International Journal of Plasticity, 21:1834–1854, 2005.

[64] P. Fredriksson and P. Gudmundson. Modelling of the interface between a thinfilm and a substrate within a strain gradient plasticity framework. Journal of theMechanics and Physics of Solids, 55:939–955, 2007.

[65] C. Fressengeas, V. Taupin, and L. Capolungo. An elasto-plastic theory of dislo-cation and disclination fields. International Journal of Solids and Structures, 48:3499–3509, 2011.

[66] C. Fressengeas, V. Taupin, and L. Capolungo. Continuous modeling of the structureof symmetric tilt boundaries. International Journal of Solids and Structures, 51:1434–1441, 2014.

[67] J. Gemperlova, A. Jacques, A. Gemperle, T. Vystavel, N. Zarubova, andM. Janecek. In-situ transmission electron microscopy observation of slip propa-gation in Σ3 bicrystals. Materials Science and Engineering A, 342:183–189, 2002.

[68] V.Y. Gertsman, A.A. Nazarov, A.E. Romanov, R.Z. Valiev, and V.I. Vladimirov.Disclination-structural unit model of grain boundaries. Philosophical Magazine A,59:1113–1118, 1989.

[69] G. Giacomin and J.L. Lebowitz. Phase segregation dynamics in particle systemswith long range interactions. I. Macroscopic limits. Journal of Statistical Physics,87:37–61, 1997.

Page 141: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

128 Bibliography

[70] N.A. Gjostein and F.N. Rhines. Absolute interfacial energies of [001] tilt and twistgrain boundaries in copper. Acta Metallurgica, 7:319–330, 1959.

[71] M.W. Grabski. Mobility of dislocations in grain boundaries. Journal de Physique(Colloque C5), 49:497–506, 1988.

[72] J.R. Greer and J.T.M. de Hosson. Plasticity in small-sized metallic systems: In-trinsic versus extrinsic size effect. Progress in Materials Science, 56:654–724, 2011.

[73] H. Grimmer, W. Bollmann, and D.H. Warrington. Coincidence-site lattices andcomplete pattern-shift lattices in cubic crystals. Acta Crystallographica A, 30:197–207, 1974.

[74] P. Gudmundson. A unified treatment of strain gradient plasticity. Journal of theMechanics and Physics of Solids, 52:1379–1406, 2004.

[75] M.E. Gurtin. Generalized Ginzburg-Landau and Cahn-Hilliard equations based ona microforce balance. Physica D, 92:178–192, 1996.

[76] M.E. Gurtin. A gradient theory of single-crystal viscoplasticity that accounts forgeometrically necessary dislocations. Journal of the Mechanics and Physics ofSolids, 50:5–32, 2002.

[77] M.E. Gurtin. A theory of grain boundaries that accounts automatically for grainmisorientation and grain-boundary orientation. Journal of the Mechanics andPhysics of Solids, 56:640–662, 2008.

[78] M.E. Gurtin and A.N. Needleman. Boundary conditions in small-deformation, sin-gle crystal plasticity that account for the burgers vector. Journal of the Mechanicsand Physics of Solids, 53:1–31, 2005.

[79] G. Hasson, J.Y. Boos, I. Herbeuval, M. Biscondi, and C. Goux. Theoretical and ex-perimental determinations of grain boundary structures and energies: Correlationwith various experimental results. Surface Science, 31:115–137, 1972.

[80] J.P. Hirth. The influence of grain boundaries on mechanical properties. Metallur-gical Transactions, 3:3047–3067, 1972.

[81] J.P. Hirth and R.W. Balluffi. On grain boundary dislocations and ledges. ActaMetallurgica, 21:929–942, 1973.

[82] J.P. Hirth and J. Lothe. Theory of Dislocations. Krieger Publishing Company,Florida, 1982.

[83] J.P. Hirth, R.C. Pond, R.G. Hoagland, X.Y. Liu, and J. Wang. Interface defects,reference spaces and the Frank-Bilby equation. Progress in Materials Science, 58:749–823, 2013.

[84] E.A. Holm, D.L. Olmsted, and S.M. Foiles. Comparing grain boundary energiesin face-centered cubic metals: Al, Au, Cu and Ni. Scripta Materialia, 63:905–908,2010.

Page 142: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

129

[85] J.M. Howe. Interfaces in Materials: Atomic Structure, Kinetics and Thermody-namics of Solid-Vapor, Solid-Liquid and Solid-Solid Interfaces. John Wiley & Sons,New York, 1997.

[86] D. Hull and D.J. Bacon. Introduction to Dislocations. Butterworth-Heinemann,Oxford, 2001.

[87] Y. Ishida and M. Mori. Theoretical studies of segregated internal interfaces. Journalde Physique (Colloque C4), 46:465–474, 1985.

[88] Y. Ishida, T. Hasegawa, and F. Nagata. Dislocation images on grain boundaryand their behavior at elevated temperatures. Transactions of the Japan Instituteof Metals (Supplement), 9:504–508, 1968.

[89] Z.H. Jin, P. Gumbsch, E. Ma, K. Albe, K. Lu, H. Hahn, and H. Gleiter. The inter-action mechanism of screw dislocations with coherent twin boundaries in differentface-centered cubic metals. Scripta Materialia, 54:1163–1168, 2006.

[90] Z.H. Jin, P. Gumbsch, K. Albe, E. Ma, K. Lu, H. Gleiter, and H. Hahn. In-teractions between non-screw lattice dislocations and coherent twin boundaries inface-centered cubic metals. Acta Materialia, 56:1126–1135, 2008.

[91] T. Johannesson and A. Tholen. The role of grain boundaries in creep deformation.Metal Science Journal, 6:189–194, 1972.

[92] H.K. Kim and S.I. Oh. An interfacial energy incorporated couple stress crystalplasticity and the finite element simulation of grain subdivision. Journal of theMechanics and Physics of Solids, 60:1815–1841, 2012.

[93] S. Kobayashi, S. Tsurekawa, and T. Watanabe. Grain boundary hardening andtriple junction hardening in polycrystalline molybdenum. Acta Materialia, 53:1051–1057, 2005.

[94] O. Kraft, P.A. Gruber, R. Monig, and D. Weygand. Plasticity in confined dimen-sions. Annual Review of Materials Research, 40:293–317, 2010.

[95] M. Kuroda and V. Tvergaard. On the formulations of higher-order strain gradientcrystal plasticity models. Journal of the Mechanics and Physics of Solids, 56:1591–1608, 2008.

[96] R.J. Kurtz, R.G. Hoagland, and J.P. Hirth. Computer simulation of extrinsic grain-boundary defects in the Σ11, 〈101〉{131} symmetric tilt boundary. PhilosophicalMagazine A, 79:683–703, 1990.

[97] E.P. Kvam and R.W. Balluffi. Observations of hierarchical grain-boundary disloca-tion structures in [001] symmetric tilt boundaries in gold. Philosophical MagazineA, 56:137–148, 1987.

[98] J.W. Kysar, Y.X. Gan, and G. Mendez-Arzuza. Cylindrical void in a rigid-ideallyplastic single crystal. Part I: Anisotropic slip line theory solution for face-centeredcubic crystals. International Journal of Plasticity, 21:1481–1520, 2005.

Page 143: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

130 Bibliography

[99] C.S. Lee, G.W. Han, R.E. Smallman, D. Feng, and J.K.L. Lai. The influence ofboron-doping on the effectiveness of grain boundary hardening in Ni3Al. ActaMaterialia, 47:1823–1830, 1999.

[100] T.C. Lee, I.M. Robertson, and H.K. Birnbaum. Prediction of slip transfer mecha-nisms across grain boundaries. Scripta Metallurgica, 23:799–803, 1989.

[101] T.C. Lee, I.M. Robertson, and H.K. Birnbaum. An in situ transmission electronmicroscope deformation study of the slip transfer mechanisms in metals. Metallur-gical Transactions A, 21:2437–2447, 1990.

[102] T.C. Lee, I.M. Robertson, and H.K. Birnbaum. TEM in situ deformation study ofthe interaction of lattice dislocations with grain boundaries in metals. PhilosophicalMagazine, 62:131–153, 1990.

[103] S. Lefebvre, B. Devincre, and T. Hoc. Yield stress strengthening in ultrafine-grained metals: A two-dimensional simulation of dislocation dynamics. Journal ofthe Mechanics and Physics of Solids, 55:788–802, 2007.

[104] J.C.M. Li. Disclination model of high angle grain boundaries. Surface Science, 31:12–26, 1972.

[105] J.C.M. Li and Y.T. Chou. The role of dislocations in the flow stress grain sizerelationships. Metallurgical Transactions, 1:1145–1159, 1970.

[106] Z. Li, C. Hou, M. Huang, and C. Ouyang. Strengthening mechanism in micro-polycrystals with penetrable grain boundaries by discrete dislocation dynamic sim-ulation and Hall-Petch effect. Computational Materials Science, 46:1124–1134,2009.

[107] B. Liu, D. Raabe, P. Eisenlohr, F. Roters, A. Arsenlis, and G. Hommes. Dislocationinteractions and low-angle grain boundary strengthening. Acta Materialia, 59:7125–7134, 2011.

[108] B. Liu, P. Eisenlohr, F. Roters, and D. Raabe. Simulation of dislocation penetrationthrough a general low-angle grain boundary. Acta Materialia, 60:5380–5390, 2012.

[109] J.S. Liu and R.W. Balluffi. Grain boundary dislocation structures in symmetric[001] tilt boundaries in aluminum. Philosophical Magazine A, 52:713–720, 1985.

[110] Z.L. Liu, Z. Zhuang, X.M. Liu, X.C. Zhao, and Z.H. Zhang. A dislocation dynamicsbased higher-order crystal plasticity model and applications on confined thin-filmplasticity. International Journal of Plasticity, 27:201–216, 2011.

[111] J.D. Livingston and B. Chalmers. Multiple slip in bicrystal deformation. ActaMetallurgica, 5:322–327, 1957.

[112] W. Lojkowski. On the spreading of grain boundary dislocations and its effect ongrain boundary properties. Acta Metallurgica et Materialia, 39:1891–1899, 1991.

Page 144: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

131

[113] A. Ma, F. Roters, and D. Raabe. On the consideration of interactions betweendislocations and grain boundaries in crystal plasticity finite element modeling -Theory, experiments and simulations. Acta Materialia, 54:2181–2194, 2006.

[114] T.J. Massart and T. Pardoen. Strain gradient plasticity analysis of the grain-size-dependent strength and ductility of polycrystals with evolving grain boundaryconfinement. Acta Materialia, 58:5768–5781, 2010.

[115] J.R. Mayeur, D.L. McDowell, and D.J. Bammann. Dislocation-based micropolarsingle crystal plasticity: Comparison of multi- and single criterion theories. Journalof the Mechanics and Physics of Solids, 59:398–422, 2011.

[116] D.L. McDowell. Viscoplasticity of heterogeneous metallic materials. MaterialsScience and Engineering R, 62:67–123, 2008.

[117] D.L. McDowell. A perspective on trends in multiscale plasticity. InternationalJournal of Plasticity, 26:1280–1309, 2010.

[118] D.L. Medlin, C.B. Carter, J.E. Angelo, and M.J. Mills. Climb and glide of a/3〈111〉dislocations in an aluminium Σ = 3 boundary. Philosophical Magazine A, 75:733–747, 1997.

[119] D.L. Medlin, S.M. Foiles, and D. Cohen. A dislocation-based description of grainboundary dissociation: Application to a 90◦ 〈110〉 tilt boundary in gold. ActaMaterialia, 49:3689–3697, 2001.

[120] K.L. Merkle. High-resolution electron microscopy of grain boundaries. InterfaceScience, 2:311–345, 1995.

[121] M.A. Meyers, A. Mishra, and D.J. Benson. Mechanical properties of nanocrystallinematerials. Progress in Materials Science, 51:427–556, 2006.

[122] M.J. Mills, M.S. Daw, G.J. Thomas, and F. Cosandey. High-resolution transmissionelectron microscopy of grain boundaries in aluminum and correlation with atomisticcalculations. Ultramicroscopy, 40:247–257, 1992.

[123] Y. Mishin, D. Farkas, M.J. Mehl, and D.A. Papaconstantopoulos. Interatomicpotentials for monoatomic metals from experimental data and ab initio calculations.Physical Review B, 59:3393–3407, 1999.

[124] Y. Mishin, M.J. Mehl, D.A. Papaconstantopoulos, A.F. Voter, and J.D. Kress.Structural stability and lattice defects in copper: Ab initio, tight-binding, andembedded-atom calculations. Physical Review B, 63:224106–1–16, 2001.

[125] N. Moelans, B. Blanpain, and P. Wollants. An introduction to phase-field modelingof microstructure evolution. Computer Coupling of Phase Diagrams and Thermo-chemistry, 32:268–294, 2008.

[126] T. Mori and K. Tangri. Gliding of boundary dislocations on coincidence boundaries.Metallurgical Transactions A, 10:733–740, 1979.

Page 145: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

132 Bibliography

[127] T. Mori, H. Miura, T. Tokita, J. Haji, and M. Kato. Determination of the en-ergies of [001] twist boundaries in Cu with the shape of boundary SiO2 particles.Philosophical Magazine Letters, 58:11–15, 1988.

[128] A.A. Nazarov. A revision of the models for the accommodation of extrinsic grainboundary dislocations. Interface Science, 8:71–76, 2000.

[129] A.A. Nazarov, A.E. Romanov, and R.Z. Valiev. On the hierarchy of dislocationdescriptions of grain boundary structures. Physica Status Solidi (a), 122:495–502,1990.

[130] A.A. Nazarov, O.A. Shenderova, and D.W. Brenner. On the disclination-structuralunit model of grain boundaries. Materials Science and Engineering A, 281:148–155,2000.

[131] D.L. Olmsted, S.M. Foiles, and E.A. Holm. Survey of computed grain boundaryproperties in face-centered cubic metals: I. Grain boundary energy. Acta Materi-alia, 57:3694–3703, 2009.

[132] A. Otsuki. Energies of [001] small angle grain boundaries in aluminium. Journalof Materials Science, 40:3219–3223, 2005.

[133] A. Otsuki and M. Mizuno. Grain boundary energy and liquid metal embrittlementof aluminium. Transactions of the Japan Institute of Metals (Supplement), 27:789–796, 1986.

[134] I. Ozdemir and T. Yalcinkaya. Modeling of dislocation-grain boundary interactionsin a strain gradient crystal plasticity framework. Computational Mechanics, 54:255–268, 2014.

[135] R.H.J. Peerlings, R. de Borst, W.A.M. Brekelmans, and J.H.P. de Vree. Gradientenhanced damage for quasi-brittle materials. International Journal for NumericalMethods in Engineering, 39:3391–3403, 1996.

[136] R.H.J. Peerlings, M.G.D. Geers, R. de Borst, and W.A.M. Brekelmans. A criticalcomparison of nonlocal and gradient-enhanced softening continua. InternationalJournal of Solids and Structures, 38:7723–7746, 2001.

[137] O. Penrose and P.C. Fife. Thermodynamically consistent models of phase-fieldtype for the kinetics of phase transitions. Physica D, 43:44–62, 1990.

[138] B.J. Pestman, J.T.M. de Hosson, V. Vitek, and F.W. Schapink. Interaction betweenlattice dislocations and grain boundaries in f.c.c. materials. Scripta Metallurgica,23:1431–1436, 1989.

[139] B.J. Pestman, J.T.M. de Hosson, V. Vitek, and F.W. Schapink. Interaction be-tween lattice dislocations and grain boundaries in f.c.c. and ordered compounds: acomputer simulation. Philosophical Magazine A, 64:951–969, 1991.

[140] R.B. Phillips. Crystals, Defects and Microstructures: Modeling Across Scales. Cam-bridge University Press, Cambridge, 2001.

Page 146: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

133

[141] C. Polizzotto. Interfacial energy effects within the framework of strain gradientplasticity. International Journal of Solids and Structures, 46:1685–1694, 2009.

[142] R.C. Pond and D.A. Smith. On the absorption of dislocations by grain boundaries.Philosophical Magazine, 36:353–366, 1977.

[143] S. Poulat, B. Decamps, and L. Priester. In-situ transmission electron microscopystudy of the dislocation accommodation in [101] tilt grain boundaries in nickelbicrystals. Philosophical Magazine A, 79:2655–2680, 1999.

[144] L. Priester. On the accommodation of extrinsic dislocations in grain boundaries.Interface Science, 4:205–219, 1997.

[145] L. Priester. Grain Boundaries: From Theory to Engineering. Springer, Dordrecht,2013.

[146] N. Provatas and K. Elder. Phase-Field Methods in Materials Science and Engi-neering. Wiley-VCH, Weinheim, 2010.

[147] P.H. Pumphrey and H. Gleiter. Annealing of dislocations in high-angle grain-boundaries. Philosophical Magazine, 30:593–602, 1974.

[148] V. Randle. The Role of the Coincidence Site Lattice in Grain Boundary Engineer-ing. Institute of Materials, London, 1996.

[149] V. Randle. The role of the grain boundary plane in cubic polycrystals. ActaMaterialia, 46:1495–1480, 1997.

[150] W.T. Read. Dislocations in Crystals. McGraw-Hill, London, 1953.

[151] W.T. Read and W. Shockley. Dislocation models of crystal grain boundaries.Physical Review, 78:275–289, 1950.

[152] J.R. Rice. Tensile crack tip fields in elastic-ideally plastic crystals. Mechanics ofMaterials, 6:317–335, 1987.

[153] J.D. Rittner and D.N. Seidman. 〈110〉 symmetric tilt grain-boundary structuresin fcc metals with low stacking-fault energies. Physical Review B, 54:6999–7015,1996.

[154] G.S. Rohrer. Grain boundary energy anisotropy: a review. Journal of MaterialsScience, 46:5881–5895, 2011.

[155] M.D. Sangid, T. Ezaz, H. Sehitoglu, and I.M. Robertson. Energy of slip transmis-sion and nucleation at grain boundaries. Acta Materialia, 59:283–296, 2011.

[156] M.D. Sangid, T. Ezaz, and H. Sehitoglu. Energetics of residual dislocations associ-ated with slip-twin and slip-GBs interactions. Materials Science and EngineeringA, 542:21–30, 2012.

[157] D. Saraev and S. Schmauder. Atomic-scale simulations of the interaction betweendislocations and tilt grain boundaries in α-iron. Physica Status Solidi (b), 240:81–90, 2003.

Page 147: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

134 Bibliography

[158] Z. Shen, R.H. Wagoner, and W.A.T. Clark. Dislocation pile-up and grain boundaryinteractions in 304 stainless steel. Scripta Metallurgica, 20:921–926, 1986.

[159] Z. Shen, R.H. Wagoner, and W.A.T. Clark. Dislocation and grain boundary inter-actions in metals. Acta Metallurgica, 36:3231–3242, 1988.

[160] V.B. Shenoy, R. Miller, E.B. Tadmor, R. Phillips, and M. Ortiz. Quasicontinuummodels of interfacial structure and deformation. Physical Review Letters, 80:742–745, 1998.

[161] A.C. Shi, C. Rottman, and Y. He. Calculation of energy of low-angle grain bound-aries. Philosophical Magazine A, 55:499–511, 1987.

[162] K.K. Shih and J.C.M. Li. Energy of grain boundaries between cusp misorientations.Surface Science, 50:109–124, 1975.

[163] L.E. Shilkrot, R.E. Miller, and W.A. Curtin. Multiscale plasticity modeling: cou-pled atomistics and discrete dislocation mechanics. Journal of the Mechanics andPhysics of Solids, 52:755–787, 2004.

[164] T. Shimokawa, T. Kinari, and S. Shintaku. Interaction mechanism between edgedislocations and asymmetrical tilt grain boundaries investigated via quasicontin-uum simulations. Physical Review B, 75:144108–1–11, 2007.

[165] S.S. Shishvan, S. Mohammadi, M. Rahimian, and E. van der Giessen. Plane-straindiscrete dislocation plasticity incorporating anisotropic elasticity. InternationalJournal of Solids and Structures, 48:374–387, 2011.

[166] D.A. Smith. Interaction of dislocations with grain boundaries. Journal de Physique(Colloque C6), 43:225–237, 1982.

[167] W.A. Soer, K.E. Aifantis, and J.T.M. de Hosson. Incipient plasticity during nanoin-dentation at grain boundaries in body-centered cubic materials. Acta Materialia,53:4665–4676, 2005.

[168] Y.M. Soifer, A. Verdyan, M. Kazakevich, and E. Rabkin. Nanohardness of copperin the vicinity of grain boundaries. Scripta Materialia, 47:799–804, 2002.

[169] P. Steinmann. On boundary potential energies in deformational and configurationalmechanics. Journal of the Mechanics and Physics of Solids, 56:772–800, 2008.

[170] S. Sun, B.L. Adams, C. Shet, S. Saigal, and W. King. Mesoscale investigation ofthe deformation field of an aluminum bicrystal. Scripta Materialia, 39:501–508,1998.

[171] S. Sun, B.L. Adams, and W.E. King. Observations of lattice curvature near theinterface of a deformed aluminium bicrystal. Philosophical Magazine, 80:9–25, 2000.

[172] T. Surholt and C. Herzig. Grain boundary self-diffusion in Cu polycrystals ofdifferent purity. Acta Materialia, 45:3817–3823, 1997.

Page 148: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

135

[173] A.P. Sutton. On the structural unit model of grain boundary structure. Philosoph-ical Magazine Letters, 59:53–59, 1989.

[174] A.P. Sutton and R.W. Balluffi. On geometric criteria for low interfacial energy.Acta Metallurgica, 35:2177–2201, 1987.

[175] A.P. Sutton and R.W. Balluffi. Interfaces in Crystalline Materials. Oxford Uni-versity Press, New York, 1995.

[176] A.P. Sutton and V. Vitek. On the structure of tilt grain boundaries in cubic metalsI. Symmetric tilt boundaries. Philosophical Transactions of the Royal Society ofLondon A, 309:1–36, 1983.

[177] E.B. Tadmor and R.E. Miller. Modeling Materials: Continuum, Atomistic andMultiscale Techniques. Cambridge University Press, Cambridge, 2011.

[178] V. Taupin, L. Capolungo, C. Fressengeas, A. Das, and M. Upadhyay. Grain bound-ary modeling using an elasto-plastic theory of dislocation and disclination fields.Journal of the Mechanics and Physics of Solids, 61:370–384, 2013.

[179] M.A. Tschopp and D.L. McDowell. Structures and energies of Σ3 asymmetric tiltgrain boundaries in copper and aluminium. Philosophical Magazine, 87:3147–3173,2007.

[180] M.A. Tschopp and D.L. McDowell. Asymmetric tilt grain boundary structure andenergy in copper and aluminium. Philosophical Magazine, 87:3871–3892, 2007.

[181] M.A. Tschopp, D.E. Spearot, and D.L. McDowell. Influence of grain boundarystructure on dislocation nucleation in FCC metals. In J. Hirth, editor, Dislocationsin Solids, volume 14, pages 43–139. North-Holland, Amsterdam, 2008.

[182] S. Tsurekawa, K. Morita, H. Nakashima, and H. Yoshinaga. Geometric structures ofgrain boundaries expected from the O-lattice theory compared with high-resolutiontransmission electron microscope images. Materials Transactions (JIM), 38:393–400, 1997.

[183] G.J. Tucker and D.L. McDowell. Non-equilibrium grain boundary structure and in-elastic deformation using atomistic simulations. International Journal of Plasticity,27:841–857, 2011.

[184] G.J. Tucker, M.A. Tschopp, and D.L. McDowell. Evolution of structure and freevolume in symmetric tilt grain boundaries during dislocation nucleation. ActaMaterialia, 58:6464–6473, 2010.

[185] R.L.J.M. Ubachs, P.J.G. Schreurs, and M.G.D. Geers. A nonlocal diffuse interfacemodel for microstructure evolution of tin-lead solder. Journal of the Mechanicsand Physics of Solids, 52:1763–1792, 2004.

[186] R.Z. Valiev, V.Y. Gertsman, and O.A. Kaibyshev. Grain boundary structure andproperties. Physica Status Solidi (a), 97:11–56, 1986.

Page 149: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

136 Bibliography

[187] P.R.M. van Beers, G.J. McShane, V.G. Kouznetsova, and M.G.D. Geers. Grainboundary interface mechanics in strain gradient crystal plasticity. Journal of theMechanics and Physics of Solids, 61:2659–2679, 2013.

[188] P.R.M. van Beers, V.G. Kouznetsova, M.G.D. Geers, M.A. Tschopp, and D.L.McDowell. A multiscale model of grain boundary structure and energy: Fromatomistics to a continuum description. Acta Materialia, 82:513–529, 2015.

[189] P.R.M. van Beers, V.G. Kouznetsova, and M.G.D. Geers. Grain boundary interfa-cial plasticity with incorporation of internal structure and energy. Submitted.

[190] P.R.M. van Beers, V.G. Kouznetsova, and M.G.D. Geers. Defect redistributionwithin a continuum grain boundary plasticity model. Submitted.

[191] H. van Swygenhoven and J.R. Weertman. Deformation in nanocrystalline metals.Materials Today, 9:24–31, 2006.

[192] R.A. Varin, J.W. Wyrzykowski, W. Lojkowski, and M.W. Grabski. Spreading ofextrinsic grain boundary dislocations in plastically deformed aluminium. PhysicaStatus Solidi (a), 45:565–569, 1978.

[193] A.J. Vattre and M.J. Demkowicz. Determining the Burgers vectors and elasticstrain energies of interface dislocation arrays using anisotropic elasticity theory.Acta Materialia, 61:5172–5187, 2013.

[194] E.S. Venkatesh and L.E. Murr. The influence of grain boundary ledge density onthe flow stress in nickel. Materials Science and Engineering, 33:69–80, 1978.

[195] G.Z. Voyiadjis, D. Faghihi, and Y. Zhang. A theory for grain boundaries withstrain-gradient plasticity. International Journal of Solids and Structures, 51:1872–1889, 2014.

[196] M. Wallin, W.A. Curtin, M. Ristinmaa, and A. Needleman. Multi-scale plasticitymodeling: Coupled discrete dislocation and continuum crystal plasticity. Journalof the Mechanics and Physics of Solids, 56:3167–3180, 2008.

[197] G.J. Wang and V. Vitek. Relationships between grain boundary structure andenergy. Acta Metallurgica, 34:951–960, 1986.

[198] G.J. Wang, A.P. Sutton, and V. Vitek. A computer simulation study of 〈001〉and 〈111〉 tilt boundaries: The multiplicity of structures. Acta Metallurgica, 32:1093–1104, 1984.

[199] M.G. Wang and A.H.W. Ngan. Indentation strain burst phenomenon induced bygrain boundaries in niobium. Journal of Materials Research, 19:2478–2486, 2004.

[200] Y. Wang, J.W. Kysar, S. Vukelic, and Y.L. Yao. Spatially resolved characterizationof geometrically necessary dislocation dependent deformation in microscale lasershock peening. Journal of Manufacturing Science and Engineering, 131:041014–1–9, 2009.

Page 150: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

137

[201] Y.B. Wang and M.L. Sui. Atomic-scale in situ observation of lattice dislocationspassing through twin boundaries. Applied Physics Letters, 94:021909–1–3, 2009.

[202] D.H. Warner, F. Sansoz, and J.F. Molinari. Atomistic based continuum investi-gation of plastic deformation in nanocrystalline copper. International Journal ofPlasticity, 22:754–774, 2006.

[203] Y. Wei and H. Gao. An elasticviscoplastic model of deformation in nanocrystallinemetals based on coupled mechanisms in grain boundaries and grain interiors. Ma-terials Science and Engineering A, 478:16–25, 2008.

[204] P.C. Wo and A.H.W. Ngan. Investigation of slip transmission behavior across grainboundaries in polycrystalline Ni3Al using nanoindentation. Journal of MaterialsResearch, 19:189–201, 2004.

[205] D. Wolf. A Read-Shockley model for high-angle grain boundaries. Scripta Metal-lurgica, 23:1713–1718, 1989.

[206] D. Wolf. Structure-energy correlation for grain boundaries in FCC metals - III.Symmetric tilt boundaries. Acta Metallurgica et Materialia, 38:781–790, 1990.

[207] D. Wolf. A broken-bond model for grain boundaries in face-centered cubic metals.Journal of Applied Physics, 68:3221–3236, 1990.

[208] D. Wolf and S. Yip. Materials Interfaces: Atomic-level Structure and Properties.Chapman & Hall, London, 1992.

[209] D. Wolf, V. Yamakov, S.R. Phillpot, A. Mukherjee, and H. Gleiter. Deforma-tion of nanocrystalline materials by molecular-dynamics simulation: relationshipto experiments? Acta Materialia, 53:1–40, 2005.

[210] S. Wulfinghoff, E. Bayerschen, and T. Bohlke. A gradient plasticity grain boundaryyield theory. International Journal of Plasticity, 51:33–46, 2013.

[211] T. Yalcinkaya, W.A.M. Brekelmans, and M.G.D. Geers. Deformation patterningdriven by rate dependent non-convex strain gradient plasticity. Journal of theMechanics and Physics of Solids, 59:1–17, 2011.

[212] B. Yang and H. Vehoff. Grain size effects on the mechanical properties of nanonickelexamined by nanoindentation. Materials Science and Engineering A, 400-401:467–470, 2005.

[213] J.B. Yang, Y. Nagai, Z.G. Yang, and M. Hasegawa. Quantization of the Frank-Bilbyequation for misfit dislocation arrays in interfaces. Acta Materialia, 57:4874–4881,2009.

[214] W. Yu and Z. Wang. Interactions between edge lattice dislocations and Σ11 sym-metrical tilt grain boundaries in copper: A quasi-continuum method study. ActaMaterialia, 60:5010–5021, 2012.

Page 151: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

138 Bibliography

[215] S. Zaefferer, J.C. Kuo, Z. Zhao, M. Winning, and D. Raabe. On the influence of thegrain boundary misorientation on the plastic deformation of aluminium bicrystals.Acta Materialia, 51:4719–4735, 2003.

[216] Y.W. Zhang and T.W. Wang. Molecular dynamics simulation of interaction of adislocation array from a crack tip with grain boundaries. Modelling and Simulationin Materials Science and Engineering, 4:231–244, 1996.

[217] Z.F. Zhang and Z.G. Wang. Comparison of fatigue cracking possibility along large-and low-angle grain boundaries. Materials Science and Engineering A, 284:285–291, 2000.

[218] Z.F. Zhang, Z.G. Wang, and J. Eckert. What types of grain boundaries can bepassed through by persistent slip bands? Journal of Materials Research, 18:1031–1034, 2003.

[219] T.T. Zhu, A.J. Bushby, and D.J. Dunstan. Materials mechanical size effects: areview. Materials Technology, 23:193–209, 2008.

[220] Y.T. Zhu and T.G. Langdon. Influence of grain size on deformation mechanisms:An extension to nanocrystalline materials. Materials Science and Engineering A,409:234–242, 2005.

Page 152: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Samenvatting

Multischaal modelleren van korrelgrensplasticiteit

Polykristallijne materialen, zoals metalen, bevatten verschillende intrinsieke materiaalde-fecten, waarvan dislocaties en korrelgrenzen er twee belangrijke zijn. Dislocaties zijnlijndefecten die de regelmaat van het kristalrooster verstoren, waarbij hun collectievebeweging de onomkeerbare plastische vervorming van het materiaal veroorzaakt. Korrel-grenzen zijn de planaire defecten tussen twee naburige kristallen met een andere orientatievan het atoomrooster. Beide microstructurele defecten hebben een grote invloed op deresulterende materiaaleigenschappen en het macroscopische gedrag, in het bijzonder alskorrelgrenzen een relatief groot deel van het materiaalvolume vormen, bijvoorbeeld in fi-jnkorrelige metalen, dunne lagen en micro-apparaten. In deze configuraties worden de rolvan de korrelgrenzen en hun interactie met dislocaties van groot belang. In de literatuurzijn verschillende types van interactiemechanismen tussen dislocaties en korrelgrenzengeıdentificeerd op basis van kleinschalige experimenten en computersimulaties: dislo-caties kunnen opstapelen, doorglijden, of geabsorbeerd worden in (of genucleeerd op) dekorrelgrenzen. Een nauwkeurige beschrijving van deze mechanismen op een korrelgrensis een van de grote uitdagingen in het modelleren van de plasticiteit van polykristallijnemetalen, wat het onderwerp is van dit proefschrift.

Het doel van dit proefschrift is het overbruggen van de schaal tussen de kleinschalige inter-actiemechanismen, geassocieerd met plasticiteit op korrelgrenzen, en het polykristallijnecontinuumniveau, waar de meeste ontwerpmethoden voor technische toepassingen zichop verlaten. Dit is tot stand gebracht door de ontwikkeling van een multischaal korrel-grensvlak plasticiteitsraamwerk, dat bruikbaar is in de context van dislocatiegebaseerdegradientverbeterde kristalplasticiteitsmodellen.

Als eerste zijn de fundamenten van het korrelgrensplasticiteitsmodel ontwikkeld doormiddel van het formuleren van balansvergelijkingen en thermodynamisch consistenteconstitutieve vergelijkingen, die zijn opgenomen in een microstructureel gemotiveerderekgradient kristalplasticiteitsraamwerk via de hogere-orde randvoorwaarden. Een be-langrijk ingredient is de vloeiwet op de korrelgrens, welke een evolutievergelijking is voor

139

Page 153: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

140 Samenvatting

een scalaire continuumschaal grootheid op elk glijvlak op de korrelgrens. Deze scalairegrootheid representeert het netto-effect van de verschillende fijnschalige mechanismen dieplaatsvinden op de korrelgrenzen. Alleen atomaire simulaties kunnen de details van destructuur en energie van de korrelgrens ontsluiten. Gedreven door de behoefte aan eenfysisch onderbouwde continuumschaal beschrijving van de korrelgrensenergie voor hetgebruik in continuummodellen, is daarom een multischaal aanpak voorgesteld die leidttot een continuumrepresentatie van de initiele korrelgrensstructuur en energie, gebaseerdop de resultaten van atomaire simulaties. Vervolgens is deze multischaal atomaire-naar-continuum aanpak benut in het raamwerk van kristalplasticiteit met het korrelgrensmodeldoor middel van een generalisatie van de continuumuitdrukking voor de initiele korrel-grensenergie. Tenslotte is een continuummodel ontwikkeld voor de beschrijving van deherverdeling van de defecten langs de korrelgrens in het korrelgrensplasticiteitsmodel.Een diffusieve evolutiewet is voorgesteld voor de herverdeling van het nettodefect binnende korrelgrens, waardoor niet-lokale relaxatie van de lokale opstapeling van het nettode-fect in de korrelgrens mogelijk is. Dit continuummodel bestaat uit een aantal vergelijkin-gen, inbegrepen een balans en constitutieve vergelijkingen voor een vectorieel nettodefectveld langs het korrelgrensvlak.

Om de mogelijkheden van het ontwikkelde korrelgrensplasticiteitsraamwerk in de verschil-lende stadia van de modelontwikkeling te onderzoeken, is het model toegepast om hetmicroscopische en macroscopische gedrag van bikristal referentieproblemen te bestuderen.Gedetailleerde numerieke analyses van de invloed van de constitutieve korrelgrensparam-eters en de kristalgeometrie laten zien dat het macroscopische gedrag aanzienlijk veran-dert ten opzichte van de conventionele modellen voor de korrelgrens, waarbij de kritischeinteracties tussen dislocaties en korrelgrenzen ontbreken.

Het multischaalmodel voor korrelgrensplasticiteit gepresenteerd in dit proefschrift biedteen basis voor een meer diepgaande analyse van polykristallijne metalen op het con-tinuumniveau. Vooral voor materialen waarin grote fracties van korrelgrenzen aanwezigzijn, is de nauwkeurige beschrijving van het gedrag op deze grensvlakken van wezenlijkbelang.

Page 154: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Acknowledgements

Looking back at the last few years, I can definitely state that it has been a true learningexperience, both academically and personally. Many people have contributed to thisexperience and to whom I would like to express some words of gratitude.

I would like to sincerely thank my promotor Marc Geers and copromotor Varvara Kouznet-sova for providing the opportunity of doing this PhD project. I am really grateful fortheir guidance, help, support and stimulating discussions over the past years. I like tothank Graham McShane for the fruitful discussions and contributions to the second chap-ter. Furthermore, I would like to direct acknowledgements to Mark Tschopp and DavidMcDowell for their constructive help, insight and contributions to the third chapter. Aspecial thanks goes to Alice for all sorts of administrative details. I am also grateful tothe committee members for taking the time and effort to read the thesis and providingvaluable comments.

I wish to thank many colleagues, in particular my office mates from over the years(Jan, Bart, Tom, Lambert, Chaowei, Andre, Danqing and Young Joon), for the pleasantworking and social atmosphere, including the variety of discussions, conference trips,necessary and unnecessary distractions.

I owe many thanks to my friends and family, especially my parents and grandparents,who supported me in various ways. Before I finish, I thank my beloved wife Leila,whose constant support, love, care, patience and encouragements kept me focussed andmotivated. Also the joy our daughter Emma brings, makes me realise what matters most.I close with the words ‘baruch hashem adonai’ to him to whom we can come through hisson, the author and finisher of my faith.

Paul van BeersEindhoven, November 2014

141

Page 155: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary
Page 156: Multiscale modelling of grain boundary plasticity · and macroscopic behaviour of bicrystal benchmark problems. Detailed numerical analy-ses of the in uence of the grain boundary

Curriculum Vitae

Paul van Beers was born on the 4th of July 1986 in Heerlen, the Netherlands. In 2004,he finished his secondary education at the Gertrudiscollege in Roosendaal, after whichhe started the Bachelor program Mechanical Engineering at the Eindhoven Universityof Technology. He obtained his Bachelor’s degree (Cum Laude) in 2007 and continuedthe Mechanical Engineering master in the group Mechanics of Materials of prof.dr.ir.M.G.D. Geers. During his master, he did his internship at the Massachusetts Instituteof Technology in the group of prof. R. Radovitzky at the department of Aeronauticsand Astronautics. Following his internship, he started his master thesis within the PhDproject, receiving his Master’s degree (Cum Laude) in 2010. The PhD project entitled‘A multiscale methodology for the analysis of metallic interfaces – WP3’ was performedat the Eindhoven University of Technology in the group Mechanics of Materials, underthe supervision of prof.dr.ir. M.G.D. Geers and dr.ir. V.G. Kouznetsova, of which theresults are presented in this dissertation. During the PhD project, he completed theeducational track of the Graduate School of Engineering Mechanics.

143